Next Article in Journal
Risks in Management of Enteral Nutrition in Intensive Care Units: A Literature Review and Narrative Synthesis
Next Article in Special Issue
Bioflavonoid Robinin from Astragalus falcatus Lam. Mildly Improves the Effect of Metothrexate in Rats with Adjuvant Arthritis
Previous Article in Journal
Low Protein Diets and Plant-Based Low Protein Diets: Do They Meet Protein Requirements of Patients with Chronic Kidney Disease?
Previous Article in Special Issue
The Effects of a Functional Food Breakfast on Gluco-Regulation, Cognitive Performance, Mood, and Satiety in Adults
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

From Preclinical Stroke Models to Humans: Polyphenols in the Prevention and Treatment of Stroke

Division of Pharmacology, Department of Molecular and Translational Medicine, University of Brescia, 25123 Brescia, Italy
*
Author to whom correspondence should be addressed.
Nutrients 2021, 13(1), 85; https://doi.org/10.3390/nu13010085
Submission received: 6 December 2020 / Accepted: 24 December 2020 / Published: 29 December 2020
(This article belongs to the Special Issue Dietary (Poly)Phenols and Health)

Abstract

:
Polyphenols are an important family of molecules of vegetal origin present in many medicinal and edible plants, which represent important alimentary sources in the human diet. Polyphenols are known for their beneficial health effects and have been investigated for their potential protective role against various pathologies, including cancer, brain dysfunctions, cardiovascular diseases and stroke. The prevention of stroke promoted by polyphenols relies mainly on their effect on cardio- and cerebrovascular systems. However, a growing body of evidence from preclinical models of stroke points out a neuroprotective role of these molecules. Notably, in many preclinical studies, the polyphenolic compounds were effective also when administered after the stroke onset, suggesting their possible use in promoting recovery of patients suffering from stroke. Here, we review the effects of the major polyphenols in cellular and in vivo models of both ischemic and hemorrhagic stroke in immature and adult brains. The results from human studies are also reported.

1. Introduction: The Burden of Stroke

Stroke is a cerebrovascular disease caused by the interruption of blood flow to the brain due to the blockage or rupture of a vessel and can affect both immature and mature brains.
Perinatal stroke occurs between the 20th week of gestation and the 28th day after birth with an incidence between 1/2300 and 1/5000 live births [1,2]. Perinatal stroke is not only a major cause of acute mortality in the early days of life, but newborn survivors may also develop neurological disabilities including cerebral palsy, cognitive deficits and behavioral disorders often lasting the entire lifetime [3,4].
Adult stroke is the second main cause of mortality and the third cause of disability worldwide [5,6]. Typical symptoms of stroke include understanding and speech issues; sudden unilateral weakness, numbness or loss of vision; ataxia; diplopia; dizziness; and nausea [5]. Moreover, stroke survivors face a significantly increased risk of developing depression and cognitive decline [7,8]. According to its etiology, stroke can be classified as ischemic, hemorrhagic or caused by subarachnoid hemorrhage [9,10,11]. Ischemic stroke, or brain ischemia, is the most frequent subtype of stroke, accounting for 85% of cases [9,10]. In ischemic stroke, the blood supply to part of the brain is reduced by the occlusion of a blood vessel either by an embolus or by local thrombosis [12]. About 10% of strokes are due to intracerebral hemorrhage (ICH), a severe neurological disorder associated with high rates of mortality and disability [13,14,15]. ICH results from the rupture of cerebral blood vessels that causes a rapidly expanding hematoma occurring within the brain parenchyma. Subarachnoid hemorrhage (SAH) accounts for approximately 5% of all strokes and is characterized by severe mortality and morbidity (more than 50%) [16,17]. In the major part of cases, SAH is caused by the rupture of an intracranial aneurysm [17].
Stroke in both immature and mature brains is a complex phenomenon that includes a series of pathological processes such as excitotoxicity, oxidative damage, apoptosis and inflammation, which eventually leads to cell death [18,19,20]. One of the main pathophysiological features of ischemic stroke is the brain–blood barrier (BBB) disruption, an event that occurs in almost two-thirds of patients in the first hours from the ischemia onset and causes vasogenic edema, hemorrhagic transformation and increased mortality [21,22]. Cerebral hemorrhage leads to a primary brain injury caused by increased intracranial pressure, followed by a secondary brain injury mediated by the physiological responses to hematoma, including inflammation [23].
Treatment options for stroke are currently very limited. The only approved pharmacological therapy for ischemic stroke is the recombinant tissue plasminogen activator (rtPA) [24,25]. Unfortunately, administration of rtPA after 4.5 h from the ischemic event is contraindicated for the risk of hemorrhagic conversion, limiting the use of this drug [25]. Neurosurgical interventions are performed to remove blood in ICH [26] or to treat brain aneurysm in SAH [27].
In light of the above considerations, there is an urgent need for the development of new therapies able to prevent or reduce stroke neuronal injury.

2. Polyphenols: Definition and Classification

Polyphenols are molecules chemically characterized by the presence of at least one aromatic ring with one or more hydroxyl groups attached [28,29]. Polyphenols are plant secondary metabolites that are thought to help plants to survive and proliferate, protecting them against microbial infections or herbivorous animals, or luring pollinators [30]. Polyphenols are found in many medicinal and edible plants which represent important alimentary sources, including fruits, vegetables, beverages (such as tea and red wine) and extra virgin oil [31].
This group of natural products includes a broad number of different compounds, ranging from simple molecules with low molecular weight to complex and large derived polyphenols [28,29]. According to their chemical structure, polyphenols can be classified into various classes including flavonoids, phenolic acids, stilbenes, curcuminoids, lignans, ellagitannins and ellagic acid and coumarins [28,29]. Flavonoids are structurally based on a skeleton of fifteen carbons, with two aromatic rings connected by a three-carbon bridge. They are the most numerous of polyphenols and are widely distributed through the plant kingdom [28,29]. The main subclasses of dietary flavonoids include flavonols, flavan-3-ols, flavanones, flavones, isoflavones, anthocyanins, dihydrochalcones and proanthocyanidins [28,29]. Among the non-flavonoid polyphenols, phenolic acids can be further divided into hydroxycinnamic acids and hydroxybenzoic acids [28,29]. Figure 1 depicts the classification of polyphenols and describes some known food sources for each molecular class.
Among the many micronutrients present in plants, polyphenols are the most numerous and particularly endowed with beneficial properties [32]. For these reasons, polyphenols have been widely investigated for the prevention and treatment of several pathological conditions, including cancer, neurodegenerative disorders, metabolic and cardiovascular diseases and stroke [33,34,35].

3. Polyphenol Metabolism: Role of Gut Microbiota

Polyphenol bioavailability is generally poor, and only 1–10% of total polyphenol intake is detectable in blood and urine samples [33]. Bioavailability is particularly low for flavones, stilbenes and curcumin and is slightly higher for tea flavan-3-ols, flavanones in citrus fruits, soy isoflavones and red wine anthocyanidins [36,37,38]. However, maximum polyphenol concentration in plasma remains extremely low and rarely exceeds 1 μM, even in individuals consuming a polyphenol-rich diet [36]. How polyphenols exert their beneficial actions despite their poor bioavailability is not clear yet. A possible explanation may rely on the fact that many polyphenol metabolites exhibit a biological activity [39].
Polyphenols are generally consumed with the diet or as supplements. A proportion of ingested polyphenolic compounds can be absorbed in the small intestine and metabolized by phase II enzymes. However, the major part of polyphenols reach the large intestine where they are degraded by intestinal microbiota. A large body of evidence indicates a fundamental role of colonic microorganisms in determining the bioavailability and activity of polyphenols by transforming them into readily absorbable molecules or biologically active metabolites [32,40,41]. The relationship between polyphenols and microbiota is bidirectional and, if the intestinal bacteria modulate polyphenol metabolism, polyphenolic compounds can in their turn influence the composition of the microbial population [32,42].
Different findings suggest the gut microbiota could modulate the activity of polyphenols potentially active against stroke. For example, flavan-3-ols, phenolic compounds characterized by a generally low bioavailability, are extensively metabolized by host and gut microbiota enzymes. Phenyl-γ-valerolactones and phenylvaleric acids, the main microbial metabolites of flavan-3-ols, might be responsible for the beneficial effects attributed to their parent compounds, including neuroprotection [43]. Daidzein, an isoflavone endowed with beneficial properties enriched in soy food, is metabolized by gut microbiota to equol, which possesses higher antioxidant activity and affinity for estrogen receptors than the parent compound. The neuroprotective flavone glycoside baicalin (baicalein 7-O-glucuronide) can only be absorbed after hydrolysis by gut microbiota β-glucuronidase to the aglycone form baicalein [44]. Similarly, the neuroprotective anthocyanin cyanidin-3-O-glucoside displays a poor bioavailability, while its microbiota degradation products are more easily absorbable [45]. Ellagic acid and ellagitannins are degraded by intestinal microorganisms to form urolithins, molecules characterized by higher bioavailability and better anti-inflammatory and antioxidant properties than their compounds of origin [46].

4. Polyphenols: Mechanisms of Action

The major mechanism of natural polyphenols in preventing stroke relies on their protective action on the cardiovascular system [34,47,48].
Many polyphenols are endowed with anticoagulant and antiplatelet activities, potentially contributing to the prevention of thrombus formation, the main cause of ischemic stroke [49,50]. For example, several coumarin derivatives exert anticoagulant properties by inhibiting the vitamin K epoxide reductase complex and are widely used as clinical anticoagulant agents [51]. Among the polyphenols endowed with antiplatelet activity, the isoflavones genistein and daidzein possess a marked and physiologically relevant cyclooxygenase-1 (COX-1) inhibitory activity [52]. Other flavonoids with antiaggregant effects, including the isoflavone tectorigenin, have been reported to act as antagonists on thromboxane receptors [52,53].
Hypertension, a long-term medical condition affecting millions of individuals worldwide [54], is an important risk factor in particular for ICH and SAH subtypes of stroke [55,56]. Dietary intake of flavonoids belonging to anthocyanin, flavone and flavan-3-ol subclasses may contribute to the prevention of hypertension [57,58]. The underlying biological mechanisms by which polyphenols regulate blood pressure include vasodilation through the regulation of nitric oxide (NO) and endothelium-derived hyperpolarizing factor (EDHF) [47,57,58].
Besides the well-documented beneficial effects of polyphenols on cardio- and cerebrovascular systems, a growing number of studies in cellular and animal stroke models indicates a direct protective effect of many polyphenols on the brain. Notably, several polyphenols exert neuroprotective actions in preclinical models even when administered after stroke induction, indicating that these molecules may be useful not only in increasing resilience to brain damage, but also for the recovery of patients suffering from stroke. Moreover, the fact that different polyphenolic compounds act on the same molecular pathways raises the possibility that they may promote synergistic effects at very low doses. Therefore, the possible synergistic effect between polyphenols with each other or with other compounds may provide the rationale to overcome the limitations caused by the poor bioavailability of these molecules.
At the mechanistic level, polyphenols exert their neuroprotective benefits by acting on several targets simultaneously. These compounds are generally strong antioxidants, working as reactive oxygen species (ROS) scavengers and metal chelators due to the presence of hydroxyl groups and neutrophilic centers [33]. Furthermore, many polyphenols are able to activate transcription factors involved in antioxidant-responsive element pathways, such as erythroid 2-related factor 2 (Nrf2), thus promoting the expression of antioxidant enzymes including superoxide dismutase (SOD), heme oxygenase-1 (HO-1), catalase, glutathione reductase and glutathione-S-transferase [59].
Apoptosis is a process that can play a primary role in various pathologies, including cardiovascular diseases and stroke [60]. Many polyphenols are able to interact with proteins regulating apoptosis, including proapoptotic (Bax, Bad) and antiapoptotic (Bcl-2, Bcl-XL) members of Bcl-2 family, p53, mitogen-activated protein kinases (MAPKs) and protein kinase B (AKT) [60]. These compounds can act as pro- or antiapoptotic agents, depending on their concentrations, cellular system and stage of pathological process [60].
The polyphenol-mediated neuroprotection not only involves a direct effect on neurons, but also modulatory effects on different inflammation players in the brain, including microglia and mast cells (MCs) [61,62]. The anti-inflammatory properties of polyphenols are based on their capability to interfere with immune cell regulation, inflammatory gene expression and the synthesis of inflammatory mediators [63]. For example, a number of polyphenols have been shown to modulate nuclear factor kappa-light-chain-enhancer of activated B cells (NF-kB), toll-like receptor (TLR) and arachidonic acid pathways, suppressing the production of tumor necrosis factor α (TNF-α), interleukin (IL)-1β, IL-6, IL-1 and IL-8, as well as cyclooxygenase-2 (COX-2), inducible nitric oxide synthase (iNOS) and nitric oxide (NO) [63].
Epigenetic modifications, including DNA methylation, histone modifications and RNA-based mechanisms, modify gene expression without altering the DNA sequence. Epigenetic modifications regulate important physiological processes in living organisms, but they have also been associated with the pathogenesis of various diseases, including stroke [64]. Various polyphenols can influence epigenetic mechanisms underlying stroke pathogenesis and progression by modulating DNA methylation and histone modifications through the interaction with histone deacetylases (HDACs) and DNA methyltransferases (DNMTs) [65].
The hypothesized mechanisms of action of individual polyphenols active in models of stroke are discussed in the following section.

5. Polyphenols and Stroke: Results from Preclinical Stroke Models

Isolated polyphenols active in preclinical stroke models are described below. In this review, polyphenols have been clustered according to their chemical structure. The major part of polyphenols exists in plants as glycosides, where different sugars are bonded to the polyphenolic structure in different positions. For simplicity, we have classified polyphenols according to the chemical structures of the aglycones.

5.1. Flavonoids

5.1.1. Flavonols

Flavonols, the most ubiquitous class of flavonoids in foods, display antioxidant and anti-inflammatory properties [66]. They are present mainly in onions, kale, leeks and broccoli, but also in red wine, tea and fruits.
Quercetin is a plant flavonol widely distributed in nature; common sources of quercetin are red onions and kale. The molecule is a strong antioxidant and anti-inflammatory agent and displays potential protective properties against hypertension and ischemic heart disease in both animal models and humans [67,68]. Notably, quercetin and its glycosides isoquercetin (3-O-glucoside) and rutin (3-O-rutinoside) have been found to promote beneficial effects in various models of brain ischemia [69,70,71,72,73,74], ICH [75,76] and SAH [77,78]. The health effect of quercetin could be attributed to its antioxidant, antiapoptotic and anti-inflammatory actions, combined with a protective effect on BBB through the inhibition of metalloproteinase (MMP) activity [70,72,73,74,75,76,77]. It has been proposed that the antioxidant and antiapoptotic activity of the molecule may be mediated by the activation of Nrf2 factor [69].
Myricetin can be found in various medicinal herbs, vegetables (tomatoes) and fruits. The molecule has been extensively studied for its multiple pharmacological activities, including antiapoptotic, anti-inflammatory and antioxidant properties [79]. Studies have revealed that myricetin acts against ischemic damage in both oxygen–glucose deprivation (OGD) cellular model [80] and in rats subjected to transient middle cerebral artery occlusion (tMCAO) [80,81]. Moreover, myricetin reduced endothelial permeability in human brain microvascular endothelial cells (BMVECs) subjected to OGD, suggesting a beneficial role in maintaining BBB function [82]. Among the mechanisms proposed to explain myricetin-induced protection, there are the inhibition of p38 MAPK and the activation of AKT and Nrf2 factors [80,81,82].
Kaempferol is present in a variety of plants and fruits including Carthamus tinctorius (safflower), beans and broccoli [83]. A protective effect of kaempferol or its glycosides (kaempferol-3-O-rutinoside and kaempferol-3-O-glucoside) has been shown in tMCAO rats [84,85,86]. These studies suggest that postischemic treatment with kaempferol prevents neuroinflammation by decreasing activation of NF-kB/RelA and STAT3 [85,86].
Fisetin is a flavonol found in various fruits and vegetables, including apple, strawberry, persimmon and onion. The molecule has been reported to exert several beneficial effects, including neuroprotective activities [87]. For example, fisetin improved outcomes in the rabbit small clot embolism (SCE) model [88] or in rats subjected to permanent middle cerebral artery occlusion (pMCAO) [71] when administered the first minutes following ischemia. Notably, treatment of tMCAO mice with fisetin even 3 h after ischemia reduced infarct size and immune cell activation and infiltration [89].
Morin, a natural flavonol found in the branches of Morus alba (white mulberry) and other Chinese medicine plants, exhibits a wide spectrum of antioxidant and antiapoptotic activities [90]. Pre- and poststroke treatment with morin ameliorated brain damage, BBB leakage and neurological deficits in tMCAO rats by reducing oxidative stress, apoptosis and inflammation [91,92].

5.1.2. Flavan-3-ols

Flavan-3-ols are particularly abundant in tea plants, as well as in cocoas and chocolates. These polyphenols are considered primarily responsible for tea-promoted health benefits thanks to their cardioprotective and neuroprotective activities [93,94]. Flavan-3-ols are effective also in protecting against ischemic insults, as suggested by studies on green-tea-based supplements in preclinical models of stroke [95,96,97]. There are also numerous pieces of evidence indicating an antistroke activity for individual flavan-3-ols.
Epigallocatechin-3-gallate (EGCG), the predominant and most studied flavan-3-ol in green tea, induced preconditioning against OGD in a cellular model of brain ischemia [97]. EGCG has been shown to ameliorate cerebral damage and neurological deficits in different rodent species subjected to brain ischemia [98,99,100,101,102,103,104,105,106]. Notably, coadministration of EGCG with rtPA reduced the side effects of delayed rtPA treatment in a rat tMACO model, suggesting a potential clinical use of EGCG as an adjuvant in stroke therapies [107]. Multiple mechanisms have been proposed as mediating the protective effect of the molecule, including suppression of MMP activation [98], antioxidant effects [100,103,104] (possibly through activation of Nrf2 [104]), attenuation of inflammation [102] and reduction of apoptosis via modulation of phosphoinositide 3-kinase (PI3K)/Akt signaling [103]. Moreover, recent findings pointed out a beneficial effect of EGCG in cellular and animal models of SAH by targeting hemoglobin (Hb)-induced mitochondrial dysfunction [108,109,110].
The chemical structure of (-)-epicatechin-3-gallate (ECG) is similar to that of EGCG. The protective effect of ECG on cells subjected to OGD was consistent with EGCG [97]. Moreover, ECG protected human BMVECs against ischemic insult by promoting neovascularization and modulating apoptosis and autophagy [111].
(-)-Epicatechin (EC) is particularly abundant in cocoa, dark chocolate and green tea. The molecule exerted protective effects in cellular and animal models of brain ischemia, namely OGD-injured neurons and rats subjected to MCAO [112,113]. Similarly, EC was effective in ICH models, protecting cultured astrocytes against Hb-induced damage and mice from intracerebral hemorrhage [114,115]. The EC protection could be partly mediated by the activation of Nrf2 signaling [112,113,114,115].

5.1.3. Flavanones

Flavanones are present in citrus fruits and tomatoes and their juices. Compared to other flavonoids, flavanones show a weaker radical scavenging activity. However, these molecules can reduce oxidative stress by targeting the Nrf2/HO-1 axis [116]. The antioxidant and anti-inflammatory properties of the flavanones suggest a potential use of these molecules for the prevention or treatment of cardiovascular diseases [117]. Various studies also support a role of flavanones against stroke.
Eriodictyol is a compound isolated from the Chinese herb Dracocephalum rupestre and citrus fruits [118]. The treatment of mice with eriodictyol prevented neuronal death, reduced infarct area and neuroinflammation and improved neurological and memory deficits induced by ischemic insult [119]. Eriodictyol-7-O-glucoside was effective in protecting astrocytes against OGD-induced ischemia and in reducing brain damage and neurological deficits in rats subjected to tMCAO by targeting Nrf2 signaling [120].
Naringenin is a natural flavanone endowed with neuroprotective properties [121]. The molecule protected culture neurons against hypoxic injury by reducing oxidative stress and mitochondrial dysfunction via activation of Nrf2 signaling [122,123]. Interestingly, naringenin nanoparticles rescued human mesenchymal stem cells (MSCs) from OGD-mediated stress, suggesting a potential use of the molecule as a strategy to improve MSC-based strategy against stroke [124]. Naringenin also showed a neuroprotective profile in an animal model of brain ischemia, reducing apoptosis, inflammation, oxidative stress and neurological deficits through the modulation of claudin-5, MMP9, Nrf2, nucleotide oligomerization domain-like receptor 2 (NOD2) and NF-κB [123,125,126]. Finally, diet supplementation of naringin, the naringenin-7-O-glycoside, prevented cerebral thrombogenesis in pial microvessels of stroke-prone spontaneously hypertensive rats [127].
Hesperetin and its glycoside hesperidin (hesperetin-7-O-glycoside) can be isolated from the rinds of some citrus species. Preclinical studies have shown the potential of these molecules in the treatment of neurological and cardiovascular pathologies [128]. Pretreatment with hesperetin ameliorated functional and histological outcomes in an MCAO rat model [129]. Similarly, hesperidin nanoparticles reduced infarct volume, inflammatory cytokines and oxidative stress in rats subjected to two-vessel occlusion (2VO, also called bilateral common carotid artery occlusion (BCCAO)) [130]. As observed for naringenin, hesperidin consumption through the diet reduced thrombotic tendency in stroke-prone spontaneously hypertensive rats [127], raising the possibility that daily ingestion of these flavonoids could promote an antithrombotic effect. Hesperidin could also be useful in treating cerebral vasospasm, as recently suggested by its beneficial effects on vessel walls and luminal diameters in an SAH rat model [131].
In recent years, there has been a growing interest in the neuroprotective agent pinocembrin, a flavanone particularly abundant in propolis [132]. Pinocembrin promoted protection against ischemic stress in neurons, BMVECs or in a cellular BBB model subjected to OGD [133,134,135]. The compound was effective also in different animal models of brain ischemia [134,136,137,138]. Notably, pinocembrin pretreatment extended the therapeutic time window of rtPA treatment in a rat model of brain ischemia [135]. Possible mechanisms for pinocembrin activity in preclinical models of brain ischemia include inhibition of MMPs, BBB protection, autophagy modulation, inhibition of apoptosis and inflammatory cascade [133,134,135,136,137,138]. Additionally, pinocembrin improved early outcomes in an ICH mouse model by inhibiting toll-like receptor 4 (TLR4) and modulating microglia activation [139].

