Next Article in Journal
ROS-Induced DNA-Damage and Autophagy in Oral Squamous Cell Carcinoma by Usnea barbata Oil Extract—An In Vitro Study
Next Article in Special Issue
Anti-Inflammatory, Antimicrobial, Antioxidant and Photoprotective Investigation of Red Propolis Extract as Sunscreen Formulation in Polawax Cream
Previous Article in Journal
BMP/Smad Pathway Is Involved in Lithium Carbonate-Induced Neural-Tube Defects in Mice and Neural Stem Cells
Previous Article in Special Issue
Effects of Luteolin on Biofilm of Trueperella pyogenes and Its Therapeutic Effect on Rat Endometritis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Neuroprotective Potentiality of Flavonoids on Alzheimer’s Disease

by
Antonella Calderaro
1,
Giuseppe Tancredi Patanè
1,
Ester Tellone
1,*,
Davide Barreca
1,*,
Silvana Ficarra
1,
Francesco Misiti
2 and
Giuseppina Laganà
1
1
Department of Chemical, Biological, Pharmaceutical and Environmental Sciences, University of Messina, Viale Ferdinando Stagno d’Alcontres 31, 98166 Messina, Italy
2
Department of Human Sciences, Society and Health, University of Cassino and Southern Lazio, V. S. Angelo, Loc. Folcara, 3043 Cassino, Italy
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(23), 14835; https://doi.org/10.3390/ijms232314835
Submission received: 17 October 2022 / Revised: 19 November 2022 / Accepted: 24 November 2022 / Published: 27 November 2022
(This article belongs to the Special Issue Flavonoids and Their Impact on Human Health)

Abstract

:
Alzheimer’s disease (AD), due to its spread, has become a global health priority, and is characterized by senile dementia and progressive disability. The main cause of AD and other neurodegenerations (Huntington, Parkinson, Amyotrophic Lateral Sclerosis) are aggregated protein accumulation and oxidative damage. Recent research on secondary metabolites of plants such as polyphenols demonstrated that they may slow the progression of AD. The flavonoids’ mechanism of action in AD involved the inhibition of acetylcholinesterase, butyrylcholinesterase, Tau protein aggregation, β-secretase, oxidative stress, inflammation, and apoptosis through modulation of signaling pathways which are implicated in cognitive and neuroprotective functions, such as ERK, PI3-kinase/Akt, NFKB, MAPKs, and endogenous antioxidant enzymatic systems. This review focuses on flavonoids and their role in AD, in terms of therapeutic potentiality for human health, antioxidant potential, and specific AD molecular targets.

1. Introduction

Alzheimer’s disease (AD) is the most common cause of senile dementia associated with progressive disability. The inherited disease, in an autosomal dominant way, generally leads to a lethal outcome after about 5–10 years from the onset of the first symptoms [1]. Its pathology is complex, characterized by a decline in memory that, in its most common form, arises after 60 years of age. More rarely, the symptomatology can begin between 40 and 50 years, and in this case, the disease has a very rapid progression [2]. Generally, the early-onset form, called “familial”, is related to specific mutations in the genes encoding presenilin1 (PS1) and 2 (PS2) and amyloid precursor protein (APP), while sporadic late-onset disease is associated with mutations in the gene encoding apolipoprotein E (ApoE), and includes several environmental risk factors. [3]. The exact etiology of AD is not yet known, but several mechanisms have been described, including cholinesterase deficiency and generation of oxidative stress [4,5]. The histological features of AD are the extracellular deposits of the amyloid beta peptide (Aβ), in the form of neuritic plaques. The intraneuronal neurofibrillary tangles (NFTs) constitute aggregates of hyperphosphorylated Tau protein [6]. The symptomatology is characterized by an initial difficulty with language, concentration, and orientation that evolves into motor difficulties and personality changes, leading to a serious impact on public health and a strong burden on the field of health [7]. To date, approximately 50 million cases of AD have been estimated in the world (this number will more than double by 2050) and it is predicted to double every 5 years. Several drugs have been selected to combat the disease, but unfortunately, due to the different natures of the pathological targets related to the progression of AD, none of these modify the disease; they confer only milder management and transient symptoms. Further studies are needed to better characterize the risk factors that predispose one to the progression of AD, and to identify drugs to counteract the evolution of the disease and/or defend against its development. In this context, diet and natural products are show great promise in helping to reduce the development of neurodegenerative diseases [8,9,10,11,12,13,14,15,16]. In fact, unlike synthetic products, which possess serious side effects, the use of natural compounds can be a good alternative therapy. The purpose of this review is to describe the state of the art of current knowledge on AD and on the biochemical targets that trigger its pathological progression, as well as to highlight the potential protective association of flavonoids for the purpose of reversing the age-related decline caused by AD.

2. Oxidative Stress and Alzheimer

Oxidative stress is defined as an imbalance between oxidants and antioxidants that causes a rise in oxidant levels [17]. According to the amyloid cascade hypothesis, accumulation of non-soluble amyloid β peptides in the Central Nervous System (CNS) is the primary cause that initiates a pathogenic cascade, leading to the complex multilayered pathology and clinical manifestation of the disease. It is, therefore, not surprising that the search for mechanisms underlying cognitive changes observed in AD has focused on the brain and Aβ-inducing oxidative stress. However, since Aβ depositions can be found in normal, non-demented elderly people and in many other pathological conditions, the amyloid cascade hypothesis was modified to claim that intraneuronal accumulation of soluble Aβ oligomers, rather than monomer or insoluble Aβ fibrils, is the first step of a fatal cascade in AD (Figure 1). Oxidative stress was initially proposed to be a major factor in AD in 1986 [18]. Overwhelming evidence exists that the cells in the Alzheimer’s brain undergo abnormally high levels of oxidative stress, and that those amyloid plaques are a focus of cellular and molecular oxidation [19]. Aβ peptides trigger oxidative stress in the brain [20,21]. In addition to mediating Aβ-induced cytotoxicity, numerous studies have suggested that oxidative stress promotes the production of Aβ. It has been demonstrated that defects in the antioxidant defense system caused elevated oxidative stress [22,23]. Previous studies have shown that oxidative stress decreases the activity of alpha-secretase while promoting the activation of a cascade of redox-sensitive cell signal pathways, including JNK, which promotes the expression of BACE1 and PS1, and eventually β- and γ-secretase, enzymes critical for the generation of Aβ from APP [24,25]. Notably, the oxidative damage appears to become pronounced following the interaction of the sulfur-free radical with methionine 35 in the Aβ peptide [26]. The brain in AD appears to sustain more oxidative damage than normal, with low levels of antioxidants [27,28].
Therefore, while the brain membrane phospholipids are composed of polyunsaturated fatty acids, this organ is particularly vulnerable to free radical attacks. Plasma levels of thiobarbituric acid are high in the early stages of AD [29]; lipid hydroperoxides are the unstable products of lipid peroxidation, and they undergo non-enzymatic decomposition to generate aldehydes such as malondialdehyde (MDA) and 4-hydroxynonenal (4-HNE). High blood hydroperoxide levels are associated with mild cognitive impairment (MCI) and AD [30].
Proteins are major targets of reactive oxygen species (ROS). Protein oxidative modifications can induce unfolding or conformational changes that can lead to the loss of specific protein function [31] and the formation of cross-linked protein aggregates, which are resistant to removal by proteinases. Increased reactive oxygen species production and oxidative modification of brain proteins are significant in AD pathogenesis [32]. Carbonyl formation of 3-nitrotyrosine (3-NT) is an important marker of protein oxidation. Protein carbonyls and 3-NT levels were increased in the frontal cortex of individuals with MCI, mild AD, and AD [33].
Nucleic acid damage also occurs early in AD. Significantly elevated levels of 8OHG and 4,6-diamino-5-formamidopyrimidine have been reported in post-mortem MCI brains relative to the age-matched controls [34]. In addition to oxidative damage, reduced antioxidant defenses have been reported in MCI and early AD [35,36,37]. Plasma glutathione levels and antioxidant enzymes, such as glutathione peroxidase, catalase, and superoxide dismutase (SOD), are significantly decreased in early AD [29,38].
Aside from its presence in CNS, Aβ can be detected in platelets [39] and blood [40], where it interacts with red blood cells (RBC). Previous studies [41,42,43] suggest that Aβ-induced oxidative stress alters RBC metabolism.

3. Flavonoids

Recent studies have tested the power of natural compounds derived from plants against AD. Among these, flavonoids are ubiquitous compounds of plants, produced by plants for growth and defense against all kinds of stress, including cold tolerance. More than 6000 different flavonoids have been identified, the primary sources of which are apples, red fruits, onions, citrus fruits, nuts, and beverages such as tea, coffee, beer, and red wine. These compounds, derived from phenol, are particularly interesting for their ability to cross the blood–brain barrier and for their multi-target activity. Several studies have described flavonoids to exhibit relevant biologic activities involving the neuronal antioxidants, as well as anti-amyloidogenic properties, acting as metal chelators, showing anti-inflammatory properties, and ameliorating cognition and neuroprotection [44,45,46,47,48,49,50,51,52]. All of these capabilities are critical to counteract neurodegeneration, as they help to safeguard the number and efficiency of neurons as well as the integrity of their synaptic connections. Epidemiological studies have highlighted a direct relationship between diets rich in flavonoids, reduced risk of dementia, and potential for improved symptomatology in patients with AD [53,54]. Several subclasses of flavonoids have been identified; all are present in the human diet, as they are very abundant in vegetables, fruits, and some beverages [55]. In general, we can find flavonoids in all parts of plants, since they are produced by the plant itself for its growth and defense against all kinds of stress, including cold tolerance [56,57].

3.1. Chemical Structure and Flavonoids Classification

From a chemical point of view, flavonoids consist of two benzene rings, called A and B, linked via a third pyranosic ring C. Flavonoids can be divided into a variety of subclasses that differ in terms of the structural characteristics of the B ring and the degree of hydroxylation and glycosylation of the third ring. Typically, ring B binds in position 2 on ring C, but can also bind in position 3 or 4. We can, therefore, distinguish the isoflavones in which the B ring binds in position 3 of the C ring and the neoflavonoids in which the B ring binds in position 4 of the C ring (Figure 2). The group of flavonoids in which the B ring is linked in position 2 of the C ring can be further divided into six subgroups, according to the structural peculiarities of the C ring: flavones, flavonols, flavanones, flavanonols, flavanols, catechins, and anthocyanins [58]. Finally, the flavonoids in which the C ring is open are called chalcones.
In general, since all flavonoids contain the same core scaffold, the functional differences between the various groups and subgroups are mainly due to the different substituent groups. These are weak polybasic acids of a polyphenolic nature, characterized by varying degrees of hydroxylation, methoxylation, glycosylation or glucuronidation, and this contributes to the great variety of biological properties of this large group of polyphenols [59,60]. In fact, a different side chain can significantly influence the activity of flavonoids on the same molecular target, and the total number of hydroxyl groups is important for the enhancement of antioxidant activity, free radical scavenging, and metal ion chelation [61,62,63,64].

3.2. Biological Activities of Flavonoids

Flavonoids are a wide group of secondary metabolites characterized by many interesting biological potentials, both in vitro and in vivo (such as anticancer, antioxidant, antiaging, anti-inflammatory, antimicrobial, and immunomodulatory activities, along with modulation of the activity of key metabolic enzymes, cytoprotective and cardioprotective potentials, and inhibition of cellular proliferation, for example) and, in the last decades, they have emerged as a promising agents for neuroprotection [58,65,66,67,68]. In a recent epidemiologic study, Shishtar et al. [69] analyzed long-term dietary flavonoid intake and risk of Alzheimer’s disease and related dementias in the Framingham Offspring Cohort, with a total of 5209 participants aged 28–62 years in the original cohort. The intake effects of six classes of dietary flavonoid (flavonols, flavones, flavanones, flavan-3-ols, anthocyanins, and flavonoid polymers) and the risk of Alzheimer’s disease and related dementias (ADRD) and Alzheimer’s disease (AD) alone were analyzed based on the data from the Framingham Heart Study Offspring Cohort. Participants were ADRD-free with a valid FFQ at baseline. Flavonoid intakes were updated at each exam in order to represent the cumulative average intake across the five exams, and were expressed as percentile categories of intake to handle their nonlinear relation with ADRD and AD. After multivariate and dietary adjustments, individuals with the highest intakes of flavonols, anthocyanins, and flavonoid polymers had a lower risk of ADRD relative to individuals with the lowest intakes. A similar trend was found also in AD for flavonols and anthocyanins, but not for flavonoid polymers, showing a direct correlation between higher long-term dietary intake of flavonoids and the risks of ADRD and AD onset in American adults.
Flavonoids are widely distributed in the plant kingdom, and are characterized by chemical structures with the presence of several substituents that allow them to assume particular activities and exert beneficial effects for the wellness of organisms, as well as for their potential for therapeutic utilization. This does not indicate that every type of flavonoid is able to show biological potential, but only that those with particular characteristics can be employed for specific roles. For instance, one of the well-known and best-studied activities of flavonoids is its antioxidant activity, which is linked to the number of hydroxyl groups on the B ring. Generally, a greater number of free hydroxyl groups corresponds to a greater scavenging effect, but their location in the skeleton of flavonoids is a crucial structural element. These hydroxyl groups, through the donation of hydrogen atoms and electrons to radical species, favor the repair of the damage caused by ROS and reactive nitrogen species (RNS), reducing their degree of reactivity [70,71]. This mechanism leads to the generation of a relatively more stable flavonoid radical, which significantly reduces the oxidative stress triggered by the interrupting free radicals. At the level of the cell membrane, there is a chain reaction of propagation of peroxylic radicals between the molecules of polyunsaturated fatty acids and other intermediates [72,73]. The scavenger action of flavonoids on ROS and RNS is important because the ROS/RNS balance is directly connected with the redox state of the cell, which is also influenced by the presence of metal ions. In addition to the direct action on free radicals, experimental data demonstrate specific chelating properties of flavonoids against transition metals, mainly iron and copper ions [74,75,76]. Ferrous and/or copper ions represent a danger because they tend to react through the Fenton reaction, with hydrogen peroxide generating hydroxyl radicals, a very reactive species which rapidly oxidizes surrounding molecules, triggering the oxidative stress cascade. Chelating activity further enhances the ability of flavonoids to protect against oxidative stress, since the chelate metal ions may not participate in the generation of ROS through the Fenton reaction and because the chelates have a more powerful scavenger action against ROS than free flavonoids [77,78,79]. The chelating activity of flavonoids is considered a key mechanism for the biological activity of flavonoids, because metallocomplexes affect several biochemical properties, such as lipophilicity, membrane transport, and interaction with biomolecules [80,81,82]. In a second antioxidant mechanism, flavonoids do not act directly on ROS, but “indirectly” interact with some proteins involved in the gene expression regulation pathway. They upregulate the endogenous antioxidant capacity of the cell, and inhibit others involved in redox balance and inflammatory processes, such as cyclooxygenase, lipoxygenase, xanthine oxidase, NADH oxidase, and myeloperoxidase [83,84,85,86,87,88,89]. Therefore, the mechanisms we have described suggest not only a direct involvement of the flavonoid molecule, but also of metabolites that result from its oxidation and the formation of flavonoid–metal complexes [90].

