Next Article in Journal
Synergistic Effect of Bioactive Monoterpenes against the Mosquito, Culex pipiens (Diptera: Culicidae)
Next Article in Special Issue
Antioxidant and Antimicrobial Activities of Erodium arborescens Aerial Part Extracts and Characterization by LC-HESI-MS2 of Its Acetone Extract
Previous Article in Journal
Optically Transparent Gold Nanoparticles for DSSC Counter-Electrode: An Electrochemical Characterization
Previous Article in Special Issue
Effects of Frankincense Compounds on Infection, Inflammation, and Oral Health
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Polyphenols for the Treatment of Ischemic Stroke: New Applications and Insights

1
Department of Neurosurgery, The First Hospital of Jilin University, Changchun 130021, China
2
Department of Orthopaedic Surgery, China-Japan Union Hospital of Jilin University, Changchun 130033, China
3
Key Laboratory of Organ Regeneration and Transplantation of Ministry of Education, The First Hospital of Jilin University, Changchun 130061, China
4
National-Local Joint Engineering Laboratory of Animal Models for Human Diseases, Changchun 130061, China
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(13), 4181; https://doi.org/10.3390/molecules27134181
Submission received: 11 June 2022 / Revised: 23 June 2022 / Accepted: 28 June 2022 / Published: 29 June 2022
(This article belongs to the Special Issue Discovery of Bioactive Ingredients from Natural Products II)

Abstract

:
Ischemic stroke (IS) is a leading cause of death and disability worldwide. Currently, the main therapeutic strategy involves the use of intravenous thrombolysis to restore cerebral blood flow to prevent the transition of the penumbra to the infarct core. However, due to various limitations and complications, including the narrow time window in which this approach is effective, less than 10% of patients benefit from such therapy. Thus, there is an urgent need for alternative therapeutic strategies, with neuroprotection against the ischemic cascade response after IS being one of the most promising options. In the past few decades, polyphenolic compounds have shown great potential in animal models of IS because of their high biocompatibility and ability to target multiple ischemic cascade signaling pathways, although low bioavailability is an issue that limits the applications of several polyphenols. Here, we review the pathophysiological changes following cerebral ischemia and summarize the research progress regarding the applications of polyphenolic compounds in the treatment of IS over the past 5 years. Furthermore, we discuss several potential strategies for improving the bioavailability of polyphenolic compounds as well as some essential issues that remain to be addressed for the translation of the related therapies to the clinic.

Graphical Abstract

1. Introduction

Stroke is a common neurological disorder—the second leading cause of death and the third leading cause of disability in adults worldwide—and affects approximately a quarter of all individuals in their lifetime [1,2,3]. With population growth, an increase in life expectancy, and increased exposure to risks such as hypertension, hyperglycemia, hyperlipidemia, environmental particulate matter pollution, and high body mass index, the prevalence of stroke is expected to further increase, thus placing a heavy burden on both individuals and societies [1,2,4].
Based on the underlying pathology, stroke can be characterized as ischemic stroke (IS) or hemorrhagic stroke [1]. Thrombosis, embolisms, and systemic hypoperfusion can result in IS, which accounts for 62.4% of all stroke events [5]. When IS occurs, blood flow in the infarcted area rapidly drops below the critical levels, and the electrical activity of neurons ceases within seconds [5,6]. Brain tissue that is supplied entirely by the blocked vessels will suffer irreversible neuronal damage or death within minutes of the perfusion cut-off—this region is known as the “infarct core” [7]. Collateral circulation supply, such as from the circle of Willis and leptomeningeal anastomoses, allows brain tissue surrounding the infarct core to maintain cell and tissue vitality for a period of time. If the blood supply can be restored timely, the damage to this potentially salvageable brain region, termed the ischemic penumbra, is reversible to some extent [2,8,9]. With the passage of time, long-term insufficiency of glucose and oxygen supply leads to an imbalance in energy supply and demand in the infarcted area. Subsequently, brain cells synergistically or sequentially initiate ischemic cascade reactions, such as excitotoxicity, oxidative stress, and inflammatory responses, which result in the transformation of the ischemic penumbra to the infarct core [2,6,8,9,10].
As the transition from the penumbra to the infarct core is progressive, brain tissues can be rescued with effective cerebral protection therapy. Currently, the main treatment strategy in the early stages of IS is to re-establish the blood supply to the infarcted area. Although reperfusion injury is inevitable in this process, it is essential to attempt to preserve the penumbra tissue and restore normal neurological functions [9,11,12]. To date, intravenous thrombolysis using alteplase, a recombinant tissue plasminogen activator (tPA), is the only pharmacological intervention approved by the FDA for the treatment of acute IS [11,13]. tPA cleaves plasminogen to release plasmin, which subsequently degrades fibrin in the thrombus and promotes blood flow restoration [8,11]. Studies have shown that intravenous tPA administration until 4.5 h from stroke onset is beneficial for patients, and the earlier the better [14,15]. After approximately 4.5 h, the risk of death due to intracranial hemorrhage induced by intravenous tPA administration increases by 5%, far exceeding the possible benefits [15,16,17]. Owing to this narrow time window and the contraindications to thrombolysis, tPA therapy is only suitable for fewer than 10% of stroke patients [18,19]. Therefore, there is an urgent need to devise alternative strategies for rescuing ischemic brain tissue.
In recent years, researchers have intensively studied the pathophysiological characteristics of the ischemic cascade and assessed pharmacological interventions for various molecular events [8,10,20,21,22]. Polyphenols, a class of bioactive compounds found widely in nature, have attracted much attention in this regard because of their high biosafety, multiple therapeutic targets, and excellent therapeutic effects [23,24,25,26,27,28,29]. However, polyphenols often have limited clinical applications due to issues with solubility, stability, and blood-brain barrier (BBB) permeability [23]. In this article, we introduce the pathophysiology of the ischemic cascade and review the literature regarding the application of polyphenols in the treatment of IS over the past 5 years. We also summarize some emerging strategies for improving the bioavailability and ability of polyphenols to resist IS, and finally discuss several important issues that should be considered in the context of the future applications of polyphenols for IS treatment.

2. Pathophysiology of IS

Once IS occurs, energy supply in regions with decreased blood supply becomes inadequate to support normal cellular function. This is followed by an ischemic cascade response involving a complex series of downstream cellular and molecular events [30]. Here, we summarize several key steps in the ischemic cascade response with a focus on the mechanisms of acute neuronal injury (Figure 1).

2.1. Excitotoxicity

One of the critical events following cerebral ischemia is that Na+-K+-ATPase activity decreases when intracellular ATP drops below 25% of normal levels. This results in the massive inward flow of sodium, outward flow of potassium, severe disruption of intracellular ion homeostasis, and presynaptic membrane depolarization [18,31]. Subsequently, a large number of neurotransmitters, including glutamate, are released into the synaptic gap [32]. Under conditions of energy exhaustion and ion homeostasis disruption, glutamate reuptake by the excitatory amino acid transporter protein (EAAT) on neurons and astrocytes is inhibited, leading to glutamate over-accumulation in synaptic gaps [33,34].
The N-methyl-D-aspartate receptor (NMDAR) is the main ionotropic glutamate receptor that enables rapid ion influx in response to glutamate stimulation [35]. In the resting state, the channel pores of the NMDAR are blocked by extracellular magnesium, preventing the influx of other ions. In the presence of excess glutamate, magnesium dissociates from NMDARs; consequently, a large volume of calcium flows rapidly into the cell through NMDARs, triggering the activation of several calcium-dependent pathways, such as calpain activation, mitochondrial damage, and free radical production, ultimately leading to the initiation of neuronal death processes such as apoptosis, necrosis, and autophagy [31,36,37]. This sequence of events is described as excitotoxicity and is considered to be the main mechanism that drives neuronal death during the hyperacute phase of IS [31,38].

2.2. Oxidative and Nitrosative Stress

Brain tissue is highly susceptible to oxidative damage owing to high oxygen consumption and iron and unsaturated lipid content and relatively low endogenous antioxidant capacity [22,39]. It is well known that oxidative and nitrosative stress is one of the downstream consequences of excitotoxicity. Intracellular calcium overload, especially in mitochondria, leads to the activation of a series of calcium-dependent protein kinases. These kinases affect the activity of enzymes such as NADPH oxidase 2 (NOX2) and cytochrome c oxidase (COX), which ultimately leads to the inhibition of the mitochondrial oxidative respiratory chain and excessive production of reactive oxygen species (ROS) and reactive nitrogen species (RNS) [40,41,42]. Meanwhile, the mitochondrial permeability transition pore (mPTP), a high-conductance voltage- and Ca2+-dependent channel, is activated under conditions of high calcium and oxidative stress and remains open for a prolonged period of time. The opening of mPTP results in loss of mitochondrial inner membrane potential, leading to mitochondrial swelling and rupture. Large amounts of intermembrane proteins and ROS/RNS are released into the cytosol, which ultimately leads to severe oxidative damage and apoptosis [40,43,44,45]. The abnormal activation of xanthine oxidase (XOD) and NOX2 outside mitochondria is also a major source of early ROS/RNS production [46]. Excessive ROS/RNS production overwhelms the endogenous antioxidant system and results in nucleic acid disruption, protein nitration, and oxidation, lipid peroxidation, and activation of multiple pro-inflammatory, pro-apoptotic, and necrotic signaling pathways [22,47]. Excess free radicals also affect the BBB, disrupting the tight junctions between cerebrovascular endothelial cells and leading to increased BBB permeability, causing further damage to brain tissue [48].

2.3. Inflammatory Response

Apart from excitotoxicity and oxidative stress, IS also triggers a complex innate immune response. After severe cellular damage occurs in the ischemic region, damaged cells release several DAMPs, including adenosine, heat shock proteins, high mobility group box 1 (HMGB1), interleukin (IL)-33, S100 proteins, and heparan sulfate, into the intercellular space [22,49]. DAMPs are detected by immune cells with corresponding pattern recognition receptors (PRRs), such as the nucleotide-binding oligomerization domain, leucine-rich repeat-containing receptors (NLRs), and toll-like receptors (TLRs), which mediate the activation of intracellular pro-inflammatory signaling pathways [50,51].
Microglia, the resident central nervous system immune cells, are among the first cells to respond to these danger signals [50,52]. Within minutes of an injury, microglia are rapidly activated, undergo morphological changes, and secrete various cytokines [53,54]. Based on their activation pathways, microglia can be categorized into the M1 and M2 phenotypes [55,56]. M1 microglia exist in a pro-inflammatory state in the brain and secrete pro-inflammatory cytokines and chemokines to recruit a variety of peripheral immune cells, including neutrophils, monocytes, and lymphocytes, which eventually results in the coordinated infiltration of immune cells into the brain parenchyma and aggravation of brain damage [22,57]. M2 microglia are anti-inflammatory and release anti-inflammatory cytokines and neurotrophic factors that contribute to brain injury repair [58,59]. Thus, modulating the conversion of microglia from the M1 to the M2 phenotype during the acute phase may be a key approach for the treatment of IS.
Cerebrovascular endothelial cells are also activated rapidly after IS and upregulate the expression of a range of adhesion and procoagulant factors [60]. Neutrophils recruited from the periphery can bind to leukocyte adhesion receptors such as P-selectin, E-selectin, and intercellular adhesion molecule-1 (ICAM-1) on activated cerebrovascular endothelial cells and attach to the endothelium to block capillaries and create further blockages, leading to the “no reflow” phenomenon in the region [61,62,63]. Neutrophils can also participate in thrombosis, form neutrophil extracellular traps, and release matrix metalloproteinases (MMPs) to promote vascular inflammation and BBB disruption, exacerbating the vascular injury and impeding revascularization [53,64,65,66].

2.4. Apoptosis

Among the various cell death pathways initiated after IS, apoptosis within the penumbra has been investigated in depth in an effort to understand how to rescue damaged neurons [67]. Apoptosis can be triggered in two ways, via the mitochondria-mediated pathway, referred to as the intrinsic apoptosis pathway, or the receptor-mediated pathway, referred to as the extrinsic apoptosis pathway [68,69]. Depending on the underlying mechanism of cell death, it can also be classified as caspase-dependent or caspase-independent [68].
The B-cell lymphoma-2 (Bcl-2) protein family is a major regulator of mitochondrial outer membrane permeability and plays a key role in regulating the endogenous apoptosis pathway [70,71]. BH3-interacting domain death agonist (BID), a pro-apoptotic member of the Bcl-2 family, is cleaved into its truncated form (tBID) by calpain [68,72]. tBID interacts with Bcl-2-associated X protein (Bax) and Bcl-2 homologous antagonist/killer (Bak) on the mitochondrial membrane and promotes the formation of mPTPs, leading to the release of several pro-apoptotic factors such as cytochrome c (Cyt c) and apoptosis-inducing factor (AIF) into the plasma [68,73,74]. Cyt c binds to apoptotic-protease-activating factor-1 (Apaf-1) and deoxyATP in the cytoplasm to form apoptotic vesicles and subsequently activate pro-caspase-9, which further cleaves pro-caspase-3, and ultimately caspase-3 degrades nuclear DNA to drive apoptosis [45,72,75,76]. In contrast to Cyt c, AIF is released from the mitochondria and translocated to the nucleus within minutes, mediating significant DNA breakage in a caspase-independent pathway of cell death [77,78,79].
The extrinsic apoptotic pathway is primarily triggered by the binding of ligands to death receptors on the cell surface [69,80]. Some of the most prominent ligand/receptor combinations include tumor necrosis factor-α (TNF-α)/TNF-receptor 1, TNF-related apoptosis-inducing ligand (TRAIL)/TRAIL-R, and first apoptosis signal ligand (FasL)/FAS(CD95) [73,81]. Upon specific binding of these ligands to their receptor, tumor necrosis factor receptor type-1-associated death domain (TRADD) and Fas-associated death domain (FADD) are recruited and initiate downstream responses [68,82]. The N-terminal region of FADD in the FasL complex contains a death effector domain, and pro-caspase-8 is recruited through homotopic domain interaction to form a complex (FasL-Fas receptor-FADD and pro-caspase-8) called the death-induced signaling complex [68,73]. This complex promotes caspase-8 activation and translocation into the cytoplasm, followed by pro-caspase-3 cleavage to activate caspase-3 by direct or mitochondria-dependent mechanisms, which eventually contribute to apoptosis [73,83].

2.5. Autophagy

Autophagy is generally activated in response to nutrient deficiency or metabolic stress. It also maintains cellular homeostasis by removing damaged organelles and extra proteins [84]. Autophagy is a highly regulated process, with mechanistic targets of rapamycin complex 1 (mTORC1) and AMP-activated protein kinase (AMPK) being the two important targets for its initiation [85,86,87].
Under nutrient-sufficient conditions, mTORC1 directly binds to and phosphorylates two subunits of the Unc-51-like kinase 1 (ULK1) complex, autophagy-related gene 13 (ATG13) and ULK1, which keeps the ULK1 complex inactive and prevents the initiation of autophagy [88]. However, once IS occurs, cerebral cells suffer from nutritional deficiency, ULK1 undergoes autophosphorylation, and mTORC1 dissociates from the ULK1 complex to relieve its suppression [88,89]. At the same time, as the intracellular AMP/ATP ratio increases after ischemia, intracellular AMPK is activated, and the tuberous sclerosis 2 complex is further activated to inhibit mTORC1 activity indirectly [90,91]. Furthermore, AMPK can also directly induce phosphorylation of the ULK1 complex, thus triggering autophagy [84,92].
When the ULK1 kinase complex is activated, it continues to phosphorylate the downstream class III phosphoinositide 3-kinase (PI3K) complex, which converts phosphatidylinositol into phosphatidylinositol-3 phosphate to promote membrane nucleation and phagophore formation [87,93,94]. Upon initiation of the autophagic cascade, pro-LC3 in the cytoplasm is cleaved by Atg4 to form LC3-I, and then, under the action of Atg7 and Atg3, it conjugates with phosphatidylethanolamine to form LC3-II [87,95]. LC3-II, together with the Atg5-Atg12-Atg16L1 complex, is involved in membrane expansion and membrane fusion of phagophores to promote the maturation of autophagosomes [87,96,97]. Finally, mature autophagosomes fuse with lysosomes and recycle nutrients to complete the autophagic process [97].
Under normal conditions, autophagy maintains intracellular homeostasis and facilitates cell survival through the removal and/or recycling of harmful cell components. However, if stimulation continues to induce excessive autophagy beyond the cell’s adaptive capacity, even essential cellular components may be destroyed, leading to cell death [68,98]. Therefore, autophagy appears to be a double-edged sword in the context of cellular self-protection, and further investigation is necessary to obtain the information necessary for modulating adaptive cellular autophagy.
Overall, the ischemic cascade response is a dynamic and complex process involving multiple different cell types and response pathways, which in turn interact with and promote each other, further exacerbating the injury (Figure 2). Further studies are still needed to clarify the interactions between different cascade reactions and to provide a guiding direction for future treatment.

3. Application of Polyphenols in the Treatment of IS

Polyphenol is the general term for an aromatic compound containing one or more phenolic hydroxyl structures. They are widely found in plants in nature [99,100,101]—to date, over 8000 polyphenol structures have been identified [23]. Polyphenols can be classified into five categories according to their chemical structures: flavonoids, phenolic acids, stilbenes, lignans, and curcumins [23]. Here, we summarize the research progress related to these five categories of polyphenols in the context of therapeutic applications for IS over the last 5 years.

3.1. Flavonoids

Generally, flavonoids contain two benzene rings and an epoxy heterocyclic ring as their typical chemical backbone [102]. Flavonoids are commonly found in fruits, grains, vegetables, and flowers and are the most abundant polyphenols in nature [103]. Based on the different structures connecting the two benzene rings, flavonoids can be divided into six categories: flavonols, isoflavones, flavones, flavanols, flavanones, and anthocyanidins [104,105]. The therapeutic effects of various flavonoids on IS are summarized in Table 1 and their mechanisms of action are discussed in detail in later sections.

3.1.1. Flavonols

Quercetin (QE), a polyphenol found widely in nature, has been shown to have promising neuroprotective properties against various neurodegenerative diseases [167]. The neuroprotective effects of QE mainly manifest as antioxidant, anti-inflammatory, anti-excitatory, anti-calcium overload, and anti-apoptotic effects [168]. Studies have reported that QE can increase the activity of the antioxidant enzymes superoxide dismutase 1 (SOD1), SOD2, catalase (CAT), and glutathione peroxidase (GPx) in ischemic brain tissue and enhance the antioxidant effects mediated by the Sirt1/Nrf2/HO-1 pathway [107,108]. In addition, there is evidence that QE pre-treatment can reduce the downregulation of two calmodulins, parvalbumin and hippocalcin, after IS, which reduces glutamate-induced toxicity and helps to maintain intraneuronal calcium homeostasis [110,112]. QE pre-treatment can also inhibit neuronal apoptosis by regulating ERK/Akt pathway phosphorylation [111]. Under conditions of ischemia and hypoxia, microglia in the brain are activated and secrete various pro-inflammatory factors such as TNF-α, IL-1β, and IL-6, which eventually aggravate brain tissue damage [23]. QE reduced pro-inflammatory cytokine production in oxygen-glucose deprivation/reoxygenation (OGD/R)-treated BV2 cells and inhibited TLR4/MyD88/NF-κB pathway signaling to protect damaged brain tissue [110]. However, despite these beneficial effects, the use of QE in IS is still limited owing to its low oral bioavailability and weak BBB permeability [169].
Isoquercetin (Q3G) is a monoglucoside derivative of QE with better bioavailability [170]. Oxidative stress and neuronal apoptosis after ischemia/reperfusion (IR) can be mitigated by Q3G through inhibition of the NOX4/ROS/NF-κB pathway [118]. It can also inhibit TLR4/NF-κB pathway activation and the phosphorylation of JNK1/2, ERK1/2, and p38 MAPK to reduce the inflammatory response and apoptosis [117].
Decreased estrogen levels are considered a key factor affecting the risk of postmenopausal stroke [171,172]. Rutin, a disaccharide rutinose derivative of QE, has also been reported to reduce post-cerebral ischemia by activating estrogen receptor-mediated brain-derived neurotrophic factor (BDNF)-TrκB and nerve growth factor (NGF)-TrkA signaling in ovariectomized rats [119].
Kaempferol (KEM), a flavonol, has been demonstrated to have anti-inflammatory and antioxidant effects that are effective for the treatment of many diseases [173]. Evidence shows that KEM treatment suppresses the production of chemokines such as MCP-1 and ICAM-1 and pro-inflammatory factors such as inducible nitric oxide synthase (iNOS) and COX-2 after stroke, reducing microglial overactivation [120]. In addition, KEM can reduce intracellular mitochondrial damage and help maintain mitochondrial function after ischemia and hypoxia by affecting various pathways, such as upregulating Sirt1 expression and inhibiting p66shc acetylation or inhibiting mitochondrial Drp1 recruitment and HK-II detachment [122,123]. Interestingly, KEM can also improve OGD/R-induced neuronal ferroptosis by activating the Nrf2/SLC7A11/GPx4 pathway, which could emerge as a new strategy for IS treatment [124].
Icariin (ICA), the main active ingredient in Epimedium genus plants, has a wide range of pharmacological activities [174,175,176]. Studies have shown that ICA inhibits NF-κB activation after IS by upregulating the PPARs/Nrf2 pathway and downregulating the JAK2/STAT3 pathway, thereby enhancing mild hypothermia-induced neuroprotection [127]. ICA can also reduce endoplasmic reticulum (ER) stress and inflammatory responses by modulating IRE1α-XBP1 pathway activation [126]. Furthermore, Liu et al. found that ICA activates the PI3K/ERK1/2 pathway and increases VEGF and BDNF secretion by mesenchymal stem cells, thereby promoting angiogenesis and neurogenesis after stroke [125].