5.1.4. Flavones

Flavones are a class of polyphenols widely distributed in the plant kingdom, including several vegetables and fruits that are components of the human diet [140]. Different flavones have been explored for neuroprotection in preclinical models.
Luteolin is a flavone found in different vegetables, fruits and medicinal herbs including celery, carrots and broccoli [141]. Luteolin and its glycosides orientin (8-C-glucoside) and luteoside (7-O-glucoside) have been reported to exert beneficial effects in cellular and animal models of neonatal hypoxic–ischemic brain injury (NHIBI) [142], brain ischemia [143,144,145,146,147] and ICH [148,149]. The molecule engages several mechanisms of action that could play a role in its antistroke action, including antiapoptotic activity, BBB stabilization by claudin-5 upregulation and MMP9 inhibition, reduction of oxidative stress and autophagy enhancement through the activation of Nr2 pathway, MCs modulation, reduction of inflammation via activation of PPARγ and downregulation of TLR4/NF-κB pathway [144,145,146,147,148,149].
Apigenin is a natural flavone present in vegetables and fruits such as celery, parsley, tea, onion and grapefruit [150]. Recent findings suggested that the molecule could also alleviate brain damage and ameliorate poststroke neurological and cognitive deficits in brain ischemia [151,152] and SAH [153] models. Apigenin protective effects seem to rely on multiple mechanisms involving the promotion of angiogenesis via caveolin-1/vascular endothelial growth factor (VEGF) pathway and reduction of TLR4-mediated inflammation [151,153,154]. The effect of apigenin in alleviating poststroke cognitive impairment was found to involve the epigenetic induction of brain-derived neurotrophic factor (BDNF) through HDAC inhibition [152].
Nobiletin is a flavone extracted from the peel of citrus fruits. The molecule is endowed with several beneficial properties, including neuroprotective activities [155]. Nobiletin also exerted neuroprotective action in animal models of stroke, reducing cerebral apoptosis and inflammation and improving learning and memory deficits following ischemia [156,157].
Tangeretin is a citrus flavone endowed with neuroprotective properties [158]. The molecule was able to protect against OGD insult by preventing activation of proapoptotic c-Jun N-terminal kinase (JNK) signaling [159]. Both nobiletin and tangeretin improved the viability of the human hepatocellular carcinoma cells (HepG2) under hypoxic conditions [160]. The two molecules have been identified as the active compounds in the extract of Pericarpium aurantii, the immature fruit of Citrus aurantium, able to attenuate brain pathology in tMCAO rats [160].
Baicalein and its glycoside baicalin are flavones extracted from Scutellaria baicalensis, a traditional Chinese herb. Over the years, a growing body of evidence has pointed out the neuroprotective and antistroke activities of these molecules [161,162]. Baicalin reduced OGD-mediated neurotoxicity in the human neuroblastoma cell line SH-SY5Y [163]. Injected after the onset of stroke, this molecule reduced infarct size and improved neurological deficits in tMCAO rats [163,164]. Additionally, poststroke baicalin administration alleviated brain damage in a gerbil model of global cerebral ischemia by antioxidative and antiapoptotic mechanisms [165].
Acacetin is a naturally occurring flavone known for its numerous pharmacological activities, including neuroprotective and anti-inflammatory properties [166]. Recent findings indicate that acacetin administration to mice subjected to tMCAO promoted neuroprotection, possibly via inhibition of microglial activation and Nod-like receptor family, pyrin domain containing 3 (NLRP3) inflammatory signaling [167].

5.1.5. Isoflavones

Isoflavones are found primarily in legumes, including soybeans and chickpeas, but also in fruits such as raisins and currants. In the human diet, the main sources of isoflavones are soybeans and soy food, which contain mainly genistein and daidzein [168,169]. For their chemical properties, isoflavones have been investigated as a therapy for cardiovascular and cerebrovascular pathologies [168,169].
A variety of preclinical studies support a beneficial role of the isoflavone genistein in protecting the brain against stroke (for complete reviews, see [170,171]). Briefly, the molecule promoted neuroprotection in cellular cultures subjected to OGD [172,173], in global and focal in vivo models of brain ischemia [174,175,176,177,178,179,180,181,182,183,184,185,186,187] and in animal models of SAH [188,189]. Genistein engages several mechanisms of action, including reduction of oxidative stress, inflammation and apoptosis through the promotion of antiapoptotic and growth factors such as Bcl2 and Nrf2 [170,171].
The isoflavone daidzein reduced injury and/or enhanced functional recovery in rats or mice subjected to MCAO when administered not only before but also after stroke onset [190,191,192]. It has been suggested that the beneficial effect of the molecule could rely on its capability to decrease oxygen free radical production [192] and to promote cholesterol homeostasis, a crucial process in injury-induced synaptic remodeling [191]. The daidzein glycoside puerarin (daidzein-8-C-glucoside), the main active compound of Pueraria lobata, has been extensively investigated for its beneficial pharmacological properties [193]. The compound was found to express protective effects in different cellular and animal models of stroke, including NHIB [194], brain ischemia [195,196,197,198,199,200] and SAH [201]. Puerarin protection may be mediated by its antiapoptotic and antioxidant activity through the modulation of factors and signaling pathways such as SOD, PI3K/Akt, MAPK and NF-κB [193].
Biochanin A, a natural isoflavonoid phytoestrogen derived from red clover or chickpea, displays a broad range of pharmacological functions, including neuroprotective activities [202]. The molecule was effective in alleviating brain damage and symptoms of rodents subjected to tMCAO [203,204] and SAH [205]. The activation of the protective factors glutamate oxaloacetate transaminase (GOT), SOD and Nrf2 and the inhibition of the NF-kB pathway may contribute to the neuroprotective effects of biochanin A [203,204,205].

5.1.6. Anthocyanins

Anthocyanins are strongly pigmented compounds present in brightly colored fruits and vegetables. For their multiple pharmacological properties, anthocyanins have been investigated for the prevention or treatment of different diseases, including stroke [206].
The cyanidin glycoside cyanidin-3-O-glucoside is one of the most common anthocyanins and accounts for >95% of the total anthocyanin in Chinese bayberry (Myrica rubra) [207]. Recent findings pointed out a protective activity of cyanidin-3-O-glucoside in a mouse model of brain ischemia through the modulation of TLR4, NF-κB, Nrf2 and NLPR3 [208]. Similarly, cyanidin-3-O-glycosides protected PC12 cells against OGD-induced injury and provided beneficial effects in rodent models of brain ischemia by modulating antioxidant factors such as HO-1 [209,210,211]. Furthermore, cyanidin promoted protection in pial microcirculation of a rat model of global ischemia [212]. The protective effects on BBB integrity were mediated by arteriolar vasodilation via NO release and reduction of ROS levels [212].

5.1.7. Dihydrochalcones

A few dihydrochalcones, a family of bicyclic flavonoids, have shown antioxidant properties [213]. In particular, phloretin, a dihydrochalcone abundant in apples and apple-derived products, displayed a neuroprotective effect by activating the Nrf2 pathway in rats subjected to tMCAO [214].

5.1.8. Proanthocyanidins

Proanthocyanidins are oligomers or polymers of monomeric flavan-3-ols, particularly catechin and epicatechin. They are present in several plants, including apples, pine barks, cinnamon, blueberry and green and black tea. Proanthocyanidins are known for their cardioprotective and neuroprotective properties [215] and have also been investigated in stroke models.
For example, grape seed proanthocyanidin extract (GSPE) pretreatment alleviated brain damage in an NHIBI model, possibly through its antiapoptotic activity [216].
Procyanidins, composed of (-)-epicatechin units, significantly attenuated BBB disruption and neurological deficits in rodent brain ischemia models [217,218]. The neuroprotection was paralleled by a reduction of apoptosis and oxidative stress and an increase in angiogenesis [217,218]. Of note, apple polyphenols, which contains approximately 64% of procyanidins, prevented the formation of cerebral vasospasm in a rabbit model of SAH [219], suggesting a possible role for procyanidins in the treatment of this disease.

5.2. Phenolic Acids

5.2.1. Hydroxycinnamic Acids

Hydroxycinnamic acids are a group of dietary phenolic compounds derivatives of cinnamic acid found abundant in cereals, legumes, oilseeds, vegetables and various beverages. They are present as four basic molecules, namely caffeic acid, ferulic acid, sinapic acid and p-coumaric acid [220].
Caffeic acid is found in thyme, sage, spearmint, sunflower seeds, yerba mate, coffee, wine, olive oil and various spices. Its derivative caffeic acid phenethyl ester (CAPE) is an active component of propolis. Thanks to their antioxidant, anti-inflammatory, cardioprotective and neuroprotective activities, these compounds have been investigated as antistroke agents [221,222]. CAPE administered either before or after ischemia reduced neonatal brain injury in an NHIBI rat model by inhibiting apoptosis and inflammation [223]. Pre- or postischemia administration of caffeic acid and CAPE also exerted protective effects in mature brains, improving outcomes in various in vivo models of brain ischemia [224,225,226,227,228]. The neuroprotection provided by these molecules was likely to be mediated through their antioxidant and anti-inflammatory actions and the inhibition of 5-lipoxygenase [224,225,226,227,228].
Ferulic acid is commonly present in the leaves, fruits and seeds of many plants such as rice, wheat, oats and giant fennel. The molecule is known for its multiple biological activities, including antioxidant, anti-inflammatory and antithrombotic actions [229]. Ferulic acid exerted protective effects in a cellular model of brain ischemia as well as in animal models of global and focal cerebral ischemia [230,231,232,233,234]. The neuroprotective effect of ferulic acid could be mediated by the anti-inflammatory and neurotrophic actions promoted by the reduction of intercellular adhesion molecule-1 (ICAM-1) and increase in brain levels of erythropoietin (EPO) and granulocyte colony-stimulating factor (G-CSF) [232,233,234].
Sinapic acid and its derivatives are orally available compounds found in spices, citrus and berry fruits, vegetables and cereals [235]. Postischemia sinapic acid treatment reduced neuronal damage and memory deficits in the four-vessel occlusion (4VO) rat model of global cerebral ischemia [236].
The main sources of p-coumaric acid are tea, coffee, wine, beer and various vegetables and fruits. Many studies have shown the beneficial properties of p-coumaric acid, including neuroprotective and anti-inflammatory effects [237]. The compound promoted neuroprotection in animal models of focal [238] and global brain ischemia [239] by hampering ROS production and apoptosis.
Chlorogenic acid, a caffeic acid–quinate conjugate, is a major component of coffee, tea and several fruits or vegetables. Studies on chlorogenic acid suggest that it may promote neuroprotection against stroke through multiple effects [240]. Chlorogenic acid (and its metabolite dihydrocaffeic acid) administered either before or after ischemia reduced brain infarct volume, BBB damage and behavioral deficits in tMCAO rats by blunting MMP activation and increasing brain levels of EPO, HIF-1a and nerve growth factor (NGF) [240,241,242]. Moreover, the compound promoted neuroprotection in rats subjected to 2VO by regulating the Nrf2 pathway [243]. Of note, the combination of chlorogenic acid with rtPA was effective in reducing behavioral deficits in the rabbit SCE model, extending the therapeutic time window for rtPA administration [244].
Rosmarinic acid is a natural antioxidant hydroxycinnamate commonly found in Lamiaceae and Boraginaceae plant families, including rosemary, sage, basil, thyme and peppermint [245]. The compound protected SH-SY5Y cells against OGD-induced cell death [246,247]. Pre- or postischemia administration of rosmarinic acid alleviated brain injury and memory impairment in MCAO animal models through the modulation of Nrf2, HO-1 and synaptophysin [248,249]. Moreover, rosmarinic acid protected diabetic rats against ischemic assault by attenuating BBB breakdown even when administered 5 h after the stroke onset [247]. The protective effects of the molecule may involve an anti-inflammatory action through the modulation of high-mobility group box1 (HMGB1) and the NF-κB signaling pathway [247].

5.2.2. Hydroxybenzoic Acids

Recent findings suggest that gallic acid, a benzoic acid found in tea leaves and red wine, could play a protective role against stroke [250]. Prestroke treatment with gallic acid promoted neuroprotection in cellular and animal models of cerebral ischemia [251,252]. Interestingly, administration of the molecule before stroke onset mitigated brain injury and behavioral deficits in a rat model of global brain ischemia exposed to particulate matter [253]. Furthermore, administration of gallic acid and its derivatives reduced depressive symptoms and oxidative stress in a mouse model of poststroke depression [254].

5.3. Stilbenes

Stilbenes are part of a vast group of natural defense compounds occurring in many plants. These molecules are endowed with a wide range of beneficial activities, including the capability to protect against oxidative stress [255].
Among stilbenes, resveratrol is by far the most widely studied for its beneficial properties [256,257]. Pre- or poststroke treatment with resveratrol or its glucoside derivative polydatin showed beneficial effects in a large number of models of stroke (for complete reviews, see [258,259]). Recent examples of stroke models that benefited from the actions of resveratrol include NHIBI models [260,261,262], cellular cultures challenged with OGD [263,264,265], rodents subjected to MCAO [264,266,267,268,269] and ICH and SAH rodent models [270,271,272,273,274]. The potential mechanisms of action underlying the effects of resveratrol against stroke are numerous since the molecule interacts with a wide range of enzymes and receptors and promotes the expression of several factors devoted to enhancing cellular stress resistance and reducing apoptosis [259]. It has been proposed that beneficial properties of resveratrol could be mediated by its modulatory action on sirtuins and AMP-activated kinase (AMPK), a serine/threonine kinase known to be a key metabolic and stress sensor/effector [65,264,266,267].

5.4. Curcuminoids

Curcuminoids consist of curcumin and its derivatives, molecules found in the rhizome of turmeric (Curcuma longa). Curcumin possesses several beneficial properties that could make it a suitable candidate for stroke prevention or treatment, including anti-inflammatory, antilipemic, antiaggregant, neuroprotective and epigenetic modulatory activities [275]. Pre- and posthypoxia treatment with curcumin was found to effectively promote neuroprotection in rat neurons challenged with OGD [276], in neonatal mice subjected to hypoxic–ischemic brain injury [277], in MCAO rodents [278,279,280,281,282,283,284] and in stroke-prone spontaneously hypertensive rats [285].

5.5. Lignans

Lignans are a large group of polyphenolic compounds present in good quantity in various plants, including flax and sesame seeds, and whole bran cereals [286]. Some dietary lignans have been suggested to have potential in the prevention of cardiovascular disease and, possibly, stroke [287,288].
For example, administration of pinoresinol, a lignan found in sesame feed, Brassica vegetables and olive oil, prevented pial circulation damage induced by 2VO in rats by reducing oxidative stress [289].

5.6. Ellagitannins and Ellagic Acid

Ellagitannins and ellagic acid are polyphenols present in different fruits, including pomegranates, strawberries, black raspberries, raspberries, walnuts and almonds. In vivo, ellagitannins are hydrolyzed to ellagic acids which in turn are metabolized by intestinal microbiota to different types of urolithins [46]. Some molecules of this polyphenol class show neuroprotective properties against stroke.
Ellagic acid protected against ischemic injury in both cellular and in vivo models of brain ischemia by regulating Bcl-2/Bax expression [290].
Among urolithins, urolithin A has been shown to mitigate OGD-induced damage in N2a neuroblastoma cells and primary neurons and to reduce ischemic brain injury in mice by inhibiting endoplasmic reticulum (ER) stress [291].
Punicalagin, a natural ellagitannin found at high concentration in pomegranates, reduced infarct volume and neurological deficits in rats subjected to tMCAO through its antioxidant, anti-inflammatory and antiapoptotic properties [292,293].

5.7. Coumarins

Coumarin and coumarins, its derivatives, can be found in many plants, including tonka beans (where they are present in high concentration), Cinnamon cassia and cherry blossom of the genus Prunus [294]. The use of coumarins in the prevention of ischemic stroke is based on the anticoagulant or antiplatelet effects exhibited by many of these molecules [295,296,297,298]. Recent findings indicate that coumarins, besides their cardiovascular effects, may promote neuroprotection in preclinical models of stroke.
For example, auraptene, a citrus coumarin endowed with anti-inflammatory properties [299], acts as neuroprotective agent in the 2VO mouse model of brain ischemia by inhibiting inflammation [300,301].
Umbelliferone, a natural coumarin derivative with antioxidant and scavenging properties [302], ameliorated neurological outcomes and brain injury in tMCAO rats partly through the inhibition of thioredoxin-interactive protein (TXNIP)/NLRP3 inflammasome and activation of peroxisome proliferator-activated receptor-γ (PPAR-γ) [303].
Esculetin, another natural coumarin compound studied for its antioxidant and anti-inflammatory activities [304], promoted neuroprotection in mice subjected to tMCAO via upregulation of Bcl-2 and downregulation of Bax, two proteins involved in apoptosis [305].
Imperatorin is a naturally occurring coumarin that over the years has gained increasing interest for its health properties [306]. The compound exerted a protective activity in the SH-SY5Y neuroblastoma cell line subjected to OGD and in a rat model of brain ischemia [307]. The beneficial effect of imperatorin was associated with a reduction of apoptosis and upregulation of BDNF [307].
Scopoletin, a coumarin compound used in traditional Chinese medicine, has been studied for its antioxidant and anti-inflammatory properties [308]. A recent study pointed out a neuroprotective effect of scopoletin in a rat model of brain ischemia [309].
Osthole is a natural coumarin derivative isolated in several medicinal plants [310]. Various studies support the neuroprotective abilities of osthole in in vivo models of brain ischemia by hampering apoptosis [311,312,313].
Similar to other coumarins, daphnetin, a natural compound extracted from medicinal herbs, has been reported to show multiple beneficial properties [302]. Daphnetin provided neuroprotection in animal models of cerebral ischemia in both immature and adult brains by reducing inflammatory cytokine production and neuronal apoptosis [314,315].
The polyphenols active in preclinical models of stroke, whether the treatment was started before, during or after stroke induction (pre, concomitant or post, respectively), effective polyphenol concentration/dose and length of the treatment, and polyphenol major effects are reported in Table 1.

6. Polyphenols and Stroke: Results from Human Studies

Clues about the relationship between dietary consumption of polyphenols and beneficial effects on human health come from epidemiological studies.
In general, the inverse association between intake of high-polyphenol content food (e.g., fruits and vegetables) and risk of stroke appears clear, even though the role of polyphenols in this protection is still debated [316]. Moreover, fruit-derived polyphenol supplementation has been shown to improve cognitive and functional recovery of ischemic stroke patients [317].
The association between the consumption of beverages rich in polyphenols (i.e., wine, beer, coffee and tea) and stroke has been investigated. Although alcohol consumption at high intakes is detrimental even when occasionally consumed, a moderate intake of wine and beer has been associated with a lower risk of cardiovascular disease and ischemic stroke (for a review, see [318]). The protective effects related to wine and beer consumption have been attributed not only to ethanol itself but also to nonalcoholic components, mainly polyphenols [319,320]. Moreover, moderate consumption of coffee has been suggested to reduce the incidence of cardiovascular diseases [321]. Conversely, the association between coffee intake and stroke is under debate, with some studies indicating the beneficial effect of this beverage and others showing positive or no associations [322]. Dose–response analyses of tea intake indicate that high consumption of green tea was related to a reduced risk of both ischemic and hemorrhagic stroke [323].
Another typically polyphenol-rich food is cocoa, mostly consumed as chocolate in Western countries. Although few prospective studies on chocolate and stroke exist, the available data suggest that chocolate consumption could reduce the risk of coronary heart disease and stroke [324].
An important source of polyphenols is represented by culinary spices and herbs that, besides their use in cooking to add flavor to food dishes, are also employed in traditional medicines to prevent or treat different conditions. The effectiveness of traditional herbal medicine in stroke prevention and treatment has been reported in a variety of preclinical and clinical studies. However, solid conclusions about the relationship between spices/herbs and the risk of stroke cannot be made due to the methodological gaps of many investigations [325].
Scientific evidence about the antistroke role of food containing specific polyphenol classes or isolated polyphenolic compounds is still scarce, although the results from the available studies are generally promising.
Results from a meta-analysis considering 11 prospective cohort studies suggested that high dietary intake of flavonoids may moderately reduce the risk of stroke [326].
Different prospective studies also examined the relationship between dietary flavonoid subclasses and stroke. Higher dietary flavonol intake has been associated with a reduced risk for stroke [327,328]. In a cohort study, the consumption of food rich in quercetin was associated with a decreased risk of thrombotic stroke [329]. The quercetin metabolite 4-methylcatechol displayed a relevant antiplatelet activity in human blood, supporting a possible use of this molecule in the prevention of thrombotic stroke [330]. Moreover, a meta-analysis of randomized controlled trials indicated a significant effect of quercetin supplementation in reducing blood pressure [331]. In a double-blind randomized clinical trial among stroke patients, fisetin was found to prolong the therapy window of rtPA treatment, likely by reducing levels of MMPs and C-reactive protein (CRP) [332].
A similar effect in extending the rtPA therapy window was observed in a clinical trial involving the use of EGCG [333]. In this study, the beneficial effect of the polyphenolic compound could also be attributed to the reduction of plasma levels of MMPs [333]. The finding that pyrogallol, a human metabolite of EGCG from green tea, inhibited platelet aggregation in human blood [330] suggests that inhibition of platelet formation could play a role in the protective effect of green tea against brain ischemia.
The increased intake of flavanones has been associated with diminished risk of brain ischemia in women [334]. The flavanone pinocembrin has been approved by the China Food and Drug Administration for the treatment of ischemic stroke, and it is currently under phase II clinical trial [132].
In an observational study, administration of luteolin in combination with the lipid amide palmitoylethanolamide (PEA) promoted clinical improvement in stroke patients, when compared with literature data of patients with similar pathological conditions that did not receive pharmacological treatment [335].
The frequent intake of soybeans and soy food like soymilk or tofu, containing high levels of genistein and other isoflavones, has been associated with a reduced risk of stroke in Japanese and Chinese populations [336,337]. The antiplatelet potential of many isoflavones might contribute to the cerebrovascular protection provided by a soy-rich diet. In support of this, the isoflavones genistein and tectorigenin showed a strong antiplatelet effect when tested in human blood [53]. The isoflavone puerarin is an important component of traditional Chinese medicine. A recent meta-analysis of 35 randomized controlled trials on the effect of puerarin injections in acute cerebral ischemia suggests a possible clinical use of the compound against stroke [338]. However, due to the poor methodological quality of some of the studies, further clinical trials are needed to verify the safety of the drug [338].
Different clinical studies have indicated that consumption of food rich in anthocyanins, including blueberries and cranberries, improves cerebral blood flow in healthy adults (for a review, see [206]). However, a recent meta-study on 19 prospective cohorts demonstrated that there was no relationship between consumption of anthocyanins and different types of strokes, although that dietary intake of anthocyanins was inversely correlated with the risk of cardiovascular diseases [339].
Some recent cohort studies indicated an inverse association between phenolic acids (hydroxybenzoic and hydroxycinnamic acids in particular) and cardiovascular diseases and hypertension (for a recent review, see [340]). Different clinical trials confirmed a moderate effect of chlorogenic acid in reducing blood pressure in mild hypertensive adults [341]. To our knowledge, no clinical studies are available on the correlation between phenolic acids and stroke.
Long-term resveratrol supplementation in patients who suffered a stroke in the previous 12 months promoted a beneficial effect on blood pressure, body mass index and lipid profile, indicating a possible role of resveratrol as an adjuvant in the secondary prevention of stroke [342]. Moreover, resveratrol administration improved the outcomes of stroke patients receiving rtPA, suggesting that this polyphenol could serve as a potential adjuvant of rtPA therapies [343]. A positive correlation between resveratrol-promoted outcomes and reduction of plasma levels of MMPs was observed [343].

7. Conclusions

The incomplete list of polyphenols active in cellular and animal models presented here strongly supports the potential role of many classes of polyphenols against different types of stroke. Interestingly, many molecules also exert beneficial effects when administered after the stroke onset, suggesting that they could be exploited in the treatment of this pathology. Of note, some compounds were able to synergize with rtPA, indicating a possible use as coadjuvant in the current treatment of stroke.
Notwithstanding the body of positive preclinical findings, conclusive evidence from human studies is still lacking. Although the consumption of food rich in polyphenols is generally associated with positive health effects, including a lower incidence of cardiovascular disease and stroke, the effectiveness of isolated polyphenols in stroke treatment is still under debate.
Individuals consume polyphenols not as isolated compounds but rather as components of their overall daily diet. The variability in food composition and the difficulty in determining the accurate quantity of polyphenols in food, the potential modulation by food matrices and culinary techniques, the possible interactions between polyphenols with each other or with other food components and the effect of gut microbiota metabolism on the bioavailability of polyphenols, make studies in human population extremely challenging. In the future, long-term, large-scale and well-designed clinical trials are required to establish the effectiveness of the most promising polyphenols that have emerged from studies on preclinical models of stroke.
Finally, it will be important to further characterize the molecular targets of polyphenolic compounds potentially active against stroke. The identification of common molecular pathways targeted by different bioactive components may lead to the formulation of novel nutritional supplements, where different polyphenols could synergize at doses much lower than those of active individual compounds.