3.3. Flavonoids in Neurodegeneration

The ability of flavonoids to cross the blood–brain barrier suggests that these compounds can feasibly have a direct effect on the brain. Numerous studies have documented the bioactivity of flavonoids against neurodegenerative disorders such as AD, Parkinson’s, Huntington’s, and other neurological disorders [51,58,64,65,66,67,68,69,91,92,93]. Regarding AD treatment, there is still no significantly efficient drug that can reduce the progression or improve the outcome of the disease [94,95]. The search for natural substances for the treatment of AD is considered key to brain health, because these compounds are often easily isolated and possess well-documented biomechanisms and safety profiles [96,97]. In addition, their abundance in vegetables and fruits has made them a major part of the human diet.

3.4. Potential Role of Flavonoids in AD Therapy

Flavonoids, including epicatechin-3-gallate, gossypetin, naringerin, quercetin, and myricetin are reported to block β-amyloid and Tau aggregation, scavenge free radicals, and sequester metal ions at clinically low concentrations [98,99]. In order to better understand flavonoids’ role in AD treatment, the pharmacological effects of some compounds from the six subclasses are described (see Figure 3).
Quercetin is a polyhydroxyflavonoid that belongs to the subclass of flavonols. Its chemical name is 3,3,4,5,7-pentahydroxyflavone, and the molecule contains five -OH groups, in positions 3, 3′, 5, 7 and 4′, which are crucial for potential biochemical–pharmacological activities. Quercetin is a natural antioxidant, widely used in healthcare for its beneficial role. Quercetin is found in flowers and fruits of edible plants; onions, apples, cherries, berries, asparagus, and red leaf lettuce have the highest levels, while tomatoes, peas, and broccoli have lower levels [92,93]. Experimental studies have shown the existence of an inverse correlation between dietary quercetin intake and risk of senile dementia. Specifically, experimental studies have identified various quercetin targets of neuronal protection that contribute to reducing the main neuronal lesions present in the brains of AD patients. Among them are hyperphosphorylation of the Tau protein, the deposition of beta amyloid, oxidative stress, inflammation, and apoptotic processes [100,101].
More specifically, Tau phosphorylation is under the control of several distinct kinases, such as Erk, Akt, p38, AMP activated protein kinase (AMPK), glycogen synthase kinase 3 beta (GSK3β), cyclin-dependent kinase 5 (cdk5), and protein phosphatase 2A (PP2A) [102,103]. Through molecular dynamic simulation studies carried out according to the molecular docking results, Zu et al. identified MAPK as core target of quercetin. Jiang et al. demonstrated the anti-apoptotic role of quercetin via MAPKs and PI3K/Akt/GSK3ß signaling pathways by preincubating HT22 cells with 5 μmol/L of the drug for 12 h [103,104]. MAPK is a heterotrimeric Ser/Thr protein kinase; its activation contributes to the hyperphosphorylation of Tau in neurons. It also checks Aβ metabolism, and is involved in cell proliferation, apoptosis, and inflammatory responses [105]. Thus, MAPK pathway inhibition can significantly improve synaptic plasticity, memory, and cognitive functions, and can be considered a valid target against AD progression [106]. GSK3β, a Ser/Thr kinase that connects numerous signaling pathways in the cell, including inflammatory responses, is another strategic target for neuroprotection. GSK3β connects numerous signaling pathways in the cell. It is a downstream enzyme of the PI3K/Akt signaling pathway, and is considered the main factor responsible for the phosphorylation of the Tau protein in AD [107]. Quercetin acts on GSK3β, decreasing the kinase activity, and therefore, quercetin indirectly has an anti-hyperphosphorylation of Tau protein action that helps to strengthen its neuroprotective effects [104]. Bao et al. have demonstrated that pretreatment for 2 h of rat pheochomocytoma PC-12 cell line with 500 µM quercetin attenuates H2O2-induced p53 expression, and also significantly reduces apoptosis and caspase 3 activation [108]. p53 regulates the action of nitric oxide synthase (NOS) and represses the transcription of peroxisome proliferator-activated receptor-γ coactivator (PGC-1α), one of the most powerful stimulators of mitochondrial biogenesis and respiration [109,110]. In AD, the decrease in the expression of PGC-1α is one of the causes related to the progression of the disease, as it is linked to the increase in the generation of the peptide Aβ [109,110]. In addition, the quercetin action on p53 affects the oxidative stress decrease, because p53 is a significant activator of the ROS-mediated apoptotic pathway [111,112]. A total of 50 μM quercetin, after 4 h of HepG2 cells incubation, has been shown to restore cellular redox homeostasis through its significant scavenger activity against ROS. Feeding the C57BL/6J mice with a 1% quercetin diet for 20 weeks increased the level of glutathione (GSH) and the expression of certain antioxidant enzymes, including SOD, catalase (CAT), and glutathione peroxidase (GPx), in hippocampal neurons [113,114,115,116].
Hung et al. demonstrated that quercetin (10mM) pretreatment of human umbilical vein endothelial cells (HUVECs) suppressed the nuclear factor- kB (NF-kB) signal, suggesting that the drug is a powerful antiatherosclerotic [117]. In addition, quercetin is a strong inhibitor of two important key enzymes involved in the pathology of AD, namely acetylcholinesterase (AchE) and butyrylcholinesterase (BchE) [118,119]. AChE and BChE contribute to hydrolytic degradation of acetylcholine (ACh), an important neurotransmitter that coordinates the excitability and activation of groups of neurons in the brain, and also influences its transmission and synaptic plasticity [120]. In the brain, a decrease in ACh corresponds to a slowdown in communication between neurons. Quercetin, through its OH groups of the phenyl ring, forms hydrogen bonds with specific amino acids in the active site of the AChE [121]. Inhibition of AChE and BChE facilitates communication between nerves and increases the activity of cholinergic pathways in the brain, relieving symptoms of memory loss [122]. In the brain, BChE also appears to play a role in the transformation of dangerous amyloid plaques to the pathogenic structures present in dementia and AD [123,124]. In Figure 4, a schematic representation of the main activities of quercetin is depicted.
Naringenin is the aglycon of naringin, and belongs to the subclass of flavanones. It is abundant in citrus fruits (especially grapefruits, to which it gives the characteristic bitter taste), as well as vegetables, and especially in grapes, tomatoes, and cherries. Naringenin can be found in two forms. One is characterized by a bond with a sugar on C7, and one derives from the action of specific enzymes which are able to cleave this glycosidic bond by releasing the aglycone [125,126]. Both forms have antioxidant activity, but naringenin has a more powerful scavenging action than its precursor naringin, because the presence of sugar in the latter causes a steric hindrance that impairs activity [127]. Naringenin (400 mM) improved learning and spatial memory in PC12 cells of rats with AD through the regulation of the PI3K/Akt/GSK-3K pathway and by reducing the hyper-phosphorylation of TAU. The intracellular mechanism that allows this neuroprotective action is related to the inhibition of caspase 3 activity, the activation of PI3K/Akt, and the modulation of the signaling pathway GSK3β, which plays a crucial role in neuronal survival [128,129,130]. The inhibition of caspase 3 also affects programmed cell death; in fact, the block of neuroapoptosis is also caused by the decreased levels of malondialdehyde and hippocampal nitrite found in socially defeated rat pups treated with naringenin (50–100 mM) [131]. Naringenin significantly regulates the (NF-kB) signaling pathway, which is implicated in inflammatory processes, and decreases tumor necrosis factor-α (TNF-α) as well as interleukin (IL)-6 and IL-1β [131]. The reduced expression of NF-kB is induced by naringenin through a significant decrease in phosphorylation and nuclear translocation of P65, a subunit of NF-kB, as well as by an increase in sirtuin 1 (SIRT1) levels in the hippocampus [132,133,134,135]. The inflammatory pathway is also repressed by the interaction of naringenin with other molecular targets, including inducible nitric oxide synthase (iNOS) and cyclooxygenase 2 (COX-2) e MAPK. iNOS and COX-2 are both inhibited in a concentration-dependent manner, while Zhang et al. demonstrated that naringenin (100 mM) inhibited MAPK signaling pathway activation by suppressing the phosphorylation of JNK and ERK1/2 in BV-2 cells [136]. Numerous studies have reported the antioxidant properties of polyphenols; it has been shown that GSH activity, one of the most efficient endogenous antioxidants, is improved by naringenin. Additionally, an increase in SOD, CAT, GPx, and glutathione reductase (GR) activity, in addition to a decrease in hydrogen peroxide (H2O2) and protein carbonyls levels, have been demonstrated [137,138,139,140]. In Figure 5, a schematic representation of the main activities of naringenin is depicted.
Epigallocatechin-3-gallate (EGCG), an ester of epigallocatechin and gallic acid, is the main bioactive polyphenol found in solid green tea extract [141,142]. EGCG has been reported to bypass the blood–brain barrier (BBB) and exert potent neuroprotective properties against AD, in a wide range of cell models [143]. Several works have shown that EGCG (5-15 mg/kg) reduces the accumulation of β amyloid able to interfere with the formation of β-sheets, the process involved in amyloid formation cascade [144,145]. EGCG promotes the non-amyloidogenic process by promoting α secretase cleavage and inhibiting β and γ secretase, by way of suppression of the ERK/NFkB pathway [146,147]. Sonawane et al. demonstrated the EGCG’s potential for dissolving pre-formed Tau filaments and oligomers in a time- and concentration-dependent manner; the IC50 for Tau aggregation by EGCG was found to be 64.2 μM [148]. Ehrnhoefer et al. have demonstrated that EGCG can convert mature Aβ fibrils into smaller forms free of toxicity, redirecting polypeptide aggregation into off-pathway protein assemblies [145]. In APP-C99-overexpressed cultured MC65 cells, EGCG (5–20 μM) is also able to suppress Aβ-induced neurotoxicity through GSK3β activation and inhibition of c-Abl/FE65 nuclear translocation [149]. In addition, it attenuates oxidative stress and mitochondrial impairment, and restores intracellular antioxidant levels in different neuronal cell lines and AD models [150]. More specifically, in EOC 13.31 microglial cell lines, EGCG (5–20 μM) suppresses the expression of TNFα, iNOS, IL-1β, and IL-6, and it inhibits the activation of the NF-Kβ and MAPK signal, but increases the synthesis of GSH. Chi et al. demonstrated significant protection of EGCG (22.5–90 μM) against oxidative damage caused by H2O2 to chicken lymphocytes. After preincubation with EGCG, the compound restored H2O2′s harmful effects, suppressing the increase of ROS and restoring the antioxidant system by mRNA expression of SOD, CAT, and GPx [151]. EGCG enhances cholinergic neurotransmission through inhibition of AChE and BChE. ACh content was significantly elevated in a dose-dependent manner, which ultimately led to an improvement in the learning and memory function of AD rats [152]. In Figure 6 a schematic representation of the main activity of epigallocatechin-3-gallate is depicted.
Myricetin is a natural flavanol widely distributed in several vegetables and fruits, mainly including blackcurrant teas, red wines, and medical herbs [153,154,155,156]. Myricetin, mainly in the form of glycoside (O-glycosides), is also known as hydroxy quercetin because of its quercetin-like structure [157,158], from which it differs by one extra hydroxyl at the 5′-OH of the B ring. The compound has a wide range of beneficial effects on human health, including antihypertensive, antiallergic, analgesic, anti-inflammatory, immunomodulatory, antiplatelets, and aggregation activities [158,159,160]. Several studies have indicated the neuroprotective properties of myricetin, which are expressed through different molecular targets. Ramezani et al. have shown that intraperitoneal injection of myricetin, at a dose of 5 or 10 mg/kg over 21 days, improves learning and memory in rat models with AD [161]. Myricetin exhibits antiamyloidogenic activities; it reduces the formation of ordered Aβ aggregation through the formation of H-bonds between its hydroxyl group and the carbonyl group on the surface of the β sheet [162,163,164]. In this way, myricetin weakens the interstrand hydrogen bonds, inhibits the extension of fibrils of Aβ, and prevents Aβ from undergoing toxic changes [164,165,166]. Antiamyloidogenic activity is also supported by the compound’s ability to interact with α and β-secretase. In more detail, in cultured rat primary cortical neurons, myricetin (10 μM) has been shown to increase the α-secretase (ADAM10) level and enzyme activity, while it inhibits the activity of β-secretase (BACE-1) with an IC50 of 2.8 µM [167,168]. Chakraborty et al. have shown that an H-bond is created between the hydroxyl group in position C7 of the A ring of myricetin and the dyad of Asp 32 and 228 of BACE-1; in this way, the enzymatic activity is strongly reduced [168]. Moreover, myricetin shows significant antioxidant and free radical scavenging effects [156,157,159,169]. It has been reported that in murine models, myricetin (40 and 80 µM) can inhibit oxidative stress generation of ROS and myeloperoxidase, and depletion of glutathione and ATP. However, it restores the levels and activity of the main antioxidant enzymes, such as SOD, CAT, and GSHpx in animal models [159,170,171]. Treatment with 80 μM myricetin for 3 h increased cell viability to 81% ± 4.2% of isolated cardiomyocytes intoxicated with 20 μg/mL aluminum phosphide (AlP) [171]. The antioxidant power of myricetin is attributable to the pyrogallol group. The molecule tends to react with free radicals to form radical semiquinones; this ability helps to interrupt the chain of reactions triggered by ROS [172,173,174]. In addition, the compound has chelating properties on metal ions such as Cu2+ and Fe2+; this, on the one hand, strongly enhances its antioxidant activity, because the Fenton reaction is inhibited and, consequently, the ROS generation is reduced. On the other hand, myricetin acts directly on Aβ complexes, reducing their toxicity through the reduction in metal ions that can interact with them [175,176,177]. Myricetin inhibits lipopolysaccharide (LPS)-induced neuroinflammation, as it reduces the levels of proinflammatory mediators, including IL, NF-kB, TNFα, iNOS and COX2, in the microglia BV2 cell line. [178]. Finally, other myricetin-like polyphenols, including gossypetin, also has inhibitory abilities against AChE [179,180]. In Figure 7, a schematic representation of the main activity of myricetin is depicted.
Gossypetin (3,5,7,8,3′,4′-hexahydroxy flavone) is a flavonol isolated from the flowers and the calyx of Hibiscus sabdariffa. Gossypetin has been shown to exert antioxidant, antimutagenic, antimicrobial, and anti-atherosclerotic activities [181,182,183,184]. Chen et al. [183], in murine macrophage cell line J774A.1, demonstrated that gossypetin (1–1000 μM) has inhibitory effects on both lipid peroxidation and lipoprotein oxidation, attenuating the formation of foam cells and lipid accumulation through PPAR pathways. The drug improves cholesterol removal from macrophages and delays atherosclerosis [183]. In addition, Lin et al. [184] demonstrated that gossypetin (0.1–0.5 μM) has inhibitory effects on abnormal vascular smooth muscle cell proliferation and migration, which could lead to the containment of atherosclerosis and other cardiovascular illnesses [184]. Inhibition of AChE and BChE activity, key enzymes in brain-related disorders, further emphasizes the therapeutic benefit of gossypetin for the treatment of AD [180,185].
Genistein (4′,5′,7-trihydroxyisoflavone) is an isoflavone distributed in several vegetables such as legumes, green peas, and peanuts, and is predominantly extracted from the Glycine max soybean [186,187]. Several researches point out genistein potential therapeutic role to delay the onset of Alzheimer’s dementia through the improve of cognitive function and synapse development [188]. In this context, Safahani et al. demonstrated that genistein supplementation modulates dopaminergic and cholinergic function, helping with memory recovery and neuroprotection in rats [189]. Genistein (10 mg/kg) protects against AD progression by reducing the production and deposition of Aβ aggregates, as well as the hyperphosphorylation of the Tau protein, in rat models [188,190,191]. Genistein protects against AD progression by reducing the generation and aggregation of Aβ. Seong et al. showed that the inhibitory activity (%) of genistein against the extent of Aβ25–35 self-aggregation, after 24 h of incubation, decreased by 34.90% when co-treated with 100 µM genistein [192]. In rat hippocampal neurons, the drug (0.375 µg/mL) downregulated presenilin levels, increased α secretase while decreasing β secretase and BACE1 activity, and, last but not least, modulated the PPARγ receptor to upregulate ApoE production [193,194,195,196]. In vitro genistein (0.391 mM) has been effectively proved to reduce oxidative stress, due to its high antioxidant power and potent ROS scavenger ability [197].
The antioxidant effects of genistein are associated with AMPK activation and the drug’s binding with estrogen receptor α (ERα), both of which promote the expression of antioxidant enzymes such as SOD, CAT, and GPx [198,199,200]. Moreover, in RAW 264.7 cell model, genistein (20 mM) prevents neuro-inflammation by regulating gene transcription of cytokines, such as TNFα, IL-1β, IL-6, and IL-12 [201,202]. Finally, Fang et al., 2014, demonstrated the inhibitory activity of AChE by two genistein derivatives, highlighting the therapeutic potentiality of this isoflavone [203].
Apigenin (4′,5,7-trihydroxyflavone) formally belongs to the flavone subclass, and is widely distributed in the plant kingdom, present principally in chamomile flowers. and in lower concentrations in vegetables, citrus fruits, herbs, and plant-based beverages (tea, beer, and wine) [204]. In a rat model of AD, apigenin (50 mg/kg) significantly reduced the hyperphosphorylation of tau levels in the hippocampus, decreasing the expression of GSK-3β, suppressing BACE1 expression, and supporting an antiamioloidogenic activity [205]. The drug (25 µM in human THP-1 monotypic cells) inhibits the production of IL-6 and IL-1β by modulating the MAPK/ERK and PI3-K/Akt signal transduction pathways associated with neuronal survival blocking [206,207,208]. Apigenin decreases ROS levels and significantly increases GSH levels, improving the cellular antioxidant defense system [209,210].
Cyanidin 3-O-glucoside belongs to the anthocyanins class; it is present in plants such as berries and soybean fruits, where it is responsible for the red, purple, and blue pigments. Cyanidin has been reported to act as neuroprotector in several disorders, such as AD, Parkinson’s disease, and multiple sclerosis [211,212,213,214,215]. The anti-AD effect is evidenced in the inhibition of Aβ amyloid accumulation. Cyanidin could directly interact with Aβ peptide through hydrogen bonds responsible for Aβ aggregation and fibrils formation [216]. In addition, Thummayot et al. (2014, 2016) demonstrated that, in human neuroblastoma (SK-N-SH) cells, cyanidin (20 μM) decreases Aβ-induced apoptosis of SK-N-SH cells by decreasing the expression levels of several proteins. This facilitates the release of the pro-apoptotic factors required for caspase cascade activation [217,218]. The neuroprotective effects of cyanidin may also be mediated via inhibition of oxidative stress and pro-inflammatory cytokines release. The compound is a scavenger of radical activity, a strong inhibitor of intracellular ROS generation, and an enhancer of the cellular antioxidant system. It has been demonstrated that pretreatment of the cells with cyanidin improved the expression of antioxidant enzymes, namely SOD, CAT, and GPx [219,220,221]. Moreover, Kaewmool et al. showed that, in LPS-stimulated BV2 microglia, cyanidin (2.5–10 mM) inhibits the signaling pathways NF-κB and p38, while MAPK suppresses the production of interleukin-1β (IL-1β) and interleukin-6 (IL-6) and subregulates the gene expressions of iNOS and COX-2 in BV2 cells [222]. In Table 1, the reported data from the in vitro and in vivo studies are described.
Unfortunately, there is still a lack of translational research and clinical evidence for these promising compounds, and we found only one clinical trial, which began in 2022 and will finish in 2024, studying the efficacy and safety of the Flos Gossypii flavonoid tablet in the treatment of Alzheimer’s disease [223]. A total of 240 patients (male and female), aged between 50 to 85 years old, who meet the diagnostic criteria of “likely AD dementia” of the National Institute on Aging—Alzheimer’s Disease Association, are primary school graduates/graduates and above, and have the ability to complete the cognitive ability test and other tests specified in the program will be enrolled. The study proposed a multicenter, randomized, double-blind, placebo-controlled, parallel method to recruit AD patients in order to confirm the efficacy and safety of the Flos Gossypii Flavonoid Tablet. This Phase II clinical trial aims to demonstrate the efficacy and safety of the Flos Gossypii Flavonoid Tablet in the treatment of mild to moderate Alzheimer’s disease (marinus sea deficiency/brain collateral stasis syndrome). The study will monitor changes in AD patients’ general cognitive and daily living activities, different cognitive domain functions, and symptom gravity. The Primary Outcome Measure is the assessment of the Alzheimer’s Disease Scale—Cognitive section (ADAS-cog/11) based on the change from baseline ADAS-cog scores to those at Week 26. It also checks the differences between the low-dose and high-dose groups in the changes in ADAS cog/11 scores (relative to baseline) at weeks 13 and 26, when compared to the placebo group. Seven components will be utilized for the assessment of ADAS-cog cognitive function: word recall, instruction, structural practice, naming, conceptual practice, orientation, and word recognition. The total score ranges from 0 to 70, with lower scores representing milder disease progression. The secondary outcomes include: mini-mental state examination, Alzheimer’s disease co-operative study—activities of daily living, clinician interview-based impression of severity, neuropsychiatric inventory, and dementia syndrome classification scale.