3.1.2. Isoflavones

Puerarin (PUE) is an isoflavone extracted from Pueraria genus plants, and its pharmacological activity has been extensively investigated [177]. Pre-treatment with PUE before IS can increase p-Akt1/p-GSK-3β/MCL-1 cascade activity to improve the survival of hippocampal neurons and alleviate motor and cognitive deficits [130]. In addition, PUE could inhibit ischemia-induced neuronal autophagy by activating the AMPK/mTOR/ULK1 pathway, downregulating the LC3-II/LC3-I ratio, and increasing p62 expression in the ischemic hippocampus, significantly reducing ischemic brain edema, infarct volume, and neurological deficits [131].
Genistein (GE) and daidzein (DAZ) are the two main isoflavones in soybeans, and are classified as phytoestrogens, because of the similarities of their structures to estrogens [178,179]. Studies have shown that GE and DAZ play unique roles in the treatment of postmenopausal cerebral ischemia. GE can reduce the production of excess ROS in ischemic regions by increasing Nrf2 and NQO1 expression in temporary-middle cerebral artery occlusion (T-MCAO) models of ovariectomized rats [133]. It can also modulate the PI3K/Akt/mTOR pathway to reduce apoptosis, which could provide a new strategy for the treatment of stroke in postmenopausal women [134]. DAZ can also inhibit neuronal apoptosis and promote BDNF and cAMP-response element binding protein (CREB) expression by regulating the PI3K/Akt/mTOR pathway, effectively stimulating neuronal regeneration after IS [135]. Thus, these two polyphenols have great potential for the treatment of postmenopausal stroke.

3.1.3. Flavones

Scutellaria baicalensis is a traditional Chinese medicinal plant, which widely used to treat patients with inflammatory cardiovascular diseases such as hypertension and atherosclerosis [180]. Baicalein (BAI) and baicalin (BG), the main components of the S. baicalensis extract, have excellent pharmacological activities [181]. It has been demonstrated that BAI could play anti-inflammatory and antioxidant roles by regulating the AMPK/Nrf2 signaling pathway and suppressing the expression of inflammatory mediators such as LOX-1, COX-2, PGE2, and NF-κB [137]. Notably, BAI treatment also inhibits STAT1 phosphorylation, promoting the conversion of ischemic penumbra microglia to the M2 type, resulting in significant reductions in infarct volume [138]. Moreover, BAI restrains the nuclear transport of AIF and macrophage migration inhibitory factor (MIF) by downregulating the activity of calpain-1 and reducing the expression of poly (ADP-ribose) polymerase 1 (PARP-1), leading to the inhibition of ROS production and apoptosis [139,140].
BG, a glycoside derivative of BAI, displays therapeutic activities similar to those of BAI. Studies have shown that BG can inactivate succinate dehydrogenase in astrocytes to suppress mitochondria-derived ROS production, protect glutamine synthase from 20S proteasomal degradation, and enhance extracellular glutamate uptake and resistance to excitotoxicity [141].
Scutellarin (SCU) is a flavonoid extracted from the traditional Chinese herb Erigeron breviscapus. SCU treatment can reduce infarct volume and brain water content [182]. Moreover, recent studies have shown that pre-treatment with scutellarin suppresses the phosphorylation of p38 MAPK and JNK1/2, attenuates microglia-mediated inflammatory responses, and effectively reduces ischemia-induced brain injury [142]. Interestingly, SCU can also bind specifically to NOX2 and effectively inhibit its activation within astrocytes after ischemic brain injury [144].
Luteolin (LTL) is a flavonoid polyphenol available from a variety of dietary sources [183]. LTL could upregulate Sirt3 expression and activate the AMPK/mTOR pathway in the brains of T-MCAO rats, effectively reducing the number of activated glial cells, improving neurological function, and reducing brain infarct volume [147]. PPARγ belongs to a receptor family of ligand-activated nuclear transcription factors that regulate the transcription and expression of several genes and play a crucial role in neuroprotection [184,185]. LTL treatment significantly increased the expression of PPARγ and modulated the downstream Nrf2/NF-κB pathway in the brain of T-MCAO rats, thus reducing I/R injury in the brain [146].

3.1.4. Flavanols

Three flavanols have been found to contribute to neuroprotection in IS. Epigallocatechin gallate (EGCG) and epicatechin gallate (ECG) are the most abundant polyphenols in tea and have been reported to have therapeutic effects on a variety of neurological disorders [186,187]. In rat p-MCAO models, EGCG promoted thioredoxin expression and increased its interaction with ASK-1 and demonstrated neuroprotective effects against glutamate toxicity and ischemic brain injury [156]. PARP is a family of signature proteins that induces apoptosis by promoting cellular AIF release [188]. Pre-treatment with EGCG was shown to effectively reduce PARP expression in ischemic cerebral tissue and to regulate the apoptotic cascade to reduce cell death after focal cerebral ischemia [155]. EGCG can also inhibit the inflammatory response and apoptosis after injury by moderating the PI3K/Akt pathway and upregulating endothelial nitric oxide synthase (eNOS), increasing the proliferation and differentiation of neural progenitor cells, and promoting neurogenesis [153,154]. ECG can significantly downregulate ROS levels in cerebrovascular endothelial cells after OGD/R, decrease apoptotic and autophagic protein expression, and promote VEGF expression and neovascularization, and thus may provide novel avenues for the treatment of IS [157].

3.1.5. Flavanones

Naringenin (NRG) and naringin (NG) are flavanone polyphenols found in citrus fruits and have excellent antioxidant and anti-inflammatory effects [189]. Recent studies have shown that NRG increases Nrf2 expression and promotes its nuclear translocation, reduces oxidative stress, and prevents apoptosis in cortical neurons [159]. NG, a glycoside derivative of NRG, has antioxidant and anti-apoptotic properties similar to those of NRG, and can scavenge ONOO and reduce excessive mitochondrial autophagy mediated by it to improve brain damage in a rat model of T-MCAO [160].

3.1.6. Anthocyanins

Anthocyanins are a group of natural plant pigments commonly found in fruits and vegetables [23,190]. In preclinical studies, cyanidin-3-glucoside (C3G) was found to inhibit TLR4/NF-κB/NLRP3 signaling and block the expression of several related inflammatory factors to reduce the inflammatory response [163]. It could also inhibit the glutamate-mediated apoptosis of HT22 neurons by suppressing ER stress through decreasing oxidative stress and increasing the expression of antioxidant proteins such as Nrf2, SOD, CAT, GPx, and GST [164].
Petunidin-3-O-rutinoside (p-coumaroyl)-5-O-glucoside is an anthocyanin purified from dried Lycium ruthenicum Murr. fruit. Studies have shown that it can significantly minimize infarct volume and cerebral edema, inhibit NF-κB/NLRP3 pathway activation while suppressing MMP9 activation, and promote the protection of the neurovascular unit [165]. Moreover, it can lower SQSTM1 expression, increases the LC3B II/LC3B I ratio, enhance autophagy, and restrict OGD-induced neural injury [166].

3.2. Phenolic Acids

Phenolic acids have a molecular structure consisting of a carboxylic group and benzene rings, in addition to one or more methoxy and/or hydroxyl groups. Based on their chemical structure, they can be divided into benzoic acid derivatives and cinnamic acid derivatives [191]. Here, we summarize the therapeutic effects and associated mechanisms of action of these two types of phenolic acids in IS (Table 2).

3.2.1. Cinnamic Acid Derivatives

Ferulic acid (FA), a component of Angelica sinensis and Ligusticum chuanxiong, has therapeutic effects against a variety of neurodegenerative diseases due to its anti-inflammatory and antioxidant properties [213,214]. Recent studies suggest that the administration of FA immediately after an ischemic attack is effective in reducing cerebral infarction and improving neurological function, which may be attributed to its upregulation of the Akt/mTOR/4E-BP1/Bcl-2 anti-apoptotic pathway [193].
Rosmarinic acid (RA) is a caffeic acid derivative extracted from the rosemary plant [215,216]. Upon systemic administration, RA modulates the PI3K/Akt pathway to promote Nrf2/OH-1 pathway activation and protects against cerebral I/R injury via activation of antioxidant and anti-apoptotic pathways [195].
Chlorogenic acid (CA), the ester of caffeic acid and quinic acid, is considered to be one of the most abundant dietary polyphenols in coffee [217]. CA reduces apoptosis mediated by the miR-23b/TAB3/NF-κB pathway and decreases the release of inflammatory factors, thus acting as an anti-neuroinflammatory and anti-apoptotic agent [198]. Another study demonstrated that CA also regulates the expression of the apoptosis-related proteins caspase-3, caspase-7, and PARP and protects neurons from cerebral ischemia [197]. CA can downregulate intercellular adhesion molecule-1 (ICAM-1) and vascular cell adhesion molecule-1 (VCAM-1) levels and upregulate targets such as erythropoietin (EPO), hypoxia-inducible factor 1α (HIF-1α), and NGF levels in brain tissue, which reduces neuronal death and promotes neuronal regeneration [196].
Salvianolic acids are a class of bioactive compounds extracted from Salvia, traditional Chinese herbs used to treat cardiovascular diseases [218]. Studies have shown that salvianolic acid A (SAA) can inhibit apoptosis and the inflammatory response by modulating the TLR2/4/MyD88 and FOXO3a/BIM pathways, which can have a neuroprotective effect [202,205]. SAA treatment can also reduce VEGFA-Src-VAV2-Rac1-PAK pathway activation and depress MMP expression in ischemic brain tissue, preventing the degradation of the tight junction proteins ZO-1, claudin-5, and occludin, which protects the BBB from damage and reduces neuronal death [204]. Dickkopf-1 (DKK1), one of the major members of the DKK family, blocks the Wnt/β-catenin signaling pathway by binding to the Wnt receptor complex low-density lipoprotein receptor-related protein 5/6, which in turn mediates downstream target gene transcription to protect against acute brain injury [219]. SAA has also been shown to modulate DKK1 to activate the Wnt/β-catenin signaling pathway to reduce I/R-induced brain injury by upregulating miR-449a levels in neuronal cells and a rat model of I/R [203]. Surprisingly, long-term administration of SAA also activated the Wnt3a/GSK3β/β-catenin pathway after IS to promote endogenous neurogenesis and inhibited apoptotic signaling to accelerate neurological recovery after injury [206].
Salvianolic acid B (SAB) is the most abundant bioactive hydrophilic compound in Salvia and has been designated by the Chinese Pharmacopoeia as a marker component of this genus [220]. SAB has been described to have potent antiplatelet activity and great therapeutic potential for the treatment of thrombotic disorders [221]. In a T-MCAO rat model, SAB decreased plasma levels of P-selectin and the CD40/CD40 ligand ratio, inhibited platelet activation and inflammatory cell recruitment, and suppressed NF-κB p65 phosphorylation and production of pro-inflammatory mediators [207]. SAB has also been reported to increase glycogen phosphorylase activity, promote astrocyte glycogenolysis, increases antioxidant levels, including those of NADPH and GSH, decrease intracellular ROS levels, and increase astrocyte and neuronal survival, leading to reduced infarct size and enhanced neurological recovery [208].

3.2.2. Benzoic Acid Derivatives

Protocatechuic acid (PA) and gallic acid (GA), natural components of green tea, are benzoic acid derivatives with strong free-radical scavenging effects [222,223]. In a rat model of focal cerebral ischemia, early administration of PA increased CREB expression in the rat brain and prevented cerebral I/R injury [209]. GA was found to induce microglial polarization to the M2 type after mouse brain I/R injury, reduce inflammatory factor secretion, and regulate tight junction-related protein expression to protect the BBB and mitigate cerebral injury [210].

3.3. Lignans

Lignans are natural products formed by two or three polymerizations of different types of phenylpropanoid groups. Their main dietary sources are oilseeds, cereals, and legumes, and they are known to have various types of pharmacological properties, such as anti-inflammatory and antioxidant activities, which are expected to be useful in IS [224,225]. Here, we summarize the research findings regarding the use of lignans for the treatment of IS over the past 5 years (Table 3).
Magnolol, a lignan-like compound isolated from the Chinese herb Magnolia officinalis, is a potent antioxidant that can effectively reduce the production of oxidative stress markers and inflammatory factors that regulate brain injury after IS. EphA2 receptors are a class of transmembrane receptor tyrosine kinases that facilitate the maintenance of the BBB tight junctional architecture [234]. In the early stages of IS, magnolol has been shown to inhibit EphA2 phosphorylation to attenuate BBB damage and reduce infarct size [226].
Schisandrins are lignans isolated from Schisandra chinensis fruit. It has been shown that pre-treatment with schisandrin A can promote neural progenitor cell regeneration, migration, and differentiation after cerebral ischemia by increasing cell division control protein 42 (Cdc42) levels, which contributes to neural regeneration after IS [230].

3.4. Stilbenes

Stilbenes are a class of compounds with two aromatic ring structures connected by an ethylene bridge, with various subclasses based on different substituents in the aromatic ring [235]. Resveratrol, a representative astragal, has attracted extensive attention from researchers (Table 4).
It is a natural stilbene found abundantly in foods such as peanuts, grapes, and red wine and has been reported to have anti-inflammatory, anti-apoptotic, and autophagy-modulating effects through various pathways, including the JAK/ERK/STAT and PI3K/Akt/mTOR pathways [237,238,239]. CD147 is a transmembrane glycoprotein that has recently been proven to be an important immune response and inflammation mediator, as it induces MMP9 expression in many cell types [253,254,255]. Timely administration of RES after injury can suppress this process and thus inhibit microglial activation [236]. In addition, RES inhibits mitochondrial respiration and sequentially activates AMPK and Sirt1, regulates acetyl coenzyme A levels to achieve mitochondrial and nuclear adaptation, and improves glycolysis efficiency, which ultimately increases basal ATP levels and promotes long-term ischemic tolerance [256,257]. A recent study on the gut-brain axis showed that RES also modulates immune cell homeostasis mediated by intestinal flora—specifically, Th1/Th2 and Treg/Th17 balance in the lamina propria of the small intestine—to suppress inflammation [240]. Interestingly, RES also activates sonic hedgehog signaling to promote neurogenesis and functional recovery after IS [246].

3.5. Curcumin

Curcumin (CUR) is a natural polyphenol with unsaturated aliphatic and aromatic moieties in its main chain. It is extracted from Curcuma longa root and has been widely used in IS treatment studies owing to its anti-inflammatory, antioxidant, and neuroprotective effects (Table 5).
CUR can inhibit post-stroke apoptosis and the inflammatory response by modulating the TLR4/p38 MAPK and MEK/ERK/CREB pathways [267,268,274]. CUR administration also activates the PI3K/Akt/mTOR pathway and reduces downstream autophagy-related protein expression to decrease autophagic activity and exert neuroprotective effects [267]. Orai1, a calcium-regulated protein, mediates the inward flow of calcium ions induced by oxidative stress [275]. CUR administration inhibits Orai1-induced inward calcium flow through upregulation of protein kinase C-θ (PKC-θ) expression, effectively maintaining BBB integrity and function, and exhibits protective effects in brain I/R injury [262]. Interestingly, CUR also exerted therapeutic effects in a diabetic stroke model by activating glucose transporter protein 1/3 (GLUT1/3) to promote glucose uptake and anti-apoptosis effects [273].

3.6. Approaches to Improve the Bioavailability of Polyphenols in IS

Despite the widespread recognition of the beneficial effects of natural polyphenol components in the prevention and treatment of IS, their bioavailability is limited by low solubility, stability, and BBB permeability [276,277,278]. Accordingly, polyphenol modification and polyphenol encapsulation have emerged as research hotspots as various attempts are being made to overcome these issues.

3.6.1. Polyphenol Modification Strategies to Improve Bioavailability

Researchers have attempted to enhance the anti-stroke efficacy of polyphenolic compounds by modifying their structure to obtain novel active ingredients with high stability, bioactivity, and bioavailability and fewer adverse effects. Mrvová et al. synthesized 3′-O-(3-chloropivaloyl) quercetin, a pivaloyl ester derivative of quercetin with increased lipophilicity and better BBB permeation efficiency, inflammation inhibition, and cell cycle regulation than quercetin [279]. Similarly, Skandik et al. prepared semisynthetic 4-O-(2-chloro-1,4-naphthoquinone-3-yloxy) quercetin, which has enhanced electrophilic and lipophilic properties compared to quercetin, and is effective in reducing microglial activation after stroke by modulating Nrf2 expression at low concentrations [280]. Zhang et al. synthesized Cur20, a CUR derivative with high hydrolytic stability and lower hydrolysis efficiency than CUR and can significantly promote angiogenesis by activating the HIF-1α/VEGF/TFEB pathway to reduce brain injury after ischemia [281].

3.6.2. Polyphenol Delivery Strategies for IS Therapy

In recent years, nanotechnology-based drug delivery systems have received considerable attention for their potential to improve drug stability and solubility and increase circulation times in vivo [30,282]. Nanoparticles made from poly(lactic-co-glycolic acid) (PLGA) are recognized as effective drug carriers because of their long circulation time, high stability and carrier capacity, and diverse delivery routes. Ghosh et al. loaded quercetin in PLGA nanoparticles and found that nano-quercetin showed better anti-inflammatory and antioxidant effects than free quercetin in both young and aged rats and significantly reduced neuronal damage [283]. Subsequently, they incorporated mitochondria-targeting triphenylphosphine into the nanomaterials, which then exhibited even better mitochondrial protection and antioxidant effects in the IS model [284].
Mesoporous silica nanoparticles (MSNPs) are highly attractive for drug delivery applications owing to their excellent biocompatibility, drug-carrying capacity, and surface functionalization properties [285]. Shen et al. developed polylactic acid (PLA)-coated MSNP as a drug carrier for RES and showed that these could bind to the ligand peptide of the low-density lipoprotein receptor to enhance its transcytosis across the BBB, effectively increasing RES concentrations in the brain and reducing oxidative stress [286].
It has been shown that polysorbate-based nanoparticles adsorb ApoE from circulating blood and are then specifically taken up into the brain by ApoE receptors at the BBB. Kakkar et al. prepared a class of Tween80-based solid lipid nanoparticles for CUR loading and showed that their use could increase the concentration of CUR in the brain by up to 30 times the previous concentration and effectively improve its bioavailability [287].
Recently, the combination of microbubbles (MBs) and transcranial low-intensity focused ultrasound (LIFU) has been considered a prospective strategy for the delivery of drugs across the BBB [288,289]. Yan et al. prepared lipid-PLGA nanobubbles loaded with CUR that could accumulate at lesions in large quantities after the BBB was briefly disrupted using focused ultrasound and release CUR to achieve its pharmacological effects [290].

4. Conclusions and Future Prospects

IS is a complex and constantly developing pathological process that involves various pathways; thus, it can be difficult to treat through a single approach. Polyphenols differ from traditional medications in that they can exert anti-stroke effects by focusing on multiple targets and pathways, which is one of their advantages. Although the prevention and therapeutic efficacy of natural polyphenol components in IS have been widely recognized, there are still problems that limit their clinical translation.
First, the mechanism of most natural polyphenols in vivo remains unclear, and further studies are required to confirm their application potential. Second, the dose range, time window, and duration of administration of polyphenol components with anti-IS effects need to be further investigated with the goal of maximizing their therapeutic effects and minimizing possible side effects. Third, current preclinical models have tended to focus on young, healthy rodents that have not been exposed to other medications, but in clinical practice, stroke patients are mostly elderly individuals with underlying conditions such as hypertension, hyperglycemia, and hyperlipidemia, and are often being administered other medications [291,292]. In addition, although the T-MCAO model is most commonly used in preclinical studies, only 10% of clinical patients can be treated with reperfusion, which may affect the neuroprotective effects of polyphenols to varying degrees. Therefore, it is necessary to further investigate the pharmacokinetic, pharmacodynamic, and toxicological properties and interactions of polyphenols with other drugs in IS using well-simulated clinical disease models. Fourth, in clinical trials, scales such as the modified Rankin scale, the Barthel Index, and the National Institute of Health Stroke Scale are used for long-term (usually 90 days), multifaceted (e.g., related to sensory, motor, and speech-related effects) assessment of neurological recovery to evaluate treatment effects [293]. However, in preclinical studies, the assessment measures, usually infarct size comparisons, are short-term (usually 24 h), and the evaluation of neurological recovery is limited. This may lead to an exaggeration of the therapeutic efficacy of polyphenols. Therefore, a more clinically appropriate evaluation system should be developed to better evaluate the therapeutic effects. Finally, although several strategies have been adopted to improve the bioavailability of polyphenols, there is a need to explore more delivery strategies based on the pathophysiological alterations in brain tissue after IS in order to achieve targeted delivery and ischemic tissue responsive release of polyphenols and achieve better therapeutic effects.
This article reviewed the pathogenesis of IS and the progress of research on the application of polyphenolic components (including flavonoids, phenolic acids, astragalus, lignans, and curcumin) in the treatment of IS. Despite their considerable therapeutic potential, the poor stability and low bioavailability of polyphenols hinder their application in vivo. Modification of polyphenols or application of nanoformulations to assist polyphenol therapy offers great advantages in terms of conferring better bioavailability to polyphenol components and achieving better therapeutic effects. This review provides a reference for exploring the applications of polyphenols in IS treatment.

