Next Article in Journal
HER-2 Expression in Colorectal Cancer and Its Correlation with Immune Cell Infiltration
Next Article in Special Issue
Erucin Exerts Cardioprotective Effects on Ischemia/Reperfusion Injury through the Modulation of mitoKATP Channels
Previous Article in Journal
Taxanes in the Treatment of Head and Neck Squamous Cell Carcinoma
Previous Article in Special Issue
Management of Obesity-Related Inflammatory and Cardiovascular Diseases by Medicinal Plants: From Traditional Uses to Therapeutic Targets
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Resveratrol and beyond: The Effect of Natural Polyphenols on the Cardiovascular System: A Narrative Review

1
Division of Cardiology, 1st Department of Medicine, Medical School, University of Pecs, 7624 Pecs, Hungary
2
Szentágothai Research Centre, University of Pecs, 7624 Pecs, Hungary
3
Department of Biochemistry and Medical Chemistry, University of Pecs, 7624 Pecs, Hungary
4
Medical School, University of Minnesota, Minneapolis, MN 55455, USA
5
Department of Medicine, University of Minnesota Medical School, Minneapolis, MN 55455, USA
6
Department of Medicine, Division of Cardiology, University of Minnesota, Minneapolis, MN 55455, USA
*
Author to whom correspondence should be addressed.
Biomedicines 2023, 11(11), 2888; https://doi.org/10.3390/biomedicines11112888
Submission received: 5 September 2023 / Revised: 11 October 2023 / Accepted: 19 October 2023 / Published: 25 October 2023
(This article belongs to the Special Issue Role of Natural Product in Cardiovascular Disease)

Abstract

:
Cardiovascular diseases (CVDs) are among the leading causes of morbidity and mortality worldwide. Unhealthy dietary habits have clearly been shown to contribute to the development of CVDs. Beyond the primary nutrients, a healthy diet is also rich in plant-derived compounds. Natural polyphenols, found in fruits, vegetables, and red wine, have a clear role in improving cardiovascular health. In this review, we strive to summarize the results of the relevant pre-clinical and clinical trials that focused on some of the most important natural polyphenols, such as resveratrol and relevant flavonoids. In addition, we aim to identify their common sources, biosynthesis, and describe their mechanism of action including their regulatory effect on signal transduction pathways. Finally, we provide scientific evidence regarding the cardiovascular benefits of moderate, long-term red wine consumption.

1. Introduction

According to the World Health Organization (WHO) definition, the term “cardiovascular diseases” (CVDs) encompasses a variety of disorders including coronary artery disease (CAD), stroke, peripheral artery disease (PAD), hypertension, cerebrovascular disease, and heart failure (HF) [1,2,3]. Not only do they represent one of the leading causes of morbidity in countries with western lifestyles, but also contributed to 17.9 million deaths in 2019 with this number expected to exceed 23.6 million annually by 2030.
A variety of risk factors, including hypertension, dyslipidemia, diabetes mellitus, obesity, smoking, alcohol consumption, sedentary lifestyle, and unhealthy diet have been shown to contribute to the development and progression of CVDs [3]. These promote endothelial cell dysfunction, oxidative stress, smooth muscle cell, and fibroblast proliferation, and the transition of macrophages to foam cells. The resultant chronic inflammation ultimately leads to atherosclerotic plaque development across the vascular bed [4]. In addition, the complex interaction between chronic inflammation, oxidative stress, lipid metabolism, and endothelial dysfunction plays a crucial role in the development of HF [4,5,6].
Lifestyle modifications, such as improving dietary habits, are traditionally considered primary determinants of cardiovascular (CV) health. The potential benefits of dietary interventions and nutrient supplementation in improving CV outcomes have been investigated extensively over the past decades. The Mediterranean diet represents a special historical nutrition pattern consumed by populations living in the Mediterranean Basin. As demonstrated by several clinical trials in recent years, prolonged and strict adherence to this diet may reduce the incidence of CVDs by improving the CV risk profile, primarily due to its anti-inflammatory and antioxidant effects [7,8]. There is evidence that modern healthy dietary habits share many similarities with the Mediterranean diet, including the high intake of vegetables, fruits, legumes, seeds, nuts, whole grains, dairy foods, fish, seafood, and vegetable oils (e.g., olive oil) combined with the limited consumption of sweets, soft drinks, and red meat [9,10,11]. While heavy alcohol use can be detrimental, the regular but modest consumption of red wine, which is part of the Mediterranean diet, has well-documented health benefits [12]. As published by Renaud et al. in the early 1990s, the French population has a lower overall incidence and mortality from CVDs, despite a diet relatively rich in saturated fat compared to other developed countries. This observation, also known as the “French paradox”, may be attributable to the higher red wine consumption per capita in France [13].
In addition to the minerals and vitamins found in healthy diets, the role of plant-derived ingredients is also important to mention, such as natural polyphenols. These bioactive phytochemicals can be found in a wide variety of plants (e.g., fruits, vegetables, seeds, nuts), and are also present in wine, tea, spices, oils, and chocolate. They act as natural UV filters and protect from a variety of biotic (e.g., infection) and abiotic stresses (e.g., oxidative, or toxic damages, injury) owing to their antioxidant/anti-inflammatory activities [14,15,16,17,18,19]. Polyphenols also contribute to the flavor, taste, and color of plants. Their dietary intake has seen an increase in recent years with the concomitant decrease in the risk of chronic and age-related disorders, such as cancer, neurodegenerative conditions, inflammation, and CV diseases. Especially in recent years, most consumers prefer using natural ingredients as additives because of their broad availability and safety profile. Grape seed and olive oil extracts rich in polyphenols may be used as food additives [20,21].
A large number of pre-clinical and clinical trials performed over the past 25 years have provided compelling evidence of the beneficial CV effects of polyphenols [22,23,24,25,26,27,28]. However, their exact mechanism of action is yet to be fully elucidated. The most unique feature of this article is that it provides an extensive review of the favorable CV effects of polyphenols from the molecular mechanisms to the clinical data. It focuses primarily on the most relevant natural polyphenols found in red wine, including stilbenes and flavonoids. We summarize their pharmacological properties, and biopharmaceutical classification, describe their main mechanisms of action, and how they may affect cellular signal transduction pathways. We briefly discuss findings from the most relevant pre-clinical and clinical trials with a focus on CV prevention. In addition, this manuscript covers topics and areas that have not been reviewed prior, such as CV outcome trials and the effect of polyphenols on CV mortality. We trust that our review will be beneficial for clinicians and scientists planning future research in the area of polyphenols.
The well-known search engine, PubMed was used to identify literature relevant to this review. The key words included the names of the included compounds, in addition to the terms “cardiovascular”, “cardiac” or “heart”. As the relevant full-text articles were reviewed, the reference section of each manuscript was further reviewed to identify potentially relevant papers. Duplicate articles, as well as small studies with results not relevant to the current paper, were excluded. Not aiming to provide a systematic meta-analysis, the format of narrative review was selected to summarize findings from relevant pre-clinical studies and human clinical trials.

2. Classification, Biosynthesis, and Metabolism of Polyphenols

2.1. Classification and Biosynthesis of Polyphenols

The classification of polyphenols is based on the number of phenolic rings and the structural elements attached to the rings. The majority are derived through the secondary metabolism of L-phenylalanine via the phenylpropanoid pathway. Alternatively, they can be synthesized through the shikimate and/or polyketide pathways, therefore approximately 10,000 structural variants may exist [14,29,30]. The bioactive polyphenols are divided into four major subclasses: (1) flavonoids; (2) non-flavonoid stilbenes; (3) phenolic acids; and (4) lignans. The main sources of relevant polyphenols are depicted in Figure 1 [31,32].
Stilbenes, including the most widely investigated resveratrol (RES), are produced via the general phenylpropanoid pathway. They contain two phenyl residues linked by a two-carbon methylene bridge (1,2-diphenylethylene nucleus). This basic structure can subsequently be glycosylated, methylated, or prenylated by specific enzymes. In addition to RES, other stilbenes include piceatannol, pinosylvin, pinostilbene, combretastatin, polydatin, mulberroside, and various oligostilbenes [33,34]. Overall, a relatively small fraction of habitual diets contain these compounds when compared to flavonoids and phenolic acids (peanuts, berries, grapes, and red wine) [35].
Flavonoids are composed of two aromatic rings and a heterocyclic ring with a typical C6-C3-C6 skeleton structure. They are generally classified into 7 subgroups based on the degree of oxidation, hydroxylation, glycosylation, and substitutions of the central heterocyclic ring: (1) flavonols; (2) flavones; (3) isoflavones; (4) anthocyanidins; (5) flavanones; (6) flavanols; (7) chalcones. Approximately two-third of the polyphenolic compounds are flavonoids and are ubiquitous to most plants [36,37,38].
Phenolic or phenolcarboxylic acids produced by the phenylpropanoid pathway contain aromatic phenolic rings and one carboxylic acid group in their structure. They are among the main groups of polyphenols and are present in one-third of plants. They are divided into two subclasses: (1) hydroxybenzoic acid (e.g., gallic acid); and (2) hydroxycinnamic acid (e.g., caffeic acid, ferulic acid). Their aromatic phenolic rings have the ability to lose an electron, which explains their high antioxidant activity [39,40,41].
Lignans are biphenolic compounds derived from the combination of two phenylpropanoid C6–C3 units synthetized through the shikimate pathway. The ether, lactone, or carbon bonds linked to the biphenolic ring are responsible for the biological activity of these moieties. The primary compounds that are also the most studied include secoisolariciresinol, matairesinol, and pinoresinol [42,43].

2.2. Metabolism of Polyphenols

The bioavailability of polyphenols is poor in general. They undergo rapid and extensive metabolism in the gastrointestinal tract and liver that alters their molecular structure. These changes are naturally responsible for the low plasma levels of intact (unmetabolized) forms [44,45] that, according to the literature, rarely exceed 1 µM after consuming 10 to 100 mg of a polyphenolic compound [46]. The main mechanisms of action of stilbenes are shown in detail in Figure 2.
Polyphenols are ingested primarily in a conjugated form and are absorbed from the stomach, intestine, and colon with the rest eliminated through feces. The metabolism of glycosylated polyphenols (O-glycosides) is initiated by oral glycosidase enzymes, which is followed by conversion into aglycone form (free form) by hydroxylation in the stomach and small intestine. This process is facilitated by intestinal microbiota (phase I metabolism) [47,48]. Aglycone polyphenols are absorbed from the lumen of the small intestine into the cytosol of enterocytes by passive diffusion or by utilizing protein carriers. However, selected polyphenols (such as hydroxycinnamic acids) are resistant to enzymatic digestion and transition unchanged into the colon, where microbiota metabolize them into aglycone form. Following absorption, aglycone polyphenols are transported to the liver via the portal vein, where they are conjugated to O-sulphate or O-glucuronide forms (phase II metabolism). Consequently, polyphenols are typically found in the circulation in conjugated forms. These are filtered by the kidneys and excreted in the urine [49,50,51]. The metabolized forms may also be biologically active, and in fact may be more active than their precursors, as is the case with RES, quercetin (Qct), catechin, and isoflavons [52,53,54,55,56,57]. Furthermore, several polyphenols can penetrate and accumulate in a variety of tissues (e.g., heart, bone, adipose tissue) both in their native and metabolized forms [58,59,60,61,62].

3. Stilbenes and CVDs

Although stilbenes represent one of the main classes of polyphenols, they are only present in a limited number of plants, including grapes, berries, and peanuts. Consequently, their intake through regular diet is limited with red wine representing the primary source. However, their potential benefits for human health, including CVD prevention, are significant and have been the subject of numerous pre-clinical and clinical trials. To date, over 500 molecular variants of stilbenes have been described, such as piceatannol, pinosylvin, and pinostilbene, yet 3,5,4-trihydroxystilbene (RES) is the most prominent and widely studied form [63,64]. RES exists as two geometric isomers: cis- (Z) and trans- (E) with the latter representing the biologically active form [9].

3.1. Mechanism of Action

Stilbenes, and particularly RES, have been extensively investigated over the past decades owing to their potent antioxidant, anti-inflammatory, anti-proliferative, anti-apoptotic and mitochondrial protective effects. Collectively, these are thought to be responsible for their derived CV benefits.
Reactive oxygen (ROS) and nitrogen species (RNS) play an important role in physiological processes (cell homeostasis, signaling pathways) when present at low concentrations. However, high ROS levels promote oxidative stress within the cells and may lead to permanent damage of critical biomolecules (lipids, proteins, DNA) [65]. Although the bioavailability and plasma concentration of stilbenes are low, several in vitro studies have demonstrated that they are important antioxidant agents. Their efficacy is explained by two separate mechanisms: (1) direct accumulation in cardiac and vascular tissues; and (2) active metabolite formation [66,67]. RES has been shown to directly scavenge excessive ROS and RNS, including superoxide (O2·−), hydroxyl radical (·OH), hydrogen-peroxide (H2O2), and peroxynitrite [68,69,70]. As reported by Leonard et al. using murine macrophage cell culture, RES effectively eliminates lipid peroxidation and DNA damage by scavenging hydroxyl, superoxide, and metal-induced radicals [63]. In subsequent experiments, it protected cardiomyocytes [71] and endothelial cells [72] against oxidative toxicity and increased the level of antioxidants. Similar results were obtained using another stilbene and RES analog, piceatannol (3,4,3′,5′-tetrahydroxy-trans-stilbene) [73,74,75,76].
In further experiments, RES was proven to attenuate ROS formation by inhibiting a range of NADPH oxidases (NOX isoforms) and by modulating the activity of the mitochondrial respiratory chain enzymes [77,78,79]. It reduces NOX isoform activity through the upregulation of SIRT1 (NAD+-dependent histone/protein deacetylase sirtuin 1) that, in turn: (1) prompts deacetylation of transcription factor NF (nuclear factor)-κB [80]; (2) upregulates the Forkhead box O1 (FOXO1) transcription factor-mediated synthesis of superoxide dismutase (SOD), as well as the expression of catalase (CAT) and glutathione peroxidase 1 (GPx1) [81,82,83]. Dysfunction of the endothelial nitric oxide synthase (eNOS) due to tetrahydrobiopterin (BH4) deactivation (uncoupling of eNOS) can also contribute to increased ROS production and consequent oxidative damage. RES can reverse eNOS uncoupling by upregulating GTP cyclohydrolase 1 (GCH1) expression, therefore increasing BH4 biosynthesis [84,85]. Furthermore, RES can effectively eliminate free radicals by enhancing the activity of the antioxidant defense system. According to in vitro experiments and in vivo trials, RES prompts nuclear respiratory factor (Nrf-2) activation. Nrf-2, in turn, increases cellular antioxidant content (e.g., GSH) through the upregulation of genes encoding various antioxidant enzymes: uridine-diphosphate-glucuronosyltransferases (UGTs), glutathione S-transferases, γ-glutamylcysteine synthase, and sulfotransferases [86,87].
The effective inhibition of inflammatory processes is a key function of stilbenes [88,89]. NF-κB and STAT (signal transducers and activators of transcription) transcription factors play an important role in the regulation of physiological and inflammatory processes as well as apoptosis in macrophages. The activation of these factors is mediated by TLRs (toll-like receptors—transmembrane proteins of macrophages) and by their downstream signaling pathways, which lead to increased expression of pro-inflammatory cytokines, including IL-1β, IL-6, TNF-α [90,91]. As a potent anti-inflammatory agent, RES can inhibit the expression of pro-inflammatory cytokines [92,93] by suppressing NF-кB and JAK/STAT through the activation of SIRT1 [94,95]. Yeung et al. demonstrated that SIRT1 can inhibit RelA acetylation prompting the silencing of NF-κB activity [96]. Similar results were reported with piceatannol [97,98]. Yang et al. found that RES can alleviate TLR-mediated chronic inflammation by direct downregulation of TLR4 expression [99,100].
The expression of adhesion molecules, such as VCAM-1, ICAM-1, and E-selectin, are important steps in leukocyte activation. According to Zhang et al., RES can successfully inhibit adhesion molecules by suppressing the TNF-α-induced NF-κB activation [101]. Moreover, it can decrease the expression of ICAM-1 and thus monocyte adhesion by Nrf-2 pathway stimulation [102]. In addition to inhibiting pro-inflammatory pathways, stilbenes can also increase the secretion and activation of anti-inflammatory cytokines (e.g., IL-10) and enzymes, such as heme oxygenase-1 (HO-1) [103,104,105].
COX-1 and COX-2 isoforms of the cyclooxygenase enzyme are key catalysts of prostaglandin (PG) biosynthesis and they play an important role in inflammatory processes. Vascular endothelial and smooth muscle cells serve as the major source of PG production. Arachidonic acid is released by endothelial cell membrane phospholipids with the cleavage of phospholipase A2 (PLA2) and then metabolized by COX enzymes to generate PGs (such as PGD2, PGE2, PGI2) and thromboxane (TX) A2. Proinflammatory cytokines (IL-1, TNF-α) play a crucial role in the activation of PG biosynthesis in states of chronic inflammation by upregulation of the NF-κB and MAPK signaling pathways. The ability of RES to inhibit the COX-1 and COX-2 enzymes can contribute to its marked anti-inflammatory (COX-2) and antiplatelet effects (COX-1) [106,107]. RES can reduce COX-2 expression by the downregulation of p65, c-Jun, Fos, and NF-κB transcription factors via SIRT1 activation, thus decreasing eicosanoid production (PGE2, TXA2) and consequently reducing inflammation. Moreover, SIRT1 can also inhibit (acetylation) the AP-1 (activated by MAPKs) transcription factors that are essential for COX-2 gene expression and for PGE2 release in inflammation as well as for endothelial proliferation [108]. In a recent human trial, RES reduced the level of pro-inflammatory cytokines (IL-1, IL-6) by inhibiting oxidative phosphorylation in leukocytes. In addition, it significantly reduced gene expression encoding B cell receptors and leukocyte extravasation signaling [66].
Stilbenes, especially RES, affect different signal transduction pathways that have the ability to modify a variety of cellular functions, including cardiac remodeling, cell survival, apoptosis, maladaptive hypertrophy, fibrosis, and mitochondrial function. It can also regulate the phosphatidylinositol 3-kinase/protein kinase B (PI3K/Akt) pro-survival pathway by activating SIRT1. Enhanced Akt activation can inhibit ROS-mediated apoptosis and improve cardiomyocyte survival through upregulation of FOXO transcription factors, inhibition of the Bcl-2 transcription factor family as well as glycogen synthase kinase-3β (GSK-3β) [109,110]. In addition, RES suppresses the ROS-mediated activation of MAPKs, which are important factors in controlling several cellular processes such as proliferation and apoptosis under both normal and pathological conditions. Overexpression of MAPKs induced by ROS has been shown to be involved in pathological cardiac hypertrophy and remodeling [111]. RES can stimulate the production of MAPK phosphatase-1 (MKP-1), which is the major negative regulator of MAPK activity [112,113]. However, it can inhibit cardiac remodeling through other pathways as well. RES can reduce Ang II-induced cardiac hypertrophy by inhibiting adverse NF-κB signaling activation [114]. Moreover, the downregulation of the serine/threonine-specific protein kinase mammalian target of rapamycin (mTOR) and activation of SIRT1 by RES may be important in inhibiting cardiac and endothelial hypertrophy. mTOR is a downstream target of many signaling pathways including AMPK and PI3K/Akt and has a central role in the regulation of cell growth, survival, and apoptosis, as well as transcription [115,116,117]. Besides maladaptive hypertrophy, cardiac fibrosis is also a key feature of remodeling defined by increased collagen secretion and deposition in the extracellular matrix (ECM) by excessive fibroblast activation [118]. Transforming growth factor β (TGF-β)/Smad2/3 signaling pathway plays a central role in the regulation of ECM production; however, RES-mediated SIRT1 activation can alleviate fibroblast activity by downregulating the TGF-β/Smad2/3 signaling pathway via overexpressing the Smad7 inhibitor protein and silencing miR-17 gene [119,120]. Moreover, RES inhibits phosphatase and tensin homolog (PTEN) degradation, which is a negative regulator for a large number of signaling pathways (PI3K/AKT/mTOR and TGF-β/Smad2/3) and a positive regulator of AMPK (AMP-activated protein kinase). Therefore, it can favorably moderate cardiac hypertrophy and fibrosis [116,121].
Endothelial cells regulate vascular tone, vascular resistance, blood pressure, and hemostasis through the release of vasoactive substances, such as PGs, NO, and vasoconstrictive endothelin-1 (ET-1). Therefore, they play an essential and dynamic role in regulating the CV system. [122,123]. Normal vascular homeostasis can be deranged by various noxious insults and under different pathologic conditions, such as chronic inflammation and oxidative stress. The abnormal upregulation of ET-1 and the reduction in NO levels were shown to be associated with atherosclerosis development and CVDs. Besides their anti-inflammatory and antioxidant properties, stilbenes can modulate the activity of several signaling pathways and improve endothelial function. The beneficial effects of stilbenes on endothelial function and vascular wall tension have been previously linked to increased NO production and decreased ET-1 synthesis. RES enhances NO synthesis in endothelial cells by upregulating eNOS expression and preventing eNOS uncoupling through multiple mechanisms, such as SIRT1 activation [124,125,126]. Furthermore, RES can inhibit the overproduction of ET-1 induced by RAAS and various pro-inflammatory cytokines [127,128]. Piceatannol can also protect endothelial function by enhancing eNOS phosphorylation and increasing NO production [106].
Mitochondria serve as a crucial source of energy for cells as they produce adenosine triphosphate (ATP) in the process known as oxidative phosphorylation [129]. The beneficial effects of RES on mitochondrial biogenesis and function are thoroughly investigated. The activation of the AMPK/SIRT1/PGC1α pathway is a major axis related to mitochondrial biogenesis and the regulation of cellular energy metabolism. AMPK is an energy sensor that controls cellular metabolic function and balances ATP generation. It is stimulated by an increased AMP-to-ATP ratio and in response, shifts cell metabolism towards ATP generation [130]. Downregulation of NOX isoforms by RES induces phosphorylation (stimulation) of AMPK and increases the level of nicotinamide adenine dinucleotide (NAD+), thereby inducing SIRT1 expression (NAD+-dependent protein deacetylase). SIRT1 stimulates deacetylation and activation of the transcriptional coactivator PGC-1α (peroxisome proliferator-activated receptor gamma coactivator 1-alpha), thereby enhancing Nrf-1 and Nrf-2 transcription factors, which regulate the expression of genes encoding certain proteins of the electron transport chain (ETC) [131,132,133]. Moreover, experimental trials demonstrated that RES effectively regulates mitochondrial dynamics (fusion and fission) and other quality control processes in cardiomyocytes through activation of the Sirt1/Sirt3/PGC-1α signaling pathways and enhancing the expression of Drp1 (a regulator protein of mitochondrial fission), among others [134,135].

3.2. Pre-Clinical and Human Clinical Trials with RES Related to CVDs

RES is found at high concentrations in grape skin thus red wine represents an important dietary source (Figure 1). A large number of pre-clinical and human clinical trials have published on the beneficial effects of this stilbene in CVDs and we aim to provide a brief review of pertinent data below (Table 1). Overall, RES is well-tolerated over a wide daily dose range (5 mg–5 g) with mild side effects described in human clinical trials when exceeding 500 mg/day intake. However, the exact cut-off dose for toxicity remains unknown to date.

3.2.1. Pre-Clinical Trials

Vascular Function, Hypertension

Based on several pre-clinical studies, RES has a potent vasorelaxant and thus antihypertensive effect. Multiple mechanisms contribute to this favorable property, including SIRT1 activation, AMPK phosphorylation, increased eNOS activity, and reduced ROS production [185,186,187,188]. Per Dolinsky et al., 320 mg/kg daily RES intake for 5 weeks significantly reduced the blood pressure in spontaneously hypertensive rats (SHRs) as well as mice receiving angiotensin II (AT-II) infusion. Improved vascular relaxation by enhanced eNOS activity and reduced oxidative stress contributed to this finding [189]. Grujic-Milanovic et al. further investigated the antihypertensive and antioxidant effects of RES in rats [190]. SHRs were fed RES at a daily dose of 10 mg/kg or placebo for 8 weeks, which led to an improvement in both systolic (172.04 ± 9.68 vs. 138.21 ± 16.05 mmHg; p < 0.01) and diastolic blood pressure (133.33 ± 12.65 vs. 110.85 ± 14.07; p < 0.01). Redox status and renal function also showed significant improvements in the RES group [190]. Similar results were found by Franco et al. in obese adult rats receiving RES supplementation at 30 mg/kg/day for 30 days [191]. In addition to the above experimental designs, RES was shown to effectively reduce blood pressure in other animal models of hypertension, including renal artery clipping (RES: 5, 10 or 20 mg/kg/day for 4 weeks) [192], hypoxia-induced (RES: 10 mg/day for 4 weeks), AT-II infusion (RES: 146 mg/day for 4 weeks) [185], and rodents fed with fructose [187,193]. Interestingly, Prysyazhna et al. described a paradox mechanism of action for RES in hypertensive mice. While RES reduced blood pressure in their study, the vasorelaxation was caused by cGMP-dependent protein kinase 1α (PKG1α) activation via direct protein oxidation. This pro-oxidant effect is likely due to free radical formation upon electron transfer to oxygen, especially in ROS-damaged tissues, thus facilitating favorable signaling directly at the site of injury [194]. The protective effect of RES on vascular function was further investigated in numerous in vivo and in vitro experimental studies [195,196,197,198].