4. Conclusions

This review provides evidence that flavonoids have potential for treating AD, and are considered drug candidates for future clinical research. Although precise mechanisms are still unclear, flavonoids regulate several important physiological responses, which may contribute to neuroprotective effects in AD. The advantage of flavonoids over conventional targeting drugs is the possibility of administering these molecules as food supplements. Supplementation with flavonoids could allow for early protection, even at a young age. They can also be used without the need for a preclinical diagnosis, due to their low toxicity. Certainly, further long-term dietary intervention studies indicating the dosage and the times of drug assumption may contribute to fully evaluating the effectiveness of flavonoids as agents for the management of AD. It will be important to incorporate bioavailability and metabolism into experimental planning at all stages of preclinical research, in order to better clarify such mechanisms in vivo.

Author Contributions

A.C., G.L., G.T.P., F.M. and S.F. performed the literature review and drafted the paper; D.B. critically revised the paper and provided funding; D.B., F.M. and E.T. conceived the study and critically revised the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

Alzheimer disease (AD);
Presenilin1 (PS1);
Presenilin 2 (PS2);
Amyloid Precursor Protein (APP);
Apolipoprotein E (ApoE);
Amyloid beta peptide (Aβ);
Intraneuronal Neurofibrillary Tangles (NFTs);
Central Nervous System (CNS);
malondialdehyde (MDA);
4-hydroxynonenal (4-HNE);
Mild Cognitive Impairment (MCI);
Reactive Oxygen Species (ROS);
Carbonyl formation 3-nitrotyrosine (3-NT);
superoxide dismutase (SOD);
red blood cells (RBC);
Reactive Nitrogen Species (RNS);
AMP Activated Protein Kinase (AMPK);
Glycogen Synthase Kinase 3 beta (GSK3β);
cyclin-dependent kinase 5 (cdk5);
Protein Phosphatase 2A (PP2A);
nitric oxide synthase (NOS);
Peroxisome proliferator-activated receptor-γ coactivator (PGC-1α);
glutathione (GSH);
catalase (CAT);
glutathione peroxidase (GPx);
Acetylcholinesterase (AchE);
Butyrylcholinesterase (BchE);
Acetylcholine (ACh);
nuclear factor-kB (NF-kB);
tumor necrosis factor-α (TNF-α);
interleukin (IL);
sirtuin 1 (SIRT1);
inducible nitric oxide synthase (iNOS);
cyclooxygenase 2 (COX-2);
glutathione reductase (GR);
hydrogen peroxide (H2O2);
Epigallocatechin-3-gallate (EGCG),
blood–brain barrier (BBB).