5. Chemical Compounds Studied in This Article

Quercetin (PubChem CID: 5280343); Isoquercetin (PubChem CID: 5280804); Rutin (PubChem CID: 5280805); Kaempferol (PubChem CID: 5280863); Icariin (PubChem CID: 5318997); Myricetin (PubChem CID: 5281672); Puerarin (PubChem CID: 5281807); Genistein (PubChem CID: 5280961); Daidzein (PubChem CID: 5281708); Baicalein (PubChem CID: 5281605); Baicalin (PubChem CID: 64982); Scutellarin (PubChem CID: 185617); Luteolin (PubChem CID: 5280445); Chrysin (PubChem CID: 5281607); Apigenin (PubChem CID: 5280443); Epigallocatechin Gallate (PubChem CID: 65064); Epicatechin gallate (PubChem CID: 107905); Procyanidin (PubChem CID: 107876); NARINGENIN (PubChem CID: 932); Naringin (PubChem CID: 442428); HESPERETIN (PubChem CID: 72281); Cyanidin-3-glucoside (PubChem CID: 441667); Ferulic acid (PubChem CID: 445858); rosmarinic acid (PubChem CID: 5281792); Chlorogenic acid (PubChem CID: 1794427); Salvianolic acid A (PubChem CID: 5281793); Salvianolic Acid B (PubChem CID: 11629084); protocatechuic acid (PubChem CID: 72); Gallic acid (PubChem CID: 370); vanillic acid (PubChem CID: 8468); Rhein (PubChem CID: 10168); Magnolol (PubChem CID: 72300); Schisandrin A (PubChem CID: 155256); Schisandrin B (PubChem CID: 108130); Sesamol (PubChem CID: 68289); Arctigenin (PubChem CID: 64981); Resveratrol (PubChem CID: 445154); Pterostilbene (PubChem CID: 5281727); Piceatannol (PubChem CID: 667639); Polydatin (PubChem CID: 5281718); Curcumin (PubChem CID: 969516).

Author Contributions

S.L.: writing—original draft, data curation, visualization. F.L.: writing—original draft, investigation, data curation. J.W.: investigation. X.P.: data curation. L.S.: writing—review & editing, supervision. W.W.: conceptualization, supervision, project administration. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Jilin Provincial Finance Program (JLSWSRCZX2020-001).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AIFapoptosis-inducing factor
BBBblood-brain barrier;
BDNFbrain-derived neurotrophic factor;
CAchlorogenic acid;
CREBCAMP-response element binding;
DAMPdamage-associated molecular pattern;
EAATexcitatory amino acid transporter;
ERendoplasmic reticulum;
FAferulic acid;
FADDFas-associated death domain;
FasLfirst apoptosis signal ligand;
GAgallic acid; IS: ischemic stroke;
LIFUlow-intensity focused ultrasound;
MIFmigration inhibitory factor;
NGFnerve growth factor;
NMDARN-methyl-D-aspartate receptor;
PAprotocatechuic acid;
PRRpattern recognition receptor;
RArosmarinic acid;
ROSreactive oxygen species;
SAAsalvianolic acid;
SDSprague Dawley;
TLRtoll-like receptor