Cardiac Function, Remodeling

It is well established that oxidative damage and chronic inflammation stimulate specific intracellular signal transduction pathways resulting in adverse cardiac remodeling characterized by fibrosis, cardiomyocyte hypertrophy, apoptosis, and necrosis [199]. Given its potent antioxidant and anti-inflammatory properties, RES was extensively evaluated in different murine models of cardiac remodeling and HF. Gu et al. studied the efficacy of RES at a dose of 2.5 mg/kg/day administered for 16 weeks in a mouse model of HF caused by myocardial infarction (surgical left coronary artery ligation). The results showed improved cardiac function and survival in the group receiving RES, potentially due to SIRT1-mediated upregulation of the AMPK signaling pathway [200]. These results are in agreement with another post-infarction, pre-clinical trial performed by Ahmed et al. [201]. It this study, long-term (10-month) RES treatment at a daily dose of 5 mg/kg significantly improved left ventricular (LV) systolic function in rats [201]. Similar results were published by Kanamori and colleagues [202]. In a subsequent study by Riba et al., RES administered at 15 mg/kg/day for two months reduced the severity of myocardial dysfunction and improved LV function in an isoproterenol-induced post-infarction HF model. These beneficial effects of RES were attributed to its ability to modify the activity of various intracellular signaling pathways (Akt-1/GSK-3β, p38-MAPK, ERK1/2, and iNOS uncoupling) [28]. In a hydroxyeicosatetraenoic acid (HETE)-induced post-infarction rat model of HF, RES supplementation at 5.82 mg/kg/day significantly increased systolic LV function and successfully reversed LV and left atrial (LA) remodeling [203]. Wojciechowski et al. demonstrated in a pressure overload-induced myocardial hypertrophy model that RES can alleviate the progression of cardiac remodeling, dysfunction, and reduce oxidative injury (lipid peroxidation) in cardiac tissue even at lower doses (2.5 mg/kg/day) [204]. In a study by Rimbaud et al., RES supplementation at 18 mg/kg/day for 2 months was associated with decreased cardiac remodeling and improved cardiac function in a high salt diet-induced hypertensive rat model of HF. In addition, RES increased the antioxidant activity of cardiomyocytes, improved mitochondrial function by enhancing PGC-1α cascade activation, and decreased mitochondrial lipid oxidation [205]. Conversely, Sung et al. published that RES supplied at a higher dose (150 mg/kg/day) did not significantly change systolic LV function after 2 weeks, but improved diastolic function, reduced LV diameters/volumes, and was able to reverse cardiac remodeling (fibrosis, hypertrophy) [206]. In another set of experiments, RES-induced SIRT1 activation was shown to represent a key protective mechanism against cardiac dysfunction. Ma et al. have shown that RES supplemented at 25 mg/kg/day inhibits myocardial hypertrophy and fibrosis in mice with HF by enhancing SIRT1 activity, thereby upregulating the PGC-1α/Nrf-1 and Nrf-2 pathways and improving mitochondrial function [133]. The experiments by Bagul et al. have provided further evidence supporting the antihypertrophic effects of RES mediated by SIRT1 when it was used at 10 mg/kg/day for 8 weeks in diabetic rats [84]. Considering outcomes beyond cardiac function, RES given at 450 mg/kg/day for 2 weeks significantly improved exercise capacity and reduced fatigue by improving flow-mediated vasodilatation and vascular function in a mouse model of HF induced by pressure overload [207]. Further pre-clinical trials suggested that RES may pre-condition the myocardium against ischemia-reperfusion (I/R) damage by activating antioxidant, anti-inflammatory, pro-survival, and anti-apoptotic pathways [208,209,210]. However, this benefit seems to be highly dependent on the dose and duration of administration [211].
In addition to the studies detailed above, many others have examined the potential beneficial effects of RES on HF provoked by pressure overload [204,212], diabetes mellitus [213], chemotherapy [214], sepsis [215], myocarditis [216], and genetic mutations [217].

Other Conditions, Risk Factors

Its anti-atherogenic effects are based primarily on the ability to improve vascular and endothelial function as well as favorably modifying the CV risk factor profile. RES supplementation has been shown to reduce serum lipid levels (total cholesterol, LDL-C, and triglyceride) [218,219,220], regulate glucose homeostasis [221,222], reduce vascular inflammation and oxidative stress [223,224] through the downregulation of NF-κB and p38-MAPK pathways and upregulating SIRT1 expression [225,226].

3.2.2. Human Clinical Trials

Hypertension

The success of RES in pre-clinical studies prompted a range of human clinical trials involving populations with various diseases. Theodotou et al. investigated the efficacy of RES supplementation in patients with essential hypertension, adding it on top of standard therapy. A total of 97 patients were enrolled, divided into two groups, and followed for 2 years: (1) RES (50 mg/day) + standard treatment; (2) standard of care alone. The long-term administration of RES led to a reduction in blood pressure when initiated on top of standard medical therapy [145]. Conversely, Walker et al. were unable to detect a significant biological effect of RES given at 1 g/day for 35 days on blood pressure in patients with metabolic syndrome [146]. Similar results were published by Gal et al. in patients with systolic HF [66]. Importantly, these two studies enrolled normotensive patients suggesting that RES supplementation does not provoke hypotension making it a potentially safe agent in those with low baseline blood pressure (such as advanced systolic HF). In a recent meta-analysis by Fogacci et al., low-dose RES did not significantly affect either systolic or diastolic blood pressure but administering it at a higher dose (>300 mg/day) reduced both values significantly [140]. In another meta-analysis, Liu et al. found that daily RES consumption exceeding 150 mg effectively reduced systolic (−11.9 mmHg, p = 0.01) but not diastolic blood pressure [141]. Several individual trials provided data regarding the antihypertensive effect of RES in the setting of different comorbidities, including diabetes [142,143], obesity [144], and fatty liver disease [227].

Vascular Protection

Although the vascular protective effects of RES are well-established in pre-clinical experiments, strong human evidence is lacking. In a pilot trial by Wong et al., RES given at 270 mg/day for 4 weeks led to a significant improvement in endothelial function as measured by flow-mediated dilatation (FMD) of the brachial artery in overweight and hypertensive individuals [147]. Similarly, in a study published by Fujitaka et al., RES (100 mg/day for 3 months) improved endothelial function as measured by FMD in 34 patients with metabolic syndrome [228]. However, no significant change was noted in blood pressure [228]. A study by Magyar et al. found that, compared to placebo, low-dose RES intake (10 mg/day given for 3 months) improved vasorelaxation significantly in a population with chronic angina pectoris [27]. While these studies established the beneficial effects of long-term RES supplementation, a single dose may also be effective. Marques et al. showed that 300 mg of RES improved endothelial function as measured by FMD but there was no change in peripheral and central (aortic) systolic blood pressure [148]. In addition to its beneficial effects on the vasculature, there is ample evidence regarding the potent anti-atherogenic and CV risk-modifying properties of RES, including lipid profile, diabetes, and inflammation [27,69,136,137,138,139,229]. Altogether, these data suggest that RES may be considered an effective compound for CV prevention.

Heart Failure

In contrast to the high number of pre-clinical studies with RES in various HF models, merely a few human clinical trials have been performed to date in populations with HF. In a double-blind, placebo-controlled, randomized clinical trial, Magyar et al. investigated the efficacy of RES intake in post-MI patients with Stage B HF and preserved ejection fraction (baseline EF: 54.77 ± 1.64% in the treated group) [27]. Not unexpectedly, 10 mg/day RES for 3 months did not significantly impact the ejection fraction. However, it significantly improved the diastolic function parameters compared to placebo [27]. In another randomized clinical trial by Militaru et al. involving patients with chronic stable angina pectoris, RES supplemented at 20 mg/day for 2 months led to a significant reduction in serum N-terminal prohormone brain natriuretic peptide (NT-proBNP) levels [230]. To date, only one double-blind, placebo-controlled, randomized clinical trial evaluated the potential benefits of RES in patients with symptomatic systolic HF [66]. The study was performed by our group at the University of Pécs. Sixty patients with New York Heart Association (NYHA) class II-III systolic HF were enrolled and randomly assigned to one of two cohorts: (1) 100 mg/day RES supplementation for 3 months; (2) placebo. All subjects received maximally tolerated guideline-directed medical therapy for HF at the time of enrollment. At the end of the study period, systolic and diastolic function as well as global longitudinal strain improved in the RES cohort with significantly lower serum biomarker (NT-proBNP and galectin-3) and inflammatory cytokine (IL-1 and IL-6) levels. Transcriptomic data indicated attenuated leukocyte activation suggesting that RES has a significant anti-inflammatory effect that may serve as a crucial mechanism mediating its benefits in HF. Important from the patient’s perspective, exercise capacity, as well as the reported quality of life (Qol) were higher in the treatment group [66].

4. Effect of Relevant Flavonoid Compounds on CVDs

Flavonoids are a large group of polyphenolic compounds found in fruits, vegetables, nuts, seeds, coffee, wine, and tea. They are strong antioxidants that are responsible for the color of fruits and flowers, protect them against biotic stressors, and act as preservatives. Several pre-clinical and clinical studies have demonstrated the beneficial cardioprotective effects of dietary flavonoid intake. The most relevant flavonoids with regard to CVD prevention include flavonols (e.g., Qct, kaempferol), flavanones (e.g., hesperidin), flavanols (e.g., catechin, epicatechin), and anthocyanidins (e.g., malvidin) [231,232].

4.1. Mechanism of Action

Flavonoids have several, extensively investigated favorable biochemical effects such as anti-inflammatory, anti-aging, and anti-cancer properties. However, the primary biological activity of these compounds is their profound antioxidant capacity exerted via multiple mechanisms: (1) elimination of ROS; (2) prevention of ROS and RNS production; (3) activation of the antioxidant systems [233,234]. Flavonoids exhibit a marked scavenger activity. The functional hydroxyl group attached to the central ring can donate an electron to peroxynitrite, hydroxyl, and peroxyl radicals through resonance, stabilize these, and create a relatively stable flavonoid radical [235,236,237]. Furthermore, certain flavonoids can directly inhibit the activity of ROS-generating enzymes such as xanthine oxidase, NOS, and NADPH oxidases (NOX isoforms) [238,239,240]. They can also increase the activity of the antioxidant defense system through functional upregulation of selected enzymes such as NADPH-quinone oxidoreductase, glutathione S-transferase, and UDP-glucuronyl transferase [241,242]. In a recent human trial, Zhang et al. demonstrated that anthocyanin significantly increased SOD activity over baseline after 6 weeks of supplementation (p < 0.05), and in a dose-dependent manner [243]. Finally, the antioxidant potential of flavonoids may be further enhanced by modification of the chemical structure, such as polymerization (hydroxyl groups) [244], or glycosylation of anthocyanidins [245].
Several trials reported that flavonoids have potent anti-inflammatory properties. They inhibit prostaglandin synthesis, thereby decreasing leukocyte infiltration and edema development [246]. Some block arachidonic acid release through the inhibition of phospholipase A2 (PLA2) as well as phospholipase C1 (PLC1). Moreover, Qct can block the synthesis of prostaglandins, leukotrienes, and thromboxanes by inhibiting the cyclooxygenase (COX) and lipoxygenase enzymes [247,248,249]. Many flavonoids (e.g., catechin, Qct) can decrease pro-inflammatory cytokine expression, such as IL-6, IL-8, TNF-α, and IL-1β, in macrophages and T-cells while enhancing the production of anti-inflammatory cytokines (e.g., IL-10) [247,250,251]. Nam et al. demonstrated in mice that rutin, a Qct derivative, is not only able to reduce pro-inflammatory cytokine levels but can also decrease NO production by inhibiting iNOS and COX-2 expression [252]. In a recent human trial, Zhang et al. found that anthocyanin supplementation significantly decreased pro-inflammatory cytokine levels, a favorable effect that was dose-dependent [243]. In addition, apigenin successfully inhibited the TNF-α-induced intercellular adhesion molecule-1 (ICAM-1), E-selectin, and vascular cell adhesion molecule-1 (VCAM-1) expression on the surface of endothelial cells [247,251].
Flavonoids have the ability to modulate numerous intracellular signaling factors and signal transduction pathways, such as NF-κB, AP-1, Nrf2, MAPKs, ERK, PI3K/Akt, to reduce oxidative stress and inflammation, thereby acting as cardioprotective agents [253,254,255,256]. NF-κB regulates the expression of several inflammation-associated genes. Inhibition of this central transcription factor can therefore be of significant benefit. Flavonoids inhibit the nuclear translocation of p50 and p65 subunits of NF-κB in macrophages thus decreasing the expression of pro-inflammatory cytokines, adhesion molecules, NOS, and COX-2 [257,258]. In addition, the flavonoid-induced blockade of IκB-α protein phosphorylation can lead to NF-κB pathway inactivation, thereby suppressing TNF-α-induced apoptosis and inflammation [259,260]. Many flavonoid compounds can block pro-inflammatory cytokine production by modulating the MAPK pathway at different levels. In some pre-clinical trials, the inhibition of MAPK phosphorylation (ERK1/2, p38-MAPK, JNK) by flavonoids prompted a decrease in the transcription and expression of pro-inflammatory cytokines, such as TNF-α [261,262,263]. On the other hand, Qct and rutin-provoked downregulation of MAPK phosphorylation led to protection against Ang II-induced hypertrophy in cardiomyocytes [264]. Moreover, flavonoid compounds have significant anti-fibrotic effects owing to the downregulation of TGF-β/SMADs signaling pathways [265,266]. In addition, they can alleviate oxidative stress-induced damage by targeting the PI3K-AKT–mTOR pathway and inhibiting eNOS uncoupling [267,268].
The protective effect of flavonoids on mitochondria is also extensively studied. They can improve mitochondrial function, and cellular energy metabolism, decrease mitochondrial ROS production, and upregulate the expression of selected genes of the respiratory chain complex I (e.g., heart-specific Ndufs4) [269,270]. In addition, they can modulate the process of mitochondrial dynamics (fission and fusion) by inhibiting OMA1 activity [271].
Finally, flavonoids can improve endothelial function and have potent vasorelaxant properties owing to their ability to suppress endothelin-1 activity and increase NO production. Many flavonoids (e.g., Qct, kaempferol, luteolin) can activate the cAMP/protein kinase A (AMPK) signaling pathway, which further upregulates eNOS, resulting in increased concentrations of endothelial NO [272,273,274]. In addition, flavonoids lower intracellular calcium ion concentration by blocking voltage-gated calcium channels [275].

4.2. Pre-Clinical and Human Clinical Trials with Selected Flavonoids Related to CVDs

4.2.1. Quercetin

One of the most widely studied flavonoids is quercetin (Qct) and its derivatives (Qct-glycosides, Qct-ethers, Qct-prenyls) found in apple, blueberry, onion, pepper, broccoli, and tea. They have potent anti-inflammatory and antioxidant effects that are primarily responsible for their positive CV effect (Table 1).

Pre-Clinical Trials

  • Cardiac function, remodeling
The cardioprotective ability of Qct has been broadly examined using in vivo experimental models. Arumugam et al. demonstrated that Qct dosed at 10 mg/kg can protect against the progression of experimental autoimmune myocarditis to dilated cardiomyopathy in rats by inhibiting oxidative stress through modulation of the endothelin 1/MAPK signaling pathways [276]. In another rat model, Qct (25 mg/kg/day) decreased myocardial remodeling and fibrosis by reducing mitochondrial ROS production and total Ca2+ content [277]. According to Kumar et al., pre-treatment with Qct (50 mg/kg/day) for 14 days in rats significantly attenuated the isoproterenol-induced acute myocardial injury and inflammation [278]. Putakala et al. showed that Qct (200 mg/kg/day) protects from cardiac oxidative damage caused by high-fructose intake [279]. In addition, it can protect against myocardial I/R injury. According to Liu et al., it decreases I/R-induced oxidative damage and cardiomyocyte apoptosis in mice through NF-κB pathway inhibition and by PPARγ activation [280]. Based on several studies, Qct can effectively reduce cardiac hypertrophy, fibrosis, and remodeling, and can protect against the development of HF [281,282,283]. In a recent trial, Chang et al. also investigated the effects of Qct on cardiac function and found that if given daily at 50 mg/kg dosage for 4 weeks, it can improve cardiac systolic function, attenuate cardiac hypertrophy, and fibrosis in rats by modulating the TGF-β signaling pathway. Moreover, Qct decreased inflammatory cytokine levels (TNF-α, IL-13, IL-18) and ROS production by improving mitochondrial function and activating SIRT5 expression [284].
  • Vascular function
The vascular endothelial benefits of Qct are also well demonstrated in pre-clinical trials. Machha et al. found in diabetic rats that the daily intake of 10 mg/kg protected vascular function via antioxidant activity and increased endothelium-derived NO levels [285]. According to a pre-clinical trial by Sánchez et al., Qct administration (10 mg/kg for 13 weeks) improved blood pressure and endothelial function in spontaneously hypertensive rats via enhancing NO production and suppressing the NADPH oxidase-mediated ROS generation [286]. Several researchers reported on the renin–angiotensin–aldosterone system inhibitory potential of Qct in pre-clinical trials that may be responsible, at least in part, for its antihypertensive properties and ability to improve endothelial dysfunction [272,287,288]. Finally, Qct and its derivatives can influence major CV risk factors as well, including diabetes mellitus [289,290,291] and dyslipidemia [292,293].

Human Clinical Trials

  • Hypertension
A limited number of human clinical trials evaluated the beneficial effects of Qct on hypertension. Edwards et al. were the first to investigate the potential benefits of this compound in humans. In a randomized, double-blind, placebo-controlled, crossover study, 19 subjects with pre-hypertension and 22 with stage 1 disease were enrolled and administered 730 mg Qct or placebo daily for 4 weeks. There was no significant blood pressure change in the pre-hypertension group, but a significant reduction was documented in those with stage 1 hypertension: systolic (−7 ± 2 mmHg) and diastolic (−5 ± 2 mmHg) [149]. In another randomized clinical trial by Zahedi et al., 500 mg/day Qct administration for 10 weeks significantly decreased systolic blood pressure compared to placebo (−8.8 ± 9.3 vs. −3.5 ± 11.7 mmHg, p = 0.04) in women with type 2 diabetes mellitus (n = 72). However, it had no effect on diastolic blood pressure or on any of the other measured variables, such as lipid profile and inflammatory markers [150]. Furthermore, in 2015 Brüll et al. demonstrated the benefits of Qct in a double-blind, placebo-controlled, crossover trial. After randomization, participants (n = 70) received 162 mg Qct daily (onion skin extract powder) or placebo for 6 weeks [151]. Systolic blood pressure in hypertensives was significantly lower in the treatment cohort vs. controls receiving placebo. In contrast, using a similar dosing scheme, Dower et al. did not find Qct to significantly reduce blood pressure at 4 weeks (n = 37) [152]. However, measurements were made using different methods (office or 24-h ambulatory blood pressure monitoring).
The effect of Qct may be dose-dependent. A meta-analysis by Serban et al. found that supplementation of 500 mg/day and above may significantly reduce both systolic (−4.45 mmHg, p = 0.007) and diastolic (−2.98 mmHg, p < 0.001) blood pressure values; however, the effect was not significant at lower doses [153]. Evaluating its effect on end-organ function, Kondratiuk et al. found that the addition of Qct to standard therapy in hypertensive patients (n = 84) for 12 months significantly improved diastolic echocardiographic parameters when compared to placebo (E/e’: −0.41 ± 0.01 vs. −0.08 ± 0.01; p = 0.001; LV mass index: −3.28 ± 0.02 vs. −2.04 ± 0.03 g/m2, p = 0.02) [157]. In a small double-blind, placebo-controlled trial, Larson et al. investigated the acute effects of Qct in humans. They enrolled 12 subjects with stage 1 hypertension and administered 1095 mg Qct in a single dose that led to a significant reduction in both systolic and diastolic blood pressure at 10 h [154]. When given at a lower dose (50 to 400 mg) as a single dose to healthy individuals, Bondonno et al. found no acute change in blood pressure or in NO-mediated endothelium-dependent vasodilation [155]. Based on these data, the efficacy of Qct on blood pressure may be dose-dependent with longer-term administration needed to observe a significant change.
  • Myocardial ischemia
The number of studies evaluating the effect of Qct in HF and CAD is limited. Chekalina et al. investigated the potential benefits of Qct on myocardial ischemia and hemodynamic parameters in a population with stable CAD receiving standard-of-care therapy in a randomized, placebo-controlled trial. A total of 85 subjects were randomized into two groups: (1) 120 mg daily Qct supplementation for 2 months; (2) continued standard treatment for CAD. LVEF increased by 4.5% and 3.2%, respectively on echocardiography which also revealed an improvement in diastolic function parameters in both groups except for mitral inflow deceleration time (DT), which only changed in the cohort receiving Qct. In addition, 24-h Holter ECG monitoring revealed a significant reduction in the number of ST segment depression episodes as well as the incidence of premature ventricular complexes in the treatment group [156].

4.2.2. Catechins

The potential CV benefits of flavanols (catechin, epicatechin, and their derivative epigallocatechin) have been extensively studied in the past decades (Table 1). The major sources of catechins are apples, pears, cocoa beans, tea, and grapes. Therefore, it is present in popular consumer products such as chocolate, tea (especially green tea), and wine. It is important to point out that flavanols are the major biologically active polyphenol components of green tea and cocoa, therefore many studies have been published using extracts from these products.

Pre-Clinical Trials

  • Vascular function, hypertension
The antihypertensive and vascular protective effects of catechins are well demonstrated in pre-clinical trials. Catechins were shown to decrease blood pressure, improve endothelial function, and induce vasorelaxation [273,274,294,295]. In a recent trial, Munoz et al. investigated the beneficial effects of green tea extract on the CV system in animal models. Supplementation for 20 weeks prevented the development of hypertension and improved endothelial function in obese mice by decreasing arterial inflammation and oxidative stress [296]. In a similar study, Gao et al. found that green tea consumption can decrease salt-induced hypertension in elderly rats by inhibiting the renin–angiotensin–aldosterone system, upregulating eNOS synthesis, improving antioxidant activity, and downregulating pro-inflammatory processes (IL-1β) [297]. In addition, Sabri et al. demonstrated that daily epigallocatechin gallate (EGCG) supplementation at 50 mg/kg for 14 days decreases blood pressure and alleviates vascular dysfunction caused by increasing NO bioavailability in hypertensive mice receiving angiotensin II infusion [298].
Several animal studies performed in the past decades have demonstrated the beneficial effect of catechins on serum lipid levels and their ability to reduce atherosclerotic plaque development by inhibiting pro-inflammatory cytokine formation (e.g., TNF-α) and oxidative stress-induced production of oxidized LDL [299,300,301,302,303].
  • Cardiac function, remodeling
Ischemia-induced oxidative damage of the myocardium leads to cardiomyocyte hypertrophy, myocardial fibrosis, and apoptosis. The myocardial protective effects of catechins during I/R insults may be explained by their antioxidant and anti-inflammatory properties. Ferenczyová et al. provided a thorough summary of the most relevant results related to I/R injury [304]. It is important to emphasize that catechin doses used in pre-clinical studies varied broadly (1–200 mg/kg/day).
Mou et al. published an experimental trial in 2022 using chronic pressure overload to induce HF [270]. Following abdominal aortic constriction (AAC) surgery, rats were fed EGCG for 8 weeks. Authors found that EGCG supplementation helped preserve cardiac function and inhibited myocardial hypertrophy/fibrosis by downregulating the TGF-β signaling pathway activity. In addition, it improved cellular energy metabolism by improving mitochondrial function [270]. Muhammad et al. published similar results with EGCG supplementation at 200 mg/kg/day for one month in aged Wistar albino rats [305]. In addition to these papers, the anti-hypertrophic, anti-fibrotic, and anti-apoptotic properties of catechins have been well described in other pre-clinical animal studies [306,307,308].

Human Clinical Trials

  • Hypertension
Several human clinical trials have investigated the vasoactive and anti-hypertensive activity of catechins [309,310,311]. Bogdanski et al. evaluated the effect of green tea extract on cardiovascular risk factors in obese, hypertensive individuals. In this double-blind, placebo-controlled study, patients (n = 56) received green tea extract (379 mg/day) or placebo for 3 months. Both systolic (−4.9 ± 5.7 mmHg) and diastolic (−4.7 ± 3.2 mmHg) blood pressure values were significantly lower in the treatment group compared to placebo at the end of the follow-up period [158].
Rostami et al. provided further evidence on the ability of catechin-rich cocoa products to reduce blood pressure in patients at high CV risk [159]. In addition, 150 mg/day EGCG supplementation effectively lowered blood pressure in obese patients as published by Chatree et al. [160]. Contrary to the above trials, a study by Dower et al. failed to show a significant change in blood pressure (in-office and 24-hour ambulatory monitoring) and vascular function with 100 mg/day epicatechin supplementation for one month [152]. In a meta-analysis, Ellinger et al. found that epicatechin successfully reduced both systolic and diastolic blood pressure values, but the effect was dose-dependent, and the daily intake had to exceed 25 mg [161]. The positive vascular effects of catechins were also confirmed in smokers. Oyama et al. found that high dose catechin intake (580 mg/day) for 2 weeks improved endothelium-dependent vasorelaxation by increasing NO levels and decreasing ROS activity in this population [162]. The dose-dependent vasodilator activity of catechins was evaluated in a short, double-blind, placebo-controlled trial. Twenty healthy males were enrolled and received different doses of epicatechin (0.1, 0.5, and 1.0 mg/kg) or water as control. There was a significant improvement in vascular function as measured by FMD one hour after 1 mg/kg epicatechin intake and two hours after 0.5 mg/kg when compared to placebo [163].
  • Atherosclerosis and myocardial ischemia
Catechins can also favorably influence the progression of atherosclerosis through CV risk factor modification, such as diabetes and lipid metabolism [152,160,164,312]. It is well known that elevated oxLDL levels induced by oxidative stress are primarily responsible for the development and progression of atherosclerosis and in general, CVDs. In a randomized, placebo-controlled, double-blind trial, Suzuki-Sugihara examined the antioxidant capacity of catechins in healthy male volunteers with 19 subjects eating a total of 1 g of catechin each. Plasma total antioxidant capacity was elevated at one hour following ingestion and LDL oxidizability was significantly reduced [165]. In a small, phase II, double-blind, randomized clinical trial (COCOA-PAD study), patients with known PAD consumed 75 mg epicatechin daily in the form of cocoa (n = 44) or placebo. The average six-minute walking distance increased significantly in the treatment group compared to the placebo at 6-month follow-up (+42.6 m; p = 0.005) [166]. In a human study published more recently, Kishimoto et al. examined the association between green tea and CAD. Investigators randomized 612 subjects between 2008 and 2017 and performed coronary angiography identifying 388 patients with CAD and 138 with MI. Individuals were divided into three groups according to the frequency of green tea consumption (1 cup/day, 1–3 cups/day, and 3 or more cups/day). Higher green tea consumption was significantly inversely associated with CAD and MI prevalence in Japanese adults compared to lower green tea intake (OR: 0.54, 95% CI: 0.30, 0.98) [167]. Wang et al. found similar benefits for catechins in a meta-analysis: green tea consumption (one cup daily) was associated with a 10% CAD risk reduction [168].
  • Heart failure
Data on the effect of catechin supplementation in HF are limited. In a recent trial, Dural et al. examined the benefits of chocolate (containing high amounts of catechins) in patients with systolic HF. They found that NT-proBNP levels decreased significantly with the use of both milk chocolate (356 ± 54.2 vs. 310 ± 72.1 pg/mL; p = 0.007) and dark chocolate (341 ± 57 vs. 301 ± 60.1 pg/mL; p = 0.028) compared to placebo. However, six-minute walking distance was unaffected [169].

4.2.3. Anthocyanins

Anthocyanins or anthocyanidins (aglycon form) include cyanidin, delphinidin, pelargonidin, petunidin, and malvidin. They are naturally occurring, water-soluble pigments responsible for the color of fruits and vegetables, especially berries (predominantly blueberry, cranberry, and strawberry), grapes, and carrots. The habitual intake of anthocyanins is variable worldwide (12–44 mg/day), but wine consumption contributes to approximately 14.4–24.5% of the total intake in Europe. Anthocyanins became of interest as natural therapeutic polyphenols owing to their potent antioxidant and anti-inflammatory properties with proven efficacy against cardiac and vascular disorders (Table 1) [313].