References

  1. Vermunt, L.; Sikkes, S.A.M.; van den Hout, A.; Handels, R.; Bos, I.; van der Flier, W.M.; Kern, S.; Ousset, P.-J.; Maruff, P.; Skoog, I.; et al. Duration of preclinical, prodromal, and dementia stages of Alzheimer’s disease in relation to age, sex, and APOE genotype. Alzheimers Dement. 2019, 15, 888–898. [Google Scholar] [PubMed]
  2. Bateman, R.J.; Xiong, C.; Benzinger, T.L.S.; Fagan, A.M.; Goate, A.; Fox, N.C.; Marcus, D.S.; Cairns, N.J.; Xie, X.; Blazey, T.M.; et al. Dominantly Inherited Alzheimer Network. Clinical and biomarker changes in dominantly inherited Alzheimer’s disease. N. Engl. J. Med. 2012, 367, 795–804. [Google Scholar] [PubMed] [Green Version]
  3. Corrêa-Velloso, J.C.; Gonçalves, M.C.B.; Naaldijk, Y.; Oliveira-Giacomelli, A.; Pillat, M.M.; Ulrich, H. Pathophysiology in the comorbidity of bipolar disorder and Alzheimer’s disease: Pharmacological and stem cell approaches. Prog. Neuro-Psychopharm. Biol. Psych. 2018, 80, 34–53. [Google Scholar]
  4. Kamal, Z.; Ullah, F.; Ayaz, M.; Sadiq, A.; Ahmad, S.; Zeb, A.; Hussain, A.; Imran, M. Anticholinesterse and antioxidant investigations of crude extracts, subsequent fractions, saponins and flavonoids of Atriplex laciniata L.: Potential effectiveness in Alzheimer’s and other neurological disorders. Biol. Res. 2015, 48, 21. [Google Scholar]
  5. Ullah, F.; Ayaz, M.; Sadiq, A.; Hussain, A.; Ahmad, S.; Imran, M.; Zeb, A. Phenolic, flavonoid contents, anticholinesterase and antioxidant evaluation of Iris germanica var; florentina. Nat. Prod. Res. 2016, 30, 1440–1444. [Google Scholar]
  6. Duyckaerts, C.; Delatour, B.; Potier, M.-C. Classification and basic pathology of Alzheimer disease. Acta Neuropathol. 2009, 118, 5–36. [Google Scholar]
  7. McKhann, G.M.; Knopman, D.S.; Chertkow, H.; Hyman, B.T.; Jack, C.R.; Kawas, C.H., Jr.; Klunk, W.E.; Koroshetz, W.J.; Manly, J.J.; Mayeux, R.; et al. The diagnosis of dementia due to Alzheimer’s disease: Recommendations from the National Institute on Aging-Alzheimer’s Association workgroups on diagnostic guidelines for Alzheimer’s disease. Alzheimers Dement. 2011, 7, 263–269. [Google Scholar]
  8. Rakesh, G.; Szabo, S.T.; Alexopoulos, G.S.; Zannas, A.S. Strategies for dementia prevention: Latest evidence and implications. Ther. Adv. Chronic. Dis. 2017, 8, 121–136. [Google Scholar]
  9. Hu, N.; Yu, J.-T.; Tan, L.; Wang, Y.-L.; Sun, L.; Tan, L. Nutrition and the risk of Alzheimer’s disease. BioMed. Res. Int. 2013, 2013, 524820. [Google Scholar]
  10. Tellone, E.; Galtieri, A.; Russo, A.; Ficarra, S. Protective effects of the caffeine against neurodegenerative diseases. Curr. Med. Chem. 2019, 25, 5137–5151. [Google Scholar]
  11. Talarek, S.; Listos, J.; Barreca, D.; Tellone, E.; Sureda, A.; Nabavi, S.F.; Braidy, N.; Nabavi, S.M. Neuroprotective effects of honokiol: From chemistry to medicine. Biofactors 2017, 43, 760–769. [Google Scholar] [CrossRef] [PubMed]
  12. Barreca, D.; Currò, M.; Bellocco, E.; Ficarra, S.; Laganà, G.; Tellone, E.; Laura Giunta, M.; Visalli, G.; Caccamo, D.; Galtieri, A.; et al. Neuroprotective effects of phloretin and its glycosylated derivative on rotenone-induced toxicity in human SH-SY5Y neuronal-like cells. Biofactors 2017, 43, 549–557. [Google Scholar] [CrossRef]
  13. Tellone, E.; Galtieri, A.; Russo, A.; Ficarra, S. How does resveratrol influence the genesis of some neurodegenerative diseases? Neural Regen Res. 2016, 11, 86–87. [Google Scholar] [CrossRef] [PubMed]
  14. Carelli-Alinovi, C.; Ficarra, S.; Russo, A.M.; Giunta, E.; Barreca, D.; Galtieri, A.; Misiti, F.; Tellone, E. Involvement of acetylcholinesterase and protein kinase C in the protective effect of caffeine against β-amyloid-induced alterations in red blood cells. Biochimie 2016, 121, 52–59. [Google Scholar] [CrossRef] [PubMed]
  15. Tellone, E.; Galtieri, A.; Russo, A.; Giardina, B.; Ficarra, S. Resveratrol: A Focus on Several Neurodegenerative Diseases. Oxidative Med. Cell Longev. 2015, 2015, 392169. [Google Scholar] [CrossRef] [Green Version]
  16. Jones, D.P. Redefining oxidative stress. Antioxid. Redox Signal. 2006, 8, 1865–1879. [Google Scholar] [CrossRef]
  17. Halliwell, B. Antioxidants and human disease: A general introduction. Nutr. Rev. 1997, 55, S44–S49. [Google Scholar] [CrossRef]
  18. Smith, M.A.; Harris, P.L.R.; Sayre, L.M.; Perry, G. Iron accumulation in Alzheimer disease is a source of redox-generated free radicals. Proc. Natl Acad. Sci. USA 1997, 94, 9866–9868. [Google Scholar] [CrossRef]
  19. Butterfield, D.A.; Boyd-Kimball, D. Amyloid beta-peptide (1-42) contributes to the oxidative stress and neurodegeneration found in Alzheimer disease brain. Brain Pathol. 2004, 14, 426–432. [Google Scholar] [CrossRef]
  20. Pratico, D.; Delanty, N. Oxidative injury in diseases of the central nervous system: Focus on Alzheimer’s disease. Am. J. Med. 2000, 109, 577–585. [Google Scholar] [CrossRef]
  21. Li, F.; Calingasan, N.Y.; Yu, F.; Mauck, W.M.; Toidze, M.; Almeida, C.G.; Takahashi, R.H.; Carlson, G.A.; Flint Beal, M.; Lin, M.T.; et al. Increased plaque burden in brains of APP mutant MnSOD heterozygous knockout mice. J. Neurochem. 2004, 89, 1308–1312. [Google Scholar] [CrossRef] [PubMed]
  22. Nishida, Y.; Yokota, T.; Takahashi, T.; Uchihara, T.; Jishage, K.-I.; Mizusawa, H. Deletion of vitamin E enhances phenotype of Alzheimer disease model mouse. Biochem. Biophys. Res. Comm. 2006, 350, 530–536. [Google Scholar] [CrossRef] [PubMed]
  23. Tamagno, E.; Bardini, P.; Obbili, A.; Vitali, A.; Borghi, R.; Zaccheo, D.; Pronzato, M.A.; Danni, O.; Smith, M.A.; Perry, G.; et al. Oxidative stress increases expression and activity of BACE in NT2 neurons. Neurobiol. Dis. 2002, 10, 279–288. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Fukumoto, H.; Cheung, B.S.; Hyman, B.T.; Irizarry, M.C. β-secretase protein and activity are increased in the neocortex in Alzheimer disease. Arch. Neurol. 2002, 59, 1381–1389. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Carelli-Alinovi, C.; Misiti, F. Methionine 35 sulphoxide reduces toxicity of Aβ in red blood cell. Eur. J. Clin. Investig. 2017, 47, 314–321. [Google Scholar] [CrossRef] [PubMed]
  26. Lovell, M.A.; Ehmann, W.D.; Mattson, M.P.; Markesbery, W.R. Elevated 4-hydroxynonenal in ventricular fluid in Alzheimer’s disease. Neurobiol. Aging 1997, 18, 457–461. [Google Scholar] [CrossRef]
  27. Zaman, Z.; Roche, S.; Fielden, P.; Frost, P.G.; Niriella, D.C.; Cayley, A.C. Plasma concentrations of vitamins A and E and carotenoids in Alzheimer’s disease. Age Ageing 1992, 21, 91–94. [Google Scholar] [CrossRef]
  28. Puertas, M.C.; Martinez-Martos, J.M.; Cobo, M.P.; Carrera, M.P.; Mayas, M.D.; Ramirez-Exposito, M.J. Plasma oxidative stress parameters in men and women with early stage Alzheimer type dementia. Exp. Gerontol. 2012, 47, 625–630. [Google Scholar] [CrossRef]
  29. Cervellati, C.; Romani, A.; Seripa, D.; Cremonini, E.; Bosi, C.; Magon, S.; Bergamini, C.M.; Valacchi, G.; Pilotto, A.; Zuliani, G. Systemic oxidative stress and conversion to dementia of elderly patients with mild cognitive impairment. Biomed. Res. Int. 2014, 2014, 309507. [Google Scholar] [CrossRef]
  30. Dean, R.T.; Fu, S.; Stocker, R.; Davies, M.J. Biochemistry and pathology of radical-mediated protein oxidation. Biochem. J. 1997, 324, 1–18. [Google Scholar] [CrossRef] [Green Version]
  31. Retz, W.; Gsell, W.; Munch, G.; Rosler, M.; Riederer, P. Free radicals in Alzheimer’s disease. J. Neural. Transm. Suppl. 1998, 54, 221–236. [Google Scholar] [PubMed]
  32. Aksenova, M.V.; Aksenov, M.Y.; Payne, R.M.; Trojanowski, J.Q.; Schmidt, K.L.; Carney, J.M.; Butterfield, D.A.; Markesbery, W.R. Oxidation of cytosolic proteins and expression of creatine kinase BB in frontal lobe in different neurodegenerative disorders. Dement. Geriatr. Cogn. Disord. 1999, 10, 158–165. [Google Scholar] [CrossRef] [PubMed]
  33. Arslan, J.; Jamshed, H.; Qureshi, H. Early Detection and Prevention of Alzheimer’s Disease: Role of Oxidative Markers and Natural Antioxidants. Front. Aging Neurosci. 2020, 12, 231. [Google Scholar] [CrossRef] [PubMed]
  34. Rinaldi, P.; Polidori, M.; Metastasio, A.; Mariani, E.; Mattioli, P.; Cherubini, A.; Catani, M.; Cecchetti, R.; Senin, U.; Mecocci, P. Plasma antioxidants are similarly depleted in mild cognitive impairment and in Alzheimer’s disease. Neurobiol. Aging 2003, 24, 915–919. [Google Scholar] [CrossRef] [PubMed]
  35. Baldeiras, I.; Santana, I.; Proença, M.T.; Garrucho, M.H.; Pascoal, R.; Rodrigues, A.; Duro, D.; Oliveira, C.R. Oxidative damage and progression to Alzheimer’s disease in patients with mild cognitive impairment. J. Alzheimers Dis. 2010, 21, 1165–1177. [Google Scholar] [CrossRef]
  36. Chico, L.; Simoncini, C.; Gerfo, A.L.; Rocchi, A.; Petrozzi, L.; Carlesi, C.; Volpi, L.; Tognoni, G.; Siciliano, G.; Bonuccelli, U. Oxidative stress and APO E polymorphisms in Alzheimer’s disease and in mild cognitive impairment. Free Radic. Res. 2013, 47, 569–576. [Google Scholar] [CrossRef]
  37. Torres, L.L.; Quaglio, N.B.; de Souza, G.T.; Garcia, R.T.; Dati, L.M.M.; Moreira, W.L.; Loureiro, A.P.D.M.; de Souza-Talarico, J.N.; Smid, J.; Porto, C.S.; et al. Peripheral oxidative stress biomarkers in mild cognitive impairment and Alzheimer’s disease. J. Alzheimers Dis. 2011, 26, 59–68. [Google Scholar] [CrossRef]
  38. Chen, M.; Inestrosa, G.S.; Ross, H.L.; Fernandez, H.L. Platelets are the primary source of amyloid β-peptide in human blood. Biochem. Biophys. Res. Commun. 1995, 213, 96–103. [Google Scholar] [CrossRef]
  39. Seubert, P.; Vigo-Pelfrey, C.; Esch, F.; Lee, M.; Dovey, H.; Davis, D.; Sinha, S.; Schiossmacher, M.; Whaley, J.; Swindlehurst, C.; et al. Isolation and quantification of soluble Alzheimer’s β-peptide from biological fluids. Nature 1992, 359, 325–327. [Google Scholar] [CrossRef]
  40. Clementi, M.E.; Giardina, B.; Colucci, D.; Galtieri, A.; Misiti, F. Amyloid-β peptide affects the oxygen dependence of RBC metabolism: A role for caspase 3. Int. J. Biochem. Cell Biol. 2007, 39, 727–735. [Google Scholar] [CrossRef]
  41. Carelli-Alinovi, C.; Pirolli, D.; Giardina, B.; Misiti, F. Protein kinase C mediates caspase 3 activation: A role for erythrocyte morphology changes. Clin. Hemorheol. Microcirc. 2015, 59, 345–354. [Google Scholar] [CrossRef] [PubMed]
  42. Jabir, N.R.; Khan, F.R.; Tabrez, S. Cholinesterase targeting by polyphenols: A therapeutic approach for the treatment of Alzheimer’s disease. CNS Neurosci Ther. 2018, 24, 753–762. [Google Scholar] [CrossRef] [PubMed]
  43. Havsteen, B.H. The biochemistry and medical significance of the flavonoids. Pharmacol. Ther. 2002, 96, 67–202. [Google Scholar] [PubMed]
  44. Grosso, C.; Valentão, P.; Ferreres, F.; Andrade, P. The use of flavonoids in central nervous system disorders. Curr. Med. Chem. 2013, 20, 4694–4719. [Google Scholar] [CrossRef] [PubMed]
  45. Kelsey, N.A.; Wilkins, H.M.; Linseman, D.A. Nutraceutical antioxidants as novel neuroprotective agents. Molecules 2010, 15, 7792–7814. [Google Scholar] [CrossRef] [Green Version]
  46. Li, J.K.; Jiang, Z.T.; Li, R.; Tan, J. Investigation of antioxidant activities and free radical scavenging of flavonoids in leaves of Polygonum multiflorum Thumb. China Food Addit. 2012, 2, 69–74. [Google Scholar]
  47. Prakash, D.; Sudhandiran, G. Dietary flavonoid fisetin regulates aluminium chloride-induced neuronal apoptosis in cortex and hippocampus of mice brain. J. Nutr. Biochem. 2015, 26, 1527–1539. [Google Scholar] [CrossRef]
  48. Ashafaq, M.; Raza, S.S.; Khan, M.M.; Ahmad, A.; Javed, H.; Ahmad, E.; Tabassum, R.; Islam, F.; Siddiqui, M.S.; Safhi, M.M.; et al. Catechin hydrate ameliorates redox imbalance and limits inflammatory response in focal cerebral ischemia. Neurochem. Res. 2012, 37, 1747–1760. [Google Scholar] [CrossRef]
  49. Williams, R.J.; Spencer, J.P. Flavonoids, cognition, and dementia: Actions, mechanisms, and potential therapeutic utility for Alzheimer disease. Free Radic. Biol. Med. 2012, 52, 35–45. [Google Scholar] [CrossRef]
  50. Bakhtiari, M.; Panahi, Y.; Ameli, J.; Darvishi, B. Protective effect of flavonoids against Alzheimer’s disease-related neural dysfunctions. Biomed. Pharmacother. 2017, 93, 218–229. [Google Scholar] [CrossRef]