Author Contributions

Literature searching, draft preparation and manuscript editing, E.P.; figure preparation and manuscript editing, C.G.; figure preparation and manuscript editing, V.P.; manuscript editing, M.B.; supervision and manuscript editing, M.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by MIUR PNR 2015–2020 PerMedNet and ITN nEUROinflammation Grant agreement ID: 607962. V.P. Researcher Fellowship is covered by Fondazione Cariplo–Giovani Ricercatori—Research Support GR 2018-0391.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Faustino-Mendes, T.; Machado-Pereira, M.; Castelo-Branco, M.; Ferreira, R. The Ischemic Immature Brain: Views on Current Experimental Models. Front. Cell. Neurosci. 2018, 12, 277. [Google Scholar] [CrossRef] [Green Version]
  2. Nelson, K.B.; Lynch, J.K. Stroke in newborn infants. Lancet Neurol. 2004, 3, 150–158. [Google Scholar] [CrossRef] [Green Version]
  3. Kirton, A.; Deveber, G. Life after perinatal stroke. Stroke 2013, 44, 3265–3271. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. McNally, M.A.; Soul, J.S. Pharmacologic Prevention and Treatment of Neonatal Brain Injury. Clin. Perinatol. 2019, 46, 311–325. [Google Scholar] [CrossRef] [PubMed]
  5. Campbell, B.C.V.; Khatri, P. Stroke. Lancet 2020, 396, 129–142. [Google Scholar] [CrossRef]
  6. Zerna, C.; Thomalla, G.; Campbell, B.C.V.; Rha, J.H.; Hill, M.D. Current practice and future directions in the diagnosis and acute treatment of ischaemic stroke. Lancet 2018, 392, 1247–1256. [Google Scholar] [CrossRef]
  7. Hakim, A.M. Depression, strokes and dementia: New biological insights into an unfortunate pathway. Cardiovasc. Psychiatry Neurol. 2011, 2011, 649629. [Google Scholar] [CrossRef] [Green Version]
  8. Kalaria, R.N.; Akinyemi, R.; Ihara, M. Stroke injury, cognitive impairment and vascular dementia. Biochim. Biophys. Acta 2016, 1862, 915–925. [Google Scholar] [CrossRef] [Green Version]
  9. Petrea, R.E.; Beiser, A.S.; Seshadri, S.; Kelly-Hayes, M.; Kase, C.S.; Wolf, P.A. Gender differences in stroke incidence and poststroke disability in the Framingham Heart Study. Stroke 2009, 40, 1032–1037. [Google Scholar] [CrossRef] [Green Version]
  10. Feigin, V.L.; Lawes, C.M.M.; Bennett, D.A.; Anderson, C.S. Stroke epidemiology: A review of population-based studies of incidence, prevalence, and case-fatality in the late 20th century. Lancet Neurol. 2003, 2, 43–53. [Google Scholar] [CrossRef]
  11. Krishnamurthi, R.V.; Barker-Collo, S.; Parag, V.; Parmar, P.; Witt, E.; Jones, A.; Mahon, S.; Anderson, C.S.; Barber, P.A.; Feigin, V.L. Stroke Incidence by Major Pathological Type and Ischemic Subtypes in the Auckland Regional Community Stroke Studies: Changes Between 2002 and 2011. Stroke 2018, 49, 3–10. [Google Scholar] [CrossRef] [PubMed]
  12. Dirnagl, U.; Iadecola, C.; Moskowitz, M.A. Pathobiology of ischaemic stroke: An integrated view. Trends Neurosci. 1999, 22, 391–397. [Google Scholar] [CrossRef]
  13. Sacco, S.; Marini, C.; Toni, D.; Olivieri, L.; Carolei, A. Incidence and 10-year survival of intracerebral hemorrhage in a population based registry. Stroke 2009, 40, 394–399. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Sansing, L. Intracerebral Hemorrhage. Semin. Neurol. 2016, 36, 223–224. [Google Scholar] [CrossRef] [PubMed]
  15. Zhou, Y.; Wang, Y.; Wang, J.; Anne Stetler, R.; Yang, Q.W. Inflammation in intracerebral hemorrhage: From mechanisms to clinical translation. Prog. Neurobiol. 2014, 115, 25–44. [Google Scholar] [CrossRef] [PubMed]
  16. Macdonald, R.L.; Schweizer, T. Spontaneous subarachnoid haemorrhage. Lancet 2017, 389, 655–666. [Google Scholar] [CrossRef]
  17. Penn, D.L.; Witte, S.R.; Komotar, R.J.; Sander Connolly, E., Jr. The role of vascular remodeling and inflammation in the pathogenesis of intracranial aneurysms. J. Clin. Neurosci. 2014, 21, 28–32. [Google Scholar] [CrossRef]
  18. Kunz, A.; Dirnagl, U.; Mergenthaler, P. Acute pathophysiological processes after ischaemic and traumatic brain injury. Best Pract. Res. Clin. Anaesthesiol. 2010, 24, 495–509. [Google Scholar] [CrossRef]
  19. Chamorro, A.; Dirnagl, U.; Urra, X.; Planas, A.M. Neuroprotection in acute stroke: Targeting excitotoxicity, oxidative and nitrosative stress, and inflammation. Lancet Neurol. 2016, 15, 869–881. [Google Scholar] [CrossRef]
  20. Abdullahi, W.; Tripathi, D.; Ronaldson, P.T. Blood-brain barrier dysfunction in ischemic stroke: Targeting tight junctions and transporters for vascular protection. Am. J. Physiol. Cell Physiol. 2018, 315, C343–C356. [Google Scholar] [CrossRef]
  21. Yang, C.; Hawkins, K.; Doré, S.; Candelario-Jalil, E. Neuroinflammatory mechanisms of blood-brain barrier damage in ischemic stroke. Am. J. Physiol. Cell Physiol. 2019, 316, C135–C153. [Google Scholar] [CrossRef] [PubMed]
  22. Ribatti, D. The crucial role of mast cells in blood–brain barrier alterations. Exp. Cell Res. 2015, 3, 119–125. [Google Scholar] [CrossRef] [PubMed]
  23. Chen, S.; Yang, Q.; Chen, G.; Zhang, J.H. An update on inflammation in the acute phase of intracerebral hemorrhage. Transl. Stroke Res. 2015, 6, 4–8. [Google Scholar] [CrossRef] [PubMed]
  24. Moretti, A.; Ferrari, F.; Villa, R.F. Neuroprotection for ischaemic stroke: Current status and challenges. Pharmacol. Ther. 2015, 146, 23–34. [Google Scholar] [CrossRef]
  25. Fisher, M.; Saver, J. Future directions of acute ischaemic stroke therapy. Lancet Neurol. 2015, 14, 758–767. [Google Scholar] [CrossRef]
  26. Vespa, P.M.; Martin, N.; Zuccarello, M.; Awad, I.; Hanley, D.F. Surgical trials in intracerebral hemorrhage. Stroke 2013, 44, S79–S82. [Google Scholar] [CrossRef] [Green Version]
  27. Steiner, T.; Juvela, S.; Unterberg, A.; Jubg, C.; Forsting, M.; Rinkel, G. European Stroke Organization guidelines for the management of intracranial aneurysms and subarachnoid haemorrhage. Cerebrovasc. Dis. 2013, 35, 93–112. [Google Scholar] [CrossRef] [Green Version]
  28. Crozier, A.; Jaganath, I. Clifford MN. Dietary phenolics: Chemistry. bioavailability and effects on health. Nat. Prod. Rep. 2009, 26, 1001–1043. [Google Scholar] [CrossRef]
  29. Tsao, R. Chemistry and biochemistry of dietary polyphenols. Nutrients 2010, 2, 1231–1246. [Google Scholar] [CrossRef]
  30. Almeida, S.; Alves, M.; Sousa, M.; Oliveira, P.F.; Silva, B.M. Are Polyphenols Strong Dietary Agents Against Neurotoxicity and Neurodegeneration? Neurotox. Res. 2016, 30, 345–366. [Google Scholar] [CrossRef]
  31. Pérez-Jiménez, J.; Neveu, V.; Vos, F.; Scalbert, A. Identification of the 100 richest dietary sources of polyphenols: An application of the Phenol-Explorer database. Eur. J. Clin. Nutr. 2010, 64, S112–S120. [Google Scholar] [CrossRef] [PubMed]
  32. Fraga, C.G.; Croft, K.; Kennedy, D.O.; Tomas-Barberan, F.A. The effects of polyphenols and other bioactives on human health. Food Funct. 2019, 10, 514–528. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Tresserra-Rimbau, A.; Lamuela-Raventos, R.; Moreno, J.J. Polyphenols, food and pharma. Current knowledge and directions for future research. Biochem. Pharmacol. 2018, 156, 186–195. [Google Scholar] [CrossRef] [PubMed]
  34. Tressera-Rimbau, A.; Arranz, S.; Eder, M.; Vallverdú-Queralt, A. Dietary Polyphenols in the Prevention of Stroke. Oxid. Med. Cell. Longev. 2017, 2017, 7467962. [Google Scholar] [CrossRef] [Green Version]
  35. Pogačnik, L.; Ota, A.; Ulrih, N.P. An Overview of Crucial Dietary Substances and Their Modes of Action for Prevention of Neurodegenerative Diseases. Cells 2020, 9, 576. [Google Scholar] [CrossRef] [Green Version]
  36. D’Archivio, M.; Filesi, C.; Di Benedetto, R.; Gargiulo, R.; Giovannini, C.; Masella, R. Polyphenols, dietary sources and bioavailability. Ann. Ist. Super. Sanità 2007, 43, 348–361. [Google Scholar]
  37. Chimento, A.; De Amicis, F.; Sirianni, R.; Sinicropi, M.S.; Puoci, F.; Casaburi, I.; Saturnino, C.; Pezzi, V. Progress to Improve Oral Bioavailability and Beneficial Effects of Resveratrol. Int. J. Mol. Sci. 2019, 20, 1381. [Google Scholar] [CrossRef] [Green Version]
  38. Lopresti, A.L. The Problem of Curcumin and Its Bioavailability: Could Its Gastrointestinal Influence Contribute to Its Overall Health-Enhancing Effects? Adv. Nutr. 2018, 9, 41–50. [Google Scholar] [CrossRef] [Green Version]
  39. Lambert, J.D.; Sang, S.; Yang, C.S. Biotransformation of green tea polyphenols and the biological activities of those metabolites. Mol. Pharm. 2007, 4, 819–825. [Google Scholar] [CrossRef]
  40. Dey, P. Gut microbiota in phytopharmacology: A comprehensive overview of concepts. reciprocal interactions. biotransformations and mode of actions. Pharmacol. Res. 2019, 147, 104367. [Google Scholar] [CrossRef]
  41. Feng, W.; Ao, H.; Peng, C.; Yan, D. Gut microbiota, a new frontier to understand traditional Chinese medicines. Pharmacol. Res. 2019, 142, 176–191. [Google Scholar] [CrossRef] [PubMed]
  42. Wiciński, M.; Gębalski, J.; Mazurek, E.; Podhorecka, M.; Śniegocki, M.; Szychta, P.; Sawicka, E.; Malinowski, B. The Influence of Polyphenol Compounds on Human Gastrointestinal Tract Microbiota. Nutrients 2020, 12, 350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Márquez Campos, E.; Stehle, P.; Simon, M.C. Microbial Metabolites of Flavan-3-Ols and Their Biological Activity. Nutrients 2019, 11, 2260. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Noh, K.; Kang, Y.; Nepal, M.R.; Jeon, K.S.; Oh, D.G.; Kang, M.J.; Lee, S.; Kang, W.; Jeong, H.G.; Jeong, T.C. Role of Intestinal Microbiota in Baicalin-Induced Drug Interaction and Its Pharmacokinetics. Molecules 2016, 21, 337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Vitaglione, P.; Donnarumma, G.; Napolitano, A.; Galvano, F.; Gallo, A.; Scalfi, L.; Fogliano, V. Protocatechuic acid is the major human metabolite of cyanidin-glucosides. J. Nutr. 2007, 137, 2043–2048. [Google Scholar] [CrossRef] [PubMed]
  46. Landete, J. Ellagitannins, ellagic acid and their derived metabolites: A review about source, metabolism, functions and health. Food Res. Int. 2011, 44, 1150–1160. [Google Scholar] [CrossRef]
  47. Galleano, M.; Pechanova, O.; Fraga, C.G. Hypertension, nitric oxide, oxidants, and dietary plant polyphenols. Curr. Pharm. Biotechnol. 2010, 11, 837–848. [Google Scholar] [CrossRef]
  48. Grootaert, C.; Kamiloglu, S.; Capanoglu, E.; Van Camp, J. Cell Systems to Investigate the Impact of Polyphenols on Cardiovascular Health. Nutrients 2015, 7, 9229–9255. [Google Scholar] [CrossRef] [Green Version]
  49. Bojić, M.; Maleš, Ž.; Antolić, A.; Babić, I.; Tomičić, M. Antithrombotic activity of flavonoids and polyphenols rich plant species. Acta Pharm. 2019, 69, 483–495. [Google Scholar] [CrossRef] [Green Version]
  50. Bijak, M.; Sut, A.; Kosiorek, A.; Saluk-Bijak, J.; Golanski, J. Dual Anticoagulant/Antiplatelet Activity of Polyphenolic Grape Seeds Extract. Nutrients 2019, 11, 93. [Google Scholar] [CrossRef] [Green Version]
  51. Kasperkiewicz, K.; Ponczek, M.B.; Owczarek, J.; Guga, P.; Budzisz, E. Antagonists of Vitamin K-Popular Coumarin Drugs and New Synthetic and Natural Coumarin Derivatives. Molecules 2020, 25, 1465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Karlíčková, J.; Říha, M.; Filipský, T.; Macáková, K.; Hrdina, R.; Mladěnka, P. Antiplatelet Effects of Flavonoids Mediated by Inhibition of Arachidonic Acid Based Pathway. Planta Med. 2016, 82, 76–83. [Google Scholar] [CrossRef] [PubMed]
  53. Applová, L.; Karlíčková, J.; Říha, M.; Filipský, T.; Macáková, K.; Spilková, J.; Mladěnka, P. The isoflavonoid tectorigenin has better antiplatelet potential than acetylsalicylic acid. Phytomedicine 2017, 35, 11–17. [Google Scholar] [CrossRef] [PubMed]
  54. Forouzanfar, M.H.; Liu, P.; Roth, G.A.; Ng, M.; Biryukov, S.; Marczak, L.; Alexander, L.; Estep, K.; Abate, K.H.; Akinyemiju, T.F.; et al. Global Burden of Hypertension and Systolic Blood Pressure of at Least 110 to 115 mm Hg, 1990–2015. JAMA 2017, 317, 165–182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Woo, D.; Haverbusch, M.; Sekar, P.; Kissela, B.; Khoury, J.; Schneider, A.; Kleindorfer, D.; Szaflarski, J.; Pancioli, A.; Jauch, E.; et al. Effect of untreated hypertension on hemorrhagic stroke. Stroke 2004, 35, 1703–1708. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Pistoia, F.; Sacco, S.; Degan, D.; Tiseo, C.; Ornello, R.; Carolei, A. Hypertension and Stroke: Epidemiological Aspects and Clinical Evaluation. High Blood Press. Cardiovasc. Prev. 2016, 23, 9–18. [Google Scholar] [CrossRef] [PubMed]
  57. Cassidy, A.; O’Reilly, É.J.; Kay, C.; Sampson, L.; Franz, M.; Forman, J.P.; Curhan, G.; Rimm, E.B. Habitual intake of flavonoid subclasses and incident hypertension in adults. Am. J. Clin. Nutr. 2011, 93, 338–347. [Google Scholar] [CrossRef] [Green Version]
  58. Medina-Remón, A.; Estruch, R.; Tresserra-Rimbau, A.; Vallverdú-Queralt, A.; Lamuela-Raventos, R.M. The effect of polyphenol consumption on blood pressure. Mini Rev. Med. Chem. 2013, 13, 1137–1149. [Google Scholar] [CrossRef]
  59. Scapagnini, G.; Vasto, S.; Abraham, N.G.; Caruso, C.; Zella, D.; Fabio, G. Modulation of Nrf2/ARE pathway by food polyphenols: A nutritional neuroprotective strategy for cognitive and neurodegenerative disorders. Mol. Neurobiol. 2011, 44, 192–201. [Google Scholar] [CrossRef] [Green Version]
  60. Giovannini, C.; Masella, R. Role of polyphenols in cell death control. Nutr. Neurosci. 2012, 15, 134–149. [Google Scholar] [CrossRef]
  61. Silva, R.F.M.; Pogačnik, L. Polyphenols from Food and Natural Products: Neuroprotection and Safety. Antioxidants 2020, 9, 61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Parrella, E.; Porrini, V.; Benarese, M.; Pizzi, M. The Role of Mast Cells in Stroke. Cells 2019, 8, 437. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Yahfoufi, N.; Alsadi, N.; Jambi, M.; Matar, C. The Immunomodulatory and Anti-Inflammatory Role of Polyphenols. Nutrients 2018, 10, 1618. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Stanzione, R.; Cotugno, M.; Bianchi, F.; Marchitti, S.; Forte, M.; Volpe, M.; Rubattu, S. Pathogenesis of Ischemic Stroke: Role of Epigenetic Mechanisms. Genes 2020, 11, 89. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Pan, M.H.; Lai, C.S.; Wu, J.C.; Ho, C.T. Epigenetic and disease targets by polyphenols. Curr. Pharm. Des. 2013, 19, 6156–6185. [Google Scholar] [CrossRef] [PubMed]
  66. Wang, L.; Tu, Y.C.; Lian, T.W.; Hung, J.T.; Yen, J.H.; Wu, M.J. Distinctive antioxidant and antiinflammatory effects of flavonols. J. Agric. Food Chem. 2006, 54, 9798–9804. [Google Scholar] [CrossRef]
  67. Batiha, G.E.; Beshbishy, A.M.; Ikram, M.; Mulla, Z.S.; El-Hack, M.E.A.; Taha, A.E.; Algammal, A.M.; Elewa, Y.H.A. The Pharmacological Activity, Biochemical Properties, and Pharmacokinetics of the Major Natural Polyphenolic Flavonoid: Quercetin. Foods 2020, 9, 374. [Google Scholar] [CrossRef] [Green Version]
  68. Najmanová, I.; Pourová, J.; Mladěnka, P. A Mixture of Phenolic Metabolites of Quercetin Can Decrease Elevated Blood Pressure of Spontaneously Hypertensive Rats Even in Low Doses. Nutrients 2020, 12, 213. [Google Scholar] [CrossRef] [Green Version]
  69. Dai, Y.; Zhang, H.; Zhang, J.; Yan, M. Isoquercetin attenuates oxidative stress and neuronal apoptosis after ischemia/reperfusion injury via Nrf2-mediated inhibition of the NOX4/ROS/NF-κB pathway. Chem. Biol. Interact. 2018, 284, 32–40. [Google Scholar] [CrossRef]
  70. Khan, M.M.; Ahmad, A.; Ishrat, T.; Khuwaja, G.; Srivastawa, P.; Khan, M.B.; Raza, S.S.; Javed, H.; Vaibhav, K.; Khan, A.; et al. Rutin protects the neural damage induced by transient focal ischemia in rats. Brain Res. 2009, 1292, 123–135. [Google Scholar] [CrossRef]
  71. Rivera, F.; Urbanavicius, J.; Gervaz, E.; Morquio, A.; Dajas, F. Some aspects of the in vivo neuroprotective capacity of flavonoids: Bioavailability and structure-activity relationship. Neurotox. Res. 2004, 6, 543–553. [Google Scholar] [CrossRef] [PubMed]
  72. Lee, J.K.; Kwak, H.J.; Piao, M.S.; Jang, J.W.; Kim, S.H.; Kim, H.S. Quercetin reduces the elevated matrix metalloproteinases-9 level and improves functional outcome after cerebral focal ischemia in rats. Acta Neurochir. 2011, 153, 1321–1329. [Google Scholar] [CrossRef] [PubMed]
  73. Jang, J.W.; Lee, J.K.; Hur, H.; Kim, T.W.; Joo, S.P.; Piao, M.S. Rutin improves functional outcome via reducing the elevated matrix metalloproteinase-9 level in a photothrombotic focal ischemic model of rats. J. Neurol. Sci. 2014, 339, 75–80. [Google Scholar] [CrossRef] [PubMed]
  74. Cho, J.Y.; Kim, I.S.; Jang, Y.H.; Kim, A.R.; Lee, S.R. Protective effect of quercetin, a natural flavonoid against neuronal damage after transient global cerebral ischemia. Neurosci. Lett. 2006, 404, 330–335. [Google Scholar] [CrossRef] [PubMed]
  75. Galho, A.R.; Cordeiro, M.F.; Ribeiro, S.A.; Marques, M.S.; Antunes, M.F.; Luz, D.C.; Hädrich, G.; Muccillo-Baisch, A.L.; Barros, D.M.; Lima, J.V.; et al. Protective role of free and quercetin-loaded nanoemulsion against damage induced by intracerebral haemorrhage in rats. Nanotechnology 2016, 27, 175101. [Google Scholar] [CrossRef]
  76. Zhang, Y.; Yi, B.; Ma, J.; Zhang, L.; Zhang, H.; Yang, Y.; Dai, Y. Quercetin promotes neuronal and behavioral recovery by suppressing inflammatory response and apoptosis in a rat model of intracerebral hemorrhage. Neurochem. Res. 2015, 40, 195–203. [Google Scholar] [CrossRef]
  77. Gül, Ş.; Aydoğmuş, E.; Bahadir, B.; Büyükuysal, M.Ç.; Güven, B. Neuroprotective Effects of Quercetin on Cerebral Vasospasm Following Experimental Subarachnoid Haemorrhage in Rats. Turk. J. Med. Sci. 2020, 50, 1106–1110. [Google Scholar] [CrossRef]
  78. Dong, Y.S.; Wang, J.L.; Feng, D.Y.; Qin, H.Z.; Wen, H.; Yin, Z.M.; Gao, G.D.; Li, C. Protective effect of quercetin against oxidative stress and brain edema in an experimental rat model of subarachnoid hemorrhage. Int. J. Med. Sci. 2014, 11, 282–290. [Google Scholar] [CrossRef] [Green Version]
  79. Semwal, D.K.; Semwal, R.B.; Combrinck, S.; Viljoen, A. Myricetin: A Dietary Molecule with Diverse Biological Activities. Nutrients 2016, 8, 90. [Google Scholar] [CrossRef] [Green Version]
  80. Wu, S.; Yue, Y.; Peng, A.; Zhang, L.; Xiang, J.; Cao, X.; Ding, H.; Yin, S. Myricetin ameliorates brain injury and neurological deficits via Nrf2 activation after experimental stroke in middle-aged rats. Food Funct. 2016, 7, 2624–2634. [Google Scholar] [CrossRef]
  81. Sun, L.; Xu, P.; Fu, T.; Huang, X.; Song, J.; Chen, M.; Tian, X.; Yin, H.; Han, J. Myricetin against ischemic cerebral injury in rat middle cerebral artery occlusion model. Mol. Med. Rep. 2018, 17, 3274–3280. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Zhang, S.; Hu, X.; Guo, S.; Shi, L.; He, Q.; Zhang, P.; Yu, S.; Zhao, R. 2019 Myricetin ameliorated ischemia/reperfusion-induced brain endothelial permeability by improvement of eNOS uncoupling and activation eNOS/NO. J. Pharmacol. Sci. 2019, 140, 62–72. [Google Scholar] [CrossRef] [PubMed]
  83. Ren, J.; Lu, Y.; Qian, Y.; Chen, B.; Wu, T.; Ji, G. Recent progress regarding kaempferol for the treatment of various diseases. Exp. Ther. Med. 2019, 18, 2759–2776. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. López-Sánchez, C.; Martín-Romero, F.J.; Sun, F.; Luis, L.; Samhan-Arias, A.K.; García-Martínez, V.; Gutiérrez-Merino, C. Blood micromolar concentrations of kaempferol afford protection against ischemia/reperfusion-induced damage in rat brain. Brain Res. 2007, 1182, 123–137. [Google Scholar] [CrossRef]
  85. Yu, L.; Chen, C.; Wang, L.F.; Kuang, X.; Liu, K.; Zhang, H.; Du, J.R. Neuroprotective effect of kaempferol glycosides against brain injury and neuroinflammation by inhibiting the activation of NF-κB and STAT3 in transient focal stroke. PLoS ONE 2013, 8, e55839. [Google Scholar] [CrossRef]
  86. Li, W.H.; Cheng, X.; Yang, Y.L.; Liu, M.; Zhang, S.S.; Wang, Y.H.; Du, G.H. Kaempferol attenuates neuroinflammation and blood brain barrier dysfunction to improve neurological deficits in cerebral ischemia/reperfusion rats. Brain Res. 2019, 1722, 146361. [Google Scholar] [CrossRef]
  87. Maher, P. How fisetin reduces the impact of age and disease on CNS function. Front. Biosci. Schol. Ed. 2015, 7, 58–82. [Google Scholar] [CrossRef]
  88. Maher, P.; Salgado, K.F.; Zivin, J.A.; Lapchak, P.A. A novel approach to screening for new neuroprotective compounds for the treatment of stroke. Brain Res. 2007, 1173, 117–125. [Google Scholar] [CrossRef] [Green Version]
  89. Gelderblom, M.; Leypoldt, F.; Lewerenz, J.; Birkenmayer, G.; Orozco, D.; Ludewig, P.; Thundyil, J.; Arumugam, T.V.; Gerloff, C.; Tolosa, E.; et al. The flavonoid fisetin attenuates postischemic immune cell infiltration, activation and infarct size after transient cerebral middle artery occlusion in mice. J. Cereb. Blood Flow Metab. 2012, 32, 835–843. [Google Scholar] [CrossRef]
  90. Campos-Esparza, M.R.; Sánchez-Gómez, M.V.; Matute, C. Molecular mechanisms of neuroprotection by two natural antioxidant polyphenols. Cell Calcium 2009, 45, 358–368. [Google Scholar] [CrossRef]
  91. Chen, Y.; Li, Y.; Xu, H.; Li, G.; Ma, Y.; Pang, Y.J. Morin mitigates oxidative stress, apoptosis and inflammation in cerebral ischemic rats. Afr. J. Tradit. Complement. Altern. Med. 2017, 14, 348–355. [Google Scholar] [CrossRef] [PubMed]
  92. Khamchai, S.; Chumboatong, W.; Hata, J.; Tocharus, C.; Suksamrarn, A.; Tocharus, J. Morin protects the blood-brain barrier integrity against cerebral ischemia reperfusion through anti-inflammatory actions in rats. Sci Rep. 2020, 10, 13379. [Google Scholar] [CrossRef] [PubMed]
  93. Khan, N.; Mukhtar, H. Tea Polyphenols in Promotion of Human Health. Nutrients 2018, 11, 39. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Pervin, M.; Unno, K.; Takagaki, A.; Isemura, M.; Nakamura, Y. Function of Green Tea Catechins in the Brain: Epigallocatechin Gallate and its Metabolites. Int. J. Mol. Sci. 2019, 20, 3630. [Google Scholar] [CrossRef] [Green Version]
  95. Faggi, L.; Porrini, V.; Lanzillotta, A.; Benarese, M.; Mota, M.; Tsoukalas, D.; Parrella, E.; Pizzi, M. A Polyphenol-Enriched Supplement Exerts Potent Epigenetic-Protective Activity in a Cell-Based Model of Brain Ischemia. Nutrients 2019, 11, 345. [Google Scholar] [CrossRef] [Green Version]
  96. Panickar, K.S.; Polansky, M.M.; Anderson, R.A. Green tea polyphenols attenuate glial swelling and mitochondrial dysfunction following oxygen-glucose deprivation in cultures. Nutr. Neurosci. 2009, 12, 105–113. [Google Scholar] [CrossRef]
  97. Gundimeda, U.; McNeill, T.H.; Elhiani, A.A.; Schiffman, J.E.; Hinton, D.R.; Gopalakrishna, R. Green tea polyphenols precondition against cell death induced by oxygen-glucose deprivation via stimulation of laminin receptor, generation of reactive oxygen species, and activation of protein kinase Cε. J. Biol. Chem. 2012, 287, 34694–34708. [Google Scholar] [CrossRef] [Green Version]
  98. Park, J.W.; Jang, Y.H.; Kim, J.M.; Lee, H.; Park, W.K.; Lim, M.B.; Chu, Y.K.; Lo, E.H.; Lee, S.R. Green tea polyphenol—Epigallocatechin gallate reduces neuronal cell damage and up-regulation of MMP-9 activity in hippocampal CA1 and CA2 areas following transient global cerebral ischemia. J. Neurosci. Res. 2009, 87, 567–575. [Google Scholar] [CrossRef]
  99. Lim, S.H.; Kim, H.S.; Kim, Y.K.; Kim, T.M.; Im, S.; Chung, M.E.; Hong, B.Y.; Ko, Y.J.; Kim, H.W.; Lee, J.I. The functional effect of epigallocatechin gallate on ischemic stroke in rats. Acta Neurobiol. Exp. (Wars) 2010, 70, 40–46. [Google Scholar]
  100. Choi, Y.B.; Kim, Y.I.; Lee, K.S.; Kim, B.S.; Kim, D.J. Protective effect of epigallocatechin gallate on brain damage after transient middle cerebral artery occlusion in rats. Brain Res. 2004, 1019, 47–54. [Google Scholar] [CrossRef]
  101. Rahman, R.M.; Nair, S.M.; Helps, S.C.; Shaw, O.M.; Sims, N.R.; Rosengren, R.J.; Appleton, I. (-)-Epigallocatechin gallate as an intervention for the acute treatment of cerebral ischemia. Neurosci. Lett. 2005, 382, 227–230. [Google Scholar] [CrossRef] [PubMed]
  102. Zhang, F.; Li, N.; Jiang, L.; Chen, L.; Huang, M. Neuroprotective Effects of (-)-Epigallocatechin-3-Gallate Against Focal Cerebral Ischemia/Reperfusion Injury in Rats Through Attenuation of Inflammation. Neurochem. Res. 2015, 40, 1691–1698. [Google Scholar] [CrossRef] [PubMed]
  103. Nan, W.; Zhonghang, X.; Keyan, C.; Tongtong, L.; Wanshu, G.; Zhongxin, X. Epigallocatechin-3-Gallate Reduces Neuronal Apoptosis in Rats after Middle Cerebral Artery Occlusion Injury via PI3K/AKT/eNOS Signaling Pathway. Biomed. Res. Int. 2018, 2018, 6473580. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Han, J.; Wang, M.; Jing, X.; Shi, H.; Ren, M.; Lou, H. (-)-Epigallocatechin gallate protects against cerebral ischemia-induced oxidative stress via Nrf2/ARE signaling. Neurochem. Res. 2014, 39, 1292–1299. [Google Scholar] [CrossRef] [PubMed]
  105. Lee, S.; Suh, S.; Kim, S. Protective effects of the green tea polyphenol (-)-epigallocatechin gallate against hippocampal neuronal damage after transient global ischemia in gerbils. Neurosci. Lett. 2000, 287, 191–194. [Google Scholar] [CrossRef]
  106. Lee, H.; Bae, J.H.; Lee, S.R. Protective effect of green tea polyphenol EGCG against neuronal damage and brain edema after unilateral cerebral ischemia in gerbils. J. Neurosci. Res. 2004, 77, 892–900. [Google Scholar] [CrossRef] [PubMed]
  107. You, Y.P. Epigallocatechin Gallate Extends the Therapeutic Window of Recombinant Tissue Plasminogen Activator Treatment in Ischemic Rats. J. Stroke Cerebrovasc. Dis. 2016, 25, 990–997. [Google Scholar] [CrossRef]
  108. Chen, Y.; Huang, L.; Wang, L.; Chen, L.; Ren, W.; Zhou, W. Differential expression of microRNAs contributed to the health efficacy of EGCG in in vitro subarachnoid hemorrhage model. Food Funct. 2017, 8, 4675–4683. [Google Scholar] [CrossRef]
  109. Chen, Y.; Huang, L.; Zhang, H.; Diao, X.; Zhao, S.; Zhou, W. Reduction in Autophagy by (-)-Epigallocatechin-3-Gallate EGCG: A Potential Mechanism of Prevention of Mitochondrial Dysfunction After Subarachnoid Hemorrhage. Mol. Neurobiol. 2017, 54, 392–405. [Google Scholar] [CrossRef]
  110. Chen, Y.; Chen, J.; Sun, X.; Shi, X.; Wang, L.; Huang, L.; Zhou, W. Evaluation of the neuroprotective effect of EGCG: A potential mechanism of mitochondrial dysfunction and mitochondrial dynamics after subarachnoid hemorrhage. Food Funct. 2018, 9, 6349–6359. [Google Scholar] [CrossRef]
  111. Fu, B.; Zeng, Q.; Zhang, Z.; Qian, M.; Chen, J.; Dong, W.; Li, M. Epicatechin Gallate Protects HBMVECs from Ischemia/Reperfusion Injury through Ameliorating Apoptosis and Autophagy and Promoting Neovascularization. Oxid. Med. Cell. Longev. 2019, 2019, 7824684. [Google Scholar] [CrossRef] [PubMed]
  112. Shah, Z.A.; Li, R.C.; Ahmad, A.S.; Kensler, T.W.; Yamamoto, M.; Biswal, S.; Doré, S. The flavanol (-)-epicatechin prevents stroke damage through the Nrf2/HO1 pathway. J. Cereb. Blood Flow Metab. 2010, 30, 1951–1961. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Leonardo, C.C.; Agrawal, M.; Singh, N.; Moore, J.R.; Biswal, S.; Doré, S. Oral administration of the flavanol (-)-epicatechin bolsters endogenous protection against focal ischemia through the Nrf2 cytoprotective pathway. Eur. J. Neurosci. 2013, 38, 3659–3668. [Google Scholar] [CrossRef] [PubMed]
  114. Lan, X.; Han, X.; Li, Q.; Wang, J. (-)-Epicatechin, a Natural Flavonoid Compound, Protects Astrocytes Against Hemoglobin Toxicity via Nrf2 and AP-1 Signaling Pathways. Mol. Neurobiol. 2017, 54, 7898–7907. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Chang, C.F.; Cho, S.; Wang, J. (-)-Epicatechin protects hemorrhagic brain via synergistic Nrf2 pathways. Ann. Clin. Transl. Neurol. 2014, 1, 258–271. [Google Scholar] [CrossRef] [PubMed]
  116. Habtemariam, S. The Nrf2/HO-1 Axis as Targets for Flavanones: Neuroprotection by Pinocembrin, Naringenin, and Eriodictyol. Oxid. Med. Cell. Longev. 2019, 2019, 4724920. [Google Scholar] [CrossRef] [PubMed]
  117. Barreca, D.; Gattuso, G.; Bellocco, E.; Calderaro, A.; Trombetta, D.; Smeriglio, A.; Laganà, G.; Daglia, M.; Meneghini, S.; Nabavi, S.M. Flavanones: Citrus phytochemical with health-promoting properties. Biofactors 2017, 43, 495–506. [Google Scholar] [CrossRef]
  118. Islam, A.; Islam, M.; Rahman, M.K.; Uddin, M.N.; Akanda, M.R. The pharmacological and biological roles of eriodictyol. Arch. Pharm. Res. 2020, 43, 582–592. [Google Scholar] [CrossRef]
  119. Ferreira, E.O.; Fernandes, M.; Lima, N.M.; Neves, K.R.; Carmo, M.R.; Lima, F.A.; Fonteles, A.A.; Menezesa, A.P.; Andrade, G.M. Neuroinflammatory response to experimental stroke is inhibited by eriodictyol. Behav. Brain Res. 2016, 312, 321–332. [Google Scholar] [CrossRef]
  120. Jing, X.; Ren, D.; Wei, X.; Shi, H.; Zhang, X.; Perez, R.G.; Lou, H.; Lou, H. Eriodictyol-7-Oglucoside activates Nrf2 and protects against cerebral ischemic injury. Toxicol. Appl. Pharmacol. 2013, 273, 672–679. [Google Scholar] [CrossRef]
  121. Nouri, Z.; Fakhri, S.; El-Senduny, F.F.; Sanadgol, N.; Abd-ElGhani, G.E.; Farzaei, M.H.; Chen, J.T. On the Neuroprotective Effects of Naringenin: Pharmacological Targets, Signaling Pathways, Molecular Mechanisms, and Clinical Perspective. Biomolecules 2019, 9, 690. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Wang, K.; Chen, Z.; Huang, L.; Meng, B.; Zhou, X.; Wen, X.; Ren, D. Naringenin reduces oxidative stress and improves mitochondrial dysfunction via activation of the Nrf2/ARE signaling pathway in neurons. Int. J. Mol. Med. 2017, 40, 1582–1590. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Wang, K.; Chen, Z.; Huang, J.; Huang, L.; Luo, N.; Liang, X.; Liang, M.; Xie, W. Naringenin prevents ischaemic stroke damage via anti-apoptotic and anti-oxidant effects. Clin. Exp. Pharmacol. Physiol. 2017, 44, 862–871. [Google Scholar] [CrossRef] [PubMed]
  124. Ahmad, A.; Fauzia, E.; Kumar, M.; Mishra, R.K.; Kumar, A.; Khan, M.A.; Raza, S.S.; Khan, R. Gelatin-Coated Polycaprolactone Nanoparticle-Mediated Naringenin Delivery Rescue Human Mesenchymal Stem Cells from Oxygen Glucose Deprivation-Induced Inflammatory Stress. ACS Biomater. Sci. Eng. 2019, 5, 683–695. [Google Scholar] [CrossRef]
  125. Raza, S.S.; Khan, M.M.; Ahmad, A.; Ashafaq, M.; Islam, F.; Wagner, A.P.; Safhi, M.M.; Islam, F. Neuroprotective effect of naringenin is mediated through suppression of NF-κB signaling pathway in experimental stroke. Neuroscience 2013, 230, 157–171. [Google Scholar] [CrossRef]
  126. Bai, X.; Zhang, X.; Chen, L.; Zhang, J.; Zhang, L.; Zhao, X.; Zhao, T.; Zhao, Y. Protective effect of naringenin in experimental ischemic stroke: Down-regulated NOD2. RIP2. NF-κB. MMP-9 and up-regulated claudin-5 expression. Neurochem. Res. 2014, 39, 1405–1415. [Google Scholar] [CrossRef]
  127. Ikemura, M.; Sasaki, Y.; Giddings, J.C.; Yamamoto, J. Preventive effects of hesperidin. glucosyl hesperidin and naringin on hypertension and cerebral thrombosis in stroke-prone spontaneously hypertensive rats. Phytother. Res. 2012, 26, 1272–1277. [Google Scholar] [CrossRef]
  128. Li, C.; Schluesener, H. Health-promoting effects of the citrus flavanone hesperidin. Crit. Rev. Food Sci. Nutr. 2017, 57, 613–631. [Google Scholar] [CrossRef]
  129. Raza, S.S.; Khan, M.M.; Ahmad, A.; Ashafaq, M.; Khuwaja, G.; Tabassum, R.; Javed, H.; Siddiqui, M.S.; Safhi, M.M.; Islam, F. Hesperidin ameliorates functional and histological outcome and reduces neuroinflammation in experimental stroke Brain Res. 2011, 1420, 93–105. 1420.
  130. Praveen Kumar, P.; Sunil Kumar, K.T.; Kavya Nainita, M.; Sai Tarun, A.; Raghu Ramudu, B.G.; Deepika, K.; Pramoda, A.; Yasmeen, C. Cerebroprotective Potential of Hesperidin Nanoparticles Against Bilateral Common Carotid Artery Occlusion Reperfusion Injury in Rats and In silico Approaches. Neurotox. Res. 2020, 37, 264–274. [Google Scholar] [CrossRef]
  131. Aydogmus, E.; Gul, S.; Bahadir, B. Neuroprotective Effects of Hesperidin on Cerebral Vasospasm After Experimental Subarachnoid Hemorrhage in Rats: Biochemical. Pathologic. and Histomorphometric Analysis. World Neurosurg. 2019, 122, e1332–e1337. [Google Scholar] [CrossRef]
  132. Shen, X.; Liu, Y.; Luo, X.; Yang, Z. Advances in Biosynthesis. Pharmacology. and Pharmacokinetics of Pinocembrin, a Promising Natural Small-Molecule Drug. Molecules 2019, 24, 2323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Liu, R.; Gao, M.; Yang, Z.H.; Du, G.H. Pinocembrin protects rat brain against oxidation and apoptosis induced by ischemia-reperfusion both in vivo and in vitro. Brain Res. 2008, 1216, 104–115. [Google Scholar] [CrossRef] [PubMed]
  134. Meng, F.; Liu, R.; Gao, M.; Wang, Y.; Yu, X.; Xuan, Z.; Sun, J.; Yang, F.; Wu, C.; Du, G. Pinocembrin attenuates blood-brain barrier injury induced by global cerebral ischemia-reperfusion in rats. Brain Res. 2011, 1391, 93–101. [Google Scholar] [CrossRef] [PubMed]
  135. Ma, Y.; Li, L.; Kong, L.; Zhu, Z.; Zhang, W.; Song, J.; Chang, J.; Du, G. Pinocembrin Protects Blood-Brain Barrier Function and Expands the Therapeutic Time Window for Tissue-Type Plasminogen Activator Treatment in a Rat Thromboembolic Stroke Model. Biomed. Res. Int. 2018, 2018, 8943210. [Google Scholar] [CrossRef]
  136. Zhao, G.; Zhang, W.; Li, L.; Wu, S.; Du, G. Pinocembrin protects the brain against ischemia-reperfusion injury and reverses the autophagy dysfunction in the penumbra area. Molecules 2014, 19, 15786–15798. [Google Scholar] [CrossRef] [Green Version]
  137. Gao, M.; Zhu, S.; Tan, C.B.; Xu, B.; Zhang, W.C.; Du, G.H. Pinocembrin protects the neurovascular unit by reducing inflammation and extracellular proteolysis in MCAO rats. J. Asian Nat. Prod. Res. 2010, 12, 407–418. [Google Scholar] [CrossRef]
  138. Saad, M.A.; Abdel Salam, R.; Kenawy, S.A.; Attia, A.S. Pinocembrin attenuates hippocampal inflammation, oxidative perturbations and apoptosis in a rat model of global cerebral ischemia reperfusion. Pharmacol. Rep. 2015, 67, 115–122. [Google Scholar] [CrossRef]
  139. Lan, X.; Han, X.; Li, Q.; Li, Q.; Gao, Y.; Cheng, T.; Wan, J.; Zhu, W.; Wang, J. Pinocembrin protects hemorrhagic brain primarily by inhibiting toll-like receptor 4 and reducing M1 phenotype microglia. Brain Behav. Immun. 2017, 61, 326–339. [Google Scholar] [CrossRef] [Green Version]
  140. Jiang, N.; Doseff, A.; Grotewold, E. Flavones: From Biosynthesis to Health Benefits. Plants 2016, 5, 27. [Google Scholar] [CrossRef]
  141. Ashaari, Z.; Hadjzadeh, M.A.; Hassanzadeh, G.; Alizamir, T.; Yousefi, B.; Keshavarzi, Z.; Mokhtari, T. The Flavone Luteolin Improves Central Nervous System Disorders by Different Mechanisms: A Review. J. Mol. Neurosci. 2018, 65, 491–506. [Google Scholar] [CrossRef]
  142. Tian, T.; Zeng, J.; Zhao, G.; Zhao, W.; Gao, S.; Liu, L. Neuroprotective effects of orientin on oxygen-glucose deprivation/reperfusion-induced cell injury in primary culture of rat cortical neurons. Exp. Biol. Med. 2018, 243, 78–86. [Google Scholar] [CrossRef] [PubMed]
  143. Parrella, E.; Porrini, V.; Iorio, R.; Benarese, M.; Lanzillotta, A.; Mota, M.; Fusco, M.; Tonin, P.; Spano, P.; Pizzi, M. PEA and luteolin synergistically reduce mast cell-mediated toxicity and elicit neuroprotection in cell-based models of brain ischemia. Brain Res. 2016, 1648 Pt A, 409–417. [Google Scholar] [CrossRef] [Green Version]
  144. Luo, S.; Li, H.; Mo, Z.; Lei, J.; Zhu, L.; Huang, Y.; Fu, R.; Li, C.; Huang, Y.; Liu, K.; et al. Connectivity map identifies luteolin as a treatment option of ischemic stroke by inhibiting MMP9 and activation of the PI3K/Akt signaling pathway. Exp. Mol. Med. 2019, 51, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Li, Q.; Tian, Z.; Wang, M.; Kou, J.; Wang, C.; Rong, X.; Li, J.; Xie, X.; Pang, X. Luteoloside attenuates neuroinflammation in focal cerebral ischemia in rats via regulation of the PPARγ/Nrf2/NF-κB signaling pathway. Int. Immunopharmacol. 2019, 66, 309–316. [Google Scholar] [CrossRef] [PubMed]
  146. Qiao, H.; Dong, L.; Zhang, X.; Zhu, C.; Wang, L.; Liu, Z.; Chen, L.; Xing, Y.; Wang, C.; Li, Y. Protective effect of luteolin in experimental ischemic stroke: Upregulated SOD1. CAT. Bcl-2 and claudin-5. down-regulated MDA and Bax expression. Neurochem. Res. 2012, 37, 2014–2024. [Google Scholar] [CrossRef] [PubMed]
  147. Qiao, H.; Zhang, X.; Zhu, C.; Dong, L.; Wang, L.; Xing, Y.; Wang, C.; Ji, Y.; Cao, X. Luteolin downregulates TLR4. TLR5. NF-kappaB and p-p38MAPK expression. upregulates the p-ERK expression. and protects rat brains against focal ischemia. Brain Res. 2012, 1448, 71–81. [Google Scholar] [CrossRef] [PubMed]
  148. Tan, X.; Yang, Y.; Xu, J.; Zhang, P.; Deng, R.; Mao, Y.; He, J.; Chen, Y.; Zhang, Y.; Ding, J.; et al. Luteolin Exerts Neuroprotection via Modulation of the p62/Keap1/Nrf2 Pathway in Intracerebral Hemorrhage. Front. Pharmacol. 2020, 10, 1551. [Google Scholar] [CrossRef]
  149. Yang, Y.; Tan, X.; Xu, J.; Wang, T.; Liang, T.; Xu, X.; Ma, C.; Xu, Z.; Wang, W.; Li, H.; et al. Luteolin alleviates neuroinflammation via downregulating the TLR4/TRAF6/NF-κB pathway after intracerebral hemorrhage. Biomed. Pharmacother. 2020, 126, 110044. [Google Scholar] [CrossRef]