Pre-Clinical Trials

  • Cardiac function, remodeling
Several pre-clinical studies have recently demonstrated the favorable effect of anthocyanins on myocardial hypertrophy, fibrosis, and remodeling by limiting oxidative stress, reducing inflammation, and modulating cellular energy metabolism. In an in vivo murine model by Chen et al., high-dose delphinidin (15 mg/day for 8 weeks) successfully reversed pressure-overload-induced cardiac hypertrophy, LV dysfunction, and dilation, as provoked by transverse aortic constriction (TAC) in mice. One of the key possible mechanisms for these findings is the attenuation of TAC-induced oxidative stress by delphinidin. Notably, no significant benefit could be detected when it was administered at a lower dose (5 mg/day) [314]. In a similar animal model, Hu et al. used 500 mg/kg daily supplementation of anthocyanin-enriched blueberry extract for 6 weeks in mice. Compared to controls, anthocyanin significantly decreased cardiac mass, myocardial hypertrophy, and fibrosis, and increased LV systolic function by alleviating TAC-induced pro-inflammatory cytokine secretion (IL-1β, TNF-α) and ROS production [315]. Aloud et al. investigated the efficacy of anthocyanins on the cardiac function of SHR. Five-week-old male rats received 10 mg/kg anthocyanin or placebo (water) for 15 weeks. Cardiac mass was lower (less cardiac hypertrophy) and diastolic function was better in the treatment group at the end of the follow-up period. However, blood pressure and LV systolic function were unaffected [316]. The cardioprotective effect of anthocyanins was also demonstrated by Chen et al. in a streptozotocin-induced type I diabetes rat model. Authors showed that 250 mg/kg/day anthocyanin extract for 4 weeks significantly attenuated cardiac hypertrophy, reduced fibrosis, and improved LV function indices [317]. Yue et al. reported using another animal model that anthocyanin reduced myocardial fibrosis and hypertrophy, protected cardiac function, and downregulated the IL-17-related chronic inflammation in diabetic mice [318]. Similar favorable results were found by Liu et al. after cyclophosphamide-induced cardiac injury in rats [319] and by Kim et al. in diabetic mice [320]. Toufektsain and colleagues were the first to provide pre-clinical evidence that prolonged (2 months) anthocyanin intake exerted a myocardial protective effect against ischemia-reperfusion injury in rats through enhancing the activity of the antioxidant defense system [321]. Similar findings were published by Guler et al. in rats [322].
  • Vascular function, hypertension
The beneficial effects of anthocyanins on vascular function have also been described. Petersen et al. demonstrated that the use of strawberry extract can limit vascular dysfunction by increasing endothelium-dependent vasorelaxation. In addition, it prompts a reduction in blood pressure by inhibiting chronic endothelial inflammation as well as ROS production [323]. Several other authors have published on the anti-inflammatory and vascular protective effects of anthocyanins [324]. For example, Nasri et al. observed that cyanidin-3-glucoside administration at 10 mg/kg for 2 months prompted a significant increase in endothelium-dependent vasodilatation in diabetic rats as compared to placebo [325]. There is further evidence that anthocyanin may lower blood pressure by affecting endothelial vasorelaxation and the activity of the renin–angiotensin–aldosterone system [326,327]. In a pre-clinical trial by Shaughnessy et al., long-term (2 months) blueberry supplementation significantly reduced systolic blood pressure as early as 4 weeks (19% reduction) with continued decline subsequently [328]. As shown by Xu et al., intracranial infusion of anthocyanin extract decreased the sympathetic nervous system activity and reduced blood pressure in salt-induced hypertensive rats [329]. Furthermore, anthocyanins administered at 50, 100, or 200 mg/kg decreased systolic and diastolic blood pressure as well as heart rate in hypertensive rats in a dose-dependent manner, at least in part through RAAS inhibition [330].
  • Atherosclerosis and risk factors
Several pre-clinical trials have reported on the potent anti-atherogenic properties of anthocyanin-rich fruit extracts, likely related to their lipid-lowering, anti-inflammatory, and antioxidant actions [331,332,333,334,335]. When considering acute CV events, such as MI or stroke, platelet activation (aggregation and adhesion) represents a pivotal event. Therefore, effective inhibition of platelet hyperactivity may be considered a favorable benefit in improving CV outcomes and reducing the risk of adverse events [239]. Importantly, anthocyanins were shown to effectively inhibit collagen- as well as ADP-induced platelet aggregation and adhesion when tested in animal experiments [331,336,337].

Human Clinical Trials

  • Myocardial ischemia
The relationship between anthocyanin supplementation and CVD prevention is relatively well-studied. Cassidy et al. followed 93,600 women (25 to 42 years of age) who were healthy at baseline from the Nurses’ Health Study and examined the association between anthocyanin intake and the risk of MI. During 18 years of follow-up, 405 MI cases were reported and higher anthocyanin intake was inversely associated with the risk of MI after multivariate adjustment (HR: 0.68; 95% CI: 0.49, 0.96; p = 0.03 highest vs. lowest quintiles). These data suggest that high habitual anthocyanin intake may effectively reduce the risk of MI in a population of predominantly young women [170]. In a similar study, Cassidy et al. also examined the effect of anthocyanin supplementation on healthy men (n = 4046). During the 24 years of follow-up, higher habitual intake of these flavonoids was associated with a reduced non-fatal MI and ischemic stroke risk [171].
In addition to the pre-clinical studies detailed above, the antiplatelet properties of anthocyanins were also evaluated in small human clinical trials. Alvarez-Suarez et al. reported that the daily intake of 500 g of strawberries for one month significantly decreased platelet activation and improved the antioxidant status [172]. Aboonabi et al. were also able to demonstrate the antiplatelet activity of 320 mg daily anthocyanin supplementation for one month in humans [173]. Similar results were reported by Santhakumar and colleagues [174] and Tian et al. published the dose-dependent effect of anthocyanins on platelet function [175].
  • Hypertension
As potent vasorelaxant and antioxidant agents, anthocyanins can improve vascular function and reduce blood pressure. In a human clinical trial (n = 1898), Jennings et al. found that anthocyanin-rich food intake led to a significant reduction in systolic blood pressure and improvement in central hemodynamic parameters, such as augmentation index (AI) and pulse wave velocity (PWV) as measured by non-invasive methods [176]. McKay et al. reported that anthocyanin-rich hibiscus tea consumption for 6 weeks by 65 adults with grade I hypertension reduced systolic blood pressure significantly when compared to placebo (−7.2 ± 11.4 vs. −1.3 ± 10.0 mmHg; p = 0.030). However, diastolic blood pressure did not change significantly [177]. Similar benefits were shown by Basu et al. after blueberry consumption [178]. In addition to the aforementioned trials, numerous short-term (less than two months) and long-term studies reported on the antihypertensive effect of an anthocyanin-rich diet (predominantly berries) in hypertensive patients with or without other CV risk factors [179,180,338,339].
Evaluating the acute effects of high anthocyanin intake (480 mL cranberry juice with 94 mg anthocyanin content), Dohadwala et al. found a beneficial effect on vascular function as measured by FMD in subjects with CAD [181]. Several other studies also reported a similar improvement in FMD parameters after short- and long-term consumption of anthocyanin-rich extracts from different berries [340,341,342,343]. In addition to their potent vascular and myocardial protective properties, many trials and meta-analyses were published providing evidence that anthocyanins can positively influence CV risk factors, including diabetes and dyslipidemia [182,183,184,344].

5. The Association between CV Mortality and Polyphenol Intake

The number of trials investigating the association between clinical cardiovascular outcomes and polyphenol use is limited. Sesso et al. performed a large, randomized, double-blind, placebo-controlled clinical trial (COcoa Supplement and Multivitamin Outcomes Study; COSMOS) aiming to evaluate the effect of cocoa intake on CVD prevention [345]. Researchers randomized 21,442 adults aged 60 years or older from the US with no prior history of MI or stroke. Participants received daily cocoa extract supplements (500 mg cocoa flavanols, including 80 mg epicatechin) and/or multivitamin supplements or matching placebo. The primary composite outcome was the incidence of stroke, MI, unstable angina requiring hospitalization, need for coronary revascularization, CV mortality, carotid artery surgery, and peripheral artery surgery. After a median follow-up of 3.6 years, 866 participants had a confirmed CV event. Although there was a favorable trend, the study failed its primary composite endpoint (HR: 0.90; 95% CI: 0.78, p = 1.02). Analyzing the secondary outcomes, there was a significant reduction in CVD deaths in the group treated with cocoa extract compared to placebo [345]. In another prospective trial, McCullough et al. also reported a mortality benefit with flavonoid supplementation. They randomized almost 100,000 individuals, divided them into quintiles based on their total flavonoid intake, and followed them for 7 years. Comparing the highest quintile group to the lowest, there was a significant reduction in CV mortality (RR: 0.82; 95% CI: 0.73, 0.92; p = 0.01). Importantly, there was a signal suggesting that even a relatively small amount of flavonoid-rich food intake may be beneficial for decreasing CV mortality [346]. Similar results were published by Grosso et al. where a high level of flavonoid supplementation was associated with reduced risk of all-cause mortality when compared to low-level consumption (RR: 0.74, 95% CI: 0.55, 0.99, p = 0.01) [347]. The PREDIMED trial evaluated the benefits of the Mediterranean Diet on primary CVD prevention with a special focus on the relationship between polyphenol intake and all-cause mortality in patients at high CV risk. Investigators randomized 7447 subjects and assigned them to quintiles based on their total polyphenol intake. During a mean follow-up of 4.8 years, 327 deaths were observed with a 37% relative reduction in all-cause mortality between patients in the highest vs. the lowest quintiles (HR: 0.63; 95% CI: 0.41, 0.97). Subgroup analyses revealed that the mortality benefit was limited to stilbenes (HR: 0.48; 95% CI: 0.25, 0.91; p-trend = 0.04) and lignan (HR: 0.60; 95% CI: 0.37, 0.95; p-trend = 0.03) [348]. However, more trials are needed to further clarify the potential role of long-term polyphenol intake in CV mortality reduction.

6. Red Wine Consumption and Its Association with CVDs

Wine is a complex hydro-alcoholic solution typically derived from fermented grapes. Since ancient times, wine, particularly red wine, has been part of various diets with known health benefits when consumed in moderate amounts [12]. Red wine contains a large variety of bioactive ingredients including polyphenols such as flavonoids (catechin, quercetin, and anthocyanins) as well as stilbenes (RES). These are thought to be responsible for its beneficial effects on the CV system [349]. The polyphenol content of wines depends strongly on the species and the interaction between environmental factors. For example, red wine contains an approximately 10-fold higher concentration of phenolic compounds, especially RES, than white wine with the discrepancy potentially explained by the winemaking method [350].
Renaud et al. were the first to report in 1992 that moderate red wine intake is inversely associated with the prevalence of ischemic heart disease and CV mortality despite the high saturated fat intake in southern France. This phenomenon is known as the “French paradox” [13]. Since then, several human clinical and epidemiological trials have confirmed that regular, long-term red wine consumption in moderate amounts exerts cardioprotective effects and can positively influence CV risk factors [351], lipid profile, and type-2 diabetes [352,353,354,355,356,357]. Our group has shown that moderate red wine consumption at dinner (200 mL for 3 weeks) has a beneficial effect on hemorheological parameters, including red blood cell aggregability and deformability, in healthy volunteers. This may have a positive effect on the microcirculation thereby reducing the risk of CVDs [358]. Naissides et al. reported that short-term red wine consumption had no impact on postprandial lipid parameters or insulin homeostasis, confirming that the benefit is derived from chronic moderate consumption [359].
Gronbaek et al. reported in 2000 that those consuming low (1–7 glasses per week) and moderate (8–21 glasses per week) amounts of red wine have 20% and 24% lower all-cause mortality compared to those not consuming wine, respectively [360]. However, a J-shaped relationship was observed between total alcohol use and mortality. A similar association was described by other investigators [361,362]. Lucerón-Lucas-Torres published a large systematic review and meta-analysis in 2023 focusing on the association between wine consumption and the incidence of CVDs, coronary artery disease, and CV death [363]. Twenty-five clinical studies from 1985 to 2021 were included in the analysis with almost 1.5 million individuals. There was a strong inverse association between red wine consumption and the risk of ischemic heart disease (RR: 0.76; 95% CI: 0.69, 0.84), CVDs (RR: 0.83; 95% CI: 0.70, 0.98), and CV mortality (RR: 0.73; 95% CI: 0.59, 0.90) [363].
The difference between wine and other alcoholic beverages was also analyzed with regard to their CV beneficial effects. For example, Torres et al. demonstrated that red wine consumption increases total antioxidant capacity and improves pro-inflammatory profile, therefore it was more effective in preventing CVD than other alcoholic drinks [364]. It is critical to emphasize, however, that excessive alcohol consumption, including red wine, represents a serious worldwide public health challenge. It is related to the high prevalence (approximately 4% of all adults), its well-established adverse health effects, and social consequences depending on the volume of alcohol intake over time, type of beverage, and consumption pattern. Ethanol exhibits teratogenic, genotoxic, carcinogenic, hepatotoxic, and neurotoxic properties. In addition, it may cause various gastrointestinal adverse effects (gastritis, ulcer formation, bleeding), and can increase the risk of CVD [365,366,367].
In summary, these findings suggest that regular, low-to-moderate intake of polyphenol-rich red wine has a significant beneficial effect in preventing CVDs and reducing CV mortality. However, further research is necessary to better understand the health benefits of red wine consumption, and to define the optimal quantity based on gender and body weight.

7. Adverse Effects of the Most Relevant Polyphenols

While polyphenols are well-tolerated compounds overall, it is important to mention a few relevant adverse effects that have been reported in clinical trials.
The side effect profile of stilbenes is the most thoroughly investigated when compared to other polyphenols. The development of adverse effects depends primarily on the duration of administration, achieved plasma concentration, as well as the primary cell/tissue type affected [368,369,370,371,372,373]. RES has been administered over a wide dose range in pre-clinical and human studies, however, the cut-off dose for toxicity remains unknown. The most commonly reported adverse effects were gastrointestinal. Brown et al. found in humans that emesis, diarrhea, and mild hepatic dysfunction may occur at a daily dose of 2.5 g and above [374]. Furthermore, Crowell et al. published that extremely high (3 g/kg/day) oral RES supplementation for 4 weeks resulted in nephrotoxicity (reported as elevated serum creatinine level and renal histopathological changes) in rats [375]. In a human study by la Porte et al., 2 g RES twice daily was overall well tolerated by healthy volunteers, however, diarrhea was observed in six of the eight subjects [376].
Only a few dose-dependent side effects were reported for flavonoids. These generally include gastrointestinal (emesis, diarrhea, abdominal pain, and bloating), in addition to insomnia, headache, and palpitations [377]. Younes et al. published that catechin taken as a dietary supplement in doses exceeding 800 mg/day induces a significant increase in serum transaminases [378]. In another trial, Cruz-Correa et al. reported that the combination of oral curcumin and quercetin for three months only led to a limited number of adverse effects, such as nausea and sour taste [379].
In addition to direct side effects, potentially harmful adverse drug interactions have also been reported with polyphenol consumption. They have been shown to reduce the gastrointestinal transport of folic acid and thiamine and to change the activity of drugs through interaction with drug transporters and digestive enzymes leading to inhibition and increased bioavailability [380]. For example, polyphenols can unfavorably influence iron absorption owing to their chelating effect [381].

8. Conclusions and Future Directions

Unhealthy dietary habits are known to contribute to the development and progression of CVDs and mortality. Numerous in vitro and in vivo pre-clinical studies as well as human clinical trials have shown that the consumption of polyphenol-rich diet, including moderate red wine consumption, may be beneficial in the prevention and treatment of CVDs. In this narrative review, we aimed to summarize the primary mechanisms of action for RES and relevant flavonoids in association with CVDs. Their antioxidant and anti-inflammatory properties, and ability to influence various intracellular processes are relatively well described. In pre-clinical trials, RES and flavonoids could effectively reduce blood pressure, enhance vascular function, and attenuate LV hypertrophy and remodeling while improving LV function. In addition, these polyphenols have potent anti-atherogenic effects and can positively influence several CV risk factors such as lipid profile, diabetes, and vascular inflammation. A large number of clinical trials have also been performed confirming the potent cardioprotective effects of these polyphenols in humans.
Despite these published reports, further randomized, interventional clinical studies are needed with extended follow-up time and large cohort enrollment. In addition, it will be critical to establish optimal dosing for each compound and describe potential interaction with other, already approved medications. It is important to emphasize that “natural products or extracts” do not inherently represent ‘safer’ substances than “synthetic products”, therefore detailed investigations are necessary to establish their side effect profile. Based on the literature review, researchers to date have only studied a small fraction of the approximately 10,000 described polyphenols leaving a vast area open for research. Given the increasing prevalence of CVDs, these may prove to be extremely beneficial for cardiovascular prevention.