References

  1. Feigin, V.L.; Stark, B.A.; Johnson, C.O.; Roth, G.A.; Bisignano, C.; Abady, G.G.; Abbasifard, M.; Abbasi-Kangevari, M.; Abd-Allah, F.; Abedi, V.; et al. Global, regional, and national burden of stroke and its risk factors, 1990–2019: A systematic analysis for the Global Burden of Disease Study 2019. Lancet Neurol. 2021, 20, 795–820. [Google Scholar] [CrossRef]
  2. Campbell, B.C.V.; Khatri, P. Stroke. Lancet 2020, 396, 129–142. [Google Scholar] [CrossRef]
  3. Feigin, V.L.; Nguyen, G.; Cercy, K.; Johnson, C.O.; Alam, T.; Parmar, P.G.; Abajobir, A.A.; Abate, K.H.; Abd-Allah, F.; The GBD 2016 Lifetime Risk of Stroke Collaborators; et al. Global, Regional, and Country-Specific Lifetime Risks of Stroke, 1990 and 2016. N. Engl. J. Med. 2018, 379, 2429–2437. [Google Scholar] [CrossRef] [PubMed]
  4. Feigin, V.L.; Vos, T.; Nichols, E.; Owolabi, M.O.; Carroll, W.M.; Dichgans, M.; Deuschl, G.; Parmar, P.; Brainin, M.; Murray, C. The global burden of neurological disorders: Translating evidence into policy. Lancet Neurol. 2020, 19, 255–265. [Google Scholar] [CrossRef]
  5. Astrup, J.; Siesjö, B.K.; Symon, L. Thresholds in cerebral ischemia—The ischemic penumbra. Stroke 1981, 12, 723–725. [Google Scholar] [CrossRef] [Green Version]
  6. Doyle, K.; Simon, R.P.; Stenzel-Poore, M.P. Mechanisms of ischemic brain damage. Neuropharmacology 2008, 55, 310–318. [Google Scholar] [CrossRef] [Green Version]
  7. Bentley, P.; Sharma, P. Pharmacological treatment of ischemic stroke. Pharmacol. Ther. 2005, 108, 334–352. [Google Scholar] [CrossRef]
  8. Campbell, B.C.V.; De Silva, D.A.; MacLeod, M.R.; Coutts, S.B.; Schwamm, L.H.; Davis, S.M.; Donnan, G.A. Ischaemic stroke. Nat. Rev. Dis. Primers 2019, 5, 70. [Google Scholar] [CrossRef]
  9. Fisher, M.; Savitz, S.I. Pharmacological brain cytoprotection in acute ischaemic stroke—Renewed hope in the reperfusion era. Nat. Rev. Neurol. 2022, 18, 193–202. [Google Scholar] [CrossRef]
  10. Xing, C.; Arai, K.; Lo, E.H.; Hommel, M. Pathophysiologic Cascades in Ischemic Stroke. Int. J. Stroke 2012, 7, 378–385. [Google Scholar] [CrossRef]
  11. Zhang, L.; Zhang, Z.G.; Chopp, M. The neurovascular unit and combination treatment strategies for stroke. Trends Pharmacol. Sci. 2012, 33, 415–422. [Google Scholar] [CrossRef] [Green Version]
  12. Ramos-Cabrer, P.; Campos, F.; Sobrino, T.; Castillo, J. Targeting the Ischemic Penumbra. Stroke 2011, 42, S7–S11. [Google Scholar] [CrossRef] [Green Version]
  13. Hacke, W.; Kaste, M.; Bluhmki, E.; Brozman, M.; Dávalos, A.; Guidetti, D.; Larrue, V.; Lees, K.R.; Medeghri, Z.; Machnig, T.; et al. Thrombolysis with Alteplase 3 to 4.5 Hours after Acute Ischemic Stroke. N. Engl. J. Med. 2008, 359, 1317–1329. [Google Scholar] [CrossRef] [Green Version]
  14. Lees, K.R.; Bluhmki, E.; von Kummer, R.; Brott, T.G.; Toni, D.; Grotta, J.C.; Albers, G.W.; Kaste, M.; Marler, J.R.; Hamilton, S.A.; et al. Time to treatment with intravenous alteplase and outcome in stroke: An updated pooled analysis of ECASS, ATLANTIS, NINDS, and EPITHET trials. Lancet 2010, 375, 1695–1703. [Google Scholar] [CrossRef]
  15. Hacke, W.; Donnan, G.; Fieschi, C.; Kaste, M.; von Kummer, R.; Broderick, J.P.; Brott, T.; Frankel, M.; Grotta, J.C.; Haley, E.C.; et al. Association of outcome with early stroke treatment: Pooled analysis of ATLANTIS, ECASS, and NINDS rt-PA stroke trials. Lancet 2004, 363, 768–774. [Google Scholar] [CrossRef]
  16. Lansberg, M.G.; Schrooten, M.; Bluhmki, E.; Thijs, V.N.; Saver, J. Treatment Time-Specific Number Needed to Treat Estimates for Tissue Plasminogen Activator Therapy in Acute Stroke Based on Shifts over the Entire Range of the Modified Rankin Scale. Stroke 2009, 40, 2079–2084. [Google Scholar] [CrossRef] [Green Version]
  17. Fisher, M.; Saver, J.L. Future directions of acute ischaemic stroke therapy. Lancet Neurol. 2015, 14, 758–767. [Google Scholar] [CrossRef]
  18. Kaplan-Arabaci, O.; Acari, A.; Ciftci, P.; Gozuacik, D. Glutamate Scavenging as a Neuroreparative Strategy in Ischemic Stroke. Front. Pharmacol. 2022, 13, 1018. [Google Scholar] [CrossRef]
  19. de Sousa, D.A.; von Martial, R.; Abilleira, S.; Gattringer, T.; Kobayashi, A.; Gallofré, M.; Fazekas, F.; Szikora, I.; Feigin, V.; Caso, V.; et al. Access to and delivery of acute ischaemic stroke treatments: A survey of national scientific societies and stroke experts in 44 European countries. Eur. Stroke J. 2018, 4, 13–28. [Google Scholar] [CrossRef] [Green Version]
  20. O’Collins, V.E.; Macleod, M.R.; Donnan, G.; Horky, L.L.; Van Der Worp, B.H.; Howells, D.W. 1,026 Experimental treatments in acute stroke. Ann. Neurol. 2006, 59, 467–477. [Google Scholar] [CrossRef]
  21. Moskowitz, M.A.; Lo, E.H.; Iadecola, C. The Science of Stroke: Mechanisms in Search of Treatments. Neuron 2010, 67, 181–198. [Google Scholar] [CrossRef] [Green Version]
  22. Chamorro, Á.; 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]
  23. Li, R.; Zhou, Y.; Zhang, S.; Li, J.; Zheng, Y.; Fan, X. The natural (poly)phenols as modulators of microglia polarization via TLR4/NF-κB pathway exert anti-inflammatory activity in ischemic stroke. Eur. J. Pharmacol. 2022, 914, 174660. [Google Scholar] [CrossRef]
  24. Annunziata, G.; Sureda, A.; Orhan, I.E.; Battino, M.; Arnone, A.; Jiménez-García, M.; Capó, X.; Cabot, J.; Sanadgol, N.; Giampieri, F.; et al. The neuroprotective effects of polyphenols, their role in innate immunity and the interplay with the microbiota. Neurosci. Biobehav. Rev. 2021, 128, 437–453. [Google Scholar] [CrossRef]
  25. 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. [Google Scholar] [CrossRef]
  26. Pacifici, F.; Rovella, V.; Pastore, D.; Bellia, A.; Abete, P.; Donadel, G.; Santini, S.; Beck, H.; Ricordi, C.; Daniele, N.; et al. Polyphenols and Ischemic Stroke: Insight into One of the Best Strategies for Prevention and Treatment. Nutrients 2021, 13, 1967. [Google Scholar] [CrossRef]
  27. Anwar, S.; Shamsi, A.; Shahbaaz, M.; Queen, A.; Khan, P.; Hasan, G.M.; Islam, A.; Alajmi, M.F.; Hussain, A.; Ahmad, F.; et al. Rosmarinic Acid Exhibits Anticancer Effects via MARK4 Inhibition. Sci. Rep. 2020, 10, 10300. [Google Scholar] [CrossRef]
  28. Shamsi, A.; Anwar, S.; Shahbaaz, M.; Mohammad, T.; Alajmi, M.F.; Hussain, A.; Hassan, I.; Ahmad, F.; Islam, A. Evaluation of Binding of Rosmarinic Acid with Human Transferrin and Its Impact on the Protein Structure: Targeting Polyphenolic Acid-Induced Protection of Neurodegenerative Disorders. Oxid. Med. Cell. Longev. 2020, 2020, 1245875. [Google Scholar] [CrossRef]
  29. Shahwan, M.; Alhumaydhi, F.; Ashraf, G.; Hasan, P.M.; Shamsi, A. Role of polyphenols in combating Type 2 Diabetes and insulin resistance. Int. J. Biol. Macromol. 2022, 206, 567–579. [Google Scholar] [CrossRef]
  30. Li, C.; Sun, T.; Jiang, C. Recent advances in nanomedicines for the treatment of ischemic stroke. Acta Pharm. Sin. B 2021, 11, 1767–1788. [Google Scholar] [CrossRef]
  31. Rossi, D.J.; Brady, J.D.; Mohr, C. Astrocyte metabolism and signaling during brain ischemia. Nat. Neurosci. 2007, 10, 1377–1386. [Google Scholar] [CrossRef] [PubMed]
  32. Obrenovitch, T.P.; Urenjak, J.; Richards, D.A.; Ueda, Y.; Curzon, G.; Symon, L. Extracellular Neuroactive Amino Acids in the Rat Striatum during Ischaemia: Comparison between Penumbral Conditions and Ischaemia with Sustained Anoxic Depolarisation. J. Neurochem. 1993, 61, 178–186. [Google Scholar] [CrossRef] [PubMed]
  33. Pál, B. Involvement of extrasynaptic glutamate in physiological and pathophysiological changes of neuronal excitability. Cell. Mol. Life Sci. 2018, 75, 2917–2949. [Google Scholar] [CrossRef] [PubMed]
  34. Bridges, R.J.; Esslinger, C.S. The excitatory amino acid transporters: Pharmacological insights on substrate and inhibitor specificity of the EAAT subtypes. Pharmacol. Ther. 2005, 107, 271–285. [Google Scholar] [CrossRef] [PubMed]
  35. Sattler, R.; Tymianski, M. Molecular mechanisms of glutamate receptor-mediated excitotoxic neuronal cell death. Mol. Neurobiol. 2001, 24, 107–129. [Google Scholar] [CrossRef]
  36. Wu, Q.J.; Tymianski, M. Targeting NMDA receptors in stroke: New hope in neuroprotection. Mol. Brain 2018, 11, 15. [Google Scholar] [CrossRef] [PubMed]
  37. Choi, D.W. Glutamate neurotoxicity and diseases of the nervous system. Neuron 1988, 1, 623–634. [Google Scholar] [CrossRef]
  38. Joy, M.T.; Carmichael, S.T. Encouraging an excitable brain state: Mechanisms of brain repair in stroke. Nat. Rev. Neurosci. 2021, 22, 38–53. [Google Scholar] [CrossRef]
  39. Saeed, S.A.; Shad, K.F.; Saleem, T.; Javed, F.; Khan, M.U. Some new prospects in the understanding of the molecular basis of the pathogenesis of stroke. Exp. Brain Res. 2007, 182, 1. [Google Scholar] [CrossRef]
  40. Cabral-Costa, J.V.; Kowaltowski, A.J. Neurological disorders and mitochondria. Mol. Asp. Med. 2020, 71, 100826. [Google Scholar] [CrossRef]
  41. Görlach, A.; Bertram, K.; Hudecova, S.; Krizanova, O. Calcium and ROS: A mutual interplay. Redox Biol. 2015, 6, 260–271. [Google Scholar] [CrossRef] [Green Version]
  42. Cobley, J.N.; Fiorello, M.L.; Bailey, D.M. 13 reasons why the brain is susceptible to oxidative stress. Redox Biol. 2018, 15, 490–503. [Google Scholar] [CrossRef]
  43. Zorov, D.B.; Juhaszova, M.; Sollott, S.J. Mitochondrial Reactive Oxygen Species (ROS) and ROS-Induced ROS Release. Physiol. Rev. 2014, 94, 909–950. [Google Scholar] [CrossRef] [Green Version]
  44. Murphy, E.; Ardehali, H.; Balaban, R.S.; DiLisa, F.; Dorn, G.W., 2nd; Kitsis, R.N.; Otsu, K.; Ping, P.; Rizzuto, R.; Sack, M.N.; et al. Mitochondrial Function, Biology, and Role in Disease. Circ. Res. 2016, 118, 1960–1991. [Google Scholar] [CrossRef]
  45. He, Z.; Ning, N.; Zhou, Q.; Khoshnam, S.E.; Farzaneh, M. Mitochondria as a therapeutic target for ischemic stroke. Free Radic. Biol. Med. 2020, 146, 45–58. [Google Scholar] [CrossRef]
  46. Abramov, A.Y.; Scorziello, A.; Duchen, M. Three Distinct Mechanisms Generate Oxygen Free Radicals in Neurons and Contribute to Cell Death during Anoxia and Reoxygenation. J. Neurosci. 2007, 27, 1129. [Google Scholar] [CrossRef]
  47. Lo, E.H.; Dalkara, T.; Moskowitz, M.A. Mechanisms, challenges and opportunities in stroke. Nat. Rev. Neurosci. 2003, 4, 399–414. [Google Scholar] [CrossRef]
  48. Parvez, S.; Kaushik, M.; Ali, M.; Alam, M.M.; Ali, J.; Tabassum, H.; Kaushik, P. Dodging blood brain barrier with “nano” warriors: Novel strategy against ischemic stroke. Theranostics 2022, 12, 689–719. [Google Scholar] [CrossRef]
  49. Shi, K.; Tian, D.-C.; Li, Z.-G.; Ducruet, A.F.; Lawton, M.T.; Shi, F.-D. Global brain inflammation in stroke. Lancet Neurol. 2019, 18, 1058–1066. [Google Scholar] [CrossRef]
  50. Endres, M.; Moro, M.A.; Nolte, C.H.; Dames, C.; Buckwalter, M.S.; Meisel, A. Immune Pathways in Etiology, Acute Phase, and Chronic Sequelae of Ischemic Stroke. Circ. Res. 2022, 130, 1167–1186. [Google Scholar] [CrossRef]
  51. Dokalis, N.; Prinz, M. Resolution of neuroinflammation: Mechanisms and potential therapeutic option. Semin. Immunopathol. 2019, 41, 699–709. [Google Scholar] [CrossRef] [PubMed]
  52. Zhang, S.R.; Phan, T.G.; Sobey, C.G. Targeting the Immune System for Ischemic Stroke. Trends Pharmacol. Sci. 2021, 42, 96–105. [Google Scholar] [CrossRef] [PubMed]
  53. Iadecola, C.; Buckwalter, M.S.; Anrather, J. Immune responses to stroke: Mechanisms, modulation, and therapeutic potential. J. Clin. Investig. 2020, 130, 2777–2788. [Google Scholar] [CrossRef] [PubMed]
  54. Jiang, C.T.; Wu, W.F.; Deng, Y.H.; Ge, J.W. Modulators of microglia activation and polarization in ischemic stroke (Review). Mol. Med. Rep. 2020, 21, 2006–2018. [Google Scholar] [CrossRef] [Green Version]
  55. Kanazawa, M.; Ninomiya, I.; Hatakeyama, M.; Takahashi, T.; Shimohata, T. Microglia and Monocytes/Macrophages Polarization Reveal Novel Therapeutic Mechanism against Stroke. Int. J. Mol. Sci. 2017, 18, 2135. [Google Scholar] [CrossRef]
  56. Mesquida-Veny, F.; Del Río, J.A.; Hervera, A. Macrophagic and microglial complexity after neuronal injury. Prog. Neurobiol. 2021, 200, 101970. [Google Scholar] [CrossRef]
  57. Pocock, J.M.; Kettenmann, H. Neurotransmitter receptors on microglia. Trends Neurosci. 2007, 30, 527–535. [Google Scholar] [CrossRef]
  58. Xiong, X.-Y.; Liu, L.; Yang, Q.-W. Functions and mechanisms of microglia/macrophages in neuroinflammation and neurogenesis after stroke. Prog. Neurobiol. 2016, 142, 23–44. [Google Scholar] [CrossRef]
  59. Ma, Y.; Wang, J.; Wang, Y.; Yang, G.-Y. The biphasic function of microglia in ischemic stroke. Prog. Neurobiol. 2017, 157, 247–272. [Google Scholar] [CrossRef]
  60. De Meyer, S.F.; Denorme, F.; Langhauser, F.; Geuss, E.; Fluri, F.; Kleinschnitz, C. Thromboinflammation in Stroke Brain Damage. Stroke 2016, 47, 1165–1172. [Google Scholar] [CrossRef] [Green Version]
  61. Del Zoppo, G.J.; Mabuchi, T. Cerebral Microvessel Responses to Focal Ischemia. J. Cereb. Blood Flow Metab. 2003, 23, 879–894. [Google Scholar] [CrossRef]
  62. Lambertsen, K.L.; Finsen, B.; Clausen, B.H. Post-stroke inflammation—Target or tool for therapy? Acta Neuropathol. 2019, 137, 693–714. [Google Scholar] [CrossRef] [Green Version]
  63. Neumann, J.; Riek-Burchardt, M.; Herz, J.; Doeppner, T.R.; König, R.; Hütten, H.; Etemire, E.; Männ, L.; Klingberg, A.; Fischer, T.; et al. Very-late-antigen-4 (VLA-4)-mediated brain invasion by neutrophils leads to interactions with microglia, increased ischemic injury and impaired behavior in experimental stroke. Acta Neuropathol. 2015, 129, 259–277. [Google Scholar] [CrossRef]
  64. Peña-Martínez, C.; Durán-Laforet, V.; Garcia-Culebras, A.; Ostos, F.; Hernández-Jiménez, M.; Bravo-Ferrer, I.; Pérez-Ruiz, A.; Ballenilla, F.; Díaz-Guzmán, J.; Pradillo, J.; et al. Pharmacological Modulation of Neutrophil Extracellular Traps Reverses Thrombotic Stroke tPA (Tissue-Type Plasminogen Activator) Resistance. Stroke 2019, 50, 3228–3237. [Google Scholar] [CrossRef]
  65. Kang, L.; Yu, H.; Yang, X.; Zhu, Y.; Bai, X.; Wang, R.; Cao, Y.; Xu, H.; Luo, H.; Lu, L.; et al. Neutrophil extracellular traps released by neutrophils impair revascularization and vascular remodeling after stroke. Nat. Commun. 2020, 11, 2488. [Google Scholar] [CrossRef]
  66. Chen, R.; Zhang, X.; Gu, L.; Zhu, H.; Zhong, Y.; Ye, Y.; Xiong, X.; Jian, Z. New Insight Into Neutrophils: A Potential Therapeutic Target for Cerebral Ischemia. Front. Immunol. 2021, 12, 692061. [Google Scholar] [CrossRef]
  67. Splichal, Z.; Jurajda, M.; Duris, K. The Role of Inflammatory Response in Stroke Associated Programmed Cell Death. Curr. Neuropharmacol. 2018, 16, 1365–1374. [Google Scholar] [CrossRef]
  68. Datta, A.; Sarmah, D.; Mounica, L.; Kaur, H.; Kesharwani, R.; Verma, G.; Veeresh, P.; Kotian, V.; Kalia, K.; Borah, A.; et al. Cell Death Pathways in Ischemic Stroke and Targeted Pharmacotherapy. Transl. Stroke Res. 2020, 11, 1185–1202. [Google Scholar] [CrossRef]
  69. Fricker, M.; Tolkovsky, A.M.; Borutaite, V.; Coleman, M.; Brown, G.C. Neuronal Cell Death. Physiol. Rev. 2018, 98, 813–880. [Google Scholar] [CrossRef]
  70. Chao, D.T.; Korsmeyer, S.J. BCL-2 FAMILY: Regulators of Cell Death. Annu. Rev. Immunol. 1998, 16, 395–419. [Google Scholar] [CrossRef]
  71. Alam, M.; Alam, S.; Shamsi, A.; Adnan, M.; Elasbali, A.M.; Abu Al-Soud, W.; Alreshidi, M.; Hawsawi, Y.M.; Tippana, A.; Pasupuleti, V.R.; et al. Bax/Bcl-2 Cascade Is Regulated by the EGFR Pathway: Therapeutic Targeting of Non-Small Cell Lung Cancer. Front. Oncol. 2022, 12, 869672. [Google Scholar] [CrossRef]
  72. Yang, J.-L.; Mukda, S.; Chen, S.-D. Diverse roles of mitochondria in ischemic stroke. Redox Biol. 2018, 16, 263–275. [Google Scholar] [CrossRef]
  73. Kist, M.; Vucic, D. Cell death pathways: Intricate connections and disease implications. EMBO J. 2021, 40, e106700. [Google Scholar] [CrossRef]
  74. Tuo, Q.-z.; Zhang, S.-t.; Lei, P. Mechanisms of neuronal cell death in ischemic stroke and their therapeutic implications. Med. Res. Rev. 2022, 42, 259–305. [Google Scholar] [CrossRef]
  75. Almeida, A. Genetic determinants of neuronal vulnerability to apoptosis. Cell. Mol. Life Sci. 2013, 70, 71–88. [Google Scholar] [CrossRef]
  76. Galluzzi, L.; Morselli, E.; Kepp, O.; Kroemer, G. Targeting post-mitochondrial effectors of apoptosis for neuroprotection. Biochim. Biophys. Acta—Bioenerg. 2009, 1787, 402–413. [Google Scholar] [CrossRef] [Green Version]
  77. Susin, S.A.; Lorenzo, H.K.; Zamzami, N.; Marzo, I.; Snow, B.E.; Brothers, G.M.; Mangion, J.; Jacotot, E.; Costantini, P.; Loeffler, M.; et al. Molecular characterization of mitochondrial apoptosis-inducing factor. Nature 1999, 397, 441–446. [Google Scholar] [CrossRef]
  78. Landshamer, S.; Hoehn, M.; Barth, N.; Duvezin-Caubet, S.; Schwake, G.; Tobaben, S.; Kazhdan, I.; Becattini, B.; Zahler, S.; Vollmar, A.; et al. Bid-induced release of AIF from mitochondria causes immediate neuronal cell death. Cell Death Differ. 2008, 15, 1553–1563. [Google Scholar] [CrossRef]
  79. Culmsee, C.; Zhu, C.; Landshamer, S.; Becattini, B.; Wagner, E.; Pellecchia, M.; Blomgren, K.; Plesnila, N. Apoptosis-Inducing Factor Triggered by Poly(ADP-Ribose) Polymerase and Bid Mediates Neuronal Cell Death after Oxygen-Glucose Deprivation and Focal Cerebral Ischemia. J. Neurosci. 2005, 25, 10262. [Google Scholar] [CrossRef] [Green Version]
  80. Elmore, S. Apoptosis: A Review of Programmed Cell Death. Toxicol. Pathol. 2007, 35, 495–516. [Google Scholar] [CrossRef]
  81. Nakka, V.P.; Gusain, A.; Mehta, S.L.; Raghubir, R. Molecular Mechanisms of Apoptosis in Cerebral Ischemia: Multiple Neuroprotective Opportunities. Mol. Neurobiol. 2008, 37, 7–38. [Google Scholar] [CrossRef] [PubMed]
  82. Huang, J.-L.; Li, Y.; Zhao, B.-L.; Li, J.-S.; Zhang, N.; Ye, Z.-H.; Sun, X.-J.; Liu, W.-W. Necroptosis Signaling Pathways in Stroke: From Mechanisms to Therapies. Curr. Neuropharmacol. 2018, 16, 1327–1339. [Google Scholar] [CrossRef]
  83. Broughton, B.R.; Reutens, D.C.; Sobey, C.G.; Sims, K.; Politei, J.; Banikazemi, M.; Lee, P. Apoptotic Mechanisms after Cerebral Ischemia. Stroke 2009, 40, e331–e339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Shi, Q.; Cheng, Q.; Chen, C. The Role of Autophagy in the Pathogenesis of Ischemic Stroke. Curr. Neuropharmacol. 2021, 19, 629–640. [Google Scholar] [CrossRef]
  85. Egan, D.F.; Shackelford, D.B.; Mihaylova, M.M.; Gelino, S.; Kohnz, R.A.; Mair, W.; Vasquez, D.S.; Joshi, A.; Gwinn, D.M.; Taylor, R.; et al. Phosphorylation of ULK1 (hATG1) by AMP-Activated Protein Kinase Connects Energy Sensing to Mitophagy. Science 2011, 331, 456–461. [Google Scholar] [CrossRef] [Green Version]
  86. Liu, G.Y.; Sabatini, D.M. mTOR at the nexus of nutrition, growth, ageing and disease. Nat. Rev. Mol. Cell Biol. 2020, 21, 183–203. [Google Scholar] [CrossRef]
  87. Wang, P.; Shao, B.-Z.; Deng, Z.; Chen, S.; Yue, Z.; Miao, C.-Y. Autophagy in ischemic stroke. Prog. Neurobiol. 2018, 163–164, 98–117. [Google Scholar] [CrossRef]
  88. Hosokawa, N.; Hara, T.; Kaizuka, T.; Kishi, C.; Takamura, A.; Miura, Y.; Iemura, S.-I.; Natsume, T.; Takehana, K.; Yamada, N.; et al. Nutrient-dependent mTORC1 Association with the ULK1–Atg13–FIP200 Complex Required for Autophagy. Mol. Biol. Cell 2009, 20, 1981–1991. [Google Scholar] [CrossRef] [Green Version]
  89. Jung, C.H.; Jun, C.B.; Ro, S.-H.; Kim, Y.-M.; Otto, N.M.; Cao, J.; Kundu, M.; Kim, D.-H. ULK-Atg13-FIP200 Complexes Mediate mTOR Signaling to the Autophagy Machinery. Mol. Biol. Cell 2009, 20, 1992–2003. [Google Scholar] [CrossRef] [Green Version]
  90. Hardie, D.G. AMP-activated/SNF1 protein kinases: Conserved guardians of cellular energy. Nat. Rev. Mol. Cell Biol. 2007, 8, 774–785. [Google Scholar] [CrossRef]
  91. Tripathi, D.N.; Chowdhury, R.; Trudel, L.J.; Tee, A.R.; Slack, R.S.; Walker, C.L.; Wogan, G.N. Reactive nitrogen species regulate autophagy through ATM-AMPK-TSC2–mediated suppression of mTORC1. Proc. Natl. Acad. Sci. USA 2013, 110, E2950–E2957. [Google Scholar] [CrossRef] [Green Version]
  92. Kim, J.; Kundu, M.; Viollet, B.; Guan, K.-L. AMPK and mTOR regulate autophagy through direct phosphorylation of Ulk1. Nat. Cell Biol. 2011, 13, 132–141. [Google Scholar] [CrossRef] [Green Version]
  93. Bento, C.F.; Renna, M.; Ghislat, G.; Puri, C.; Ashkenazi, A.; Vicinanza, M.; Menzies, F.M.; Rubinsztein, D.C. Mammalian Autophagy: How Does It Work? Annu. Rev. Biochem. 2016, 85, 685–713. [Google Scholar] [CrossRef]
  94. Russell, R.C.; Tian, Y.; Yuan, H.; Park, H.W.; Chang, Y.-Y.; Kim, J.; Kim, H.; Neufeld, T.P.; Dillin, A.; Guan, K.-L. ULK1 induces autophagy by phosphorylating Beclin-1 and activating VPS34 lipid kinase. Nat. Cell Biol. 2013, 15, 741–750. [Google Scholar] [CrossRef] [Green Version]
  95. Mizushima, N.; Yoshimori, T.; Ohsumi, Y. The Role of Atg Proteins in Autophagosome Formation. Annu. Rev. Cell Dev. Biol. 2011, 27, 107–132. [Google Scholar] [CrossRef]
  96. Kim, K.-A.; Shin, D.; Kim, J.-H.; Shin, Y.-J.; Rajanikant, G.; Majid, A.; Baek, S.-H.; Bae, O.-N. Role of Autophagy in Endothelial Damage and Blood–Brain Barrier Disruption in Ischemic Stroke. Stroke 2018, 49, 1571–1579. [Google Scholar] [CrossRef]
  97. Dikic, I.; Elazar, Z. Mechanism and medical implications of mammalian autophagy. Nat. Rev. Mol. Cell Biol. 2018, 19, 349–364. [Google Scholar] [CrossRef]
  98. Ajoolabady, A.; Wang, S.; Kroemer, G.; Penninger, J.M.; Uversky, V.N.; Pratico, D.; Henninger, N.; Reiter, R.J.; Bruno, A.; Joshipura, K.; et al. Targeting autophagy in ischemic stroke: From molecular mechanisms to clinical therapeutics. Pharmacol. Ther. 2021, 225, 107848. [Google Scholar] [CrossRef]
  99. Zhang, X.; Li, Z.; Yang, P.; Duan, G.; Liu, X.; Gu, Z.; Li, Y. Polyphenol scaffolds in tissue engineering. Mater. Horiz. 2021, 8, 145–167. [Google Scholar] [CrossRef]
  100. Yousefian, M.; Shakour, N.; Hosseinzadeh, H.; Hayes, A.W.; Hadizadeh, F.; Karimi, G. The natural phenolic compounds as modulators of NADPH oxidases in hypertension. Phytomedicine 2019, 55, 200–213. [Google Scholar] [CrossRef]
  101. Tresserra-Rimbau, A.; Lamuela-Raventos, R.M.; Moreno, J.J. Polyphenols, food and pharma. Current knowledge and directions for future research. Biochem. Pharmacol. 2018, 156, 186–195. [Google Scholar] [CrossRef]
  102. Zhang, Z.; Xie, L.; Ju, Y.; Dai, Y. Recent Advances in Metal-Phenolic Networks for Cancer Theranostics. Small 2021, 17, 2100314. [Google Scholar] [CrossRef]
  103. Shamsi, A.; Anwar, S.; Mohammad, T.; Shahwan, M.; Hassan, I.; Islam, A. Therapeutic Potential of Polyphenols in Alzheimer’s Therapy: Broad-Spectrum and Minimal Side Effects as Key Aspects; Springer: Singapore, 2021; pp. 111–133. [Google Scholar] [CrossRef]
  104. Xu, H.; Wang, E.; Chen, F.; Xiao, J.; Wang, M. Neuroprotective Phytochemicals in Experimental Ischemic Stroke: Mechanisms and Potential Clinical Applications. Oxid. Med. Cell. Longev. 2021, 2021, 6687386. [Google Scholar] [CrossRef]