  51. Mohebali, N.; Shahzadeh Fazeli, S.A.; Ghafoori, H.; Farahmand, Z.; MohammadKhani, E.; Vakhshiteh, F.; Ghamarian, A.; Farhangniya, M.; Sanati, M.H. Effect of flavonoid rich extract of Capparis spinosa on inflammatory involved genes in amyloid beta peptide injected rat model of Alzheimer’s disease. Nutr. Neurosci. 2018, 21, 143–150. [Google Scholar] [CrossRef] [PubMed]
  52. Morris, M.C.; Tangney, C.C.; Wang, Y.; Sacks, F.M.; Bennett, D.A.; Aggarwal, N.T. MIND diet associated with reduced incidence of Alzheimer’s disease. Alzheimer’s Dement. 2015, 11, 1007–1014. [Google Scholar] [CrossRef] [Green Version]
  53. Hill, E.; Goodwill, A.M.; Gorelik, A.; Szoeke, C. Diet and biomarkers of Alzheimer’s disease: A systematic review and meta-analysis. Neurobiol. Aging 2019, 76, 45–52. [Google Scholar] [CrossRef] [PubMed]
  54. Manach, C.; Scalbert, A.; Morand, C.; Rémésy, C.; Jiménez, L. Polyphenols: Food sources and bioavailability. Am. J. Clin. Nutr. 2004, 79, 727–747. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Dewick, P.M. The shikimate pathway: Aromatic amino acids and phenylpropanoids. In Medicinal Natural Products: A Biosynthetic Approach, 2nd ed.; Dewick, P.M., Ed.; John Wiley: Chichester, UK, 2001; pp. 137–186. [Google Scholar]
  56. Takahashi, A.; Ohnishi, T. The significance of the study about the biological effects of solar ultraviolet radiation using the exposed facility on the international space station. Biol. Sci. Space 2004, 18, 255–260. [Google Scholar] [CrossRef] [Green Version]
  57. Agati, G.; Azzarello, E.; Pollastri, S.; Tattini, M. Flavonoids as antioxidants in plants: Location and functional significance. Plant Sci. 2012, 196, 67–76. [Google Scholar] [CrossRef]
  58. Ayaz, M.; Sadiq, A.; Junaid, M.; Ullah, F.; Ovais, M.; Ullah, I.; Ahmed, J.; Shahid, M. Flavonoids as Prospective Neuroprotectants and Their Therapeutic Propensity in Aging Associated Neurological Disorders. Front. Aging Neurosci. 2019, 11, 115. [Google Scholar] [CrossRef] [Green Version]
  59. Gil, E.S.; Cout, R.O. Flavonoid electrochemistry: A review on the electroanalytical applications. Rev. Bras. De Farmacogn. 2013, 23, 542–558. [Google Scholar] [CrossRef] [Green Version]
  60. Qiu, T.; Wu, D.; Yang, L.; Ye, H.; Wang, Q.; Cao, Z.; Tang, K. Exploring the Mechanism of Flavonoids Through Systematic Bioinformatics Analysis. Front. Pharmacol. 2018, 9, 918. [Google Scholar] [CrossRef]
  61. Pandey, A.K.; Mishra, A.K.; Mishra, A. Antifungal and antioxidative potential of oil and extracts derived from leaves of Indian spice plant Cinnamomum tamala. Cell. Mol. Biol. 2012, 58, 142–147. [Google Scholar]
  62. Husain, S.R.; Cillard, J.; Cillard, P. Hydroxyl radical scavenging activity of flavonoids. Phytochemistry 1987, 26, 2489–2491. [Google Scholar] [CrossRef]
  63. Kumar, S.; Mishra, A.; Pandey, A.K. Antioxidant mediated protective effect of Parthenium hysterophorus against oxidative damage using in vitro models. BMC Compl. Altern. Med. 2013, 13, 120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Sekher Pannala, A.; Chan, T.S.; O’Brien, P.J.; Rice-Evans, C.A. Flavonoid B-ring chemistry and antioxidant activity: Fast reaction kinetics. Biochem. Biophys. Res. Commun. 2001, 282, 1161–1168. [Google Scholar] [CrossRef] [PubMed]
  65. Hole, K.L.; Williams, R.J. Flavonoids as an Intervention for Alzheimer’s Disease: Progress and Hurdles Towards Defining a Mechanism of Action. Brain Plast 2021, 6, 167–192. [Google Scholar] [CrossRef]
  66. Minocha, M.; Birla, H.; Obaid, A.A.; Rai, V.; Sushma, P.; Shivamallu, C.; Moustafa, M.; Al-Shehri, M.; Al-Emam, A.; Tikhonova, M.A.; et al. Flavonoids as Promising Neuroprotectants and Their Therapeutic Potential against Alzheimer’s Disease. Oxidative Med. Cell. Longev. 2022, 2022, 6038996. [Google Scholar] [CrossRef]
  67. de Andrade Teles, R.B.; Diniz, T.C.; Costa Pinto, T.C.; de Oliveira Junior, R.G.; Gama, E.S.M.; de Lavor, E.M.; Fernandes, A.W.C.; de Oliveira, A.P.; de Almeida Ribeiro, F.P.R.; da Silva, A.A.M.; et al. Flavonoids as Therapeutic Agents in Alzheimer’s and Parkinson’s Diseases: A Systematic Review of Preclinical Evidences. Oxidative Med. Cell. Longev. 2018, 2018, 7043213. [Google Scholar] [CrossRef] [Green Version]
  68. Chauhan, P.S.; Yadav, D.; Arukha, A.P. Dietary Nutrients and Prevention of Alzheimer’s Disease. CNS Neurol. Disord. Drug. Targets 2022, 21, 217–227. [Google Scholar] [CrossRef]
  69. Shishtar, E.; Rogers, G.T.; Blumberg, J.B.; Au, R.; Jacques, P.F. Long-term dietary flavonoid intake and risk of Alzheimer disease and related dementias in the Framingham Offspring Cohort. Am. J. Clin. Nutr. 2020, 112, 343–353. [Google Scholar] [CrossRef]
  70. Heijnen, C.G.; Haenen, G.R.; van Acker, F.A.; van der Vijgh, W.J.; Bast, A. Flavonoids as peroxynitrite scavengers: The role of the hydroxyl groups. Toxicol In Vitro 2001, 15, 3–6. [Google Scholar] [CrossRef]
  71. Di Meo, F.; Lemaur, V.; Cornil, J.; Lazzaroni, R.; Duroux, J.L.; Olivier, Y.; Trouillas, P. Free radical scavenging by natural polyphenols: Atom versus electron transfer. J. Phys. Chem. A 2013, 117, 2082–2092. [Google Scholar] [CrossRef]
  72. Alov, P.; Tsakovska, I.; Pajeva, I. Computational studies of free radical-scavenging properties of phenolic compounds. Curr. Top. Med. Chem. 2015, 15, 85–104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Guo, M.L.; Perez, C.; Wei, Y.B.; Rapoza, E.; Su, G.; Bou-Abdallah, F.; Chasteen, N.D. Iron-binding properties of plant phenolics and cranberry’s bio-effects. Dalton Trans. 2007, 10, 4951–4961. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Horniblow, R.D.; Henesy, D.; Iqbal, T.H.; Tselepis, C. Modulation of iron transport, metabolism and reactive oxygen status by quercetin-iron complexes in vitro. Mol. Nutr. Food Res. 2016, 61, 1600692. [Google Scholar] [CrossRef] [PubMed]
  75. Milicevic, A.; Raos, N. Modelling of Protective Mechanism of Iron(II)-polyphenol Binding with OH-related Molecular Descriptors. Croat. Chem. Acta 2016, 89, 89. [Google Scholar] [CrossRef]
  76. Symonowicz, M.; Kolanek, M. Flavonoids and their properties to form chelate complexes. Biotechnol Food Sci. 2012, 76, 35–41. [Google Scholar]
  77. Cherrak, S.A.; Mokhtari-Soulimane, N.; Berroukeche, F.; Bensenane, B.; Cherbonnel, A.; Merzouk, H.; Elhabiri, M. In Vitro Antioxidant versus Metal Ion Chelating Properties of Flavonoids: A Structure-Activity Investigation. PLoS ONE 2016, 11, e0165575. [Google Scholar] [CrossRef] [Green Version]
  78. Perez, C.A.; Wei, Y.; Guo, M. Iron-binding and anti-Fenton properties of baicalein and baicalin. J. Inorg. Biochem. 2009, 103, 326–332. [Google Scholar] [CrossRef] [Green Version]
  79. Kim, Y.A.; Tarahovsky, Y.S.; Yagolnik, E.A.; Kuznetsova, S.M.; Muzafarov, E.N. Lipophilicity of flavonoid complexes with iron(II) and their interaction with liposomes. Biochem. Biophys. Res. Commun. 2013, 431, 680–685. [Google Scholar] [CrossRef]
  80. Kim, Y.A.; Tarahovsky, Y.S.; Yagolnik, E.A.; Kuznetsova, S.M.; Muzafarov, E.N. Integration of Quercetin-Iron Complexes intoPh osphatidylcholine or Phosphatidylethanolamine Liposomes. Appl. Biochem. Biotech. 2015, 176, 1904–1913. [Google Scholar] [CrossRef]
  81. Kostyuk, V.A.; Potapovich, A.; Vladykovskaya, E.; Korkina, L.; Afanas’ev, I. Influence of Metal Ions on Flavonoid Protection against Asbestos-Induced Cell Injury. Arch. Biochem. Biophys. 2001, 385, 129–137. [Google Scholar] [CrossRef]
  82. Erlank, H.; Elmann, A.; Kohen, R.; Kanner, J. Polyphenols activate nrf2 in astrocytes via h2o2, semiquinones, and quinones. Free Radic. Biol. Med. 2011, 51, 2319–2327. [Google Scholar] [CrossRef] [PubMed]
  83. Lee-Hilz, Y.Y.; Boerboom, A.M.J.; Westphal, A.H.; van Berkeln, W.J.; Aarts, J.M.; Rietjens, I.M. Pro-oxidant activity of flavonoids induces epre-mediated gene expression. Chem. Res. Toxicol. 2006, 19, 1499–1505. [Google Scholar] [CrossRef] [PubMed]
  84. Speisky, H.; Shahidi, F.; Costa de Camargo, A.; Fuentes, J. Revisiting the oxidation of flavonoids: Loss, conservation or enhancement of their antioxidant properties. Antioxidants 2022, 11, 133. [Google Scholar] [CrossRef] [PubMed]
  85. Mahesha, H.; Singh, S.A.; Rao, A.A. Inhibition of lipoxygenase by soy isoflavones: Evidence of isoflavones as redox inhibitors. Arch. Biochem. Biophys. 2007, 461, 176–185. [Google Scholar] [CrossRef]
  86. Ribeiro, D.; Freitas, M.; Tomé, S.M.; Silva, A.M.S.; Laufer, S.; Lima, J.L.F.C.; Fernandes, E. Flavonoids Inhibit COX-1 and COX-2 enzymes and cytokine/chemokine production in human whole blood. Inflammation 2014, 38, 858–870. [Google Scholar] [CrossRef]
  87. Nagao, A.; Seki, M.; Kobayashi, H. Inhibition of xanthine oxidase by flavonoids. Biosci. Biotechnol. Biochem. 1999, 63, 1787–1790. [Google Scholar] [CrossRef]
  88. Bohmont, C.; Aaronson, L.M.; Mann, K.; Pardini, R.S. Inhibition of mitochondrial NADH oxidase, succinoxidase, and ATPase by naturally occurring flavonoids. J. Nat. Prod. 1987, 50, 427–433. [Google Scholar] [CrossRef]
  89. Kejík, Z.; Kaplánek, R.; Masarík, M.; Babula, P.; Matkowski, A.; Filipenský, P.; Veselá, K.; Gburek, J.; Sýkora, D.; Martásek, P.; et al. Iron Complexes of Flavonoids-Antioxidant Capacity and beyond. Int. J. Mol. Sci. 2021, 22, 646. [Google Scholar] [CrossRef]
  90. Spencer, J.P.; Vafeiadou, K.; Williams, R.J.; Vauzour, D. Neuroinflammation: Modulation by flavonoids and mechanisms of action. Mol. Asp. Med. 2012, 33, 83–97. [Google Scholar] [CrossRef]
  91. Babaei, F.; Mirzababaei, M.; Nassiri-Asl, M. Quercetin in food: Possible mechanisms of its effect on memory. J. Food Sci. 2018, 83, 2280–2287. [Google Scholar] [CrossRef] [Green Version]
  92. Costa, L.G.; Garrick, J.M.; Roquè, P.J.; Pellacani, C. Mechanisms of neuroprotection by quercetin: Counteracting oxidative stress andmore. Oxidative Med. Cell. Long. 2016, 2016, 2986796. [Google Scholar]
  93. Atri, A. Current and future treatments in Alzheimer’s disease. Semin. Neurol. 2019, 39, 227–240. [Google Scholar] [CrossRef] [PubMed]
  94. Storr, T. Multifunctional compounds for the treatment of Alzheimer’s disease. Can. J. Chem. 2021, 99, 1–9. [Google Scholar] [CrossRef]
  95. Elumalai, P.; Lakshmi, S. Role of Quercetin Benefits in Neurodegeneration. Adv. Neurobiol. 2016, 12, 229–245. [Google Scholar] [PubMed]
  96. Andrade, S.; Ramalho, M.J.; Loureiro, J.A.; Pereira, M.D.C. Natural compounds for Alzheimer’s disease therapy: A systematic review of preclinical and clinical studies. Int. J. Mol. Sci. 2019, 10, 2313. [Google Scholar] [CrossRef] [Green Version]
  97. Ono, K.; Yoshiike, Y.; Takashima, A.; Hasegawa, K.; Naiki, H.; Yamada, M. Potent anti-amyloidogenic and fibril-destabilizing effects of polyphenols in vitro: Implications for the prevention and therapeutics of Alzheimer’s disease. J. Neurochem. 2003, 87, 172–181. [Google Scholar] [CrossRef] [PubMed]
  98. Weinreb, O.; Mandel, S.; Amit, T.; Youdim, M.B. Neurological mechanisms of green tea polyphenols in Alzheimer’s and Parkinson’s diseases. Nutr. Biochem. 2004, 15, 506–516. [Google Scholar] [CrossRef] [PubMed]
  99. Khan, A.; Ali, T.; Rehman, S.U.; Khan, M.S.; Alam, S.I.; Ikram, M.; Muhammad, T.; Saeed, K.; Badshah, H.; Kim, M.O. Neuroprotective effect of quercetin against the detrimental effects of LPS in the adult mouse brain. Front. Pharmacol. 2018, 9, 1–16. [Google Scholar] [CrossRef] [Green Version]
  100. Sandhir, R.; Mehrotra, A. Quercetin supplementation is effective in improving mitochondrial dysfunctions induced by 3-nitropropionic acid: Implications in Huntington’s disease. Bioch. Biophys. Acta 2013, 1832, 421–430. [Google Scholar] [CrossRef] [Green Version]
  101. Hébert, S.S.; Papadopoulou, A.S.; Smith, P.; Galas, M.C.; Planel, E.; Silahtaroglu, A.N.; Sergeant, N.; Buée, L.; De Strooper, B. Genetic ablation of Dicer in adult forebrain neurons results in abnormal tau hyperphosphorylation and neurodegeneration. Hum. Mol. Genet. 2010, 19, 3959–3969. [Google Scholar] [CrossRef] [Green Version]
  102. Sergeant, N.; Bretteville, A.; Hamdane, M.; Caillet-Boudin, M.L.; Grognet, P.; Bombois, S.; Blum, D.; Delacourt, A.; Pasquier, F.; Vanmechelen, E.; et al. Biochemistry of tau in Alzheimer’s disease and related neurological disorders. Expert Rev. Proteom. 2008, 5, 207–224. [Google Scholar] [CrossRef] [PubMed]
  103. Zu, G.; Sun, K.; Li, L.; Zu, X.; Han, T.; Huang, H. Mechanism of quercetin therapeutic targets for Alzheimer disease and type 2 diabetes mellitus. Nat. Sci. Rep. 2021, 11, 22959. [Google Scholar] [CrossRef] [PubMed]
  104. Jiang, W.; Luo, T.; Li, S.; Zhou, Y.; Shen, X.Y.; He, F.; Xu, J.; Wang, H.Q. Quercetin protects against okadaic acid-induced injury via MAPK and PI3K/Akt/GSK3β signaling pathways in HT22 hippocampal neurons. PLoS ONE. 2016, 11, e0152371. [Google Scholar] [CrossRef] [PubMed]
  105. Vingtdeux, V.; Giliberto, L.; Zhao, H.; Chandakkar, P.; Wu, Q.; Simon, J.E.; Janle, E.M.; Lobo, J.; Ferruzzi, M.G.; Davies, P.; et al. AMP-activated protein kinase signaling activation by resveratrol modulates amyloid-beta peptide metabolism. J. Biol. Chem. 2010, 285, 9100–9113. [Google Scholar] [CrossRef] [Green Version]
  106. Samuels, I.S.; Karlo, J.C.; Faruzzi, A.N.; Pickering, K.; Herrup, K.; Sweatt, J.D.; Saitta, S.C.; Landreth, G.E. Deletion of ERK2 mitogen-activated protein kinase identifies its key roles in cortical neurogenesis and cognitive function. J. Neurosci. 2008, 28, 6983–6995. [Google Scholar] [CrossRef] [Green Version]
  107. Hu, S.; Cui, W.; Mak, S.; Tang, J.; Choi, C.; Pang, Y.; Han, Y. Bis(propyl)-cognitin protects against glutamate induced neuro-excitotoxicity via concurrent regulation of NO, MAPK/ERK and PI3-K/Akt/GSK3β pathways. Neurochem. Int. 2013, 62, 468–477. [Google Scholar] [CrossRef]
  108. Bao, D.; Wang, J.; Pang, X.; Liu, H. Protective effect of quercetin against oxidative stress-induced cytotoxicity in rat pheochromocytoma (PC-12) cells. Molecules 2017, 22, 1122. [Google Scholar] [CrossRef] [Green Version]
  109. St-Pierre, J.; Drori, S.; Uldry, M.; Silvaggi, J.M.; Rhee, J.; Jäger, S.; Handschin, C.; Zheng, K.; Lin, J.; Yang, W.; et al. Suppression of reactive oxygen species and neurodegeneration by the PGC-1 transcriptional coactivators. Cell 2006, 127, 397–408. [Google Scholar] [CrossRef] [Green Version]
  110. Qin, W.; Haroutunian, V.; Katsel, P.; Cardozo, C.P.; Ho, L.; Buxbaum, J.D.; Pasinetti, G.M. PGC-1alpha expression decreases in the Alzheimer disease brain as a function of dementia. Arch. Neurol. 2009, 66, 352–361. [Google Scholar] [CrossRef]
  111. Sen, N.; Satija, Y.K.; Das, S. PGC-1a, a Key Modulator of p53, Promotes cell survival upon metabolic stress. Mol. Cell 2011, 44, 621–634. [Google Scholar] [CrossRef] [Green Version]
  112. Vigneron, A.; Vousde, K.H. p53, ROS and senescence in the control of aging. Aging 2010, 2, 471–474. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Chen, B.H.; Park, J.H.; Ahn, J.H.; Cho, J.H.; Kim, I.H.; Lee, J.C.; Won, M.H.; Lee, C.H.; Hwang, I.K.; Kim, J.D.; 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] [PubMed]
  114. Prasad, J.; Baitharu, I.; Sharma, A.K.; Dutta, R.; Prasad, D.; Singh, S.B. Quercetin reverses hypobaric hypoxia-induced hippocampal neurodegeneration and improves memory function in the rat. High Alt. Med. Biol. 2013, 14, 383–394. [Google Scholar] [CrossRef] [PubMed]
  115. Kobori, M.; Takahashi, Y.; Akimoto, Y.; Sakurai, M.; Matsunaga, I.; Nishimuro, H.; Ippoushi, K.; Oike, H.; Ohnishi-Kameyama, M. Chronic high intake of quercetin reduces oxidative stress and induces expression of the antioxidant enzymes in the liver and visceral adipose tissues in mice. J. Funct. Foods. 2015, 15, 551–560. [Google Scholar] [CrossRef]
  116. Belen Granado-Serrano, A.; Angeles Martin, M.; Bravo, L.; Goya, L.; Ramos, S. Quercetin modulates Nrf2 and glutathione-related defenses in HepG2 cells: Involvement of p38. Chem. Biol. Interact. 2012, 195, 154–164. [Google Scholar] [CrossRef] [Green Version]
  117. Hung, C.H.; Chan, S.H.; Chu, P.M.; Tsai, K.L. Quercetin is a potent anti-atherosclerotic compound by activation of SIRT1 signaling under oxLDL stimulation. Mol. Nutr. Food Res. 2015, 59, 1905–1917. [Google Scholar] [CrossRef]
  118. Khan, M.T.; Orhan, I.; Senol, F.S.; Kartal, M.; Sener, B.; Dvorská, M.; Smejkal, K.; Slapetová, T. Cholinesterase inhibitory activities of some flavonoid derivatives and chosen xanthone and their molecular docking studies. Chem. Biol. Interact. 2009, 181, 383–389. [Google Scholar] [CrossRef]
  119. Perry, E.K.; Perry, R.H.; Blessed, G.; Tomlinson, B.E. Changes in brain cholinesterases in senile dementia of Alzheimer type. Neuropath. Appl. Neurobiol. 1978, 4, 273–277. [Google Scholar] [CrossRef]
  120. Nordberg, A.; Ballard, C.; Bullock, R.; Darreh-Shori, T.; Somogyi, M. A review of butyrylcholinesterase as a therapeutic target in the treatment of Alzheimer’s disease. Prim. Care. Companion CNS Disord. 2013, 15, PCC.12r01412. [Google Scholar] [CrossRef]
  121. Ademosun, A.O.; Oboh, G.; Bello, F.; Ayeni, P.O. Antioxidative properties and effect of quercetin and its glycosylated form (rutin) on acetylcholinesterase and butyrylcholinesterase activities. J. Evid. Based Complem. Altern. Med. 2016, 21, NP11–NP17. [Google Scholar] [CrossRef]
  122. Mehta, M.; Adem, A.; Sabbagh, M. New acetylcholinesterase inhibitors for Alzheimer’s disease. Int. J. Alzheimers Dis. 2012, 2012, 728983. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Guillozet, A.L.; Mesulam, M.M.; Smiley, J.F.; Mash, D.C. Butyrylcholinesterase in the life cycle of amyloid plaques. Ann. Neurol. 1997, 42, 909–918. [Google Scholar] [CrossRef] [PubMed]
  124. Darvesh, S.; Cash, M.K.; Reid, G.A.; Martin, E.; Mitnitski, A.; Geula, C. Butyrylcholinesterase is associated with β-amyloid plaques in the transgenic APPSWE/PSEN1dE9 mouse model of Alzheimer Disease. J. Neuropathol. Exp. Neurol. 2012, 71, 2–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Hernández-Aquino, E.; Muriel, P. Naringenin and the liver. In Liver Pathophysiology; Elsevier: Amsterdam, The Netherlands, 2017; pp. 633–651. [Google Scholar]
  126. Hernández-Aquino, E.; Muriel, P. Beneficial effects of naringenin in liver diseases: Molecular mechanisms. World J. Gastroenterol. 2018, 24, 1679–1707. [Google Scholar] [CrossRef] [PubMed]
  127. Al-Ghamdi, N.A.M.; Virk, P.; Hendi, A.; Awad, M.; Elobeid, M. Antioxidant potential of bulk and nanoparticles of naringenin against cadmium-induced oxidative stress in Nile tilapia, Oreochromis niloticus. Green Process. Synth. 2021, 10, 392–402. [Google Scholar] [CrossRef]
  128. Zhang, N.; Hu, Z.; Zhang, Z.; Liu, G.; Wang, Y.; Ren, Y.; Wu, X.; Geng, F. Protective Role of Naringenin Against A 25-35-Caused Damage via ER and PI3K/Akt-Mediated Pathways. Cell. Mol. Neurobiol. 2018, 38, 549–557. [Google Scholar] [CrossRef]
  129. Liu, P.; Cheng, H.; Roberts, T.M.; Zhao, J.J. Targeting the phosphoinositide 3-kinase pathway in cancer. Nat. Rev. Drug Discov. 2009, 8, 627–644. [Google Scholar] [CrossRef] [Green Version]
  130. Brunet, A.; Datta, S.R.; Greenberg, M.E. Transcription-dependent and-independent control of neuronal survival by the PI3K-Akt signaling pathway. Curr. Opin. Neurobiol. 2001, 11, 297–305. [Google Scholar] [CrossRef]
  131. Hua, F.Z.; Ying, J.; Zhang, J.; Wang, X.F.; Hu, Y.H.; Liang, Y.P.; Liu, Q.; Xu, G.H. Naringenin pre-treatment inhibits neuroapoptosis and ameliorates cognitive impairment in rats exposed to isoflurane anesthesia by regulating the PI3/Akt/PTEN signalling pathway and suppressing NF-κB-mediated inflammation. Int. J. Mol. Med. 2016, 38, 1271–1280. [Google Scholar] [CrossRef] [Green Version]
  132. Vafeiadou, K.; Vauzour, D.; Lee, H.Y.; Rodriguez-Mateos, A.; Williams, R.J.; Spencer, J.P. The citrus flavanon naringenin inhibits inflammatory signalling in glial cells and protects against neuroinflammatory injury. Arch. Biochem. Biophys. 2009, 484, 100–109. [Google Scholar] [CrossRef]
  133. Santa-Cecília, F.V.; Socias, B.; Ouidja, M.O.; Sepulveda-Diaz, J.E.; Acuna, L.; Silva, R.L.; Michel, P.P.; Del-Bel, E.; Cunha, T.M.; Raisman-Vozari, R. Doxycycline suppresses microglial activation by inhibiting the p38 MAPK and NF-kB signaling pathways. Neurotox. Res. 2016, 29, 447–459. [Google Scholar] [CrossRef] [PubMed]
  134. Sarubbo, F.; Ramis, M.; Kienzer, C.; Aparicio, S.; Esteban, S.; Miralles, A.; Moranta, D.J. Chronic silymarin, quercetin and naringenin treatments increase monoamines synthesis and hippocampal Sirt1 levels improving cognition in aged rats. J. Neuroimmune Pharmacol. 2018, 13, 24–38. [Google Scholar] [CrossRef] [PubMed]
  135. Gao, J.; Wang, W.-Y.; Mao, Y.-W.; Grä, J.; Guan, J.-S.; Pan, L.; Mak, G.; Kim, D.; Su, S.C.; Tsai, L.H. A novel pathway regulates memory and plasticity via SIRT1 and miR-134. Nature 2010, 466, 1105–1109. [Google Scholar] [CrossRef] [Green Version]
  136. Zhang, B.; Wei, Y.Z.; Wang, G.Q.; Li, D.D.; Shi, J.S.; Zhang, F. Targeting MAPK pathways by naringenin modulates microglia M1/M2 polarization in lipopolysaccharide-stimulated cultures. Front. Cell Neurosci. 2019, 12, 531. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Mani, S.; Sekar, S.; Barathidasan, R.; Manivasagam, T.; Thenmozhi, A.J.; Sevanan, M.; Chidambaram, S.B.; Essa, M.M.; Guillemin, G.J.; Sakharkar, M.K. Naringenin decreases synuclein expression and neuroinflammation in MPTP-induced Parkinson’s disease model in mice. Neurotox. Res. 2018, 33, 656–670. [Google Scholar] [CrossRef]
  138. Sugumar, M.; Sevanan, M.; Sekar, S. Neuroprotective effect of naringenin against MPTP-induced oxidative stress. Int. J. Neurosci. 2019, 129, 534–539. [Google Scholar] [CrossRef]
  139. Zaki, H.F.; Abd-El-Fattah, M.A.; Attia, A.S. Naringenin protects against scopolamine-induced dementia in rats. Bull. Fac. Pharm. Cairo Univ. 2014, 52, 15–25. [Google Scholar] [CrossRef] [Green Version]
  140. Chtourou, Y.; Fetoui, H.; Gdoura, R. Protective effects of naringenin on iron-overload-induced cerebral cortex neurotoxicity correlated with oxidative stress. Biol. Trace Elem. Res. 2014, 158, 376–383. [Google Scholar] [CrossRef]
  141. Rady, I.; Mohamed, H.; Rady, M.; Siddiqui, I.A.; Mukhtar, H. Cancer preventive and therapeutic effects of EGCG, the major polyphenol in green tea, Egypt. J. Basic Appl. Sci. 2019, 5, 1–23. [Google Scholar]
  142. Khan, N.; Afaq, F.; Saleem, M.; Ahmad, N.; Mukhtar, H. Targeting multiple signaling pathways by green tea polyphenol (−)-epigallocatechin-3-gallate. Cancer Res. 2006, 66, 2500–2505. [Google Scholar] [CrossRef] [Green Version]
  143. Singh, M.; Arseneault, M.; Sanderson, T.; Murthy, V.; Ramassamy, C. Challenges for research on polyphenols from foods in Alzheimer’s disease: Bioavailability, metabolism, and cellular and molecular mechanisms. J Agric Food Chem. 2008, 56, 4855–4873. [Google Scholar] [CrossRef] [PubMed]
  144. Chang, X.; Rong, C.; Chen, Y.; Yang, C.; Hu, Q.; Mo, Y.; Zhang, C.; Gu, X.; Zhang, L.; He, W.; et al. (−)-Epigallocatechin-3-gallate attenuates cognitive deterioration in Alzheimer’s disease model mice by upregulating neprilysin expression. Exp. Cell Res. 2015, 334, 136–145. [Google Scholar] [CrossRef] [PubMed]
  145. Ehrnhoefer, D.E.; Bieschke, J.; Boeddrich, A.; Herbst, M.; Masino, L.; Lurz, R.; Engemann, S.; Pastore, A.; Wanker, E.E. EGCG redirects amyloidogenic polypeptides into unstructured, off-pathway oligomers. Nat. Struct. Mol. Biol. 2008, 15, 558–566. [Google Scholar] [CrossRef] [PubMed]
  146. Rezai-Zadeh, K.; Shytle, D.; Sun, N.; Mori, T.; Hou, H.; Jeanniton, D.; Ehrhart, J.; Townsend, K.; Zeng, J.; Morgan, D.; et al. Green tea epigallocatechin-3-gallate (EGCG) modulates amyloid precursor protein cleavage and reduces cerebral amyloidosis in Alzheimer transgenic mice. J. Neurosci. 2005, 25, 8807–8814. [Google Scholar] [CrossRef] [Green Version]
  147. Lee, Y.K.; Yuk, D.Y.; Lee, J.W.; Lee, S.Y.; Ha, T.Y.; Oh, K.W.; Yun, Y.P.; Hong, J.T. (-)-Epigallocatechin-3-gallate prevents lipopolysaccharide-induced elevation of beta-amyloid generation and memory deficiency. Brain Res. 2009, 1250, 164–174. [Google Scholar] [CrossRef]
  148. Sonawane, S.K.; Chidambaram, H.; Boral, D.; Gorantla, N.V.; Balmik, A.A.; Dangi, A.; Ramasamy, S.; Marelli, U.K.; Chinnathambi, S. EGCG impedes human Tau aggregation and interacts with Tau. Sci. Rep. 2020, 10, 12579. [Google Scholar] [CrossRef]
  149. Lin, C.L.; Chen, T.F.; Chiu, M.J.; Way, T.D.; Lin, J.K. Epigallocatechin gallate (EGCG) suppresses beta-amyloid-induced neurotoxicity through inhibiting c-Abl/FE65 nuclear translocation and GSK3 beta activation. Neurobiol. Aging 2009, 30, 81–92. [Google Scholar] [CrossRef]
  150. Wei, J.C.C.; Huang, H.C.; Chen, W.J.; Huang, C.N.; Peng, C.H.; Lin, C.L. Epigallocatechin gallate attenuates amyloid β-induced inflammation and neurotoxicity in EOC 13.31 microglia. Eur. J. Pharmacol. 2016, 770, 16–24. [Google Scholar]
  151. Chi, X.; Ma, X.; Li, Z.; Zhang, Y.; Wang, Y.; Yuan, L.; Wu, Y.; Xu, W.; Hu, S. Protective Effect of Epigallocatechin-3-Gallate in Hydrogen Peroxide-Induced Oxidative Damage in Chicken Lymphocytes. Oxidative Med. Cell. Longev. 2020, 2020, 7386239. [Google Scholar] [CrossRef]
  152. Ali, B.; Jamal, Q.M.; Shams, S.; Al-Wabel, N.A.; Siddiqui, M.U.; Alzohairy, M.A.; Al Karaawi, M.A.; Kesari, K.K.; Mushtaq, G.; Kamal, M.A. In silico analysis of green tea polyphenols as inhibitors of AChE and BChE enzymes in Alzheimer’s disease treatment. CNS Neurol. Disord. Drug Targets 2016, 15, 624–628. [Google Scholar] [CrossRef]
  153. Nardini, M.; Garaguso, I. Characterization of bioactive compounds and antioxidant activity of fruit beers. Food Chem. 2020, 305, 125437. [Google Scholar] [CrossRef] [PubMed]
  154. Imran, M.; Saeed, F.; Hussain, G.; Imran, A.; Mehmood, Z.; Gondal, T.A.; El-Ghorab, A.; Ahmad, I.; Pezzani, R.; Arshad, M.U.; et al. Myricetin: A comprehensive review on its biological potentials. Food Sci. Nutr. 2021, 9, 5854–5868. [Google Scholar] [CrossRef] [PubMed]
  155. Häkkinen, S.H.; Kärenlampi, S.O.; Heinonen, I.M.; Mykkänen, H.M.; Törrönen, A.R. Content of the flavonols quercetin, myricetin, and kaempferol in 25 edible berries. J. Agric. Food. Chem. 1999, 47, 2274–2279. [Google Scholar] [CrossRef] [PubMed]
  156. Ong, K.C.; Khoo, H.E. Biological effects of myricetin. Gen. Pharmacol. 1997, 29, 121–126. [Google Scholar] [CrossRef] [PubMed]
  157. Semwal, D.K.; Semwal, R.B.; Combrinck, S.; Viljoen, A. Myricetin: A dietary molecule with diverse biological activities. Nutrients 2016, 8, 90. [Google Scholar] [CrossRef] [Green Version]
  158. Walker, E.H.; Pacold, M.E.; Perisic, O.; Stephens, L.; Hawkins, P.T.; Wymann, M.P.; Williams, R.L. Structural determinants of phosphoinositide 3-kinase inhibition by wortmannin, LY294002, quercetin, myricetin, and staurosporine. Mol. Cell. 2000, 6, 909–919. [Google Scholar] [CrossRef]
  159. Taheri, Y.; Suleria, H.A.R.; Martins, N.; Sytar, O.; Beyatli, A.; Yeskaliyeva, B.; Seitimova, G.; Salehi, B.; Semwal, P.; Painuli, S.; et al. Myricetin bioactive effects: Moving from preclinical evidence to potential clinical applications. BMC Complement. Med. Ther. 2020, 20, 241. [Google Scholar] [CrossRef]
  160. Xu, C.; Liu, Y.L.; Gao, Z.W.; Jiang, H.M.; Xu, C.J.; Li, X. Pharmacological activities of myricetin and its glycosides. Zhongguo Zhong Yao Za Zhi 2020, 45, 3575–3583. [Google Scholar]
  161. Ramezani, M.; Darbandi, N.; Khodagholi, F.; Hashemi, A. Myricetin protects hippocampal CA3 pyramidal neurons and improves learning and memory impairments in rats with Alzheimer’s disease. Neural. Regen. Res. 2016, 11, 1976. [Google Scholar] [CrossRef]
  162. Fiori, J.; Naldi, M.; Bartolini, M.; Andrisano, V. Disclosure of a fundamental clue for the elucidation of the myricetin mechanism of action as amyloid aggregation inhibitor by mass spectrometry. Electrophoresis 2012, 33, 3380–3386. [Google Scholar] [CrossRef]
  163. Andarzi Gargari, S.; Barzegar, A.; Tarinejad, A. The role of phenolic OH groups of flavonoid compounds with H-bond formation ability to suppress amyloid mature fibrils by destabilizing β-sheet conformation of monomeric Aβ17-42. PLoS ONE 2018, 13, e0199541. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Hirohata, M.; Hasegawa, K.; Tsutsumi-Yasuhara, S.; Ohhashi, Y.; Ookoshi, T.; Ono, K.; Yamada, M.; Naiki, H. The anti-amyloidogenic effect is exerted against Alzheimer’s β-amyloid fibrils in vitro by preferential and reversible binding of flavonoids to the amyloid fibril structure? Biochemistry 2007, 46, 1888–1899. [Google Scholar] [CrossRef] [PubMed]
  165. Berhanu, W.M.; Masunov, A.E. Natural polyphenols as inhibitors of amyloid aggregation. Molecular dynamics study of GNNQQNY heptapeptide decamer. Biophys. Chem. 2010, 149, 12–21. [Google Scholar] [CrossRef]
  166. Naldi, M.; Fiori, J.; Pistolozzi, M.; Drake, A.F.; Bertucci, C.; Wu, R.; Mlynarczyk, K.; Filipek, S.; De Simone, A.; Andrisano, V. Amyloid beta-peptide 25-35 selfassembly and its inhibition: A model undecapeptide system to gain atomistic and secondary structure details of the Alzheimer’s disease process and treatment. ACS Chem. Neurosci. 2012, 3, 952–962. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Shimmyo, Y.; Kihara, T.; Akaike, A.; Niidome, T.; Sugimoto, H. Multifunction of myricetin on A beta: Neuroprotection via a conformational change of A beta and reduction of A beta via the interference of secretases. J. Neurosci. Res. 2008, 86, 368–377. [Google Scholar] [CrossRef] [PubMed]
  168. Chakraborty, S.; Kumar, S.; Basu, S. Conformational transition in the substrate binding domain of β-secretase exploited by NMA and its implication in inhibitor recognition: BACE1-myricetin a case study. Neurochem. Int. 2011, 58, 914–923. [Google Scholar] [CrossRef] [PubMed]
  169. Barzegar, A. Antioxidant activity of polyphenolic myricetin in vitro cell free and cell-based systems. Mol. Biol. Res. Commun. 2016, 5, 87–95. [Google Scholar] [PubMed]
  170. Zhao, J.; Hong, T.; Dong, M.; Meng, Y.; Mu, J. Protective Effect of Myricetin in Dextran Sulphate Sodium-Induced Murine Ulcerative Colitis. Mol. Med. Rep. 2013, 7, 565–570. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Salimi, A.; Jamali, Z.; Shabani, M. Antioxidant Potential and Inhibition of Mitochondrial Permeability Transition Pore by Myricetin Reduces Aluminium Phosphide-Induced Cytotoxicity and Mitochondrial Impairments. Front. Pharmacol. 2021, 12, 719081. [Google Scholar] [CrossRef]
  172. Kenouche, S.; Sandoval-Yañez, C.; Martínez-Araya, J.I. The antioxidant capacity of myricetin. A molecular electrostatic potential analysis based on DFT calculations. Chem. Phys. Lett. 2022, 801, 139708. [Google Scholar] [CrossRef]
  173. Zhang, K.; Ma, Z.; Wang, J.; Xie, A.; Xie, J. Myricetin attenuated MPP+-induced cytotoxicity by anti-oxidation and inhibition of MKK4 and JNK activation in MES23. 5 cells. Neuropharmacology 2011, 61, 329–335. [Google Scholar] [CrossRef] [PubMed]
  174. Mendes, V.; Vilaça, R.; de Freitas, V.; Ferreira, P.M.; Mateus, N.; Costa, V. Effect of myricetin, pyrogallol, and phloroglucinol on yeast resistance to oxidative stress. Oxidat. Med. Cell. Long. 2015, 2015, 782504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Mira, L.; Fernandez, M.T.; Santos, M.; Rocha, R.; Florêncio, M.H.; Jennings, K.R. Interactions of flavonoids with iron and copper ions: A mechanism for their antioxidant activity. Free Radic. Res. 2002, 36, 1199–1208. [Google Scholar] [CrossRef] [PubMed]
  176. Agraharam, G.; Girigoswami, A.; Girigoswami, K. Myricetin: A Multifunctional Flavonol in Biomedicine. Curr. Pharmacol. Rep. 2022, 8, 48–61. [Google Scholar] [CrossRef] [PubMed]
  177. Korshavn, K.J.; Jang, M.; Kwak, Y.J.; Kochi, A.; Vertuani, S.; Bhunia, A.; Manfredini, S.; Ramamoorthy, A.; Lim, M.H. Reactivity of Metal-Free and Metal-Associated Amyloid-β with Glycosylated Polyphenols and Their Esterified Derivatives. Sci. Rep. 2015, 5, 17842. [Google Scholar] [CrossRef] [PubMed]
  178. Jang, J.H.; Lee, S.H.; Jung, K.; Yoo, H.; Park, G. Inhibitory Effects of Myricetin on Lipopolysaccharide-Induced Neuroinflammation. Brain Sci. 2020, 10, 32. [Google Scholar] [CrossRef] [Green Version]
  179. Kou, X.; Liu, X.; Chen, X.; Li, J.; Yang, X.; Fan, J.; Yang, Y.; Chen, N. Ampelopsin attenuates brain aging of D-gal-induced rats through miR-34a-mediated SIRT1/mTOR signal pathway. Oncotarget 2016, 7, 74484–74495. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Patel, D.K.; Patel, K. P-MD005. Neuroprotective effects of gossypetin in alzheimer’s disease: Therapeutic approaches to evaluate the acetylcholinesterase and butyl cholinesterase inhibitory potential. Clin. Neurophysiol. 2021, 132, e97–e98. [Google Scholar] [CrossRef]
  181. Francis, A.R.; Shetty, T.K.; Bhattacharya, R.K. Modulating effect of plant flavonoids on the mutagenicity of N-methyl-N′-nitro-N-nitrosoguanidine. Carcinogenesis 1989, 10, 1953–1955. [Google Scholar] [CrossRef]
  182. Miceli, N.; Trovato, A.; Dugo, P.; Cacciola, F.; Donato, P.; Marino, A.; Bellinghieri, V.; La Barbera, T.M.; Güvenç, A.; Taviano, M.F. Comparative analysis of flavonoid profile, antioxidant and antimicrobial activity of the berries of Juniperus communis L. var. communis and Juniperus communis L. var. saxatilis Pall. from Turkey. J. Agric. Food Chem. 2009, 57, 6570–6577. [Google Scholar] [CrossRef]
  183. Chen, J.H.; Tsai, C.W.; Wang, C.P.; Lin, H.H. Anti-atherosclerotic potential of gossypetin via inhibiting LDL oxidation and foam cell formation. Toxicol. Appl. Pharmacol. 2013, 272, 313–324. [Google Scholar] [CrossRef] [PubMed]
  184. Lin, H.H. In Vitro and In Vivo Atheroprotective Effects of Gossypetin against Endothelial Cell Injury by Induction of Autophagy. Chem. Res. Toxicol. 2015, 28, 202–215. [Google Scholar] [CrossRef] [PubMed]
  185. Hillhouse, B.; Ming, D.S.; French, C.; Towers, G.H. Acetylcholine Esterase Inhibitors in Rhodiola rosea. Pharm. Biol. 2004, 42, 68–72. [Google Scholar] [CrossRef]
  186. Zubik, L.; Meydani, M. Bioavailability of soybean isoflavones from aglycone and glucoside forms in American women. Am. J. Clin. Nutr. 2003, 77, 1459–1465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Dixon, R.A.; Ferreira, D. Genistein. Phytochemistry 2002, 60, 205–211. [Google Scholar] [CrossRef]
  188. Bagheri, M.; Joghataei, M.T.; Mohseni, S.; Roghani, M. Genistein ameliorates learning and memory deficits in amyloid β (1–40) rat model of Alzheimer’s disease. Neurobiol. Learn. Mem. 2011, 95, 270–276. [Google Scholar] [CrossRef] [Green Version]
  189. Safahani, M.; Amani, R.; Aligholi, H.; Sarkaki, A.; Badavi, M.; Zand Moghaddam, A.; Haghighizadeh, M.H. Effect of different doses of soy isoflavones on spatial learning and memory in ovariectomized rats. Basic Clin. Neurosci. 2011, 2, 12–18. [Google Scholar]
  190. Zeng, H.; Chen, Q.; Zhao, B. Genistein ameliorates β-amyloid peptide (25–35) -induced hippocampal neuronal apoptosis. Free Radic. Bio. Med. 2004, 36, 180–188. [Google Scholar] [CrossRef]
  191. Petry, F.D.S.; Hoppe, J.B.; Klein, C.P.; Dos Santos, B.G.; Hözer, R.M.; Bifi, F.; Matté, C.; Salbego, C.G.; Trindade, V.M.T. Genistein attenuates amyloid-beta-induced cognitive impairment in rats by modulation of hippocampal synaptotoxicity and hyperphosphorylation of Tau. J. Nutr. Biochem. 2021, 87, 108525. [Google Scholar] [CrossRef]
  192. Seong, S.H.; Kim, B.R.; Cho, M.L.; Kim, T.S.; Im, S.; Han, S.; Jeong, J.W.; Jung, H.A.; Choi, J.S. Phytoestrogen Coumestrol Selectively Inhibits Monoamine Oxidase-A and Amyloid β Self-Aggregation. Nutrients 2022, 14, 3822. [Google Scholar] [CrossRef]
  193. Okumura, N.; Yoshida, H.; Nishimura, Y.; Murakami, M.; Kitagishi, Y.; Matsuda, S. Genistein downregulates presenilin 1 and ubiquilin 1 expression. Mol. Med. Rep. 2012, 5, 559–561. [Google Scholar]
  194. Bonet-Costa, V.; Herranz-Perez, V.; Blanco-Gandia, M.; Mas-Bargues, C.; Ingles, M.; Garcia-Tarraga, P.; Rodriguez-Arias, M.; Minarro, J.; Borras, C.; Garcia-Verdugo, J.M.; et al. Clearing amyloid-beta through PPARgamma/ApoE activation by genistein is a treatment of experimental Alzheimer’s disease. J. Alzheimer's Dis. 2016, 51, 701–711. [Google Scholar] [CrossRef] [PubMed]
  195. Youn, K.; Park, J.H.; Lee, S.; Lee, J.; Yun, E.Y.; Jeong, W.S.; Jun, M. BACE1 inhibition by genistein: Biological evaluation, kinetic analysis, and molecular docking simulation. J. Med. Food 2018, 21, 416–420. [Google Scholar] [CrossRef] [PubMed]
  196. Liao, W.; Jin, G.; Zhao, M.; Yang, H. The effect of genistein on the content and activity of α- and β-secretase and protein kinase C in Aβ-injured hippocampal neurons. Basic Clin. Pharmacol. Toxicol. 2013, 112, 182–185. [Google Scholar] [CrossRef] [PubMed]
  197. Kładna, A.; Berczy’nski, P.; Kruk, I.; Piechowska, T.; Aboul-Enein, H.Y. Studies on the antioxidant properties of some phytoestrogens. Luminescence 2016, 31, 1201–1206. [Google Scholar] [CrossRef]
  198. Suzuki, K.; Koike, H.; Matsui, H.; Ono, Y.; Hasumi, M.; Nakazato, H.; Okugi, H.; Sekine, Y.; Oki, K.; Ito, K.; et al. Genistein, a soy isoflavone, induces glutathione peroxidase in the human prostate cancer cell lines LNCaP and PC-3. Int. J. Cancer 2002, 99, 846–852. [Google Scholar] [CrossRef]
  199. Park, C.E.; Yun, H.; Lee, E.B.; Min, B.I.; Bae, H.; Choe, W.; Kang, I.; Kim, S.S.; Ha, J. The antioxidant effects of genistein are associated with AMP-activated protein kinase activation and PTEN induction in prostate cancer cells. J. Med. Food. 2010, 13, 815–820. [Google Scholar] [CrossRef]
  200. Borrás, C.; Gambini, J.; Gómez-Cabrera, M.C.; Sastre, J.; Pallardó, F.V.; Mann, G.E.; Viña, J. Genistein, a soy isoflavone, up-regulates expression of antioxidant genes: Involvement of estrogen receptors, ERK1/2, and NFkappaB. FASEB J. 2006, 20, 2136–2138. [Google Scholar] [CrossRef]
  201. Verdrengh, M.; Jonsson, I.M.; Holmdahl, R.; Tarkowski, A. Genistein as an antiinflammatory agent. Inflamm. Res. 2003, 52, 341–346. [Google Scholar] [CrossRef]
  202. Blay, M.; Espinel, A.E.; Delgado, M.A.; Baiges, I.; Bladé, C.; Arola, L.; Salvadó, J. Isoflavone effect on gene expression profile and biomarkers of inflammation. J. Pharm. Biomed. Anal. 2010, 51, 382–390. [Google Scholar] [CrossRef]
  203. Fang, J.; Wu, P.; Yang, R.; Gao, L.; Li, C.; Wang, D.; Wu, S.; Liu, A.L.; Du, G.H. Inhibition of acetylcholinesterase by two genistein derivatives: Kinetic analysis, molecular docking and molecular dynamics simulation. Acta Pharm. Sin. B 2014, 4, 430–437. [Google Scholar] [CrossRef] [Green Version]
  204. McKay, D.L.; Blumberg, J.B. A review of the bioactivity and potential health benefits of chamomile tea (Matricaria recutita L.). Phytother. Res. 2006, 20, 519–530. [Google Scholar] [CrossRef]
  205. Alsadat, A.M.; Nikbakht, F.; Nia, H.H.; Golab, F.; Khadem, Y.; Barati, M.; Vazifekhah, S. GSK-3β as a target for apigenin-induced neuroprotection against Aβ 25–35 in a rat model of Alzheimer’s disease. Neuropeptides 2021, 90, 102200. [Google Scholar] [CrossRef]
  206. Zhang, X.; Wang, G.; Gurley, E.C.; Zhou, H. Flavonoid apigenin inhibits lipopolysaccharide-induced inflammatory response through multiple mechanisms in macrophages. PLoS ONE 2014, 9, e107072. [Google Scholar] [CrossRef] [Green Version]
  207. Tong, L.; Balazs, R.; Soiampornkul, R.; Thangnipon, W.; Cotman, C.W. Interleukin-1 b impairs brain derived neurotrophic factor-induced signal transduction. Neurobiol. Aging 2008, 29, 1380–1393. [Google Scholar] [CrossRef] [Green Version]
  208. Dourado, N.S.; Souza, C.D.S.; de Almeida, M.M.A.; Bispo da Silva, A.; Dos Santos, B.L.; Silva, V.D.A.; De Assis, A.M.; da Silva, J.S.; Souza, D.O.; Costa, M.F.D.; et al. Neuroimmunomodulatory and neuroprotective effects of the flavonoid apigenin in in vitro models of neuroinflammation associated with Alzheimer’s disease. Front. Aging Neurosci. 2020, 12, 119. [Google Scholar] [CrossRef]
  209. Wang, N.; Yi, W.J.; Tan, L.; Zhang, J.H.; Xu, J.; Chen, Y.; Qin, M.; Yu, S.; Guan, J.; Zhang, R. Apigenin attenuates streptozotocin-induced pancreatic β cell damage by its protective effects on cellular antioxidant defense. In Vitro Cell Dev. Biol. Anim. 2017, 53, 554–563. [Google Scholar] [CrossRef]
  210. Sánchez-Marzo, N.; Pérez-Sánchez, A.; Ruiz-Torres, V.; Martínez-Tébar, A.; Castillo, J.; Herranz-López, M.; Barrajón-Catalán, E. Antioxidant and Photoprotective Activity of Apigenin and its Potassium Salt Derivative in Human Keratinocytes and Absorption in Caco-2 Cell Monolayers. Int. J. Mol. Sci. 2019, 20, 2148. [Google Scholar] [CrossRef] [Green Version]
  211. Yamakawa, M.Y.; Uchino, K.; Watanabe, Y.; Adachi, T.; Nakanishi, M.; Ichino, H.; Hongo, K.; Mizobata, T.; Kobayashi, S.; Nakashima, K.; et al. Anthocyanin suppresses the toxicity of Aβ deposits through diversion of molecular forms in in vitro and in vivo models of Alzheimer’s disease. Nutr. Neurosci. 2016, 19, 32–42. [Google Scholar] [CrossRef]
  212. Tarozzi, A.; Morroni, F.; Merlicco, A.; Bolondi, C.; Teti, G.; Falconi, M.; Cantelli-Forti, G.; Hrelia, P. Neuroprotective effects of cyanidin 3-O-glucopyranoside on amyloid beta (25–35) oligomer-induced toxicity. Neurosci. Lett. 2010, 473, 72–76. [Google Scholar] [CrossRef]
  213. Shih, P.H.; Wu, C.H.; Yeh, C.T.; Yen, G.C. Protective effects of anthocyanins against amyloid β-peptide-induced damage in neuro-2A, Cells. J. Agric. Food Chem. 2011, 59, 1683–1689. [Google Scholar] [CrossRef] [PubMed]
  214. Strathearn, K.E.; Yousef, G.G.; Grace, M.H.; Roy, S.L.; Tambe, M.A.; Ferruzzi, M.G.; Wu, Q.L.; Simon, J.E.; Lila, M.A.; Rochet, J.C. Neuroprotective effects of anthocyanin- and proanthocyanidin-rich extracts in cellular models of Parkinson’s disease. Brain Res. 2014, 1555, 60–77. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Wang, W.; Zhu, G.; Wang, Y.; Li, W.; Yi, S.; Wang, K.; Fan, L.; Tang, J.; Chen, R. Multi-Omics Integration in Mice with Parkinson’s Disease and the Intervention Effect of Cyanidin-3-O-Glucoside. Front. Aging Neurosci. 2022, 14, 877078. [Google Scholar] [CrossRef] [PubMed]
  216. Pike, C.J.; Walencewicz-Wasserman, A.J.; Kosmoski, J.; Cribbs, D.H.; Glabe, C.G.; Cotman, C.W. Structure-activity analyses of beta-amyloid peptides: Contributions of the beta 25–35 region to aggregation and neurotoxicity. J. Neurochem. 1995, 64, 253–265. [Google Scholar] [CrossRef]
  217. Thummayot, S.; Tocharus, C.; Pinkaew, D.; Viwatpinyo, K.; Sringarm, K.; Tocharus, J. Neuroprotective effect of purple rice extract and its constituent against amyloid beta induced neuronal cell death in SK-N-SH cells. Neurotoxicology 2014, 45, 149–158. [Google Scholar] [CrossRef]
  218. Thummayot, S.; Tocharus, C.; Suksamrarn, A.; Tocharus, J. Neuroprotective effects of cyanidin against Ab-induced oxidative and ER stress in SK-N-SH cells. Neurochem. Int. 2016, 101, 15–21. [Google Scholar] [CrossRef]
  219. Behl, C.; Moosmann, B. Antioxidant neuroprotection in Alzheimer’s disease as preventive and therapeutic approach. Free Radic. Biol. Med. 2002, 33, 182–191. [Google Scholar] [CrossRef]
  220. Essa, M.M.; Vijayan, R.K.; Castellano-Gonzalez, G.; Memon, M.A.; Braidy, N.; Guillemin, G.J. Neuroprotective effect of natural products against Alzheimer’s disease. Neurochem. Res. 2012, 37, 1829–1842. [Google Scholar] [CrossRef]
  221. Leong, P.K.; Chiu, P.Y.; Chen, N.; Leung, H.; Ko, K.M. Schisandrin B elicits a glutathione antioxidant response and protects against apoptosis via the redox-sensitive ERK/Nrf2 pathway in AML12 hepatocytes. Free Radic. Res. 2011, 45, 483–495. [Google Scholar] [CrossRef]
  222. Kaewmool, C.; Udomruk, S.; Phitak, T.; Pothacharoen, P.; Kongtawelert, P. Cyanidin-3-O-Glucoside Protects PC12 Cells Against Neuronal Apoptosis Mediated by LPS-Stimulated BV2 Microglial Activation. Neurotox Res. 2020, 37, 111–125. [Google Scholar] [CrossRef]
  223. ClinicalTrials.gov Identifier: NCT05269173; Efficacy and Safety of Flos Gossypii Flavonoids Tablet in the Treatment of Alzheimer’s Disease—Full Text View—ClinicalTrials.gov. Available online: https://www.clinicaltrials.gov/ct2/show/NCT05269173 (accessed on 17 October 2022).
Figure 1. Schematic representation of the main features of Alzheimer’s disease. ( increase; decrease).
Figure 1. Schematic representation of the main features of Alzheimer’s disease. ( increase; decrease).
Ijms 23 14835 g001
Figure 2. Schematic representation of the basic structures of flavonoid subclasses.
Figure 2. Schematic representation of the basic structures of flavonoid subclasses.
Ijms 23 14835 g002
Figure 3. Schematic effects of described flavonoids on molecular targets of AD.
Figure 3. Schematic effects of described flavonoids on molecular targets of AD.
Ijms 23 14835 g003
Figure 4. Schematic effects of quercetin on molecular targets of AD.
Figure 4. Schematic effects of quercetin on molecular targets of AD.
Ijms 23 14835 g004
Figure 5. Schematic effects of naringenin on molecular targets of AD.
Figure 5. Schematic effects of naringenin on molecular targets of AD.
Ijms 23 14835 g005
Figure 6. Schematic effects of epigallocatechin-3-gallate on molecular targets of AD.
Figure 6. Schematic effects of epigallocatechin-3-gallate on molecular targets of AD.
Ijms 23 14835 g006
Figure 7. Schematic effects of myricetin on molecular targets of AD.
Figure 7. Schematic effects of myricetin on molecular targets of AD.
Ijms 23 14835 g007
Table 1. Preclinical studies of flavonoids and their neuroprotective role against Alzheimer’s disease.
Table 1. Preclinical studies of flavonoids and their neuroprotective role against Alzheimer’s disease.
Flavonoids Molecular TargetsModelDoseReferences
QuercetinRegulates MAPK signalingHT22 cells 5 μmol/L [103]
Decreases phosphorylation of Tau protein HT22 cells 5–10 μmol/L[104]
Reduces apoptosis and caspase 3 activationPC-12 cell line500 μM[108]
Restores antioxidant cellular defensesGerbilli’s CA1 pyramidal neurons, HepG2 cells, C57BL/6J mice20 mg/kg, 50 μM
1% quercetin diet
[113,114,115,116]
Inhibits AChE and BChE AChE (EC 3.1.1.7 Sigma)
BChE (EC 3.1.1.8, Sigma)
1 mg/mL[118,119]
NaringerinDecreases phosphorylation of Tau proteinPC12 cells400 μM[128,129,130]
Reduces apoptosis and caspase 3 activationRat pups50–100 mM[128,129,130,131]
Decreases the inflammatory pathway Male rats, glial cells20 mg/kg/day, 0.1–0.3 μmol/L[132,133,134,135]
Regulates the MAPK signaling pathway BV-2 microglial cell line100 mM[136]
Improves the antioxidant systemC57BL/6J mice 25–100 mg/kg[137,138,139,140]
Epigallocatechin-3-gallateReduces the accumulation of b amyloid mice P8 (SAMP8), SweAPP N2 a cells, mouse model, MC65 cells 5–15 mg/kg/day, 20 mM,
1–3 mg/kg,
5–20 μM
[144,145,146,147,148]
Restores antioxidant cellular defensesEOC 13.31 microglial cell line, chicken lymphocytes5–20 μM,
22.5–90 μM
[149,150]
MyricetinImproves learning and memory Rat models 5 or 10 mg/kg[160]
Decreases Ab aggregation [156,157,158,159,160] [162,163,164,165,166]
Regulates a and b secretase activityrat primary cortical neurons 10 μM[166]
Inhibits oxidative stressMurine models40–80 µM[158,169,170]
GossypetinInhibits lipid peroxidationMurine macrophage cell line J774A.1 1–1000 μM[182]
Fights against Atherosclerosisvascular smooth muscle cells 0.1–0.5 μM[183]
GenisteinReduces the production and deposition of Ab aggregates Rat model10 mg/kg[187,188,189]
Prevents Tau hyperphosphorylationRat model10 mg/kg[190]
Regulates a and b secretase activityRat hippocampal neurons0.375 µg/mL[194]
Prevents neuro-inflammation RAW 264.7 cell model20 μM[199]
ApigeninReduces Tau hyperphosphorylationRat model 50 mg/kg[203]
Inhibits the production of IL-6 and IL-1bHuman THP-1 monotypic cells25 µM [204,205,206]
CyanidinRegulates NF-κB and p38 MAPK signaling pathways LPS-stimulated BV2 microglia 2.5–10 mM[220]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Calderaro, A.; Patanè, G.T.; Tellone, E.; Barreca, D.; Ficarra, S.; Misiti, F.; Laganà, G. The Neuroprotective Potentiality of Flavonoids on Alzheimer’s Disease. Int. J. Mol. Sci. 2022, 23, 14835. https://doi.org/10.3390/ijms232314835

AMA Style

Calderaro A, Patanè GT, Tellone E, Barreca D, Ficarra S, Misiti F, Laganà G. The Neuroprotective Potentiality of Flavonoids on Alzheimer’s Disease. International Journal of Molecular Sciences. 2022; 23(23):14835. https://doi.org/10.3390/ijms232314835

Chicago/Turabian Style

Calderaro, Antonella, Giuseppe Tancredi Patanè, Ester Tellone, Davide Barreca, Silvana Ficarra, Francesco Misiti, and Giuseppina Laganà. 2022. "The Neuroprotective Potentiality of Flavonoids on Alzheimer’s Disease" International Journal of Molecular Sciences 23, no. 23: 14835. https://doi.org/10.3390/ijms232314835

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

Article Metrics

Back to TopTop