Author Contributions

Conceptualization: R.H., T.A., F.G.J. and L.C.; first draft preparation: R.G.; review and editing: L.C., R.H., T.A., M.T., P.M. and K.T. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by NKFIH within the framework of the University of Pécs project TKP2021-EGA-17.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Laslett, L.J.; Alagona, P., Jr.; Clark, B.A.; Drozda, J.P., Jr.; Saldivar, F.; Wilson, S.R.; Poe, C.; Hart, M. The worldwide environment of cardiovascular disease: Prevalence, diagnosis, therapy, and policy issues: A report from the American College of Cardiology. J. Am. Coll. Cardiol. 2012, 60, S1–S49. [Google Scholar] [CrossRef] [PubMed]
  2. Yu, C.; Moore, B.M.; Kotchetkova, I.; Cordina, R.L.; Celermajer, D.S. Causes of death in a contemporary adult congenital heart disease cohort. Heart 2018, 104, 1678–1682. [Google Scholar] [CrossRef] [PubMed]
  3. Cardiovascular Diseases (CVDs)—World Health Organization. Available online: https://www.who.int/news-room/fact-sheets/detail/cardiovascular-diseases-(cvds) (accessed on 4 September 2023).
  4. Chen, J.; Zhang, X.; Millican, R.; Sherwood, J.; Martin, S.; Jo, H.; Yoon, Y.-S.; Brott, B.C.; Jun, H.-W. Recent advances in nanomaterials for therapy and diagnosis for atherosclerosis. Adv. Drug Deliv. Rev. 2021, 170, 142–199. [Google Scholar] [CrossRef] [PubMed]
  5. Banez, M.J.; Geluz, M.I.; Chandra, A.; Hamdan, T.; Biswas, O.S.; Bryan, N.S.; Von Schwarz, E.R. A systemic review on the antioxidant and anti-inflammatory effects of resveratrol, curcumin, and dietary nitric oxide supplementation on human cardiovascular health. Nutr. Res. 2020, 78, 11–26. [Google Scholar] [CrossRef] [PubMed]
  6. Steven, S.; Frenis, K.; Oelze, M.; Kalinovic, S.; Kuntic, M.; Jimenez, M.T.B.; Vujacic-Mirski, K.; Helmstädter, J.; Kröller-Schön, S.; Münzel, T.; et al. Vascular Inflammation and Oxidative Stress: Major Triggers for Cardiovascular Disease. Oxid. Med. Cell. Longev. 2019, 2019, 7092151. [Google Scholar] [CrossRef] [PubMed]
  7. Finicelli, M.; Di Salle, A.; Galderisi, U.; Peluso, G. The Mediterranean Diet: An Update of the Clinical Trials. Nutrients 2022, 14, 2956. [Google Scholar] [CrossRef] [PubMed]
  8. Martínez-González, M.A.; Gea, A.; Ruiz-Canela, M. The Mediterranean Diet and Cardiovascular Health. Circ. Res. 2019, 124, 779–798. [Google Scholar] [CrossRef] [PubMed]
  9. Mozaffarian, D. Dietary and policy priorities for cardiovascular disease, diabetes, and obesity: A comprehensive review. Circulation 2016, 133, 187–225. [Google Scholar] [CrossRef]
  10. Silveira, B.K.S.; Oliveira, T.M.S.; Andrade, P.A.; Hermsdorff, H.H.M.; Rosa, C.O.B.; Franceschini, S.D.C.C. Dietary pattern and macronutrients profile on the variation of inflammatory biomarkers: Scientific Update. Cardiol. Res. Pract. 2018, 2018, 4762575. [Google Scholar]
  11. Visseren, F.L.J.; Mach, F.; Smulders, Y.M.; Carballo, D.; Koskinas, K.C. 2021 ESC Guidelines on cardiovascular disease prevention in clinical practice. Eur. Heart J. 2021, 42, 3227–3337. [Google Scholar] [CrossRef]
  12. Willett, W.C.; Sacks, F.; Trichopoulou, A.; Drescher, G.; Ferro-Luzzi, A.; Helsing, E.; Trichopoulos, D. Mediterranean diet pyramid: A cultural model for healthy eating. Am. J. Clin. Nutr. 1995, 61, 1402S–1406S. [Google Scholar] [CrossRef] [PubMed]
  13. Renaud, S.; De Lorgeril, M. Wine, alcohol, platelets, and the French paradox for coronary heart disease. Lancet 1992, 339, 1523–1526. [Google Scholar] [CrossRef] [PubMed]
  14. Quideau, S.; Deffieux, D.; Douat-Casassus, C.; Pouységu, L. Plant polyphenols: Chemical properties, biological activities, and synthesis. Angew. Chem. Int. Ed. 2011, 50, 586–621. [Google Scholar] [CrossRef] [PubMed]
  15. Hooper, L.; Kroon, P.A.; Rimm, E.B.; Cohn, J.S.; Harvey, I.; Le Cornu, K.A.; Ryder, J.J.; Hall, W.L.; Cassidy, A. Flavonoids, flavonoid-rich foods, and cardiovascular risk: A meta-analysis of randomized controlled trials. Am. J. Clin. Nutr. 2008, 88, 38–50. [Google Scholar] [CrossRef] [PubMed]
  16. Tungmunnithum, D.; Thongboonyou, A.; Pholboon, A.; Yangsabai, A. Flavonoids and Other Phenolic Compounds from Medicinal Plants for Pharmaceutical and Medical Aspects: An Overview. Medicines 2018, 5, 93. [Google Scholar] [CrossRef] [PubMed]
  17. Chen, D.; Daniel, K.G.; Kuhn, D.J.; Kazi, A.; Bhuiyan, M.; Li, L.; Dou, Q.P. Green tea and tea polyphenols in cancer prevention. Front. Biosci. 2004, 9, 2618. [Google Scholar] [CrossRef] [PubMed]
  18. Chen, G.L.; Fan, M.X.; Wu, J.L.; Li, N.; Guo, M.Q. Antioxidant and anti-inflammatory properties of flavonoids from lotus plumule. Food Chem. 2019, 277, 706–712. [Google Scholar] [CrossRef] [PubMed]
  19. Sarker, U.; Oba, S. Polyphenol and flavonoid profiles and radical scavenging activity in leafy vegetable Amaranthus gangeticus. BMC Plant Biol. 2020, 20, 499. [Google Scholar] [CrossRef]
  20. Martillanes, S.; Rocha-Pimienta, J.; Cabrera-Bañegil, M.; Martín-Vertedor, D.; Delgado-Adámez, J. Application of Phenolic Compounds for Food Preservation: Food Additive and Active Packaging. In Phenolic Compounds-Biological Activity; BoD–Books on Demand: London, UK, 2017. [Google Scholar] [CrossRef]
  21. Maqsood, S.; Benjakul, S.; Shahidi, F. Emerging Role of Phenolic Compounds as Natural Food Additives in Fish and Fish Products. Crit. Rev. Food Sci. Nutr. 2013, 53, 162–179. [Google Scholar] [CrossRef]
  22. Andriantsitohaina, R.; Auger, C.; Chataigneau, T.; Étenne-Selloum, N.; Li, H.; Martínez, M.C.; Schini-Kerth, V.B.; Laher, I. Molecular Mechanisms of the Cardiovascular Protective Effects of Polyphenols. Br. J. Nutr. 2012, 108, 1532–1549. [Google Scholar] [CrossRef]
  23. Gal, R.; Praksch, D.; Kenyeres, P.; Rabai, M.; Toth, K.; Halmosi, R.; Habon, T. Hemorheological Alterations in Patients with Heart Failure with Reduced Ejection Fraction Treated by Resveratrol. Cardiovasc. Ther. 2020, 2020, 7262474. [Google Scholar] [CrossRef] [PubMed]
  24. Castaldo, L.; Narváez, A.; Izzo, L.; Graziani, G.; Gaspari, A.; Minno, G.; Ritieni, A. Red Wine Consumption and Cardiovascular Health. Molecules 2019, 24, 3626. [Google Scholar] [CrossRef] [PubMed]
  25. Fraga, C.G.; Croft, K.D.; Kennedye, D.O.; Tomás-Barberán, F.A. The effects of polyphenols and other bioactives on human health. Food Funct. 2019, 10, 514–528. [Google Scholar] [CrossRef] [PubMed]
  26. Recio, M.; Andujar, I.; Rios, J. Anti-Inflammatory Agents from Plants: Progress and Potential. Curr. Med. Chem. 2012, 19, 2088–2103. [Google Scholar] [CrossRef] [PubMed]
  27. 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. Hemorheol. Microcirc. 2012, 50, 179–187. [Google Scholar] [CrossRef] [PubMed]
  28. Riba, A.; Deres, L.; Sumegi, B.; Toth, K.; Szabados, E.; Halmosi, R. Cardioprotective Effect of Resveratrol in a Postinfarction Heart Failure Model. Oxid. Med. Cell. Longev. 2017, 2017, 6819281. [Google Scholar] [CrossRef] [PubMed]
  29. Li, A.N.; Li, S.; Zhang, Y.J.; Xu, X.R.; Chen, Y.M.; Li, H. Resources and biological activities of natural polyphenols. Nutrients 2014, 6, 6020–6047. [Google Scholar] [CrossRef]
  30. Rana, A.; Samtiya, M.; Dhewa, T.; Mishra, V.; Aluko, R.E. Health benefits of polyphenols: A concise review. J. Food Biochem. 2022, 46, e14264. [Google Scholar] [CrossRef]
  31. Manach, C.; Williamson, G.; Morand, C.; Scalbert, A.; Remesy, C. Bioavailability and bioefficacy of polyphenols in humans. I. Review of 97 bioavailability studies. Am. J. Clin. Nutr. 2005, 81, 230s–242s. [Google Scholar] [CrossRef]
  32. Bešlo, D.; Došlić, G.; Agić, D.; Rastija, V.; Šperanda, M.; Gantner, V.; Lučić, B. Polyphenols in Ruminant Nutrition and Their Effects on Reproduction. Antioxidants 2022, 11, 970. [Google Scholar] [CrossRef]
  33. Zhang, Y.; Butelli, E.; Alseekh, S.; Tohge, T.; Rallapalli, G.; Luo, J.; Kawar, P.G. Multi-level engineering facilitates the production of phenylpropanoid compounds in tomato. Nat. Commun. 2015, 6, 8635. [Google Scholar] [CrossRef] [PubMed]
  34. Dubrovina, A.S.; Kiselev, K.V. Regulation of stilbene biosynthesis in plants. Planta 2017, 246, 597–623. [Google Scholar] [CrossRef] [PubMed]
  35. El Khawand, T.; Courtois, A.; Valls, J.; Richard, T.; Krisa, S. A review of dietary stilbenes: Sources and bioavailability. Phytochem. Rev. 2018, 17, 1007–1029. [Google Scholar] [CrossRef]
  36. Pandey, K.B.; Rizv, S.I. Plant polyphenols as dietary antioxidants in human health and disease. Oxid. Med. Cell. Longev. 2009, 2, 270–278. [Google Scholar] [CrossRef] [PubMed]
  37. Dretcanu, G.; Stirbu, I.; Leopold, N.; Cruceriu, D.; Danciu, C.; Stănilă, A. Chemical Structure, Sources and Role of Bioactive Flavonoids in Cancer Prevention: A Review. Plants 2022, 11, 1117. [Google Scholar] [CrossRef] [PubMed]
  38. Santos, E.; Maia, B.; Ferriani, A.; Teixeira, S. Flavonoids: Classification, Biosynthesis and Chemical Ecology; InTechOpen: London, UK, 2017. [Google Scholar] [CrossRef]
  39. Tsao, R.; Deng, D. Separation procedures for naturally occurring antioxidant hytochemicals. J. Chromatogr. B Analyt. Technol. Biomed. Life Sci. 2004, 812, 85–99. [Google Scholar] [CrossRef] [PubMed]
  40. Sevgi, K.; Tepe, B.; Sarikurkcu, C. Antioxidant and DNA damage protection potentials of selected phenolic acids. Food Chem. Toxicol. 2015, 77, 12–21. [Google Scholar] [CrossRef] [PubMed]
  41. Kim, K.H.; Tsao, R.; Yang, R.; Cui, S.W. Phenolic acid profiles and antioxidant activities of wheat bran extracts and the effect of hydrolysis conditions. Food Chem. 2006, 95, 466–473. [Google Scholar] [CrossRef]
  42. Durazzo, A. Lignans. In Phenolic Compounds in Food: Characterization and Analysis (Food Analysis and Properties); Leo, M.L.N., Ed.; CRC Press: Boca Raton, FL, USA, 2018. [Google Scholar]
  43. Lewis, N.G.; Davin, L.B. Lignans: Biosynthesis and function. In Comprehensive Natural Products Chemistry; Barton, D., Nakanishi, K., Meth-Cohn, O., Eds.; Elsevier: Amsterdam, The Netherlands, 1999; pp. 639–712. [Google Scholar]
  44. D’Archivio, M.; Filesi, C.; Di Benedetto, R.; Gargiulo, R.; Giovannini, C.; Masella, R. Polyphenols, dietary sources and bioavailability. Ann. dell’Istituto Super. Sanita 2007, 43, 348–361. [Google Scholar]
  45. Gambini, J.; Inglés, M.; Olaso, G.; Lopez-Grueso, R.; Bonet-Costa, V.; Abdelaziz, K.M. Properties of resveratrol: In vitro and in vivo studies about metabolism, bioavailability, and biological effects in animal models and humans. Oxid. Med. Cell. Longev. 2015, 2015, 837042. [Google Scholar] [CrossRef]
  46. Scalbert, A.; Williamson, G. Dietary Intake and Bioavailability of Polyphenols. J. Nutr. 2000, 130, 2073S–2085S. [Google Scholar] [CrossRef] [PubMed]
  47. De-Souza, J.E.; Casanova, L.M.; Costa, S.S. Bioavailability of phenolic compounds: A major challenge for drug development? Rev. Fitos 2015, 9, 55–67. [Google Scholar]
  48. Lippolis, T.; Cofano, M.; Caponio, G.R.; De Nunzio, V.; Notarnicola, M. Bioaccessibility and bioavailability of diet polyphenols and their modulation of gut microbiota. Int. J. Mol. Sci. 2023, 24, 3813. [Google Scholar] [CrossRef] [PubMed]
  49. Rein, M.J.; Renouf, M.; Cruz-Hernandez, C.; Actis-Goretta, L.; Thakkar, S.K.; da Silva Pinto, M. Bioavailability of bioactive food compounds: A challenging journey to bioefficacy. Br. J. Clin. Pharmacol. 2013, 75, 588–602. [Google Scholar] [CrossRef] [PubMed]
  50. Stevens, J.F.; Maier, C.S. The chemistry of gut microbial metabolism of polyphenols. Phytochem. Rev. 2016, 15, 425–444. [Google Scholar] [CrossRef] [PubMed]
  51. Montané, X.; Kowalczyk, O.; Reig-Vano, B.; Bajek, A.; Roszkowski, K.; Tomczyk, R. Current perspectives of the applications of polyphenols and flavonoids in cancer therapy. Molecules 2020, 25, 3342. [Google Scholar] [CrossRef] [PubMed]
  52. Kim, D.-H.; Jung, E.-A.; Sohng, I.-S.; Han, J.-A.; Kim, T.-H.; Han, M.J. Intestinal bacterial metabolism of flavonoids and its relation to some biological activities. Arch. Pharmacol. Res. 1998, 21, 17. [Google Scholar] [CrossRef] [PubMed]
  53. Walle, T. Bioavailability of resveratrol. Ann. N. Y. Acad. Sci. 2011, 1215, 9–15. [Google Scholar] [CrossRef]
  54. Patel, K.R.; Brown, V.A.; Jones, D.J.; Britton, R.G.; Hemingway, D.; Miller, A.S.; West, K.P.; Booth, T.D.; Perloff, M.; Crowell, J.A.; et al. Clinical pharmacology of resveratrol and its metabolites in colorectal cancer patients. Cancer Res. 2010, 70, 7392–7399. [Google Scholar] [CrossRef]
  55. Walle, T. Methylation of Dietary Flavones Increases Their Metabolic Stability and Chemopreventive Effects. Int. J. Mol. Sci. 2009, 10, 5002–5019. [Google Scholar] [CrossRef]
  56. Zeng, X.; Su, W.; Bai, Y.; Chen, T.; Yan, Z.; Wang, J.; Su, M.; Zheng, Y.; Peng, W.; Hao, H. Urinary metabolite profiling of flavonoids in Chinese volunteers after consumption of orange juice by UFLC-Q-TOF-MS/MS. J. Chromatogr. B Analyt. Technol. Biomed. Life Sci. 2017, 1061–1062, 79–88. [Google Scholar] [CrossRef] [PubMed]
  57. Borges, G.; Ottaviani, J.I.; van der Hooft, J.J.J.; Schroeter, H.; Crozier, A. Absorption, metabolism, distribution and excretion of (−)-epicatechin: A review of recent findings. Mol. Asp. Med. 2018, 61, 18–30. [Google Scholar] [CrossRef] [PubMed]
  58. Tung, W.C.; Rizzo, B.; Dabbagh, Y.; Saraswat, S.; Romanczyk, M.; Codorniu-Hernández, E.; Rebollido-Rios, R. Polyphenols bind to low density lipoprotein at biologically relevant concentrations that are protective for heart disease. Arch. Biochem. Biophys. 2020, 694, 108589. [Google Scholar] [CrossRef]
  59. Sergides, C.; Chirilă, M.; Silvestro, L.; Pitta, D.; Pittas, A. Bioavailability and safety study of resveratrol 500 mg tablets in healthy male and female volunteers. Exp. Ther. Med. 2016, 11, 164–170. [Google Scholar] [CrossRef] [PubMed]
  60. Alldritt, I.; Whitham-Agut, B.; Sipin, M.; Studholme, J.; Trentacoste, A.; Tripp, J.A. Metabolomics reveals diet-derived plant polyphenols accumulate in physiological bone. Sci. Rep. 2019, 9, 8047. [Google Scholar] [CrossRef] [PubMed]
  61. Wang, L.; Shen, X.; Mi, L.; Jing, J.; Gai, S. Simultaneous determinations of four major bioactive components in Acacia catechu (L.f.) Willd and Scutellaria baicalensis Georgi extracts by LC-MS/MS: Application to its herb-herb interactions based on pharmacokinetic, tissue distribution and excretion studies in rats. Phytomedicine 2019, 56, 64–73. [Google Scholar] [PubMed]
  62. Böhmdorfer, M.; Szakmary, A.; Schiestl, R.H.; Vaquero, J.; Riha, J.; Brenner, S.; Thalhammer, T.; Szekeres, T.; Jäger, W. Involvement of UDP-glucuronyltransferases and Sulfotransferases in the Excretion and Tissue Distribution of Resveratrol in Mice. Nutrients 2017, 9, 1347. [Google Scholar] [CrossRef] [PubMed]
  63. Leonard, S.S.; Xia, C.; Jiang, B.; Stinefelt, B.; Klandorf, H.; Harris, G.K.; Shi, X. Resveratrol scavenges reactive oxygen species and effects radical-induced cellular responses. Biochem. Biophys. Res. Commun. 2003, 309, 1017–1026. [Google Scholar] [CrossRef]
  64. Zamora-Ros, R.; Andres-Lacueva, C.; Lamuela-Raventós, R.; Berenguer, T.; Jakszyn, P.; Martínez, C.; Sánchez, M.J. Concentrations of resveratrol and derivatives in foods and estimation of dietary intake in a Spanish population: European Prospective Investigation into Cancer and Nutrition (EPIC)-Spain cohort. Br. J. Nutr. 2008, 100, 188–196. [Google Scholar] [CrossRef]
  65. Li, H.; Horke, S.; Forstermann, U. Vascular oxidative stress, nitric oxide and atherosclerosis. Atherosclerosis 2014, 237, 208–219. [Google Scholar] [CrossRef]
  66. Gal, R.; Deres, L.; Horvath, O.; Eros, K.; Sandor, B.; Urban, P.; Soos, S.; Marton, Z.; Sumegi, B.; Toth, K.; et al. Resveratrol Improves Heart Function by Moderating Inflammatory Processes in Patients with Systolic Heart Failure. Antioxidants 2020, 9, 1108. [Google Scholar] [CrossRef] [PubMed]
  67. Su, M.; Dong, D.; Wan, J.; Zhou, M. Pharmacokinetics, tissue distribution and excretion study of trans-resveratrol-3-O-glucoside and its two metabolites in rats. Phytomedicine 2019, 58, 152882. [Google Scholar] [CrossRef] [PubMed]
  68. Jia, Z.; Zhu, H.; Misra, B.R.; Mahaney, J.E.; Li, Y.; Misra, H.P. EPR studies on the superoxide-scavenging capacity of the nutraceutical resveratrol. Mol. Cell. Biochem. 2008, 313, 187–194. [Google Scholar] [CrossRef]
  69. Holthoff, J.H.; Woodling, K.A.; Doerge, D.R.; Burns, S.T.; Hinson, J.A.; Mayeux, P.R. Resveratrol, a dietary polyphenolic phytoalexin, is a functional scavenger of peroxynitrite. Biochem. Pharmacol. 2010, 80, 1260–1265. [Google Scholar] [CrossRef] [PubMed]
  70. Mahal, H.S.; Mukherjee, T. Scavenging of reactive oxygen radicals by Resveratrol: Antioxidant effect. Res. Chem. Intermed. 2006, 32, 59–71. [Google Scholar] [CrossRef]
  71. Wang, J.; He, D.; Zhang, Q.; Han, Y.; Jin, S.; Qi, F. Resveratrol protects against Cisplatin-induced cardiotoxicity by alleviating oxidative damage. Cancer Biother. Radiopharm. 2009, 24, 675–680. [Google Scholar] [CrossRef] [PubMed]
  72. Ungvari, Z.; Orosz, Z.; Rivera, A.; Labinskyy, N.; Xiangmin, Z.; Olson, S.; Podlutsky, A.; Csiszar, A. Resveratrol increases vascular oxidative stress resistance. Am. J. Physiol. Heart Circ. Physiol. 2007, 292, H2417–H2424. [Google Scholar] [CrossRef] [PubMed]
  73. Hosoda, R.; Hamada, H.; Uesugi, D.; Iwahara, N.; Nojima, I.; Horio, Y.; Kuno, A. Different Antioxidative and Antiapoptotic Effects of Piceatannol and Resveratrol. J. Pharmacol. Exp. Ther. 2021, 376, 385–396. [Google Scholar] [CrossRef]
  74. Les, F.; Deleruyelle, S.; Cassagnes, L.-E.; Boutin, J.A.; Balogh, B.; Arbones-Mainar, J.M.; Biron, S.; Marceau, P.; Richard, D.; Nepveu, F.; et al. Piceatannol and resveratrol share inhibitory effects on hydrogen peroxide release, monoamine oxidase and lipogenic activities in adipose tissue, but differ in their antilipolytic properties. Chem. Biol. Interact. 2016, 258, 115–125. [Google Scholar] [CrossRef]
  75. Murias, M.; Jager, W.; Handler, N.; Erker, T.; Horvath, Z.; Szekeres, Z.; Nohl, H.; Gille, L. Antioxidant, prooxidant and cytotoxic activity of hydroxylated resveratrol analogues: Structure-activity relationship. Biochem. Pharmacol. 2005, 69, 903–912. [Google Scholar] [CrossRef]
  76. Yokozawa, T.; Kim, Y.J. Piceatannol inhibits melanogenesis by its antioxidative actions. Biol. Pharm. Bull. 2007, 30, 2007–2011. [Google Scholar] [CrossRef] [PubMed]
  77. Xia, N.; Forstermann, U.; Li, H. Resveratrol as a gene regulator in the vasculature. Curr. Pharm. Biotechnol. 2014, 15, 401–408. [Google Scholar] [CrossRef] [PubMed]
  78. Ungvari, Z.; Labinskyy, N.; Mukhopadhyay, P.; Pinto, J.T.; Bagi, Z.; Ballabh, P.; Zhang, C.; Pacher, P.; Csiszar, A.; Tarantini, S.; et al. Resveratrol attenuates mitochondrial oxidative stress in coronary arterial endothelial cells. Am. J. Physiol. Heart Circ. Physiol. 2009, 297, H1876–H1881. [Google Scholar] [CrossRef] [PubMed]
  79. Xia, N.; Daiber, A.; Förstermann, U.; Li, H. Antioxidant effects of resveratrol in the cardiovascular system. Br. J. Pharmacol. 2017, 174, 1633–1646. [Google Scholar] [CrossRef] [PubMed]
  80. Bagul, B.K.; Deepthi, N.; Sultana, R.; Banerjee, S.K. Resveratrol ameliorates cardiac oxidative stress in diabetes through deacetylation of NFkB-p65 and histone 3. J. Nutr. Biochem. 2015, 26, 1298–1307. [Google Scholar] [CrossRef] [PubMed]
  81. Spanier, G.; Xu, H.; Xia, N.; Tobias, S.; Deng, S.; Wojnowski, L.; Forstermann, U.; Li, H. Resveratrol reduces endothelial oxidative stress by modulating the gene expression of superoxide dismutase 1 (SOD1), glutathione peroxidase 1 (GPx1) and NADPH oxidase subunit (Nox4). J. Physiol. Pharmacol. 2009, 60, 111–116. [Google Scholar] [PubMed]
  82. Hsu, C.P.; Zhai, P.; Yamamoto, T.; Maejima, Y.; Matsushima, S.; Hariharan, N.; Shao, D.; Takagi, H.; Oka, S.; Sadoshimaet, J. Silent information regulator 1 protects the heart from ischemia/reperfusion. Circulation 2010, 122, 2170–2182. [Google Scholar] [CrossRef] [PubMed]
  83. Khan, A.M.; Chen, H.; Wan, X.; Tania, M.; Xu, A.; Chen, F.; Zhang, D. Regulatory effects of resveratrol on antioxidant enzymes: A mechanism of growth inhibition and apoptosis induction in cancer cells. Mol. Cells 2013, 35, 219–225. [Google Scholar] [CrossRef]
  84. Li, H.; Forstermann, U. Pharmacological Prevention of eNOS Uncoupling. Curr. Pharm. Des. 2014, 20, 3595–3606. [Google Scholar] [CrossRef]
  85. Xia, N.; Daiber, A.; Habermeier, A.; Closs, E.I.; Thum, T.; Spanier, G.; Lu, Q.; Oelze, M.; Torzewski, M.; Lackner, K.J.; et al. Resveratrol reverses endothelial nitric-oxide synthase uncoupling in apolipoprotein E knockout mice. J. Pharmacol. Exp. Ther. 2010, 335, 149–154. [Google Scholar] [CrossRef]
  86. Li, S.; Zlia, G.; Chen, L.; Ding, Y.; Lian, J.; Hong, G.; Lu, Z. Resveratrol protects mice from paraquat-induced lung injury: The important role of SIRT1 and NRF2 antioxidant pathways. Mol. Med. Rep. 2016, 13, 1833–1838. [Google Scholar] [CrossRef] [PubMed]
  87. Meng, Q.; Guo, T.; Li, G.; Sun, S.; He, S.; Cheng, B.; Shi, B.; Shan, A. Dietary resveratrol improves antioxidant status of sows and piglets and regulates antioxidant gene expression in placenta by Keap1-Nrf2 pathway and Sirt1. J. Anim. Sci. Biotechnol. 2018, 9, 34. [Google Scholar] [CrossRef] [PubMed]
  88. Malaguarnera, L. Influence of resveratrol on the immune response. Nutrients 2019, 11, 946. [Google Scholar] [CrossRef] [PubMed]
  89. Dvorakova, M.; Landa, P. Anti-inflammatory activity of natural stilbenoids: A review. Pharmacol. Res. 2017, 124, 126–145. [Google Scholar] [CrossRef] [PubMed]
  90. Marko, L.; Vigolo, E.; Hinze, C.; Park, J.K.; Roel, G.; Balogh, A.; Choi, M.; Wubken, A.; Cording, J.; Blasig, I.E.; et al. Tubular Epithelial NF-kappaB Activity Regulates Ischemic AKI. J. Am. Soc. Nephrol. 2016, 27, 2658–2669. [Google Scholar] [CrossRef] [PubMed]
  91. Severgnini, M.; Takahashi, S.; Rozo, L.M.; Homer, R.J.; Kuhn, C.; Jhung, J.W.; Perides, G.; Steer, M.; Hassoun, P.M.; Fanburg, B.L.; et al. Activation of the STAT pathway in acute lung injury. Am. J. Physiol. Lung Cell. Mol. Physiol. 2004, 286, L1282–L1292. [Google Scholar] [CrossRef] [PubMed]
  92. Csiszar, A. Anti-inflammatory effects of resveratrol: Possible role in prevention of age-related cardiovascular disease. Ann. N. Y. Acad. Sci. 2011, 1215, 117–122. [Google Scholar] [CrossRef] [PubMed]
  93. Wang, M.; Weng, X.; Chen, H.; Chen, Z.; Liu, X. Resveratrol inhibits TNF-α-induced inflammation to protect against renal ischemia/reperfusion injury in diabetic rats. Acta Cir. Bras. 2020, 35, e202000506. [Google Scholar] [CrossRef]
  94. Ma, C.; Wang, Y.; Dong, L.; Li, M.; Cai, W. Anti-inflammatory effect of resveratrol through the suppression of NF-kB and JAK/STAT signaling pathways. Acta Bioch. Bioph. Sin. 2015, 47, 207–213. [Google Scholar] [CrossRef]