  105. Hamsalakshmi; Alex, A.M.; Marappa, M.A.; Joghee, S.; Chidambaram, S.B. Therapeutic benefits of flavonoids against neuroinflammation: A systematic review. Inflammopharmacology 2022, 30, 111–136. [CrossRef]
  106. Ulya, T.; Ardianto, C.; Anggreini, P.; Budiatin, A.S.; Setyawan, D.; Khotib, J. Quercetin promotes behavioral recovery and biomolecular changes of melanocortin-4 receptor in mice with ischemic stroke. J. Basic Clin. Physiol. Pharmacol. 2021, 32, 349–355. [Google Scholar] [CrossRef]
  107. Lee, Y.L.; Park, S.M.; Chen, B.H.; Park, J.H.; Ahn, J.H.; Cho, J.H.; Kim, I.H.; Lee, J.C.; Won, M.-H.; Lee, C.-H.; et al. Pretreated quercetin protects gerbil hippocampal CA1 pyramidal neurons from transient cerebral ischemic injury by increasing the expression of antioxidant enzymes. Neural Regen. Res. 2017, 12, 220–227. [Google Scholar] [CrossRef] [PubMed]
  108. Yang, R.; Shen, Y.-J.; Chen, M.; Zhao, J.-Y.; Chen, S.-H.; Zhang, W.; Song, J.-K.; Li, L.; Du, G.-H. Quercetin attenuates ischemia reperfusion injury by protecting the blood-brain barrier through Sirt1 in MCAO rats. J. Asian Nat. Prod. Res. 2022, 24, 278–289. [Google Scholar] [CrossRef] [PubMed]
  109. Park, D.-J.; Kang, J.-B.; Shah, F.-A.; Jin, Y.-B.; Koh, P.-O. Quercetin Attenuates Decrease of Thioredoxin Expression Following Focal Cerebral Ischemia and Glutamate-induced Neuronal Cell Damage. Neuroscience 2020, 428, 38–49. [Google Scholar] [CrossRef] [PubMed]
  110. Park, D.-J.; Jeon, S.-J.; Kang, J.-B.; Koh, P.-O. Quercetin Reduces Ischemic Brain Injury by Preventing Ischemia-induced Decreases in the Neuronal Calcium Sensor Protein Hippocalcin. Neuroscience 2020, 430, 47–62. [Google Scholar] [CrossRef]
  111. Wang, Y.-Y.; Chang, C.-Y.; Lin, S.-Y.; Wang, J.-D.; Wu, C.-C.; Chen, W.-Y.; Kuan, Y.-H.; Liao, S.-L.; Wang, W.-Y.; Chen, C.-J. Quercetin protects against cerebral ischemia/reperfusion and oxygen glucose deprivation/reoxygenation neurotoxicity. J. Nutr. Biochem. 2020, 83, 108436. [Google Scholar] [CrossRef]
  112. Park, D.-J.; Kang, J.-B.; Shah, F.-A.; Koh, P.-O. Quercetin attenuates the reduction of parvalbumin in middle cerebral artery occlusion animal model. Lab. Anim. Res. 2021, 37, 9. [Google Scholar] [CrossRef]
  113. Park, D.-J.; Shah, F.-A.; Koh, P.-O. Quercetin attenuates neuronal cells damage in a middle cerebral artery occlusion animal model. J. Vet. Med. Sci. 2018, 80, 676–683. [Google Scholar] [CrossRef] [Green Version]
  114. Park, D.-J.; Kang, J.-B.; Shah, M.-A.; Koh, P.-O. Quercetin alleviates the injury-induced decrease of protein phosphatase 2A subunit B in cerebral ischemic animal model and glutamate-exposed HT22 cells. J. Vet. Med. Sci. 2019, 81, 1047–1054. [Google Scholar] [CrossRef] [Green Version]
  115. Le, K.; Song, Z.; Deng, J.; Peng, X.; Zhang, J.; Wang, L.; Zhou, L.; Bi, H.; Liao, Z.; Feng, Z. Quercetin alleviates neonatal hypoxic-ischemic brain injury by inhibiting microglia-derived oxidative stress and TLR4-mediated inflammation. Inflamm. Res. 2020, 69, 1201–1213. [Google Scholar] [CrossRef]
  116. Jeon, S.-J.; Kim, M.-O.; Ali-Shah, F.; Koh, P.-O. Quercetin attenuates the injury-induced reduction of γ-enolase expression in a middle cerebral artery occlusion animal model. LAR 2017, 33, 308–314. [Google Scholar] [CrossRef] [Green Version]
  117. Wang, C.-P.; Shi, Y.-W.; Tang, M.; Zhang, X.-C.; Gu, Y.; Liang, X.-M.; Wang, Z.-W.; Ding, F. Isoquercetin Ameliorates Cerebral Impairment in Focal Ischemia through Anti-Oxidative, Anti-Inflammatory, and Anti-Apoptotic Effects in Primary Culture of Rat Hippocampal Neurons and Hippocampal CA1 Region of Rats. Mol. Neurobiol. 2017, 54, 2126–2142. [Google Scholar] [CrossRef]
  118. 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]
  119. Liu, H.; Zhong, L.; Zhang, Y.; Liu, X.; Li, J. Rutin attenuates cerebral ischemia–reperfusion injury in ovariectomized rats via estrogen-receptor-mediated BDNF–TrkB and NGF–TrkA signaling. Biochem. Cell Biol. 2018, 96, 672–681. [Google Scholar] [CrossRef]
  120. 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]
  121. Wang, J.; Mao, J.; Wang, R.; Li, S.; Wu, B.; Yuan, Y. Kaempferol Protects against Cerebral Ischemia Reperfusion Injury through Intervening Oxidative and Inflammatory Stress Induced Apoptosis. Front. Pharmacol. 2020, 11, 424. [Google Scholar] [CrossRef] [Green Version]
  122. Zhou, Y.-p.; Li, G.-c. Kaempferol Protects Cell Damage in In Vitro Ischemia Reperfusion Model in Rat Neuronal PC12 Cells. BioMed Res. Int. 2020, 2020, 2461079. [Google Scholar] [CrossRef]
  123. Wu, B.; Luo, H.; Zhou, X.; Cheng, C.-Y.; Lin, L.; Liu, B.-L.; Liu, K.; Li, P.; Yang, H. Succinate-induced neuronal mitochondrial fission and hexokinase II malfunction in ischemic stroke: Therapeutical effects of kaempferol. Biochim. Biophys. Acta—Mol. Basis Dis. 2017, 1863, 2307–2318. [Google Scholar] [CrossRef]
  124. Yuan, Y.; Zhai, Y.; Chen, J.; Xu, X.; Wang, H. Kaempferol Ameliorates Oxygen-Glucose Deprivation/Reoxygenation-Induced Neuronal Ferroptosis by Activating Nrf2/SLC7A11/GPX4 Axis. Biomolecules 2021, 11, 923. [Google Scholar] [CrossRef]
  125. Liu, D.; Ye, Y.; Xu, L.; Yuan, W.; Zhang, Q. Icariin and mesenchymal stem cells synergistically promote angiogenesis and neurogenesis after cerebral ischemia via PI3K and ERK1/2 pathways. Biomed. Pharmacother. 2018, 108, 663–669. [Google Scholar] [CrossRef]
  126. Mo, Z.-T.; Zheng, J.; Liao, Y.-L. Icariin inhibits the expression of IL-1β, IL-6 and TNF-α induced by OGD/R through the IRE1/XBP1s pathway in microglia. Pharm. Biol. 2021, 59, 1471–1477. [Google Scholar] [CrossRef]
  127. Dai, M.; Chen, B.; Wang, X.; Gao, C.; Yu, H. Icariin enhance mild hypothermia-induced neuroprotection via inhibiting the activation of NF-κB in experimental ischemic stroke. Metab. Brain Dis. 2021, 36, 1779–1790. [Google Scholar] [CrossRef]
  128. 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] [Green Version]
  129. Zhang, S.; Hu, X.; Guo, S.; Shi, L.; He, Q.; Zhang, P.; Yu, S.; Zhao, R. 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]
  130. 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]
  131. Mei, Z.-G.; Feng, Z.-T.; Wang, J.-F.; Fu, Y.; Yang, S.-B.; Zhang, S.-Z.; Huang, W.-F.; Xiong, L.; Zhou, H.-J.; Tao, W. 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] [CrossRef]
  132. Wang, S.; Wang, J.; Wei, H.; Gu, T.; Wang, J.; Wu, Z.; Yang, Q. Genistein Attenuates Acute Cerebral Ischemic Damage by Inhibiting the NLRP3 Inflammasome in Reproductively Senescent Mice. Front. Aging Neurosci. 2020, 12, 153. [Google Scholar] [CrossRef] [PubMed]
  133. 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] [PubMed]
  134. Lu, L.-Y.; Liu, Y.; Gong, Y.-F.; Zheng, X.-Y. A preliminary report: Genistein attenuates cerebral ischemia injury in ovariectomized rats via regulation of the PI3K-Akt-mTOR pathway. Gen. Physiol. Biophys. 2019, 38, 389–397. [Google Scholar] [CrossRef] [PubMed]
  135. Zheng, M.; Zhou, M.; Chen, M.; Lu, Y.; Shi, D.; Wang, J.; Liu, C. Neuroprotective Effect of Daidzein Extracted from Pueraria lobate Radix in a Stroke Model Via the Akt/mTOR/BDNF Channel. Front. Pharmacol. 2022, 12, 772485. [Google Scholar] [CrossRef]
  136. Yang, S.; Wang, H.; Yang, Y.; Wang, R.; Wang, Y.; Wu, C.; Du, G. Baicalein administered in the subacute phase ameliorates ischemia-reperfusion-induced brain injury by reducing neuroinflammation and neuronal damage. Biomed. Pharmacother. 2019, 117, 109102. [Google Scholar] [CrossRef]
  137. Yuan, Y.; Men, W.; Shan, X.; Zhai, H.; Qiao, X.; Geng, L.; Li, C. Baicalein exerts neuroprotective effect against ischaemic/reperfusion injury via alteration of NF-kB and LOX and AMPK/Nrf2 pathway. Inflammopharmacology 2020, 28, 1327–1341. [Google Scholar] [CrossRef]
  138. Ran, Y.; Qie, S.; Gao, F.; Ding, Z.; Yang, S.; Tian, G.; Liu, Z.; Xi, J. Baicalein ameliorates ischemic brain damage through suppressing proinflammatory microglia polarization via inhibiting the TLR4/NF-κB and STAT1 pathway. Brain Res. 2021, 1770, 147626. [Google Scholar] [CrossRef]
  139. Li, W.-H.; Yang, Y.-L.; Cheng, X.; Liu, M.; Zhang, S.-S.; Wang, Y.-H.; Du, G.-H. Baicalein attenuates caspase-independent cells death via inhibiting PARP-1 activation and AIF nuclear translocation in cerebral ischemia/reperfusion rats. Apoptosis 2020, 25, 354–369. [Google Scholar] [CrossRef]
  140. Li, S.; Zhang, Y.; Fei, L.; Zhang, Y.; Pang, J.; Gao, W.; Fan, F.; Xing, Y.; Li, X. Baicalein-ameliorated cerebral ischemia-reperfusion injury dependent on calpain 1/AIF pathway. Biosci. Biotechnol. Biochem. 2022, 86, 305–312. [Google Scholar] [CrossRef]
  141. Song, X.; Gong, Z.; Liu, K.; Kou, J.; Liu, B.; Liu, K. Baicalin combats glutamate excitotoxicity via protecting glutamine synthetase from ROS-induced 20S proteasomal degradation. Redox Biol. 2020, 34, 101559. [Google Scholar] [CrossRef]
  142. Chen, H.-L.; Jia, W.-J.; Li, H.-E.; Han, H.; Li, F.; Zhang, X.-L.; Li, J.-J.; Yuan, Y.; Wu, C.-Y. Scutellarin Exerts Anti-Inflammatory Effects in Activated Microglia/Brain Macrophage in Cerebral Ischemia and in Activated BV-2 Microglia through Regulation of MAPKs Signaling Pathway. Neuromol. Med. 2020, 22, 264–277. [Google Scholar] [CrossRef]
  143. Li, Y.; Li, S.; Li, D. Breviscapine Alleviates Cognitive Impairments Induced by Transient Cerebral Ischemia/Reperfusion through Its Anti-Inflammatory and Anti-Oxidant Properties in a Rat Model. ACS Chem. Neurosci. 2020, 11, 4489–4498. [Google Scholar] [CrossRef]
  144. Sun, J.-B.; Li, Y.; Cai, Y.-F.; Huang, Y.; Liu, S.; Yeung, P.K.; Deng, M.-Z.; Sun, G.-S.; Zilundu, P.L.; Hu, Q.-S.; et al. Scutellarin protects oxygen/glucose-deprived astrocytes and reduces focal cerebral ischemic injury. Neural Regen. Res. 2018, 13, 1396–1407. [Google Scholar] [CrossRef]
  145. Zhang, P.; Guo, T.; He, H.; Yang, L.; Deng, Y. Breviscapine confers a neuroprotective efficacy against transient focal cerebral ischemia by attenuating neuronal and astrocytic autophagy in the penumbra. Biomed. Pharmacother. 2017, 90, 69–76. [Google Scholar] [CrossRef]
  146. 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]
  147. Liu, S.; Su, Y.; Sun, B.; Hao, R.; Pan, S.; Gao, X.; Dong, X.; Ismail, A.M.; Han, B. Luteolin Protects Against CIRI, Potentially via Regulation of the SIRT3/AMPK/mTOR Signaling Pathway. Neurochem. Res. 2020, 45, 2499–2515. [Google Scholar] [CrossRef]
  148. Sarkaki, A.; Farbood, Y.; Mansouri, M.T.; Badavi, M.; Khorsandi, L.; Dehcheshmeh, M.G.; Shooshtari, M.K. Chrysin prevents cognitive and hippocampal long-term potentiation deficits and inflammation in rat with cerebral hypoperfusion and reperfusion injury. Life Sci. 2019, 226, 202–209. [Google Scholar] [CrossRef]
  149. Li, T.-F.; Ma, J.; Han, X.-W.; Jia, Y.-X.; Yuan, H.-F.; Shui, S.-F.; Guo, D.; Yan, L. Chrysin ameliorates cerebral ischemia/reperfusion (I/R) injury in rats by regulating the PI3K/Akt/mTOR pathway. Neurochem. Int. 2019, 129, 104496. [Google Scholar] [CrossRef]
  150. El Khashab, I.H.; Abdelsalam, R.M.; Elbrairy, A.I.; Attia, A.S. Chrysin attenuates global cerebral ischemic reperfusion injury via suppression of oxidative stress, inflammation and apoptosis. Biomed. Pharmacother. 2019, 112, 108619. [Google Scholar] [CrossRef]
  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. Ling, C.; Lei, C.; Zou, M.; Cai, X.; Xiang, Y.; Xie, Y.; Li, X.; Huang, D.; Wang, Y. Neuroprotective effect of apigenin against cerebral ischemia/reperfusion injury. J. Int. Med. Res. 2020, 48, 0300060520945859. [Google Scholar] [CrossRef]
  153. Wang, N.; Xu, Z.; Chen, K.; Liu, T.; Guo, W.; Xu, Z. 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] [Green Version]
  154. Zhang, J.-C.; Xu, H.; Yuan, Y.; Chen, J.-Y.; Zhang, Y.-J.; Lin, Y.; Yuan, S.-Y. Delayed Treatment with Green Tea Polyphenol EGCG Promotes Neurogenesis after Ischemic Stroke in Adult Mice. Mol. Neurobiol. 2017, 54, 3652–3664. [Google Scholar] [CrossRef]
  155. Park, D.-J.; Kang, J.-B.; Koh, P.-O. Epigallocatechin gallate alleviates neuronal cell damage against focal cerebral ischemia in rats. J. Vet. Med. Sci. 2020, 82, 639–645. [Google Scholar] [CrossRef] [Green Version]
  156. Park, D.-J.; Kang, J.-B.; Shah, M.-A.; Koh, P.-O. Epigallocatechin Gallate Alleviates Down-Regulation of Thioredoxin in Ischemic Brain Damage and Glutamate-Exposed Neuron. Neurochem. Res. 2021, 46, 3035–3049. [Google Scholar] [CrossRef]
  157. 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]
  158. Yang, B.; Sun, Y.; Lv, C.; Zhang, W.; Chen, Y. Procyanidins exhibits neuroprotective activities against cerebral ischemia reperfusion injury by inhibiting TLR4-NLRP3 inflammasome signal pathway. Psychopharmacology 2020, 237, 3283–3293. [Google Scholar] [CrossRef]
  159. 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]
  160. Feng, J.; Chen, X.; Lu, S.; Li, W.; Yang, D.; Su, W.; Wang, X.; Shen, J. Naringin Attenuates Cerebral Ischemia-Reperfusion Injury through Inhibiting Peroxynitrite-Mediated Mitophagy Activation. Mol. Neurobiol. 2018, 55, 9029–9042. [Google Scholar] [CrossRef]
  161. Zhang, J.; Jiang, H.; Wu, F.; Chi, X.; Pang, Y.; Jin, H.; Sun, Y.; Zhang, S. Neuroprotective Effects of Hesperetin in Regulating Microglia Polarization after Ischemic Stroke by Inhibiting TLR4/NF-κB Pathway. J. Healthc. Eng. 2021, 2021, 9938874. [Google Scholar] [CrossRef]
  162. Xu, B.; He, X.; Sui, Y.; Wang, X.; Wang, X.; Ren, L.; Zhai, Y.-X. Ginkgetin aglycone attenuates neuroinflammation and neuronal injury in the rats with ischemic stroke by modulating STAT3/JAK2/SIRT1. Folia Neuropathol. 2019, 57, 16–23. [Google Scholar] [CrossRef] [PubMed]
  163. Cui, H.-X.; Chen, J.-H.; 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, 17880. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Sukprasansap, M.; Chanvorachote, P.; Tencomnao, T. Cyanidin-3-glucoside activates Nrf2-antioxidant response element and protects against glutamate-induced oxidative and endoplasmic reticulum stress in HT22 hippocampal neuronal cells. BMC Complement. Med. Ther. 2020, 20, 46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Pan, Z.; Cui, M.; Dai, G.; Yuan, T.; Li, Y.; Ji, T.; Pan, Y. Protective Effect of Anthocyanin on Neurovascular Unit in Cerebral Ischemia/Reperfusion Injury in Rats. Front. Neurosci. 2018, 12, 947. [Google Scholar] [CrossRef]
  166. Cai, Y.; Li, X.; Pan, Z.; Zhu, Y.; Tuo, J.; Meng, Q.; Dai, G.; Yang, G.; Pan, Y. Anthocyanin ameliorates hypoxia and ischemia induced inflammation and apoptosis by increasing autophagic flux in SH-SY5Y cells. Eur. J. Pharmacol. 2020, 883, 173360. [Google Scholar] [CrossRef]
  167. Amanzadeh, E.; Esmaeili, A.; Rahgozar, S.; Nourbakhshnia, M. Application of quercetin in neurological disorders: From nutrition to nanomedicine. Rev. Neurosci. 2019, 30, 555–572. [Google Scholar] [CrossRef]
  168. Kaur, R.; Sood, A.; Lang, D.K.; Bhatia, S.; Al-Harrasi, A.; Aleya, L.; Behl, T. Potential of flavonoids as anti-Alzheimer’s agents: Bench to bedside. Environ. Sci. Pollut. Res. 2022, 29, 26063–26077. [Google Scholar] [CrossRef]
  169. Kumar, P.; Sharma, G.; Kumar, R.; Singh, B.; Malik, R.; Katare, O.P.; Raza, K. Promises of a biocompatible nanocarrier in improved brain delivery of quercetin: Biochemical, pharmacokinetic and biodistribution evidences. Int. J. Pharm. 2016, 515, 307–314. [Google Scholar] [CrossRef]
  170. Morand, C.; Manach, C.; Crespy, V.; Remesy, C. Quercetin 3-O-β-glucoside is better absorbed than other quercetin forms and is not present in rat plasma. Free Radic. Res. 2000, 33, 667–676. [Google Scholar] [CrossRef]
  171. Scott, E.; Zhang, Q.-G.; Wang, R.; Vadlamudi, R.; Brann, D. Estrogen neuroprotection and the critical period hypothesis. Front. Neuroendocrinol. 2012, 33, 85–104. [Google Scholar] [CrossRef] [Green Version]
  172. Renoux, C.; Suissa, S. Hormone Therapy Administration in Postmenopausal Women and Risk of Stroke. Women’s Health 2011, 7, 355–361. [Google Scholar] [CrossRef]
  173. Rajendran, P.; Rengarajan, T.; Nandakumar, N.; Palaniswami, R.; Nishigaki, Y.; Nishigaki, I. Kaempferol, a potential cytostatic and cure for inflammatory disorders. Eur. J. Med. Chem. 2014, 86, 103–112. [Google Scholar] [CrossRef]
  174. Zeng, Y.; Xiong, Y.; Yang, T.; Wang, Y.; Zeng, J.; Zhou, S.; Luo, Y.; Li, L. Icariin and its metabolites as potential protective phytochemicals against cardiovascular disease: From effects to molecular mechanisms. Biomed. Pharmacother. 2022, 147, 112642. [Google Scholar] [CrossRef]
  175. Yuan, J.-Y.; Tong, Z.-Y.; Dong, Y.-C.; Zhao, J.-Y.; Shang, Y. Research progress on icariin, a traditional Chinese medicine extract, in the treatment of asthma. Allergol. Immunopathol. 2022, 50, 9–16. [Google Scholar] [CrossRef]
  176. Chuang, Y.; Van, I.; Zhao, Y.; Xu, Y. Icariin ameliorate Alzheimer’s disease by influencing SIRT1 and inhibiting Aβ cascade pathogenesis. J. Chem. Neuroanat. 2021, 117, 102014. [Google Scholar] [CrossRef]
  177. 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]
  178. Leonard, L.M.; Choi, M.S.; Cross, T.-W.L. Maximizing the Estrogenic Potential of Soy Isoflavones through the Gut Microbiome: Implication for Cardiometabolic Health in Postmenopausal Women. Nutrients 2022, 14, 553. [Google Scholar] [CrossRef]
  179. Lee, A.W.; Poynor, V.; McEligot, A.J. Urinary Phytoestrogen Levels Are Associated with Female Hormonal Cancers: An Analysis of NHANES Data from 1999 to 2010. Nutr. Cancer 2022, 1–9. [Google Scholar] [CrossRef]
  180. Li-Weber, M. New therapeutic aspects of flavones: The anticancer properties of Scutellaria and its main active constituents Wogonin, Baicalein and Baicalin. Cancer Treat. Rev. 2009, 35, 57–68. [Google Scholar] [CrossRef]
  181. Liang, W.; Huang, X.; Chen, W. The Effects of Baicalin and Baicalein on Cerebral Ischemia: A Review. Aging Dis. 2017, 8, 850–867. [Google Scholar] [CrossRef] [Green Version]
  182. Yuan, Y.; Rangarajan, P.; Kan, E.M.; Wu, Y.; Wu, C.; Ling, E.-A. Scutellarin regulates the Notch pathway and affects the migration and morphological transformation of activated microglia in experimentally induced cerebral ischemia in rats and in activated BV-2 microglia. J. Neuroinflamm. 2015, 12, 11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Lopez-Lazaro, M. Distribution and biological activities of the flavonoid luteolin. Mini Rev. Med. Chem. 2009, 9, 31–59. [Google Scholar] [CrossRef] [PubMed]
  184. Bordet, R.; Ouk, T.; Petrault, O.; Gelé, P.; Gautier, S.; Laprais, M.; Deplanque, D.; Duriez, P.; Staels, B.; Fruchart, J.C.; et al. PPAR: A new pharmacological target for neuroprotection in stroke and neurodegenerative diseases. Biochem. Soc. Trans. 2006, 34, 1341–1346. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Cai, W.; Yang, T.; Liu, H.; Han, L.; Zhang, K.; Hu, X.; Zhang, X.; Yin, K.-J.; Gao, Y.; Bennett, M.V.; et al. Peroxisome proliferator-activated receptor γ (PPARγ): A master gatekeeper in CNS injury and repair. Prog. Neurobiol. 2018, 163–164, 27–58. [Google Scholar] [CrossRef] [PubMed]
  186. Kim, H.-S.; Quon, M.; Kim, J.-A. New insights into the mechanisms of polyphenols beyond antioxidant properties; lessons from the green tea polyphenol, epigallocatechin 3-gallate. Redox Biol. 2014, 2, 187–195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Payne, A.; Nahashon, S.; Taka, E.; Adinew, G.M.; Soliman, K.F.A. Epigallocatechin-3-Gallate (EGCG): New Therapeutic Perspectives for Neuroprotection, Aging, and Neuroinflammation for the Modern Age. Biomolecules 2022, 12, 371. [Google Scholar] [CrossRef]
  188. Yu, S.-W.; Andrabi, S.A.; Wang, H.; Kim, N.S.; Poirier, G.G.; Dawson, T.M.; Dawson, V.L. Apoptosis-inducing factor mediates poly(ADP-ribose) (PAR) polymer-induced cell death. Proc. Natl. Acad. Sci. USA 2006, 103, 18314–18319. [Google Scholar] [CrossRef] [Green Version]
  189. Erlund, I. Review of the flavonoids quercetin, hesperetin, and naringenin. Dietary sources, bioactivities, bioavailability, and epidemiology. Nutr. Res. 2004, 24, 851–874. [Google Scholar] [CrossRef]
  190. Henriques, J.F.; Serra, D.; Dinis, T.C.P.; Almeida, L.M. The Anti-Neuroinflammatory Role of Anthocyanins and Their Metabolites for the Prevention and Treatment of Brain Disorders. Int. J. Mol. Sci. 2020, 21, 8653. [Google Scholar] [CrossRef]
  191. Chen, Z.; Farag, M.A.; Zhong, Z.; Zhang, C.; Yang, Y.; Wang, S.; Wang, Y. Multifaceted role of phyto-derived polyphenols in nanodrug delivery systems. Adv. Drug Deliv. Rev. 2021, 176, 113870. [Google Scholar] [CrossRef]
  192. Chen, J.-L.; Duan, W.-J.; Luo, S.; Li, S.; Ma, X.-H.; Hou, B.-N.; Cheng, S.-Y.; Fang, S.-H.; Wang, Q.; Huang, S.-Q.; et al. Ferulic acid attenuates brain microvascular endothelial cells damage caused by oxygen-glucose deprivation via punctate-mitochondria-dependent mitophagy. Brain Res. 2017, 1666, 17–26. [Google Scholar] [CrossRef]
  193. Cheng, C.; Kao, S.; Lee, Y. Ferulic acid ameliorates cerebral infarction by activating Akt/mTOR/4E-BP1/Bcl-2 anti-apoptotic signaling in the penumbral cortex following permanent cerebral ischemia in rats. Mol. Med. Rep. 2019, 19, 792–804. [Google Scholar] [CrossRef] [Green Version]
  194. Cheng, C.-Y.; Kao, S.-T.; Lee, Y.-C. Ferulic Acid Exerts Anti-apoptotic Effects against Ischemic Injury by Activating HSP70/Bcl-2- and HSP70/Autophagy-Mediated Signaling after Permanent Focal Cerebral Ischemia in Rats. Am. J. Chin. Med. 2019, 47, 39–61. [Google Scholar] [CrossRef]
  195. Zhang, X.-J.; Cui, H.-Y.; 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] [CrossRef]
  196. 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]
  197. Shah, M.-A.; Kang, J.-B.; Park, D.-J.; Kim, M.-O.; Koh, P.-O. Chlorogenic acid alleviates neurobehavioral disorders and brain damage in focal ischemia animal models. Neurosci. Lett. 2021, 760, 136085. [Google Scholar] [CrossRef]
  198. Roshan-Milani, S.; Sattari, P.; Ghaderi-Pakdel, F.; Naderi, R. miR-23b/TAB3/NF-κB/p53 axis is involved in hippocampus injury induced by cerebral ischemia–reperfusion in rats: The protective effect of chlorogenic acid. BioFactors 2022, in press. [Google Scholar] [CrossRef]
  199. 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] [Green Version]
  200. Shah, M.-A.; Kang, J.-B.; Park, D.-J.; Kim, M.-O.; Koh, P.-O. Chlorogenic acid alleviates cerebral ischemia-induced neuroinflammation via attenuating nuclear factor kappa B activation. Neurosci. Lett. 2022, 773, 136495. [Google Scholar] [CrossRef]
  201. Zhang, W.; Song, J.-K.; Zhang, X.; Zhou, Q.-M.; He, G.-R.; Xu, X.-N.; Rong, Y.; Zhou, W.-X.; Du, G.-H. Salvianolic acid A attenuates ischemia reperfusion induced rat brain damage by protecting the blood brain barrier through MMP-9 inhibition and anti-inflammation. Chin. J. Nat. Med. 2018, 16, 184–193. [Google Scholar] [CrossRef]