  150. Salehi, B.; Venditti, A.; Sharifi-Rad, M.; Kręgiel, D.; Sharifi-Rad, J.; Durazzo, A.; Lucarini, M.; Santini, A.; Souto, E.B.; Novellino, E.; et al. The Therapeutic Potential of Apigenin. Int. J. Mol. Sci. 2019, 20, 1305. [Google Scholar] [CrossRef] [Green Version]
  151. Pang, Q.; Zhao, Y.; Chen, X.; Zhao, K.; Zhai, Q.; Tu, F. Apigenin Protects the Brain against Ischemia/Reperfusion Injury via Caveolin-1/VEGF In Vitro and In Vivo. Oxid. Med. Cell. Longev. 2018, 2018, 7017204. [Google Scholar] [CrossRef]
  152. Tu, F.; Pang, Q.; Huang, T.; Zhao, Y.; Liu, M.; Chen, X. Apigenin Ameliorates Post-Stroke Cognitive Deficits in Rats Through Histone Acetylation-Mediated Neurochemical Alterations. Med. Sci. Monit. 2017, 23, 4004–4013. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Zhang, T.; Su, J.; Guo, B.; Wang, K.; Li, X.; Liang, G. Apigenin protects blood-brain barrier and ameliorates early brain injury by inhibiting TLR4-mediated inflammatory pathway in subarachnoid hemorrhage rats. Int. Immunopharmacol. 2015, 28, 79–87. [Google Scholar] [CrossRef] [PubMed]
  154. Tu, F.; Pang, Q.; Chen, X.; Huang, T.; Liu, M.; Zhai, Q. Angiogenic effects of apigenin on endothelial cells after hypoxia-reoxygenation via the caveolin-1 pathway. Int. J. Mol. Med. 2017, 40, 1639–1648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Huang, H.; Li, L.; Shi, W.; Liu, H.; Yang, J.; Yuan, X.; Wu, L. The Multifunctional Effects of Nobiletin and Its Metabolites In Vivo and In Vitro. Evid. Based Complement. Altern. Med. 2016, 2016, 2918796. [Google Scholar]
  156. Yamamoto, Y.; Shioda, N.; Han, F.; Moriguchi, S.; Nakajima, A.; Yokosuka, A.; Mimaki, Y.; Sashida, Y.; Yamakuni, T.; Ohizumi, Y.; et al. Nobiletin improves brain ischemia-induced learning and memory deficits through stimulation of CaMKII and CREB phosphorylation. Brain Res. 2009, 1295, 218–229. [Google Scholar] [CrossRef]
  157. Yasuda, N.; Ishii, T.; Oyama, D.; Fukuta, T.; Agato, Y.; Sato, A.; Shimizu, K.; Asai, T.; Asakawa, T.; Kan, T.; et al. Neuroprotective effect of nobiletin on cerebral ischemia-reperfusion injury in transient middle cerebral artery-occluded rats. Brain Res. 2014, 1559, 46–54. [Google Scholar] [CrossRef]
  158. Ashrafizadeh, M.; Ahmadi, Z.; Mohammadinejad, R.; Ghasemipour Afshar, E. Tangeretin: A mechanistic review of its pharmacological and therapeutic effects. J. Basic Clin. Physiol. Pharmacol. 2020, 31. [Google Scholar] [CrossRef]
  159. Wu, C.; Zhao, J.; Chen, Y.; Li, T.; Zhu, R.; Zhu, B.; Zhang, Y. Tangeretin protects human brain microvascular endothelial cells against oxygen-glucose deprivation-induced injury. J. Cell. Biochem. 2019, 120, 4883–4891. [Google Scholar] [CrossRef]
  160. Yang, E.J.; Lim, S.; Song, K.S.; Han, H.S.; Lee, J. Identification of active compounds from Aurantii Immatri Pericarpium attenuating brain injury in a rat model of ischemia-reperfusion. Food Chem. 2013, 138, 663–670. [Google Scholar] [CrossRef]
  161. Liang, W.; Huang, X.; Chen, W. Effects of Baicalin and Baicalein on Cerebral Ischemia: A Review. Aging Dis. 2017, 8, 850–867. [Google Scholar] [CrossRef] [Green Version]
  162. Sowndhararajan, K.; Deepa, P.; Kim, M.; Park, S.J.; Kim, S. Neuroprotective and Cognitive Enhancement Potentials of Baicalin: A Review. Brain Sci. 2018, 8, 104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Xu, M.; Chen, X.; Gu, Y.; Peng, T.; Yang, D.; Chang, R.C.; So, K.F.; Liu, K.; Shen, J. Baicalin can scavenge peroxynitrite and ameliorate endogenous peroxynitrite-mediated neurotoxicity in cerebral ischemia-reperfusion injury. J. Ethnopharmacol. 2013, 150, 116–124. [Google Scholar] [CrossRef] [PubMed]
  164. Xue, X.; Qu, X.J.; Yang, Y.; Sheng, X.H.; Cheng, F.; Jiang, E.N.; Wang, J.H.; Bu, W.; Liu, Z.P. Baicalin attenuates focal cerebral ischemic reperfusion injury through inhibition of nuclear factor κB p65 activation. Biochem. Biophys. Res. Commun. 2010, 403, 398–404. [Google Scholar] [CrossRef] [PubMed]
  165. Cao, Y.; Mao, X.; Sun, C.; Zheng, P.; Gao, J.; Wang, X.; Min, D.; Sun, H.; Xie, N.; Cai, J. Baicalin attenuates global cerebral ischemia/reperfusion injury in gerbils via anti-oxidative and anti-apoptotic pathways. Brain Res. Bull. 2011, 85, 396–402. [Google Scholar] [CrossRef] [PubMed]
  166. Semwal, R.B.; Semwal, D.; Combrinck, S.; Trill, J.; Gibbons, S.; Viljoenad, A. Acacetin—A simple flavone exhibiting diverse pharmacological activities. Phytochem. Lett. 2019, 32, 56–65. [Google Scholar] [CrossRef]
  167. Bu, J.; Shi, S.; Wang, H.Q.; Niu, X.S.; Zhao, Z.F.; Wu, W.D.; Zhang, X.L.; Ma, Z.; Zhang, Y.J.; Zhang, H.; et al. Acacetin protects against cerebral ischemia-reperfusion injury via the NLRP3 signaling pathway. Neural Regen. Res. 2019, 14, 605–612. [Google Scholar]
  168. Pabich, M.; Materska, M. Biological Effect of Soy Isoflavones in the Prevention of Civilization Diseases. Nutrients 2019, 11, 1660. [Google Scholar] [CrossRef] [Green Version]
  169. Křížová, L.; Dadáková, K.; Kašparovská, J.; Kašparovský, T. Isoflavones. Molecules 2019, 24, 1076. [Google Scholar] [CrossRef] [Green Version]
  170. Schreihofer, D.A.; Oppong-Gyebi, A. Genistein: Mechanisms of action for a pleiotropic neuroprotective agent in stroke. Nutr. Neurosci. 2019, 22, 375–391. [Google Scholar] [CrossRef]
  171. Nabavi, S.F.; Daglia, M.; Tundis, R.; Loizzo, M.R.; Sobarzo-Sanchez, E.; Orhan, I.E.; Nabavi, S.M. Genistein: A Boon for Mitigating Ischemic Stroke. Curr. Top. Med. Chem. 2015, 15, 1714–1721. [Google Scholar] [CrossRef]
  172. Ma, X.L.; Zhang, F.; Wang, Y.X.; He, C.C.; Tian, K.; Wang, H.G.; An, D.; Heng, B.; Liu, Y.Q. Genistein inhibition of OGD-induced brain neuron death correlates with its modulation of apoptosis, voltage-gated potassium and sodium currents and glutamate signal pathway. Chem. Biol. Interact. 2016, 254, 73–82. [Google Scholar] [CrossRef] [PubMed]
  173. Wang, Y.X.; Tian, K.; He, C.C.; Ma, X.L.; Zhang, F.; Wang, H.G.; An, D.; Heng, B.; Jiang, Y.G.; Liu, Y.Q. Genistein inhibits hypoxia, ischemic-induced death, and apoptosis in PC12 cells. Environ. Toxicol. Pharmacol. 2017, 50, 227–233. [Google Scholar] [CrossRef] [PubMed]
  174. Qian, Y.; Guan, T.; Huang, M.; Cao, L.; Li, Y.; Cheng, H.; Jin, H.; Yu, D. Neuroprotection by the soy isoflavone, genistein, via inhibition of mitochondria-dependent apoptosis pathways and reactive oxygen induced-NF-κB activation in a cerebral ischemia mouse model. Neurochem. Int. 2012, 60, 759–767. [Google Scholar] [CrossRef] [PubMed]
  175. Wang, S.; Wei, H.; Cai, M.; Lu, Y.; Hou, W.; Yang, Q.; Dong, H.; Xiong, L. Genistein attenuates brain damage induced by transient cerebral ischemia through up-regulation of ERK activity in ovariectomized mice. Int. J. Biol. Sci. 2014, 10, 457–465. [Google Scholar] [CrossRef] [Green Version]
  176. Ma, Y.; Sullivan, J.; Schreihofer, D.A. Dietary genistein and equol 4’, 7 isoflavandiol reduce oxidative stress and protect rats against focal cerebral ischemia. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2010, 299, R871–R877. [Google Scholar] [CrossRef]
  177. Prongay, K.D.; Lewis, A.; Hurn, P.D.; Murphy, S.J. Dietary soy may not confound acute experimental stroke infarct volume outcomes in ovariectomized female rats. Lab. Anim. 2010, 44, 238–246. [Google Scholar] [CrossRef] [Green Version]
  178. Cortina, B.; Torregrosa, G.; Castelló-Ruiz, M.; Burguete, M.C.; Moscardó, A.; Latorre, A.; Salom, J.B.; Vallés, J.; Santos, M.T.; Alborch, E. Improvement of the circulatory function partially accounts for the neuroprotective action of the phytoestrogen genistein in experimental ischemic stroke. Eur. J. Pharmacol. 2013, 708, 88–94. [Google Scholar] [CrossRef]
  179. Li, L.; Xue, J.; Liu, R.; Li, X.; Lai, L.; Xie, J.; Huang, Z.; Huang, C. Neuroprotective effects of genistein-3’-sodium sulfonate on focal cerebral ischemia in rats. Neurosci. Lett. 2017, 646, 43–48. [Google Scholar] [CrossRef]
  180. Miao, Z.Y.; Xia, X.; Che, L.; Song, Y.T. Genistein attenuates brain damage induced by transient cerebral ischemia through up-regulation of Nrf2 expression in ovariectomized rats. Neurol. Res. 2018, 40, 689–695. [Google Scholar] [CrossRef]
  181. Aras, A.B.; Guven, M.; Akman, T.; Alacam, H.; Kalkan, Y.; Silan, C.; Cosar, M. Genistein exerts neuroprotective effect on focal cerebral ischemia injury in rats. Inflammation 2015, 38, 1311–1321. [Google Scholar] [CrossRef]
  182. Trieu, V.N.; Uckun, F. Genistein is neuroprotective in murine models of familial amyotrophic lateral sclerosis and stroke. Biochem. Biophys. Res. Commun. 1999, 258, 685–688. [Google Scholar] [CrossRef] [PubMed]
  183. Rajput, M.S.; Sarkar, P.; Nirmal, N.P. Inhibition of DPP-4 Activity and Neuronal Atrophy with Genistein Attenuates Neurological Deficits Induced by Transient Global Cerebral Ischemia and Reperfusion in Streptozotocin-Induced Diabetic Mice. Inflammation 2017, 40, 623–635. [Google Scholar] [CrossRef] [PubMed]
  184. Shi, R.; Wang, S.; Qi, X.; Chen, S.; Chen, P.; Zhang, Q. Lose dose genistein inhibits glucocorticoid receptor and ischemic brain injury in female rats. Neurochem. Int. 2014, 65, 14–22. [Google Scholar] [CrossRef] [PubMed]
  185. Wang, R.; Tu, J.; Zhang, Q.; Zhang, X.; Zhu, Y.; Ma, W.; Cheng, C.; Brann, D.W.; Yang, F. Genistein attenuates ischemic oxidative damage and behavioral deficits via eNOS/Nrf2/HO-1 signaling. Hippocampus 2013, 23, 634–647. [Google Scholar] [CrossRef]
  186. Liang, H.W.; Qiu, S.F.; Shen, J.; Sun, L.N.; Wang, J.Y.; Bruce, I.C.; Xia, Q. Genistein attenuates oxidative stress and neuronal damage following transient global cerebral ischemia in rat hippocampus. Neurosci. Lett. 2008, 438, 116–120. [Google Scholar] [CrossRef]
  187. Donzelli, A.; Braida, D.; Finardi, A.; Capurro, V.; Valsecchi, A.E.; Colleoni, M.; Sala, M. Neuroprotective effects of genistein in Mongolian gerbils: Estrogen receptor-β involvement. J. Pharmacol. Sci. 2010, 114, 158–167. [Google Scholar] [CrossRef] [Green Version]
  188. Shin, H.K.; Lee, J.H.; Kim, K.Y.; Kim, C.D.; Lee, W.S.; Rhim, B.Y.; Hong, K.W. Impairment of autoregulatory vasodilation by NADPH oxidase-dependent superoxide generation during acute stage of subarachnoid hemorrhage in rat pial artery. J. Cereb. Blood Flow Metab. 2002, 22, 869–877. [Google Scholar] [CrossRef] [Green Version]
  189. Fujikawa, H.; Tani, E.; Yamaura, I.; Ozaki, I.; Miyaji, K.; Sato, M.; Takahashi, K.; Imajoh-Ohmi, S. Activation of protein kinases in canine basilar artery in vasospasm. J. Cereb. Blood Flow Metab. 1999, 19, 44–52. [Google Scholar] [CrossRef]
  190. Stout, J.M.; Knapp, A.; Banz, W.J.; Wallace, D.G.; Cheatwood, J.L. Subcutaneous daidzein administration enhances recovery of skilled ladder rung walking performance following stroke in rats. Behav. Brain Res. 2013, 256, 428–431. [Google Scholar] [CrossRef]
  191. Kim, E.; Woo, M.S.; Qin, L.; Ma, T.; Beltran, C.D.; Bao, Y.; Bailey, J.A.; Corbett, D.; Ratan, R.R.; Lahiri, D.K.; et al. Daidzein Augments Cholesterol Homeostasis via ApoE to Promote Functional Recovery in Chronic Stroke. J. Neurosci. 2015, 35, 15113–15126. [Google Scholar] [CrossRef] [Green Version]
  192. Aras, A.B.; Guven, M.; Akman, T.; Ozkan, A.; Sen, H.M.; Duz, U.; Kalkan, Y.; Coskun Silan, C.; Cosar, M. Neuroprotective effects of daidzein on focal cerebral ischemia injury in rats. Neural Regen. Res. 2015, 10, 146–152. [Google Scholar] [CrossRef] [PubMed]
  193. Wei, S.Y.; Chen, Y.; Xu, X.Y. Progress on the pharmacological research of puerarin: A review. Chin. J. Nat. Med. 2014, 12, 407–414. [Google Scholar] [CrossRef]
  194. Wang, B.; Ma, W.; Yang, H. Puerarin attenuates hypoxia-resulted damages in neural stem cells by up-regulating microRNA-214. Artif. Cells Nanomed. Biotechnol. 2019, 47, 2746–2753. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Wei, S.; Tong, J.; Xue, Q.; Liu, Y.; Xu, X. Effect of puerarin on transcriptome of astrocyte during oxygen-glucose deprivation/reoxygenation injury. Mol. Cell Biochem. 2017, 425, 113–123. [Google Scholar] [CrossRef]
  196. Wang, J.F.; Mei, Z.G.; Fu, Y.; Yang, S.B.; Zhang, S.Z.; Huang, W.F.; Xiong, L.; Zhou, H.J.; Tao, W.; Feng, Z.T. Puerarin protects rat brain against ischemia/reperfusion injury by suppressing autophagy via the AMPK-mTOR-ULK1 signaling pathway. Neural Regen. Res. 2018, 13, 989–998. [Google Scholar]
  197. Liu, X.; Mei, Z.; Qian, J.; Zeng, Y.; Wang, M. Puerarin partly counteracts the inflammatory response after cerebral ischemia/reperfusion via activating the cholinergic anti-inflammatory pathway. Neural Regen. Res. 2013, 8, 3203–3215. [Google Scholar]
  198. Tao, J.; Cui, Y.; Duan, Y.; Zhang, N.; Wang, C.; Zhang, F. Puerarin attenuates locomotor and cognitive deficits as well as hippocampal neuronal injury through the PI3K/Akt1/GSK-3β signaling pathway in an in vivo model of cerebral ischemia. Oncotarget 2017, 8, 106283–106295. [Google Scholar] [CrossRef] [Green Version]
  199. Wu, M.; Liang, S.; Ma, L.; Han, Y.; Zhang, X.; Xu, C. Effects of delayed puerarin treatment in long-term neurological outcomes of focal ischemic stroke in rats. Indian J. Pharmacol. 2014, 46, 157–160. [Google Scholar] [CrossRef]
  200. Wu, X.D.; Wang, C.; Zhang, Z.Y.; Fu, Y.; Liu, F.Y.; Liu, X.H. Puerarin attenuates cerebral damage by improving cerebral microcirculation in spontaneously hypertensive rats. Evid. Based Complement. Altern. Med. 2014, 2014, 408501. [Google Scholar]
  201. Zhang, Y.; Yang, X.; Ge, X.; Zhang, F. Puerarin attenuates neurological deficits via Bcl-2/Bax/cleaved caspase-3 and Sirt3/SOD2 apoptotic pathways in subarachnoid hemorrhage mice. Biomed. Pharmacother. 2019, 109, 726–733. [Google Scholar] [CrossRef]
  202. Sarfraz, A.; Javeed, M.; Shah, M.A.; Hussain, G.; Shafiq, N.; Sarfraz, I.; Riaz, A.; Sadiqa, A.; Zara, R.; Zafar, S.; et al. Biochanin A: A novel bioactive multifunctional compound from nature. Sci. Total Environ. 2020, 722, 137907. [Google Scholar] [CrossRef] [PubMed]
  203. Khanna, S.; Stewart, R.; Gnyawali, S.; Harris, H.; Balch, M.; Spieldenner, J.; Sen, C.K.; Rink, C. Phytoestrogen isoflavone intervention to engage the neuroprotective effect of glutamate oxaloacetate transaminase against stroke. FASEB J. 2017, 31, 4533–4544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Guo, M.; Lu, H.; Qin, J.; Qu, S.; Wang, W.; Guo, Y.; Liao, W.; Song, M.; Chen, J.; Wang, Y. Biochanin A Provides Neuroprotection Against Cerebral Ischemia/Reperfusion Injury by Nrf2-Mediated Inhibition of Oxidative Stress and Inflammation Signaling Pathway in Rats. Med. Sci. Monit. 2019, 25, 8975–8983. [Google Scholar] [CrossRef] [PubMed]
  205. Wu, L.Y.; Ye, Z.N.; Zhuang, Z.; Gao, Y.; Tang, C.; Zhou, C.H.; Wang, C.X.; Zhang, X.S.; Xie, G.B.; Liu, J.P.; et al. Biochanin A Reduces Inflammatory Injury and Neuronal Apoptosis following Subarachnoid Hemorrhage via Suppression of the TLRs/TIRAP/MyD88/NF-κB Pathway. Behav. Neurol. 2018, 2018, 1960106. [Google Scholar] [CrossRef] [Green Version]
  206. Manolescu, B.N.; Oprea, E.; Mititelu, M.; Ruta, L.L.; Farcasanu, I.C. Dietary Anthocyanins and Stroke: A Review of Pharmacokinetic and Pharmacodynamic Studies. Nutrients 2019, 11, 1479. [Google Scholar] [CrossRef] [Green Version]
  207. Sun, C.; Huang, H.; Xu, C.; Li, X.; Chen, K. Biological activities of extracts from Chinese bayberry Myrica rubra Sieb. et Zucc.: A review. Plant. Foods Hum. Nutr. 2013, 68, 97–106. [Google Scholar] [CrossRef]
  208. Cui, H.X.; Chen, J.; Li, J.W.; Cheng, F.R.; Yuan, K. Protection of Anthocyanin from Myrica rubra against Cerebral Ischemia-Reperfusion Injury via Modulation of the TLR4/NF-κB and NLRP3 Pathways. Molecules 2018, 23, 1788. [Google Scholar] [CrossRef] [Green Version]
  209. Kang, T.H.; Hur, J.; Kim, H.B.; Ryu, J.H.; Kim, S.Y. Neuroprotective effects of the cyanidin-3-O-beta-d-glucopyranoside isolated from mulberry fruit against cerebral ischemia. Neurosci. Lett. 2006, 391, 122–126. [Google Scholar] [CrossRef]
  210. Di Giacomo, C.; Acquaviva, R.; Santangelo, R.; Sorrenti, V.; Vanella, L.; Li Volti, G.; D’Orazio, N.; Vanella, A.; Galvano, F. Effect of Treatment with Cyanidin-3-O-β-D-Glucoside on Rat Ischemic/Reperfusion Brain Damage. Evid. Based Complement. Altern. Med. 2012, 2012, 285750. [Google Scholar] [CrossRef] [Green Version]
  211. Min, J.; Yu, S.W.; Baek, S.H.; Nair, K.M.; Bae, O.N.; Bhatt, A.; Kassab, M.; Nair, M.G.; Majid, A. Neuroprotective effect of cyanidin-3-O-glucoside anthocyanin in mice with focal cerebral ischemia. Neurosci. Lett. 2011, 500, 157–161. [Google Scholar] [CrossRef]
  212. Mastantuono, T.; Di Maro, M.; Chiurazzi, M.; Battiloro, L.; Muscariello, E.; Nasti, G.; Starita, N.; Colantuoni, A.; Lapi, D. Rat Pial Microvascular Changes During Cerebral Blood Flow Decrease and Recovery: Effects of Cyanidin Administration. Front. Physiol. 2018, 9, 540. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Stompor, M.; Broda, D.; Bajek-Bil, A. Dihydrochalcones: Methods of Acquisition and Pharmacological Properties-A First Systematic Review. Molecules 2019, 24, 4468. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Liu, Y.; Zhang, L.; Liang, J. Activation of the Nrf2 defense pathway contributes to neuroprotective effects of phloretin on oxidative stress injury after cerebral ischemia/reperfusion in rats. J. Neurol. Sci. 2015, 351, 88–92. [Google Scholar] [CrossRef] [PubMed]
  215. Rauf, A.; Imran, M.; Abu-Izneid, T.; Patel, S.; Pan, X.; Naz, S.; Silva, A.S.; Saeed, F.; Suleria, H.A. Proanthocyanidins: A comprehensive review. Biomed. Pharmacother. 2019, 116, 108999. [Google Scholar] [CrossRef] [PubMed]
  216. Tu, X.; Wang, M.; Liu, Y.; Zhao, W.; Ren, X.; Li, Y.; Liu, H.; Gu, Z.; Jia, H.; Liu, J.; et al. Pretreatment of Grape Seed Proanthocyanidin Extract Exerts Neuroprotective Effect in Murine Model of Neonatal Hypoxic-ischemic Brain Injury by Its Antiapoptotic Property. Cell. Mol. Neurobiol. 2019, 39, 953–961. [Google Scholar] [CrossRef] [PubMed]
  217. Wu, S.; Yue, Y.; Li, J.; Li, Z.; Li, X.; Niu, Y.; Xiang, J.; Ding, H. Procyanidin B2 attenuates neurological deficits and blood-brain barrier disruption in a rat model of cerebral ischemia. Mol. Nutr. Food Res. 2015, 59, 1930–1941. [Google Scholar] [CrossRef] [PubMed]
  218. Kong, X.; Guan, J.; Gong, S.; Wang, R. Neuroprotective Effects of Grape Seed Procyanidin Extract on Ischemia-Reperfusion Brain Injury. Chin. Med. Sci. J. 2017, 32, 92–99. [Google Scholar]
  219. Matsuda, N.; Ohkuma, H.; Naraoka, M.; Munakata, A.; Shimamura, N.; Asano, K. Role of oxidized LDL and lectin-like oxidized LDL receptor-1 in cerebral vasospasm after subarachnoid hemorrhage. J. Neurosurg. 2014, 121, 621–630. [Google Scholar] [CrossRef] [Green Version]
  220. Sova, M.; Saso, L. Natural Sources, Pharmacokinetics, Biological Activities and Health Benefits of Hydroxycinnamic Acids and Their Metabolites. Nutrients 2020, 12, 2190. [Google Scholar] [CrossRef]
  221. Tolba, M.F.; Omar, H.A.; Azab, S.S.; Khalifa, A.E.; Abdel-Naim, A.B.; Abdel-Rahman, S.Z. Caffeic Acid Phenethyl Ester: A Review of Its Antioxidant Activity, Protective Effects against Ischemia-reperfusion Injury and Drug Adverse Reactions. Crit. Rev. Food Sci. Nutr. 2016, 56, 2183–2190. [Google Scholar] [CrossRef]
  222. Tolba, M.F.; Azab, S.S.; Khalifa, A.E.; Abdel-Rahman, S.Z.; Abdel-Naim, A.B. Caffeic acid phenethyl ester, a promising component of propolis with a plethora of biological activities: A review on its anti-inflammatory, neuroprotective, hepatoprotective, and cardioprotective effects. IUBMB Life 2013, 65, 699–709. [Google Scholar] [CrossRef] [PubMed]
  223. Wei, X.; Zhao, L.; Ma, Z.; Holtzman, D.M.; Yan, C.; Dodel, R.C.; Hampel, H.; Oertel, W.; Farlow, M.R.; Du, Y. Caffeic acid phenethyl ester prevents neonatal hypoxic-ischaemic brain injury. Brain 2004, 127, 2629–2635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Zhou, Y.; Fang, S.H.; Ye, Y.L.; Chu, L.S.; Zhang, W.P.; Wang, M.L.; Wei, E.Q. Caffeic acid ameliorates early and delayed brain injuries after focal cerebral ischemia in rats. Acta Pharmacol. Sin. 2006, 27, 1103–1110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Hwang, S.A.; Kim, C.D.; Lee, W.S. Caffeic acid phenethyl ester protects against photothrombotic cortical ischemic injury in mice. Korean J. Physiol. Pharmacol. 2018, 22, 101–110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Altuğ, M.E.; Serarslan, Y.; Bal, R.; Kontaş, T.; Ekici, F.; Melek, I.M.; Aslan, H.; Duman, T. Caffeic acid phenethyl ester protects rabbit brains against permanent focal ischemia by antioxidant action: A biochemical and planimetric study. Brain Res. 2008, 1201, 135–142. [Google Scholar] [CrossRef]
  227. Tsai, S.K.; Lin, M.J.; Liao, P.H.; Yang, C.Y.; Lin, S.M.; Liu, S.M.; Lin, R.H.; Chih, C.L.; Huang, S.S. Caffeic acid phenethyl ester ameliorates cerebral infarction in rats subjected to focal cerebral ischemia. Life Sci. 2006, 78, 2758–2762. [Google Scholar] [CrossRef]
  228. Liang, G.; Shi, B.; Luo, W.; Yang, J. The protective effect of caffeic acid on global cerebral ischemia-reperfusion injury in rats. Behav. Brain Funct. 2015, 11, 18. [Google Scholar] [CrossRef] [Green Version]
  229. Kim, J.K.; Park, S.U. A recent overview on the biological and pharmacological activities of ferulic acid. EXCLI J. 2019, 18, 132–138. [Google Scholar]
  230. Hassanzadeh, P.; Arbabi, E.; Atyabi, F.; Dinarvand, R. Ferulic acid-loaded nanostructured lipid carriers: A promising nanoformulation against the ischemic neural injuries. Life Sci. 2018, 193, 64–76. [Google Scholar] [CrossRef]
  231. Ren, Z.; Zhang, R.; Li, Y.; Li, Y.; Yang, Z.; Yang, H. Ferulic acid exerts neuroprotective effects against cerebral ischemia/reperfusion-induced injury via antioxidant and anti-apoptotic mechanisms in vitro and in vivo. Int. J. Mol. Med. 2017, 40, 1444–1456. [Google Scholar] [CrossRef] [Green Version]
  232. Zhang, L.; Wang, H.; Wang, T.; Jiang, N.; Yu, P.; Chong, Y.; Fu, F. Ferulic acid ameliorates nerve injury induced by cerebral ischemia in rats. Exp. Ther. Med. 2015, 9, 972–976. [Google Scholar] [CrossRef] [PubMed]
  233. Cheng, C.Y.; Ho, T.Y.; Lee, E.J.; Su, S.Y.; Tang, N.Y.; Hsieh, C.L. Ferulic acid reduces cerebral infarct through its antioxidative and anti-inflammatory effects following transient focal cerebral ischemia in rats. Am. J. Chin. Med. 2008, 36, 1105–1119. [Google Scholar] [CrossRef] [PubMed]
  234. Cheng, C.Y.; Su, S.Y.; Tang, N.Y.; Ho, T.Y.; Chiang, S.Y.; Hsieh, C.L. Ferulic acid provides neuroprotection against oxidative stress-related apoptosis after cerebral ischemia/reperfusion injury by inhibiting ICAM-1 mRNA expression in rats. Brain Res. 2008, 1209, 136–150. [Google Scholar] [CrossRef] [PubMed]
  235. Chen, C. Sinapic Acid and Its Derivatives as Medicine in Oxidative Stress-Induced Diseases and Aging. Oxid. Med. Cell. Longev. 2016, 2016, 3571614. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Kim, Y.O.; Lee, S.W.; Oh, M.S.; Lee, H.J. Effects of sinapic Acid of 4 vessel occlusion model-induced ischemia and cognitive impairments in the rat. Clin. Psychopharmacol. Neurosci. 2011, 9, 86–90. [Google Scholar] [CrossRef] [Green Version]
  237. Pei, K.; Ou, J.; Huang, J.; Ou, S. p-Coumaric acid and its conjugates: Dietary sources, pharmacokinetic properties and biological activities. J. Sci. Food Agric. 2016, 96, 2952–2962. [Google Scholar] [CrossRef]
  238. Guven, M.; Aras, A.B.; Akman, T.; Sen, H.M.; Ozkan, A.; Salis, O.; Sehitoglu, I.; Kalkan, Y.; Silan, C.; Deniz, M.; et al. Neuroprotective effect of p-coumaric acid in rat model of embolic cerebral ischemia. Iran. J. Basic Med. Sci. 2015, 18, 356–363. [Google Scholar]
  239. Sakamula, R.; Thong-Asa, W. Neuroprotective effect of p-coumaric acid in mice with cerebral ischemia reperfusion injuries. Metab. Brain Dis. 2018, 33, 765–773. [Google Scholar] [CrossRef]
  240. Lee, K.; Lee, J.S.; Jang, H.J.; Kim, S.M.; Chang, M.S.; Park, S.H.; Kim, K.S.; Bae, J.; Park, J.W.; Lee, B.; et al. Chlorogenic acid ameliorates brain damage and edema by inhibiting matrix metalloproteinase-2 and 9 in a rat model of focal cerebral ischemia. Eur. J. Pharmacol. 2012, 689, 89–95. [Google Scholar] [CrossRef]
  241. Lee, K.; Lee, B.J.; Bu, Y. Protective Effects of Dihydrocaffeic Acid, a Coffee Component Metabolite, on a Focal Cerebral Ischemia Rat Model. Molecules 2015, 20, 11930–11940. [Google Scholar] [CrossRef] [Green Version]
  242. Miao, M.; Cao, L.; Li, R.; Fang, X.; Miao, Y. Protective effect of chlorogenic acid on the focal cerebral ischemia reperfusion rat models. Saudi Pharm. J. 2017, 25, 556–563. [Google Scholar] [CrossRef] [PubMed]
  243. Liu, D.; Wang, H.; Zhang, Y.; Zhang, Z. Protective Effects of Chlorogenic Acid on Cerebral Ischemia/Reperfusion Injury Rats by Regulating Oxidative Stress-Related Nrf2 Pathway. Drug Des. Dev. Ther. 2020, 14, 51–60. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Lapchak, P. The phenylpropanoid micronutrient chlorogenic acid improves clinical rating scores in rabbits following multiple infarct ischemic strokes: Synergism with tissue plasminogen activator. Exp. Neurol. 2007, 205, 407–413. [Google Scholar] [CrossRef] [PubMed]
  245. Petersen, M.; Simmonds, M.S. Rosmarinic acid. Phytochemistry 2003, 62, 121–125. [Google Scholar] [CrossRef]
  246. Wu, L.; Wang, H.M.; Li, J.L.; Feng, H.X.; Zhao, W.M.; Zhang, H.Y. Dual anti-ischemic effects of rosmarinic acid n-butyl ester via alleviation of DAPK-p53-mediated neuronal damage and microglial inflammation. Acta Pharmacol. Sin. 2017, 38, 459–468. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  247. Luan, H.; Kan, Z.; Xu, Y.; Lv, C.; Jiang, W. Rosmarinic acid protects against experimental diabetes with cerebral ischemia: Relation to inflammation response. J. Neuroinflamm. 2013, 10, 28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  248. Cui, H.Y.; Zhang, X.J.; Yang, Y.; Zhang, C.; Zhu, C.H.; Miao, J.Y.; Chen, R. Rosmarinic acid elicits neuroprotection in ischemic stroke via Nrf2 and heme oxygenase 1 signaling. Neural Regen. Res. 2018, 13, 2119–2128. [Google Scholar]
  249. Fonteles, A.A.; de Souza, C.M.; de Sousa Neves, J.C.; Menezes, A.P.; Santos do Carmo, M.R.; Fernandes, F.D.; de Araújo, P.R.; de Andrade, G.M. Rosmarinic acid prevents against memory deficits in ischemic mice. Behav. Brain Res. 2016, 297, 91–103. [Google Scholar] [CrossRef]
  250. Shabani, S.; Rabiei, Z.; Amini-Khoei, H. Exploring the multifaceted neuroprotective actions of gallic acid: A review. Int. J. Food Prop. 2020, 23, 736–752. [Google Scholar] [CrossRef]
  251. Zhao, Y.; Li, D.; Zhu, Z.; Sun, Y. Improved Neuroprotective Effects of Gallic Acid-Loaded Chitosan Nanoparticles Against Ischemic Stroke. Rejuvenation Res. 2019, 23, 284–292. [Google Scholar] [CrossRef]
  252. Sun, J.; Li, Y.Z.; Ding, Y.H.; Wang, J.; Geng, J.; Yang, H.; Ren, J.; Tang, J.Y.; Gao, J. Neuroprotective effects of gallic acid against hypoxia/reoxygenation-induced mitochondrial dysfunctions in vitro and cerebral ischemia/reperfusion injury in vivo. Brain Res. 2014, 1589, 126–139. [Google Scholar] [CrossRef] [PubMed]
  253. Mirshekari Jahangiri, H.; Sarkaki, A.; Farbood, Y.; Dianat, M.; Goudarzi, G. Gallic acid affects blood-brain barrier permeability, behaviors, hippocampus local EEG, and brain oxidative stress in ischemic rats exposed to dusty particulate matter. Environ. Sci. Pollut. Res. Int. 2020, 27, 5281–5292. [Google Scholar] [CrossRef] [PubMed]
  254. Nabavi, S.F.; Habtemariam, S.; Di Lorenzo, A.; Sureda, A.; Khanjani, S.; Nabavi, S.M.; Daglia, M. Post-Stroke Depression Modulation and in Vivo Antioxidant Activity of Gallic Acid and Its Synthetic Derivatives in a Murine Model System. Nutrients 2016, 8, 248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Reinisalo, M.; Kårlund, A.; Koskela, A.; Kaarniranta, K.; Karjalainen, R.O. Polyphenol Stilbenes: Molecular Mechanisms of Defence against Oxidative Stress and Aging-Related Diseases. Oxid. Med. Cell. Longev. 2015, 2015, 340520. [Google Scholar] [CrossRef] [Green Version]
  256. Rauf, A.; Imran, M.; Suleria, H.A.R.; Ahmad, B.; Peters, D.G.; Mubarak, M.S. A comprehensive review of the health perspectives of resveratrol. Food Funct. 2017, 8, 4284–4305. [Google Scholar] [CrossRef]
  257. Ramírez-Garza, S.L.; Laveriano-Santos, E.P.; Marhuenda-Muñoz, M.; Storniolo, C.E.; Tresserra-Rimbau, A.; Vallverdú-Queralt, A.; Lamuela-Raventós, R.M. Health Effects of Resveratrol: Results from Human Intervention Trials. Nutrients 2018, 10, 1892. [Google Scholar] [CrossRef] [Green Version]
  258. Lopez, M.S.; Dempsey, R.J.; Vemuganti, R. Resveratrol neuroprotection in stroke and traumatic CNS injury. Neurochem. Int. 2015, 89, 75–82. [Google Scholar] [CrossRef] [Green Version]
  259. Singh, N.; Agrawal, M.; Doré, S. Neuroprotective properties and mechanisms of resveratrol in in vitro and in vivo experimental cerebral stroke models. ACS Chem. Neurosci. 2013, 4, 1151–1162. [Google Scholar] [CrossRef] [Green Version]
  260. Pan, S.; Li, S.; Hu, Y.; Zhang, H.; Liu, Y.; Jiang, H.; Fang, M.; Li, Z.; Xu, K.; Zhang, H.; et al. Resveratrol post-treatment protects against neonatal brain injury after hypoxia-ischemia. Oncotarget 2016, 7, 79247–79261. [Google Scholar] [CrossRef] [Green Version]
  261. West, T.; Atzeva, M.; Holtzman, D.M. Pomegranate polyphenols and resveratrol protect the neonatal brain against hypoxic-ischemic injury. Dev. Neurosci. 2007, 29, 363–372. [Google Scholar] [CrossRef] [Green Version]
  262. Arteaga, O.; Revuelta, M.; Urigüen, L.; Álvarez, A.; Montalvo, H.; Hilario, E. Pretreatment with Resveratrol Prevents Neuronal Injury and Cognitive Deficits Induced by Perinatal Hypoxia-Ischemia in Rats. PLoS ONE 2015, 10, e0142424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  263. Agrawal, M.; Kumar, V.; Kashyap, M.P.; Khanna, V.K.; Randhawa, G.S.; Pant, A.B. Ischemic insult induced apoptotic changes in PC12 cells: Protection by trans resveratrol. Eur. J. Pharmacol. 2011, 666, 5–11. [Google Scholar] [CrossRef] [PubMed]
  264. Lanzillotta, A.; Pignataro, G.; Branca, C.; Cuomo, O.; Sarnico, I.; Benarese, M.; Annunziato, L.; Spano, P.; Pizzi, M. Targeted acetylation of NF-kappaB/RelA and histones by epigenetic drugs reduces post-ischemic brain injury in mice with an extended therapeutic window. Neurobiol. Dis. 2013, 49, 177–189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Yang, J.; Huang, J.; Shen, C.; Cheng, W.; Yu, P.; Wang, L.; Tang, F.; Guo, S.; Yang, Q.; Zhang, J. Resveratrol Treatment in Different Time-Attenuated Neuronal Apoptosis After Oxygen and Glucose Deprivation/Reoxygenation via Enhancing the Activation of Nrf-2 Signaling Pathway In Vitro. Cell Transplant. 2018, 27, 1789–1797. [Google Scholar] [CrossRef] [PubMed]
  266. Faggi, L.; Pignataro, G.; Parrella, E.; Porrini, V.; Vinciguerra, A.; Cepparulo, P.; Cuomo, O.; Lanzillotta, A.; Mota, M.; Benarese, M.; et al. Synergistic Association of Valproate and Resveratrol Reduces Brain Injury in Ischemic Stroke. Int. J. Mol. Sci. 2018, 19, 172. [Google Scholar] [CrossRef] [Green Version]
  267. Wan, D.; Zhou, Y.; Wang, K.; Hou, Y.; Hou, R.; Ye, X. Resveratrol provides neuroprotection by inhibiting phosphodiesterases and regulating the cAMP/AMPK/SIRT1 pathway after stroke in rats. Brain Res. Bull. 2016, 121, 255–262. [Google Scholar] [CrossRef]
  268. Yu, P.; Wang, L.; Tang, F.; Zeng, L.; Zhou, L.; Song, X.; Jia, W.; Chen, J.; Yang, Q. Resveratrol Pretreatment Decreases Ischemic Injury and Improves Neurological Function Via Sonic Hedgehog Signaling After Stroke in Rats. Mol. Neurobiol. 2017, 54, 212–226. [Google Scholar] [CrossRef]
  269. Alquisiras-Burgos, I.; Ortiz-Plata, A.; Franco-Pérez, J.; Millán, A.; Aguilera, P. Resveratrol reduces cerebral edema through inhibition of de novo SUR1 expression induced after focal ischemia. Exp. Neurol. 2020, 330, 113353. [Google Scholar] [CrossRef]
  270. Bonsack, F.; Alleyne, C.H., Jr.; Sukumari-Ramesh, S. Resveratrol Attenuates Neurodegeneration and Improves Neurological Outcomes after Intracerebral Hemorrhage in Mice. Front. Cell. Neurosci. 2017, 11, 228. [Google Scholar] [CrossRef]
  271. Karaoglan, A.; Akdemir, O.; Barut, S.; Kokturk, S.; Uzun, H.; Tasyurekli, M.; Colak, A. The effects of resveratrol on vasospasm after experimental subarachnoidal hemorrhage in rats. Surg. Neurol. 2008, 70, 337–343. [Google Scholar] [CrossRef]
  272. Zhou, X.M.; Zhou, M.L.; Zhang, X.S.; Zhuang, Z.; Li, T.; Shi, J.X.; Zhang, X. Resveratrol prevents neuronal apoptosis in an early brain injury model. J. Surg. Res. 2014, 189, 159–165. [Google Scholar] [CrossRef] [PubMed]
  273. Shao, A.W.; Wu, H.J.; Chen, S.; Ammar, A.B.; Zhang, J.M.; Hong, Y. Resveratrol attenuates early brain injury after subarachnoid hemorrhage through inhibition of NF-κB-dependent inflammatory/MMP-9 pathway. CNS Neurosci. Ther. 2014, 20, 182–185. [Google Scholar] [CrossRef] [PubMed]
  274. Qian, C.; Jin, J.; Chen, J.; Li, J.; Yu, X.; Mo, H.; Chen, G. SIRT1 activation by resveratrol reduces brain edema and neuronal apoptosis in an experimental rat subarachnoid hemorrhage model. Mol. Med. Rep. 2017, 16, 9627–9635. [Google Scholar] [CrossRef] [PubMed]
  275. Kalani, A.; Kamat, P.K.; Kalani, K.; Tyagi, N. Epigenetic impact of curcumin on stroke prevention. Metab. Brain Dis. 2015, 30, 427–435. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  276. Zhou, J.; Wu, N.; Lin, L. Curcumin Suppresses Apoptosis and Inflammation in Hypoxia/Reperfusion-Exposed Neurons via Wnt Signaling Pathway. Med. Sci. Monit. 2020, 26, e920445-1. [Google Scholar]
  277. Rocha-Ferreira, E.; Sisa, C.; Bright, S.; Fautz, T.; Harris, M.; Contreras Riquelme, I.; Agwu, C.; Kurulday, T.; Mistry, B.; Hill, D.; et al. Curcumin: Novel Treatment in Neonatal Hypoxic-Ischemic Brain Injury. Front. Physiol. 2019, 10, 1351. [Google Scholar] [CrossRef] [PubMed]
  278. Thiyagarajan, M.; Sharma, S.S. Neuroprotective effect of curcumin in middle cerebral artery occlusion induced focal cerebral ischemia in rats. Life Sci. 2004, 74, 969–985. [Google Scholar] [CrossRef]
  279. Li, W.; Suwanwela, N.C.; Patumraj, S. Curcumin by down-regulating NF-kB and elevating Nrf2, reduces brain edema and neurological dysfunction after cerebral I/R. Microvasc. Res. 2016, 106, 117–126. [Google Scholar] [CrossRef]
  280. Li, W.; Suwanwela, N.C.; Patumraj, S. Curcumin prevents reperfusion injury following ischemic stroke in rats via inhibition of NF-κB, ICAM-1, MMP-9 and caspase-3 expression. Mol. Med. Rep. 2017, 16, 4710–4720. [Google Scholar] [CrossRef] [Green Version]
  281. Liu, S.; Cao, Y.; Qu, M.; Zhang, Z.; Feng, L.; Ye, Z.; Xiao, M.; Hou, S.T.; Zheng, R.; Han, Z. Curcumin protects against stroke and increases levels of Notch intracellular domain. Neurol. Res. 2016, 38, 553–559. [Google Scholar] [CrossRef]
  282. Zhao, J.; Yu, S.; Zheng, W.; Feng, G.; Luo, G.; Wang, L.; Zhao, Y. Curcumin improves outcomes and attenuates focal cerebral ischemic injury via antiapoptotic mechanisms in rats. Neurochem. Res. 2010, 35, 374–379. [Google Scholar] [CrossRef] [PubMed]
  283. Shah, F.A.; Gim, S.A.; Sung, J.H.; Jeon, S.J.; Kim, M.O.; Koh, P.O. Identification of proteins regulated by curcumin in cerebral ischemia. J. Surg. Res. 2016, 201, 141–148. [Google Scholar] [CrossRef] [PubMed]
  284. Liu, Z.; Ran, Y.; Huang, S.; Wen, S.; Zhang, W.; Liu, X.; Ji, Z.; Geng, X.; Ji, X.; Du, H.; et al. Curcumin Protects against Ischemic Stroke by Titrating Microglia/Macrophage Polarization. Front. Aging Neurosci. 2017, 9, 233. [Google Scholar] [CrossRef] [PubMed]
  285. Lan, C.; Chen, X.; Zhang, Y.; Wang, W.; Wang, W.E.; Liu, Y.; Cai, Y.; Ren, H.; Zheng, S.; Zhou, L.; et al. Curcumin prevents strokes in stroke-prone spontaneously hypertensive rats by improving vascular endothelial function. BMC Cardiovasc. Disord. 2018, 18, 43. [Google Scholar] [CrossRef] [Green Version]
  286. Zálešák, F.; Bon, D.J.D.; Pospíšil, J. Lignans and Neolignans: Plant secondary metabolites as a reservoir of biologically active substances. Pharmacol. Res. 2019, 146, 104284. [Google Scholar] [CrossRef]
  287. Adolphe, J.L.; Whiting, S.J.; Juurlink, B.H.; Thorpe, L.U.; Alcorn, J. Health effects with consumption of the flax lignan secoisolariciresinol diglucoside. Br. J. Nutr. 2010, 103, 929–938. [Google Scholar]
  288. Peterson, J.; Dwyer, J.; Adlercreutz, H.; Scalbert, A.; Jacques, P.; McCullough, M.L. Dietary lignans: Physiology and potential for cardiovascular disease risk reduction. Nutr. Rev. 2010, 68, 571–603. [Google Scholar]
  289. Lapi, D.; Di Maro, M.; Mastantuono, T.; Battiloro, L.; Sabatino, L.; Muscariello, E.; Colantuoni, A. Effects of oleuropein and pinoresinol on microvascular damage induced by hypoperfusion and reperfusion in rat pial circulation. Microcirculation 2015, 22, 79–90. [Google Scholar] [CrossRef] [Green Version]
  290. Liu, Q.S.; Deng, R.; Li, S.; Li, X.; Li, K.; Kebaituli, G.; Li, X.; Liu, R. Ellagic acid protects against neuron damage in ischemic stroke through regulating the ratio of Bcl-2/Bax expression. Appl. Physiol. Nutr. Metab. 2017, 42, 855–860. [Google Scholar] [CrossRef]
  291. Ahsan, A.; Zheng, Y.R.; Wu, X.L.; Tang, W.D.; Liu, M.R.; Ma, S.J.; Jiang, L.; Hu, W.W.; Zhang, X.N.; Chen, Z. Urolithin A-activated autophagy but not mitophagy protects against ischemic neuronal injury by inhibiting ER stress in vitro and in vivo. CNS Neurosci. Ther. 2019, 25, 976–986. [Google Scholar] [CrossRef] [Green Version]
  292. Yaidikar, L.; Byna, B.; Thakur, S.R. Neuroprotective effect of punicalagin against cerebral ischemia reperfusion-induced oxidative brain injury in rats. J. Stroke Cerebrovasc. Dis. 2014, 23, 2869–2878. [Google Scholar] [CrossRef] [PubMed]
  293. Yaidikar, L.; Thakur, S. Punicalagin attenuated cerebral ischemia-reperfusion insult via inhibition of proinflammatory cytokines, up-regulation of Bcl-2, down-regulation of Bax, and caspase-3. Mol. Cell. Biochem. 2015, 402, 141–148. [Google Scholar] [CrossRef] [PubMed]
  294. Lončar, M.; Jakovljević, M.; Šubarić, D.; Pavlić, M.; Buzjak Služek, V.; Cindrić, I.; Molnar, M. Coumarins in Food and Methods of Their Determination. Foods 2020, 9, 645. [Google Scholar] [CrossRef] [PubMed]
  295. Zaragozá, C.; Monserrat, J.; Mantecón, C.; Villaescusa, L.; Zaragozá, F.; Álvarez-Mon, M. Antiplatelet activity of flavonoid and coumarin drugs. Vasc. Pharmacol. 2016, 87, 139–149. [Google Scholar] [CrossRef] [PubMed]
  296. Kontogiorgis, C.; Nicolotti, O.; Mangiatordi, G.F.; Tognolini, M.; Karalaki, F.; Giorgio, C.; Patsilinakos, A.; Carotti, A.; Hadjipavlou-Litina, D.; Barocelli, E. Studies on the antiplatelet and antithrombotic profile of anti-inflammatory coumarin derivatives. J. Enzym. Inhib. Med. Chem. 2015, 30, 925–933. [Google Scholar] [CrossRef] [PubMed]
  297. Macáková, K.; Řeháková, Z.; Mladěnka, P.; Karlíčková, J.; Filipský, T.; Říha, M.; Prasad, A.K.; Parmar, V.S.; Jahodář, L.; Pávek, P.; et al. In vitro platelet antiaggregatory properties of 4-methylcoumarins. Biochimie 2012, 94, 2681–2686. [Google Scholar] [CrossRef]
  298. Najmanová, I.; Doseděl, M.; Hrdina, R.; Anzenbacher, P.; Filipský, T.; Říha, M.; Mladěnka, P. Cardiovascular effects of coumarins besides their antioxidant activity. Curr. Top. Med. Chem. 2015, 15, 830–849. [Google Scholar] [CrossRef]
  299. Bibak, B.; Shakeri, F.; Barreto, G.E.; Keshavarzi, Z.; Sathyapalan, T.; Sahebkar, A. A review of the pharmacological and therapeutic effects of auraptene. Biofactors 2019, 45, 867–879. [Google Scholar] [CrossRef]
  300. Okuyama, S.; Minami, S.; Shimada, N.; Makihata, N.; Nakajima, M.; Furukawa, Y. Anti-inflammatory and neuroprotective effects of auraptene, a citrus coumarin, following cerebral global ischemia in mice. Eur. J. Pharmacol. 2013, 699, 118–123. [Google Scholar] [CrossRef]
  301. Okuyama, S.; Morita, M.; Kaji, M.; Amakura, Y.; Yoshimura, M.; Shimamoto, K.; Ookido, Y.; Nakajima, M.; Furukawa, Y. Auraptene Acts as an Anti-Inflammatory Agent in the Mouse Brain. Molecules 2015, 20, 20230–20239. [Google Scholar] [CrossRef] [Green Version]
  302. Fylaktakidou, K.C.; Hadjipavlou-Litina, D.J.; Litinas, K.E.; Nicolaides, D.N. Natural and synthetic coumarin derivatives with anti-inflammatory/antioxidant activities. Curr. Pharm. Des. 2004, 10, 3813–3833. [Google Scholar] [CrossRef] [PubMed]
  303. Wang, X.; Li, R.; Wang, X.; Fu, Q.; Ma, S. Umbelliferone ameliorates cerebral ischemia-reperfusion injury via upregulating the PPAR gamma expression and suppressing TXNIP/NLRP3 inflammasome. Neurosci. Lett. 2015, 600, 182–187. [Google Scholar] [CrossRef] [PubMed]
  304. Liang, C.; Ju, W.; Pei, S.; Tang, Y.; Xiao, Y. Pharmacological Activities and Synthesis of Esculetin and Its Derivatives: A Mini-Review. Molecules 2017, 22, 387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  305. Wang, C.; Pei, A.; Chen, J.; Yu, H.; Sun, M.L.; Liu, C.F.; Xu, X. A natural coumarin derivative esculetin offers neuroprotection on cerebral ischemia/reperfusion injury in mice. J. Neurochem. 2012, 121, 1007–1013. [Google Scholar] [CrossRef]
  306. Deng, M.; Xie, L.; Zhong, L.; Liao, Y.; Liu, L.; Li, X. Imperatorin: A review of its pharmacology, toxicity and pharmacokinetics. Eur. J. Pharmacol. 2020, 879, 173124. [Google Scholar] [CrossRef]
  307. Wang, N.; Wu, L.; Cao, Y.; Wang, Y.; Zhang, Y. The protective activity of imperatorin in cultured neural cells exposed to hypoxia re-oxygenation injury via anti-apoptosis. Fitoterapia 2013, 90, 38–43. [Google Scholar] [CrossRef]
  308. Mogana, R.; Teng-Jin, K.; Wiart, C. Anti-inflammatory, anticholinesterase, and antioxidant potential of scopoletin isolated from canarium patentinervium miq. burseaceae kunth. Evid. Based Complement. Altern. Med. 2013, 2013, 734824. [Google Scholar] [CrossRef] [Green Version]
  309. Connell, B.J.; Saleh, M.; Rajagopal, D.; Saleh, T.M. UPEI-400, a conjugate of lipoic acid and scopoletin, mediates neuroprotection in a rat model of ischemia/reperfusion. Food Chem. Toxicol. 2017, 100, 175–182. [Google Scholar] [CrossRef]
  310. Zafar, S.; Sarfraz, I.; Rasul, A.; Shah, M.A.; Hussain, G.; Zahoor, M.K.; Shafiq, N.; Riaz, A.; Selamoglu, Z.; Sarker, S. Osthole: A Multifunctional Natural Compound with Potential Anticancer, Antioxidant and Anti-inflammatory Activities. Mini Rev. Med. Chem. 2020. [Google Scholar] [CrossRef]
  311. Mao, X.; Yin, W.; Liu, M.; Ye, M.; Liu, P.; Liu, J.; Lian, Q.; Xu, S.; Pi, R. Osthole, a natural coumarin, improves neurobehavioral functions and reduces infarct volume and matrix metalloproteinase-9 activity after transient focal cerebral ischemia in rats. Brain Res. 2011, 1385, 275–280. [Google Scholar] [CrossRef]
  312. Li, K.; Ding, D.; Zhang, M. Neuroprotection of Osthole against Cerebral Ischemia/Reperfusion Injury through an Anti-apoptotic Pathway in Rats. Biol. Pharm. Bull. 2016, 39, 336–342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  313. Chao, X.; Zhou, J.; Chen, T.; Liu, W.; Dong, W.; Qu, Y.; Jiang, X.; Ji, X.; Zhen, H.; Fei, Z. Neuroprotective effect of osthole against acute ischemic stroke on middle cerebral ischemia occlusion in rats. Brain Res. 2010, 1363, 206–211. [Google Scholar] [CrossRef] [PubMed]
  314. Du, G.; Tu, H.; Li, X.; Pei, A.; Chen, J.; Miao, Z.; Li, J.; Wang, C.; Xie, H.; Xu, X.; et al. Daphnetin, a natural coumarin derivative, provides the neuroprotection against glutamate-induced toxicity in HT22 cells and ischemic brain injury. Neurochem. Res. 2014, 39, 269–275. [Google Scholar] [CrossRef] [PubMed]
  315. Liu, J.; Chen, Q.; Jian, Z.; Xiong, X.; Shao, L.; Jin, T.; Zhu, X.; Wang, L. Daphnetin Protects against Cerebral Ischemia/Reperfusion Injury in Mice via Inhibition of TLR4/NF-κB Signaling Pathway. Biomed. Res. Int. 2016, 2016, 2816056. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  316. He, F.J.; Nowson, C.A.; MacGregor, G.A. Fruit and vegetable consumption and stroke: Meta-analysis of cohort studies. Lancet 2006, 367, 320–326. [Google Scholar] [CrossRef]
  317. Bellone, J.A.; Murray, J.R.; Jorge, P.; Fogel, T.G.; Kim, M.; Wallace, D.R.; Hartman, R.E. Pomegranate supplementation improves cognitive and functional recovery following ischemic stroke: A randomized trial. Nutr. Neurosci. 2019, 22, 738–743. [Google Scholar] [CrossRef]
  318. Chiva-Blanch, G.; Badimon, L. Benefits and Risks of Moderate Alcohol Consumption on Cardiovascular Disease: Current Findings and Controversies. Nutrients 2019, 12, 108. [Google Scholar] [CrossRef] [Green Version]
  319. Asmaro, K.; Fu, P.; Ding, Y. Neuroprotection & mechanism of ethanol in stroke and traumatic brain injury therapy: New prospects for an ancient drug. Curr. Drug Targets 2013, 14, 74–80. [Google Scholar]
  320. Arranz, S.; Chiva-Blanch, G.; Valderas-Martínez, P.; Medina-Remón, A.; Lamuela-Raventós, R.M.; Estruch, R. Wine, beer, alcohol and polyphenols on cardiovascular disease and cancer. Nutrients 2012, 4, 759–781. [Google Scholar] [CrossRef] [Green Version]
  321. Rodríguez-Artalejo, F.; López-García, E. Coffee Consumption and Cardiovascular Disease: A Condensed Review of Epidemiological Evidence and Mechanisms. J. Agric. Food Chem. 2018, 66, 5257–5263. [Google Scholar] [CrossRef]
  322. Qian, Y.; Ye, D.; Huang, H.; Wu, D.J.H.; Zhuang, Y.; Jiang, X.; Mao, Y. Coffee Consumption and Risk of Stroke: A Mendelian Randomization Study. Ann. Neurol. 2020, 87, 525–532. [Google Scholar] [CrossRef] [PubMed]
  323. Yi, M.; Wu, X.; Zhuang, W.; Xia, L.; Chen, Y.; Zhao, R.; Wan, Q.; Du, L.; Zhou, Y. Tea Consumption and Health Outcomes: Umbrella Review of Meta-Analyses of Observational Studies in Humans. Mol. Nutr. Food Res. 2019, 63, 1900389. [Google Scholar] [CrossRef] [PubMed]
  324. Morze, J.; Schwedhelm, C.; Bencic, A.; Hoffmann, G.; Boeing, H.; Przybylowicz, K.; Schwingshackl, L. Chocolate and risk of chronic disease: A systematic review and dose-response meta-analysis. Eur. J. Nutr. 2020, 59, 389–397. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  325. Han, C.H.; Kim, M.; Cho, S.Y.; Jung, W.S.; Moon, S.K.; Park, J.M.; Ko, C.N.; Cho, K.H.; Kwon, S. Adjunctive herbal medicine treatment for patients with acute ischemic stroke: A systematic review and meta-analysis. Complement. Ther. Clin. Pract. 2018, 33, 124–137. [Google Scholar] [CrossRef]
  326. Tang, Z.; Li, M.; Zhang, X.; Hou, W. Dietary flavonoid intake and the risk of stroke: A dose-response meta-analysis of prospective cohort studies. BMJ Open 2016, 6, e008680. [Google Scholar] [CrossRef]
  327. Wang, Z.M.; Zhao, D.; Nie, Z.L.; Zhao, H.; Zhou, B.; Gao, W.; Wang, L.S.; Yang, Z.J. Flavonol intake and stroke risk: A meta-analysis of cohort studies. Nutrition 2014, 30, 518–523. [Google Scholar] [CrossRef]
  328. Hollman, P.C.; Geelen, A.; Kromhout, D. Dietary flavonol intake may lower stroke risk in men and women. J. Nutr. 2010, 140, 600–604. [Google Scholar] [CrossRef] [Green Version]
  329. Knekt, P.; Isotupa, S.; Rissanen, H.; Heliövaara, M.; Järvinen, R.; Häkkinen, S.; Aromaa, A.; Reunanen, A. Quercetin intake and the incidence of cerebrovascular disease. Eur. J. Clin. Nutr. 2000, 54, 415–417. [Google Scholar] [CrossRef] [Green Version]
  330. Applová, L.; Karlíčková, J.; Warncke, P.; Macáková, K.; Hrubša, M.; Macháček, M.; Tvrdý, V.; Fischer, D.; Mladěnka, P. 4-Methylcatechol, a Flavonoid Metabolite with Potent Antiplatelet Effects. Mol. Nutr. Food Res. 2019, 63, 1900261. [Google Scholar] [CrossRef]
  331. Serban, M.C.; Sahebkar, A.; Zanchetti, A.; Mikhailidis, D.P.; 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] [Green Version]
  332. Wang, L.; Cao, D.; Wu, H.; Jia, H.; Yang, C.; Zhang, L. Fisetin Prolongs Therapy Window of Brain Ischemic Stroke Using Tissue Plasminogen Activator: A Double-Blind Randomized Placebo-Controlled Clinical Trial. Clin. Appl. Thromb. Hemost. 2019, 25, 1076029619871359. [Google Scholar] [CrossRef] [PubMed]
  333. Wang, X.H.; You, Y.P. Epigallocatechin Gallate Extends Therapeutic Window of Recombinant Tissue Plasminogen Activator Treatment for Brain Ischemic Stroke: A Randomized Double-Blind and Placebo-Controlled Trial. Clin. Neuropharmacol. 2017, 40, 24–28. [Google Scholar] [CrossRef] [PubMed]
  334. Cassidy, A.; Rimm, E.B.; O’Reilly, E.J.; Logroscino, G.; Kay, C.; Chiuve, S.E.; Rexrode, K.M. Dietary flavonoids and risk of stroke in women. Stroke 2012, 43, 946–951. [Google Scholar] [CrossRef] [PubMed]
  335. Caltagirone, C.; Cisari, C.; Schievano, C.; Di Paola, R.; Cordaro, M.; Bruschetta, G.; Esposito, E.; Cuzzocrea, S.; Stroke Study Group. Coultramicronized palmitoylethanolamide/luteolin in the treatment of cerebral ischemia: From rodent to man. Transl. Stroke Res. 2016, 7, 54–69. [Google Scholar] [CrossRef] [Green Version]
  336. Kokubo, Y.; Iso, H.; Ishihara, J.; Okada, K.; Inoue, M.; Tsugane, S.; JPHC Study Group. Association of dietary intake of soy, beans, and isoflavones with risk of cerebral and myocardial infarctions in Japanese populations: The Japan Public Health Center-based (JPHC) study cohort I. Circulation 2007, 116, 2553–2562. [Google Scholar] [CrossRef] [Green Version]
  337. Liang, W.; Lee, A.H.; Binns, C.W.; Huang, R.; Hu, D.; Shao, H. Soy consumption reduces risk of ischemic stroke: A case-control study in southern china. Neuroepidemiology 2009, 33, 111–116. [Google Scholar] [CrossRef]
  338. Zheng, Q.H.; Li, X.L.; Mei, Z.G.; Xiong, L.; Mei, Q.X.; Wang, J.F.; Tan, L.J.; Yang, S.B.; Feng, Z.T. Efficacy and safety of puerarin injection in curing acute ischemic stroke: A meta-analysis of randomized controlled trials. Medicine 2017, 96, e5803. [Google Scholar] [CrossRef]
  339. Kimble, R.; Keane, K.M.; Lodge, J.K.; Howatson, G. Dietary intake of anthocyanins and risk of cardiovascular disease: A systematic review and meta-analysis of prospective cohort studies. Crit. Rev. Food Sci. Nutr. 2019, 59, 3032–3043. [Google Scholar] [CrossRef]
  340. Coman, V.; Vodnar, D.C. Hydroxycinnamic acids and human health: Recent advances. J. Sci. Food Agric. 2020, 100, 483–499. [Google Scholar] [CrossRef]
  341. Onakpoya, I.J.; Spencer, E.A.; Thompson, M.J.; Heneghan, C.J. The effect of chlorogenic acid on blood pressure: A systematic review and meta-analysis of randomized clinical trials. J. Hum. Hypertens. 2015, 29, 77–81. [Google Scholar] [CrossRef]
  342. Fodor, K.; Tit, D.M.; Pasca, B.; Bustea, C.; Uivarosan, D.; Endres, L.; Iovan, C.; Abdel-Daim, M.M.; Bungau, S. Long-Term Resveratrol Supplementation as a Secondary Prophylaxis for Stroke. Oxid. Med. Cell. Longev. 2018, 2018, 4147320. [Google Scholar] [CrossRef] [PubMed]
  343. Chen, J.; Bai, Q.; Zhao, Z.; Sui, H.; Xie, X. Resveratrol improves delayed r-tPA treatment outcome by reducing MMPs. Acta Neurol. Scand. 2016, 134, 54–60. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Major polyphenol classes and their most important food sources.
Figure 1. Major polyphenol classes and their most important food sources.
Nutrients 13 00085 g001
Table 1. Polyphenols active in preclinical stroke models.
Table 1. Polyphenols active in preclinical stroke models.
Polyphenol ClassPolyphenolType of StrokeExperimental ModelEffective Dose and TreatmentFindingsReferences
FlavonolsQuercetinIschemic StroketMCAO rat modelPost; 20 mg/kg; daily for 3 days
Pre; 25 mg/kg; daily for 21 days
↓brain damage, oxidative stress, apoptosis, neurological deficits[69,70]
Ischemic StrokepMCAO rat modelPost; 30 mg/kg; single administration↓brain damage[71]
Ischemic StrokePhotothrombotic rat modelPost; 25 µmol/kg; every 12 h for 3 days
Post; 50 mg/kg; every 12 h for 3 days
↓BBB damage, neurological deficits[72,73]
Ischemic Stroke2VO mouse modelPre; 50 mg/kg; 30 min before and immediately after stroke, then daily for 2 days↓ brain damage, BBB damage[74]
ICHCollagenase infusion rat modelPost; 30 mg/kg; single administration
Post; 50 mg/kg; single administration
↓brain damage, oxidative stress, apoptosis, inflammation, neurological deficits[75,76]
SAHBlood infusion rat modelPost; 10 mg/kg; every 8 h for 2 days
Post; 50 mg/kg; 30 min, 12 h and 24 h after SAH
↓brain damage, oxidative stress, vasospasm, apoptosis, neurological deficits[77,78]
MyricetinIschemic StrokeOGD SH-SY5Y cellsPre; 0.1 nM↓toxicity[80]
Ischemic StrokeOGD human BMVECsPre; 30 µM↓oxidative stress, inflammation, endothelial permeability[82]
Ischemic StroketMCAO rat modelPre; 20 mg/kg; 2 h before and daily for 2 days after stroke
Pre; 25 mg/kg; daily for 7 days
↓brain damage, oxidative stress, apoptosis, inflammation, neurological deficits[80,81]
KaempferolIschemic StroketMCAO rat modelPre; 100 and 200 µM; 30 min before and immediately after reperfusion
Post; 10 and 7.5 mg/kg; single administration
Post; 100 mg/kg; daily for 7 days
↓brain damage, oxidative stress, apoptosis, inflammation, BBB damage, neurological deficits[84,85,86]
FisetinIschemic StroketMCAO mouse modelPre; 50 mg/kg; single administration
Post; 50 mg/kg; single administration
↓brain damage, inflammation[89]
Ischemic StrokepMCAO rat modelPost; 30 mg/kg; single administration↓brain damage[71]
Ischemic StrokeSCE rabbit modelPost; 50 mg/kg; single administration↓ neurological deficits[88]
MorinIschemic StroketMCAO rat modelPost; 30 mg/kg; daily for 7 days
Post; 30 mg/kg; single administration
↓brain damage, BBB damage, inflammation, oxidative stress, apoptosis, neurological deficits[91,92]
Flavan-3-olsEGCGIschemic StrokeOGD PC12 cellsPre; 2 µM↓toxicity, apoptosis[97]
Ischemic StroketMCAO mouse modelPost; 50 mg/kg; single administration↓brain damage, MPPs[98]
Ischemic StroketMCAO rat modelPost; 50 mg/kg; single administration
Post; 50 mg/kg; single administration
Post; 50 mg/kg; daily for 3 days
Post; 50 mg/kg; single administration
Post; 20 mg/kg; single administration
Pre; 40 mg/kg; daily for 3 days
↓brain damage, oxidative stress, apoptosis, inflammation, neurological deficits[99,100,101,102,103,104]
Ischemic Stroke2VO gerbil modelPost; 25 and 30 mg/kg; single administration
Pre; 50 mg/kg; 30 min before and immediately after stroke
↓brain damage, oxidative stress[105,106]
Ischemic StroketMCAO rat model treated with rtPAPost; 20 mg/kg; single administration↓brain damage, BBB damage, neurological deficits[107]
SAHHb PC12 cellsPre; 1–50 µM
Pre; 1–50 µM
Pre, 50 µM
↓cell proliferation[108,109,110]
SAHHb mouse model Pre; 50 mg/kg; daily for 14 days
Pre; 50 mg/kg; daily for 14 days
↓brain damage, oxidative stress, mitochondrial dysfunction, apoptosis, neurological deficits[109,110]
ECGIschemic StrokeOGD PC12 cellsPre; 2 µM↓toxicity, apoptosis[97]
Ischemic StrokeOGD human BMVECsPre; 2 µM ↓oxidative stress, apoptosis[111]
ECIschemic StrokeOGD mouse neuronsPre; 100 µM↓toxicity, oxidative stress[113]
Ischemic StroketMCAO mouse modelPre; 30 mg/kg; single administration
Post; 30 mg/kg; single administration
↓brain damage, neurological deficits[112]
Ischemic StrokepMCAO mouse modelPre; 15 mg/kg; single administration↓brain damage, inflammation[113]
ICHHb mouse astrocytesPre; 100 µM↓oxidative stress[114]
ICHCollagenase infusion mouse modelPost; 15 mg/kg; 3 h after ICH and daily for 3 days↓brain damage, oxidative stress, neurological deficits[115]
ICHBlood infusion mouse modelPost; 15 mg/kg; 3 h after ICH and daily for 3 days↓brain damage, oxidative stress, neurological deficits[115]
ICHThrombin infusion mouse modelPost; 15 mg/kg; 3 h after ICH and daily for 3 days↓brain damage, oxidative stress, neurological deficits[115]
FlavanonesEriodictyolIschemic StrokeOGD rat astrocytesPre; 20–80 µM↓toxicity[120]
Ischemic StroketMCAO rat modelPre; 30 mg/kg; daily for 5 days↓brain damage, neurological deficits[120]
Ischemic StrokepMCAO mouse modelPre; 4 mg/kg; 30 min before, 2 h and daily for 5 days after stroke↓brain damage, inflammation, neurological and memory deficits[119]
NaringeninIschemic StrokeOGD rat neuronsPre; 80 µM
Post; 80 µM
↓toxicity, oxidative stress, apoptosis, mitochondrial dysfunction[122,123]
Ischemic StrokeOGD human MSCsPost; 40 and 80 µM↓ inflammation[124]
Ischemic StrokepMACO rat modelPre; 100 mg/kg; daily for 4 days↓brain damage, inflammation, neurological deficits[126]
Ischemic StroketMCAO rat modelPost; 80 µM; single administration
Pre; 50 mg/kg; daily for 21 days
↓brain damage, oxidative stress, apoptosis, inflammation, neurological deficits[123,125]
Ischemic StrokeStroke-prone spontaneously hypertensive rats9.0 and 17.7 mg/kg; daily for 4 weeks↓blood pressure, thrombotic tendency, oxidative stress[127]
HesperetinIschemic StroketMCAO rat modelPre; 50 mg/kg; daily for 15 days↓brain damage, oxidative stress, inflammation, neurological and behavioral deficits[129]
Ischemic Stroke2VO rat modelPre; 20 mg/kg; daily for 14 days↓brain damage, oxidative stress, inflammation[130]
Ischemic StrokeStroke-prone spontaneously hypertensive rats14.5 (hesperidin), 16.2 and 31.6 (glucosyl hesperidin) mg/kg; daily for 4 weeks↓blood pressure, thrombotic tendency, oxidative stress[127]
SAHBlood infusion rat modelPost; 50 and 100 mg/kg; daily for 2 days↓vessel wall thickness, ↑vessel luminal diameter[131]
PinocembrinIschemic StrokeOGD rat neuronsPost; 0.1, 1 and 10 µM↓toxicity, oxidative stress, apoptosis[133]
Ischemic StrokeOGD rat BMVECsPost; 1 and 10 µM↓toxicity[134]
Ischemic StrokeOGD human BMECs and astrocytesPost; 1 µM↓endothelial permeability[135]
Ischemic StroketMCAO rat modelPost; 3–30 mg/kg; single administration
Post; 1–10 mg/kg; single administration
↓brain damage, oxidative stress, apoptosis, neurological deficits[133,136]
Ischemic StrokepMCAO rat modelPost; 3, 10 and 30 mg/kg; single administration↓brain damage, apoptosis, inflammation, neurological deficits[137]
Ischemic Stroke2VO rat modelPre; 10 mg/kg; daily for 7 days↓brain damage, oxidative stress, apoptosis, inflammation[138]
Ischemic Stroke4VO rat modelPost; 1–10 mg/kg; single administration↓brain and BBB damage, neurological deficits[134]
Ischemic StrokeThromboembolic rat model treated with rtPAPost; 10 mg/kg; single administration or daily for 7 days↓brain and BBB damage, neurological deficits[135]
ICHCollagenase infusion mouse modelPost; 5 mg/kg; 2 h after ICH and every 12 h for 3 days↓brain damage, inflammation, neurological deficits[139]
FlavonesLuteolinNHIBOGD rat neuronsPre; 10–30 µM↓toxicity, oxidative stress, apoptosis[142]
Ischemic StrokeOGD mouse neuronsPost; 0.1–1 µM↓toxicity[143]
Ischemic StrokeOGD mouse MCsPre; 10 and 100 nM ↓MC degranulation [143]
Ischemic StrokeOGD human BMVECsPost; 90 µM↓toxicity, apoptosis[144]
Ischemic StroketMCAO rat modelPost; 20–80 mg/kg; 0 and 12 h after stroke↓brain damage, inflammation, neurological deficits[145]
Ischemic StrokepMCAO rat modelPost; 10 and 25 mg/kg; single administration
Post; 10 and 25 mg/kg; single administration
Post; 10 and 25 mg/kg; 0 h and daily for 3 days after stroke
↓brain damage, oxidative stress, apoptosis, neurological deficits [144,146,147]
ICHHb rat neuronsPost; 10 µM↓toxicity[148]
ICHHb rat neurons and microgliaPost; 10 and 20 µM↓ microglia activation and induced neurotoxicity[149]
ICHBlood infusion rat modelPost; 10 and 20 mg/kg; single administration
Post; 10 and 20 mg/kg; 30 min, 12 h and 24 h after ICH
↓brain damage, inflammation, neurological and memory deficits[148,149]
ApigeninIschemic StrokeOGD human BMVECsPre; 5 µM ↓toxicity, apoptosis, autophagy[151]
Ischemic StroketMCAO rat modelPost; 25 mg/kg; daily for 7 or 14 days
Post; 20 and 40 mg/kg; daily for 28 days
↓brain damage, neurological and memory deficits[151,152]
SAHEndovascular perforation rat modelPost; 20 mg/kg; single administration↓brain and BBB damage, inflammation, neurological deficits[153]
NobiletinIschemic StrokeHypoxia HepG2 cellsPost; 100 µg/mL ↓toxicity[160]
Ischemic StroketMCAO rat modelPost; 15 mg/kg; 0 and 1 h after stroke↓brain damage, apoptosis, inflammation, neurological deficits[157,160]
Ischemic Stroke2VO mouse modelPre; 50 mg/kg; daily for 7 days before and 7 days after stroke↓brain damage, memory deficits[156]
TangeretinIschemic StrokeOGD human BMVECsPost; 2.5, 5 and 10 µM↓toxicity, oxidative stress, apoptosis[159]
Ischemic StrokeHypoxia HepG2 cellsPost; 10 and 100 µg/mL↓toxicity[160]
BaicaleinIschemic StrokeOGD SH-SY5Y cellsPre; 0.1, 1 and 10 µM↓toxicity[163]
Ischemic StroketMCAO rat modelPost; 25 and 50 mg/kg; single administration
Post; 50, 100 and 200 mg/kg; single administration
↓brain damage, oxidative stress, apoptosis, neurological deficits[163,164]
Ischemic Stroke2VO gerbil modelPost; 100 and 200 mg/kg; single administration↓brain damage, oxidative stress, apoptosis[165]
AcacetinIschemic StroketMCAO mouse modelPost; 25 mg/kg; single administration↓brain damage, inflammation, neurological deficits[167]
IsoflavoneGenisteinIschemic StrokeOGD rat neuronsConcomitant, 1 mM↓toxicity, apoptosis[172]
Ischemic StrokeOGD PC12 cells Concomitant, 30 µM↓toxicity, apoptosis[173]
Ischemic StroketMCAO mouse and rat modelsPre; 5 and 10 mg/kg; daily for 14 days
Pre; 10 mg/kg; daily for 14 days
Pre; 500 mg/kg (genistein), 250 mg/kg (equol); daily for 14 days
Pre; 500 mg/kg; daily for 28 days before and 1 day after stroke
Post; 10 mg/kg; daily for 3 days
Post; 1 and 2 mg/kg; single administration
Pre; 10 mg/kg; daily for 14 days
↓brain damage, oxidative stress, apoptosis, neurological deficits ↑circulatory function[174,175,176,177,178,179,180]
Ischemic StrokepMCAO rat modelPost; 10 mg/kg; single administration↓brain damage, oxidative stress, apoptosis, neurological deficits[181]
Ischemic StrokePhotothrombotic rat modelPre; 16 mg/kg; every 6 h from 24 h before to 24 h after stroke↓brain and BBB damage[182]
Ischemic Stroke2VO mouse model treated with streptozotocinPre; 5 and 10 mg/kg; daily for 14 days before and 1 day after stroke↓brain damage, oxidative stress, apoptosis, neurological deficits[183]
Ischemic Stroke4VO rat modelPre; 0.1 mg/kg; daily for 7 days before and 7 days after stroke
Post; 1 mg/kg; single administration
Pre; 15 mg/kg; 30 min before and 24 h after stroke
↓brain damage, oxidative stress, apoptosis, neurological deficits[184,185,186]
Ischemic Stroke2VO gerbil modelPost; 3 and 10 mg/kg; single administration ↓brain damage, oxidative stress, behavioral and memory deficits[187]
SAHBlood infusion rat and dog modelsPost; 10 µM; single administration
Post; 14.0 and 17.2 µM; single administration
↓vasospasm[188,189]
DaidzeinNHIBIOD rat NSCsPre; 20–100 µM↓toxicity, apoptosis[194]
Ischemic StrokeOGD rat astrocytes Concomitant, 12 and 48 µg/mL↓toxicity, apoptosis[195]
Ischemic StroketMCAO mouse and rat modelsPost; 10 mg/kg; daily for 7 days, then every other day up to 1 month
Pre; 50 and 100 mg/kg; daily for 7 days
Pre; 36 and 54 mg/kg; daily for 5 days
Pre; 50 and 100 mg/kg; single administration
Post; 50 and 100 mg/kg; daily for 14 days
↑cholesterol homeostasis, ↓brain damage, inflammation, neurological, motor and memory deficits[191,196,197,198,199]
Ischemic StrokepMCAO mouse and rat modelsPost; 0.10 mg/kg; daily for 14 days
Post; 10 mg/kg; single administration
↓brain damage, oxidative stress, apoptosis, behavioral deficits[190,192]
Ischemic StrokeStroke-prone spontaneously hypertensive rats100 mg/kg; daily for 14 days↓arterial dysfunction, blood pressure[200]
SAHEndovascular perforation mouse modelPre; 100 mg/kg; single administration↓brain and BBB damage, oxidative stress, apoptosis, neurological deficits [201]
Biochanin AIschemic StroketMCAO mouse and rat modelsPre; 5 and 10 mg/kg; daily for 4 weeks
Pre; 10, 20 and 40 mg/kg; daily for 14 days
↓brain damage, oxidative stress, neurological deficits[203,204]
SAHBlood infusion rat modelPost; 20 and 40 mg/kg; single administration↓brain damage, apoptosis, inflammation, neurological and memory deficits[205]
AnthocyaninsCyanidinIschemic StrokeOGD PC12 cellsConcomitant, 10 µg/mL↓toxicity[209]
Ischemic StroketMCAO mouse modelPost; 150 and 300 mg/kg; daily for 7 days
Post; 10 mg/kg; single administration
↓brain damage, oxidative stress [208,209]
Ischemic Stroke2VO rat modelPre; 10 mg/kg; 1 h before stroke/post; 10 mg/kg; 0.5 h after reperfusion
Pre; 10 and 20 mg/kg; 10 min before stroke and during reperfusion
↓brain damage, oxidative stress, microvascular permeability[210,212]
Ischemic StrokepMCAO mouse modelPre; 2 and 5 mg/kg; 1 h before stroke/post; 2 mg/kg; 3 h after stroke↓brain damage, oxidative stress, neurological deficits[211]
PhloretinIschemic StroketMCAO rat modelPre; 40 and 80 mg/kg; daily for 14 days ↓brain damage, oxidative stress, neurological deficits[214]
DihydrochalconesGSPE NHIBICarotid ligation mouse modelPre; 30 mg/kg; single administration↓brain damage, apoptosis, neurological deficits[216]
Proanthocyanidins ProcyanidinsIschemic StroketMCAO mouse, rat modelsPost; 20 and 40 mg/kg; daily for 2–14 days
Pre; 50 mg/kg; daily for 14 days
↓brain and BBB damage, oxidative stress, apoptosis, neurological deficits[217,218]
SAHBlood infusion rabbit modelPost; 10 and 50 mg/kg; daily for 3 days↓vasospasm[219]
Hydroxycinnamic acidsCaffeic acidNHIBICarotid ligation mouse modelPre; 40 mg/kg; 30 min before and daily for 7 days after stroke/post; 40 mg/kg; single administration↓brain damage, apoptosis, inflammation[223]
Ischemic StroketMCAO rat modelPre; 50 mg/kg; 30 min before, 0 h, 1 h and 2 h after stroke, then every 12 h for 4 days
Pre; 0.1–10 µg/kg; single administration
↓brain damage, inflammation, neurological deficits[224,227]
Ischemic StrokepMCAO rabbit modelPre; 10 µmol/kg; daily for 7 days↓brain damage, oxidative stress, neurological deficits[226]
Ischemic StrokePhotothrombotic mouse model Post; 2 and 5 mg/kg; 1 and 6 h after stroke↓brain damage, inflammation[225]
Ischemic Stroke2VO rat model with hypotensionPost; 10, 30 and 50 mg/kg; single administration↓brain damage, oxidative stress, memory deficits[228]
Ferulic acidIschemic StrokeOGD PC12 cellsPre; 80 and 100 µM↓toxicity, oxidative stress[230]
Ischemic StroketMCAO rat modelPost; 28, 56 and 112 mg/kg; daily for 5 days
Post; 100 and 200 mg/kg; daily for 7 days
Concomitant; 100 mg/kg; single administration
Concomitant, 80 and 100 mg/kg/post; 100 mg/kg; single administrations
↓brain damage, oxidative stress, apoptosis, inflammation, neurological and memory deficits[231,232,233,234]
Ischemic Stroke4VO rat modelPre; 20 and 25 mg/kg; daily 4 days before, immediately after stroke and during reperfusion↓brain damage, oxidative stress, neurological and memory deficits[230]
Sinapic acidIschemic Stroke4VO rat modelPost; 3 and 10 mg/kg; 0 and 90 min after stroke or daily for 14 days ↓brain damage, memory deficits[236]
p-Coumaric acidIschemic StrokepMCAO rat modelPost; 100 mg/kg; single administration↓brain damage, oxidative stress, apoptosis, neurological deficits[238]
Ischemic Stroke2VO mouse modelPre; 100 mg/kg; daily for 14 days↓brain damage, oxidative stress[239]
Chlorogenic acidIschemic StroketMCAO rat modelPost; 30 mg/kg; 0 and 2 h after stroke
Post; 30 mg/kg; 0 and 2 h after stroke
Pre; 15, 30 and 60 mg/kg; daily for 7 days
↓brain and BBB damage, oxidative stress, MMPs level, behavioral deficits[240,241,242]
Ischemic Stroke2VO rat modelPost, 100 and 500 mg/kg↓brain damage, oxidative stress, apoptosis, MMPs level, memory deficits[243]
Ischemic StrokeSCE rabbit model treated with rtPAPost; 50 mg/kg; single administration↓behavioral deficits[244]
Rosmarinic acid Ischemic StrokeOGD SH-SY5Y cellsPre; 1 and 10 µM
Post; 3–81 µM
↓toxicity, apoptosis[246,247]
Ischemic StroketMCAO mouse modelPost; 20 and 40 mg/kg; single administration↓brain damage, oxidative stress, apoptosis [248]
Ischemic StroketMCAO rat model treated with streptozotocinPost; 50 mg/kg; single administration↓brain and BBB damage, neurological deficits[247]
Ischemic StrokepMCAO mouse modelPre; 1 and 20 mg/kg; 30 min before, 1 h after stroke and then daily for 5 days↓brain damage, inflammation, neurological and memory deficits[249]
HydroxybenzoicacidsGallic acidIschemic StrokeOGD rat neuronsConcomitant, 50 µM↓toxicity[251]
Ischemic StroketMCAO rat modelPre; 50 mg/kg; daily for 7 days
Pre; 50 mg/kg; single administration
↓brain damage, oxidative stress, apoptosis, inflammation, mitochondrial dysfunction, neurological deficits[251,252]
Ischemic Stroke4VO rat model exposed to particulate matterPre; 100 mg/kg; daily for 10 days↓BBB damage, oxidative stress, behavioral deficits[253]
Ischemic Stroke2VO mouse modelPost; 25 and 50 mg/kg; daily for 7 days↓oxidative stress, depressive symptoms [254]
StilbenesResveratrolNHIBICarotid ligation mouse and rat modelsPost; 100 mg/kg; 0, 8 and 18 h after stroke
Pre; 0.2 and 20 mg/kg; single administration/post; 20 mg/kg; single administration
Pre; 20 mg/kg; single administration
↓brain damage, apoptosis, inflammation, behavioral deficits[260,261,262]
Ischemic StrokeOGD mouse and rat neurons, PC12 cellsPre, post, pre and post; 5–25 µM
Post; 30 µM
Pre and post, post; 10–80 µM
Pre; 40 µM
↓toxicity, oxidative stress, apoptosis[263,264,265,268]
Ischemic StroketMCAO mouse, rat modelsPost; 6.8 mg/kg; single administration
Post; 6.8 mg/kg; single administration
Pre; 20 and 30 mg/kg; daily for 5 days
Pre; 30 mg/kg; daily for 7 days and 30 min before stroke
Post; 1.9 mg/kg; single administration
↓brain and BBB damage, apoptosis, neurological deficits[264,266,267,268,269]
ICHCollagenase infusion mouse modelPost; 10 mg/kg; single administration↓brain damage, apoptosis, inflammation, neurological deficits[270]
SAHBlood infusion rat modelPost; 10 mg/kg; daily for 3 days
Post; 60 mg/kg; 2 and 24 h after SAH
↓vasospasm, apoptosis[271,272]
SAHEndovascular perforation rat modelPost; 30 mg/kg; 0 and 6 h after SAH
Pre; 100 mg/kg; single administration
↓brain and BBB damage, apoptosis, neurological deficits[273,274]
CurcuminoidsCurcuminNHIBICarotid ligation mouse modelPre; 100 µg/kg; single administration/post; 50–200 µg/kg; single administration↓ brain damage, oxidative stress, apoptosis, inflammation[277]
Ischemic StrokeOGD rat neuronsPre; 0.5–8 µM ↓toxicity, apoptosis, inflammation[276]
Ischemic StroketMCAO rat modelPost; 100 and 300 mg/kg; single administration
Post; 300 mg/kg; single administration
Post; 300 mg/kg; single administration
Post; 300 mg/kg; daily for 7 days
Post; 100, 300 and 500 mg/kg; single administration
↓brain damage, oxidative stress, apoptosis, inflammation, neurological deficits[278,279,280,281,282]
Ischemic StrokepMCAO mouse and rat modelsPost; 50 mg/kg; single administration
Post; 150 mg/kg; 0 and 24 h after stroke
↓brain damage, inflammation, neurological and behavioral deficits[283,284]
Ischemic StrokeStroke-prone spontaneously hypertensive rats100 mg/kg; daily for 4 weeks↓arterial dysfunction, oxidative stress, ↑survival[285]
LignansPinoresinolIschemic Stroke2VO rat modelPre; 1 and 2 mg/kg; 10 min before stroke and during reperfusion↓microvascular damage, oxidative stress[289]
Ellagitannins and ellagic acidEllagic acidIschemic StrokeOGD rat neuronsPost; 10 and 30 µM↓ toxicity, apoptosis[290]
Ischemic StrokePhotothrombotic rat modelPre; 10 and 30 mg/kg; 24 h before and immediately after stroke↓neurological deficits[290]
Urolithin AIschemic StrokeOGD mouse neurons and N2a cells Pre; 3–30 µM↓ toxicity[291]
Ischemic StroketMCAO mouse modelPre; 2.5 and 5.0 mg/kg; 24 and 1 h before stroke↓ brain damage, neurological deficits[291]
PunicalaginIschemic StroketMCAO rat modelPre; 15 and 30 mg/kg; daily for 7 days
Pre; 15 and 30 mg/kg; daily for 7 days
↓brain damage, oxidative stress, apoptosis, inflammation, neurological deficits[292,293]
CoumarinsAurapteneIschemic Stroke2VO mouse modelPost; 25 mg/kg; daily for 8 days
Pre; 10 and 25 mg/kg; daily for 5 days
↓brain damage, inflammation[300,301]
UmbelliferoneIschemic StroketMCAO rat modelPre; 15 and 30 mg/kg; daily for 7 days↓brain damage, oxidative stress, inflammation, neurological deficits[303]
EsculetinIschemic StroketMCAO mouse modelPre; 50 and 100 mg/kg;
single administration/post; 100 mg/kg; single administration
↓brain damage, apoptosis, neurological deficits[305]
ImperatorinIschemic StrokeOGD SH-SY5Y cellsConcomitant, 2.56 µM↓apoptosis[307]
Ischemic StroketMCAO rat modelPre; 10 mg/kg; single administration↓brain damage, neurological deficits[307]
ScopoletinIschemic StroketMCAO rat modelPre; 1 mg/kg; single administration↓brain damage[309]
OstholeIschemic StroketMCAO rat modelPre; 100 mg/kg; 30 min before stroke and immediately after reperfusion
Pre; 40 mg/kg; single administration
Pre; 20 and 40 mg/kg; single administration
↓brain damage, oxidative stress, apoptosis, MMPs levels, neurological deficits[311,312,313]
DaphnetinNHIBICarotid ligation rat modelPre; 10 mg/kg; single administration/post; 10 mg/kg; single administration↓brain damage[314]
Ischemic StroketMCAO mouse modelPre; 1 mg/kg; single administration↓brain damage, apoptosis, inflammation, neurological deficits[314,315]
BMVECs: brain microvascular endothelial cells; EC: (−)-epicatechin; ECG: (−)-epicatechin-3-gallate; EGCG: epigallocatechin-3-gallate; GSPE: grape seed proanthocyanidin extract; Hb: hemoglobin; ICH: intracerebral hemorrhage; MCs: mast cells; MMPs: metalloproteinases; MSCs: mesenchymal stem cells; NHIBI: neonatal hypoxic–ischemic brain injury; NSCs: neural stem cells; OD: oxygen deprivation; OGD: oxygen and glucose deprivation; pMCAO: permanent middle cerebral artery occlusion; rtPA: recombinant tissue plasminogen activator; SAH: subarachnoid hemorrhage; SCE: small clot embolism; tMCAO: middle cerebral artery occlusion; 2VO: two-vessel occlusion; 4VO: four-vessel occlusion.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Parrella, E.; Gussago, C.; Porrini, V.; Benarese, M.; Pizzi, M. From Preclinical Stroke Models to Humans: Polyphenols in the Prevention and Treatment of Stroke. Nutrients 2021, 13, 85. https://doi.org/10.3390/nu13010085

AMA Style

Parrella E, Gussago C, Porrini V, Benarese M, Pizzi M. From Preclinical Stroke Models to Humans: Polyphenols in the Prevention and Treatment of Stroke. Nutrients. 2021; 13(1):85. https://doi.org/10.3390/nu13010085

Chicago/Turabian Style

Parrella, Edoardo, Cristina Gussago, Vanessa Porrini, Marina Benarese, and Marina Pizzi. 2021. "From Preclinical Stroke Models to Humans: Polyphenols in the Prevention and Treatment of Stroke" Nutrients 13, no. 1: 85. https://doi.org/10.3390/nu13010085

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