  95. Pinheiro, D.; de Oliveira, A.; Coutinho, L.G.; Fontes, F.L.; de Medeiros, O.R.; Oliveira, T.T.; Faustino, A.; Lira, D.S.V.; de Melo, C.J.; Lajus, T.; et al. Resveratrol decreases the expression of genes involved in inflammation through transcriptional regulation. Free Radic. Biol. Med. 2019, 130, 8–22. [Google Scholar] [CrossRef]
  96. Yeung, F.; Hoberg, J.E.; Ramsey, C.S.; Keller, M.D.; Jones, D.R.; Frye, R.A.; Mayo, M.W. Modulation of NF-κB-dependent transcription and cell survival by the SIRT1 deacetylase. EMBO J. 2004, 23, 2369–2380. [Google Scholar] [CrossRef] [PubMed]
  97. Jin, C.Y.; Moon, D.O.; Lee, K.J.; Kim, M.O.; Lee, J.D.; Choi, Y.H.; Park, Y.M.; Kim, G.Y. Piceatannol attenuates lipopolysaccharide-induced NF-κB activation and NF-κB -related proinflammatory mediators in BV2 microglia. Pharmacol. Res. 2006, 54, 461–467. [Google Scholar] [CrossRef] [PubMed]
  98. Su, L.; David, M. Distinct mechanisms of STAT phosphorylation via the interferon-a/b receptor. Selective inhibition of STAT3 and STAT5 by piceatannol. J. Biol. Chem. 2000, 275, 12661–12666. [Google Scholar] [CrossRef] [PubMed]
  99. Yang, Y.; Li, S.; Yang, Q.; Shi, Y.; Zheng, M.; Liu, Y.; Chen, F.; Song, G.; Xu, H.; Wan, T.; et al. Resveratrol reduces the proinflammatory effects and lipopolysaccharide- induced expression of HMGB1 and TLR4 in RAW264.7 cells. Cell. Physiol. Biochem. 2014, 33, 1283–1292. [Google Scholar] [CrossRef] [PubMed]
  100. Narayanankutty, A. Toll like receptors as a novel therapeutic target for natural products against chronic diseases. Curr. Drug Target. 2019, 20, 1068–1080. [Google Scholar] [CrossRef]
  101. Zhang, Y.; Liu, H.; Tang, W.; Qiu, Q.; Peng, J. Resveratrol prevents TNF- α-induced VCAM-1 and ICAM-1 upregulation in endothelial progenitor cells via reduction of NF- κB activation. Int. Med. Res. 2020, 48, 300060520945131. [Google Scholar] [CrossRef] [PubMed]
  102. Seo, Y.; Park, J.; Choi, W.; Ju Son, D.; Sung Kim, Y.; Kim, M.; Yoon, B.; Pyee, J.; Tae Hong, J.; Go, Y.; et al. Antiatherogenic effect of resveratrol attributed to decreased expression of ICAM-1 (intercellular adhesion molecule-1). Arterioscler. Thromb. Vasc. Biol. 2019, 39, 675–684. [Google Scholar] [CrossRef] [PubMed]
  103. Palacz-Wrobel, M.; Borkowska, P.; Paul-Samojedny, M.; Kowalczyk, M.; Fila-Danilow, A.; Suchanek-Raif, R.; Kowalski, J. Effect of apigenin, kaempferol and resveratrol on the gene expression and protein secretion of tumor necrosis factor alpha (TNF-α) and interleukin-10 (IL-10) in RAW-264.7 macrophages. Biomed. Pharmacother. 2017, 93, 1205–1212. [Google Scholar] [CrossRef]
  104. Son, Y.; Chung, H.; Pae, H. Differential effects of resveratrol and its natural analogues, piceatannol and 3,5,40-trans-trimethoxystilbene, on anti-inflammatory heme oxigenase-1 expression in RAW264.7 macrophages. BioFactors 2014, 40, 138–145. [Google Scholar] [CrossRef]
  105. Jeong, S.O.; Son, Y.; Lee, J.H.; Cheong, Y.K.; Park, S.H.; Chung, T.H.; Pae, H.O. Resveratrol analogue piceatannol restores the palmitic acid-induced impairment of insulin signaling and production of endothelial nitric oxide via activation of anti-inflammatory and antioxidative heme oxygenase-1 in human endothelial cells. Mol. Med. Rep. 2015, 12, 937–944. [Google Scholar] [CrossRef]
  106. Szewczuk, L.M.; Forti, L.; Stivala, L.A.; Penning, T.M. Resveratrol is a peroxidase mediated inactivator of COX-1 but not COX-2: A mechanistic approach to the design of COX-1 selective agents. J. Biol. Chem. 2004, 279, 22727–22737. [Google Scholar] [CrossRef] [PubMed]
  107. Meng, T.; Xiao, D.; Muhammed, A.; Deng, J.; Chen, L.; He, J. Anti-Inflammatory Action and Mechanisms of Resveratrol. Molecules 2021, 26, 229. [Google Scholar] [CrossRef] [PubMed]
  108. Yang, C.M.; Chen, Y.W.; Chi, P.L.; Lin, C.C.; Hsiao, L.D. Resveratrol inhibits BK-induced COX-2 transcription by suppressing acetylation of AP-1 and NF-κB in human rheumatoid arthritis synovial fibroblasts. Biochem. Pharmacol. 2017, 132, 77–91. [Google Scholar] [CrossRef] [PubMed]
  109. Matsui, T.; Tao, J.; del Monte, F.; Lee, K.H.; Li, L.; Picard, M.; Force, T.L.; Franke, T.F.; Hajjar, R.J.; Rosenzweig, A. Akt activation preserves cardiac function and prevents injury after transient cardiac ischemia in vivo. Circulation 2001, 104, 330–335. [Google Scholar] [CrossRef] [PubMed]
  110. Miyamoto, S.; Murphy, A.N.; Brown, J.H. Akt mediated mitochondrial protection in the heart: Metabolic and survival pathways to the rescue. J. Bioenerg. Biomemb. 2009, 41, 169–180. [Google Scholar] [CrossRef] [PubMed]
  111. Muslin, A.J. MAPK signalling in cardiovascular health and disease: Molecular mechanisms and therapeutic targets. Clin. Sci. (Lond.) 2008, 115, 203–218. [Google Scholar] [CrossRef] [PubMed]
  112. Singh, A.K.; Vinayak, M. Resveratrol alleviates inflammatory hyperalgesia by modulation of reactive oxygen species (ROS), antioxidant enzymes and ERK activation. Inflamm. Res. 2017, 66, 911–921. [Google Scholar] [CrossRef] [PubMed]
  113. Kondohand, K.; Nishida, E. Regulation of MAPkinases by MAP kinase phosphatases. Biochim. Biophys. Acta Mol. Cell Res. 2007, 1773, 1227–1237. [Google Scholar] [CrossRef]
  114. Ma, E.; Wu, C.; Chen, J.; Wo, D.; Ren, D.; Yan, H.; Peng, L.; Zhu, W. Resveratrol prevents Ang II-induced cardiac hypertrophy by inhibition of NF-κB signaling. Biomed. Pharmacother. 2023, 165, 115275. [Google Scholar] [CrossRef]
  115. Song, Y.; Huang, W.; Zheng, J.; Yan, M.; Cheng, R.; Zhao, L.; Chen, C.; Hu, C.; Jia, W. Resveratrol reduces intracellular reactive oxygen species levels by inducing autophagy through the AMPK-mTOR pathway. Front. Med. 2018, 12, 697–706. [Google Scholar] [CrossRef]
  116. Sanches-Silva, A.; Testai, L.; Nabavi, S.F.; Battino, M.; Pandima Devi, K.P.; Tejada, S.; Sureda, A.; Xu, S.; Yousefi, B.; Majidinia, M.; et al. Therapeutic potential of polyphenols in cardiovascular diseases: Regulation of mTOR signaling pathway. Pharmacol. Res. 2020, 152, 104626. [Google Scholar] [CrossRef] [PubMed]
  117. Hosoda, R.; Nakashima, R.; Yano, M.; Iwahara, N.; Asakura, S.; Nojima, I.; Saga, Y.; Kunimoto, R.; Horio, Y.; Kuno, A. Resveratrol, a SIRT1 activator, attenuates aging-associated alterations in skeletal muscle and heart in mice. J. Pharmacol. Sci. 2023, 152, 112–122. [Google Scholar] [CrossRef] [PubMed]
  118. Travers, J.G.; Kamal, F.A.; Robbins, J.; Yutzey, K.E.; Blaxall, B.C. Cardiac Fibrosis: The Fibroblast Awakens. Circ. Res. 2016, 118, 1021–1040. [Google Scholar] [CrossRef] [PubMed]
  119. Liu, Z.H.; Zhang, Y.; Wang, X.; Fan, X.F.; Zhang, Y.; Li, X.; Gong, Y.S.; Han, L.P. SIRT1 activation attenuates cardiac fibrosis by endothelial-to-mesenchymal transition. Biomed. Pharmacother. 2019, 118, 109227. [Google Scholar] [CrossRef] [PubMed]
  120. Zhang, Y.; Lu, Y.; Ong’achwa, M.J.; Ge, L.; Qian, Y.; Chen, L.; Hu, X.; Li, F.; Wei, H.; Zhang, C.; et al. Resveratrol Inhibits the TGF-β1-Induced Proliferation of Cardiac Fibroblasts and Collagen Secretion by Downregulating miR-17 in Rat. BioMed. Res. Int. 2018, 2018, 8730593. [Google Scholar] [CrossRef] [PubMed]
  121. Dolinsky, V.W.; Soltys, C.L.M.; Rogan, K.R.; Chan, A.Y.M.; Nagendran, J.; Wang, S.; Dyck, J.R.B. Resveratrol prevents pathological but not physiological cardiac hypertrophy. J. Mol. Med. (Berl.) 2015, 93, 413–425. [Google Scholar] [CrossRef] [PubMed]
  122. Zhao, Y.; Vanhoutte, P.M.; Leung, S.W. Vascular nitric oxide: Beyond eNOS. J. Pharmacol. Sci. 2015, 129, 83–94. [Google Scholar] [CrossRef] [PubMed]
  123. Forstermann, U.; Xia, N.; Li, H. Roles of Vascular Oxidative Stress and Nitric Oxide in the Pathogenesis of Atherosclerosis. Circ. Res. 2017, 120, 713–735. [Google Scholar] [CrossRef]
  124. Parsamanesh, N.; Asghari, A.; Sardari, S.; Tasbandi, A.; Jamialahmadi, T.; Xu, S.; Sahebkar, A. Resveratrol and endothelial function: A literature review. Pharmacol. Res. 2021, 170, 105725. [Google Scholar] [CrossRef]
  125. Xia, N.; Forstermann, U.; Li, H. Resveratrol and endothelial nitric oxide. Molecules 2014, 19, 16102–16121. [Google Scholar] [CrossRef]
  126. Carrizzo, A.; Puca, A.; Damato, A.; Marino, M.; Franco, E.; Pompeo, F.; Traficante, A.; Civitillo, F.; Santini, L.; Trimarco, V.; et al. Resveratrol improves vascular function in patients with hypertension and dyslipidemia by modulating NO metabolism. Hypertension 2013, 62, 359–366. [Google Scholar] [CrossRef] [PubMed]
  127. Hostenbach, S.; D’Haeseleer, M.; Kooijman, R.; De Keyser, J. Modulation of Cytokine-Induced Astrocytic Endothelin-1 Production as a Possible New Approach to the Treatment of Multiple Sclerosis. Front. Pharmacol. 2020, 10, 1491. [Google Scholar] [CrossRef] [PubMed]
  128. Zhang, X.; Wang, Y.; Yang, W.; Hou, X.; Zou, J.; Cao, K. Resveratrol inhibits angiotensin II-induced ERK1/2 activation by downregulating quinone reductase 2 in rat vascular smooth muscle cells. J. Biomed. Res. 2012, 26, 103–109. [Google Scholar] [CrossRef] [PubMed]
  129. Bhatti, J.S.; Bhatti, G.K.; Reddy, P.H. Mitochondrial dysfunction and oxidative stress in metabolic disorders—A step towards mitochondria based therapeutic strategies. Biochim. Biophys. Acta Mol. Basis Dis. 2017, 1863, 1066–1077. [Google Scholar] [CrossRef] [PubMed]
  130. Steinberg, G.R.; Hardie, D.G. New insights into activation and function of the AMPK. Nat. Rev. Mol. Cell Biol. 2023, 24, 255–272. [Google Scholar] [CrossRef] [PubMed]
  131. Guo, S.; Yao, Q.; Ke, Z.; Chen, H.; Wu, J.; Liu, C. Resveratrol attenuates high glucose-induced oxidative stress and cardiomyocyte apoptosis through AMPK. Mol. Cell. Endocrinol. 2015, 412, 85–94. [Google Scholar] [CrossRef] [PubMed]
  132. Huang, Y.; Lu, J.; Zhan, L.; Wang, M.; Shi, R.; Yuan, X.; Gao, X.; Liu, X.; Zang, J.; Liu, W.; et al. Resveratrol-induced Sirt1 phosphorylation by LKB1 mediates mitochondrial metabolism. J. Biol. Chem. 2021, 297, 100929. [Google Scholar] [CrossRef] [PubMed]
  133. Ma, S.; Feng, J.; Zhang, R.; Chen, J.; Han, D.; Li, X.; Yang, B.; Li, X.; Fan, M.; Li, C.; et al. SIRT1 Activation by Resveratrol Alleviates Cardiac Dysfunction via Mitochondrial Regulation in Diabetic Cardiomyopathy Mice. Oxid. Med. Cell. Longev. 2017, 2017, 4602715. [Google Scholar] [CrossRef]
  134. Ren, X.; Chen, L.; Xie, J.; Zhang, Z.; Dong, G.; Liang, J.; Liu, L.; Zhou, H.; Luo, P. Resveratrol Ameliorates Mitochondrial Elongation via Drp1/Parkin/PINK1 Signaling in Senescent-Like Cardiomyocytes. Oxid. Med. Cell. Longev. 2017, 2017, 4175353. [Google Scholar] [CrossRef]
  135. Zheng, M.; Bai, Y.; Sun, X.; Fu, R.; Liu, L.; Liu, M.; Li, Z.; Huang, X. Resveratrol Reestablishes Mitochondrial Quality Control in Myocardial Ischemia/Reperfusion Injury through Sirt1/Sirt3-Mfn2-Parkin-PGC-1 Pathway. Molecules 2022, 27, 5545. [Google Scholar] [CrossRef]
  136. Boa, S.; Ponzoa, V.; Cicconeb, G.; Evangelistab, A.; Sabaa, F.; Goitrea, I.; Procopioa, M.; PaganoaM, G.F.; Cassadera, M.; Gambinoa, R. Six months of resveratrol supplementation has no measurable effect in type 2 diabetic patients. A randomized, double blind, placebo-controlled trial. Pharmacol. Res. 2016, 111, 896–905. [Google Scholar] [CrossRef] [PubMed]
  137. Asgary, S.; Karimi, R.; Momtaz, S.; Naseri, R.; Farzaei, M.H. Effect of resveratrol on metabolic syndrome components: A systematic review and meta-analysis. Rev. Endocr. Metab. Disord. 2019, 20, 173–186. [Google Scholar] [CrossRef] [PubMed]
  138. Hoseini, A.; Namazi, G.; Farrokhian, A.; Reiner, Z.; Aghadavod, E.; Bahmani, F.; Asemi, Z. The effects of resveratrol on metabolic status in patients with type 2 diabetes mellitus and coronary heart disease. Food Funct. 2019, 10, 6042–6051. [Google Scholar] [CrossRef] [PubMed]
  139. Simental-Mendía, L.E.; Guerrero-Romero, F. Effect of resveratrol supplementation on lipid profile in subjects with dyslipidemia: A randomized double-blind, placebo-controlled trial. Nutrition 2019, 58, 7–10. [Google Scholar] [CrossRef] [PubMed]
  140. Fogacci, F.; Tocci, G.; Presta, V.; Fratter, A.; Borghi, C.; Cicero, A.F.G. Effect of resveratrol on blood pressure: A systematic review and meta-analysis of randomized, controlled, clinical trials. Crit. Rev. Food Sci. Nutr. 2019, 59, 1605–1618. [Google Scholar] [CrossRef] [PubMed]
  141. Liu, Y.; Ma, W.; Zhang, P.; He, S.; Huang, D. Effect of resveratrol on blood pressure: A meta-analysis of randomized controlled trials. Clin. Nutr. 2015, 34, 27–34. [Google Scholar] [CrossRef] [PubMed]
  142. Bhatt, J.K.; Thomas, S.; Nanjan, M.J. Resveratrol supplementation improves glycemic control in type 2 diabetes mellitus. Nutr. Res. 2012, 32, 537–541. [Google Scholar] [CrossRef] [PubMed]
  143. Imamura, H.; Yamaguchi, T.; Nagayama, D.; Saiki, A.; Shirai, K.; Tatsuno, I. Resveratrol ameliorates arterial stiffness assessed by cardio-ankle vascular index in patients with type 2 diabetes mellitus. Int. Heart J. 2017, 58, 577–583. [Google Scholar] [CrossRef]
  144. 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.; et al. Calorie restriction-like effects of 30 days of resveratrol supplementation on energy metabolism and metabolic profile in obese humans. Cell Metab. 2011, 14, 612–622. [Google Scholar] [CrossRef]
  145. Theodotou, M.; Fokianos, K.; Mouzouridou, A.; Konstantinou, C.; Aristotelous, A.; Prodromou, D.; Chrysikou, A. The effect of resveratrol on hypertension: A clinical trial. Exp. Ther. Med. 2017, 13, 295–301. [Google Scholar] [CrossRef]
  146. Walker, J.M.; Eckardt, P.; Aleman, J.O.; Da Rosa, J.C.; Liang, Y.; Iizumi, T.; Etheve, S.; Blaser, M.J.; Breslow, J.L.; Holt, P.R. The effects of trans-resveratrol on insulin resistance, inflammation, and microbiota in men with the metabolic syndrome: A pilot randomized, placebo-controlled clinical trial. J. Clin. Transl. Res. 2019, 4, 122–135. [Google Scholar] [PubMed]
  147. Wong, R.H.X.; Howe, P.R.C.; Buckley, J.D.; Coates, A.M.; Kunz, I.; Berry, N.M. Acute resveratrol supplementation improves flow-mediated dilatation in overweight/obese individuals with mildly elevated blood pressure. Nutr. Metab. Cardiovasc. Dis. 2011, 21, 851–856. [Google Scholar] [CrossRef] [PubMed]
  148. Marques, B.C.A.A.; Trindade, M.; Aquino, J.C.F.; Cunha, A.R.; Gismondi, R.O.; Neves, M.F.; Oigman, W. Beneficial effects of acute trans-resveratrol supplementation in treated hypertensive patients with endothelial dysfunction. Clin. Exp. Hypertens. 2018, 40, 218–223. [Google Scholar] [CrossRef] [PubMed]
  149. Edwards, R.; Lyon, T.; Litwin, S.; Rabovsky, A.; Symons, J. Quercetin Reduces Blood Pressure in Hypertensive Subjects. J. Nutr. 2007, 137, 2405–2411. [Google Scholar] [CrossRef] [PubMed]
  150. Zahedi, M.; Ghiasvand, R.; Feizi, A.; Asgari, G.; Darvish, L. Does quercetin improve cardiovascular risk factors and inflammatory biomarkers in women with type 2 diabetes: A double-blind randomized controlled clinical trial. Int. J. Prev. Med. 2013, 4, 777–785. [Google Scholar] [PubMed]
  151. Brüll, V.; Burak, C.; Stoffel-Wagner, B.; Wolffram, S.; Nickenig, G.; Müller, C.; Langguth, P.; Alteheld, B.; Fimmers, R.; Naaf, S.; et al. Effects of a quercetin-rich onion skin extract on 24 h ambulatory blood pressure and endothelial function in overweight-to-obese patients with pre-hypertension: A randomised double-blinded placebo-controlled cross-over trial. Br. J. Nutr. 2015, 114, 1263–1277. [Google Scholar] [CrossRef] [PubMed]
  152. Dower, J.I.; Geleijnse, J.M.; Gijsbers, L.; Zock, P.L.; Kromhout, D.; Hollman, P.C.H. Effects of the pure flavonoids epicatechin and quercetin on vascular function and cardiometabolic health: A randomized, double-blind, placebo-controlled, crossover trial. Am. J. Clin. Nutr. 2015, 101, 914–921. [Google Scholar] [CrossRef]
  153. Serban, M.; Sahebkar, A.; Zanchetti, A.; Mikhailidis, D.; Howard, G.; Antal, D.; Andrica, F.; Ahmed, A.; Aronow, W.S.; Muntner, P.; et al. Effects of Quercetin on Blood Pressure: A Systematic Review and Meta-Analysis of Randomized Controlled Trials. J. Am. Heart Assoc. 2016, 5, e002713. [Google Scholar] [CrossRef]
  154. Larson, A.; Witman, M.; Guo, Y.; Ives, S.; Richardson, R.; Bruno, R.; Jalili, T.; Symons, J.D. Acute, quercetin-induced reductions in blood pressure in hypertensive individuals are not secondary to lower plasma angiotensin-converting enzyme activity or endothelin-1: Nitric oxide. Nutr. Res. 2012, 32, 557–564. [Google Scholar] [CrossRef]
  155. Bondonno, N.; Bondonno, C.; Rich, L.; Mas, E.; Shinde, S.; Ward, N.; Hodgson, J.M.; Croft, K.D. Acute effects of quercetin-3-O-glucoside on endothelial function and blood pressure: A randomized dose-response study. Am. J. Clin. Nutr. 2016, 104, 97–103. [Google Scholar] [CrossRef]
  156. Chekalina, N.I.; Shut, S.V.; Trybrat, T.A.; Burmak, Y.H.; Petrov, Y.Y.; Manusha, Y.I.; Kazakov, Y.M. Effect of quercetin on parameters of central hemodynamics and myocardial ischemia in patients with stable coronary heart disease. Wiadomości Lek. 2017, 70, 707–711. [Google Scholar]
  157. Kondratiuk, V.E.; Synytsia, Y.P. Effect of quercetin on the echocardiographic parameters of left ventricular diastolic function in patients with gout and essential hypertension. Wiadomości Lek. 2018, 71, 1554–1559. [Google Scholar]
  158. Bogdanski, P.; Suliburska, J.; Szulinska, M.; Stepien, M.; Pupek-Musialik, D.; Jablecka, A. Green tea extract reduces blood pressure, inflammatory biomarkers, and oxidative stress and improves parameters associated with insulin resistance in obese, hypertensive patients. Nutr. Res. 2012, 32, 421–427. [Google Scholar] [CrossRef] [PubMed]
  159. Rostami, A.; Khalili, M.; Haghighat, N.; Eghtesadi, S.; Shidfar, F.; Heidari, I.; Ebrahimpour-Koujan, S.; Eghtesadi, M. High-cocoa polyphenol-rich chocolate improves blood pressure in patients with diabetes and hypertension. ARYA Atheroscler. 2015, 11, 21–29. [Google Scholar] [PubMed]
  160. Chatree, S.; Sitticharoon, C.; Maikaew, P.; Pongwattanapakin, K.; Keadkraichaiwat, I.; Churintaraphan, M.; Sripong, C.; Sririwichitchai, R.; Tapechum, S. Epigallocatechin gallate decreases plasma triglyceride, blood pressure, and serum kisspeptin in obese human subjects. Exp. Biol. Med. (Maywood) 2021, 246, 163–176. [Google Scholar] [CrossRef] [PubMed]
  161. Ellinger, S.; Reusch, A.; Stehle, P.; Helfrich, H.-P. Epicatechin ingested via cocoa products reduces blood pressure in humans: A nonlinear regression model with a Bayesian approach. Am. J. Clin. Nutr. 2012, 95, 1365–1377. [Google Scholar] [CrossRef] [PubMed]
  162. Oyama, J.-I.; Maeda, T.; Kouzuma, K.; Ochiai, R.; Tokimitsu, I.; Higuchi, Y.; Sugano, M.; Makino, N. Green tea catechins improve human forearm endothelial dysfunction and have antiatherosclerotic effects in smokers. Circ. J. 2010, 74, 578–588. [Google Scholar] [CrossRef] [PubMed]
  163. Alañón, M.E.; Castle, S.M.; Serra, G.; Lévèques, A.; Poquet, L.; Actis-Goretta, L.; Spencer, J.P.L. Acute study of dose-dependent effects of (−)-epicatechin on vascular function in healthy male volunteers: A randomized controlled trial. Clin. Nutr. 2020, 39, 746–754. [Google Scholar] [CrossRef]
  164. Samavat, H.; Newman, A.R.; Wang, R.; Yuan, J.-M.; Wu, A.H.; Kurzer, M.S. Effects of green tea catechin extract on serum lipids in postmenopausal women: A randomized, placebo-controlled clinical trial. Am. J. Clin. Nutr. 2016, 104, 1671–1682. [Google Scholar] [CrossRef]
  165. Suzuki-Sugihara, N.; Kishimoto, Y.; Saita, E.; Taguchi, C.; Kobayashi, M.; Ichitani, M.; Ukawa, Y.; Sagesaka, Y.M.; Suzuki, E.; Kondo, K. Green tea catechins prevent low-density lipoprotein oxidation via their accumulation in low-density lipoprotein particles in humans. Nutr. Res. 2016, 36, 16–23. [Google Scholar] [CrossRef]
  166. McDermott, M.M.; Criqui, M.H.; Domanchuk, K.; Ferrucci, L.; Guralnik, J.M.; Kibbe, M.R.; Kosmac, K. Cocoa to Improve Walking Performance in Older People with Peripheral Artery Disease: The COCOA-PAD Pilot Randomized Clinical Trial. Circ. Res. 2020, 126, 589–599. [Google Scholar] [CrossRef] [PubMed]
  167. Kishimoto, Y.; Saita, E.; Taguchi, C.; Aoyama, M.; Ikegami, Y.; Ohmori, R.; Kondo, K.; Momiyama, Y. Associations between green tea consumption and coffee consumption and the prevalence of coronary artery disease. J. Nutr. Sci. Vitaminol. 2020, 66, 237–245. [Google Scholar] [CrossRef] [PubMed]
  168. Wang, Z.-M.; Zhou, B.; Wang, Y.-S.; Gong, Q.-Y.; Wang, Q.-M.; Yan, J.-J.; Gao, W.; Wang, L.-S. Black and green tea consumption and the risk of coronary artery disease: A meta-analysis. Am. J. Clin. Nutr. 2011, 93, 506–515. [Google Scholar] [CrossRef] [PubMed]
  169. Dural, I.E.; Onrat, E.; Celik, S.; Vurmaz, A.; Yalim, Z.; Gokaslan, S.; Gokaslan, C.; Korucu, C.; Aksu, U. The Relationships Between Chocolate Consumption and Endothelial Dysfunction in Patients with Heart Failure. Türk Kardiyol. Derneği Arşivi 2022, 50, 334–339. [Google Scholar] [CrossRef] [PubMed]
  170. Cassidy, A.; Mukamal, K.J.; Liu, L.; Franz, M.; Eliassen, A.H.; Rimm, E.B. High anthocyanin intake is associated with a reduced risk of myocardial infarction in young and middle-aged women. Circulation 2013, 127, 188–196. [Google Scholar] [CrossRef] [PubMed]
  171. Cassidy, A.; Bertoia, M.; Chiuve, S.; Flint, A.; Forman, J.; Rimm, E.B. Habitual intake of anthocyanins and flavanones and risk of cardiovascular disease in men. Am. J. Clin. Nutr. 2016, 104, 587–594. [Google Scholar] [CrossRef] [PubMed]
  172. Alvarez-Suarez, J.M.; Giampieri, F.; Tulipani, S.; Casoli, T. One month strawberry-rich anthocyanin supplementation ameliorates cardiovascular risk, oxidative stress markers and platelet activation in humans. J. Nutr. Biochem. 2014, 25, 289–294. [Google Scholar] [CrossRef]
  173. Aboonabi, A.; Meyer, R.R.; Gaiz, A.; Singh, I. Anthocyanins in berries exhibited anti-atherogenicity and antiplatelet activities in a metabolic syndrome population. Nutr. Res. 2020, 76, 82–93. [Google Scholar] [CrossRef]
  174. Santhakumar, A.; Kundur, A.; Fanning, K.; Netzel, M.; Stanley, R.; Singh, I. Consumption of anthocyanin-rich Queen Garnet plum juice reduces platelet activation related thrombogenesis in healthy volunteers. J. Func. Foods 2015, 12, 11–22. [Google Scholar] [CrossRef]
  175. Tian, Z.; Li, K.; Fan, D.; Zhao, Y.; Gao, X.; Ma, X.; Xu, L.; Shi, Y.; Ya, F.; Zou, J.; et al. Dose-dependent effects of anthocyanin supplementation on platelet function in subjects with dyslipidemia: A randomized clinical trial. BioMedicine 2021, 70, 103533. [Google Scholar] [CrossRef]
  176. Jennings, A.; Welch, A.A.; Fairweather-Tait, S.J.; Kay, C.; Minihane, A.; Chowienczyk, P. Higher anthocyanin intake is associated with lower arterial stiffness and central blood pressure in women. Am. J. Clin. Nutr. 2012, 96, 781–788. [Google Scholar] [CrossRef] [PubMed]