  202. Song, J.; Zhang, W.; Wang, J.; Yang, H.; Zhou, Q.; Wang, H.; Li, L.; Du, G. Inhibition of FOXO3a/BIM signaling pathway contributes to the protective effect of salvianolic acid A against cerebral ischemia/reperfusion injury. Acta Pharm. Sin. B 2019, 9, 505–515. [Google Scholar] [CrossRef] [PubMed]
  203. Zhao, J.; Li, L.; Fang, G. Salvianolic acid A attenuates cerebral ischemia/reperfusion injury induced rat brain damage, inflammation and apoptosis by regulating miR-499a/DDK1. Am. J. Transl. Res. 2020, 12, 3288–3301. [Google Scholar] [PubMed]
  204. Liu, C.-D.; Liu, N.-N.; Zhang, S.; Ma, G.-D.; Yang, H.-G.; Kong, L.-L.; Du, G.-H. Salvianolic acid A prevented cerebrovascular endothelial injury caused by acute ischemic stroke through inhibiting the Src signaling pathway. Acta Pharmacol. Sin. 2021, 42, 370–381. [Google Scholar] [CrossRef] [PubMed]
  205. Ling, Y.; Jin, L.; Ma, Q.; Huang, Y.; Yang, Q.; Chen, M.; Shou, Q. Salvianolic acid A alleviated inflammatory response mediated by microglia through inhibiting the activation of TLR2/4 in acute cerebral ischemia-reperfusion. Phytomedicine 2021, 87, 153569. [Google Scholar] [CrossRef]
  206. Zhang, S.; Kong, D.-W.; Ma, G.-D.; Liu, C.-D.; Yang, Y.-J.; Liu, S.; Jiang, N.; Pan, Z.-R.; Zhang, W.; Kong, L.-L.; et al. Long-term administration of salvianolic acid A promotes endogenous neurogenesis in ischemic stroke rats through activating Wnt3a/GSK3β/β-catenin signaling pathway. Acta Pharmacol. Sin. 2022. [Google Scholar] [CrossRef]
  207. Xu, S.; Zhong, A.; Ma, H.; Li, D.; Hu, Y.; Xu, Y.; Zhang, J. Neuroprotective effect of salvianolic acid B against cerebral ischemic injury in rats via the CD40/NF-κB pathway associated with suppression of platelets activation and neuroinflammation. Brain Res. 2017, 1661, 37–48. [Google Scholar] [CrossRef]
  208. Guo, H.; Zhang, Z.; Gu, T.; Yu, D.; Shi, Y.; Gao, Z.; Wang, Z.; Liu, W.; Fan, Z.; Hou, W.; et al. Astrocytic glycogen mobilization participates in salvianolic acid B-mediated neuroprotection against reperfusion injury after ischemic stroke. Exp. Neurol. 2022, 349, 113966. [Google Scholar] [CrossRef]
  209. Kale, S.; Sarode, L.P.; Kharat, A.; Ambulkar, S.; Prakash, A.; Sakharkar, A.J.; Ugale, R.R. Protocatechuic Acid Prevents Early Hour Ischemic Reperfusion Brain Damage by Restoring Imbalance of Neuronal Cell Death and Survival Proteins. J. Stroke Cerebrovasc. Dis. 2021, 30, 105507. [Google Scholar] [CrossRef]
  210. Qu, Y.; Wang, L.; Mao, Y. Gallic acid attenuates cerebral ischemia/re-perfusion-induced blood–brain barrier injury by modifying polarization of microglia. J. Immunotoxicol. 2022, 19, 17–26. [Google Scholar] [CrossRef]
  211. Khoshnam, S.E.; Sarkaki, A.; Rashno, M.; Farbood, Y. Memory deficits and hippocampal inflammation in cerebral hypoperfusion and reperfusion in male rats: Neuroprotective role of vanillic acid. Life Sci. 2018, 211, 126–132. [Google Scholar] [CrossRef]
  212. Zhao, Q.; Wang, X.; Chen, A.; Cheng, X.; Zhang, G.; Sun, J.; Zhao, Y.; Huang, Y.; Zhu, Y. Rhein protects against cerebral ischemic-/reperfusion-induced oxidative stress and apoptosis in rats. Int. J. Mol. Med. 2018, 41, 2802–2812. [Google Scholar] [CrossRef]
  213. Neto-Neves, E.M.; Filho, C.D.S.M.B.; Dejani, N.N.; de Sousa, D.P. Ferulic Acid and Cardiovascular Health: Therapeutic and Preventive Potential. Mini-Rev. Med. Chem. 2021, 21, 1625–1637. [Google Scholar] [CrossRef]
  214. Chaudhary, A.; Jaswal, V.S.; Choudhary, S.; Sonika; Sharma, A.; Beniwal, V.; Tuli, H.S.; Sharma, S. Ferulic Acid: A Promising Therapeutic Phytochemical and Recent Patents Advances. Recent Pat. Inflamm. Allergy Drug Discov. 2019, 13, 115–123. [Google Scholar] [CrossRef]
  215. Chung, C.H.; Jung, W.; Keum, H.; Kim, T.W.; Jon, S. Nanoparticles Derived from the Natural Antioxidant Rosmarinic Acid Ameliorate Acute Inflammatory Bowel Disease. ACS Nano 2020, 14, 6887–6896. [Google Scholar] [CrossRef]
  216. Huerta-Madroñal, M.; Caro-León, J.; Espinosa-Cano, E.; Aguilar, M.R.; Vázquez-Lasa, B. Chitosan—Rosmarinic acid conjugates with antioxidant, anti-inflammatory and photoprotective properties. Carbohydr. Polym. 2021, 273, 118619. [Google Scholar] [CrossRef]
  217. Tajik, N.; Tajik, M.; Mack, I.; Enck, P. The potential effects of chlorogenic acid, the main phenolic components in coffee, on health: A comprehensive review of the literature. Eur. J. Nutr. 2017, 56, 2215–2244. [Google Scholar] [CrossRef]
  218. Yang, M.; Orgah, J.; Zhu, J.; Fan, G.; Han, J.; Wang, X.; Zhang, B.; Zhu, Y. Danhong injection attenuates cardiac injury induced by ischemic and reperfused neuronal cells through regulating arginine vasopressin expression and secretion. Brain Res. 2016, 1642, 516–523. [Google Scholar] [CrossRef]
  219. Wang, K.; Zhang, Y.; Li, X.; Chen, L.; Wang, H.; Wu, J.; Zheng, J.; Wu, D. Characterization of the Kremen-binding Site on Dkk1 and Elucidation of the Role of Kremen in Dkk-mediated Wnt Antagonism*. J. Biol. Chem. 2008, 283, 23371–23375. [Google Scholar] [CrossRef] [Green Version]
  220. Hu, P.; Liang, Q.-L.; Luo, G.-A.; Zhao, Z.-Z.; Jiang, Z.-H. Multi-component HPLC Fingerprinting of Radix Salviae Miltiorrhizae and Its LC-MS-MS Identification. Chem. Pharm. Bull. 2005, 53, 677–683. [Google Scholar] [CrossRef] [Green Version]
  221. Wu, Y.P.; Zhao, X.M.; Pan, S.D.; Guo, D.A.; Wei, R.; Han, J.J.; Kainoh, M.; Xia, Z.L.; de Groot, P.G.; Lisman, T. Salvianolic Acid B inhibits platelet adhesion under conditions of flow by a mechanism involving the collagen receptor α2β1. Thromb. Res. 2008, 123, 298–305. [Google Scholar] [CrossRef]
  222. Sun, Y.; Kotani, A.; Machida, K.; Yamamoto, K.; Hakamata, H. Determination of Phenolic Compounds in Beverages by Three-Flow Channel Isocratic HPLC with Electrochemical Detections Using a Column-Switching Technique. Chem. Pharm. Bull. 2022, 70, 43–49. [Google Scholar] [CrossRef]
  223. Loarca-Piña, G.; Mendoza, S.; Ramos-Gómez, M.; Reynoso, R. Antioxidant, Antimutagenic, and Antidiabetic Activities of Edible Leaves from Cnidoscolus chayamansa Mc. Vaugh. J. Food Sci. 2010, 75, H68–H72. [Google Scholar] [CrossRef]
  224. Durazzo, A.; Lucarini, M.; Camilli, E.; Marconi, S.; Gabrielli, P.; Lisciani, S.; Gambelli, L.; Aguzzi, A.; Novellino, E.; Santini, A.; et al. Dietary Lignans: Definition, Description and Research Trends in Databases Development. Molecules 2018, 23, 3251. [Google Scholar] [CrossRef] [Green Version]
  225. Zhou, Y.; Men, L.; Sun, Y.; Wei, M.; Fan, X. Pharmacodynamic effects and molecular mechanisms of lignans from Schisandra chinensis Turcz. (Baill.), a current review. Eur. J. Pharmacol. 2021, 892, 173796. [Google Scholar] [CrossRef]
  226. Liu, X.; Chen, X.; Zhu, Y.; Wang, K.; Wang, Y. Effect of magnolol on cerebral injury and blood brain barrier dysfunction induced by ischemia-reperfusion in vivo and in vitro. Metab. Brain Dis. 2017, 32, 1109–1118. [Google Scholar] [CrossRef]
  227. Kou, D.-Q.; Jiang, Y.-L.; Qin, J.-H.; Huang, Y.-H. Magnolol attenuates the inflammation and apoptosis through the activation of SIRT1 in experimental stroke rats. Pharmacol. Rep. 2017, 69, 642–647. [Google Scholar] [CrossRef]
  228. Huang, S.; Tai, S.; Chang, C.; Tu, Y.; Chang, C.; Lee, E. Magnolol protects against ischemic-reperfusion brain damage following oxygen-glucose deprivation and transient focal cerebral ischemia. Int. J. Mol. Med. 2018, 41, 2252–2262. [Google Scholar] [CrossRef] [Green Version]
  229. Liu, Z.; Xie, J.; Lin, K.; Qi, L. Influencing mechanism of magnolol on expression of BDNF and Bax in rats with cerebral ischemic stroke. Exp. Ther. Med. 2018, 16, 4423–4428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  230. Zong, W.; Gouda, M.; Cai, E.; Wang, R.; Xu, W.; Wu, Y.; Munekata, P.E.S.; Lorenzo, J.M. The Antioxidant Phytochemical Schisandrin A Promotes Neural Cell Proliferation and Differentiation after Ischemic Brain Injury. Molecules 2021, 26, 7466. [Google Scholar] [CrossRef] [PubMed]
  231. Fan, X.; Elkin, K.; Shi, Y.; Zhang, Z.; Cheng, Y.; Gu, J.; Liang, J.; Wang, C.; Ji, X. Schisandrin B improves cerebral ischemia and reduces reperfusion injury in rats through TLR4/NF-κB signaling pathway inhibition. Neurol. Res. 2020, 42, 693–702. [Google Scholar] [CrossRef] [PubMed]
  232. Gao, X.-J.; Xie, G.-N.; Liu, L.; Fu, Z.-J.; Zhang, Z.-W.; Teng, L.-Z. Sesamol attenuates oxidative stress, apoptosis and inflammation in focal cerebral ischemia/reperfusion injury. Exp. Ther. Med. 2017, 14, 841–847. [Google Scholar] [CrossRef] [Green Version]
  233. Zhang, S.; Jiang, L.; Che, F.; Lu, Y.; Xie, Z.; Wang, H. Arctigenin attenuates ischemic stroke via SIRT1-dependent inhibition of NLRP3 inflammasome. Biochem. Biophys. Res. Commun. 2017, 493, 821–826. [Google Scholar] [CrossRef]
  234. Thundyil, J.; Manzanero, S.; Pavlovski, D.; Cully, T.R.; Lok, K.-Z.; Widiapradja, A.; Chunduri, P.; Jo, D.-G.; Naruse, C.; Asano, M.; et al. Evidence That the EphA2 Receptor Exacerbates Ischemic Brain Injury. PLoS ONE 2013, 8, e53528. [Google Scholar] [CrossRef] [Green Version]
  235. Dvorakova, M.; Landa, P. Anti-inflammatory activity of natural stilbenoids: A review. Pharmacol. Res. 2017, 124, 126–145. [Google Scholar] [CrossRef]
  236. Zhang, H.; Zhao, W. Resveratrol Alleviates Ischemic Brain Injury by Inhibiting the Activation of Pro-Inflammatory Microglia Via the CD147/MMP-9 Pathway. J. Stroke Cerebrovasc. Dis. 2022, 31, 106307. [Google Scholar] [CrossRef]
  237. Chang, C.; Zhao, Y.; Song, G.; She, K. Resveratrol protects hippocampal neurons against cerebral ischemia-reperfusion injury via modulating JAK/ERK/STAT signaling pathway in rats. J. Neuroimmunol. 2018, 315, 9–14. [Google Scholar] [CrossRef]
  238. Lei, J.; Chen, Q. Resveratrol attenuates brain damage in permanent focal cerebral ischemia via activation of PI3K/Akt signaling pathway in rats. Neurol. Res. 2018, 40, 1014–1020. [Google Scholar] [CrossRef]
  239. Hou, Y.; Wang, K.; Wan, W.; Cheng, Y.; Pu, X.; Ye, X. Resveratrol provides neuroprotection by regulating the JAK2/STAT3/PI3K/AKT/mTOR pathway after stroke in rats. Genes Dis. 2018, 5, 245–255. [Google Scholar] [CrossRef]
  240. Dou, Z.; Rong, X.; Zhao, E.; Zhang, L.; Lv, Y. Neuroprotection of Resveratrol against Focal Cerebral Ischemia/Reperfusion Injury in Mice through a Mechanism Targeting Gut-Brain Axis. Cell. Mol. Neurobiol. 2019, 39, 883–898. [Google Scholar] [CrossRef]
  241. Grewal, A.K.; Singh, N.; Singh, T.G. Effects of resveratrol postconditioning on cerebral ischemia in mice: Role of the sirtuin-1 pathway. Can. J. Physiol. Pharmacol. 2019, 97, 1094–1101. [Google Scholar] [CrossRef] [Green Version]
  242. Pineda-Ramírez, N.; Alquisiras-Burgos, I.; Ortiz-Plata, A.; Ruiz-Tachiquín, M.-E.; Espinoza-Rojo, M.; Aguilera, P. Resveratrol Activates Neuronal Autophagy through AMPK in the Ischemic Brain. Mol. Neurobiol. 2020, 57, 1055–1069. [Google Scholar] [CrossRef]
  243. Teertam, S.K.; Jha, S.; Babu, P.P. Up-regulation of Sirt1/miR-149-5p signaling may play a role in resveratrol induced protection against ischemia via p53 in rat brain. J. Clin. Neurosci. 2020, 72, 402–411. [Google Scholar] [CrossRef]
  244. Ma, S.; Fan, L.; Li, J.; Zhang, B.; Yan, Z. Resveratrol promoted the M2 polarization of microglia and reduced neuroinflammation after cerebral ischemia by inhibiting miR-155. Int. J. Neurosci. 2020, 130, 817–825. [Google Scholar] [CrossRef]
  245. Gao, J.; Wang, H.; Li, Y.; Li, W. Resveratrol attenuates cerebral ischaemia reperfusion injury via modulating mitochondrial dynamics homeostasis and activating AMPK-Mfn1 pathway. Int. J. Exp. Pathol. 2019, 100, 337–349. [Google Scholar] [CrossRef]
  246. Yu, P.; Wang, L.; Tang, F.; Guo, S.; Liao, H.; Fan, C.; Yang, Q. Resveratrol-mediated neurorestoration after cerebral ischemic injury—Sonic Hedgehog signaling pathway. Life Sci. 2021, 280, 119715. [Google Scholar] [CrossRef] [PubMed]
  247. Yan, W.; Ren, D.; Feng, X.; Huang, J.; Wang, D.; Li, T.; Zhang, D. Neuroprotective and Anti-Inflammatory Effect of Pterostilbene Against Cerebral Ischemia/Reperfusion Injury via Suppression of COX-2. Front. Pharmacol. 2021, 12, 770329. [Google Scholar] [CrossRef] [PubMed]
  248. Liu, J.; Xu, J.; Mi, Y.; Yang, Y.; Li, Q.; Zhou, D.; Wei, K.; Chen, G.; Li, N.; Hou, Y. Pterostilbene alleviates cerebral ischemia and reperfusion injury in rats by modulating microglial activation. Food Funct. 2020, 11, 5432–5445. [Google Scholar] [CrossRef] [PubMed]
  249. Liu, H.; Wu, X.; Luo, J.; Wang, X.; Guo, H.; Feng, D.; Zhao, L.; Bai, H.; Song, M.; Liu, X.; et al. Pterostilbene Attenuates Astrocytic Inflammation and Neuronal Oxidative Injury after Ischemia-Reperfusion by Inhibiting NF-κB Phosphorylation. Front. Immunol. 2019, 10, 2408. [Google Scholar] [CrossRef] [PubMed]
  250. Wang, K.; Zhang, W.; Liu, J.; Cui, Y.; Cui, J. Piceatannol protects against cerebral ischemia/reperfusion-induced apoptosis and oxidative stress via the Sirt1/FoxO1 signaling pathway. Mol. Med. Rep. 2020, 22, 5399–5411. [Google Scholar] [CrossRef]
  251. Wang, L.; Guo, Y.; Ye, J.; Pan, Z.; Hu, P.; Zhong, X.; Qiu, F.; Zhang, D.; Huang, Z. Protective Effect of Piceatannol against Cerebral Ischaemia–Reperfusion Injury Via Regulating Nrf2/HO-1 Pathway In Vivo and Vitro. Neurochem. Res. 2021, 46, 1869–1880. [Google Scholar] [CrossRef]
  252. Ruan, W.; Li, J.; Xu, Y.; Wang, Y.; Zhao, F.; Yang, X.; Jiang, H.; Zhang, L.; Saavedra, J.M.; Shi, L.; et al. MALAT1 Up-Regulator Polydatin Protects Brain Microvascular Integrity and Ameliorates Stroke through C/EBPβ/MALAT1/CREB/PGC-1α/PPARγ Pathway. Cell. Mol. Neurobiol. 2019, 39, 265–286. [Google Scholar] [CrossRef]
  253. Zhu, X.; Song, Z.; Zhang, S.; Nanda, A.; Li, G. CD147: A novel modulator of inflammatory and immune disorders. Curr. Med. Chem. 2014, 21, 2138–2145. [Google Scholar] [CrossRef]
  254. Shi, Y.; Yan, W.; Lin, Q.; Wang, W. Icariin influences cardiac remodeling following myocardial infarction by regulating the CD147/MMP-9 pathway. J. Int. Med. Res. 2018, 46, 2371–2385. [Google Scholar] [CrossRef] [Green Version]
  255. Jin, R.; Xiao, A.Y.; Chen, R.; Granger, D.N.; Li, G. Inhibition of CD147 (Cluster of Differentiation 147) Ameliorates Acute Ischemic Stroke in Mice by Reducing Thromboinflammation. Stroke 2017, 48, 3356–3365. [Google Scholar] [CrossRef]
  256. Khoury, N.; Xu, J.; Stegelmann, S.D.; Jackson, C.W.; Koronowski, K.B.; Dave, K.R.; Young, J.I.; Perez-Pinzon, M.A. Resveratrol Preconditioning Induces Genomic and Metabolic Adaptations within the Long-Term Window of Cerebral Ischemic Tolerance Leading to Bioenergetic Efficiency. Mol. Neurobiol. 2019, 56, 4549–4565. [Google Scholar] [CrossRef]
  257. Koronowski, K.B.; Khoury, N.; Saul, I.; Loris, Z.B.; Cohan, C.H.; Stradecki-Cohan, H.M.; Dave, K.R.; Young, J.I.; Perez-Pinzon, M.A. Neuronal SIRT1 (Silent Information Regulator 2 Homologue 1) Regulates Glycolysis and Mediates Resveratrol-Induced Ischemic Tolerance. Stroke 2017, 48, 3117–3125. [Google Scholar] [CrossRef]
  258. Zhang, T.; Chen, X.; Qu, Y.; Ding, Y. Curcumin Alleviates Oxygen-Glucose-Deprivation/Reperfusion-Induced Oxidative Damage by Regulating miR-1287-5p/LONP2 Axis in SH-SY5Y Cells. Anal. Cell. Pathol. 2021, 2021, 5548706. [Google Scholar] [CrossRef]
  259. Ran, Y.; Su, W.; Gao, F.; Ding, Z.; Yang, S.; Ye, L.; Chen, X.; Tian, G.; Xi, J.; Liu, Z. Curcumin Ameliorates White Matter Injury after Ischemic Stroke by Inhibiting Microglia/Macrophage Pyroptosis through NF-κB Suppression and NLRP3 Inflammasome Inhibition. Oxid. Med. Cell. Longev. 2021, 2021, 1552127. [Google Scholar] [CrossRef]
  260. Wang, C.; Yang, Y.-H.; Zhou, L.; Ding, X.-L.; Meng, Y.-C.; Han, K. Curcumin alleviates OGD/R-induced PC12 cell damage via repressing CCL3 and inactivating TLR4/MyD88/MAPK/NF-κB to suppress inflammation and apoptosis. J. Pharm. Pharmacol. 2020, 72, 1176–1185. [Google Scholar] [CrossRef]
  261. Wang, W.; Xu, J. Curcumin Attenuates Cerebral Ischemia-reperfusion Injury Through Regulating Mitophagy and Preserving Mitochondrial Function. Curr. Neurovasc. Res. 2020, 17, 113–122. [Google Scholar] [CrossRef]
  262. Mo, Y.; Yue, E.; Shi, N.; Liu, K. The protective effects of curcumin in cerebral ischemia and reperfusion injury through PKC-θ signaling. Cell Cycle 2021, 20, 550–560. [Google Scholar] [CrossRef] [PubMed]
  263. Wu, S.; Guo, T.; Qi, W.; Li, Y.; Gu, J.; Liu, C.; Sha, Y.; Yang, B.; Hu, S.; Zong, X. Curcumin ameliorates ischemic stroke injury in rats by protecting the integrity of the blood-brain barrier. Exp. Ther. Med. 2021, 22, 783. [Google Scholar] [CrossRef] [PubMed]
  264. 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]
  265. 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] [PubMed] [Green Version]
  266. Zhu, H.; Fan, Y.; Sun, H.; Chen, L.; Man, X. Curcumin inhibits endoplasmic reticulum stress induced by cerebral ischemia-reperfusion injury in rats. Exp. Ther. Med. 2017, 14, 4047–4052. [Google Scholar] [CrossRef] [Green Version]
  267. Huang, L.; Chen, C.; Zhang, X.; Li, X.; Chen, Z.; Yang, C.; Liang, X.; Zhu, G.; Xu, Z. Neuroprotective Effect of Curcumin against Cerebral Ischemia-Reperfusion Via Mediating Autophagy and Inflammation. J. Mol. Neurosci. 2018, 64, 129–139. [Google Scholar] [CrossRef]
  268. Xu, L.; Ding, L.; Su, Y.; Shao, R.; Liu, J.; Huang, Y. Neuroprotective effects of curcumin against rats with focal cerebral ischemia-reperfusion injury. Int. J. Mol. Med. 2019, 43, 1879–1887. [Google Scholar] [CrossRef]
  269. Xu, H.; Nie, B.; Liu, L.; Zhang, C.; Zhang, Z.; Xu, M.; Mei, Y. Curcumin Prevents Brain Damage and Cognitive Dysfunction During Ischemic-reperfusion through the Regulation of miR-7-5p. Curr. Neurovasc. Res. 2019, 16, 441–454. [Google Scholar] [CrossRef]
  270. Hou, Y.; Wang, J.; Feng, J. The neuroprotective effects of curcumin are associated with the regulation of the reciprocal function between autophagy and HIF-1 alpha in cerebral ischemia-reperfusion injury. Drug Des. Dev. Ther. 2019, 13, 1135–1144. [Google Scholar] [CrossRef] [Green Version]
  271. Xie, C.-J.; Gu, A.-P.; Cai, J.; Wu, Y.; Chen, R.-C. Curcumin protects neural cells against ischemic injury in N2a cells and mouse brain with ischemic stroke. Brain Behav. 2018, 8, e00921. [Google Scholar] [CrossRef] [Green Version]
  272. Wang, Y.; Zhang, Y.; Yang, L.; Yuan, J.; Jia, J.; Yang, S. SOD2 Mediates Curcumin-Induced Protection against Oxygen-Glucose Deprivation/Reoxygenation Injury in HT22 Cells. Evid.-Based Complement. Altern. Med. 2019, 2019, 2160642. [Google Scholar] [CrossRef]
  273. Xia, M.; Ye, Z.; Shi, Y.; Zhou, L.; Hua, Y. Curcumin improves diabetes mellitus-associated cerebral infarction by increasing the expression of GLUT1 and GLUT3. Mol. Med. Rep. 2018, 17, 1963–1969. [Google Scholar] [CrossRef] [Green Version]
  274. Subedi, L.; Gaire, B.P. Neuroprotective Effects of Curcumin in Cerebral Ischemia: Cellular and Molecular Mechanisms. ACS Chem. Neurosci. 2021, 12, 2562–2572. [Google Scholar] [CrossRef]
  275. Henke, N.; Albrecht, P.; Bouchachia, I.; Ryazantseva, M.; Knoll, K.; Lewerenz, J.; Kaznacheyeva, E.; Maher, P.; Methner, A. The plasma membrane channel ORAI1 mediates detrimental calcium influx caused by endogenous oxidative stress. Cell Death Dis. 2013, 4, e470. [Google Scholar] [CrossRef] [Green Version]
  276. Dima, C.; Assadpour, E.; Dima, S.; Jafari, S.M. Nutraceutical nanodelivery; an insight into the bioaccessibility/bioavailability of different bioactive compounds loaded within nanocarriers. Crit. Rev. Food Sci. Nutr. 2021, 61, 3031–3065. [Google Scholar] [CrossRef]
  277. Zhang, L.; McClements, D.J.; Wei, Z.; Wang, G.; Liu, X.; Liu, F. Delivery of synergistic polyphenol combinations using biopolymer-based systems: Advances in physicochemical properties, stability and bioavailability. Crit. Rev. Food Sci. Nutr. 2020, 60, 2083–2097. [Google Scholar] [CrossRef]
  278. Zhao, D.; Simon, J.E.; Wu, Q. A critical review on grape polyphenols for neuroprotection: Strategies to enhance bioefficacy. Crit. Rev. Food Sci. Nutr. 2020, 60, 597–625. [Google Scholar] [CrossRef]
  279. Mrvová, N.; Škandík, M.; Kuniaková, M.; Račková, L. Modulation of BV-2 microglia functions by novel quercetin pivaloyl ester. Neurochem. Int. 2015, 90, 246–254. [Google Scholar] [CrossRef]
  280. Škandík, M.; Mrvová, N.; Bezek, Š.; Račková, L. Semisynthetic quercetin-quinone mitigates BV-2 microglia activation through modulation of Nrf2 pathway. Free Radic. Biol. Med. 2020, 152, 18–32. [Google Scholar] [CrossRef]
  281. Zhang, R.; Zhao, T.; Zheng, B.; Zhang, Y.; Li, X.; Zhang, F.; Cen, J.; Duan, S. Curcumin Derivative Cur20 Attenuated Cerebral Ischemic Injury by Antioxidant Effect and HIF-1α/VEGF/TFEB-Activated Angiogenesis. Front. Pharmacol. 2021, 12, 648107. [Google Scholar] [CrossRef]
  282. Yan, G.; Wang, Y.; Han, X.; Zhang, Q.; Xie, H.; Chen, J.; Ji, D.; Mao, C.; Lu, T. A Modern Technology Applied in Traditional Chinese Medicine: Progress and Future of the Nanotechnology in TCM. Dose-Response 2019, 17, 1559325819872854. [Google Scholar] [CrossRef] [PubMed]
  283. Ghosh, A.; Sarkar, S.; Mandal, A.K.; Das, N. Neuroprotective role of nanoencapsulated quercetin in combating ischemia-reperfusion induced neuronal damage in young and aged rats. PLoS ONE 2013, 8, e57735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  284. Ghosh, S.; Sarkar, S.; Choudhury, S.T.; Ghosh, T.; Das, N. Triphenyl phosphonium coated nano-quercetin for oral delivery: Neuroprotective effects in attenuating age related global moderate cerebral ischemia reperfusion injury in rats. Nanomed. Nanotechnol. Biol. Med. 2017, 13, 2439–2450. [Google Scholar] [CrossRef] [PubMed]
  285. Wang, Y.; Zhao, Q.; Han, N.; Bai, L.; Li, J.; Liu, J.; Che, E.; Hu, L.; Zhang, Q.; Jiang, T.; et al. Mesoporous silica nanoparticles in drug delivery and biomedical applications. Nanomed. Nanotechnol. Biol. Med. 2015, 11, 313–327. [Google Scholar] [CrossRef]
  286. Shen, Y.; Cao, B.; Snyder, N.R.; Woeppel, K.M.; Eles, J.R.; Cui, X.T. ROS responsive resveratrol delivery from LDLR peptide conjugated PLA-coated mesoporous silica nanoparticles across the blood–brain barrier. J. Nanobiotechnol. 2018, 16, 13. [Google Scholar] [CrossRef] [Green Version]
  287. Kakkar, V.; Muppu, S.K.; Chopra, K.; Kaur, I.P. Curcumin loaded solid lipid nanoparticles: An efficient formulation approach for cerebral ischemic reperfusion injury in rats. Eur. J. Pharm. Biopharm. 2013, 85, 339–345. [Google Scholar] [CrossRef]
  288. Kinfe, T.; Stadlbauer, A.; Winder, K.; Hurlemann, R.; Buchfelder, M. Incisionless MR-guided focused ultrasound: Technical considerations and current therapeutic approaches in psychiatric disorders. Expert Rev. Neurother. 2020, 20, 687–696. [Google Scholar] [CrossRef]
  289. McDannold, N.; Zhang, Y.; Supko, J.G.; Power, C.; Sun, T.; Peng, C.; Vykhodtseva, N.; Golby, A.J.; Reardon, D.A. Acoustic feedback enables safe and reliable carboplatin delivery across the blood-brain barrier with a clinical focused ultrasound system and improves survival in a rat glioma model. Theranostics 2019, 9, 6284–6299. [Google Scholar] [CrossRef]