  177. McKay, D.L.; Chen, C.-Y.O.; Saltzman, E.; Blumberg, J.B. Hibiscus sabdaria L. Tea (Tisane) lowers blood pressure in prehypertensive and mildly hypertensive adults. J. Nutr. 2009, 140, 298–303. [Google Scholar] [CrossRef] [PubMed]
  178. Basu, A.; Du, M.; Leyva, M.J.; Sanchez, K.; Betts, N.M.; Wu, M. Blueberries decrease cardiovascular risk factors in obese men and women with metabolic syndrome. J. Nutr. 2010, 140, 1582–1587. [Google Scholar] [CrossRef] [PubMed]
  179. Johnson, S.A.; Figueroa, A.; Navaei, N.; Wong, A.; Kalfon, R.; Ormsbee, L.T. Daily blueberry consumption improves blood pressure and arterial stiffness in postmenopausal women with pre- and stage 1-hypertension: A randomized, double-blind, placebo-controlled clinical trial. J. Acad. Nutr. Diet. 2015, 115, 369–377. [Google Scholar] [CrossRef] [PubMed]
  180. Emamat, H.; Zahedmehr, A.; Asadian, S.; Nasrollahzadeh, J. The effect of purple-black barberry (Berberis integerrima) on blood pressure in subjects with cardiovascular risk factors: A randomized controlled trial. J. Ethnopharmacol. 2022, 289, 115097. [Google Scholar] [CrossRef] [PubMed]
  181. Dohadwala, M.M.; Holbrook, M.; Hamburg, N.M.; Shenouda, S.M.; Chung, W.B.; Titas, M.; Kluge, M.A.; Na Wang, N.; Palmisano, J.; Milbury, P.E.; et al. Effects of cranberry juice consumption on vascular function in patients with coronary artery disease. Am. J. Clin. Nutr. 2011, 93, 934–940. [Google Scholar] [CrossRef] [PubMed]
  182. Huang, H.; Chen, G.; Liao, D.; Zhu, Y.; Xue, X. Effects of berries consumption on cardiovascular risk factors: A meta-analysis with trial sequential analysis of randomized controlled trials. Sci. Rep. 2016, 6, 23625. [Google Scholar] [CrossRef] [PubMed]
  183. Yang, L.; Ling, W.; Yang, Y.; Chen, Y.; Tian, Z.; Du, Z. Role of purified anthocyanins in improving cardiometabolic risk factors in Chinese men and women with prediabetes or early untreated diabetes—A randomized controlled trial. Nutrients 2017, 9, 1104. [Google Scholar] [CrossRef]
  184. Liu, C.; Sun, J.; Lu, Y.; Bo, Y. Effects of anthocyanin on serum lipids in dyslipidemia patients: A systematic review and meta-analysis. PLoS ONE 2016, 11, e0162089. [Google Scholar] [CrossRef]
  185. Gordish, K.L.; Beierwaltes, W.H. Chronic resveratrol reverses a mild angiotensin II-induced pressor effect in a rat model. Integr. Blood Press. Control 2016, 9, 23–31. [Google Scholar]
  186. Zhang, L.-X.; Li, C.-X.; Kakar, M.U.; Khan, M.S.; Wu, P.-F.; Amir, R.M.; Dai, D.-F.; Naveed, M.; Li, Q.-Y.; Saeed, M.; et al. Resveratrol (RV): A pharmacological review and call for further research. Biomed. Pharmacother. 2021, 143, 112164. [Google Scholar] [CrossRef] [PubMed]
  187. Cheng, P.W.; Ho, W.Y.; Su, Y.T.; Lu, P.J.; Chen, B.Z.; Cheng, W.H.; Lu, W.H.; Sun, G.C.; Yeh, T.C.; Hsiao, M.; et al. Resveratrol decreases fructose-induced oxidative stress, mediated by NADPH oxidase via an AMPK-dependent mechanism. Br. J. Pharmacol. 2014, 171, 2739–2750. [Google Scholar] [CrossRef] [PubMed]
  188. Yu, L.; Tu, Y.; Jia, X.; Fang, K.; Liu, L.; Wan, L.; Xiang, C.; Wang, Y.; Sun, X.; Liu, T.; et al. Resveratrol protects against pulmonary arterial hypertension in rats via activation of silent information regulator 1. Cell Physiol. Biochem. 2017, 42, 55–67. [Google Scholar] [CrossRef]
  189. Dolinsky, V.W.; Chakrabarti, S.; Pereira, T.J.; Oka, T.; Levasseur, J.; Beker, D. Resveratrol prevents hypertension and cardiac hypertrophy in hypertensive rats and mice. Biochim. Biophys. Acta 2013, 1832, 1723–1733. [Google Scholar] [CrossRef] [PubMed]
  190. Grujić-Milanović, J.; Jaćević, V.; Miloradović, Z.; Milanović, S.D.; Jovović, D.; Ivanov, M.; Karanović, D.; Vajić, U.-J.; Mihailović-Stanojević, N. Resveratrol improved kidney function and structure in malignantly hypertensive rats by restoration of antioxidant capacity and nitric oxide bioavailability. Biomed. Pharmacother. 2022, 154, 113642. [Google Scholar] [CrossRef] [PubMed]
  191. Franco, J.G.; Lisboa, P.C.; Lima, N.S.; Amaral, T.A.; Peixoto-Silva, N.; Resende, A.C.; Oliveira, E.; Passos, M.C.; Moura, E.G. Resveratrol attenuates oxidative stress and prevents steatosis and hypertension in obese rats programmed by early weaning. J. Nutr. Biochem. 2013, 24, 960–966. [Google Scholar] [CrossRef] [PubMed]
  192. Mozafari, M.; Nekooeian, A.A.; Panjeshahin, M.R.; Zare, H.R. The effects of resveratrol in rats with simultaneous type 2 diabetes and renal hypertension: A study of antihypertensive mechanisms. Iran. J. Med. Sci. 2015, 40, 152–160. [Google Scholar] [PubMed]
  193. Tain, Y.L.; Lee, W.C.; Wu, K.L.H.; Leu, S.; Chan, J.Y.H. Resveratrol prevents the development of hypertension programmed by maternal plus post-weaning high-fructose consumption through modulation of oxidative stress, nutrient-sensing signals, and gut microbiota. Mol. Nutr. Food Res. 2018, 62, e1800066. [Google Scholar] [CrossRef]
  194. Prysyazhna, O.; Wolhuter, K.; Switzer, C.; Santos, C.; Yang, X.; Lynham, S.; Shah, A.M.; Eaton, P.; Burgoyne, J.R. Blood Pressure-Lowering by the Antioxidant Resveratrol Is Counterintuitively Mediated by Oxidation of cGMP-Dependent Protein Kinase. Circulation 2019, 140, 126–137. [Google Scholar] [CrossRef]
  195. Li, H.; Xia, N.; Hasselwander, S.; Daiber, A. Resveratrol and vascular function. Int. J. Mol. Sci. 2019, 20, 2155. [Google Scholar] [CrossRef]
  196. Lee, D.I.; Acosta, C.; Anderson, C.M.; Anderson, H.D. Peripheral and cerebral resistance arteries in the spontaneously hypertensive heart failure rat: Effects of stilbenoid polyphenols. Molecules 2017, 22, 380. [Google Scholar] [CrossRef]
  197. Li, H.; Förstermann, U. Uncoupling of endothelial NO synthase in atherosclerosis and vascular disease. Curr. Opin. Pharmacol. 2013, 13, 161–167. [Google Scholar] [CrossRef] [PubMed]
  198. Fabricio, V.; Oishi, J.C.; Biffe, B.G.; Ruffoni, L.D.; Silva, K.A.; Nonaka, K.O.; Rodrigues, G.J. Resveratrol treatment normalizes the endothelial function and blood pressure in ovariectomized rats. Arq. Bras. Cardiol. 2017, 108, 116–121. [Google Scholar] [CrossRef] [PubMed]
  199. Tsutsui, H.; Kinugawa, S.; Matsushimam, S. Oxidative stress and heart failure. Am. J. Physiol. Heart Circ. Physiol. 2011, 301, H2181–H2190. [Google Scholar] [CrossRef] [PubMed]
  200. Gu, X.; Wang, Z.; Ye, Z.; Lei, J.; Li, L.; Su, D.; Zheng, X. Resveratrol, an activator of SIRT1, upregulates AMPK and improves cardiac function in heart failure. Genet. Mol. Res. 2014, 13, 323–335. [Google Scholar] [CrossRef] [PubMed]
  201. Ahmet, I.; Tae, H.J.; Lakatta, E.G.; Talan, M. Long-term low dose dietary resveratrol supplement reduces cardiovascular structural and functional deterioration in chronic heart failure in rats. Can. J. Physiol. Pharmacol. 2017, 95, 268–274. [Google Scholar] [CrossRef] [PubMed]
  202. Kanamori, H.; Takemura, G.; Goto, K.; Tsujimoto, A.; Ogino, A.; Takeyama, T.; Kawaguchi, T.; Watanabe, T.; Morishita, K.; Kawasaki, M.; et al. Resveratrol reverses remodeling in hearts with large, old myocardial infarctions through enhanced autophagy-activating AMP kinase pathway. Am. J. Pathol. 2013, 182, 701–713. [Google Scholar] [CrossRef] [PubMed]
  203. Matsumura, N.; Takahara, S.; Maayah, Z.H.; Parajuli, N.; Byrne, N.J.; Shoieb, S.M.; Soltys, C.M.; Beker, D.L.; Masson, G.; El-Kadi, A.O.S.; et al. Resveratrol improves cardiac function and exercise performance in MI-induced heart failure through the inhibition of cardiotoxic HETE metabolites. J. Mol. Cell. Cardiol. 2018, 125, 162–173. [Google Scholar] [CrossRef] [PubMed]
  204. Wojciechowski, P.; Juric, D.; Louis, X.L.; Thandapilly, S.J.; Yu, L.; Taylor, C.; Netticadan, T. Resveratrol arrests and regresses the development of pressure overload- but not volume overload-induced cardiac hypertrophy in rats. J. Nutr. 2010, 140, 962–968. [Google Scholar] [CrossRef]
  205. 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, e26391. [Google Scholar] [CrossRef]
  206. Sung, M.M.; Das, S.K.; Levasseur, J.; Byrne, N.J.; Fung, D.; Kim, T.T.; Masson, G.; Boisvenue, J.; Soltys, C.-L. Resveratrol treatment of mice with pressure-overload-induced heart failure improves diastolic function and cardiac energy metabolism. Circ. Heart Fail. 2015, 8, 128–137. [Google Scholar] [CrossRef] [PubMed]
  207. Sung, M.M.; Byrne, N.J.; Robertson, I.M.; Kim, T.T.; Samokhvalov, V.; Levasseur, J.; Soltys, C.-L.; Fung, D.; Tyreman, N.; Denou, E.; et al. Resveratrol improves exercise performance and skeletal muscle oxidative capacity in heart failure. Am. J. Physiol. 2017, 312, H842–H853. [Google Scholar] [CrossRef] [PubMed]
  208. Feng, H.; Mou, S.-Q.; Li, W.-J.; Zhang, N.; Zhou, Z.-Y.; Ding, W.; Bian, Z.-Y.; Liao, H.-H. Resveratrol Inhibits Ischemia-Induced Myocardial Senescence Signals and NLRP3 Inflammasome Activation. Oxid. Med. Cell. Longev. 2020, 2020, 2647807. [Google Scholar] [CrossRef] [PubMed]
  209. Fourny, N.; Lan, C.; Eric, S.; Bernard, M.; Desrois, M. Protective effect of resveratrol against ischemia-reperfusion injury via enhanced high energy compounds and eNOS-SIRT1 expression in type 2 diabetic female rat heart. Nutrients 2019, 11, 105. [Google Scholar] [CrossRef] [PubMed]
  210. Mao, Z.-J.; Lin, H.; Hou, J.-W.; Zhou, Q.; Wang, Q.; Chen, Y.-H. A meta-analysis of resveratrol protects against myocardial ischemia/reperfusion injury: Evidence from small animal studies and insight into molecular mechanisms. Oxid. Med. Cell. Longev. 2019, 2019, 5793867. [Google Scholar] [CrossRef] [PubMed]
  211. Gurusamy, N.; Lekli, I.; Mukherjee, S.; Ray, D.; Ahsan, M.K.; Gherghiceanu, M.; Popescu, L.M.; Das, D.K. Cardioprotection by resveratrol: A novel mechanism via autophagy involving the mTORC2 pathway. Cardiovasc. Res. 2010, 86, 103–112. [Google Scholar] [CrossRef] [PubMed]
  212. Gupta, P.K.; DiPette, D.J.; Supowit, S.C. Protective effect of resveratrol against pressure overload-induced heart failure. Food Sci. Nutr. 2014, 2, 218–229. [Google Scholar] [CrossRef] [PubMed]
  213. Gao, Y.; Kang, L.; Li, C.; Liu, R.; Wang, J. Resveratrol ameliorates diabetes-induced cardiac dysfunction through AT1R-ERK/p38 MAPK signaling pathway. Cardiovasc. Toxicol. 2016, 16, 130–137. [Google Scholar] [CrossRef]
  214. Dolinsky, V.W.; Rogan, K.J.; Sung, M.M.; Zordoky, B.N.; Haykowsky, M.J.; Young, M.E.; Jones, L.W.; Dyck, J.R.B. Both aerobic exercise and resveratrol supplementation attenuate doxorubicin-induced cardiac injury in mice. Am. J. Physiol. Endocrinol. Metab. 2013, 305, E243–E253. [Google Scholar] [CrossRef]
  215. Zeng, Y.; Cao, G.; Lin, L.; Zhang, Y.; Luo, X.; Ma, X.; Aiyisake, A.; Cheng, Q. Resveratrol Attenuates Sepsis-Induced Cardiomyopathy in Rats through Anti-Ferroptosis via the Sirt1/Nrf2 Pathway. J. Investig. Surg. 2023, 36, 2157521. [Google Scholar] [CrossRef]
  216. Yoshida, Y.; Shioi, T.; Izumi, T. Resveratrol ameliorates experimental autoimmune myocarditis. Circulation 2007, 71, 397–404. [Google Scholar] [CrossRef]
  217. Tanno, M.; Kuno, A.; Yano, T.; Miura, T.; Hisahara, S.; Ishikawa, S.; Shimamoto, K.; Horio, Y. Induction of manganese superoxide dismutase by nuclear translocation and activation of SIRT1 promotes cell survival in chronic heart failure. J. Biol. Chem. 2010, 285, 8375–8382. [Google Scholar] [CrossRef] [PubMed]
  218. Pan, Q.R.; Ren, Y.L.; Liu, W.X.; Hu, Y.J.; Zheng, J.S.; Xu, Y.; Wang, G. Resveratrol prevents hepatic steatosis and endoplasmic reticulum stress and regulates the expression of genes involved in lipid metabolism, insulin resistance, and inflammation in rats. Nutr. Res. 2015, 35, 576–584. [Google Scholar] [CrossRef] [PubMed]
  219. Chen, Q.; Wang, E.; Ma, L.; Zhai, P. Dietary resveratrol increases the expression of hepatic 7α-hydroxylase and ameliorates hypercholesterolemia in high-fat fed C57BL/6J mice. Lipids Health Dis. 2012, 11, 56. [Google Scholar] [CrossRef] [PubMed]
  220. Tanko, Y.; Jimoh, A.; Ahmed, A.; Mohammed, A.; Ayo, J.O. Resveratrol Protects Rabbits Against Cholesterol Diet-Induced Hyperlipidaemia. Niger. J. Physiol. Sci. 2016, 31, 71–75. [Google Scholar] [PubMed]
  221. Singh, A.P.; Singh, R.; Verma, S.S.; Rai, V.; Kaschula, C.H.; Maiti, P.; Gupta, S.C. Health benefits of resveratrol: Evidence from clinical studies. Med. Res. Rev. 2019, 39, 1851–1891. [Google Scholar] [CrossRef] [PubMed]
  222. Zhao, W.; Li, A.; Feng, X.; Hou, T.; Liu, K.; Liu, B.; Zhang, N. Metformin and resveratrol ameliorate muscle insulin resistance through preventing lipolysis and inflammation in hypoxic adipose tissue. Cell Signal. 2016, 28, 1401–1411. [Google Scholar] [CrossRef] [PubMed]
  223. Berrougui, H.; Grenier, G.; Loued, S.; Drouin, G.; Khalil, A. A new insight into resveratrol as an atheroprotective compound: Inhibition of lipid peroxidation and enhancement of cholesterol efflux. Atherosclerosis 2009, 207, 420–427. [Google Scholar] [CrossRef]
  224. Liu, Y.; Chen, X.; Li, J. Resveratrol protects against oxidized low-density lipoprotein-induced human umbilical vein endothelial cell apoptosis via inhibition of mitochondrial-derived oxidative stress. Mol. Med. Rep. 2017, 15, 2457–2464. [Google Scholar] [CrossRef]
  225. Deng, Z.Y.; Hu, M.M.; Xin, Y.F.; Gang, C. Resveratrol alleviates vascular inflammatory injury by inhibiting inflammasome activation in rats with hypercholesterolemia and vitamin D2 treatment. Inflamm Res. 2015, 64, 321–332. [Google Scholar] [CrossRef]
  226. Brawerman, G.M.; Kereliuk, S.M.; Brar, N.; Cole, L.K.; Seshadri, N.; Pereira, T.J.; Xiang, B.; Hunt, K.L.; Fonseca, M.A.; Hatch, G.M.; et al. Maternal resveratrol administration protects against gestational diabetes-induced glucose intolerance and islet dysfunction in the rat on spring. J. Physiol. 2019, 597, 4175–4192. [Google Scholar] [CrossRef] [PubMed]
  227. Heebøll, S.; Kreuzfeldt, M.; Hamilton-Dutoit, S.; Kjær Poulsen, M.; Stødkilde-Jørgensen, H.; Møller, H.J.; Jessen, N.; Thorsen, K.; Kristina Hellberg, Y.; Bønløkke Pedersen, S.; et al. Placebo-controlled, randomised clinical trial: High-dose resveratrol treatment for non-alcoholic fatty liver disease. Scand. J. Gastroenterol. 2016, 51, 456–464. [Google Scholar] [CrossRef] [PubMed]
  228. Fujitaka, K.; Otani, H.; Jo, F.; Jo, H.; Nomura, E.; Iwasaki, M.; Nishikawa, M.; Iwasaka, T.; Das, D.K. Modified resveratrol Longevinex improves endothelial function in adults with metabolic syndrome receiving standard treatment. Nutr. Res. 2011, 31, 842–847. [Google Scholar] [CrossRef] [PubMed]
  229. Agarwal, B.; Campen, M.J.; Channell, M.M.; Wherry, S.J.; Varamini, B.; Davis, J.G.; Baur, J.A.; Smoliga, J.A. Resveratrol for primary prevention of atherosclerosis: Clinical trial evidence for improved gene expression in vascular endothelium. Int. J. Cardiol. 2013, 166, 246–248. [Google Scholar] [CrossRef] [PubMed]
  230. Militaru, C.; Donoiu, I.; Craciun, A.; Scorei, I.D.; Bulearca, A.M.; Scorei, R.I. Oral resveratrol and calcium fructoborate supplementation in subjects with stable angina pectoris: Effects on lipid profiles, inflammation markers, and quality of life. Nutrition 2013, 29, 178–183. [Google Scholar] [CrossRef] [PubMed]
  231. Guven, H.; Arici, A.; Simsek, O. Flavonoids in our foods: A short review. J. Basic Clin. Health Sci. 2019, 3, 96–106. [Google Scholar] [CrossRef]
  232. Graf, B.A.; Milbury, P.E.; Blumberg, J.B. Flavonols, flavones, flavanones, and human health: Epidemiological evidence. J. Med. Food. 2005, 8, 281–290. [Google Scholar] [CrossRef]
  233. Gentile, D.; Fornai, M.; Pellegrini, C.; Colucci, R.; Blandizzi, C. Dietary flavonoids as a potential intervention to improve redox balance in obesity and related co-morbidities: A review. Nutr. Res. Rev. 2018, 31, 239–247. [Google Scholar] [CrossRef]
  234. Ginwala, R.; Bhavsar, R.; Chigbu, D.G.I.; Jain, P.; Khan, Z.K. Potential role of flavonoids in treating chronic inflammatory diseases with a special focus on the anti-inflammatory activity of apigenin. Antioxidants 2019, 8, 35. [Google Scholar] [CrossRef]
  235. Biela, M.; Rimarčík, J.; Senajová, E.; Kleinová, A.; Klein, E. Antioxidant action of deprotonated flavonoids: Thermodynamics of sequential proton-loss electron-transfer. Phytochemistry 2020, 180, 112528. [Google Scholar] [CrossRef]
  236. Dias, M.C.; Pinto, D.C.G.A.; Freitas, H.; Santos, C.; Silva, A.M.S. The antioxidant system in Olea europaea to enhanced UV-B radiation also depends on flavonoids and secoiridoids. Phytochemistry 2020, 170, 112199. [Google Scholar] [CrossRef] [PubMed]
  237. D’Amelia, V.; Aversano, R.; Chiaiese, P.; Carputo, D. The antioxidant properties of plant flavonoids: Their exploitation by molecular plant breeding. Phytochem. Rev. 2018, 17, 611–625. [Google Scholar] [CrossRef]
  238. Hamid, A.A.; Aminuddin, A.; Yunus, M.H.M.; Murthy, J.K.; Hui, C.K. Antioxidative and anti-inflammatory activities of Polygonum minus: A review of literature. Rev. Cardiovasc. Med. 2020, 21, 275–287. [Google Scholar] [PubMed]
  239. Zeng, N.; Zhang, G.; Hu, X.; Pan, J.; Gong, D. Mechanism of fisetin suppressing superoxide anion and xanthine oxidase activity. J. Funct. Foods 2019, 58, 1–10. [Google Scholar] [CrossRef]
  240. Arts, I.C.W.; Sesink, A.L.A.; Faassen-Peters, M.; Hollman, P.C.H. The type of sugar moiety is a major determinant of the small intestinal uptake and subsequent biliary excretion of dietary quercetin glycosides. Br. J. Nutr. 2004, 91, 841–847. [Google Scholar] [CrossRef] [PubMed]
  241. Kumar, S.; Pandey, A.K. Chemistry and biological activities of flavonoids: An Overview. Sci. World J. 2013, 2013, 162750. [Google Scholar] [CrossRef] [PubMed]
  242. Ahmed, N.A.; Radwan, N.M.; Aboul Ezz, H.S.; Salama, N.A. The antioxidant effect of green tea mega EGCG against electromagnetic radiation-induced oxidative stress in the hippocampus and striatum of rats. Electromagn. Biol. Med. 2017, 36, 63–73. [Google Scholar] [CrossRef]
  243. Zhang, H.; Xu, Z.; Zhao, H.; Wang, X.; Pang, J. Anthocyanin supplementation improves anti-oxidative and anti-inflammatory capacity in a dose-response manner in subjects with dyslipidemia. Redox Biol. 2020, 32, 101474. [Google Scholar] [CrossRef]
  244. Nagarajan, S.; Nagarajan, R.; Kumar, J.; Salemme, A.; Togna, A.R.; Saso, L.; Bruno, F. Antioxidant activity of synthetic polymers of phenolic compounds. Polymers 2020, 12, 1646. [Google Scholar] [CrossRef]
  245. Yin, Z.; Wu, Y.; Chen, Y.; Qie, X.; Zeng, M.; Wang, Z.; Qin, F.; Chen, J.; He, Z. Analysis of the interaction between cyanidin-3-O-glucoside and casein hydrolysates and its effect on the antioxidant ability of the complexes. Food Chem. 2021, 340, 127915. [Google Scholar] [CrossRef]
  246. Read, M. Flavonoids: Naturally occurring anti-inflammatory agents. Am. J. Pathol. 1995, 147, 235–237. [Google Scholar] [PubMed]
  247. Damon, M.; Flandre, O.; Michel, F.; Perdrix, L.; Labrid, C. Effect of chronic treatment with a purified flavonoid fraction on inflammatory granuloma in the rat. Study of prostaglandin E2 and F2 alpha and thromboxane B2 release and histological changes. Arzneimittel-Forschung 1987, 37, 1149–1153. [Google Scholar] [PubMed]
  248. Lee, H.S.; Ko, H.R.; Ryu, S.Y.; Oh, W.K.; Kim, B.Y. Inhibition of phospholipase C Gamma 1 by the prenylated flavonoids from Sophora flavescens. Planta Med. 1997, 63, 266–268. [Google Scholar] [CrossRef] [PubMed]
  249. Kimata, M.; Shichijo, M.; Miura, T.; Serizawa, I.; Inagaki, N.; Nagai, H. Effects of luteolin, quercetin and baicalein on immunoglobulin E-mediated mediator release from human cultured mast cells. Clin. Exp. Allergy 2000, 30, 501–508. [Google Scholar] [CrossRef] [PubMed]
  250. Al-Khayri, J.M.; Sahana, G.R.; Nagella, P.; Joseph, B.V.; Alessa, F.M. Flavonoids as Potential Anti-Inflammatory Molecules: A Review. Molecules 2022, 27, 2901. [Google Scholar] [CrossRef] [PubMed]
  251. Serafini, M.; Peluso, I.; Raguzzini, A. Flavonoids as Anti-Inflammatory Agents. Proc. Nutr. Soc. 2010, 69, 273–278. [Google Scholar] [CrossRef] [PubMed]
  252. Nam, T.G.; Lim, T.-G.; Lee, B.H.; Lim, S.; Kang, H. Comparison of Anti-Inflammatory Effects of Flavonoid-Rich Common and Tartary Buckwheat Sprout Extracts in Lipopolysaccharide-Stimulated RAW264.7 and Peritoneal Macrophages. Oxid. Med. Cell. Longev. 2017, 2017, 9658030. [Google Scholar] [CrossRef]
  253. Ullah, A.; Munir, S.; Badshah, S.L.; Khan, N.; Ghani, L. Important Flavonoids and Their Role as a Therapeutic Agent. Molecules 2020, 25, 5243. [Google Scholar] [CrossRef]
  254. Li, C.; Wang, T.; Zhang, C.; Xuan, J.; Su, C.; Wang, Y. Quercetin attenuates cardiomyocyte apoptosis via inhibition of JNK and p38 mitogen-activated protein kinase signaling pathways. Gene 2016, 577, 275–280. [Google Scholar] [CrossRef]
  255. Tang, L.; Peng, Y.; Xu, T.; Yi, X.; Liu, Y. The effects of quercetin protect cardiomyocytes from A/R injury is related to its capability to increasing expression and activity of PKCε protein. Mol. Cell. Biochem. 2013, 382, 145–152. [Google Scholar] [CrossRef]
  256. Shu, Z.; Yang, Y.; Yang, L.; Jiang, H.; Yu, X. Cardioprotective effects of dihydroquercetin against ischemia reperfusion injury by inhibiting oxidative stress and endoplasmic reticulum stress-induced apoptosis via the PI3K/Akt pathway. Food Funct. 2019, 10, 203–215. [Google Scholar] [CrossRef]
  257. Rahman, I.; Biswas, S.; Kirkham, P. Regulation of Inflammation and Redox Signaling by Dietary Polyphenols. Biochem. Pharmacol. 2006, 72, 1439–1452. [Google Scholar] [CrossRef] [PubMed]
  258. Rahman, I.; Marwick, J.; Kirkham, P. Redox Modulation of Chromatin Remodeling: Impact on Histone Acetylation and Deacetylation, NF-KappaB and pro-Inflammatory Gene Expression. Biochem. Pharmacol. 2004, 68, 1255–1267. [Google Scholar] [CrossRef] [PubMed]
  259. Min, Y.; Choi, C.; Bark, H.; Son, H.; Park, H. Quercetin Inhibits Expression of Inflammatory Cytokines through Attenuation of NFkappaB and P38 MAPK in HMC-1 Human Mast Cell Line. Inflamm. Res. 2007, 56, 210–215. [Google Scholar] [CrossRef] [PubMed]
  260. Chen, T.; Zhang, X.; Zhu, G.; Liu, H.; Chen, J.; Wang, Y.; He, X. Quercetin inhibits TNF-α induced HUVECs apoptosis and inflammation via downregulating NF-kB and AP-1 signaling pathway in vitro. Medicine 2020, 99, e22241. [Google Scholar] [CrossRef] [PubMed]
  261. Kundu, J.K.; Surh, Y.J. Epigallocatechin Gallate Inhibits Phorbol Ester-Induced Activation of NF-KB and CREB in Mouse Skin Role of P38 MAPK. Ann. N. Y. Acad. Sci. 2007, 1095, 504–512. [Google Scholar] [CrossRef] [PubMed]
  262. Cho, S.; Park, S.; Kwon, M.; Jeong, T.; Bok, S.; Choi, W.Y.; Jeong, W.I.; Ryu, S.Y. Quercetin Suppresses Proinflammatory Cytokines Production through MAP Kinases And NF-Kappa B Pathway in Lipopolysaccharide-Stimulated Macrophage. Mol. Cell. Biochem. 2003, 243, 153–160. [Google Scholar] [CrossRef]
  263. Yahfoufi, N.; Alsadi, N.; Jambi, M.; Matar, C. The Immunomodulatory and Anti-Inflammatory Role of Polyphenols. Nutrients 2018, 10, 1618. [Google Scholar] [CrossRef]
  264. Siti, H.N.; Jalil, J.; Asmadi, A.Y.; Kamisah, Y. Rutin Modulates MAPK Pathway Differently from Quercetin in Angiotensin II-Induced H9c2 Cardiomyocyte Hypertrophy. Int. J. Mol. Sci. 2021, 22, 5063. [Google Scholar] [CrossRef]
  265. Andugulapati, S.B.; Gourishetti, K.; Tirunavalli, S.K.; Shaikh, T.B.; Sistla, R. Biochanin-A ameliorates pulmonary fibrosis by suppressing the TGF-β mediated EMT, myofibroblasts differentiation and collagen deposition in in vitro and in vivo systems. Phytomedicine 2020, 78, 153298. [Google Scholar] [CrossRef]
  266. Wang, J.; Ge, S.; Wang, Y.; Liu, Y.; Qiu, L.; Li, J. Puerarin Alleviates UUO-Induced Inflammation and Fibrosis by Regulating the NF-κB P65/STAT3 and TGFβ1/Smads Signaling Pathways. Drug Des. Dev. Ther. 2021, 15, 3697–3708. [Google Scholar] [CrossRef] [PubMed]
  267. Meng, J.; Chen, Y.; Wang, J.; Qiu, J.; Chang, C.; Bi, F.; Wu, X.; Liu, W. EGCG protects vascular endothelial cells from oxidative stress-induced damage by targeting the autophagy-dependent PI3K-AKT-mTOR pathway. Ann. Transl. Med. 2020, 8, 200. [Google Scholar] [CrossRef] [PubMed]
  268. Zhang, Z.; Zhang, D. (−)-Epigallocatechin-3-gallate inhibits eNOS uncoupling and alleviates high glucose-induced dysfunction and apoptosis of human umbilical vein endothelial cells by PI3K/AKT/eNOS pathway. Diabetes Metab. Syndr. Obes. 2020, 13, 2495–2504. [Google Scholar] [CrossRef] [PubMed]
  269. Zhang, C.; Liao, P.; Liang, R.; Zheng, X.; Jian, J. Epigallocatechin gallate prevents mitochondrial impairment and cell apoptosis by regulating miR-30a/p53 axis. Phytomedicine 2019, 61, 152845. [Google Scholar] [CrossRef] [PubMed]
  270. Mou, Q.; Jia, Z.; Luo, M.; Liu, L.; Huang, X. Epigallocatechin-3-gallate exerts cardioprotective effects related to energy metabolism in pressure overload-induced cardiac dysfunction. Arch. Biochem. Biophys. 2022, 723, 109217. [Google Scholar] [CrossRef]
  271. Nan, J.; Nan, C.; Ye, J.; Qian, L.; Geng, Y. EGCG protects cardiomyocytes against hypoxia-reperfusion injury through inhibition of OMA1 activation. J. Cell Sci. 2019, 132, jcs220871. [Google Scholar] [PubMed]
  272. Larson, A.J.; Symons, J.D.; Jalili, T. Therapeutic potential of quercetin to decrease blood pressure: Review of efficacy and mechanisms. Adv. Nutr. 2012, 3, 39–46. [Google Scholar] [CrossRef] [PubMed]
  273. Igarashi, K.; Honma, K.; Yoshinari, O.; Nanjo, F.; Hara, Y. Effects of dietary catechins on glucose tolerance, blood pressure and oxidative status in Goto-Kakizaki rats. J. Nutr. Sci. Vitaminol. 2007, 53, 496–500. [Google Scholar] [CrossRef]
  274. Huang, Y.; Zhang, A.; Lau, C.W.; Chen, Z.Y. Vasorelaxant effects of purified green tea epicatechin derivatives in rat mesenteric artery. Life Sci. 1998, 63, 275–283. [Google Scholar] [CrossRef]
  275. Liu, Y.; Niu, L.; Cui, L.; Hou, X.; Li, J. Hesperetin inhibits rat coronary constriction by inhibiting Ca2+ influx and enhancing voltage-gated K+ channel currents of the myocytes. Eur. J. Pharmacol. 2014, 735, 193–201. [Google Scholar] [CrossRef]
  276. Arumugam, S.; Thandavarayan, R.A.; Arozal, W.; Sari, F.R.; Giridharan, V.V.; Soetikno, V.; Palaniyandi, S.S.; Harima, M. Quercetin offers cardioprotection against progression of experimental autoimmune myocarditis by suppression of oxidative and endoplasmic reticulum stress via endothelin-1/MAPK signaling. Free Radic. Res. 2012, 46, 154–163. [Google Scholar] [CrossRef] [PubMed]
  277. Shahbaz, A.U.; Kamalov, G.; Zhao, W.; Zhao, T.; Johnson, P.L. Mitochondria-targeted Cardioprotection in Aldosteronism. J Cardiovasc. Pharmacol. 2011, 57, 37–43. [Google Scholar] [CrossRef] [PubMed]
  278. Kumar, M.; Kasala, E.R.; Bodduluru, L.N.; Kumar, V.; Lahkar, M. Molecular and biochemical evidence on the protective effects of quercetin in isoproterenol-induced acute myocardial injury in rats. J. Biochem. Mol. Toxicol. 2017, 31, e21832. [Google Scholar] [CrossRef] [PubMed]
  279. Putakala, M.; Gujjala, S.; Nukala, S.; Bongu, S.B.R.; Chintakunta, N.; Desireddy, S. Cardioprotective effect of Phyllanthus amarus against high fructose diet induced myocardial and aortic stress in rat model. Biomed. Pharmacother. 2017, 95, 1359–1368. [Google Scholar] [CrossRef] [PubMed]
  280. Liu, X.; Yu, Z.; Huang, X.; Gao, Y.; Wang, X.; Gu, J.; Xue, X. Peroxisome proliferator-activated receptor γ (PPARγ) mediates the protective effect of quercetin against myocardial ischemia-reperfusion injury via suppressing the NF-κB pathway. Am. J. Transl. Res. 2016, 8, 5169. [Google Scholar] [PubMed]
  281. Han, J.J.; Hao, J.; Kim, C.H.; Hong, J.S.; Ahn, H.Y. Quercetin prevents cardiac hypertrophy induced by pressure overload in rats. J. Vet. Med. Sci. 2009, 71, 737–743. [Google Scholar] [CrossRef] [PubMed]
  282. Wang, L.; Tan, A.; An, X.; Xia, Y.; Xie, Y. Quercetin dihydrate inhibition of cardiac fibrosis induced by angiotensin II in vivo and in vitro. Biomed. Pharmacother. 2020, 127, 110205. [Google Scholar] [CrossRef]
  283. Chen, K.; Rekep, M.; Wei, W.; Wu, Q.; Xue, Q.; Li, S.; Tian, J.; Yi, Q.; Zhang, G.; Zhang, G.; et al. Quercetin prevents in vivo and in vitro myocardial hypertrophy through the proteasome-GSK-3 pathway. Cardiovasc. Drugs Ther. 2018, 32, 5–21. [Google Scholar] [CrossRef]
  284. Chang, X.; Zhang, T.; Wang, J.; Liu, Y.; Yan, P.; Meng, Q.; Yin, Y.; Wang, S. SIRT5-related Desuccinylation modification contributes to quercetin-induced protection against heart failure and high-glucose-prompted cardiomyocytes injured through regulation of mitochondrial quality surveillance. Oxid. Med. Cell. Longev. 2021, 2021, 5876841. [Google Scholar] [CrossRef]
  285. Machha, A.; Achike, F.I.; Mustafa, A.M.; Mustafa, M.R. Quercetin, a flavonoid antioxidant, modulates endothelium-derived nitric oxide bioavailability in diabetic rat aortas. Nitric Oxide 2007, 16, 442–447. [Google Scholar] [CrossRef]
  286. Sánchez, M.; Galisteo, M.; Vera, R.; Villar, I.C.; Zarzuelo, A.; Tamargo, J.; Pérez-Vizcaíno, F.; Duarte, J. Quercetin downregulates NADPI I oxidase, increases eNOS activity and prevents endothelial dysfunction in spontaneously hypertensive rats. J. Hypertens. 2006, 24, 75–84. [Google Scholar] [CrossRef]
  287. Hackl, L.P.; Cuttle, G.; Dovichi, S.S.; Lima-Landman, M.T.; Nicolau, M. Inhibition of angiotensin-converting enzyme by quercetin alters the vascular response to bradykinin and angiotensin. J. Pharmacol. 2002, 65, 182–186. [Google Scholar] [CrossRef] [PubMed]
  288. Linz, W.; Wiemer, G.; Gohike, P.; Unger, T.; Scholkens, B.A. Contribution of kinins to the cardiovascular actions of angiotensin-converting enzyme inhibitors. Pharmacol Rev. 1995, 47, 25–49. [Google Scholar]
  289. Soman, S.; Rajamanickam, C.; Rauf, A.A.; Madambath, I. Molecular mechanisms of the antiglycative and cardioprotective activities of Psidium guajava leaves in the rat diabetic myocardium. Pharm. Biol. 2016, 54, 3078–3085. [Google Scholar] [CrossRef]
  290. Roslan, J.; Giribabu, N.; Karim, K.; Salleh, N. Quercetin ameliorates oxidative stress, inflammation, and apoptosis in the heart of streptozotocin-nicotinamide-induced adult male diabetic rats. Biomed. Pharmacother. 2017, 86, 570–582. [Google Scholar] [CrossRef] [PubMed]
  291. Najafi, M.; Noroozi, E.; Javadi, A.; Badalzadeh, R. Anti-arrhythmogenic and anti-inflammatory effects of troxerutin in ischemia/reperfusion injury of diabetic myocardium. Biomed. Pharmacother. 2018, 102, 385–391. [Google Scholar] [CrossRef] [PubMed]
  292. Ulasova, E.; Perez, J.; Hill, B.G.; Bradley, W.E.; Garber, D.W.; Landar, A.; Barnes, S.; Prasain, J.; Parks, D.A.; Dell’Italia, L.J.; et al. Quercetin prevents left ventricular hypertrophy in the Apo E knockout mouse. Redox Biol. 2013, 1, 381–386. [Google Scholar] [CrossRef]
  293. Zhang, J.; Liang, R.; Wang, L.; Yan, R.; Hou, R.; Gao, S.; Yang, B. Effects of an aqueous extract of Crataegus pinnatifida Bge. var. major N.E.Br. fruit on experimental atherosclerosis in rats. J. Ethnopharmacol. 2013, 148, 563–569. [Google Scholar]
  294. Garcia, M.L.; Pontes, R.B.; Nishi, E.E.; Ibuki, F.K.; Oliveira, V.; Sawaya, A.C.; Carvalho, P.O. The antioxidant effects of green tea reduces blood pressure and sympathoexcitation in an experimental model of hypertension. J. Hypertens. 2017, 35, 348–354. [Google Scholar] [CrossRef]
  295. Huang, Y.; Chan, N.W.; Lau, C.W.; Yao, X.Q.; Chan, F.L.; Chen, Z.Y. Involvement of endothelium/nitric oxide in vasorelaxation induced by purified green tea (-) epicatechin. Biochim. Biophys. Acta 1999, 1427, 322–328. [Google Scholar] [CrossRef]
  296. Muñoz, M.d.l.F.; Fernández, M.d.l.F.; Román-Carmena, M.; Cruz, M.d.C.I.d.l.; Amor, S.; Martorell, P.; Enrique-López, M.; García-Villalón, A.L.; Inarejos-García, A.M.; Granado, M. Supplementation with two new standardized tea extracts prevents the development of hypertension in mice with metabolic syndrome. Antioxidants 2022, 11, 1573. [Google Scholar] [CrossRef] [PubMed]
  297. Gao, J.; Akbari, A.; Wang, T. Green tea could improve elderly hypertension by modulating arterial stiffness, the activity of the renin/angiotensin/aldosterone axis, and the sodium-potassium pumps in old male rats. J. Food Biochem. 2022, 46, e14398. [Google Scholar] [CrossRef] [PubMed]
  298. Sabri, N.A.M.; Lee, S.-K.; Murugan, D.D.; Ling, W.C. Epigallocatechin gallate (EGCG) alleviates vascular dysfunction in angiotensin II-infused hypertensive mice by modulating oxidative stress and eNOS. Sci. Rep. 2022, 12, 17633. [Google Scholar] [CrossRef] [PubMed]
  299. Liu, X.X.; Lin, Q.L.; Shi, Z.P. Effect of catechin and epicatechin on serum lipid level in mice. J. Hunan Agric. Univ. 2002, 28, 232–233. [Google Scholar]
  300. Ding, S.; Jiang, J.; Yu, P.; Zhang, G.; Zhang, G.; Liu, X. Green tea polyphenol treatment attenuates atherosclerosis in high-fat diet-fed apolipoprotein E-knockout mice via alleviating dyslipidemia and up-regulating autophagy. PLoS ONE 2017, 12, e0181666. [Google Scholar] [CrossRef] [PubMed]
  301. Wahyudi, S.; Sargowo, D. Green tea polyphenols inhibit oxidized LDL-induced NF-KB activation in human umbilical vein endothelial cells. Acta Med. Indones. 2007, 39, 66–70. [Google Scholar] [PubMed]
  302. Liao, Z.-L.; Zeng, B.-H.; Wang, W.; Li, G.-H.; Wu, F.; Wang, L.; Zhong, Q.-P.; Wei, H.; Fang, X. Impact of the consumption of tea polyphenols on early atherosclerotic lesion formation and intestinal Bifidobacteria in high-fat-fed ApoE−/− mice. Front Nutr. 2016, 3, 42. [Google Scholar] [CrossRef] [PubMed]
  303. Yu, J.; Li, W.; Xiao, X.; Huang, Q.; Yu, J.; Yang, Y.; Han, T.; Zhang, D.; Niu, X. (−)-Epicatechin gallate blocks the development of atherosclerosis by regulating oxidative stress in vivo and in vitro. Food Funct. 2021, 12, 8715–8727. [Google Scholar] [CrossRef]
  304. Ferenczyová, K.; Kindernay, L.; Vlkovicová, J.; Kalocayová, B.; Rajtík, T.; Barteková, M. Pharmacology of Catechins in Ischemia-Reperfusion Injury of the Heart. Antioxidants 2021, 10, 1390. [Google Scholar] [CrossRef]
  305. Muhammed, I.; Sankar, S.; Govindaraj, S. Ameliorative effect of epigallocatechin gallate on cardiac hypertrophy and fibrosis in aged rats. J. Cardiovasc. Pharmacol. 2018, 71, 65–75. [Google Scholar] [CrossRef]
  306. Sheng, R.; Gu, Z.L.; Xie, M.L. Epigallocatechin gallate, the major component of polyphenols in green tea, inhibits telomere attrition mediated cardiomyocyte apoptosis in cardiac hypertrophy. Int. J. Cardiol. 2013, 162, 199–209. [Google Scholar] [CrossRef] [PubMed]
  307. Shen, C.L.; Samathanam, C.; Tatum, O.L.; Graham, S.; Tubb, C.; Cao, J.J.; Dunn, D.M.; Wang, J.S. Green tea polyphenols avert chronic inflammation-induced myocardial fibrosis of female rats. Inflamm. Res. 2011, 60, 665–672. [Google Scholar] [CrossRef] [PubMed]
  308. Cai, Y.; Yu, S.S.; He, Y.; Bi, X.Y.; Gao, S.; Yan, T.D.; Zheng, G.D.; Chen, T.T.; Ye, J.T.; Liu, P.Q. EGCG inhibits pressure overload-induced cardiac hypertrophy via the PSMB5/Nmnat2/SIRT6-dependent signalling pathways. Acta Physiol. 2021, 231, e13602. [Google Scholar] [CrossRef] [PubMed]
  309. Ried, K.; Fakler, P.; Stocks, N.P.; Cochrane Hypertension Group. Effect of cocoa on blood pressure. Cochrane Database Syst. Rev. 2017, 2017, CD008893. [Google Scholar] [CrossRef] [PubMed]
  310. Rodriguez-Mateos, A.; Weber, T.; Skene, S.S.; Ottaviani, J.I.; Crozier, A.; Kelm, M. Assessing the respective contributions of dietary flavanol monomers and procyanidins in mediating cardiovascular effects in humans: Randomized, controlled, double-masked intervention trial. Am. J. Clin. Nutr. 2018, 108, 1229–1237. [Google Scholar] [CrossRef] [PubMed]
  311. Sun, Y.; Zimmermann, D.; De Castro, C.A.; Actis-Goretta, L. Dose-response relationship between cocoa flavanols and human endothelial function: A systematic review and meta-analysis of randomized trials. Food Funct. 2019, 10, 6322–6330. [Google Scholar] [CrossRef] [PubMed]
  312. Huang, S.; Li, J.; Wu, Y.; Ranjbar, S.; Xing, A.; Zhao, H.; Wang, Y.; Shearer, G.C.; Bao, L.; Lichtenstein, A.H.; et al. Tea consumption and longitudinal change in high-density lipoprotein cholesterol concentration in Chinese adults. J. Am. Heart Assoc. 2018, 7, e008814. [Google Scholar] [CrossRef]
  313. Zamora-Ros, R.; Knaze, V.; Lujan-Barroso, L.; Slimani, N.; Romieu, I.; Touillaud, M. Estimation of the intake of anthocyanidins and their food sources in the European Prospective Investigation into Cancer and Nutrition (EPIC) study. Br. J. Nutr. 2011, 106, 1090–1099. [Google Scholar] [CrossRef]
  314. Chen, Y.; Ge, Z.; Huang, S.; Zhou, L.; Zhai, C.; Chen, Y.; Hu, Q.; Cao, W.; Weng, Y.; Li, Y. Delphinidin attenuates pathological cardiac hypertrophy via the AMPK/NOX/MAPK signaling pathway. Aging 2020, 12, 5362. [Google Scholar] [CrossRef]
  315. Hu, W.; Wang, W.; Ma, Q.; Liu, T.; Zhang, J.; Zhang, J. Blueberry anthocyanin-enriched extract ameliorates transverse aortic constriction-induced myocardial dysfunction via the DDAH1/ADMA/NO signaling pathway in mice. Mol. Med. Rep. 2020, 21, 454–462. [Google Scholar] [CrossRef]
  316. Aloud, B.M.; Raj, P.; McCallum, J.; Kirby, C.; Louis, X.L.; Jahan, F.; Yu, L.; Hiebert, B.; Duhamel, T.A.; Wigle, J.T.; et al. Cyanidin 3-O-glucoside prevents the development of maladaptive cardiac hypertrophy and diastolic heart dysfunction in 20-week-old spontaneously hypertensive rats. Food Funct. 2018, 9, 3466–3480. [Google Scholar] [CrossRef] [PubMed]
  317. Chen, Y.-F.; Shibu, M.A.; Fan, M.-J.; Chen, M.-C.; Viswanadha, V.P.; Lin, Y.-L.; Lai, C.-H.; Lin, K.-H.; Ho, T.-J.; Kuo, W.-W.; et al. Purple rice anthocyanin extract protects cardiac function in STZ-induced diabetes rat hearts by inhibiting cardiac hypertrophy and fibrosis. J. Nutr. Biochem. 2016, 31, 98–105. [Google Scholar] [CrossRef] [PubMed]
  318. Yue, E.; Yu, Y.; Wang, X.; Liu, B.; Bai, Y.; Yang, B. Anthocyanin protects cardiac function and cardiac fibroblasts from high-glucose induced inflammation and myocardial fibrosis by inhibiting IL-17. Front. Pharmacol. 2021, 11, 593633. [Google Scholar] [CrossRef] [PubMed]
  319. Liu, Y.; Tan, D.; Shi, L.; Liu, X.; Zhang, Y.; Tong, C.; Song, D.; Hou, M. Blueberry Anthocyanins-Enriched Extracts Attenuate Cyclophosphamide-Induced Cardiac Injury. PLoS ONE 2015, 10, e0127813. [Google Scholar] [CrossRef] [PubMed]
  320. Kim, H.-Y.; Yoon, J.-J.; Lee, H.-K.; Tai, A.-L.; Lee, Y.-J.; Kim, D.-S.; Kang, D.-G.; Lee, H.-S. Blackcurrant Improves Diabetic Cardiovascular Dysfunction by Reducing Inflammatory Cytokines in Type 2 Diabetes Mellitus Mice. Nutrients 2021, 13, 4177. [Google Scholar] [CrossRef] [PubMed]
  321. Toufektsian, M.-C.; de Lorgeril, M.; Nagy, N.; Salen, P.; Donati, M.B.; Giordano, L.; Mock, H.-P.; Peterek, S.; Matros, A.; Petroni, K.; et al. Chronic dietary intake of plant-derived anthocyanins protects the rat heart against ischemia-reperfusion injury. J. Nutr. 2008, 138, 747–752. [Google Scholar] [CrossRef] [PubMed]
  322. Guler, A.; Sahin, M.A.; Yucel, O.; Yokusoglu, M.; Gamsizkan, M.; Ozal, E.; Demirkilic, U.; Arslan, M. Proanthocyanidin prevents myocardial ischemic injury in adult rats. Med. Sci. Monit. 2011, 17, BR326–BR331. [Google Scholar] [CrossRef] [PubMed]
  323. Petersen, C.; Bharat, D.; Cutler, B.R.; Gholami, S.; Denetso, C.; Mueller, J.E.; Cho, J.M.; Kim, J.-S.; Symons, J.D.; Babu, P.V.A. Circulating metabolites of strawberry mediate reductions in vascular inflammation and endothelial dysfunction in db/db mice. Int. J. Cardiol. 2018, 263, 111–117. [Google Scholar] [CrossRef]
  324. Joo, H.K.; Choi, S.; Lee, Y.R.; Lee, E.O.; Park, M.S.; Park, K.B.; Kim, C.-S.; Lim, Y.P.; Park, J.-T.; Jeon, B.H. Anthocyanin-Rich Extract from Red Chinese Cabbage Alleviates Vascular Inflammation in Endothelial Cells and Apo E−/− Mice. Int. J. Mol. Sci. 2018, 19, 816. [Google Scholar] [CrossRef]
  325. Nasri, S.; Roghani, M.; Baluchnejadmojarad, T.; Rabani, T.; Balvardi, M. Vascular mechanisms of cyanidin-3-glucoside response in streptozotocin-diabetic rats. Pathophysiology 2011, 18, 273–278. [Google Scholar] [CrossRef]
  326. Edirisinghe, I.; Banaszewski, K.; Cappozzo, J.; McCarthy, D.; Burton-Freeman, B.M. Effect of black currant anthocyanins on the activation of endothelial nitric oxide synthase (eNOS) in vitro in human endothelial cells. J. Agric. Food Chem. 2011, 59, 8616–8624. [Google Scholar] [CrossRef] [PubMed]
  327. Ojeda, D.; Jiménez-Ferrer, E.; Zamilpa, A.; Herrera-Arellano, A.; Tortoriello, J.; Alvarez, L. Inhibition of angiotensin converting enzyme (ACE) activity by the anthocyanins delphinidin- and cyanidin-3-o-sambubiosides from Hibiscus sabdariffa. J. Ethnopharmacol. 2010, 127, 7–10. [Google Scholar] [CrossRef] [PubMed]
  328. Shaughnessy, K.S.; Boswall, I.A.; Scanlan, A.P.; Gottschall-Pass, K.T.; Sweeney, M.I. Diets containing blueberry extract lower blood pressure in spontaneously hypertensive stroke-prone rats. Nutr. Res. 2009, 29, 130–138. [Google Scholar] [CrossRef] [PubMed]
  329. Xu, C.; Zhu, J.; Gong, G.; Guo, L.; Zhang, Y.; Zhang, Z.; Ma, C. Anthocyanin attenuates high salt-induced hypertension via inhibiting the hyperactivity of the sympathetic nervous system. Clin. Exp. Hypertens. 2023, 45, 2233717. [Google Scholar] [CrossRef] [PubMed]
  330. Ugwu, P.; Ubom, R.; Madueke, P.; Okorie, P.; Nwachukwu, D. Anti-Hypertensive Effects of Anthocyanins from Hibiscus sabdarifa Calyx on the Renin-Angiotensin-Aldosterone System in Wistar Rats. Niger. J. Physiol. Sci. 2022, 37, 113–117. [Google Scholar] [CrossRef] [PubMed]
  331. Yang, Y.; Andrews, M.C.; Hu, Y.; Wang, D.; Qin, Y.; Zhu, Y.; Ni, H.; Ling, W. Anthocyanin extract from black rice significantly ameliorates platelet hyperactivity and hypertriglyceridemia in dyslipidemic rats induced by high fat diets. J. Agric. Food Chem. 2011, 59, 6759–6764. [Google Scholar] [CrossRef] [PubMed]
  332. Jiang, Y.; Dai, M.; Nie, W.-J.; Yang, X.-R.; Zeng, X.-C. Effects of the ethanol extract of black mulberry (Morus nigra L.) fruit on experimental atherosclerosis in rats. J. Ethnopharmacol. 2017, 200, 228–235. [Google Scholar] [CrossRef] [PubMed]
  333. Suh, J.-H.; Romain, C.; González-Barrio, R.; Cristol, J.-P.; Teissèdre, P.-L.; Crozier, A.; Rouanet, J.-M. Raspberry juice consumption, oxidative stress and reduction of atherosclerosis risk factors in hypercholesterolemic golden Syrian hamsters. Food Funct. 2011, 2, 400. [Google Scholar] [CrossRef]
  334. Wang, L.; Zhu, H.; Zhao, Y.; Jiao, R.; Lei, L.; Chen, J.; Wang, X.; Zhang, Z.; Huang, Y.; Wang, T.; et al. Cranberry anthocyanin as an herbal medicine lowers plasma cholesterol by increasing excretion of fecal sterols. Phytomedicine 2018, 38, 98–106. [Google Scholar] [CrossRef]
  335. Vendrame, S.; Daugherty, A.; Kristo, A.S.; Klimis-Zacas, D. Wild blueberry (Vaccinium angustifolium)-enriched diet improves dyslipidaemia and modulates the expression of genes related to lipid metabolism in obese Zucker rats. Br. J. Nutr. 2014, 111, 194–200. [Google Scholar] [CrossRef]
  336. Gaiz, A.A.; Mosawy, S.; Colson, N.; Singh, I. Potential of Anthocyanin to Prevent Cardiovascular Disease in Diabetes. Altern. Ther. Health Med. 2018, 24, 40–47. [Google Scholar]
  337. Yao, Y.; Zhang, X.; Xu, Y.; Zhao, Y.; Song, F.; Tian, Z.; Zhao, M. Cyanidin-3- O-β-Glucoside Attenuates Platelet Chemokines and Their Receptors in Atherosclerotic Inflammation of ApoE−/− Mice. J. Agric. Food Chem. 2022, 70, 8254–8263. [Google Scholar] [CrossRef] [PubMed]
  338. Erlund, I.; Koli, R.; Alfthan, G.; Marniemi, J.; Puukka, P.; Mustonen, P. Favorable effects of berry consumption on platelet function, blood pressure, and HDL cholesterol. Am. J. Clin. Nutr. 2008, 87, 323–331. [Google Scholar] [CrossRef] [PubMed]
  339. Vendrame, S.; Klimis-Zacas, D. Potential Factors Influencing the Effects of Anthocyanins on Blood Pressure Regulation in Humans: A Review. Nutrients 2019, 11, 1431. [Google Scholar] [CrossRef]
  340. Curtis, P.J.; van der Velpen, V.; Berends, L.; Jennings, A.; Feelisch, M.; Umpleby, A.M.; Evans, M. Blueberries improve biomarkers of cardiometabolic function in participants with metabolic syndrome—Results from a 6-month, double-blind, randomized controlled trial. Am. J. Clin Nutr. 2019, 109, 1535–1545. [Google Scholar] [CrossRef] [PubMed]
  341. Huang, L.; Di Xiao, D.; Zhang, X.; Sandhu, A.K.; Chandra, P.; Kay, C.; Edirisinghe, I.; Burton-Freeman, B. Strawberry Consumption, Cardiometabolic Risk Factors, and Vascular Function: A Randomized Controlled Trial in Adults with Moderate Hypercholesterolemia. J. Nutr. 2021, 151, 1517–1526. [Google Scholar] [CrossRef] [PubMed]
  342. Rodriguez-Mateos, A.; Istas, G.; Boschek, L.; Feliciano, R.P.; Mills, C.E.; Boby, C.; Gómez-Alonso, S.; Milenkovic, D.; Heiss, C. Circulating Anthocyanin Metabolites Mediate Vascular Benefits of Blueberries: Insights from Randomized Controlled Trials, Metabolomics, and Nutrigenomics. J. Gerontol. A Biol. Sci. Med. Sci. 2019, 74, 967–976. [Google Scholar] [CrossRef] [PubMed]
  343. Heiss, C.; Istas, G.; Feliciano, R.P.; Weber, T.; Wang, B.; Favari, C.; Mena, P.; Del Rio, D.; Rodriguez-Mateos, A. Daily consumption of cranberry improves endothelial function in healthy adults: A double blind randomized controlled trial. Food Funct. 2022, 13, 3812–3824. [Google Scholar] [CrossRef]