  290. Yan, Y.; Chen, Y.; Liu, Z.; Cai, F.; Niu, W.; Song, L.; Liang, H.; Su, Z.; Yu, B.; Yan, F. Brain Delivery of Curcumin through Low-Intensity Ultrasound-Induced Blood-Brain Barrier Opening via Lipid-PLGA Nanobubbles. Int. J. Nanomed. 2021, 16, 7433–7447. [Google Scholar] [CrossRef]
  291. Neuhaus, A.; Couch, Y.; Hadley, G.; Buchan, A.M. Neuroprotection in stroke: The importance of collaboration and reproducibility. Brain 2017, 140, 2079–2092. [Google Scholar] [CrossRef] [Green Version]
  292. Kalra, P.; Khan, H.; Kaur, A.; Singh, T.G. Mechanistic Insight on Autophagy Modulated Molecular Pathways in Cerebral Ischemic Injury: From Preclinical to Clinical Perspective. Neurochem. Res. 2022, 47, 825–843. [Google Scholar] [CrossRef]
  293. Moretti, A.; Ferrari, F.; Villa, R.F. Neuroprotection for ischaemic stroke: Current status and challenges. Pharmacol. Ther. 2015, 146, 23–34. [Google Scholar] [CrossRef]
Figure 1. Ischemic cascade response in the acute phase after stroke. Once blood flow is disrupted, neuronal cells and astrocytes release large amounts of glutamate, resulting in excitotoxicity (1) and prompting cell death. The dead cells release damage-associated molecular patterns (DAMPs), which further stimulate microglia polarization, astrocyte activation, and release of pro-inflammatory factors (2). Blood-derived neutrophils and macrophages migrate to the injured area (3), further amplifying the ischemic cascade response. MMPs, matrix metalloproteinases; LFA-1, lymphocyte function-associated antigen-1; ICAM-1, intercellular adhesion molecule-1; BBB, blood-brain barrier; M1, M1 phenotype microglia; M2, M2 phenotype microglia.
Figure 1. Ischemic cascade response in the acute phase after stroke. Once blood flow is disrupted, neuronal cells and astrocytes release large amounts of glutamate, resulting in excitotoxicity (1) and prompting cell death. The dead cells release damage-associated molecular patterns (DAMPs), which further stimulate microglia polarization, astrocyte activation, and release of pro-inflammatory factors (2). Blood-derived neutrophils and macrophages migrate to the injured area (3), further amplifying the ischemic cascade response. MMPs, matrix metalloproteinases; LFA-1, lymphocyte function-associated antigen-1; ICAM-1, intercellular adhesion molecule-1; BBB, blood-brain barrier; M1, M1 phenotype microglia; M2, M2 phenotype microglia.
Molecules 27 04181 g001
Figure 2. Selected molecular mechanisms involved in the pathophysiological processes of stroke. Excitotoxicity: the impaired energy supply depolarizes presynaptic neurons, and massive glutamate release promotes inward calcium flow. Oxidative stress: the mitochondrial oxidative respiratory chain is inhibited, generating excess ROS/RNS and triggering a cascade of downstream responses. Inflammatory response: neuronal immune cells respond to external stimuli via receptors such as TLRs and IL-4R, mediating the synthesis and secretion of a series of inflammation-related proteins. Apoptosis: apoptosis-related proteins are activated via different pathways and ultimately drive apoptosis. Autophagy: regulated by both mTOR and AMPK proteins, the ULK1 kinase complex is activated, promotes autophagosome maturation, and completes the autophagic process step by step.
Figure 2. Selected molecular mechanisms involved in the pathophysiological processes of stroke. Excitotoxicity: the impaired energy supply depolarizes presynaptic neurons, and massive glutamate release promotes inward calcium flow. Oxidative stress: the mitochondrial oxidative respiratory chain is inhibited, generating excess ROS/RNS and triggering a cascade of downstream responses. Inflammatory response: neuronal immune cells respond to external stimuli via receptors such as TLRs and IL-4R, mediating the synthesis and secretion of a series of inflammation-related proteins. Apoptosis: apoptosis-related proteins are activated via different pathways and ultimately drive apoptosis. Autophagy: regulated by both mTOR and AMPK proteins, the ULK1 kinase complex is activated, promotes autophagosome maturation, and completes the autophagic process step by step.
Molecules 27 04181 g002
Table 1. Summary of studies on the anti-IS effects of flavonoids.
Table 1. Summary of studies on the anti-IS effects of flavonoids.
Polyphenolic CompoundChemical StructureModels and TreatmentsObserved EffectsMechanismsReference
In VitroIn Vivo
Flavonols
Quercetin Molecules 27 04181 i001(Not available) NAICR mice
Transient-middle cerebral artery occlusion (T-MCAO)
intraperitoneal injection (i.p.) for 7 days before T-MCAO
100, 150, 200 mg/kg
Improved behavioral functions
Reduced BBB permeability
↑mRNA MC4R[106]
NAGerbils
T-MCAO
Intragastric injection (i.g.) for 15 days before T-MCAO
20 mg/kg
Anti-oxidative stress↑SOD1
↑SOD2
↑CAT
↑GPx
[107]
NAWistar rats
T-MCAO
i.p. immediately after reperfusion
10, 30, 50 mg/kg
Anti-oxidative stress
Reduced BBB permeability
↑Sirt1/Nrf2/HO-1[108]
HT22 cell line and primary cortical neurons
Glutamate stimulated
1, 3, 5 µM
Sprague Dawley (SD) rats
permanent-MCAO (P-MCAO)
i.p. 0.5 h before P-MCAO
10 mg/kg
Suppressed glutamate-induced oxidative stress↑Thioredoxin/ASK-1
↓Caspase-3
[109]
HT22 cell line
Glutamate stimulated
(1, 3, 5 µM)
SD rats
P-MCAO
i.p. 0.5 h before P-MCAO
10 mg/kg
Alleviated intracellular calcium overload
Anti-apoptosis
↑Hippocalcin
↓Caspase-3
↓Bax
↑Bcl-2
[110]
Hippocampal slices and
neuron/glia co-culture
Oxygen-glucose deprivation/reoxygenation (OGD/R)
0, 10 μM
SD rats
T-MCAO
i.p. 21 days before T-MCAO
25 mg/kg
Alleviated neurological deficits, brain infarction, and BBB disruption↑p-ERK
↑p-Akt
↓Protein phosphatase
[111]
NASD rats
P-MCAO
i.p. 0.5 h before P-MCAO
10 mg/kg
Anti-apoptosis
Increased neuronal activity
↑Parvalbumin[112]
NASD rats
P-MCAO
i.p. 1 h before P-MCAO
30 mg/kg
Anti-apoptosis↓PARP
↓Caspase-3
[113]
NASD rats
P-MCAO
i.p. 0.5 h before P-MCAO
10 mg/kg
Downregulated glutamate toxicity↑PP2A subunit B[114]
BV2 cells
OGD/R
20, 40 μM
ICR mice
Hypoxic-ischemic brain injury
i.p. for 2 days after injury
50 mg/kg
Anti-inflammatory
Mitigated cognitive and motor function deficits
↓TLR4/MyD88/NF-κB[115]
NASD rats
P-MCAO
i.p. 1 h before p-MCAO
50 mg/kg
Improved energy metabolism↑γ-Enolase[116]
Isoquercetin Molecules 27 04181 i002Primary hippocampal neurons
OGD/R
20, 40, 80 μg/mL
SD rats
T-MCAO
i.g. for 3 days after T-MCAO
5, 10, 20 mg/kg
Reduced infarct size
Anti-apoptosis
↓TLR4-NF-κB
↓p-JNK1/2
↓p-p38 MAPK
[117]
NASD rats
T-MCAO
i.g. for 3 days before T-MCAO
5, 10, 20 mg/kg
Attenuated oxidative stress
Anti-apoptosis
↑Nrf2
↓NOX4/ROS/NF-κB
[118]
Rutin Molecules 27 04181 i003NAOvariectomized (OVX) SD rats
T-MCAO
i.p. for 5 days before T-MCAO
100 mg/kg
Decreased infarct size Attenuated neuron loss
Improved sensorimotor performance and recognition memory
↑BDNF-TrκB
↑NGF-TrkA
[119]
Kaempferol Molecules 27 04181 i004NASD rats
T-MCAO
i.g. for 7 days after T-MCAO
25, 50, 100 mg/kg
Anti-inflammatory
Attenuated BBB dysfunction
↓p-p65
↓MMP3
[120]
NASD rats
T-MCAO
i.g. for 7 days before T-MCAO
0.5, 1, 2 mg/kg
Reduced infarct volume
Anti-apoptosis
↑p-Akt
↑Nrf-2
↓p-NF-κB
[121]
PC12 cell line
OGD/R
5, 10, 20 μM
NAAmeliorated OGD-induced mitochondrial dysfunction↑Sirt1
↓p66shc
[122]
Primary cortical neurons
OGD
10 μM
C57BL/6 mice
T-MCAO
i.g. for 7 days before T-MCAO
50, 100, 200 mg/kg
Prevented HK-II detachment from mitochondria
Ameliorated mitochondrial dysfunction
↑p-Akt
↓Drp1
[123]
Primary cortical neurons
OGD/R
10 μM
NADecreased neuronal ferroptosis↑Nrf2/SLC7A11/GPx4[124]
Icariin Molecules 27 04181 i005NASD rats
T-MCAO
i.g. for 28 days after T-MCAO
60 mg/kg
Promoted angiogenesis and neurogenesis↑PI3K/ERK1/2
↑VEGF
↑BDNF
[125]
Primary microglia
OGD/R
0.37, 0.74, 1.48 μM
NADecreased ER stress
Anti-inflammatory
↓IRE1/XBP1s[126]
NASD rats
T-MCAO
i.g. for 28 days before T-MCAO
60 mg/kg
Promoted mild hypothermia-induced neuroprotection↑PPARs/Nrf2
↓JAK2/STAT3/NF-κB
[127]
Myricetin Molecules 27 04181 i006NASD rats
P-MCAO
i.g. for 7 days before T-MCAO
25 mg/kg
Reduced infarct volume
Anti-apoptosis
↓p-p38 MAPK
↓p-NF-κB
↑p-Akt
[128]
Human brain microvessel endothelial cells (HBMECs)
OGD/R
10, 30, 60 μM
NADecreased enhancement of endothelial permeability
Anti-inflammatory
↑eNOS
↑p-Akt
[129]
Isoflavones
Puerarin Molecules 27 04181 i007NASD rats
T-MCAO
i.p. 2 h before T-MCAO
50, 100 mg/kg
Alleviated neurological deficits
Anti-apoptosis
↑p-Akt1/p-GSK-3β/MCL-1[130]
NASD rats
T-MCAO
i.g. for 14 days before T-MCAO
50, 100 mg/kg
Suppressed excessive autophagy↓AMPK
↓ps317-ULK1
↑mTOR
↑ps757-ULK1
[131]
Genistein Molecules 27 04181 i008N9/HT22 co-culture
Primary microglia/primary cortical neuron co-culture
OGD/R
5 μg/mL
C57BL/6J mice
T-MCAO
i.p. for 14 days before T-MCAO
10 mg/kg
Anti-inflammatory
Anti-apoptosis
↓NLRP3[132]
NAOVX SD rats
T-MCAO
i.p. for 14 days before T-MCAO
10 mg/kg
Anti-oxidative stress
Anti-apoptosis
↑Nrf2
↑NQO1
↓Caspase-3
[133]
NAOVX SD rats
T-MCAO
i.p. for 14 days before T-MCAO
10 mg/kg
Anti-apoptosis↑PI3K-Akt-mTOR[134]
Daidzein Molecules 27 04181 i009NAICR mice
T-MCAO
i.p. 10 min after T-MCAO
10, 20, 30 mg/kg
Alleviated neuron impairment
Anti-apoptosis
↑PI3K/Akt/mTOR
↑BDNF/CREB
[135]
Flavones
Baicalein Molecules 27 04181 i010NASD rats
T-MCAO
i.p. for 7 days after T-MCAO
200 mg/kg
Promoted M2
polarization of microglia
Suppressed excessive autophagy
↓MAPK/NF-κB
↑PI3K/Akt/mTOR
[136]
SH-SY5Y cell line
OGD/R
0.1–8 μM
SD rats
T-MCAO
intravenous injection (i.v.) just before refusion
2.5, 5, 10 mg/kg
Anti-oxidative stress
Anti-inflammatory
↑Nrf2
↓NF-κB
↓LOX-1
↓AMPK
[137]
BV2 cell line
LPS/IFNr stimulation or OGD/R
45 μM
C57BL/6J mice
T-MCAO
i.g. for 3 days after T-MCAO
100 mg/kg
Anti-inflammatory
Promoted M2
polarization of microglia
↓TLR4/NF-κB
↓p-STAT1
[138]
SH-SY5Y cell line
OGD/R
1, 5, 10, 15, 20 μM
SD rats
T-MCAO
i.g. for 7 days after T-MCAO
100 mg/kg
Anti-apoptosis
Reduced infarct volume
↓PARP-1
↓Nuclear translocation of MIF and AIF
[139]
PC12 cell line
OGD/R
0.02, 0.1, 0.5 µM
SD rats
T-MCAO
i.g. for 7 days after T-MCAO
100 mg/kg
Anti-oxidative stress
Anti-apoptosis
↓Calpain 1
↓Nuclear translocation of AIF
[140]
Baicalin Molecules 27 04181 i011Primary astrocytes
OGD/R
1, 10 μM
SD
T-MCAO
i.p. 0.5 h before refusion
50 mg/kg
Anti-excitotoxic↓SDH
↑GS
[141]
Scutellarin Molecules 27 04181 i012BV2 cell line
LPS stimulation
0.54 μM
SD rats
P-MCAO
i.p. 2 h before MCAO and 12, 24, 36, 48, 60 h after MCAO
100 mg/kg
Decreased microglial activation
Anti-inflammatory
↓p-p38 MAPK
↓p-JNK
[142]
NASD rats
P-MCAO
i.p. for 7 days before P-MCAO
20, 50, 100 mg/kg
Alleviated cognitive impairments
Anti-inflammatory
↓PARP-1/NF-κB[143]
Primary astrocytes
OGD/R
10, 50 μM
SD rats
T-MCAO
i.p. 2 h before MCAO and 12, 24, 36, 48, 60 h after MCAO
50, 100 mg/kg
Anti-oxidative stress↓NOX2
↑Cx43
[144]
NASD rats
T-MCAO
i.v. for 7 days after T-MCAO
0.33 mg/kg
Suppressed excessive autophagy↓LC3-II/LC3-I[145]
Luteolin Molecules 27 04181 i013NASD rats
T-MCAO
i.p. 0, 12 h after T-MCAO
20, 40, 80 mg/kg
Alleviated neurologic deficits and cerebral edema
Anti-inflammatory
↑Nrf2
↓PPARγ/NF-κB
[146]
NASD rats
T-MCAO
i.p. for 7 days after T-MCAO
15, 30, 60 mg/kg
Suppressed excessive autophagy
Ameliorated mitochondrial dysfunction
↑Sirt3/AMPK/mTOR[147]
Chrysin Molecules 27 04181 i014NAWistar rats
T-MCAO
i.g. for 21 days before T-MCAO
10, 30, 100 mg/kg
Prevented cognitive and hippocampal LTP impairments↓IL-1β
↓TNF-α
[148]
SH-SY5Y cell line
OGD/R
10, 20, 40, 80, 160, 320 μM
SD rats
T-MCAO
i.g. for 7 days after T-MCAO
10, 20 mg/kg
Anti-inflammatory
Anti-apoptosis
↑PI3K/Akt/mTOR[149]
NASD rats
T-MCAO
i.g. for 14 days before T-MCAO
30 mg/kg
Anti-inflammatory↓iNOS
↓TNF-α
[150]
Apigenin Molecules 27 04181 i015HBMEC
OGD/R
2.5, 5 μM
SD rats
T-MCAO
i.p. for 14 days after T-MCAO
25 mg/kg
Suppressed excessive autophagy
Promoted neovascularization
↑Caveolin-1
↑VEGF
[151]
PC12 cell line
1.2 mM CoCl2 stimulation
0–200 µg/mL
SD rats
T-MCAO
i.p. for 7 days after T-MCAO
25 mg/kg
Ameliorated mitochondrial dysfunction↓ROS[152]
Flavanols
(-)-Epigallocatechin gallate Molecules 27 04181 i016NASD rats
T-MCAO
i.p. immediately after T-MCAO
20 mg/kg
Anti-apoptosis↑PI3K/Akt/eNOS[153]
Neurosphere culture
10 ng/mL LPS stimulation
10, 20, 40 μM for 7 days
C57BL/6 mice
T-MCAO
Injection into left ventricle for 14 days starting at 14 days post injury
2 μg
Promoted the M2
polarization of microglia Promoted neurogenesis
↑PI3K/Akt[154]
NASD rats
P-MCAO
i.p. just before P-MCAO
50 mg/kg
Anti-apoptosis↓Caspase-3
↓PARP
[155]
HT22 cell line
Glutamate stimulation
10, 20, 40 μM
SD rats
P-MCAO
i.p. just before P-MCAO
50 mg/kg
Alleviated neurological deficits
Suppressed glutamate-induced oxidative stress
↑Thioredoxin/ASK-1[156]
(-)-Epicatechin gallate Molecules 27 04181 i017HBMEC
OGD/R
0.5, 1, 2, 4 μM
NAPromoted neovascularization
Alleviated apoptosis and autophagy
↑VEGF
↓LC3-I/II
↑mTOR
↑Bcl-2/Bax
[157]
Procyanidin Molecules 27 04181 i018BV2 cell line
OGD
10 μM
SD rats
T-MCAO
i.p. 1 h before P-MCAO
20, 40, 80 mg/kg
Ameliorated neurological deficits
Anti-inflammatory
↓TLR4-p38-NF-κB-NLRP3[158]
Flavanones
Naringenin Molecules 27 04181 i019Primary cortical neuron
OGD/R
20, 40, 80 μM
SD rats
T-MCAO
i.p. immediately before T-MCAO
20 mg/kg
Anti-oxidative stress↑Nrf2[159]
Naringin Molecules 27 04181 i020SH-SY5Y cell line
OGD/R
100 μM/200 μM
SD rats
T-MCAO
i.v. immediately before T-MCAO
80, 120, 160 mg/kg
Inhibited mitophagy↓ONOO
↓3-NT
↓LC3-II/LC3-I
[160]
Hesperetin Molecules 27 04181 i021BV2 cell line
OGD/R
100 μM
C57BL/6 mice
T-MCAO
i.p. for 7 days after T-MCAO
30 mg/kg
Anti-inflammatory
Promoted the M2
polarization of microglia
↓TLR4/NF-κB[161]
Ginkgetin aglycone-NAWistar rats
T-MCAO
i.p. for 5 days before T-MCAO
100, 200 mg/kg
Anti-oxidative stress
Anti-inflammatory
↓JAK2/STAT3/Sirt1[162]
Anthocyanins
Cyanidin-3-glucoside Molecules 27 04181 i022NAICR mice
T-MCAO
i.g. for 7 days before T-MCAO
100, 150, 200 mg/kg
Anti-inflammatory↓TLR4/NF-κB
↓NLRP3
[163]
HT22 cell line
Glutamate stimulation
0.05–1 μM
NAAnti-oxidative stress induced-ER stress↑ERK/Nrf2
↑Antioxidant enzyme
[164]
Petunidin-3-O-rutinoside (p-coumaroyl)-5-O-glucoside-NASD rats
T-MCAO
i.p. for 7 days before T-MCAO
50, 100, 200 mg/kg
Protected neurovascular unit↓NLRP3
↓NF-κB
↓MMP9
[165]
SH-SY5Y cell line
OGD/R
10, 100, 1000 μg/mL
NAEnhanced autophagic flux↓SQSTM1
↑LC3B II/LC3B I
[166]
Symbols: (↑) increase; (↓) decrease.
Table 2. Summary of recent studies on the anti-IS effects of phenolic acids.
Table 2. Summary of recent studies on the anti-IS effects of phenolic acids.
Polyphenolic CompoundChemical StructureModels and TreatmentsObserved EffectsMechanismsReference
In VitroIn Vivo
Cinnamic Acid Derivatives
Ferulic acid Molecules 27 04181 i023BMEC
OGD/R
100, 200, 300, 400 μM
NAIncreased punctate-mitochondria-dependent mitophagy↑LC3-II[192]
NASD rats
P-MCAO
i.v. immediately after P-MCAO
60, 80, 100 mg/kg
Anti-apoptosis↑Akt/mTOR/4E-BP1/Bcl-2[193]
NASD rats
P-MCAO
i.v. 0.5 h after P-MCAO
80, 100 mg/kg
Anti-apoptosis
Promoted autophagy
↑HSP70/Bcl-2
↑Beclin 1
[194]
Rosmarinic acid Molecules 27 04181 i024NACD-1 mice
T-MCAO
i.p. immediately after T-MCAO
10, 20, 40 mg/kg
Anti-oxidative stress
Anti-apoptosis
↑Akt/Nrf2/HO-1[195]
Chlorogenic acid Molecules 27 04181 i025NAWistar rats
T-MCAO
i.g. for 3 days before T-MCAO
15, 30, 60 mg/kg
Decreased mortality
Reduced infarct volume
↓ICAM-1
↓VCAM-1
↑EPO
↑HIF-1α
↑NGF
[196]
NASD rats
P-MCAO
i.p. 2 h after P-MCAO
30 mg/kg
Alleviated neurobehavioral symptoms
Anti-apoptosis
↓Caspase-3
↓Caspase-7
↓PARP
[197]
NAWistar rats
T-MCAO
i.p. 10 min before T-MCAO and 10 min after refusion
10 mg/kg
Anti-apoptosis↑miR-23b
↓TAB3/NF-κB/p53
[198]
NASD rats
T-MCAO
i.p. 7 days before T-MCAO
20, 100, 500 mg/kg
Anti-oxidative stress↑Nrf2/NQO-1/HO-1[199]
NASD rats
P-MCAO
i.p. 2 h after P-MCAO
30 mg/kg
Anti-inflammatory
Inhibited the activation of astrocytes and microglia
↓NF-κB[200]
Salvianolic acid A Molecules 27 04181 i026NASD rats
T-MCAO
i.p. immediately after T-MCAO
5, 10, 20 mg/kg
Protect BBB
Anti-inflammatory
↓MMP9
↓NF-κB p65
[201]
SH-SY5Y cell line
OGD/R
0.05, 0.5, 5 μM
SD rats
T-MCAO
i.v. immediately after T-MCAO
5, 10, 20 mg/kg
Improved neurological function
Anti-apoptosis
↑Akt/FOXO3a
↓BIM/Caspase-3
[202]
PC12 cell line
OGD/R
5 μM
SD rats
T-MCAO
i.p. 0, 6 h after T-MCAO
20 mg/kg
Anti-apoptosis
Anti-inflammatory
↑miR-499a
↑Wnt3a/β-catenin
↓DDK1
[203]
HBMEC
OGD
1, 3, 10 μM
SD rats
Autologous thrombus stroke model
i.g. for 5 days before stroke
10 mg/kg
Alleviated intracerebral hemorrhage
Suppressed vascular endothelial dysfunction
↓VEGFA-Src-VAV2-Rac1-PAK[204]
NASD rats
T-MCAO
i.p. 15 min before T-MCAO
5, 10 mg/kg
Anti-inflammatory↓TLRs/MyD88[205]
NASD rats
autologous thrombus stroke model
i.p. for 14 days after stroke
10 mg/kg
Promoted endogenous neurogenesis↑Wnt3a/β-catenin
↓GSK-3β
[206]
Salvianolic acid B Molecules 27 04181 i027NAWistar rats
T-MCAO
i.p. for 3 days before T-MCAO
3, 6, 12 mg/kg
Inhibited platelet activation
Anti-inflammatory
↓P-selection
↓CD40L
↓CD40/NF-κB
[207]
Primary astrocytes/primary cortical neurons
OGD/R
800 ng/mL
C57BL/6J mice
T-MCAO
i.p. immediately after refusion
12 mg/kg
Promoted glycogenolysis
Anti-oxidative stress
↑GP activity
↑NADPH
↑GSH
[208]
Benzoic acid derivatives
Protocatechuic acid Molecules 27 04181 i028NAWistar rats
T-MCAO
i.p. 1.5 h after T-MCAO
10, 30, 50 mg/kg
Anti-apoptosis
Inhibited microglial activation
↑CREB
↓Caspase-3
[209]
Gallic acid Molecules 27 04181 i029NAC57BL/6J mice
T-MCAO
i.p. 1, 12, 24, 48, 72 h after T-MCAO
50, 100, 200 mg/kg
Inhibited microglial M1 polarization
Protect BBB
↓MMP9[210]
Vanillic acid Molecules 27 04181 i030NAWistar rats
Transient bilateral common carotid artery occlusion and reperfusion (TBCCAO/R)
i.g. for 14 days before TBCCAO/R
100 mg/kg
Restored spatial memory
Anti-inflammatory
↓IL-6
↓TNF-α
↑IL-10
[211]
Rhein Molecules 27 04181 i031NASD rats
T-MCAO
i.g. for 3 days after T-MCAO
25, 50, 100 mg/kg
Anti-oxidative stress
Anti-apoptosis
↑SOD
↑Bcl-2
↓Bax
↓Caspase-3
↓Caspase-9
[212]
Symbols: (↑) increase; (↓) decrease.
Table 3. Summary of recent studies on anti-IS effects of lignans.
Table 3. Summary of recent studies on anti-IS effects of lignans.
LignanChemical StructureModels and TreatmentsObserved EffectsMechanismsReference
In VitroIn Vivo
Magnolol Molecules 27 04181 i032BMEC
OGD/R 1, 10 μM
Primary microglia
LPS stimulation
0.01, 0.1, 1, 10 μM
Kunming mice
T-MCAO
i.v. 0, 1, 2 h after T-MCAO
1.4, 7.0, 35.0 μg/kg
Protect BBB
Anti-inflammatory
↓p-EphA2
↑ZO-1
↑Occludin
↓TNF-α
[226]
NASD rats
T-MCAO
i.p. immediately after T-MCAO
25 mg/kg
Anti-inflammatory
Anti-apoptosis
↑Sirt1
↑Bcl-2
↓Bax
↓Ac-FOXO1
↓TNF-α
[227]
BV2/RAW264.7 cell line
LPS stimulation
0.1–50 μM
SD rats
T-MCAO
i.p. 30 min before T-MCAO/2 h after T-MCAO
0.01, 0.1, 1 mg/kg
Anti-oxidative stress
Anti-inflammatory
↓TNF-α
↓IL-6
↓NOX
[228]
NASD rats
T-MCAO
i.p. for 7 days before T-MCAO
75 mg/kg
Anti-apoptosis↑BDNF
↓Bax
[229]
Schisandrin A Molecules 27 04181 i033Neural progenitor cell line/primary cortical neurons
0.1, 1, 10 μM
C57/BL6 mice
photothrombosis model
i.p. 21 days after infarction
12 mg/kg
Promoted neural cell proliferation and differentiation↑Cdc42
↑GTPase
[230]
Schisandrin B Molecules 27 04181 i034NASD rats
T-MCAO
i.v. for 3 days before T-MCAO
15, 30 mg/kg
Anti-inflammatory↓TLR4/NF-κB[231]
Sesamol Molecules 27 04181 i035NASD rats
T-MCAO
i.p. for 7 days before T-MCAO
25 mg/kg
Anti-apoptosis↑Bcl-2
↓Bax
↓Caspase-3
[232]
Arctigenin Molecules 27 04181 i036Primary cortical neurons
OGD/R
100 ng/mL
SD rats
T-MCAO
i.p. for 3 days before T-MCAO
20 mg/kg
Anti-inflammatory↑Sirt1
↓NLRP3
[233]
Symbols: (↑) increase; (↓) decrease.
Table 4. Summary of recent studies on anti-IS effects of stilbenes.
Table 4. Summary of recent studies on anti-IS effects of stilbenes.
StilbeneChemical StructureModels and TreatmentsObserved EffectsMechanismsReference
In VitroIn Vivo
Resveratrol Molecules 27 04181 i037Primary microglia
OGD/R
5 μM
C57BL/6 mice
T-MCAO
i.g. for 3 days after T-MCAO
200 mg/kg
Inhibited pro-inflammatory microglia activation↓CD147/MMP9[236]
NASD rats
T-MCAO
i.p. for 7 days after T-MCAO
20 mg/kg
Alleviated cognitive impairment
Reduced neuronal loss
↓JAK/ERK/STAT[237]
NASD rats
P-MCAO
i.p. 2/12 h after P-MCAO
100 mg/kg
Reduced neurological deficits and cerebral water content↑PI3K/Akt[238]
NASD rats
T-MCAO
i.p. for 7 days before T-MCAO
30 mg/kg
Improved neurological function
Reduced infarct size
↑p-JAK2/p-STAT3
↑PI3K/Akt/mTOR
[239]
NAC57BL/6 mice
T-MCAO
i.p. for 3 days after T-MCAO
200 mg/kg
Modulated the gut-brain axis↑IL-4/Th2
↑IL-10/Treg
[240]
NASwiss albino mice
T-MCAO
i.p. 5 min before refusion
30 mg/kg
Anti-oxidative stress
Reduced AChE activity
↑Sirt1[241]
Primary cortical neurons
Glutamate stimulated excitotoxicity
0.004–400 μM
Wistar rats
T-MCAO
i.v. immediately after T-MCAO
1.8 mg/kg
Promoted autophagy↑p-AMPK
↑Beclin 1
[242]
NASD rats
T-MCAO
i.p. 30 min before refusion
20 mg/kg
Anti-apoptosis↑Sirt1/miR-149–5p/p53[243]
NAC57/BL mice
T-MCAO
i.p. 0, 8, 18 h after T-MCAO
100 mg/kg
Promoted the M2
polarization of microglia
↓miR-155[244]
N2a cell line
OGD/R
10, 20 μM
NAModulated mitochondrial homeostasis↑p-AMPK-Mfn1[245]
Primary cortical neurons
OGD/R
5 μM
SD rats
T-MCAO
i.p. for 7 days before T-MCAO
30 mg/kg
Promoted nerve regeneration↑Sonic hedgehog[246]
Pterostilbene Molecules 27 04181 i038NAWistar rats
T-MCAO
i.g. for 30 days before T-MCAO
35 mg/kg
Inhibited inflammatory cell infiltration
Anti-inflammatory
↓COX-2
↓PGE2
↓NF-κB
[247]
BV2 cell line
LPS stimulation
1, 10 μM
SD rats
T-MCAO
i.p. immediately after refusion
7, 14, 28 mg/kg
Anti-oxidative stress
Anti-inflammatory
↓NADPH
↓NF-κB
[248]
HT22/U251 co-culture
OGD/R
2.5, 5 μM
SD rats
T-MCAO
i.p. 1 h after T-MCAO
7, 14, 28 mg/kg
Anti-inflammatory↓NF-κB[249]
Piceatannol Molecules 27 04181 i039NAC57BL/6 mice
T-MCAO
i.g. for 6 days after T-MCAO
10,20 mg/kg
Anti-oxidative stress
Anti-apoptosis
↑Sirt1/FoxO1[250]
PC12 cell line
OGD/R
2.5, 10, 40, 160 μM
ICR mice
T-MCAO
i.p. Immediately after refusion
5, 10, 20 mg/kg
Anti-oxidative stress↑Nrf2/HO-1[251]
Polydatin Molecules 27 04181 i040HUVEC/BMEC
OGD
20 μM
SD rats
T-MCAO
i.v. 10 min before T-MCAO
30 mg/kg
Protected cerebrovascular endothelial cells and enhanced BBB integrity↑C/EBPβ/MALAT1/CREB/PGC-1α/PPARγ[252]
Symbols: (↑) increase; (↓) decrease.
Table 5. Summary of recent studies on anti-IS effects of curcumin.
Table 5. Summary of recent studies on anti-IS effects of curcumin.
Chemical StructureModels and TreatmentsObserved EffectsMechanismsReference
In VitroIn Vivo
Curcumin Molecules 27 04181 i041SH-SY5Y cell line
OGD/R
25 μM
NAAnti-oxidative stress↑miR-1287–5p

↓LONP2
[258]
Primary microglia
LPS stimulation
12.5 μM
C57BL/6J mice
T-MCAO
i.p. for 7 days after T-MCAO
150 mg/kg
Inhibited pyroptosis↓NLRP3
↓GSDMD-N
[259]
PC12 cell line
OGD/R
20 μM
NAAnti-inflammatory↓CCL3
↓TLR4/MyD88/MAPK/NF-κB
[260]
Primary cortical neurons
OGD/R
5 μM
SD rats
T-MCAO
i.p. immediately after refusion
100 mg/kg
Preserved mitochondrial function
Enhanced mitophagy
↑LC3-II/LC3-I[261]
PC12 cell line
OGD/R
20 μM
SD rats
T-MCAO
i.p. 24 and 1 h before T-MCAO
100 mg/kg
Decreased intracellular calcium ion concentration
BBB protection
↑PKC-θ
↓ORai1
[262]
NASD rats
T-MCAO
i.p. 30 min before T-MCAO
300 mg/kg
Anti-inflammatory
BBB protection
↓NF-κB
↓MMP9
[263]
BV2 cell line
LPS stimulation
12.5, 25 μM
C57BL/6 mice
P-MCAO
i.p. 0 and 24 h after T-MCAO
150 mg/kg
Promoted microglial M2
polarization
↓TNF-α
↓iNOS
↑CD206
[264]
NAWistar rats
T-MCAO
i.p. 30 min before T-MCAO
300 mg/kg
Reduced neurological dysfunction
Anti-inflammatory
Anti-apoptosis
↓ICAM-1
↓MMP9
↓Caspase-3
↓NF-κB
[265]
NAWistar rats
T-MCAO
i.p. 2 h before T-MCAO
150 mg/kg
Inhibited ER stress↓GADD153
↓Caspase-12
[266]
NASD rats
T-MCAO
i.p. 30 min after T-MCAO
200 mg/kg
Anti-inflammatory
Suppressed excessive autophagy
↑PI3K/Akt/mTOR
↓TLR4/p38 MAPK
[267]
Primary astrocytes
OGD/R
5, 10, 20 μM
SD rats
T-MCAO
i.p. 30 min before T-MCAO
300 mg/kg
Decreased infarct size
Anti-apoptosis
↑MEK/ERK/CREB[268]
PC12 cell line
OGD/R
10, 20 μM
SD rats
T-MCAO
i.g. for 3 days after T-MCAO
100, 300 mg/kg
Anti-oxidative stress
Anti-inflammatory
Anti-apoptosis
↑miR-7–5p
↓RelA p65
[269]
PC12 cell line
OGD/R
5 μM
NASuppressed excessive autophagy↓HIF-1α
↓LC3-II
↓p62
[270]
N2A cell line
OGD/R
5, 15, 25, 35 μM
C57BL/6 mice
T-MCAO
i.p. 1 h after T-MCAO
100, 200, 300, 400 mg/kg
Anti-apoptosis
Ameliorated mitochondrial dysfunction
↑Bcl-2
↓Bax
↓Caspase-3
[271]
HT22 cell line
OGD/R
100 ng/mL
NAAnti-oxidative stress↑SOD2[272]
NADiabetic SD rats
T-MCAO
i.g. after T-MCAO
40 mg/kg
Decreased infarct size
Regulated glucose uptake
Anti-apoptosis
↑GLUT1
↑GLUT3
[273]
Symbols: (↑) increase; (↓) decrease.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, S.; Lin, F.; Wang, J.; Pan, X.; Sun, L.; Wu, W. Polyphenols for the Treatment of Ischemic Stroke: New Applications and Insights. Molecules 2022, 27, 4181. https://doi.org/10.3390/molecules27134181

AMA Style

Liu S, Lin F, Wang J, Pan X, Sun L, Wu W. Polyphenols for the Treatment of Ischemic Stroke: New Applications and Insights. Molecules. 2022; 27(13):4181. https://doi.org/10.3390/molecules27134181

Chicago/Turabian Style

Liu, Shuhan, Feng Lin, Jian Wang, Xiaoqiang Pan, Liguang Sun, and Wei Wu. 2022. "Polyphenols for the Treatment of Ischemic Stroke: New Applications and Insights" Molecules 27, no. 13: 4181. https://doi.org/10.3390/molecules27134181

Article Metrics

Back to TopTop