  344. Yang, L.; Ling, W.; Du, Z.; Chen, Y.; Li, D.; Deng, S.; Liu, Z.; Yang, L. Effects of anthocyanins on cardiometabolic health: A systematic review and meta-analysis of randomized controlled trials. Adv. Nutr. 2017, 8, 684–693. [Google Scholar] [CrossRef]
  345. Sesso, H.D.; Manson, J.E.; Aragaki, A.K.; Rist, P.M.; Johnson, L.G.; Friedenberg, G.; Copeland, T.; Clar, A.; Mora, S.; Moorthy, M.V.; et al. COSMOS Research Group. Effect of cocoa flavanol supplementation for the prevention of cardiovascular disease events: The COcoa Supplement and Multivitamin Outcomes Study (COSMOS) randomized clinical trial. Am. J. Clin. Nutr. 2022, 115, 1490–1500. [Google Scholar] [CrossRef]
  346. McCullough, M.L.; Peterson, J.J.; Patel, R.; Jacques, P.F.; Shah, R.; Dwyer, J.T. Flavonoid intake and cardiovascular disease mortality in a prospective cohort of US adults. Am. J. Clin. Nutr. 2012, 95, 454–464. [Google Scholar] [CrossRef] [PubMed]
  347. Grosso, G.; Micek, A.; Godos, J.; Pajak, A.; Sciacca, S.; Galvano, F.; Giovannucci, E.L. Dietary flavonoid and Lignan intake and mortality in prospective cohort studies: Systematic review and dose-response meta-analysis. Am. J. Epidemiol. 2017, 185, 1304–1316. [Google Scholar] [CrossRef] [PubMed]
  348. Tresserra-Rimbau, A.; Rimm, E.B.; Medina-Remón, A.; Martínez-González, M.A.; López-Sabater, M.C.; Covas, M.I. Polyphenol intake and mortality risk: A re-analysis of the PREDIMED trial. BMC Med. 2014, 12, 77. [Google Scholar] [CrossRef] [PubMed]
  349. Castellarin, S.D.; Bavaresco, L.; Falginella, L.; Van Zeller Gonçalves, M.I.; Di Gaspero, G. Phenolics in Grape Berry and Key Antioxidants. In The Biochemistry of the Grape Berry; Gerós, H., Chavez, M.M., Delrot, S., Eds.; Bentham e Books; Bentham Science: Sharjah, United Arab Emirates, 2012; pp. 89–110. [Google Scholar]
  350. Visioli, F.; Panaite, S.A.; Tomé-Carneiro, J. Wine’s Phenolic Compounds and Health: A Pythagorean View. Molecules 2020, 25, 4105. [Google Scholar] [CrossRef] [PubMed]
  351. Haseeb, S.; Alexander, B.; Baranchuk, A. Wine and cardiovascular health: A comprehensive review. Circulation 2017, 136, 1434–1448. [Google Scholar] [CrossRef] [PubMed]
  352. Di Renzo, L.; Marsella, L.T.; Carraro, A.; Valente, R.; Gualtieri, P.; Gratteri, S.; Tomasi, D.; Gaiotti, F.; De Lorenzo, A. Changes in LDL Oxidative Status and Oxidative and Inflammatory Gene Expression after Red Wine Intake in Healthy People: A Randomized Trial. Mediat. Inflamm. 2015, 2015, 317348. [Google Scholar] [CrossRef] [PubMed]
  353. Nova, E.; San Mauro-Martín, I.; Díaz-Prieto, L.E.; Marcos, A. Wine and beer within a moderate alcohol intake is associated with higher levels of HDL-c and adiponectin. Nutr. Res. 2019, 63, 42–50. [Google Scholar] [CrossRef] [PubMed]
  354. Golan, R.; Gepner, Y.; Shai, I.J.E. Wine and Health–New Evidence. Eur. J. Clin. Nutr. 2019, 72, 55–59. [Google Scholar] [CrossRef]
  355. Rasines-Perea, Z.; Teissedre, P.-L. Grape Polyphenols’ Effects in Human Cardiovascular Diseases and Diabetes. Molecules 2017, 22, 68. [Google Scholar] [CrossRef]
  356. Chiva-Blanch, G.; Urpi-Sarda, M.; Ros, E.; Valderas-Martinez, P.; Casas, R.; Arranz, S.; Guillén, M.; Lamuela-Raventós, R.M.; Llorach, R.; Andres-Lacueva, C.; et al. Effects of red wine polyphenols and alcohol on glucose metabolism and the lipid profile: A randomized clinical trial. Clin. Nutr. 2013, 32, 200–206. [Google Scholar] [CrossRef]
  357. Navarro-García, F.; Ponce-Ruíz, N.; Rojas-García, A.E.; Ávila-Villarreal, G.; Herrera-Moreno, J.F.; Barrón-Vivanco, B.S.; Bernal-Hernández, Y.Y.; González-Arias, C.A.; Medina-Díaz, I.M. The Role of Nutritional Habits and Moderate Red Wine Consumption in PON1 Status in Healthy Population. Appl. Sci. 2021, 11, 9503. [Google Scholar] [CrossRef]
  358. Toth, A.; Sandor, B.; Papp, J.; Rabai, M.; Botor, D.; Horvath, Z.; Kenyeres, P.; Juricskay, I.; Toth, K.; Czopf, L. Moderate red wine consumption improves hemorheological parameters in healthy volunteers. Clin. Hemorheol. Microcirc. 2014, 56, 13–23. [Google Scholar] [CrossRef] [PubMed]
  359. Naissides, M.; Mamo, J.C.; James, A.P.; Pal, S. The effect of acute red wine polyphenol consumption on postprandial lipaemia in postmenopausal women. Atherosclerosis 2004, 177, 401–408. [Google Scholar] [CrossRef] [PubMed]
  360. Grønbaek, M.; Becker, U.; Johansen, D.; Gottschau, A.; Schnohr, P.; Hein, H.O.; Jensen, G.; Sørensen, T.I. Type of alcohol consumed and mortality from all causes, coronary heart disease, and cancer. Ann. Intern. Med. 2000, 133, 411–419. [Google Scholar] [CrossRef]
  361. Stampfer, M.J.; Colditz, G.A.; Willett, W.C.; Speizer, F.E.; Hennekens, C.H. A prospective study of moderate alcohol consumption and the risk of coronary disease and stroke in women. N. Engl. J. Med. 1988, 319, 267–273. [Google Scholar] [CrossRef] [PubMed]
  362. Levantesi, G.; Marfisi, R.; Mozaffarian, D.; Franzosi, M.G.; Maggioni, A.; Nicolosi, G.L.; Schweiger, C.; Silletta, M.; Tavazzi, L.; Tognoni, G.; et al. Wine consumption and risk of cardiovascular events after myocardial infarction: Results from the GISSI-Prevenzione trial. Int. J. Cardiol. 2013, 163, 282–287. [Google Scholar] [CrossRef] [PubMed]
  363. Lucerón-Lucas-Torres, M.; Saz-Lara, A.; Díez-Fernández, A.; Martínez-García, I.; Martínez-Vizcaíno, V.; Cavero-Redondo, I.; Álvarez-Bueno, C. Association between Wine Consumption with Cardiovascular Disease and Cardiovascular Mortality: A Systematic Review and Meta-Analysis Maribel. Nutrients 2023, 15, 2785. [Google Scholar] [CrossRef]
  364. Torres, A.; Cachofeiro, V.; Millán, J.; Lahera, V.; Nieto, M.; Martin, R.; Bello, E.; Alvarez-Sala, L.A. Red wine intake but not other alcoholic beverages increases total antioxidant capacity and improves pro-inflammatory profile after an oral fat diet in healthy volunteers. Rev. Clín. Esp. 2015, 215, 486–494. [Google Scholar] [CrossRef]
  365. GBD 2020 Alcohol Collaborators. Population-level risks of alcohol consumption by amount, geography, age, sex, and year: A systematic analysis for the Global Burden of Disease Study 2020. Lancet 2022, 400, 185–235. [Google Scholar] [CrossRef]
  366. Anderson, P. The Impact of Alcoholic Beverages on Human Health. Nutrients 2021, 13, 4417. [Google Scholar] [CrossRef]
  367. WHO. Alcohol. Available online: https://www.who.int/news-room/fact-sheets/detail/alcohol (accessed on 4 September 2023).
  368. Rocha, K.K.R.; Souza, G.A.; Ebaid, G.X.; Seiva, F.R.F.; Cataneo, A.C.; Novelli, E.L.B. Resveratrol toxicity: Effects on risk factors for atherosclerosis and hepatic oxidative stress in standard and high-fat diets. Food Chem. Toxicol. 2009, 47, 1362–1367. [Google Scholar] [CrossRef]
  369. De la Lastra, C.A.; Villegas, I. Resveratrol as an antioxidant and pro-oxidant agent: Mechanisms and clinical implications. Biochem. Soc. Trans. 2007, 35 Pt 5, 1156–1160. [Google Scholar] [CrossRef] [PubMed]
  370. Macedo, T.S.J.; Barros, V.R.P.; Monte, A.P.O.; Gouveia, B.B.; Bezerra, M.É.S.; Cavalcante, A.Y.P.; Barberino, R.S.; Menezes, V.G.; Matos, M.H.T. Resveratrol has dose-dependent effects on DNA fragmentation and mitochondrial activity of ovine secondary follicles cultured in vitro. Zygote 2017, 25, 434–442. [Google Scholar] [CrossRef] [PubMed]
  371. Pignitter, M.; Schueller, K.; Burkon, A.; Knorr, V.; Esefelder, L.; Doberer, D.; Wolzt, M.; Somoza, V. Concentration-dependent effects of resveratrol and metabolites on the redox status of human erythrocytes in single-dose studies. J. Nutr. Biochem. 2016, 27, 164–170. [Google Scholar] [CrossRef]
  372. Berardi, V.; Ricci, F.; Castelli, M.; Galati, G.; Risuleo, G. Resveratrol exhibits a strong cytotoxic activity in cultured cells and has an antiviral action against polyomavirus: Potential clinical use. J. Exp. Clin. Cancer Res. 2009, 28, 96. [Google Scholar] [CrossRef] [PubMed]
  373. Martins, L.A.M.; Coelho, B.P.C.; Behr, G.; Pettenuzzo, L.F.; Souza, I.C.C.; Moreira, J.C.F.; Borojevic, R.; Gottfried, C.; Guma, F.C.R. Resveratrol induces pro-oxidant effects and time-dependent resistance to cytotoxicity in activated hepatic stellate cells. Cell Biochem. Biophys. 2014, 68, 247–257. [Google Scholar] [CrossRef] [PubMed]
  374. Brown, V.A.; Patel, K.R.; Viskaduraki, M.; Crowell, J.A.; Perloff, M.; Booth, T.D.; Vasilinin, G.; Sen, A.; Schinas, A.M.; Piccirilli, G.; et al. Repeat dose study of the cancer chemopreventive agent resveratrol in healthy volunteers: Safety, pharmacokinetics, and effect on the insulin-like growth factor axis. Cancer Res. 2010, 70, 9003–9011. [Google Scholar] [CrossRef] [PubMed]
  375. Crowell, J.A.; Korytko, P.J.; Morrissey, R.L.; Booth, T.D.; Levine, B.S. Resveratrol-associated renal toxicity. Toxicol. Sci. 2004, 82, 614–619. [Google Scholar] [CrossRef]
  376. la Porte, C.; Voduc, N.; Zhang, G.; Seguin, I.; Tardiff, D.; Singhal, N.; Cameron, D.W. Steady-State pharmacokinetics and tolerability of trans-resveratrol 2000 mg twice daily with food, quercetin and alcohol (ethanol) in healthy human subjects. Clin. Pharmacokinet. 2010, 49, 449–454. [Google Scholar] [CrossRef]
  377. Jiang, H.; Webster, D.; Cao, J.; Shao, A. The Safety of Green Tea and Green Tea Extract Consumption in Adults-Results of a Systematic Review. Regul. Toxicol. Pharmacol. 2018, 95, 412–433. [Google Scholar]
  378. Younes, M.; Aggett, P.; Aguilar, F.; Crebelli, R.; Dusemund, B.; Filipič, M.; Frutos, M.J.; Galtier, P.; Gott, D. Scientific Opinion on the Safety of Green Tea Catechins. EFSA J. 2018, 16, e05239. [Google Scholar] [PubMed]
  379. Cruz-Correa, M.; Shoskes, D.A.; Sanchez, P.; Zhao, R.; Hylind, L.M.; Wexner, S.D.; Giardiello, F.M. Combination treatment with curcumin and quercetin of adenomas in familial adenomatous polyposis. Clin. Gastroenterol. Hepatol. 2006, 4, 1035–1038. [Google Scholar] [CrossRef] [PubMed]
  380. Zhang, H.; Tsao, R. Dietary polyphenols, oxidative stress and antioxidant and anti-inflammatory effects. Curr. Opin. Food Sci. 2016, 8, 33–42. [Google Scholar] [CrossRef]
  381. Hurrell, R.; Egli, I. Iron bioavailability and dietary reference values. Am. J. Clin. Nutr. 2010, 91, 1461S–1467S. [Google Scholar] [CrossRef]
Figure 1. The primary sources and physiological effects of relevant polyphenols.
Figure 1. The primary sources and physiological effects of relevant polyphenols.
Biomedicines 11 02888 g001
Figure 2. The main mechanisms of action of stilbenes. Akt: protein kinase B; AMPK: AMP-activated protein kinase; COX: cyclooxygenase; Bcl-2: B-cell lymphoma 2, DNA: deoxyribonucleic acid; Drp1: dynamin-related protein 1; ETC: electron transport chain; FOXO: Forkhead box O; GPx: glutathione peroxidase; GSH: glutathione; GSK-3β: glycogen synthase kinase-3 beta; eNOS: endothelial nitric oxide synthase; HO-1: heme oxygenase-1; ICAM: intracellular cell adhesion molecule; IL: interleukin; JAK: Janus kinase; MAPK: mitogen-activated protein kinase; mTOR: mammalian target of rapamycin; NF-κB: nuclear factor kappa B; NOX: NADPH oxidases; Nrf: nuclear respiratory factor; PI3K: phosphatidylinositol 3-kinase; PGC-1α: peroxisome proliferator-activated receptor gamma coactivator 1-alpha; ROS: reactive oxygen species; SIRT1: NAD-dependent deacetylase sirtuin-1; SOD: superoxide dismutase; STAT: signal transducer and activator of transcription proteins; TGF-β: transforming growth factor; TLR4: toll-like receptor 4; TNF: tumor necrosis factor; VCAM: vascular cell adhesion molecule.
Figure 2. The main mechanisms of action of stilbenes. Akt: protein kinase B; AMPK: AMP-activated protein kinase; COX: cyclooxygenase; Bcl-2: B-cell lymphoma 2, DNA: deoxyribonucleic acid; Drp1: dynamin-related protein 1; ETC: electron transport chain; FOXO: Forkhead box O; GPx: glutathione peroxidase; GSH: glutathione; GSK-3β: glycogen synthase kinase-3 beta; eNOS: endothelial nitric oxide synthase; HO-1: heme oxygenase-1; ICAM: intracellular cell adhesion molecule; IL: interleukin; JAK: Janus kinase; MAPK: mitogen-activated protein kinase; mTOR: mammalian target of rapamycin; NF-κB: nuclear factor kappa B; NOX: NADPH oxidases; Nrf: nuclear respiratory factor; PI3K: phosphatidylinositol 3-kinase; PGC-1α: peroxisome proliferator-activated receptor gamma coactivator 1-alpha; ROS: reactive oxygen species; SIRT1: NAD-dependent deacetylase sirtuin-1; SOD: superoxide dismutase; STAT: signal transducer and activator of transcription proteins; TGF-β: transforming growth factor; TLR4: toll-like receptor 4; TNF: tumor necrosis factor; VCAM: vascular cell adhesion molecule.
Biomedicines 11 02888 g002
Table 1. The effect of various polyphenols on the cardiovascular system: selected randomized human clinical studies and meta-analyses.
Table 1. The effect of various polyphenols on the cardiovascular system: selected randomized human clinical studies and meta-analyses.
Resveratrol
AuthorsYearArea of InterestStudyStudy DesignMain FindingsRef.
Boa et al.2016Lipid profileRCT
  • 120 patients with DM;
  • 500 mg PO daily for 24 weeks
  • Increased TC levels
  • No change in LDL-C, HDL-C, and TG
[136]
Asgary et al.2019Lipid profileMeta-analysis
  • 396 subjects;
  • 100–3000 mg PO daily
  • No change in TC levels
  • Increase in HDL-C
[137]
Magyar et al.2012Lipid profile,
endothelial function
RCT
  • 40 patients with CAD;
  • 10 mg daily for 3 months, oral capsule
  • Decrease in LDL-C levels only
  • Improvement in FMD
[27]
Hoseini et al.2019Metabolic statusRCT
  • 56 patients with CAD + DM;
  • 500 mg PO daily for 4 weeks
  • Beneficial effects on TC/HDL-C ratio, HDL-C level, glycemic control
[138]
Simental-Mendia et al.2019Lipid profileRCT
  • 71 patients with dyslipidemia;
  • 100 mg PO daily for 2 months
  • Reduction in TC levels
[139]
Fogacci et al.2018Blood pressureMeta-analysis
  • 681 obese subjects;
  • >300 mg daily
  • Decrease in SBP
[140]
Liu et al.2015Blood pressureMeta-analysis
  • 247 subjects;
  • >150 mg daily
  • Decrease in SBP
[141]
Bhatt et al.2012Blood pressure, diabetes mellitusRCT
  • 62 patients with DM;
  • 250 mg PO daily for 3 months
  • Improved SBP and glycemic control
[142]
Imamura et al.2017Blood pressure, arterial stiffnessRCT
  • 50 patients with DM;
  • 100 mg PO daily for 12 weeks
  • Improved arterial stiffness and blood pressure
[143]
Timmers et al.2011Blood pressureRCT
  • 11 healthy, obese men;
  • 150 mg PO daily for 30 days
  • Reduction in SBP
[144]
Theodotou et al.2017Blood pressureRCT
  • 97 patients with HT;
  • 50 mg/day PO for 2 years
  • Reduction in SBP when added to standard therapy
[145]
Walker et al.2019Blood pressure, diabetesRCT
  • 28 men with MS;
  • 2 g PO daily for 30 days
  • No effect on blood pressure
  • Improved glucose homeostasis
[146]
Wong et al.2011Endothelial functionRCT
  • 19 postmenopausal women and obese men with HT;
  • single PO dose (270 mg)
  • Improved FMD after acute supplementation
[147]
Marques et al.2018Endothelial functionRCT
  • 24 subjects with HT;
  • single PO dose (300 mg)
  • Improved FMD in women after acute supplementation
[148]
Gal et al.2020Heart failure, lipid profileRCT
  • 60 patients with HFrEF;
  • 100 mg PO daily for 3 months
  • Improved cardiac function, exercise tolerance, QoL, and biomarkers
  • Reduced TC and LDL-C levels
[66]
Quercetin
AuthorsYearArea of InterestStudyStudy designMain findingsRef.
Edwards et al.2007Blood pressureFirst RCT
  • 41 subjects with HT;
  • 730 mg/day PO for 4 weeks
  • Reduced SBP and DBP
[149]
Zahedi et al.2013Blood pressure, lipid profileRCT
  • 72 women with diabetes;
  • 500 mg PO daily for 10 weeks
  • Reduced SBP
  • No effect on lipid profile and DBP
[150]
Brüll et al.2015Blood pressureRCT
  • 70 patients;
  • 162 mg PO daily for 6 weeks
  • Decreased daytime and nighttime SBP
[151]
Dower et al.2015Blood pressureRCT
  • 37 subjects with HT;
  • 160 mg/day PO for 4 weeks
  • Reduced BP measured by ABPM
[152]
Serban et al.2016Blood pressureMeta-analysis
  • 587 patients;
  • >500 mg vs. <500 mg/day
  • Reduced SBP and DBP at doses above 500 mg/day
[153]
Larson et al.2012Blood pressureRCT
  • 12 subjects with HT;
  • 1095 mg in a single dose PO
  • Acute reduction of SBP and DBP
[154]
Bondonno2016Blood pressureRCT
  • 15 health patients;
  • single dose <400 mg PO
  • No acute effect on SBP and DBP
[155]
Chekalina2017Heart functionRCT
  • 85 subjects with CAD;
  • 120 mg PO daily for 2 months
  • Improved both systolic and diastolic LV function
[156]
Kondratiuk et al.2018Heart functionRCT
  • 84 males with HT;
  • 12 months PO
  • Improved diastolic function(LV mass and e/e’)
[157]
Catechins
AuthorsYearArea of interestStudyStudy designMain findingsRef.
Bogdansy et al.2012Blood pressureRCT
  • 56 subjects;
  • 379 mg/day PO for 3 months green tea extract
  • Reduced SBP and DBP
[158]
Rostami et al.2015Blood pressure, lipid profileRCT
  • 60 subjects, high CV risk;
  • 25 g dark chocolate for 8 weeks PO
  • Reduced SBP and DBP
  • Improved TG level
[159]
Dower et al.2015Blood pressureRCT
  • 37 subjects with HT;
  • 100 mg/day PO epicatechin for 4 weeks
  • No effect on BP and vascular function
[152]
Chatree et al.2021Blood pressureRCT
  • 30 subjects;
  • 150 mg PO EGCG for 8 weeks
  • Reduced SBP and DBP
  • Improved TG level
[160]
Ellinger et al.2012Blood pressureMeta-analysis
  • 31 clinical trials;
  • cocoa extracts (epicatechin)
  • Dose-dependent effect on BP (>25 mg/day epi)
[161]
Oyama et al.2010Vascular functionRCT
  • 30 patients;
  • 580 mg/day green tea extract for 2 weeks PO
  • Improved the endothelium-dependent vasorelaxation
[162]
Alañón et al.2020Vascular functionRCT
  • 20 healthy volunteers;
  • different doses of epicatechin
  • Dose-dependent effect on vascular relaxation (>0.5 mg/kg epi)
[163]
Samavat et al.2016Lipid profileRCT
  • 1075 women;
  • green tea extract (EGCG 843 mg/day) for 12 months PO
  • EGCG reduced TC and LDL-C level
[164]
Suzuki-Sugihara et al.2016Redox statusRCT
  • 19 healthy male subjects;
  • 1 g catechin in a single PO dose
  • Increased total antioxidant capacity
  • Decreased ox-LDL-C level
[165]
McDermott et al.2020PADRCT
  • 44 subjects with PAD;
  • 75 mg/day PO epi for 6 months
  • Improved six-minute walk distance
[166]
Kishimoto et al.2020CADRCT
  • 612 subjects +/− CAD;
  • green tea extracts PO
  • Green tea is inversely associated with the prevalence of CAD and MI
[167]
Wang et al.2011CADMeta-analysis
  • 5 studies on green tea
  • 10% reduction in the risk of developing CAD
[168]
Dural et al.2022Heart failureRCT
  • 20 patients with HFrEF;
  • dark or milk chocolate PO
  • Reduced NT-proBNP levels
  • Improved vascular function
[169]
Anthocyanins
AuthorsYearArea of interestStudyStudy designMain findingsRef.
Cassidy et al.2013Risk of MIOBS
  • 93,000 healthy subjects;
  • AC-rich diet for 18 years PO
  • 47% reduction in MI risk
[170]
Cassidy et al.2016Risk of MI and ischemic strokeOBS
  • 4046 males;
  • AC-rich diet for 24 years PO
  • 13% reduction in MI risk
[171]
Alvarez-Suarez et al.2014Platelet function, lipid profileRCT
  • 23 healthy volunteers;
  • 500 g daily PO intake of strawberries for 1 month
  • Decreased platelet activation
  • Improved lipid profile
  • Enhanced antioxidant status
[172]
Aboonabi et al.2020CV risk factors, Platelet functionRCT
  • 55 subjects with metabolic syndrome;
  • 320 mg AC-rich extract PO for 4 weeks
  • Reduced thrombogenicity
  • Improved lipid and glucose homeostasis
[173]
Santhakumar et al.2015Platelet functionRCT
  • 21 healthy subjects;
  • AC-rich extract for 28 days PO
  • Decreased platelet activation
[174]
Tian et al.2021Platelet functionRCT
  • 93 subjects with dyslipidemia;
  • various doses for 12 weeks
  • Dose-dependent attenuation of platelet function (>80 mg/day)
[175]
Jennings et al.2012Blood pressureRCT
  • 1898 subjects;
  • AC-rich PO diet
  • Improved arterial stiffness and blood pressure
[176]
McKay et al.2009Blood pressureRCT
  • 65 subjects;
  • hibiscus tea PO for 6 weeks
  • Decreased the SBP
[177]
Basu et al.2010Blood pressureRCT
  • 48 subjects with MS
  • blueberry PO for 8 weeks
  • Decreased SBP and DBP
[178]
Johnson et al.2015Blood pressureRCT
  • 48 women with HT;
  • 22 g freeze-dried blueberry for 8 weeks PO
  • Reduced blood pressure and arterial stiffness
[179]
Emamat et al.2022Blood pressureRCT
  • 85 subjects with HT;
  • 10 g/day dried purple-black barberry for 2 months PO
  • Improved SBP
[180]
Dohadwala et al.2010Vascular functionRCT
  • 59 subjects with CAD;
  • 480 mL cranberry juice (94 mg AC) for 4 weeks PO
  • Improved FMD
[181]
Huang et al.2016CV risk factorsMeta-analysis
  • 22 trials with 1251 healthy subjects or with CAD
  • Reduced LDL-C, glucose levels
  • Improved SBP
[182]
Yang et al.2017CV risk factorsRCT
  • 160 participants with DM;
  • 320 mg/day PO for 12 weeks
  • Improved lipid and glucose profile
[183]
Liu et al.2016Lipid profileMeta-analysis
  • six trials with 586 subjects with high CV risk
  • Reduced TC, TG and LDL-C levels
  • Increased HDL-C
[184]
ABPM: ambulatory blood pressure monitoring; AC: anthocyanins; CAD: coronary artery disease; CV: cardiovascular; DBP: diastolic blood pressure; DM: diabetes mellitus; EGCG: epigallocatechin-gallate; EPI: epicatechin; FMD: flow-mediated dilation; HDL-C: high-density lipoprotein cholesterol; HFrEF: heart failure with reduced ejection fraction; HT: hypertension; LDL-C: low-density lipoprotein cholesterol; LV: left ventricle; MI: myocardial infarction; MS: metabolic syndrome; OBS: observational trial; PAD: peripheral arterial disease; PO: by mouth (per os); Qct: quercetin; Qol: quality of life; RCT: randomized controlled trial; RES: resveratrol; SBP: systolic blood pressure; TC: total cholesterol.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gál, R.; Halmosi, R.; Gallyas, F., Jr.; Tschida, M.; Mutirangura, P.; Tóth, K.; Alexy, T.; Czopf, L. Resveratrol and beyond: The Effect of Natural Polyphenols on the Cardiovascular System: A Narrative Review. Biomedicines 2023, 11, 2888. https://doi.org/10.3390/biomedicines11112888

AMA Style

Gál R, Halmosi R, Gallyas F Jr., Tschida M, Mutirangura P, Tóth K, Alexy T, Czopf L. Resveratrol and beyond: The Effect of Natural Polyphenols on the Cardiovascular System: A Narrative Review. Biomedicines. 2023; 11(11):2888. https://doi.org/10.3390/biomedicines11112888

Chicago/Turabian Style

Gál, Roland, Róbert Halmosi, Ferenc Gallyas, Jr., Michael Tschida, Pornthira Mutirangura, Kálmán Tóth, Tamás Alexy, and László Czopf. 2023. "Resveratrol and beyond: The Effect of Natural Polyphenols on the Cardiovascular System: A Narrative Review" Biomedicines 11, no. 11: 2888. https://doi.org/10.3390/biomedicines11112888

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop