Next Article in Journal
Impact of Phenolic Acid Derivatives on the Oxidative Stability of β-Lactoglobulin-Stabilized Emulsions
Next Article in Special Issue
Ether Lipid-Mediated Antioxidant Defense in Alzheimer’s Disease
Previous Article in Journal
Evidence of Increased Oxidative Stress in the Placental Tissue of Women Who Suffered an Episode of Psychosis during Pregnancy
Previous Article in Special Issue
Therapeutics of Alzheimer’s Disease: Recent Developments
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Non-Enzymatic Antioxidants against Alzheimer’s Disease: Prevention, Diagnosis and Therapy

by
Angelica Varesi
1,†,
Lucrezia Irene Maria Campagnoli
2,†,
Adelaide Carrara
3,
Ilaria Pola
4,
Elena Floris
3,
Giovanni Ricevuti
4,
Salvatore Chirumbolo
5 and
Alessia Pascale
2,*
1
Department of Biology and Biotechnology, University of Pavia, 27100 Pavia, Italy
2
Department of Drug Sciences, Section of Pharmacology, University of Pavia, 27100 Pavia, Italy
3
Department of Internal Medicine and Therapeutics, University of Pavia, 27100 Pavia, Italy
4
Department of Drug Sciences, University of Pavia, 27100 Pavia, Italy
5
Department of Neurosciences, Biomedicine and Movement Sciences, University of Verona, 37129 Verona, Italy
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Antioxidants 2023, 12(1), 180; https://doi.org/10.3390/antiox12010180
Submission received: 22 December 2022 / Revised: 6 January 2023 / Accepted: 8 January 2023 / Published: 12 January 2023
(This article belongs to the Special Issue Oxidative Stress in Alzheimer's Disease)

Abstract

:
Alzheimer’s disease (AD) is a neurodegenerative disorder characterized by progressive memory loss and cognitive decline. Although substantial research has been conducted to elucidate the complex pathophysiology of AD, the therapeutic approach still has limited efficacy in clinical practice. Oxidative stress (OS) has been established as an early driver of several age-related diseases, including neurodegeneration. In AD, increased levels of reactive oxygen species mediate neuronal lipid, protein, and nucleic acid peroxidation, mitochondrial dysfunction, synaptic damage, and inflammation. Thus, the identification of novel antioxidant molecules capable of detecting, preventing, and counteracting AD onset and progression is of the utmost importance. However, although several studies have been published, comprehensive and up-to-date overviews of the principal anti-AD agents harboring antioxidant properties remain scarce. In this narrative review, we summarize the role of vitamins, minerals, flavonoids, non-flavonoids, mitochondria-targeting molecules, organosulfur compounds, and carotenoids as non-enzymatic antioxidants with AD diagnostic, preventative, and therapeutic potential, thereby offering insights into the relationship between OS and neurodegeneration.

1. Introduction

Alzheimer’s disease (AD) is a progressive neurodegenerative disorder characterized by cognitive decline, memory loss, inability to complete simple tasks, and mood alterations [1]. To date, approximately 50 million AD cases have been registered worldwide, but this number is predicted to rise exponentially within the next few years [2]. Although several medications have been proposed, no disease-modifying therapy has been proven to be effective when tested in clinical trials [3]. Molecularly, AD presents a multifactorial etiology, in which genetic and environmental factors contribute to the formation of senile amyloid plaques, composed of amyloid beta (Aβ) fibrils, and intracellular neurofibrillary tangles (primarily characterized by the abnormal accumulation of the microtubule-associated tau protein, which is hyperphosphorylated) in different brain areas [4]. Among the risk factors, oxidative stress (OS) turns out to be one of the primary causes of AD, playing a key role in its pathophysiology and progression [5,6,7,8,9,10,11,12,13,14]. OS normally occurs because of an imbalance between pro-oxidant and antioxidant status, which leads to an excessive accumulation of reactive oxygen species (ROS), in turn causing oxidative damage at the level of biological macromolecules, such as lipids, proteins, and nucleic acids [15]. Concerning AD, there is evidence that an immoderate amount of ROS can increase Aβ fibril production and aggregation, tau phosphorylation, and neuronal cell death, as well as trigger a whole series of events leading to neurodegeneration (such as mitochondrial dysfunction and glial cell activation) [16,17]. In this context, it has been reported that oxidative damage to enzymes involved in glucose metabolism causes impaired adenosine triphosphate(ATP) biosynthesis and limits the brain’s energy availability in mild cognitive impairment (MCI) and AD patients [18,19]. At the same time, the decreased expression of the main antioxidant enzymes catalase (CAT), superoxide dismutase (SOD), glutathione peroxidase (GPx), and glutathione reductase reported in cognitively impaired individuals prevents the activity of the proper detoxification machinery [19,20,21]. This condition of oxidative imbalance, coupled with the consequent overexpression of nuclear factor kappa-light-chain-enhancer (NF-kB) and the release of various inflammatory mediators (i.e., interleukin 1 beta (IL-1β), IL-6, tumor necrosis factor alpha (TNF-α), and transforming growth factor beta (TGF-β)), concur to establish an age-related pro-inflammatory microenvironment that favors the onset of neurodegenerative disorders [21,22]. Once established, treatments aimed at curing the disease are often unsuccessful due to poor blood–brain barrier (BBB) penetration and the scarce bioavailability of the proposed drugs [21]. Thus, considering the harmful role of OS in AD, preventing the excessive formation of ROS may represent a useful potential strategy to counteract the onset and progression of this disorder, prior to its establishment. In this regard, the present review summarizes the preventative, diagnostic and therapeutic potential of several classes of non-enzymatic antioxidants, including carotenoids, vitamins, flavonoids, non-flavonoids, organosulfur compounds, mitochondria-targeted antioxidants, and minerals, with the aim of highlighting early biomarkers and promising preventative treatments for this devastating neurodegenerative disease (Figure 1).

2. Results

2.1. Carotenoids

Carotenoids are colorful pigments that are found in fruits, vegetables, and seaweeds, with important anti-inflammatory, antioxidant, and anti-apoptotic activity [23,24]. Chemically, carotenoids are composed of a polyene chain, enriched with conjugated double-carbon bonds, which are the basis of their redox potential [24]. Among the enormous variety of known carotenoids, α-carotene, β-carotene, lutein, zeaxanthin, lycopene and β-cryptoxanthin remain the most characterized and studied for their involvement in different pathologies and conditions [25,26]. Over the past few years, the discovery of their enrichment in certain brain areas [27] has led to the investigation of their diagnostic and therapeutic potential within the context of neurodegeneration [23,28]. Generally, low levels of blood carotenoids are detected among AD patients when compared to healthy controls [29,30], and plant-based diets are considered neuroprotective [31]. Among them, circulating levels of lutein, zeaxanthin, and lycopene are the most predictive of AD development, severity, and mortality among non-demented individuals, even when accounting for age, sex, genetic background, social status, and lifestyle, with a great potential for future application as prodromal AD biomarkers [31,32,33,34,35]. Moreover, since lutein and zeaxanthin are fundamental for retinal function, low serum levels of these carotenoids often accompany visual impairment and age-related macular degeneration (AMD) among AD patients [36]. Together with β-carotene, plasma lutein concentration is reported to distinguish severe AD patients from milder cases and healthy controls, thereby serving as a potential biomarker for patient stratification [37]. On the other hand, a higher circulating amount of the oxidized carotenoid, β-cryptoxanthin, is associated with better cognitive performance [38]. There is also evidence that circulating α-carotene and β-carotene may serve as non-invasive disease biomarkers, as they are significantly decreased in AD [39,40,41,42]. Accordingly, an analysis of 40 patients revealed that individuals presenting lower levels of plasma β-carotene showed increased concentrations of Aβ1–42 and total tau in the cerebrospinal fluid (CSF), thus linking peripheral β-carotene to established neurochemical AD markers [43]. Of note, plasma β-carotene correlation with telomerase activity may explain the association between AD development and aging [39]. However, inconsistent results and conflicting evidence are hindering clinical development. Indeed, neither differences in the circulating levels of α-carotene, β-carotene, lycopene, and β-cryptoxanthin between AD and controls, nor their association with dementia risk, were reported in other studies, thereby questioning the feasibility of these carotenoids for exploitation as disease biomarkers [34,35,40,44].
Besides diagnosis, a therapeutic supplementation of carotenoids has been proposed as an alternative curative as well as a preventative strategy for AD.
In terms of α-carotene and β-carotene, it has been observed that streptozotocin-induced AD mice receiving β-carotene show better cognitive function and a reduced Aβ pathology, thanks to a diminished OS and the reduced activity of acetylcholinesterase (AChE), one of the most relevant proteins involved in AD [45]. However, a recent systematic review of the literature shows inconsistency regarding the use of β-carotene supplements to prevent MCI or AD in humans, thus requiring further investigation [46]. As for β-carotene, data from the “Modifying the Incidence of Delirium” (MIND) randomized controlled trial reported improved cognitive ability in individuals with higher levels of plasma α-carotene, although validation studies are still needed [47].
Lycopene. It has long been known that tomato consumption reduces the risk of developing cancer and cardiovascular disease, due to its high lycopene content [48]. Thanks to its antioxidant and neuroprotective functions, several studies proposed lycopene intake as a possible intervention against cognitive decline [48,49]. In this respect, preclinical evidence reported that lycopene supplementation is sufficient to improve learning and memory (evaluated via the Y-maze and Morris water maze tests) as well as to reduce Aβ accumulation and tau hyperphosphorylation in various mice and rat models of AD, even as a preventative strategy [50,51,52,53,54]. These benefits are mediated by decreased β-secretase expression, phosphatidylinositol 3-kinase (PI3K)/protein kinase B (Akt) signaling induction, and the stimulation of neurogenesis [52,53,55,56]. A reduction in OS, associated with decreased ROS production and enhanced antioxidant capacity (measured by nuclear factor erythroid 2-related factor 2 (Nrf2) activity, a reduced glutathione/oxidized glutathione (GSH/GSSG) ratio, malondialdehyde (MDA) levels, and GPx activity), was also observed across studies [52,54,55,56,57,58]. Often, lycopene-mediated restoration in oxidative homeostasis is accompanied by re-established mitochondrial morphology and functions [56,57,59]. All these events ensure cell viability and protect against apoptosis since in vitro lycopene treatment has been associated with reduced levels of cleaved caspase-3 and cytochrome c (markers for apoptosis), as well as with an increased expression of anti-apoptotic proteins at the expense of the pro-apoptotic proteins [55,56,59,60]. Moreover, the ability of neural stem cells to secrete nerve growth factor (NGF), brain-derived neurotrophic factor (BDNF), and vascular endothelial growth factor (VEGF) upon lycopene pre-treatment may further improve cell viability [61]. One of the key mechanisms of action of lycopene is its ability to counteract neuroinflammation. Indeed, AD rats receiving lycopene displayed reduced levels of serum pro-inflammatory cytokines (TNFα, IL-1β, IL-6β) and attenuated choroid plexus expression of the inflammatory mediators, toll-like receptor 4 (TLR4) and NF-kB, as well as increased CSF and hippocampal levels of the anti-inflammatory cytokines, IL-10 and TGF-β, even when treated at early disease stages [50,51,62]. Consistently, the same anti-inflammatory effects were observed upon the lycopene pre-treatment of mice subsequently injected with lipopolysaccharide (LPS) to trigger AD, suggesting that lycopene-based diets may be effective in disease prevention [52]. When tested in humans, results from an open-label interventional study involving 918 cognitively healthy subjects show that lycopene intake, combined with omega-3 fatty acids and Ginkgo biloba extracts, attenuated the risk of AD development later in life [63]. However, although some studies report an association between lycopene and cognitive function, human data remain limited [64]. Furthermore, innovative formulations based on lycopene-loaded microemulsions may also help to improve the antioxidant and neuroprotective properties of this carotenoid, as was recently shown in rats [65].
Lutein, zeaxanthin, and meso-zeaxanthin. AMD is often associated with AD [66], and macular pigment integration is fundamental to ensure both visual and cognitive function [67]. Accordingly, some studies show that lutein and zeaxanthin levels in the blood, macula, or brain are inversely correlated to MCI and AD occurrence, while their intake guarantees better cognitive reserve [68]. In vitro, the pretreatment of AD-mimicking neuronal cell lines with lutein and zeaxanthin extracts is reported to reduce neurotoxicity, limit apoptosis, prevent ROS release, and reestablish redox homeostasis [69,70]. In agreement with this concept, rats receiving zeaxanthin prior to Aβ1–42 exposure show reduced cerebrovascular inflammation, attenuated OS, and mitigation of the AD-related alterations in Aβ-metabolism [71]. These benefits may, at least in part, be explained by the ability of zeaxanthin to attenuate endoplasmic reticulum stress and mitigate tau hyperphosphorylation through modulation of the glycogen synthase kinase 3 beta (GSK-3β) pathway [72]. The preventative and protective features associated with macular pigment intake have also been confirmed in clinical studies. For example, cognitively healthy subjects receiving a mixture of lutein, zeaxanthin, and its stereoisomer, meso-zeaxanthin, for one year display marked memory improvements, thus potentially reducing the risk of AD development later in life [73]. Even more strikingly, 59 individuals, aged between 18 and 25 years, who received macular carotenoids for 6 months showed improved memory functions (both composite and verbal), increased attention, and enhanced cognitive performance compared to controls [74]. These effects seem to be mediated by reduced blood IL-1β, as well as by increased serum BDNF and macular carotenoid concentrations, which resulted in systemic antioxidant defense [74]. Nevertheless, results from the “Age-Related Eye Disease Study 2” (AREDS2) randomized trial report no significant benefits of lutein and zeaxanthin supplementation in terms of cognitive performance among individuals with a high risk of AMD development [75]. In another randomized double-blind clinical trial, the treatment of 31 AD patients with macular xanthophylls was not sufficient to observe any cognitive improvement, despite increased levels of blood lutein and zeaxanthin that were associated with the intake of these antioxidants [76]. Again, no amelioration in serum lipid oxidation was observed in AD patients under supplementation with lutein, zeaxanthin, and meso-zeaxanthin [77]. However, increased beneficial effects of macular carotenoids were observed upon their intake together with fish oil and omega-3 fatty acids, suggesting a dietary synergism [78]. Accordingly, cognitively healthy subjects aged over 65 years, on a diet supplemented for 2 years with a combination of fish oil, vitamin E, and macular pigments, showed improved cognitive ability, measured by working memory test performance, and increased levels of tissue carotenoids, as well as systemic xanthophylls and omega-3 fatty acid concentrations [79]. New formulations may also help carotenoids to achieve a higher therapeutic impact. In this respect, cationic biopolymer nanoparticles, endowed with efficient BBB permeation and brain localization, have been proposed as a vehicle for the intranasal delivery of carotenoids, and constitute a novel strategy to promote in situ antioxidant activity [80]. These innovative and targeted approaches may be used in new clinical trials to better clarify the preventative role of macular pigments within the AD context.

2.2. Vitamins

The role of vitamins in AD and cognitive disorders has been extensively reviewed in recent years, wherein their function is usually associated with the ability to obstruct the impact of OS on neuroinflammation and neurodegeneration [81,82,83,84,85,86]. AD and its cognitive decline have been associated mainly with deficiencies in vitamins A, D, K, and E [85], while the most recent debate deals with the role of vitamin E in AD pathogenesis and progression [87].
Vitamin E is represented by a family of eight homologs that are synthesized by plants, starting from homogentisic acid (a phenolic acid). The family includes four tocopherols and four tocotrienols, divided into the α, β, γ, and δ forms, based on the methyl substitution in the aromatic ring [88]. Alpha-tocopherol is the predominant form in tissues and is, currently, the most extensively studied form [89,90,91]. Recent studies on the mechanism of action of these molecules indicate that γ-tocopherol, δ-tocopherol and γ-tocotrienol, besides influencing the immune function and cell signaling, together with lowering cholesterol levels, have unique antioxidant and anti-inflammatory properties, which are superior to those of α-tocopherol in the prevention and treatment of chronic diseases [92,93]. These forms of vitamin E are all scavengers of reactive nitrogen species (RNS), in addition to ROS; they inhibit cyclooxygenase (COX) and 5-lipoxygenase (5-LOX), catalyze the biosynthesis of eicosanoids, and suppress proinflammatory signaling. For this reason, these forms were extensively studied within the oxidative metabolism via in vitro investigation for their anti-inflammatory effect and, also, for their efficacy in preclinical models and in interventional human clinical trials [94,95]. Many biological activities that are ascribed in the literature to vitamin E offer a theoretical basis for an understanding of its beneficial effects in the treatment of AD. In this regard, recent evidence reports that vitamin E is particularly suited to preventing AD development. A possible explanation may come from the observation that vitamin E deficiency, when caused, for example, by mutations in the α-tocopherol transfer protein A (TTPA) in familial ataxia syndrome, triggers neurological disorders [96], thus indicating that this vitamin is particularly crucial to ensure an antioxidant response in the brain [97]. Undoubtedly, the role of vitamin E is paramount in reducing Aβ accumulation in the central nervous system (CNS), as observed in TTPA knock-out (Ttpa-/-) mice [98,99,100,101].
As mentioned, increased OS and inflammation are among the possible mechanisms involved in the pathogenesis of AD. Within this context, vitamin E has been proven to have positive effects on cognitive function [102]. When administered to AD mice, vitamin E was correlated with a reduction in Aβ in the brains of young mice, but not in aged animals [103]. Other studies carried out in animal models have shown that supplementation with α-tocopherol attenuates alterations of the Aβ metabolism in aged animals and prevents deficits in memory function [104]. Similar effects were observed for a metabolite of α-tocopherol (α-tocopherol quinine), the oral administration of which improves learning in amyloid precursor protein/presenilin-1 (APP/PS1)-transgenic mice, leading to a reduction in the brain levels of Aβ oligomers and determining a decrease in oxidative stress and in the production of inflammatory mediators [105]. In agreement with these findings, α-tocopherol was able to counteract the formation of Aβ oligomers and the toxicity induced by Aβ itself, reducing the inflammatory processes, the generation of ROS, and the oxidation of lipids in cell cultures [105,106,107].
Vitamin E was also documented as playing a neuroprotective role through the modulation of specific cell-signaling pathways. Indeed, studies performed on rats have shown that, at the hippocampal level, the deprivation of vitamin E is associated with the expression of numerous genes related to the development and progression of AD. These genes have been identified as important regulators of hormonal metabolism, apoptosis, growth, neurotransmission, and Aβ metabolism [108]. Consistently, low levels of cerebral α-tocopherol have been observed to induce downregulation of the genes involved in myelination and synaptogenesis, neuronal vesicle transport, and glial functions [109].
In addition, vitamin E inhibits several enzymes involved in neuroinflammation and oxidative damage, which are typical features of AD [110,111]. Furthermore, vitamin E activates protein phosphatase 2 (a phosphatase that plays a significant role in tau protein homeostasis), which has been shown to be downregulated in the brain of AD patients [112]. Recently, vitamin E has also been reported to be able to decrease cholesterol levels by affecting the pathway of sterol-regulating proteins [113]. Notably, numerous studies on cell cultures have shown that a reduced cholesterol content is associated with a decrease in the production of Aβ, while an increase in cholesterol has the opposite effect [114,115]. Accordingly, a strong positive correlation between hypercholesterolemia and increased Aβ levels has also been observed in animal models [115,116] and it has been documented that these effects are due to the direct stimulation of β- and γ-secretase activity by cholesterol [117,118]. Furthermore, the high cellular levels of cholesterol stimulate the internalization of APP, leading to an over-production of the substrate for β- and γ-secretases [119].
Besides having an impact on Aβ production, vitamin E and its derivatives also play an important role in tau functioning. Indeed, the treatment of neuronal cultures with vitamin E prevents the Aβ-induced hyperphosphorylation of tau, although the data are still conflicting [120]. The effect of vitamin E on tau was also analyzed in vivo in different animal models, where it was found that tau-transgenic mice supplemented with α-tocopherol display a reduction in disease development and improved health [121].
Concerning humans, low plasma levels of vitamin E have been observed in both MCI and AD subjects [102,122,123,124], while high plasma concentrations and a diet enriched in vitamin E have been associated with a reduced risk of developing AD [41,125,126]. On the other hand, elevated concentrations of vitamin E in the brain of AD patients suggest a possible compensatory response to the damaging oxidative stress conditions found in the CNS of these subjects [81]. Therefore, numerous studies have analyzed the effect of vitamin E supplementation on the progression of AD; however, they reached conflicting results. Actually, although some investigations have observed a slowing down in the progression of AD in patients treated with α-tocopherol [127], other authors have observed that vitamin E supplementation does not show any benefit in either MCI or AD patients [128,129].
Another vitamin that plays a key role in the brain is vitamin A (i.e., retinol, retinoic acid, and retinal and β-carotene); a deficiency in vitamin A is considered to be a fundamental biomarker of cognitive disorders [130]. Interestingly, vitamin A has been reported to be able to counteract the formation of Aβ in the AD brain [131,132,133,134]. Other vitamins, such as vitamin K, play a significant role in AD prevention and therapy [135]. For instance, low cerebral levels of vitamin K have been associated with worse cognitive decline and with the development of neurodegenerative diseases [136,137]. A deficiency in fundamental vitamins may also have a significant action on mitochondria [138,139]. A deficit in vitamin D, for example, impairs many of the functions associated with mitochondrial biology [140]. As a matter of fact, the depletion of vitamin D and the vitamin D receptor is fundamental for mitochondria activity [138,139], and this can surely help to explain why vitamin D is fundamental to preventing cognitive disorders, such as those occurring in AD [141,142].
The B family of vitamins was recently put in the spotlight because serum homocysteine is considered a potential risk factor for neurodegenerative disorders and the B vitamins, in particular vitamin B6 and vitamin B12, besides folic acid, are able to lower the serum levels of homocysteine, thus contributing to counteracting cognitive impairment [84,143,144,145,146]. By contrast, Li and colleagues reported no significant results in a meta-analysis about the effect of B vitamins on AD [147], whereas the findings of a previous meta-analysis, comprising 21 randomized controlled trials and 7571 participants, supported the intake of vitamin B to prevent cognitive impairment in adults [148]. Premature aging and neurodegeneration are the fundamental hallmarks of AD, which can be addressed by powerful antioxidant molecules, such as L-ascorbate [149]. A quantitative meta-analysis on the use of vitamin C in AD, performed on 12 studies of 1100 patients, concluded that a deficiency in vitamin C uptake is involved in the progression of AD [150], wherein a deficit in vitamin C is also associated with non-cognitive modifications, including anhedonia, decreased motivation, and sleep disorders due to a reduction of dopamine in the nucleus accumbens [151].
Fundamentally, vitamins in AD are able to reduce the impact of ROS and RNS as ignition factors for the mitochondria-released inflammasome, NRLP3, and the induction of neuroinflammation, acting either as antioxidant complexes or cofactors in the scavenger enzymes, thereby turning the debate on the crucial role of proper dietary habits in the elderly and during the course of one’s life [152].

2.3. Flavonoids

Flavonoids, a class of polyphenolic compounds produced by plants, are abundant in fruits and vegetables, as well as in tea and wine [153]. These compounds possess many biological properties, including antioxidant, anti-inflammatory, and anticarcinogenic activities [154,155], as a result of which they can potentially prevent neurodegenerative diseases [156,157]. In this regard, the intake of food rich in flavonoids, such as chocolate, wine, and green tea, was seen to be efficacious in terms of cognitive function in elderly people [158,159]. However, flavonoids’ benefits for humans are dependent on their method of intake and bioavailability [160]; indeed, considering that, in vivo, their bioavailability is low and they do not reach the concentrations necessary for exerting their antioxidant activity [155,161], the neuroprotection that they provide could be linked to additional mechanisms, including their effect on mitochondria and apoptosis [155,162].
Chemically speaking, flavonoids are divided into six subclasses: flavanols, flavanones, flavones, flavonols, isoflavones, and anthocyanins [163].
Catechins, which belong to the flavanol family, are antioxidant molecules found in various kinds of fruits (apricots, black grapes, and strawberries), vegetables (beans), as well as in tea and red wine [160]; they are able to scavenge free radicals, chelate metal ions, and induce antioxidant enzyme activity [164]. Based on their structure, eight main catechins have been identified, including epigallocatechin gallate, which is abundant in grapes, the seeds of leguminous plants, and tea [160,165].
Concerning AD, some in vitro studies have reported the beneficial role of flavanols in alleviating its pathological signs. They are able to reduce the formation of Aβ senile plaques and neurofibrillary tangles, prevent neuronal apoptosis [166], inhibit AChE activity, and activate the PI3K/Akt pathway, resulting in the inhibition of tau hyperphosphorylation (p-tau) [167]. Preclinical studies have highlighted the positive effects of different flavanols. For instance, OligonolR, a flavanol-rich extract from lychee fruit, may improve cognitive impairment [168], while icariin, a flavanol contained in the medicinal herb Epimedium sagittatum, may attenuate synaptic plasticity and cognitive deficits through the activation of the BDNF/tropomycin receptor kinase B (TrkB)/Akt pathway [169], and cocoa extract administration may reduce Aβ oligomer formation [170]. Nevertheless, until now, no clinical trials have been performed; therefore, studies involving these flavanols are mandatory to better understand their role in AD. One of the most promising flavanols for the development of AD therapies is epigallocatechin gallate (EGCG), the main component in green tea, with neuroprotective mechanisms of action that have been widely studied, both in vitro and in vivo [171]. In the context of AD, some experimental studies have reported its anti-amyloidogenic and anti-inflammatory properties. EGCG may inhibit Aβ production by increasing the levels of the α-secretase, ADAM10, which is involved in the non-amyloidogenic pathway of APP, and converting mature fibrils into nontoxic protein aggregates, thereby reducing the overexpression of Aβ1–40 and Aβ1–42 [172,173,174,175,176]. Furthermore, its ability to modulate tau hyperphosphorylation has also been demonstrated [176,177,178]. As an anti-inflammatory compound, it is able to attenuate microglia activation and neuroinflammation by inhibiting the TLR4/NF-κB inflammatory pathway and alleviate neurotoxicity by reducing neuronal loss [179,180]. All these beneficial effects are able to counteract cognitive impairments in AD mice [174,175,181]. Interestingly, the combination of EGCG with other compounds endowed with a similar action has already shown encouraging results. For instance, the administration of hyaluronic acid with EGCG and curcumin inhibited Aβ aggregation [182], while treatment with EGCG and ferulic acid improved cognitive impairment and decreased Aβ levels [183]. However, although studies in AD animal models have shown promising results, clinical trials are still limited. In this respect, a randomized, double-blind, controlled clinical trial is currently evaluating the combination of lifestyle intervention with EGCG intake [184]. The results obtained from this study will be crucial in designing further clinical trials, with the aim of clarifying the actual impact that EGCG can have on the prevention and course of AD in humans.
Flavanones are mostly found in citrus fruit and, at lesser concentrations, in tomatoes and mint, while flavones are abundant in celery and parsley [160].
Hesperedin (Hes) is a lipophilic flavanone that is present in oranges and lemons and is able to cross the BBB and provide neuroprotection [185]. Its aglycone, hesperetin (HPT), is known to be more bioavailable [160,186]. Thanks to their properties, including their neuroprotective effect [187,188,189], Hes and HPT may be potential candidates to manage neurodegenerative diseases [190]. Therefore, many studies focused their attention on these two flavanones as promising therapeutic agents for AD. Indeed, some in vivo animal studies highlighted the beneficial role of Hes in AD hallmarks, resulting in an improvement in memory deficit and behavioral impairments [191,192]. This flavanone is also able to reduce APP, Aβ1–42, and p-tau levels, and attenuate AChE activity, as well as prevent neuronal loss [185,193,194,195,196]. Furthermore, it may also act as an antioxidant and anti-inflammatory agent [188,197]. In AD mice, Hes reduces ROS, lipid peroxidation, and protein carbonyl, while increasing heme oxygenase 1 (HO-1), SOD, CAT, and GPx activities via activation of the Akt/Nrf2 signaling pathway and the inhibition of the receptor for advanced glycation end-products (RAGE)/NF-κB signaling pathway. Since RAGE is also a receptor for Aβ and the activation of this pathway results in neuronal OS and neuroinflammation, its downregulation seems to engender neuroprotection [198]. In another AD mouse model, Hes inhibited the overexpression of inflammatory markers, including NF-kB, inducible nitric oxide synthase (iNOS), cyclooxygenase-2 (COX-2), and the glial fibrillary acidic protein (GFAP) [199]. Despite the encouraging results obtained following these animal studies, only a few clinical studies have tried to elucidate the role of Hes in human neurodegenerative diseases. In a study conducted on healthy older adults, the consumption of Hes-rich orange juice improved cognitive function [200]; however, another study suggested that citrus intake more than two times a week is associated with a minor risk of developing dementia [201]. Nevertheless, the mechanism of action of Hes in the human body is still unclear; therefore, further clinical trials are needed to clarify and confirm the efficacy of this flavanone in neuroprotection and AD treatment.
Naringenin (NGN) is one of the most abundant flavanones in citrus fruits, also found in grapefruits and tomatoes. Due to its lipophilic structure, NGN can cross the BBB [202] and some in vitro studies have shown that it is endowed with a neuroprotective role. NGN may inhibit pro-inflammatory cytokine, iNOS, and COX-2 expression through adenosine monophosphate-activated protein kinase α (AMPKα)/protein kinase C δ (PKCδ) signaling pathway activation [203]. In the context of AD, this flavanone can induce autophagy by activating AMPK/unc-51-like kinase (ULK1) axis, which leads to Aβ1–42 degradation [202], resulting in alleviation of the neurotoxic effects of Aβ-protein [204]. Furthermore, the administration of NGN nanoemulsion in SH-SY5Y cells that were exposed to Aβ downregulated APP and the expression of beta-site APP cleaving enzyme 1 (BACE1, which is the major secretase involved in Aβ production), as well as tau phosphorylation [205]. To support the hypothesis of NGN protection against neurodegeneration, further preclinical studies, performed using AD animal models, evaluated its effects. In this regard, the administration of NGN significantly enhanced cognitive deficits by reducing Aβ production, tau-hyperphosphorylation, AChE activity, neuroinflammation, and OS in the brain [206,207,208,209]. A clinical trial also studied the safety and pharmacokinetics of NGN in healthy adults and reported that the ingestion of 150 to 900 mg is safe [210]. However, there is a lack of clinical trials regarding the efficacy of NGN in AD. Nevertheless, from the evidence obtained in animal studies, NGN seems a plausible candidate for the management of AD and is worthy of further investigation.
Nobiletin (NOB) is a flavone extracted from citrus peels, such as mandarins, sweet oranges, and lemons. Its beneficial properties, including its antioxidant, anti-inflammatory, and antiapoptotic effects, as well as its potential role in neurodegenerative diseases, such as AD and Parkinson’s disease (PD), have already been reported in the literature [211]. In this regard, several studies have highlighted its ability to suppress the expression of IL-1β, TNFα, iNOS, and COX-2, inhibit microglia activation via the mitogen-activated protein kinase (MAPK) and NF-κB signaling pathways, and increase the activities of GPx and manganese-superoxide dismutase (MnSOD) [212,213,214]. Concerning AD, several preclinical studies suggest the effects of NOB on the pathological features of AD. For instance, NOB can attenuate hippocampal neuroinflammation by lowering the levels of inflammatory cytokines and the transcription factor NF-kB, OS, and lipid peroxidation, and by strengthening the defense mechanisms against astrogliosis-associated neuroinflammation [212,215,216,217]. Moreover, NOB may also reduce Aβ1–40 and Aβ1–42 levels, as well as tau phosphorylation, resulting in the improvement of memory deficits and learning abilities [212,213,218,219]. Given the encouraging results obtained in animal studies, some clinical trials have also been performed. A randomized, double-blind, placebo-controlled study investigated the benefits and safety of this flavone in the elderly, finding that the administration of NOB-containing food had a beneficial effect on memory in people with MCI, compared to the placebo group [220]. Nevertheless, this study was not conducted in AD patients. In this regard, one observational study performed in AD subjects showed no significant changes in cognitive functions in patients treated with donepezil (an acetylcholinesterase inhibitor that is commonly used for AD treatment) in association with a herbal medicine containing NOB-rich Citrus reticulate, compared to the group only treated with the drug [221]. Conversely, another study conducted on 6 patients reported the beneficial effect of Nchinpi, a traditional drug containing NOB, combined with donepezil in patients with mild-to-moderate AD. Indeed, after one year, the cognitive state worsened only in the control group, pointing out the potential effectiveness of this combination in preventing the progression of cognitive impairments in AD patients [222].
Apigenin (API) is a flavone found in various plants, including chamomile, Melissa officinalis, parsley, thyme, and oregano, in vegetables, such as onions and celery, and in citrus, offering anti-inflammatory, antioxidant, anti-amyloidogenic, and neuroprotective properties [223,224]. Thanks to these effects, API could be a potential candidate to prevent or slow down the progression of AD. Indeed, several in vitro and in vivo preclinical studies reported its ability to attenuate AD hallmarks. API may be endowed with a protective role against neuronal damage by reducing neuroinflammation through the inhibition of OS and the TLR4/NF-κB inflammatory pathway [225,226]. Moreover, it has been demonstrated that this flavone may reduce Aβ formation and tau hyperphosphorylation by decreasing the BACE1 and GSK-3β levels, respectively, thus favoring an improvement in memory and learning deficits [223,224,227,228]. Interestingly, vitexin, an API flavone glycoside, was seen to reduce Aβ peptide expression and increase the lifespan in Caenorhabditis elegans with AD [229]. To assess the effects of API in humans with AD, Balez et al. used a human induced pluripotent stem-cell model of AD, in which skin cells derived from both patients and healthy individuals were cultured and differentiated into neurons. They noticed that API administration reduced neuronal apoptosis and inhibited the activation of cytokines and nitric oxide (NO) production [230]. Overall, considering the substantial evidence for API’s neuroprotective efficacy, both in vitro and in vivo, further studies and clinical trials should be encouraged.
The most widely studied flavonols, antioxidant molecules that are abundant in multiple fruits and vegetables, as well as in tea and red wine, are quercetin, fisetin, kaempferol, and myricetin [231].
Quercetin is an aglycone form of flavonoid glycosides extracted from grapes, onions, berries, broccoli, and citrus [232]. Several studies have indicated its protective role in different pathologies associated with dementia, such as stroke and cardiovascular disease, as well as in aging. In the context of AD, quercetin is found to improve memory deficits and learning functions in several rodent models [232,233,234,235,236]. Furthermore, the efficacy of the intake of quercetin-rich onions in improving memory in early-stage AD patients has been reported [237], although the study by Holland et al. showed no enhancement in cognitive activities after quercetin administration [238]. Quercetin also possesses anti-inflammatory, antioxidant, and anti-amyloidogenic properties [232,234,236,239]. Regarding this last point, some in vitro and preclinical studies have highlighted quercetin’s ability to inhibit Aβ aggregation [240,241] and reduce Aβ accumulation and plaque generation by decreasing the BACE1 protein levels [236]. Despite this evidence, a study by Huebbe et al. [242] reported no effects of quercetin on the mRNA levels of genes involved in AD, especially BACE1. As an anti-inflammatory agent, quercetin may modulate neuroinflammation by suppressing the expression of the ionized calcium-binding adapter molecule 1 (Iba-1) and GFAP, which are typical proteins of activated microglia and astrocytes, respectively. Furthermore, it was reported that quercetin inhibited the TLR4/NFKB pathway involved in inflammatory signaling, as well as multiple inflammatory mediators such as COX-2, NOS-2, IL-1β, prostaglandins, and leukotrienes [232,236]. In addition, as an antioxidant molecule, quercetin is capable of decreasing MDA levels and increasing SOD and glutathione (GSH) expression, as well as modulating the Nrf2 transcription factor, with the consequent gene expression of several anti-antioxidant enzymes [232,236]. Moreover, quercetin is able to reduce tau phosphorylation [234,235], modulate mitochondrial dysfunction by increasing the mitochondrial membrane potential and ATP synthesis, decreasing ROS expression; it also displays a neuroprotective antiapoptotic function [236,241]. Evidence has indicated that even though quercetin crosses the BBB, its permeability is poor and its concentration in the brain is low. This may be due to the fact that quercetin is a substrate for the BBB efflux transporter, P-glycoprotein [242]. Therefore, promising delivery systems, including quercetin lipid nanoparticles, have been investigated and have proven efficient as potential future pharmaceutical approaches [243,244,245].
Another important flavonol investigated as a valuable neuroprotective compound for AD is fisetin, which is not particularly abundant in fruits (strawberries, kiwi, peaches, and grapes) and vegetables (tomatoes and cucumbers) [246]. Concerning AD, several animal studies have reported the effectiveness of fisetin in alleviating AD signs. On a molecular level, the administration of this flavonol may decrease Aβ production and aggregation by reducing BACE1 expression, tau hyperphosphorylation [247,248], neuroinflammation, and GFAP [246,247]. Moreover, as an antioxidant agent, fisetin is able to counteract protein carbonylation and lipid peroxidation [246,249], as well as to favor Nrf2 activation, with a consequent increase in GSH levels [246,248]. Thanks to these beneficial properties, the ability of fisetin to counteract cognitive impairment [246,247,249] and locomotor deficits has been reported [250]. Nevertheless, given that dietary fisetin sources are inadequate for effectiveness in humans, advanced delivery systems and derived compounds are needed [246].
Interestingly, many current studies indicate the potential effect of other compounds derived from flavonols in AD treatment. An in vitro study by Jung et al. reported the role of sophoflavescenol, a prenylated flavonol extracted from Sophora flavescens, in inhibiting BACE1 and cholinesterase activities [251]. In addition, treatment with icaritin, another prenylated flavonol, improved cognitive dysfunction, and reduced amyloid deposition, as well as tau phosphorylation, in AD mice [252]. Finally, the study conducted by Calderon-Garciduenas et al. among urban children suggested the beneficial effects of cocoa and dark chocolate administration against OS and neuroinflammation [253]. However, some studies reported controversial results concerning the impact of flavonols on AD hallmarks. For instance, the treatment with Ginkgo biloba extracts did not show any effect on Aβ aggregation and BACE1 activity [254,255].
Isoflavonoids are compounds mainly found in soybeans and other leguminous plants. While daidzein, daidzin, genistin show benefits narrowly in hormone-related diseases, genistein has been widely studied as a treatment in neurodegenerative disorders [231].
Regarding this last compound, genistein (GNT), the main phytoestrogen of the Leguminosae, has several properties, including antioxidant, antimicrobial, antitumor, and neuroprotective effects [256,257,258], and could be a potential candidate for AD treatment [257]. In this regard, some in vitro studies have shown the ability of this isoflavonoid to reduce Aβ production and deposition, reduce APP phosphorylation, and inhibit the activity of AChE [258,259,260]. Nevertheless, a study on SHSY5Y cells reported an increase in both APP and β-secretase mRNA and protein expression, thus going against the concept of a positive role for GNT in AD [261]. Additionally, a study on an LPS mouse model showed reduced AChE activity in the hippocampus after GNT oral administration [262], while Yu-Xiang Wang et al. reported its neuroprotective effect against Aβ25–35, via the modulation of choline acetyltransferase (ChAT) expression [263]. In terms of AD, several in vitro and in vivo studies have shown the beneficial role of GNT administration in ameliorating the symptoms of the disease. For instance, GNT can inhibit APP expression, thus reducing Aβ production and its following aggregation, and modulate tau phosphorylation, with a consequent improvement in memory and learning abilities [258,259,260,264,265,266,267,268,269]. However, a recent clinical trial reported controversial results after soy isoflavone treatment; indeed, Gleason et al. found no cognitive changes in AD patients [270]. Besides these beneficial properties, GNT has also been proven to have antioxidant and anti-inflammatory effects. In this regard, GNT has been reported to be able to counteract Aβ-induced oxidative injury by decreasing ROS accumulation and MDA levels, as well as by increasing antioxidant enzyme activity (in SOD and GSH) [262,269], presumably through the activation of the PI3K/Akt/Nrf2 signaling cascade [271]. Moreover, as an anti-inflammatory agent, GNT seems to be involved in the inhibition of the pro-inflammatory cytokines, IL-6 and TNFα, and other inflammatory or inflammation-associated factors, such as NF-kB, COX-2, iNOS, and GFAP [262,272]. Since GNT has poor oral availability, with limited absorption and considerable metabolization by the gut microbiota [257], recent studies have explored the use of nanoconjugates, in order to increase GNT’s BBB crossing and better modulate brain distribution [273]. Finally, interesting studies have also focused on the use of GNT-O-alkylamine derivatives as a potential AD treatment, finding antioxidant and AChE-inhibitory effects [274,275].
Anthocyanins are a class of pigments that are extracted from plants, in particular, soybean seeds and berries [276]. The main dietary anthocyanins are cyanidin, delphinidin, malvidin, peonidin, pelargonidin, and petunidin [277]. These compounds have been shown to display multiple activities, including antioxidant, anti-inflammatory, and anti-apoptotic effects [278]. These properties, as well as their effective role in tau hyperphosphorylation and Aβ amyloidosis regulation, could make them potential candidates for AD treatment [278]. A clinical study by Shishtar et al. [279] reported the efficacy of long-term anthocyanin intake in reducing the risk of AD and other related dementias. In addition, an anthocyanin-enriched juice has been found to have beneficial effects in improving cognitive function in middle-aged women [280], indicating the potential role of anthocyanins, not only in AD-related dementia but also in the aging process [277]. The neuroprotective effects of anthocyanins have also been documented in AD mouse models, with consequent improvement in learning and memory abilities [281,282,283,284,285,286,287,288]. Some preclinical studies have also reported their anti-Aβ aggregation activity [289,290,291]; additionally, they are able to reduce the amount of Aβ1–40 and Aβ1–42 and alter the APP metabolism by increasing the cleavage toward the production of soluble amyloid precursor protein-α (sAPPα), a neurotrophic factor [276,277,292]. Furthermore, treatment with anthocyanins was demonstrated to significantly reverse and regulate BACE1 expression and protect against Aβ toxicity [276,286,292,293]. Regarding anthocyanin’s effects on tau-proteins, the results are controversial; some research carried out on AD mouse models reported a slight improvement in tau protein expression [286] and a reduction in p-tau hyperphosphorylation [292], while other studies showed unaltered levels of phosphorylated tau after anthocyanin administration [276]. These pigments also possess antioxidant and anti-inflammatory properties. Both the in vitro and in vivo studies demonstrated that anthocyanins are able to decrease lipid peroxidation and raise Nrf2 nuclear translocation, as well as raise the levels of antioxidant enzymes, including SOD, CAT, and GPx [277,289,294,295,296,297]. As anti-inflammatory agents, they may reverse microglia and astrocyte activation, resulting in the downregulation of the pro-inflammatory cytokines (TNFα, IL-1β, IL-6), COX-2, and iNOS [282,286]. Mechanistic studies have reported anthocyanins as molecules that are also able to suppress neuroinflammation, through the inhibition of NF-kB and Jun N-terminal kinase (JNK) activation [277,284]. Moreover, anthocyanins have been shown to modulate the mitochondrial apoptotic pathway by regulating different enzymes (caspases 3, 7, 9) and pro-apoptotic proteins in several AD mouse models [276,282,292,294]. Although anthocyanins are absorbed as glycosides at very low levels, they have been documented to cross the BBB in both rodents and humans, and exert sufficient biological activities, including gene regulation and cell signaling [276,277]. Within this context, nanodrug delivery systems have been developed for the purpose of achieving better absorption, bioavailability, and effectiveness. For instance, anophytosome formulations targeting the mitochondria, and polyethylene glycol–gold nanoparticles loaded with anthocyanins, have proven more effective than treatments using free anthocyanin [293,297,298].
Overall, flavonoids comprise a broad range of antioxidant compounds, with great translational potential as supportive AD treatments. The results from new, randomized trials involving AD and MCI patients should better elucidate the optimal time for treatment and offer insights into their possible use as preventative strategies for this devastating disease.

2.4. Non-flavonoids

Another potential approach to preventing or treating AD could be the assumption of non-flavonoid substances, natural compounds that are mainly present in fruits, vegetables, green tea, and whole grains, with effective antioxidant properties [299,300]. The ability to attenuate OS and scavenge free radicals could make them possible candidates for counteracting neurodegeneration and improving the typical signs of neurodegenerative diseases [301]. In terms of AD, several studies reported the beneficial effects of non-flavonoids in managing this pathology (Table 1) [302,303,304,305,306,307,308,309,310,311,312,313,314,315,316,317,318,319,320,321,322,323].
Non-flavonoid compounds include phenolic acids, stilbenes (resveratrol), curcuminoids (curcumin), and lignans [299].
Phenolic acids, a class of secondary metabolites that are widely present in herbs and, more generally, in the plant kingdom, are divided into two groups, based on their chemical structure: hydroxycinnamic acids (including ferulic, caffeic, and p-coumaric acids) and hydroxybenzoic acids (including gallic and ellagic acids) [324,325]. In addition to possessing anti-inflammatory, antibacterial, anticarcinogenic, and neuroprotective properties [326,327], they are considered excellent as both direct and indirect antioxidants, able to scavenge excessive free radicals and activate the endogenous antioxidant pathways and enzymes, respectively [328].
Ferulic acid (FA), commonly abundant in oranges, tomatoes, spinach, grains, and wheat bran, and largely used in the cosmetics and food industries, is one of the most studied phenolic acids for its neuroprotective, antioxidant, and anti-inflammatory effects [329,330,331,332,333]. Due to its structure, FA is a strong free-radical scavenger; moreover, it also inhibits ROS formation by reducing pro-oxidant enzyme levels and promoting antioxidant enzyme activities [330].
The neuroprotective effect of this compound has been widely reported in many experimental studies on PD, cerebral ischemia/reperfusion injury, and multiple sclerosis (MS) [333,334,335,336]. In the context of AD, using in vitro and in vivo animal models, FA has been proven to have anti-amyloid, antioxidant, and anti-inflammatory effects, resulting in the improvement of cognitive impairment and the alleviation of neuropathological signs [337]. In this regard, FA can inhibit Aβ fibril aggregation and the resulting Aβ senile plaque formation by reducing the expression of BACE1 and APP levels [302,338,339,340]. It can also reduce tau levels and disassociate the pre-formed neurofibrillary tangles [337]. As antioxidant agents, FA and its derivatives may attenuate ROS formation by increasing SOD activity and reducing MDA levels [303,341,342,343]. In particular, butyl ferulate, a lipophilic FA derivative, has been shown to inhibit Aβ1–42 aggregation, reduce ROS accumulation, and increase antioxidant enzyme expression (such as HO-1, thioredoxin 1, and GSH) by activating the Nrf2 antioxidant signaling pathway, the most important cellular antioxidant defense cascade [344]. Furthermore, as anti-inflammatory compounds, they may alleviate neuroinflammation by reducing some of the pro-inflammatory enzymes (COX-2 and 5-LOX) and cytokines (TNFα, IL-1β, and IL-6) in the brains of AD mice [17,338,345]. In addition, given that, in the presence of neuronal damage, astrocytes and microglia overproduce GFAP and Iba1, two markers of astrogliosis and microglia activation, respectively, it has been reported that FA administration can reduce the levels of these proteins [183,342]. However, notwithstanding these beneficial properties, to date, no clinical trials have been performed to study the effects of FA on AD patients; therefore, future studies are mandatory to confirm these valuable results, obtained from both in vitro and in vivo models.
Caffeic acid (CA), which is abundant in fruits (strawberries and blueberries), vegetables (carrots and tomatoes), and beverages (wine, tea, apple juice, and coffee) [346], has a broad spectrum of functions, including anti-inflammatory, antioxidant, anticancer, and immunomodulatory effects [347,348,349]. CA can work as an antioxidant and is able to prevent ROS production, scavenge free radicals, and chelate metals (such as ferrous ions) [350,351]. Besides these important activities, its neuroprotective role in cerebral ischemia/reperfusion injury, PD, brain lesions, and MS has been reported [352,353,354,355].
To date, few in vitro and in vivo studies have been performed to evaluate its anti-AD effects. In particular, a neuroprotective role of caffeic acid phenethyl ester (CAPE), abundant in the honeybee, and caffeic acid phenethyl ester 4-O-glucoside (FA-97), its synthetic derivative, has been demonstrated. Their administration may counteract OS and reduce ROS production by inducing the activation of the Nrf2/HO-1 pathway, resulting in increased SOD and GSH levels [306,356]. CA has also demonstrated the ability to inhibit AChE activity, restoring the levels of acetylcholine (ACh), a neurotransmitter with an important role in cognitive function [357,358]. This modulation alleviates cholinergic neuronal loss, improves cognitive functions, and attenuates learning and memory deficit in AD animal models [305,346,357]. Furthermore, it has been demonstrated that CA, with its anti-amyloidogenic activity, can exert a neuroprotective effect against β-amyloid-induced neurotoxicity [304].
Nevertheless, despite the encouraging results obtained from these studies, even for this compound, clinical trials are needed to better assess its potential therapeutic role in managing AD.
p-Coumaric acid (p-CA), a plant metabolite that is widely found in vegetables (carrots, tomatoes, garlic, and navy beans), fruits (apples, grapes, and oranges), peanuts, and cereals [359,360], is endowed with antioxidant, anti-inflammatory, antiangiogenic, and antitumor potential [361,362,363,364]. As an antioxidant agent, p-CA can reduce OS by activating antioxidant enzymes, and constitutes an effective free radical scavenger and metal chelator [365,366]. A powerful neuroprotective effect against embolic cerebral ischemia, cerebral ischemia-reperfusion injuries, and brain damage has also been reported [367,368,369].
Regarding its potential role in AD, few in vitro and preclinical studies have highlighted p-CA’s anti-amyloidogenic, antioxidant, and anti-inflammatory effects [307,308,370,371]. Indeed, this compound may reduce Aβ generation by inhibiting BACE1 activity, relieve ROS generation by increasing SOD and GSH expression, and modulate neuroinflammation via NF-kB pathway inactivation [307,308]. Since NF-kB plays an important role in the inflammatory response, its suppression causes the under-expression of two pro-inflammatory enzymes, iNOS, and COX-2, with a consequent attenuation in Aβ-induced toxicity [307]. Moreover, p-CA can also protect against learning and memory impairments and neuronal loss by reducing brain AChE activity [308]. Interestingly, it has been shown that maltolyl p-coumaric, a synthesized compound, is able to counteract cognitive deficits and exert neuroprotective effects in a rat model by reducing ROS and neuronal loss [309].
All this evidence suggests the potential therapeutic use of p-CA against AD; however, clinical studies are necessary to better elucidate its effects in humans and the molecular mechanisms underlying its possible pharmacological employment against AD.
Gallic acid (GA) is a natural secondary metabolite that is abundant in the plant kingdom, especially in fruits (grapes, blackberries, strawberries, plums, and pomegranates) and nuts, as well as in tea and wine [372,373]. This compound and its derivatives are largely used in the food and pharmaceutical industries, thanks to their powerful properties, including their antioxidant, anti-inflammatory, anticancer, and antiviral effects [374,375,376,377,378,379]. Besides these activities, it has also been observed that GA plays a neuroprotective role in traumatic brain injury, PD, amyotrophic lateral sclerosis (ALS), and cerebral ischemia-reperfusion, as well as in psychiatric disorders [380,381,382,383].
Concerning AD, few in vitro and in vivo studies report the potential therapeutic role of GA in improving the neuropathological hallmarks of the disease. It has been documented that GA administration can stabilize OS and alleviate neuroinflammation, resulting in improved neurodegeneration by increasing antioxidant enzyme brain expression (CAT, GSH, SOD, GPx1, and SOD1) and decreasing reactive astrocytes and microglia cell activation in AD animal models. Moreover, its efficiency has been observed in reducing Aβ1–42 aggregation and amyloid plaque formation through β-secretase modulation [310,311,384]. In addition, this molecule was found to be capable of ameliorating cognitive impairments and preventing neuronal apoptosis in AD rats [310,385]. Interestingly, tannic acid, a glucoside polymer of GA, prevents the reduction of antioxidant enzyme activities, as well as an increase in AChE activity, in the proinflammatory cytokine, IL-6, and also TNFα levels, as seen in a model of sporadic dementia of the Alzheimer’s type (SDAT). Furthermore, this compound was also described as being able to favor neuronal survival by re-establishing the levels of Akt and pAkt that are involved in this process [386]. Although these studies have highlighted the neuroprotective effects of this phenolic acid and its derivatives in AD, clinical studies are needed to investigate their efficacy and safety in humans.
Ellagic acid (EA), a natural polyphenol that is abundant in fruits (strawberries, raspberries, and pomegranates) and nuts, demonstrates antioxidant and anti-inflammatory properties that indicate a potential neuroprotective agent against several neurological disorders [387,388,389,390]. The antioxidant capacity is given by its ability to scavenge free radicals, chelate ions, inhibit lipid peroxidation, and promote antioxidant enzyme activity [391,392,393]; instead, as an anti-inflammatory compound, EA may reduce pro-inflammatory factor expression [393].
To date, few studies report EA’s potential effects in alleviating AD symptoms [312,313,314]. Using different in vivo AD animal models, its effectiveness has been demonstrated in improving learning and memory deficits, due to its ability to relieve OS (↑CAT, SOD, and GSH), neuroinflammation (↓NF-kB), AChE activity, and tau phosphorylation, this latter event occurring via the modulation of the Akt/GSK-3β signaling pathway [312,313]. Furthermore, this compound may also have an anti-amyloidogenic effect. Indeed, it is capable of decreasing Aβ deposition and plaque formation by reducing BACE1 activity and APP phosphorylation [312,314].
Despite these studies reporting the beneficial effects of EA as a potential AD treatment, more in vivo experimental studies, beyond all the clinical trials, are mandatory to confirm these findings.
Among the stilbenes family, resveratrol (RSV) represents the most common polyphenolic compound, which is mainly found in grapes, mulberries, and peanuts, as well as in wine and grape juice [394,395]. It is well known for its ability to act as a strong antioxidant and anti-inflammatory agent, making it effective against different neurological diseases, including epilepsy, PD, Huntington’s disease, ALS, and neuronal injury [395]. RSV is able to scavenge free radicals and upregulate antioxidant enzyme expression, as well as inhibit pro-inflammatory molecules [396,397,398]. In terms of AD, the administration of RSV seems to have positive effects in alleviating the symptoms of this neurodegenerative disorder. For instance, some preclinical studies reported its potential effect in improving cognitive impairment, due to its capability to inhibit Aβ aggregation and tau phosphorylation, decrease pro-inflammatory protein expression, and increase Nrf2 nuclear translocation, with the consequent upregulation of HO-1, SOD, CAT, and glutathione peroxidase (GSH-Px) activities [315,317]. It has been shown that its neuroprotective effect could be also due to the RSV-mediated activation of sirtuin 1 (SIRT1), an essential protein for cognitive function and neuronal survival [394].
Although these studies underscore the beneficial effects of RSV in managing the disease, due to the few clinical trials and the poor bioavailability of the molecule, as yet, no firm conclusions have been drawn regarding its neuroprotective role and its mechanism of action in humans. For instance, a human study, performed using high doses of RSV, reported a greater reduction in Aβ1–40 and Aβ1–42 CSF levels in the treated group compared to the baseline [399]; conversely, Gu et al. (employing a lower dose of trans-RSV) and Turner et al. found no changes in Aβ1–42 CSF and plasma levels in patients treated with this compound [316,400]. These results suggest the importance of further studies with larger samples, different routes of administration, and pharmaceutical formulations, such as nanoencapsulation [401], in order to improve RSV bioavailability and, hence, its effect on cognitive impairment.
Interestingly, within the diagnostic field, RSV could be used as a potential tool for positron emission tomography (PET), since it is able to map Aβ plaques in the human brain [402].
Curcumin (CUR), a yellow phenolic pigment belonging to the curcuminoid family, is widely used as a spice, dye, and food additive, and also as a herbal medicine [403]. Several studies report its powerful biological effects, including its antioxidant, anti-inflammatory, and neuroprotective properties, through which CUR may help to manage different inflammatory and neurological disorders [404,405,406,407,408]. As an antioxidant agent, CUR is able to scavenge free radicals, inhibit lipid peroxidation, and favor HO-1, SOD, and CAT expression; its anti-inflammatory activity is ascribed to a reduction in the levels of several inflammatory cytokines (TNFα, IL-1β, IL-6, IL-8), as well as a reduction in the expression of NF-kB, COX-2, and iNOS [409,410].
In the context of AD, numerous in vitro and preclinical studies have shown anti-Aβ-deposition and aggregation, antioxidant, and anti-inflammatory CUR effects. For instance, CUR may inhibit Aβ formation by lowering BACE1 activation and Aβ aggregation (perhaps by directly binding it to the Aβ oligomers), and by promoting Aβ disaggregation [411,412,413]. Furthermore, it has been reported its ability to prevent tau phosphorylation [414] and neuroinflammation by suppressing microglial activation, as well as by reducing TNFα, IL-1β, and IL-6 production [415,416]. In addition, several animal studies report that CUR may improve cognitive function in a dose-dependent manner [417].
However, due to CUR’s poor absorption and low bioavailability, AD human studies did not report any improvement in cognitive function or a reduction in Aβ and tau levels [320,418]. In this regard, CUR analogs that were either delivered by nanoparticles or administered with a specific adjuvant could represent possible strategies to improve the effectiveness of this molecule against AD symptoms [319,419,420,421]. Therefore, clinical studies, especially those performed with the support of nanotechnology, are necessary to better understand the properties of CUR and its potential use in AD prevention or treatment. Interestingly, CUR could be used as a tool for AD diagnosis, since it is a fluorochrome that is able to bind Aβ aggregates [422].
Lignans, a group of polyphenolic compounds that are abundant in the plant kingdom, possess several biological properties, including anti-inflammatory [423], antioxidant [424], and antitumor activities [425]. Their described neuroprotective role [426] makes them, potentially, natural candidates for AD prevention or treatment. In this regard, some preclinical studies suggested the lignans’ ability to improve learning and memory decline in mice with dementia [427,428].
It should be emphasized that among all lignans, especially those isolated from Schisandra chinensis, due to their numerous pharmacological effects, they are the most studied compounds, in light of their being a possible therapeutic choice to prevent or slow down the symptoms of AD [322,323,429,430,431,432]. Indeed, the studies that were carried out highlight their ability to enhance cognitive performance by decreasing AChE activity, to reduce Aβ levels by inhibiting BACE1 activity, to prevent tau hyperphosphorylation, as well as to restore SOD and GSH-Px activities. For instance, the possible mechanism by which schisantherin B, one of these lignans, exerts its neuroprotective role could be the activation of the glutamate transporter type 1 (GLT-1) and the consequent inhibition of GSK-3β, involved in tau phosphorylation [429]. A neuroprotective effect of sesamol, sesamin, and other lignans from Prunus tomentosa seeds has also been reported [321,433,434]. For instance, sesamol, extracted from sesame oil, was seen to decrease Aβ aggregation, inhibit microglia activation, and reduce TNFα and IL-1β expression in AD mice, with a consequent improvement in memory and learning abilities [433].
Overall, considering these encouraging findings in animals, clinical trials are mandatory to investigate the possible effects of lignans in attenuating AD signs and their potential use in AD management.

2.5. Organosulfur Compounds

Glutathione. The antioxidant role of the natural tripeptide, γ-l-glutamyl-l-cysteinylglycine, better known as glutathione, is well established [435]. Under normal conditions, redox homeostasis is maintained by the continuous conversion between the oxidized (GS-SG) and the reduced (GSH) forms of glutathione by the enzymes, GPx (which contributes one molecule of glutathione disulfide via oxidation, starting from two molecules of reduced monomeric glutathione) and glutathione reductase (which exploits nicotinamide adenine dinucleotide phosphate (NAPDH) to reduce one GS-SG molecule into two of GSH) [436]. Due to its reducing potential, GSH acts as the main intracellular antioxidant buffer and its synthesis, degradation, transport, metabolism, and interconversion are finely regulated [435,436]. In accordance with a reduction in the total antioxidant capacity during senescence [300], GSH depletion has been observed during aging and age-related diseases, including cardiovascular disease, immune dysfunction, cancer, and neurodegeneration [437,438]. In terms of AD, GSH and GSH-related compounds have been proposed, both as biomarkers and also as therapeutic molecules. It has been reported that both MCI and AD patients show an overall reduction in blood GSH levels, with the intracellular GSH content being particularly decreased in MCI individuals [439]. In line with these data, a comparison between 49 MCI and 19 healthy control subjects revealed that reduced plasma GSH content is associated with worsened cognitive function over a 2-year follow-up period, confirming its potential as a prodromal AD biomarker [440]. Besides peripheral sampling, the possibility of non-invasively monitoring the brain’s GSH content via magnetic resonance spectroscopy imaging is also of interest [441], although the data remain discordant. Accordingly, while significant reductions of cortical GSH levels were reported in female and male AD patients, compared to their healthy counterparts [441], an increase in its content in the anterior and posterior cingulate cerebral areas distinguished MCI individuals from healthy controls in an independent study [442]. Although subsequent studies confirmed a positive correlation between hippocampal/cortical GSH depletion and cognitive impairment [443,444], the high variability, due to inherent sample heterogeneity, differences in the experimental methods, and non-standardized measurement techniques remain a matter of concern [439]. According to Mandal et al., the imaging measurement of cortical GSH is capable of differentially diagnosing MCI and AD with 100% specificity and 91.7% sensitivity, while its hippocampal content diagnoses MCI early, with 87.5% sensitivity and 100% specificity, although the relatively small sample size requires confirmatory studies [443]. Overall, these results underline the importance of further trials involving the time-point monitoring of GSH levels in larger AD, MCI, and control cohorts.
Given the important role of GSH in regulating the total antioxidant capacity, several in vivo studies have investigated the potential of GSH supplementation as a preventative or additional treatment for AD. For example, a three-week oral intake of GSH in the AppNL-G-F/NL-G-F knock-in mouse model, which reproduces the features of AD, is sufficient to replenish the brain GSH levels and reduce the hippocampal expression of the OS marker, 4-hydroxynonenal [445]. Similar results were also obtained in preclinical AD models following treatment with different GSH-related molecules, including the GSH mimetic D609, the GSH precursor, S-adenosyl methionine (SAM), and a GSH analog that is resistant to γ-glutamyl-transpeptidase activity, which usually degrades GSH [446,447,448]. The reported improvements include diminished OS-related products, malondialdehyde and protein carbonyls, the potentiation of the antioxidant machinery by increasing GSH levels and SOD activity, the prevention of insoluble Aβ deposition, and improved spatial learning and memory processes [446,448]. Notably, the re-establishment of OS homeostasis was often accompanied by reduced microgliosis and neuroinflammation, mediated by an increased expression of the anti-inflammatory cytokine, IL-10, at the expense of the pro-inflammatory cytokines (i.e., TNFα, IL-6, and IL-1β), as shown by the in vitro and in vivo data [445,449]. As a further confirmation of the importance of GSH metabolism in modulating neurodegeneration, the restoration of GPx in the cortical neurons and Aβ1–42-treated mice blocked neurotoxicity, attenuated memory decline, and activated the PKC βII-mediated ERK pathway, which is implicated in AD onset [450,451]. Besides neurons, the protective effect exerted by GSH on the brain endothelial cells is also of interest for neurodegenerative diseases, since it counteracts H2O2-mediated NO, ROS, and 8-hydroxy-2′-deoxyguanosine production, strengthens the tight junction proteins, and promotes the activity of the antioxidant Nrf2 pathway [452].
Despite yielding promising data, the limited uptake of GSH or GSH-precursor supplementations, combined with the inherent age-related GSH homeostasis dysregulation, calls for novel prodrug formulations, innovative GSH-related compounds, and delivery optimization [449].
Allicin, allyl sulfide indoles, and allylcysteines. Onion and garlic are excellent sources of a variety of sulfur-containing compounds with known antioxidant, immunomodulatory, antitumoral, and anti-inflammatory properties, including allicin, alliin, allyl sulfide indoles, and allylcysteines [453,454,455,456]. Mainly found in fresh and aged garlic extracts, these compounds have been shown to prevent neurotoxicity by reducing cerebral Aβ deposition while disaggregating preformed fibrils in a dose-dependent manner [457,458,459,460,461]. Improved hippocampal memory has also been reported in transgenic AD mice supplemented with aged garlic extract [462], which displays a higher total antioxidant capacity compared to fresh garlic extract [463]. Accordingly, the pretreatment of cognitively normal adult male rats with aged garlic extract showed improved short-term recognition and working and reference memory, compared to their untreated counterparts [464,465]. These benefits are mediated by reduced microgliosis, a diminished expression of the pro-inflammatory cytokine, IL-1β, attenuated cholinergic neuronal loss, and increased hippocampal levels of the vesicular glutamate transporter 1 and glutamate decarboxylase [454,464,465], which regulate glutamate transport and metabolism and are normally downregulated in AD [466,467]. Similar results were obtained upon the administration of S-allyl cysteine (SAC), one of the main ingredients of aged garlic extract [455,468]. Elsewhere, studies showed that the SAC pretreatment of streptozotocin-infused mice (an animal model of diabetes) drastically improved cognitive function, antioxidant capacity (measured by GSH, GPx, and glutathione reductase levels) and prevented apoptosis (assessed by DNA damage and Bcl-2 and p53 expression) [469]. However, the aged garlic extract’s capacity to protect against synaptic degeneration, inhibit Aβ and neurofibrillary tangle deposition, reduce neuroinflammation, and counteract ROS formation appears to be greater than SAC alone, likely suggesting a synergistic effect among the various sulfur-containing compounds found in aged garlic extract [470].
Another main constituent of garlic known for its phytochemical properties is allicin, a hydrophobic molecule produced from alliin when garlic is chopped or damaged [471]. Being a reactive sulfur species (RSS), its antioxidant activities include Nrf2 upregulation, ROS quenching, and neuroinflammation dampening [471,472]. Moreover, it exerts an inhibitory effect against various pro-inflammatory and pro-oxidative signaling behaviors, including the TLR4/myeloid differentiation primary response 88 (MyD88)/NF-kB, and the Jun N-terminal kinase (JNK), p38 MAPK, and AChE pathways [471,472]. To date, in vivo studies prove that allicin supplementation improves cognitive function while reducing Aβ expression and accumulation in various preclinical AD models, including aluminum chloride-treated rats, APP/PS1 mice, and Aβ1–42 treated mice [473,474,475,476]. Mechanistically, these benefits are mediated by oxidative homeostasis re-establishment through increased antioxidant potential versus OS, reduced pro-inflammatory cytokine (TNFα, IL-6, and IL-1β) release, improved mitochondrial function, neurotransmitter (GABA, dopamine, serotonin, glutamate, and norepinephrine) concentration, and decreased tau-phosphorylation [473,474,475,476]. Interestingly, allicin is also protective against endoplasmic reticulum (ER) stress-mediated neurotoxicity, as it increases, for instance, the hippocampal expression of the ER-resident kinase involved in the ER stress response, PERK (protein kinase RNA-like endoplasmic reticulum kinase), together with its substrate, Nrf2 [476]. Still, the lack of data in humans and the need for a larger sample size and validation studies constrain clinical development.
α-lipoic acid. α-lipoic acid (LA), or thioic acid, is a natural antioxidant that is derived from caprylic acid and belongs to the group of organosulfur compounds [477]. Thanks to its anti-inflammatory, antioxidant, neuroprotective, and anti-amyloidogenic properties, LA has recently been considered as a treatment agent for various neurological conditions, including neurodegeneration [477,478]. In terms of AD, preclinical evidence shows that LA administration to mouse and rat models of the disease increases memory and learning ability, reduces cognitive decline, and ameliorates motor functions, compared to their untreated counterparts [479,480,481,482,483,484]. These benefits are mediated by decreased oxidative stress (measured by lower levels of MDA in favor of the antioxidant enzymes’ enhanced expression of GSH, SOD-1, and CAT), reduced neuronal death, diminished neuroinflammation (measured by the number of reactive astrocytes), and impaired Tau hyperphosphorylation [479,480,481,482]. Notably, even better results are obtained upon pairing the intake of LA with physical exercise or with the administration of anti-inflammatory drugs, such as ibuprofen [482,483,484,485]. Despite these improvements, however, some evidence reports that the administration of LA to AD mice has no impact on amyloid pathology and may even reduce the lifespan, suggesting that more studies are needed to better evaluate LA’s safety and efficacy [481,482,483]. When tested in humans, results from an open-label study that was conducted on 43 patients showed that daily LA intake significantly lowered disease progression in both MCI and AD subjects, thus confirming its neuroprotective properties [486]. Moreover, data obtained from a randomized controlled trial involving 39 AD patients confirmed slightly slower AD progression in the group administered with LA, with better outcomes reported upon combining LA with omega-3 fatty acids intake [487]. Overall, these data point to LA as a promising anti-AD agent, but further studies are needed to further assess its therapeutic efficiency.

2.6. Mitochondria-Targeted Antioxidants

Impaired glucose metabolism, oxidative stress, and mitochondrial dysfunction are all conditions that characterize human aging and age-related disorders [488]. Concerning neurodegeneration, mitochondria dysfunction contributes to AD pathophysiology through insufficient ATP biogenesis, diminished oxidative phosphorylation, imbalanced mitochondria biogenesis-mitophagy, and impaired antioxidant defenses [489,490,491,492]. To date, several mitochondria-targeted antioxidants have been proposed as alternative therapeutic molecules in AD, including CoQ10, SS31, mitoquinone (MitoQ), methylene blue (MB), and SkQs (Table 2) [493,494,495,496,497,498,499,500,501,502,503,504,505,506,507,508,509,510,511,512,513,514].
Among mitochondria-targeted molecules, CoQ10 is a lipid-soluble antioxidant belonging to the electron transport chain, which plays a key role in mitochondrial oxidative phosphorylation, inflammation, metabolism, and gene expression [515]. When it is lacking, CoQ10 has been linked to several disorders, such as cancer, fibromyalgia, cardiovascular disease, diabetes, and neurodegeneration [516]. In the context of AD, CFS samples obtained from AD patients display an enrichment in oxidized CoQ10, compared to those from aged-matched healthy controls, thus suggesting that mitochondrial oxidative damage may be a helpful parameter for disease diagnosis [517]. However, the use of CoQ10 as an early or prodromal AD biomarker remains unlikely, as no association between serum CoQ10 levels and MCI was reported in the randomized controlled trial of the ActiFE study [518]. Nevertheless, despite the diagnostic potential of CoQ10 remaining unclear, several preclinical studies are consistent in showing a therapeutic advantage from CoQ10 intake (Table 2). Accordingly, various rat and mouse models of AD CoQ10 administration resulted in improved memory, behavior, and cognitive performance, as well as the prevention of AD-associated hippocampal long-term potentiation impairment [503,506,509,512]. These effects are mediated by increased antioxidant capacity (GSH, CAT, and SOD), improved mitochondrial respiratory enzyme activity, and reduced levels of lipid peroxides, serum MDA, and brain protein carbonyls, which are considered the hallmarks of OS [503,505,506,509,511,512]. Notably, amyloid pathology (as measured by the brain levels of Aβ1–42 and the amyloid plaque burden) was consistently decreased across studies, indicating a promising role of CoQ10 in preventing Aβ-induced neurotoxicity [503,505,506,514]. Although the free radical scavenging activity of CoQ10 may appear as the only mediator of these benefits, the anti-apoptotic and anti-inflammatory properties of this antioxidant are equally important in the AD context [503,519]. In this respect, there is evidence that CoQ10 promotes neurogenesis through the PI3K pathway and limits neuroinflammation by decreasing the levels of the pro-inflammatory molecules, COX-2, prostaglandin E2 (PGE2), NF-kB, TNFα, IL-6, IL-1β, and apolipoprotein E [512,520,521,522]. In line with this evidence, the treatment of AD rats with a combination of CoQ10 and minocycline, a microglial inhibitor, further ameliorated AD symptomatology and improved cognitive performance [512]. Other combinations, such as CoQ10 with vinpocetine (a synthetic derivative of the Vinca alkaloid, vincamine) and physical/mental practice, or CoQ10 with omega-3 intake are also promising [504,523]. In the latter case, significant improvements in the OS, inflammatory, amyloidogenic, and cholinergic parameters were reported in hypercholesterolemia-induced AD rats and were correlated with increased memory and brain functions [504].
Despite the encouraging therapeutic potential of CoQ10, poor brain accumulation remains a matter of concern. Indeed, although CoQ10 can pass through the BBB, its opposite transport to the blood may limit its therapeutic efficacy [524]. In this respect, CoQ10 analogs that passively cross the BBB, such as idebenone, have been reported to reduce tau hyperphosphorylation, caspase 3 activity, and the amyloid burden through the upregulation of the α-secretase and neprilysin enzymes, which reduce the Aβ production/burden [508]. Moreover, a new water-dispersible formulation, based on ubisol-Q10, in which an amphiphilic molecule is linked to CoQ10, enabling the formation of nanomicelles, has recently been proven to be effective in reducing Aβ plaques and increasing long-term and spatial working memory in transgenic AD mice [507]. Compared to CoQ10, the ubisol-Q10-mediated upregulation of the autophagy pathway may explain its potent neuroprotection, even if delivered after, as well as before, or concomitantly with, neurotoxin exposure [510,525]. Currently, new delivery methods are being studied to increase its targeting and efficiency. For example, Sheykhhasan et al. recently showed that CoQ10 delivery via adipose-derived stem cell exosomes is more efficient than CoQ10 alone in stimulating the memory and increasing the hippocampal expression of BDNF and SRY-box transcription factor 2 (SOX2) in AD rats, while additional studies might be published shortly [513].
Overall, although the use of CoQ10 for diagnostic purposes remains unclear, its therapeutic potential has been repeatedly proven in different preclinical models, paving the way for clinical trials.
MitoQ is a lipid-soluble mitochondria-targeted molecule composed of the CoQ10 ubiquinone, which is linked to the lipophilic triphenyl phosphonium (TPP) cation, making it a powerful antioxidant [526,527], although its effect is only directed toward the mitochondrial membrane [497]. When tested in several in vitro models of AD, MitoQ’s ability to limit ROS generation, re-establish the optimal mitochondrial membrane potential, and promote mitochondrial function (as measured by reduced cyclophilin D and increased peroxiredoxins) has been proven capable of stimulating neurite outgrowth, preventing tau-dimerization, and protecting against Aβ-induced neurotoxicity [496,528,529]. In line with these data, nematode models of AD treated with MitoQ showed an extended lifespan and increased mitochondria cardiolipin [497], an essential component of the mitochondrial membrane [530]. Similarly, 3xTg AD mice that received MitoQ in drinking water for 5 months exhibited a reduction in astrogliosis, synapse loss, Aβ aggregation, microglial activation, and tau hyperphosphorylation, which was associated with improved memory ability and delayed death [495,496,531]. However, despite promising preclinical data, to date, no clinical studies have been reported on MitoQ efficacy in AD patients, thus emphasizing the need for further investigations.
SkQ1 is a mitochondria-targeted antioxidant belonging to the class of SkQs molecules, which are composed of a lipophilic cation, a plastoquinone molecule that functions as an antioxidant moiety, and an alkane linker region [300,502]. To date, data obtained mainly from preclinical studies have demonstrated the effectiveness of SkQ1 against aging and AD [532]. Indeed, it has been observed that hippocampal slices obtained from rats pre-treated with SkQ1 show the restored induction of LTP, an indicator of synaptic plasticity, compared to slices treated with Aβ alone [502]. In vivo, lessened neurodegeneration and reduced anxiety, together with improved locomotor and exploratory activity have been reported in OXYS rats upon receiving an SkQ1 intake, although learning ability and neurogenesis were not promoted [498,501,533]. These beneficial effects are mediated by a reduction in mitochondrial and synaptic damage, hippocampal Aβ1–40 and Aβ1–42 accumulation, neuronal loss, and tau hyperphosphorylation, together with enhanced levels of the synaptic proteins (which regulate the release of neurotransmitters) and neurotrophic factors [499,533]. In addition, the anti-inflammatory properties of SkQ1 help to slow down AD progression by inhibiting the p38 MAPK signaling pathway and shifting the activated microglia toward a resting state, thus limiting the neurotoxicity [498,500]. Outside the brain, there is evidence that SkQ1 administration prevents the development of retinopathy due to Aβ accumulation, but more data are needed to better assess its beneficial role against retinal damage [501,534].

2.7. Minerals

Since the pathogenesis of AD appears to be closely related to the impact of OS on the promotion of neurodegenerative mechanisms, the role of minerals in AD has been explored in the literature since they are essential for the antioxidant activity of many enzymes [535,536]. In fact, the detection of a reduced plasmatic concentration of trace elements, such as selenium, zinc, iron, and copper, in AD patients may suggest a possible target of the pathology that could, therefore, be partially counteracted by means of a specific dietary intervention in these individuals [537,538]. In this regard, several studies reported the numerous beneficial effects of minerals against AD (Table 3) [539,540,541,542,543,544,545,546,547,548,549,550,551,552,553,554,555,556,557,558,559,560,561,562,563,564,565,566,567,568,569,570,571,572,573,574,575,576,577,578,579,580,581,582,583].
Selenium. The importance of selenium (Se) is mainly related to its presence at the active center of the so-called selenoproteins, the neuroprotective role of which is mainly related to their antioxidant activity via the reduction of ROS and RNS, the regulation of calcium transport, and their anti-inflammatory effects [584]. Se deficiency was observed to be in direct correlation with the reduction of GPx in patients with AD [585]. Evidence from the literature shows lower plasma levels of Se, both in elderly patients with general cognitive decline and in patients diagnosed with MCI or AD, as well as in the brain and the CSF of AD individuals [585,586,587,588]. From these considerations, there follows the hypothesis of Se supplementation as a supportive therapy for AD; in fact, its implementation in patients via sodium selenate, which elevates Se levels in the CSF, or via Se-rich nanoparticles, which reduce ROS and, consequently, counteract Aβ deposition, was observed to have a potential therapeutic role for AD [540,589]. In addition, studies in rats pointed to the existence of molecular mechanisms supporting the use of Se supplementation to counteract AD: (i) Se prevents the reduction of antioxidant enzymes such as GPx, (ii) it reduces membrane permeability for Ca2+ by acting as an NMDA receptor antagonist, (iii) it contributes to rising membrane phospholipid levels, which act as an indirect marker of the synaptogenetic process [541,542,543]. All these effects were also noticed in mice treated with the bioactive organic form of selenomethionine (Se-Met), leading to better cognitive performance; additionally, Se-Met was observed to decrease the tau levels, both in its total amount and in its phosphorylated form, to regulate the autophagic process, and to be able to reverse synaptic dysfunction [544,545,546,547]. Furthermore, sodium selenate may represent a promising supportive therapy because of its ability to decrease tau aggregation and, more generally, to modify the cortical proteome contrasting AD in murine models [548,549,590]. Finally, Se also seems to possess beneficial effects in co-supplementation with resveratrol, folic acid, and probiotics, with which the antioxidant effects appear to be even more significant in comparison to Se when administered alone [550,551,552,591].
The effects of Se, particularly when referring to selenoprotein P, also seem to have an impact on zinc balance, the alteration of which was observed to be linked to tau hyperphosphorylation in SELENOP1 knockout mice [592].
Zinc. There is some evidence in the literature supporting the finding that zinc (Zn) can be considered a neuroprotective factor against AD, which is indeed often accompanied by reduced levels of Zn itself. In particular, this trace element was observed to be avidly bound by amyloid plaques, exerting an antioxidant effect, and its supplementation was linked to a reduction in monomeric Aβ and to significant inhibiting activity on anticholinesterase enzymes in the in vitro studies [554,555,556]. The in vivo studies in murine models show that Zn treatment, delivered by nanoparticles, is associated with a downsizing of the plaques [557]. Furthermore, Zn was demonstrated to effectively counteract the neural damage induced by other metals, such as aluminum (Al) and copper (Cu). In more detail, zinc sulfate administration is able to restore normal levels of tau, APP, and α-synuclein and to ameliorate the histological architecture altered by Al in mice; furthermore, Zn counteracts the toxic effects of Cu (which fuels ROS formation) at the brain level, limiting its intestinal absorption in AD patients [558,559,560,593]. Zn was also observed to alleviate cognitive impairment in rats when co-administered with Se, as these substances seem to stabilize the mitochondrial membranes and protect them from oxidative stress by increasing the SOD and GPx levels; in addition, Zn, together with Se and fish oil (eicosapentaenoic acid and docosahexaenoic acid), inhibits APP cleavage [561,562].
Interestingly, in vivo and in vitro studies suggest that an excess of Zn can represent a risk factor for the development of AD, as it promotes APP expression, β-secretase cleavage, and Aβ deposition and aggravates tauopathy, resulting in learning and memory impairment [594,595]. Based on this rationale, Zn chelation could be considered a therapeutic strategy in this condition: in fact, clioquinol has been reported to reduce Aβ plaques [563,596]. The same mechanism seems to be exploited by the S100A6 protein, which is synthesized in astrocytes surrounding Aβ deposits and shows a disaggregating capacity due to Zn sequestration [564].
Notwithstanding, the literature still shows conflicting data; for instance, chronic Zn supplementation did not modify Aβ and tau deposition in a study on murine models [597]. Therefore, further studies are needed to solve this controversy regarding the role of Zn in AD pathophysiology and its homeostasis.
Iron. Iron excess and its local deposition, particularly in the basal ganglia, was observed to have a pro-inflammatory and pro-aggregating effect, leading to Aβ plaque formation; additionally, its accumulation impacts cerebral perfusion and, also, on synapse plasticity via a decrease in the levels of furin, which is an enzyme that regulates the maturation of the BDNF and other proteins linked to synaptogenesis [598,599,600,601]. These observations support the hypothesis that iron chelators could represent a valid therapeutic tool, as was also suggested by experiments demonstrating that deferoxamine and deferiprone (iron chelators) are able to reduce Aβ aggregation in the brains of murine AD models and attenuate cognitive impairment [565,566,567,568]. The same mechanism is exploited not only by well-known antioxidants, such as hesperidin and naringin, but also by coumarin, curcumin, capsaicin, and S-allylcysteine, which decrease the Aβ burden by binding iron [569,570,571]. New formulations of iron chelators that target multiple factors involved in AD pathogenesis include M30, a compound that also decreases APP expression and tau phosphorylation, and aroyl nicotinoyl hydrazones (especially SNH6), which also reduce oxidative stress via the elevation of NAD+ levels and the subsequent implementation of sirtuin activity, which results in protection against axonal damage [572,573,574]. Another strategy to contrast with local iron deposition consists of using chelators, such as PBT434 (which hinders iron uptake by the endothelial cells within the BBB by binding the metal in the interstitium and, at the same time, stimulating iron efflux via an increase in ferrous iron in the intracellular compartment), and by using nanoparticles as carriers of the chelators themselves, which can then cross the BBB [575,576].
Despite all these findings, it is interesting to note that the literature also reported the beneficial effects related to iron administration, which seems to inhibit Aβ1–42 accumulation in mice models, thus suggesting the need for a balanced amount of this trace element, in order to promote a healthy condition [577].
Copper. Another contributor to oxidative stress is Cu, the overload of which was observed to determine tau hyperphosphorylation and, consequently, the formation of neurofibrillary tangles in AD patients. In addition, Cu tends to bind certain regions of Aβ when released following APP processing, exacerbating fibril deposition [602,603]. The pathogenic role of Cu is also supported by the finding that the supplementation of this trace element deteriorates the mitochondrial functioning in the hippocampus and induces axonal damage in AD mice models by means of an altered phosphorylation of the CAMK2α and ERK1/2 kinases [604]. Based on this rationale, molecules with Cu chelation activity, such as PA1637 and TDMQ20, have been developed and tested in murine models, showing the capacity to ameliorate memory deficits and generally improve cognitive and behavioral performance, respectively [578,579]. Interestingly, other well-known antioxidants, such as the flavonoid, fisetin, show the ability to neutralize the negative effects of Cu overload by binding this metal with iron [580]. Despite the cognitive improvements, the literature reports the limited impact of Cu chelators on the molecular mechanisms underlying tau neuropathology [581]. However, the consideration that AD is a multifactorial disease leads current researchers to focus on the development of multi-target and multi-functional ligands. Within this context, compounds characterized by the ability to combine the beneficial effects of Cu chelators with other activities, such as acetylcholinesterase and butyrylcholinesterase inhibition, could represent a promising therapeutic strategy [582,583].

3. Discussion

With the current exponential increase in the elderly population, the search for predictive and prognostic biomarkers for AD’s onset and progression, as well as the implementation of innovative approaches to prevent and treat neurodegeneration, are of the utmost importance. Notably, lifestyle and nutrition have turned out to be crucial regulators of the human lifespan [605,606,607,608]. It was shown that physical and mental exercise, dietary habits, antioxidant intake, vitamin supplementation, and protection from pesticide exposure reduce the risk of developing sporadic AD later in life [609,610,611]. Although not effective enough as therapeutics, these tools should be considered lifelong preventative approaches to antagonize aging and age-related neurodegeneration, with profound social implications [612]. Nevertheless, most single-factor interventions that have been tested so far turned out to be ineffective, likely due to the complex and multifactorial etiology of AD [612]. In this context, a multivariate preventative intervention, targeting several disease-causing mechanisms simultaneously, would probably be more beneficial [612,613]. Accordingly, the results from the randomized controlled FINGER trial showed that multidomain lifestyle interventions could improve cognition among old people who are at high risk of developing dementia, encouraging further investigation [612,613].
OS has long been considered to participate actively in the pathophysiology of AD, making it an excellent diagnostic and therapeutic tool [614]. For this reason, in this narrative review, we have carefully summarized the potential of carotenoids, vitamins, flavonoids, non-flavonoids, organosulfur compounds, mitochondria-targeted antioxidants, and minerals to serve as biomarkers, as well as antioxidant buffers against AD onset and progression. However, despite the promising results, some limitations still exist and need to be addressed before moving forward to the clinical application. For instance, one of the greatest concerns is the definition of the duration of the therapeutic treatment. Indeed, while most studies limited the antioxidant administration to a defined time window, evidence of the effect of long-life treatment remains poor. In this context, since OS actively participates in the AD etiology [615], an early and lasting intervention could be more effective, if not even preventive, than a late and heavy treatment [616]. Moreover, the metabolic changes that normally occur during aging [617] should be taken into account when designing the dosage and administration window. Nevertheless, it is often difficult to clearly identify the optimal dosage of antioxidants, which depends not only upon the defined treatment but also on dietary and supplement intake. In this respect, it has been reported that antioxidants can act as pro-oxidants when accumulated in excessive amounts [300,618], underlying the importance of patient-to-patient evaluation. This holds particularly true when considering the fact that older people make use of several chronic medications, which may end up interfering with the expected activity of the recommended supplementation. If antioxidant combinations are then considered, careful assessment of synergism and antagonism among various compounds should be conducted, as the simultaneous administration of several antioxidants does not always represent the best option [619]. Even when taking into account each compound alone, storage conditions and environmental factors may greatly affect the reducing potential [620], partially explaining the inconsistency in the data that are sometimes reported by independent studies.

4. Conclusions

Considering clinical translation, a major limitation of this work remains the lack of human studies for most of the antioxidant compounds presented in this review. The abundance of data coming from murine or in vitro approaches compared to human investigations hinders clinical development. Further research should focus on identifying the optimal dosage and treatment window, so as to plan large randomized clinical trials aimed at better assessing the diagnostic and therapeutic potential of these innovative strategies.

Author Contributions

Conceptualization, A.V. and A.P.; writing—original draft preparation, A.V., L.I.M.C., A.C., I.P., E.F. and S.C.; writing—review and editing, A.V., L.I.M.C. and A.P.; supervision, G.R. and A.P. 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.

References

  1. DeTure, M.A.; Dickson, D.W. The neuropathological diagnosis of Alzheimer’s disease. Mol. Neurodegener. 2019, 14, 32. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Hodson, R. Alzheimer’s disease. Nature 2018, 559, S1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Husna Ibrahim, N.; Yahaya, M.F.; Mohamed, W.; Teoh, S.L.; Hui, C.K.; Kumar, J. Pharmacotherapy of Alzheimer’s Disease: Seeking Clarity in a Time of Uncertainty. Front. Pharmacol. 2020, 11, 261. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Breijyeh, Z.; Karaman, R. Comprehensive Review on Alzheimer’s Disease: Causes and Treatment. Molecules 2020, 25, 5789. [Google Scholar] [CrossRef] [PubMed]
  5. Huang, W.-J.; Zhang, X.; Chen, W.-W. Role of oxidative stress in Alzheimer’s disease. Biomed. Rep. 2016, 4, 519–522. [Google Scholar] [CrossRef] [Green Version]
  6. Nunomura, A.; Castellani, R.J.; Zhu, X.; Moreira, P.I.; Perry, G.; Smith, M.A. Involvement of Oxidative Stress in Alzheimer Disease. J. Neuropathol. Exp. Neurol. 2006, 65, 631–641. [Google Scholar] [CrossRef] [Green Version]
  7. Buccellato, F.R.; D’Anca, M.; Fenoglio, C.; Scarpini, E.; Galimberti, D. Role of Oxidative Damage in Alzheimer’s Disease and Neurodegeneration: From Pathogenic Mechanisms to Biomarker Discovery. Antioxidants 2021, 10, 1353. [Google Scholar] [CrossRef]
  8. Baldeiras, I.; Santana, I.; Proença, M.T.; Garrucho, M.H.; Pascoal, R.; Rodrigues, A.; Duro, D.; Oliveira, C.R. Peripheral Oxidative Damage in Mild Cognitive Impairment and Mild Alzheimer’s Disease. J. Alzheimers Dis. 2008, 15, 117–128. [Google Scholar] [CrossRef]
  9. Greilberger, J.; Koidl, C.; Greilberger, M.; Lamprecht, M.; Schroecksnadel, K.; Leblhuber, F.; Fuchs, D.; Oettl, K. Malondialdehyde, carbonyl proteins and albumin-disulphide as useful oxidative markers in mild cognitive impairment and Alzheimer’s disease. Free Radic. Res. 2008, 42, 633–638. [Google Scholar] [CrossRef]
  10. Melo, J.B.; Agostinho, P.; Oliveira, C.R. Involvement of oxidative stress in the enhancement of acetylcholinesterase activity induced by amyloid beta-peptide. Neurosci. Res. 2003, 45, 117–127. [Google Scholar] [CrossRef]
  11. A Massaad, C. Neuronal and Vascular Oxidative Stress in Alzheimers Disease. Curr. Neuropharmacol. 2011, 9, 662–673. [Google Scholar] [CrossRef] [Green Version]
  12. Manoharan, S.; Guillemin, G.J.; Abiramasundari, R.S.; Essa, M.M.; Akbar, M.; Akbar, M.D. The Role of Reactive Oxygen Species in the Pathogenesis of Alzheimer’s Disease, Parkinson’s Disease, and Huntington’s Disease: A Mini Review. Oxid. Med. Cell. Longev. 2016, 2016, 8590578. [Google Scholar] [CrossRef]
  13. Kim, T.-S.; Pae, C.-U.; Yoon, S.-J.; Jang, W.-Y.; Lee, N.J.; Kim, J.-J.; Lee, S.-J.; Lee, C.; Paik, I.-H.; Lee, C.-U. Decreased plasma antioxidants in patients with Alzheimer’s disease. Int. J. Geriatr. Psychiatry 2006, 21, 344–348. [Google Scholar] [CrossRef]
  14. Tönnies, E.; Trushina, E. Oxidative Stress, Synaptic Dysfunction, and Alzheimer’s Disease. J. Alzheimers Dis. 2017, 57, 1105–1121. [Google Scholar] [CrossRef] [Green Version]
  15. Rahman, T.; Hosen, I.; Islam, M.M.T.; Shekhar, H.U. Oxidative stress and human health. Adv. Biosci. Biotechnol. 2012, 03, 997–1019. [Google Scholar] [CrossRef] [Green Version]
  16. Cheignon, C.; Tomas, M.; Bonnefont-Rousselot, D.; Faller, P.; Hureau, C.; Collin, F. Oxidative stress and the amyloid beta peptide in Alzheimer’s disease. Redox Biol. 2018, 14, 450–464. [Google Scholar] [CrossRef]
  17. Montaser, A.; Huttunen, J.; Ibrahim, S.A.; Huttunen, K.M. Astrocyte-Targeted Transporter-Utilizing Derivatives of Ferulic Acid Can Have Multifunctional Effects Ameliorating Inflammation and Oxidative Stress in the Brain. Oxid. Med. Cell. Longev. 2019, 2019, 3528148. [Google Scholar] [CrossRef] [Green Version]
  18. Butterfield, D.A.; Halliwell, B. Oxidative stress, dysfunctional glucose metabolism and Alzheimer disease. Nat. Rev. Neurosci. 2019, 20, 148–160. [Google Scholar] [CrossRef]
  19. Elfawy, H.A.; Das, B. Crosstalk between mitochondrial dysfunction, oxidative stress, and age related neurodegenerative disease: Etiologies and therapeutic strategies. Life Sci. 2019, 218, 165–184. [Google Scholar] [CrossRef]
  20. Papagiouvannis, G.; Theodosis-Nobelos, P.; Kourounakis, P.N.; Rekka, E.A. Multi-Target Directed Compounds with Antioxidant and/or Anti-Inflammatory Properties as Potent Agents for Alzheimer’s Disease. Med. Chem. 2021, 17, 1086–1103. [Google Scholar] [CrossRef]
  21. Winiarska-Mieczan, A.; Baranowska-Wójcik, E.; Kwiecień, M.; Grela, E.R.; Szwajgier, D.; Kwiatkowska, K.; Kiczorowska, B. The Role of Dietary Antioxidants in the Pathogenesis of Neurodegenerative Diseases and Their Impact on Cerebral Oxidoreductive Balance. Nutrients 2020, 12, 435. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Zheng, C.; Zhou, X.-W.; Wang, J.-Z. The dual roles of cytokines in Alzheimer’s disease: Update on interleukins, TNF-α, TGF-β and IFN-γ. Transl. Neurodegener. 2016, 5, 7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Park, H.-A.; Hayden, M.M.; Bannerman, S.; Jansen, J.; Crowe-White, K.M. Anti-Apoptotic Effects of Carotenoids in Neurodegeneration. Molecules 2020, 25, 3453. [Google Scholar] [CrossRef] [PubMed]
  24. Young, A.; Lowe, G. Carotenoids—Antioxidant Properties. Antioxidants 2018, 7, 28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Leermakers, E.T.; Darweesh, S.K.; Baena, C.P.; Moreira, E.M.; Melo van Lent, D.; Tielemans, M.J.; Muka, T.; Vitezova, A.; Chowdhury, R.; Bramer, W.M.; et al. The effects of lutein on cardiometabolic health across the life course: A systematic review and meta-analysis1,2. Am. J. Clin. Nutr. 2016, 103, 481–494. [Google Scholar] [CrossRef] [Green Version]
  26. Tan, B.L.; Norhaizan, M.E. Carotenoids: How Effective Are They to Prevent Age-Related Diseases? Molecules 2019, 24, 1801. [Google Scholar] [CrossRef] [Green Version]
  27. Craft, N.E.; Haitema, T.B.; Garnett, K.M.; Fitch, K.A.; Dorey, C.K. Carotenoid, tocopherol, and retinol concentrations in elderly human brain. J. Nutr. Health Aging 2004, 8, 156–162. [Google Scholar]
  28. Kabir, M.T.; Rahman, M.H.; Shah, M.; Jamiruddin, M.R.; Basak, D.; Al-Harrasi, A.; Bhatia, S.; Ashraf, G.M.; Najda, A.; El-kott, A.F.; et al. Therapeutic promise of carotenoids as antioxidants and anti-inflammatory agents in neurodegenerative disorders. Biomed. Pharmacother. 2022, 146, 112610. [Google Scholar] [CrossRef]
  29. Mullan, K.; Williams, M.A.; Cardwell, C.R.; McGuinness, B.; Passmore, P.; Silvestri, G.; Woodside, J.V.; McKay, G.J. Serum concentrations of vitamin E and carotenoids are altered in Alzheimer’s disease: A case-control study. Alzheimers Dement. Transl. Res. Clin. Interv. 2017, 3, 432–439. [Google Scholar] [CrossRef] [Green Version]
  30. Polidori, M.C.; Mecocci, P. Plasma susceptibility to free radical-induced antioxidant consumption and lipid peroxidation is increased in very old subjects with Alzheimer disease. J. Alzheimers Dis. 2002, 4, 517–522. [Google Scholar] [CrossRef]
  31. Yuan, C.; Chen, H.; Wang, Y.; Schneider, J.A.; Willett, W.C.; Morris, M.C. Dietary carotenoids related to risk of incident Alzheimer dementia (AD) and brain AD neuropathology: A community-based cohort of older adults. Am. J. Clin. Nutr. 2021, 113, 200–208. [Google Scholar] [CrossRef]
  32. Beydoun, M.A.; Beydoun, H.A.; Fanelli-Kuczmarski, M.T.; Weiss, J.; Hossain, S.; Canas, J.A.; Evans, M.K.; Zonderman, A.B. Association of Serum Antioxidant Vitamins and Carotenoids with Incident Alzheimer Disease and All-Cause Dementia Among US Adults. Neurology 2022, 98, e2150–e2162. [Google Scholar] [CrossRef]
  33. Feart, C.; Letenneur, L.; Helmer, C.; Samieri, C.; Schalch, W.; Etheve, S.; Delcourt, C.; Dartigues, J.-F.; Barberger-Gateau, P. Plasma Carotenoids Are Inversely Associated With Dementia Risk in an Elderly French Cohort. J. Gerontol. Ser. A 2016, 71, 683–688. [Google Scholar] [CrossRef] [Green Version]
  34. Min, J.; Min, K. Serum Lycopene, Lutein and Zeaxanthin, and the Risk of Alzheimer’s Disease Mortality in Older Adults. Dement. Geriatr. Cogn. Disord. 2014, 37, 246–256. [Google Scholar] [CrossRef]
  35. Qu, M.; Shi, H.; Wang, K.; Wang, X.; Yu, N.; Guo, B. The Associations of Plasma/Serum Carotenoids with Alzheimer’s Disease: A Systematic Review and Meta-Analysis. J. Alzheimers Dis. 2021, 82, 1055–1066. [Google Scholar] [CrossRef]
  36. Nolan, J.M.; Loskutova, E.; Howard, A.N.; Moran, R.; Mulcahy, R.; Stack, J.; Bolger, M.; Dennison, J.; Akuffo, K.O.; Owens, N.; et al. Macular Pigment, Visual Function, and Macular Disease among Subjects with Alzheimer’s Disease: An Exploratory Study. J. Alzheimers Dis. 2014, 42, 1191–1202. [Google Scholar] [CrossRef] [Green Version]
  37. Wang, W.; Shinto, L.; Connor, W.E.; Quinn, J.F. Nutritional Biomarkers in Alzheimer’s Disease: The Association between Carotenoids, n-3 Fatty Acids, and Dementia Severity. J. Alzheimers Dis. 2008, 13, 31–38. [Google Scholar] [CrossRef]
  38. Palacios, N.; Lee, J.S.; Scott, T.; Kelly, R.S.; Bhupathiraju, S.N.; Bigornia, S.J.; Tucker, K.L. Circulating Plasma Metabolites and Cognitive Function in a Puerto Rican Cohort. J. Alzheimers Dis. 2020, 76, 1267–1280. [Google Scholar] [CrossRef]
  39. Boccardi, V.; Arosio, B.; Cari, L.; Bastiani, P.; Scamosci, M.; Casati, M.; Ferri, E.; Bertagnoli, L.; Ciccone, S.; Rossi, P.D.; et al. Beta-carotene, telomerase activity and Alzheimer’s disease in old age subjects. Eur. J. Nutr. 2020, 59, 119–126. [Google Scholar] [CrossRef]
  40. Jimenez-Jimenez, F.J.; Molina, J.A.; Bustos, F.; Orti-Pareja, M.; Benito-Leon, J.; Tallon-Barranco, A.; Gasalla, T.; Porta, J.; Arenas, J. Serum levels of beta-carotene, alpha-carotene and vitamin A in patients with Alzheimer’s disease. Eur. J. Neurol. 1999, 6, 495–497. [Google Scholar] [CrossRef]
  41. Li, F.-J.; Shen, L.; Ji, H.-F. Dietary Intakes of Vitamin E, Vitamin C, and β-Carotene and Risk of Alzheimer’s Disease: A Meta-Analysis. J. Alzheimers Dis. 2012, 31, 253–258. [Google Scholar] [CrossRef] [PubMed]
  42. Schippling, S.; Kontush, A.; Arlt, S.; Buhmann, C.; Stürenburg, H.-J.; Mann, U.; Müller-Thomsen, T.; Beisiegel, U. Increased lipoprotein oxidation in alzheimer’s disease. Free Radic. Biol. Med. 2000, 28, 351–360. [Google Scholar] [CrossRef] [PubMed]
  43. Stuerenburg, H.J.; Ganzer, S.; Müller-Thomsen, T. Plasma beta carotene in Alzheimer’s disease. Association with cerebrospinal fluid beta-amyloid 1-40, (Abeta40), beta-amyloid 1-42 (Abeta42) and total Tau. Neuro Endocrinol. Lett. 2005, 26, 696–698. [Google Scholar] [PubMed]
  44. Koch, M.; Furtado, J.D.; Cronjé, H.T.; DeKosky, S.T.; Fitzpatrick, A.L.; Lopez, O.L.; Kuller, L.H.; Mukamal, K.J.; Jensen, M.K. Plasma antioxidants and risk of dementia in older adults. Alzheimers Dement. Transl. Res. Clin. Interv. 2021, 7, e12208. [Google Scholar] [CrossRef] [PubMed]
  45. Hira, S.; Saleem, U.; Anwar, F.; Sohail, M.F.; Raza, Z.; Ahmad, B. β-Carotene: A Natural Compound Improves Cognitive Impairment and Oxidative Stress in a Mouse Model of Streptozotocin-Induced Alzheimer’s Disease. Biomolecules 2019, 9, 441. [Google Scholar] [CrossRef] [Green Version]
  46. Butler, M.; Nelson, V.A.; Davila, H.; Ratner, E.; Fink, H.A.; Hemmy, L.S.; McCarten, J.R.; Barclay, T.R.; Brasure, M.; Kane, R.L. Over-the-Counter Supplement Interventions to Prevent Cognitive Decline, Mild Cognitive Impairment, and Clinical Alzheimer-Type Dementia: A Systematic Review. Ann. Intern. Med. 2018, 168, 52. [Google Scholar] [CrossRef]
  47. Liu, X.; Dhana, K.; Furtado, J.D.; Agarwal, P.; Aggarwal, N.T.; Tangney, C.; Laranjo, N.; Carey, V.; Barnes, L.L.; Sacks, F.M. Higher circulating α-carotene was associated with better cognitive function: An evaluation among the MIND trial participants. J. Nutr. Sci. 2021, 10, e64. [Google Scholar] [CrossRef]
  48. Ratto, F.; Franchini, F.; Musicco, M.; Caruso, G.; Di Santo, S.G. A narrative review on the potential of tomato and lycopene for the prevention of Alzheimer’s disease and other dementias. Crit. Rev. Food Sci. Nutr. 2022, 62, 4970–4981. [Google Scholar] [CrossRef]
  49. Chen, W.; Mao, L.; Xing, H.; Xu, L.; Fu, X.; Huang, L.; Huang, D.; Pu, Z.; Li, Q. Lycopene attenuates Aβ1–42 secretion and its toxicity in human cell and Caenorhabditis elegans models of Alzheimer disease. Neurosci. Lett. 2015, 608, 28–33. [Google Scholar] [CrossRef]
  50. Liu, C.-B.; Wang, R.; Yi, Y.-F.; Gao, Z.; Chen, Y.-Z. Lycopene mitigates β-amyloid induced inflammatory response and inhibits NF-κB signaling at the choroid plexus in early stages of Alzheimer’s disease rats. J. Nutr. Biochem. 2018, 53, 66–71. [Google Scholar] [CrossRef]
  51. Sachdeva, A.K.; Chopra, K. Lycopene abrogates Aβ(1–42)-mediated neuroinflammatory cascade in an experimental model of Alzheimer’s disease. J. Nutr. Biochem. 2015, 26, 736–744. [Google Scholar] [CrossRef]
  52. Wang, J.; Li, L.; Wang, Z.; Cui, Y.; Tan, X.; Yuan, T.; Liu, Q.; Liu, Z.; Liu, X. Supplementation of lycopene attenuates lipopolysaccharide-induced amyloidogenesis and cognitive impairments via mediating neuroinflammation and oxidative stress. J. Nutr. Biochem. 2018, 56, 16–25. [Google Scholar] [CrossRef]
  53. Xu, X.; Teng, Y.; Zou, J.; Ye, Y.; Song, H.; Wang, Z. Effects of lycopene on vascular remodeling through the LXR–PI3K–AKT signaling pathway in APP/PS1 mice. Biochem. Biophys. Res. Commun. 2020, 526, 699–705. [Google Scholar] [CrossRef]
  54. Yu, L.; Wang, W.; Pang, W.; Xiao, Z.; Jiang, Y.; Hong, Y. Dietary Lycopene Supplementation Improves Cognitive Performances in Tau Transgenic Mice Expressing P301L Mutation via Inhibiting Oxidative Stress and Tau Hyperphosphorylation. J. Alzheimers Dis. 2017, 57, 475–482. [Google Scholar] [CrossRef]
  55. Fang, Y.; Ou, S.; Wu, T.; Zhou, L.; Tang, H.; Jiang, M.; Xu, J.; Guo, K. Lycopene alleviates oxidative stress via the PI3K/Akt/Nrf2pathway in a cell model of Alzheimer’s disease. PeerJ 2020, 8, e9308. [Google Scholar] [CrossRef]
  56. Huang, C.; Wen, C.; Yang, M.; Gan, D.; Fan, C.; Li, A.; Li, Q.; Zhao, J.; Zhu, L.; Lu, D. Lycopene protects against t-BHP-induced neuronal oxidative damage and apoptosis via activation of the PI3K/Akt pathway. Mol. Biol. Rep. 2019, 46, 3387–3397. [Google Scholar] [CrossRef]
  57. Lim, S.; Hwang, S.; Yu, J.H.; Lim, J.W.; Kim, H. Lycopene inhibits regulator of calcineurin 1-mediated apoptosis by reducing oxidative stress and down-regulating Nucling in neuronal cells. Mol. Nutr. Food Res. 2017, 61, 1600530. [Google Scholar] [CrossRef]
  58. Qu, M.; Jiang, Z.; Liao, Y.; Song, Z.; Nan, X. Lycopene Prevents Amyloid [Beta]-Induced Mitochondrial Oxidative Stress and Dysfunctions in Cultured Rat Cortical Neurons. Neurochem. Res. 2016, 41, 1354–1364. [Google Scholar] [CrossRef]
  59. Qu, M.; Li, L.; Chen, C.; Li, M.; Pei, L.; Chu, F.; Yang, J.; Yu, Z.; Wang, D.; Zhou, Z. Protective effects of lycopene against amyloid β-induced neurotoxicity in cultured rat cortical neurons. Neurosci. Lett. 2011, 505, 286–290. [Google Scholar] [CrossRef]
  60. Hwang, S.; Lim, J.; Kim, H. Inhibitory Effect of Lycopene on Amyloid-β-Induced Apoptosis in Neuronal Cells. Nutrients 2017, 9, 883. [Google Scholar] [CrossRef] [Green Version]
  61. Huang, C.; Gan, D.; Fan, C.; Wen, C.; Li, A.; Li, Q.; Zhao, J.; Wang, Z.; Zhu, L.; Lu, D. The Secretion from Neural Stem Cells Pretreated with Lycopene Protects against tert-Butyl Hydroperoxide-Induced Neuron Oxidative Damage. Oxid. Med. Cell. Longev. 2018, 2018, 5490218. [Google Scholar] [CrossRef] [PubMed]
  62. Xu, Z.; Liu, C.; Wang, R.; Gao, X.; Hao, C.; Liu, C. A combination of lycopene and human amniotic epithelial cells can ameliorate cognitive deficits and suppress neuroinflammatory signaling by choroid plexus in Alzheimer’s disease rat. J. Nutr. Biochem. 2021, 88, 108558. [Google Scholar] [CrossRef] [PubMed]
  63. Bun, S.; Ikejima, C.; Kida, J.; Yoshimura, A.; Lebowitz, A.J.; Kakuma, T.; Asada, T. A Combination of Supplements May Reduce the Risk of Alzheimer’s Disease in Elderly Japanese with Normal Cognition. J. Alzheimers Dis. 2015, 45, 15–25. [Google Scholar] [CrossRef] [PubMed]
  64. Crowe-White, K.M.; Phillips, T.A.; Ellis, A.C. Lycopene and cognitive function. J. Nutr. Sci. 2019, 8, e20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Ning, W.; Lv, R.; Xu, N.; Hou, X.; Shen, C.; Guo, Y.; Fan, Z.; Cao, N.; Liu, X.-P. Lycopene-Loaded Microemulsion Regulates Neurogenesis in Rats with Aβ-Induced Alzheimer’s Disease Rats Based on the Wnt/β-catenin Pathway. Neural Plast. 2021, 2021, 5519330. [Google Scholar] [CrossRef]
  66. Ashok, A.; Singh, N.; Chaudhary, S.; Bellamkonda, V.; Kritikos, A.E.; Wise, A.S.; Rana, N.; McDonald, D.; Ayyagari, R. Retinal Degeneration and Alzheimer’s Disease: An Evolving Link. Int. J. Mol. Sci. 2020, 21, 7290. [Google Scholar] [CrossRef]
  67. Hammond, B.R.; Renzi-Hammond, L. The influence of the macular carotenoids on women’s eye and brain health. Nutr. Neurosci. 2022, 1–7. [Google Scholar] [CrossRef]
  68. Wang, H.; Wang, G.; Billings, R.; Li, D.; Haase, S.R.; Wheeler, P.F.; Vance, D.E.; Li, W. Can Diet Supplements of Macular Pigment of Lutein, Zeaxanthin, and Meso-zeaxanthin Affect Cognition? J. Alzheimers Dis. 2022, 87, 1079–1087. [Google Scholar] [CrossRef]
  69. Ademowo, O.S.; Dias, I.H.K.; Diaz-Sanchez, L.; Sanchez-Aranguren, L.; Stahl, W.; Griffiths, H.R. Partial Mitigation of Oxidized Phospholipid-Mediated Mitochondrial Dysfunction in Neuronal Cells by Oxocarotenoids. J. Alzheimers Dis. 2020, 74, 113–126. [Google Scholar] [CrossRef] [Green Version]
  70. Singhrang, N.; Tocharus, C.; Thummayot, S.; Sutheerawattananonda, M.; Tocharus, J. Protective effects of silk lutein extract from Bombyx mori cocoons on β-Amyloid peptide-induced apoptosis in PC12 cells. Biomed. Pharmacother. 2018, 103, 582–587. [Google Scholar] [CrossRef]
  71. Li, X.; Zhang, P.; Li, H.; Yu, H.; Xi, Y. The Protective Effects of Zeaxanthin on Amyloid-β Peptide 1–42-Induced Impairment of Learning and Memory Ability in Rats. Front. Behav. Neurosci. 2022, 16, 912896. [Google Scholar] [CrossRef]
  72. Zhang, L.-N.; Li, M.-J.; Shang, Y.-H.; Liu, Y.-R.; Han-Chang, H.; Lao, F.-X. Zeaxanthin Attenuates the Vicious Circle Between Endoplasmic Reticulum Stress and Tau Phosphorylation: Involvement of GSK-3β Activation. J. Alzheimers Dis. 2022, 86, 191–204. [Google Scholar] [CrossRef]
  73. Power, R.; Coen, R.F.; Beatty, S.; Mulcahy, R.; Moran, R.; Stack, J.; Howard, A.N.; Nolan, J.M. Supplemental Retinal Carotenoids Enhance Memory in Healthy Individuals with Low Levels of Macular Pigment in A Randomized, Double-Blind, Placebo-Controlled Clinical Trial. J. Alzheimers Dis. 2018, 61, 947–961. [Google Scholar] [CrossRef] [Green Version]
  74. Stringham, N.T.; Holmes, P.V.; Stringham, J.M. Effects of macular xanthophyll supplementation on brain-derived neurotrophic factor, pro-inflammatory cytokines, and cognitive performance. Physiol. Behav. 2019, 211, 112650. [Google Scholar] [CrossRef]
  75. Chew, E.Y.; Clemons, T.E.; Agrón, E.; Launer, L.J.; Grodstein, F.; Bernstein, P.S. Effect of Omega-3 Fatty Acids, Lutein/Zeaxanthin, or Other Nutrient Supplementation on Cognitive Function: The AREDS2 Randomized Clinical Trial. JAMA 2015, 314, 791. [Google Scholar] [CrossRef]
  76. Nolan, J.M.; Loskutova, E.; Howard, A.; Mulcahy, R.; Moran, R.; Stack, J.; Bolger, M.; Coen, R.F.; Dennison, J.; Akuffo, K.O.; et al. The Impact of Supplemental Macular Carotenoids in Alzheimer’s Disease: A Randomized Clinical Trial. J. Alzheimers Dis. 2015, 44, 1157–1169. [Google Scholar] [CrossRef] [Green Version]
  77. Ademowo, O.S.; Dias, H.K.I.; Milic, I.; Devitt, A.; Moran, R.; Mulcahy, R.; Howard, A.N.; Nolan, J.M.; Griffiths, H.R. Phospholipid oxidation and carotenoid supplementation in Alzheimer’s disease patients. Free Radic. Biol. Med. 2017, 108, 77–85. [Google Scholar] [CrossRef] [Green Version]
  78. Nolan, J.M.; Mulcahy, R.; Power, R.; Moran, R.; Howard, A.N. Nutritional Intervention to Prevent Alzheimer’s Disease: Potential Benefits of Xanthophyll Carotenoids and Omega-3 Fatty Acids Combined. J. Alzheimers Dis. 2018, 64, 367–378. [Google Scholar] [CrossRef] [Green Version]
  79. Power, R.; Nolan, J.M.; Prado-Cabrero, A.; Roche, W.; Coen, R.; Power, T.; Mulcahy, R. Omega-3 fatty acid, carotenoid and vitamin E supplementation improves working memory in older adults: A randomised clinical trial. Clin. Nutr. 2022, 41, 405–414. [Google Scholar] [CrossRef]
  80. Dhas, N.; Mehta, T. Cationic biopolymer functionalized nanoparticles encapsulating lutein to attenuate oxidative stress in effective treatment of Alzheimer’s disease: A non-invasive approach. Int. J. Pharm. 2020, 586, 119553. [Google Scholar] [CrossRef]
  81. Grundman, M. Vitamin E and Alzheimer disease: The basis for additional clinical trials. Am. J. Clin. Nutr. 2000, 71, 630S–636S. [Google Scholar] [CrossRef] [PubMed]
  82. Boothby, L.A.; Doering, P.L. Vitamin C and Vitamin E for Alzheimer’s Disease. Ann. Pharmacother. 2005, 39, 2073–2080. [Google Scholar] [CrossRef] [PubMed]
  83. Aisen, P.S.; Schneider, L.S.; Sano, M.; Diaz-Arrastia, R.; van Dyck, C.H.; Weiner, M.F.; Bottiglieri, T.; Jin, S.; Stokes, K.S.; Thomas, R.G.; et al. High-Dose B Vitamin Supplementation and Cognitive Decline in Alzheimer Disease: A Randomized Controlled Trial. JAMA 2008, 300, 1774. [Google Scholar] [CrossRef] [PubMed]
  84. Zhang, D.-M.; Ye, J.-X.; Mu, J.-S.; Cui, X.-P. Efficacy of Vitamin B Supplementation on Cognition in Elderly Patients with Cognitive-Related Diseases: A Systematic Review and Meta-Analysis. J. Geriatr. Psychiatry Neurol. 2017, 30, 50–59. [Google Scholar] [CrossRef] [PubMed]
  85. Mielech, A.; Puścion-Jakubik, A.; Markiewicz-Żukowska, R.; Socha, K. Vitamins in Alzheimer’s Disease—Review of the Latest Reports. Nutrients 2020, 12, 3458. [Google Scholar] [CrossRef]
  86. Fei, H.-X.; Qian, C.-F.; Wu, X.-M.; Wei, Y.-H.; Huang, J.-Y.; Wei, L.-H. Role of micronutrients in Alzheimer’s disease: Review of available evidence. World J. Clin. Cases 2022, 10, 7631–7641. [Google Scholar] [CrossRef]
  87. Browne, D.; McGuinness, B.; Woodside, J.V.; McKay, G.J. Vitamin E and Alzheimer’s disease: What do we know so far? Clin. Interv. Aging 2019, 14, 1303–1317. [Google Scholar] [CrossRef] [Green Version]
  88. Shahidi, F.; Pinaffi-Langley, A.C.C.; Fuentes, J.; Speisky, H.; de Camargo, A.C. Vitamin E as an essential micronutrient for human health: Common, novel, and unexplored dietary sources. Free Radic. Biol. Med. 2021, 176, 312–321. [Google Scholar] [CrossRef]
  89. Jiang, Q. Natural forms of vitamin E: Metabolism, antioxidant, and anti-inflammatory activities and their role in disease prevention and therapy. Free Radic. Biol. Med. 2014, 72, 76–90. [Google Scholar] [CrossRef] [Green Version]
  90. Gugliandolo, A.; Bramanti, P.; Mazzon, E. Role of Vitamin E in the Treatment of Alzheimer’s Disease: Evidence from Animal Models. Int. J. Mol. Sci. 2017, 18, 2504. [Google Scholar] [CrossRef] [Green Version]
  91. Dorey, C.K.; Gierhart, D.; Fitch, K.A.; Crandell, I.; Craft, N.E. Low Xanthophylls, Retinol, Lycopene, and Tocopherols in Grey and White Matter of Brains with Alzheimer’s Disease. J. Alzheimers Dis. 2022, 1–16. [Google Scholar] [CrossRef]
  92. Singh, U.; Jialal, I. Anti-inflammatory Effects of α-Tocopherol. Ann. N. Y. Acad. Sci. 2004, 1031, 195–203. [Google Scholar] [CrossRef]
  93. Singh, U.; Devaraj, S.; Jialal, I. Vitamin E, Oxidative Stress, and Inflammation. Annu. Rev. Nutr. 2005, 25, 151–174. [Google Scholar] [CrossRef]
  94. Ciarcià, G.; Bianchi, S.; Tomasello, B.; Acquaviva, R.; Malfa, G.A.; Naletova, I.; La Mantia, A.; Di Giacomo, C. Vitamin E and Non-Communicable Diseases: A Review. Biomedicines 2022, 10, 2473. [Google Scholar] [CrossRef]
  95. Donnelly, J.; Appathurai, A.; Yeoh, H.-L.; Driscoll, K.; Faisal, W. Vitamin E in Cancer Treatment: A Review of Clinical Applications in Randomized Control Trials. Nutrients 2022, 14, 4329. [Google Scholar] [CrossRef]
  96. Di Donato, I.; Bianchi, S.; Federico, A. Ataxia with vitamin E deficiency: Update of molecular diagnosis. Neurol. Sci. 2010, 31, 511–515. [Google Scholar] [CrossRef]
  97. Ulatowski, L.M.; Manor, D. Vitamin E and neurodegeneration. Neurobiol. Dis. 2015, 84, 78–83. [Google Scholar] [CrossRef]
  98. Nishida, Y.; Ito, S.; Ohtsuki, S.; Yamamoto, N.; Takahashi, T.; Iwata, N.; Jishage, K.; Yamada, H.; Sasaguri, H.; Yokota, S.; et al. Depletion of Vitamin E Increases Amyloid β Accumulation by Decreasing Its Clearances from Brain and Blood in a Mouse Model of Alzheimer Disease. J. Biol. Chem. 2009, 284, 33400–33408. [Google Scholar] [CrossRef] [Green Version]
  99. Grimm, M.O.W.; Stahlmann, C.P.; Mett, J.; Haupenthal, V.J.; Zimmer, V.C.; Lehmann, J.; Hundsdörfer, B.; Endres, K.; Grimm, H.S.; Hartmann, T. Vitamin E: Curse or benefit in Alzheimer’s disease? A systematic investigation of the impact of α-, γ- and δ-tocopherol on Aβ generation and degradation in neuroblastoma cells. J. Nutr. Health Aging 2015, 19, 646–654. [Google Scholar] [CrossRef]
  100. Grimm, M.; Regner, L.; Mett, J.; Stahlmann, C.; Schorr, P.; Nelke, C.; Streidenberger, O.; Stoetzel, H.; Winkler, J.; Zaidan, S.; et al. Tocotrienol Affects Oxidative Stress, Cholesterol Homeostasis and the Amyloidogenic Pathway in Neuroblastoma Cells: Consequences for Alzheimer’s Disease. Int. J. Mol. Sci. 2016, 17, 1809. [Google Scholar] [CrossRef] [Green Version]
  101. Chin, K.-Y.; Tay, S. A Review on the Relationship between Tocotrienol and Alzheimer Disease. Nutrients 2018, 10, 881. [Google Scholar] [CrossRef] [PubMed]
  102. Rinaldi, P.; Polidori, M.C.; 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]
  103. Sung, S.; Yao, Y.; Uryu, K.; Yang, H.; Lee, V.M.-Y.; Trojanowski, J.Q.; Praticò, D. Early Vitamin E supplementation in young but not aged mice reduces Aβ levels and amyloid deposition in a transgenic model of Alzheimer’s disease. FASEB J. 2004, 18, 323–325. [Google Scholar] [CrossRef] [PubMed]
  104. Sinha, M.; Bir, A.; Banerjee, A.; Bhowmick, P.; Chakrabarti, S. Multiple mechanisms of age-dependent accumulation of amyloid beta protein in rat brain: Prevention by dietary supplementation with N-acetylcysteine, α-lipoic acid and α-tocopherol. Neurochem. Int. 2016, 95, 92–99. [Google Scholar] [CrossRef] [PubMed]
  105. Wang, S.; Yang, S.; Liu, W.; Zhang, Y.; Xu, P.; Wang, T.; Ling, T.; Liu, R. Alpha-tocopherol quinine ameliorates spatial memory deficits by reducing beta-amyloid oligomers, neuroinflammation and oxidative stress in transgenic mice with Alzheimer’s disease. Behav. Brain Res. 2016, 296, 109–117. [Google Scholar] [CrossRef]
  106. Qi, X.-L.; Xiu, J.; Shan, K.-R.; Xiao, Y.; Gu, R.; Liu, R.-Y.; Guan, Z.-Z. Oxidative stress induced by beta-amyloid peptide1–42 is involved in the altered composition of cellular membrane lipids and the decreased expression of nicotinic receptors in human SH-SY5Y neuroblastoma cells. Neurochem. Int. 2005, 46, 613–621. [Google Scholar] [CrossRef]
  107. Dai, X.; Sun, Y.; Jiang, Z. Protective Effects of Vitamin E against Oxidative Damage Induced by Aβ1–40 Cu(II) Complexes. Acta Biochim. Biophys. Sin. 2007, 39, 123–130. [Google Scholar] [CrossRef] [Green Version]
  108. Rota, C.; Rimbach, G.; Minihane, A.-M.; Stoecklin, E.; Barella, L. Dietary vitamin E modulates differential gene expression in the rat hippocampus: Potential implications for its neuroprotective properties. Nutr. Neurosci. 2005, 8, 21–29. [Google Scholar] [CrossRef]
  109. Gohil, K.; Schock, B.C.; Chakraborty, A.A.; Terasawa, Y.; Raber, J.; Farese, R.V.; Packer, L.; Cross, C.E.; Traber, M.G. Gene expression profile of oxidant stress and neurodegeneration in transgenic mice deficient in α-tocopherol transfer protein. Free Radic. Biol. Med. 2003, 35, 1343–1354. [Google Scholar] [CrossRef]
  110. Block, M.L. NADPH oxidase as a therapeutic target in Alzheimer’s disease. BMC Neurosci. 2008, 9, S8. [Google Scholar] [CrossRef] [Green Version]
  111. Chu, J.; Praticò, D. 5-lipoxygenase as an endogenous modulator of amyloid beta formation in vivo. Ann. Neurol. 2011, 69, 34–46. [Google Scholar] [CrossRef]
  112. Voronkov, M.; Braithwaite, S.P.; Stock, J.B. Phosphoprotein phosphatase 2A: A novel druggable target for Alzheimer’s disease. Future Med. Chem. 2011, 3, 821–833. [Google Scholar] [CrossRef] [Green Version]
  113. Nohturfft, A.; Yabe, D.; Goldstein, J.L.; Brown, M.S.; Espenshade, P.J. Regulated Step in Cholesterol Feedback Localized to Budding of SCAP from ER Membranes. Cell 2000, 102, 315–323. [Google Scholar] [CrossRef] [Green Version]
  114. Fassbender, K.; Simons, M.; Bergmann, C.; Stroick, M.; Lütjohann, D.; Keller, P.; Runz, H.; Kühl, S.; Bertsch, T.; von Bergmann, K.; et al. Simvastatin strongly reduces levels of Alzheimer’s disease β-amyloid peptides Aβ42 and Aβ40 in vitro and in vivo. Proc. Natl. Acad. Sci. USA 2001, 98, 5856–5861. [Google Scholar] [CrossRef] [Green Version]
  115. Maulik, M.; Westaway, D.; Jhamandas, J.H.; Kar, S. Role of Cholesterol in APP Metabolism and Its Significance in Alzheimer’s Disease Pathogenesis. Mol. Neurobiol. 2013, 47, 37–63. [Google Scholar] [CrossRef]
  116. Refolo, L.M.; Pappolla, M.A.; Malester, B.; LaFrancois, J.; Bryant-Thomas, T.; Wang, R.; Tint, G.S.; Sambamurti, K.; Duff, K. Hypercholesterolemia Accelerates the Alzheimer’s Amyloid Pathology in a Transgenic Mouse Model. Neurobiol. Dis. 2000, 7, 321–331. [Google Scholar] [CrossRef]
  117. Grimm, M.O.W.; Grimm, H.S.; Tomic, I.; Beyreuther, K.; Hartmann, T.; Bergmann, C. Independent Inhibition of Alzheimer Disease β- and γ-Secretase Cleavage by Lowered Cholesterol Levels. J. Biol. Chem. 2008, 283, 11302–11311. [Google Scholar] [CrossRef] [Green Version]
  118. Osenkowski, P.; Ye, W.; Wang, R.; Wolfe, M.S.; Selkoe, D.J. Direct and Potent Regulation of γ-Secretase by Its Lipid Microenvironment. J. Biol. Chem. 2008, 283, 22529–22540. [Google Scholar] [CrossRef] [Green Version]
  119. Cossec, J.-C.; Simon, A.; Marquer, C.; Moldrich, R.X.; Leterrier, C.; Rossier, J.; Duyckaerts, C.; Lenkei, Z.; Potier, M.-C. Clathrin-dependent APP endocytosis and Aβ secretion are highly sensitive to the level of plasma membrane cholesterol. Biochim. Biophys. Acta BBA—Mol. Cell Biol. Lipids 2010, 1801, 846–852. [Google Scholar] [CrossRef]
  120. Giraldo, E.; Lloret, A.; Fuchsberger, T.; Viña, J. Aβ and tau toxicities in Alzheimer’s are linked via oxidative stress-induced p38 activation: Protective role of vitamin E. Redox Biol. 2014, 2, 873–877. [Google Scholar] [CrossRef] [Green Version]
  121. Nakashima, H.; Ishihara, T.; Yokota, O.; Terada, S.; Trojanowski, J.Q.; Lee, V.M.-Y.; Kuroda, S. Effects of α-tocopherol on an animal model of tauopathies. Free Radic. Biol. Med. 2004, 37, 176–186. [Google Scholar] [CrossRef] [PubMed]
  122. Bourdel-Marchasson, I.; Delmas-Beauvieux, M.C.; Peuchant, E.; Richard-Harston, S.; Decamps, A.; Reignier, B.; Emeriau, J.P.; Rainfray, M. Antioxidant defences and oxidative stress markers in erythrocytes and plasma from normally nourished elderly Alzheimer patients. Age Ageing 2001, 30, 235–241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Mangialasche, F.; Xu, W.; Kivipelto, M.; Costanzi, E.; Ercolani, S.; Pigliautile, M.; Cecchetti, R.; Baglioni, M.; Simmons, A.; Soininen, H.; et al. Tocopherols and tocotrienols plasma levels are associated with cognitive impairment. Neurobiol. Aging 2012, 33, 2282–2290. [Google Scholar] [CrossRef] [PubMed]
  124. Lopes da Silva, S.; Vellas, B.; Elemans, S.; Luchsinger, J.; Kamphuis, P.; Yaffe, K.; Sijben, J.; Groenendijk, M.; Stijnen, T. Plasma nutrient status of patients with Alzheimer’s disease: Systematic review and meta-analysis. Alzheimers Dement. 2014, 10, 485–502. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Morris, M.C.; Evans, D.A.; Tangney, C.C.; Bienias, J.L.; Wilson, R.S.; Aggarwal, N.T.; Scherr, P.A. Relation of the tocopherol forms to incident Alzheimer disease and to cognitive change. Am. J. Clin. Nutr. 2005, 81, 508–514. [Google Scholar] [CrossRef] [Green Version]
  126. Mangialasche, F.; Kivipelto, M.; Mecocci, P.; Rizzuto, D.; Palmer, K.; Winblad, B.; Fratiglioni, L. High Plasma Levels of Vitamin E Forms and Reduced Alzheimer’s Disease Risk in Advanced Age. J. Alzheimers Dis. 2010, 20, 1029–1037. [Google Scholar] [CrossRef] [Green Version]
  127. Dysken, M.W.; Sano, M.; Asthana, S.; Vertrees, J.E.; Pallaki, M.; Llorente, M.; Love, S.; Schellenberg, G.D.; McCarten, J.R.; Malphurs, J.; et al. Effect of Vitamin E and Memantine on Functional Decline in Alzheimer Disease: The TEAM-AD VA Cooperative Randomized Trial. JAMA 2014, 311, 33. [Google Scholar] [CrossRef]
  128. Petersen, R.C.; Thomas, R.G.; Grundman, M.; Bennett, D.; Doody, R.; Ferris, S.; Galasko, D.; Jin, S.; Kaye, J.; Levey, A.; et al. Vitamin E and Donepezil for the Treatment of Mild Cognitive Impairment. N. Engl. J. Med. 2005, 352, 2379–2388. [Google Scholar] [CrossRef] [Green Version]
  129. Galasko, D.R.; Peskind, E.; Clark, C.M.; Quinn, J.F.; Ringman, J.M.; Jicha, G.A.; Cotman, C.; Cottrell, B.; Montine, T.J.; Thomas, R.G.; et al. Antioxidants for Alzheimer Disease: A Randomized Clinical Trial With Cerebrospinal Fluid Biomarker Measures. Arch. Neurol. 2012, 69, 836–841. [Google Scholar] [CrossRef] [Green Version]
  130. Wołoszynowska-Fraser, M.U.; Kouchmeshky, A.; McCaffery, P. Vitamin A and Retinoic Acid in Cognition and Cognitive Disease. Annu. Rev. Nutr. 2020, 40, 247–272. [Google Scholar] [CrossRef]
  131. Ono, K.; Yoshiike, Y.; Takashima, A.; Hasegawa, K.; Naiki, H.; Yamada, M. Vitamin A exhibits potent antiamyloidogenic and fibril-destabilizing effects in vitro. Exp. Neurol. 2004, 189, 380–392. [Google Scholar] [CrossRef]
  132. Ono, K.; Yamada, M. Vitamin A potently destabilizes preformed α-synuclein fibrils in vitro: Implications for Lewy body diseases. Neurobiol. Dis. 2007, 25, 446–454. [Google Scholar] [CrossRef]
  133. Ono, K.; Yamada, M. Vitamin A and Alzheimer’s disease: Vitamin A and Alzheimer’s disease. Geriatr. Gerontol. Int. 2012, 12, 180–188. [Google Scholar] [CrossRef]
  134. Takasaki, J.; Ono, K.; Yoshiike, Y.; Hirohata, M.; Ikeda, T.; Morinaga, A.; Takashima, A.; Yamada, M. Vitamin A has Anti-Oligomerization Effects on Amyloid-β In Vitro. J. Alzheimers Dis. 2011, 27, 271–280. [Google Scholar] [CrossRef] [Green Version]
  135. Huang, S.-H.; Fang, S.-T.; Chen, Y.-C. Molecular Mechanism of Vitamin K2 Protection against Amyloid-β-Induced Cytotoxicity. Biomolecules 2021, 11, 423. [Google Scholar] [CrossRef]
  136. Booth, S.L.; Shea, M.K.; Barger, K.; Leurgans, S.E.; James, B.D.; Holland, T.M.; Agarwal, P.; Fu, X.; Wang, J.; Matuszek, G.; et al. Association of vitamin K with cognitive decline and neuropathology in community-dwelling older persons. Alzheimers Dement. Transl. Res. Clin. Interv. 2022, 8, e12255. [Google Scholar] [CrossRef]
  137. Emekli-Alturfan, E.; Alturfan, A.A. The emerging relationship between vitamin K and neurodegenerative diseases: A review of current evidence. Mol. Biol. Rep. 2022. [Google Scholar] [CrossRef]
  138. Wimalawansa, S.J. Vitamin D Deficiency: Effects on Oxidative Stress, Epigenetics, Gene Regulation, and Aging. Biology 2019, 8, 30. [Google Scholar] [CrossRef] [Green Version]
  139. Mantle, D.; Hargreaves, I.P. Mitochondrial Dysfunction and Neurodegenerative Disorders: Role of Nutritional Supplementation. Int. J. Mol. Sci. 2022, 23, 12603. [Google Scholar] [CrossRef]
  140. Ricca, C.; Aillon, A.; Bergandi, L.; Alotto, D.; Castagnoli, C.; Silvagno, F. Vitamin D Receptor Is Necessary for Mitochondrial Function and Cell Health. Int. J. Mol. Sci. 2018, 19, 1672. [Google Scholar] [CrossRef] [Green Version]
  141. Le Floch, M.; Gautier, J.; Annweiler, C. Vitamin D Concentration and Motoric Cognitive Risk in Older Adults: Results from the Gait and Alzheimer Interactions Tracking (GAIT) Cohort. Int. J. Environ. Res. Public. Health 2022, 19, 13086. [Google Scholar] [CrossRef] [PubMed]
  142. Patel, P.; Shah, J. Vitamin D3 supplementation ameliorates cognitive impairment and alters neurodegenerative and inflammatory markers in scopolamine induced rat model. Metab. Brain Dis. 2022, 37, 2653–2667. [Google Scholar] [CrossRef] [PubMed]
  143. Malouf, R.; Grimley Evans, J.; Areosa Sastre, A. Folic acid with or without vitamin B12 for cognition and dementia. In The Cochrane Database of Systematic Reviews; The Cochrane Collaboration, Ed.; John Wiley & Sons, Ltd.: Chichester, UK, 2003; p. CD004514. [Google Scholar]
  144. Malouf, R.; Grimley Evans, J. Folic acid with or without vitamin B12 for the prevention and treatment of healthy elderly and demented people. Cochrane Database Syst. Rev. 2008, 8, CD004514. [Google Scholar] [CrossRef] [PubMed]
  145. Jager, C.A.; Oulhaj, A.; Jacoby, R.; Refsum, H.; Smith, A.D. Cognitive and clinical outcomes of homocysteine-lowering B-vitamin treatment in mild cognitive impairment: A randomized controlled trial: Treatment of mild cognitive impairment. Int. J. Geriatr. Psychiatry 2012, 27, 592–600. [Google Scholar] [CrossRef] [PubMed]
  146. Song, Y.; Quan, M.; Li, T.; Jia, J. Serum Homocysteine, Vitamin B12, Folate, and Their Association with Mild Cognitive Impairment and Subtypes of Dementia. J. Alzheimers Dis. 2022, 90, 681–691. [Google Scholar] [CrossRef]
  147. Li, M.-M.; Yu, J.-T.; Wang, H.-F.; Jiang, T.; Wang, J.; Meng, X.-F.; Tan, C.-C.; Wang, C.; Tan, L. Efficacy of vitamins B supplementation on mild cognitive impairment and Alzheimer’s disease: A systematic review and meta-analysis. Curr. Alzheimer Res. 2014, 11, 844–852. [Google Scholar]
  148. Li, S.; Guo, Y.; Men, J.; Fu, H.; Xu, T. The preventive efficacy of vitamin B supplements on the cognitive decline of elderly adults: A systematic review and meta-analysis. BMC Geriatr. 2021, 21, 367. [Google Scholar] [CrossRef]
  149. Monacelli, F.; Acquarone, E.; Giannotti, C.; Borghi, R.; Nencioni, A. Vitamin C, Aging and Alzheimer’s Disease. Nutrients 2017, 9, 670. [Google Scholar] [CrossRef] [Green Version]
  150. Hamid, M.; Mansoor, S.; Amber, S.; Zahid, S. A quantitative meta-analysis of vitamin C in the pathophysiology of Alzheimer’s disease. Front. Aging Neurosci. 2022, 14, 970263. [Google Scholar] [CrossRef]
  151. Consoli, D.C.; Brady, L.J.; Bowman, A.B.; Calipari, E.S.; Harrison, F.E. Ascorbate deficiency decreases dopamine release in gulo −/− and APP/PSEN1 mice. J. Neurochem. 2021, 157, 656–665. [Google Scholar] [CrossRef]
  152. Stefaniak, O.; Dobrzyńska, M.; Drzymała-Czyż, S.; Przysławski, J. Diet in the Prevention of Alzheimer’s Disease: Current Knowledge and Future Research Requirements. Nutrients 2022, 14, 4564. [Google Scholar] [CrossRef]
  153. Nishiumi, S.; Miyamoto, S.; Kawabata, K.; Ohnishi, K.; Mukai, R.; Murakami, A.; Ashida, H.; Terao, J. Dietary flavonoids as cancer-preventive and therapeutic biofactors. Front. Biosci. 2011, S3, 1332. [Google Scholar] [CrossRef]
  154. Shahidi, F.; Ambigaipalan, P. Phenolics and polyphenolics in foods, beverages and spices: Antioxidant activity and health effects—A review. J. Funct. Foods 2015, 18, 820–897. [Google Scholar] [CrossRef]
  155. Cho, I.; Song, H.; Cho, J.H. Flavonoids mitigate neurodegeneration in aged Caenorhabditis elegans by mitochondrial uncoupling. Food Sci. Nutr. 2020, 8, 6633–6642. [Google Scholar] [CrossRef]
  156. Commenges, D.; Scotet, V.; Renaud, S.; Jacqmin-Gadda, H.; Barberger-Gateau, P.; Dartigues, J.-F. Intake of flavonoids and risk of dementia. Eur. J. Epidemiol. 2000, 16, 357–363. [Google Scholar] [CrossRef]
  157. Bondonno, C.P.; Bondonno, N.P.; Dalgaard, F.; Murray, K.; Gardener, S.L.; Martins, R.N.; Rainey-Smith, S.R.; Cassidy, A.; Lewis, J.R.; Croft, K.D.; et al. Flavonoid intake and incident dementia in the Danish Diet, Cancer, and Health cohort. Alzheimers Dement. Transl. Res. Clin. Interv. 2021, 7, e12175. [Google Scholar] [CrossRef]
  158. Kuriyama, S.; Hozawa, A.; Ohmori, K.; Shimazu, T.; Matsui, T.; Ebihara, S.; Awata, S.; Nagatomi, R.; Arai, H.; Tsuji, I. Green tea consumption and cognitive function: A cross-sectional study from the Tsurugaya Project. Am. J. Clin. Nutr. 2006, 83, 355–361. [Google Scholar] [CrossRef] [Green Version]
  159. Nurk, E.; Refsum, H.; Drevon, C.A.; Tell, G.S.; Nygaard, H.A.; Engedal, K.; Smith, A.D. Intake of Flavonoid-Rich Wine, Tea, and Chocolate by Elderly Men and Women Is Associated with Better Cognitive Test Performance. J. Nutr. 2009, 139, 120–127. [Google Scholar] [CrossRef] [Green Version]
  160. 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] [Green Version]
  161. Naoi, M.; Wu, Y.; Shamoto-Nagai, M.; Maruyama, W. Mitochondria in Neuroprotection by Phytochemicals: Bioactive Polyphenols Modulate Mitochondrial Apoptosis System, Function and Structure. Int. J. Mol. Sci. 2019, 20, 2451. [Google Scholar] [CrossRef] [Green Version]
  162. Dragicevic, N.; Smith, A.; Lin, X.; Yuan, F.; Copes, N.; Delic, V.; Tan, J.; Cao, C.; Shytle, R.D.; Bradshaw, P.C. Green Tea Epigallocatechin-3-Gallate (EGCG) and Other Flavonoids Reduce Alzheimer’s Amyloid-Induced Mitochondrial Dysfunction. J. Alzheimers Dis. 2011, 26, 507–521. [Google Scholar] [CrossRef] [PubMed]
  163. Kozłowska, A.; Szostak-Wegierek, D. Flavonoids--food sources and health benefits. Rocz. Panstw. Zakl. Hig. 2014, 65, 79–85. [Google Scholar] [PubMed]
  164. Bernatoniene, J.; Kopustinskiene, D. The Role of Catechins in Cellular Responses to Oxidative Stress. Molecules 2018, 23, 965. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Bae, J.; Kim, N.; Shin, Y.; Kim, S.-Y.; Kim, Y.-J. Activity of catechins and their applications. Biomed. Dermatol. 2020, 4, 8. [Google Scholar] [CrossRef] [Green Version]
  166. Kocahan, S.; Doğan, Z. Mechanisms of Alzheimer’s Disease Pathogenesis and Prevention: The Brain, Neural Pathology, N-methyl-D-aspartate Receptors, Tau Protein and Other Risk Factors. Clin. Psychopharmacol. Neurosci. 2017, 15, 1–8. [Google Scholar] [CrossRef] [PubMed]
  167. Lou, H.; Fan, P.; Perez, R.G.; Lou, H. Neuroprotective effects of linarin through activation of the PI3K/Akt pathway in amyloid-β-induced neuronal cell death. Bioorg. Med. Chem. 2011, 19, 4021–4027. [Google Scholar] [CrossRef] [PubMed]
  168. Chen, X.; Xu, B.; Nie, L.; He, K.; Zhou, L.; Huang, X.; Spencer, P.; Yang, X.; Liu, J. Flavanol-rich lychee fruit extract substantially reduces progressive cognitive and molecular deficits in a triple-transgenic animal model of Alzheimer disease. Nutr. Neurosci. 2021, 24, 720–734. [Google Scholar] [CrossRef]
  169. Sheng, C.; Xu, P.; Zhou, K.; Deng, D.; Zhang, C.; Wang, Z. Icariin Attenuates Synaptic and Cognitive Deficits in an A β 1–42-Induced Rat Model of Alzheimer’s Disease. BioMed Res. Int. 2017, 2017, 7464872. [Google Scholar] [CrossRef] [Green Version]
  170. Wang, J.; Varghese, M.; Ono, K.; Yamada, M.; Levine, S.; Tzavaras, N.; Gong, B.; Hurst, W.J.; Blitzer, R.D.; Pasinetti, G.M. Cocoa Extracts Reduce Oligomerization of Amyloid-β: Implications for Cognitive Improvement in Alzheimer’s Disease. J. Alzheimers Dis. 2014, 41, 643–650. [Google Scholar] [CrossRef]
  171. Spencer, J.P.E. The interactions of flavonoids within neuronal signalling pathways. Genes Nutr. 2007, 2, 257–273. [Google Scholar] [CrossRef] [Green Version]
  172. 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]
  173. Bieschke, J.; Russ, J.; Friedrich, R.P.; Ehrnhoefer, D.E.; Wobst, H.; Neugebauer, K.; Wanker, E.E. EGCG remodels mature α-synuclein and amyloid-β fibrils and reduces cellular toxicity. Proc. Natl. Acad. Sci. USA 2010, 107, 7710–7715. [Google Scholar] [CrossRef]
  174. Walker, J.M.; Klakotskaia, D.; Ajit, D.; Weisman, G.A.; Wood, W.G.; Sun, G.Y.; Serfozo, P.; Simonyi, A.; Schachtman, T.R. Beneficial Effects of Dietary EGCG and Voluntary Exercise on Behavior in an Alzheimer’s Disease Mouse Model. J. Alzheimers Dis. 2015, 44, 561–572. [Google Scholar] [CrossRef]
  175. Ettcheto, M.; Cano, A.; Manzine, P.R.; Busquets, O.; Verdaguer, E.; Castro-Torres, R.D.; García, M.L.; Beas-Zarate, C.; Olloquequi, J.; Auladell, C.; et al. Epigallocatechin-3-Gallate (EGCG) Improves Cognitive Deficits Aggravated by an Obesogenic Diet Through Modulation of Unfolded Protein Response in APPswe/PS1dE9 Mice. Mol. Neurobiol. 2020, 57, 1814–1827. [Google Scholar] [CrossRef]
  176. Nan, S.; Wang, P.; Zhang, Y.; Fan, J. Epigallocatechin-3-Gallate Provides Protection Against Alzheimer’s Disease-Induced Learning and Memory Impairments in Rats. Drug Des. Dev. Ther. 2021, 15, 2013–2024. [Google Scholar] [CrossRef]
  177. Rezai-Zadeh, K.; Arendash, G.W.; Hou, H.; Fernandez, F.; Jensen, M.; Runfeldt, M.; Shytle, R.D.; Tan, J. Green tea epigallocatechin-3-gallate (EGCG) reduces β-amyloid mediated cognitive impairment and modulates tau pathology in Alzheimer transgenic mice. Brain Res. 2008, 1214, 177–187. [Google Scholar] [CrossRef]
  178. 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] [PubMed]
  179. Du, K.; Liu, M.; Zhong, X.; Yao, W.; Xiao, Q.; Wen, Q.; Yang, B.; Wei, M. Epigallocatechin Gallate Reduces Amyloid β-Induced Neurotoxicity via Inhibiting Endoplasmic Reticulum Stress-Mediated Apoptosis. Mol. Nutr. Food Res. 2018, 62, 1700890. [Google Scholar] [CrossRef]
  180. Zhong, X.; Liu, M.; Yao, W.; Du, K.; He, M.; Jin, X.; Jiao, L.; Ma, G.; Wei, B.; Wei, M. Epigallocatechin-3-Gallate Attenuates Microglial Inflammation and Neurotoxicity by Suppressing the Activation of Canonical and Noncanonical Inflammasome via TLR4/NF-κB Pathway. Mol. Nutr. Food Res. 2019, 63, 1801230. [Google Scholar] [CrossRef]
  181. Liu, M.; Chen, F.; Sha, L.; Wang, S.; Tao, L.; Yao, L.; He, M.; Yao, Z.; Liu, H.; Zhu, Z.; et al. (−)-Epigallocatechin-3-Gallate Ameliorates Learning and Memory Deficits by Adjusting the Balance of TrkA/p75NTR Signaling in APP/PS1 Transgenic Mice. Mol. Neurobiol. 2014, 49, 1350–1363. [Google Scholar] [CrossRef] [Green Version]
  182. Jiang, Z.; Dong, X.; Yan, X.; Liu, Y.; Zhang, L.; Sun, Y. Nanogels of dual inhibitor-modified hyaluronic acid function as a potent inhibitor of amyloid β-protein aggregation and cytotoxicity. Sci. Rep. 2018, 8, 3505. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Mori, T.; Koyama, N.; Tan, J.; Segawa, T.; Maeda, M.; Town, T. Combination therapy with octyl gallate and ferulic acid improves cognition and neurodegeneration in a transgenic mouse model of Alzheimer’s disease. J. Biol. Chem. 2017, 292, 11310–11325. [Google Scholar] [CrossRef] [PubMed]
  184. Forcano, L.; Fauria, K.; Soldevila-Domenech, N.; Minguillón, C.; Lorenzo, T.; Cuenca-Royo, A.; Menezes-Cabral, S.; Pizarro, N.; Boronat, A.; Molinuevo, J.L.; et al. Prevention of cognitive decline in subjective cognitive decline APOE ε4 carriers after EGCG and a multimodal intervention (PENSA): Study design. Alzheimers Dement. Transl. Res. Clin. Interv. 2021, 7, e12155. [Google Scholar] [CrossRef] [PubMed]
  185. Thenmozhi, A.J.; Raja, T.R.W.; Janakiraman, U.; Manivasagam, T. Neuroprotective Effect of Hesperidin on Aluminium Chloride Induced Alzheimer’s Disease in Wistar Rats. Neurochem. Res. 2015, 40, 767–776. [Google Scholar] [CrossRef] [PubMed]
  186. Evans, J.A.; Mendonca, P.; Soliman, K.F.A. Neuroprotective Effects and Therapeutic Potential of the Citrus Flavonoid Hesperetin in Neurodegenerative Diseases. Nutrients 2022, 14, 2228. [Google Scholar] [CrossRef]
  187. Hajialyani, M.; Hosein Farzaei, M.; Echeverría, J.; Nabavi, S.; Uriarte, E.; Sobarzo-Sánchez, E. Hesperidin as a Neuroprotective Agent: A Review of Animal and Clinical Evidence. Molecules 2019, 24, 648. [Google Scholar] [CrossRef] [Green Version]
  188. Ikram, M.; Muhammad, T.; Rehman, S.U.; Khan, A.; Jo, M.G.; Ali, T.; Kim, M.O. Hesperetin Confers Neuroprotection by Regulating Nrf2/TLR4/NF-κB Signaling in an Aβ Mouse Model. Mol. Neurobiol. 2019, 56, 6293–6309. [Google Scholar] [CrossRef]
  189. Kuşi, M.; Becer, E.; Vatansever, H.S.; Yücecan, S. Neuroprotective Effects of Hesperidin and Naringin in SK-N-AS Cell as an in vitro Model for Alzheimer’s Disease. J. Am. Nutr. Assoc. 2022, 1–9. [Google Scholar] [CrossRef]
  190. Khan, A.; Ikram, M.; Hahm, J.R.; Kim, M.O. Antioxidant and Anti-Inflammatory Effects of Citrus Flavonoid Hesperetin: Special Focus on Neurological Disorders. Antioxidants 2020, 9, 609. [Google Scholar] [CrossRef]
  191. Li, C.; Zug, C.; Qu, H.; Schluesener, H.; Zhang, Z. Hesperidin ameliorates behavioral impairments and neuropathology of transgenic APP/PS1 mice. Behav. Brain Res. 2015, 281, 32–42. [Google Scholar] [CrossRef]
  192. Luo, Y.; Fan, H.; Tan, X.; Li, Z. Hesperetin rescues emotional memory and synaptic plasticity deficit in aged rats. Behav. Neurosci. 2021, 135, 721–731. [Google Scholar] [CrossRef] [PubMed]
  193. Justin Thenmozhi, A.; William Raja, T.R.; Manivasagam, T.; Janakiraman, U.; Essa, M.M. Hesperidin ameliorates cognitive dysfunction, oxidative stress and apoptosis against aluminium chloride induced rat model of Alzheimer’s disease. Nutr. Neurosci. 2017, 20, 360–368. [Google Scholar] [CrossRef] [PubMed]
  194. Justin-Thenmozhi, A.; Dhivya Bharathi, M.; Kiruthika, R.; Manivasagam, T.; Borah, A.; Essa, M.M. Attenuation of Aluminum Chloride-Induced Neuroinflammation and Caspase Activation Through the AKT/GSK-3β Pathway by Hesperidin in Wistar Rats. Neurotox. Res. 2018, 34, 463–476. [Google Scholar] [CrossRef] [PubMed]
  195. Abou Baker, D.H.; Ibrahim, B.M.M.; Hassan, N.S.; Yousuf, A.F.; Gengaihi, S.E. Exploiting Citrus aurantium seeds and their secondary metabolites in the management of Alzheimer disease. Toxicol. Rep. 2020, 7, 723–729. [Google Scholar] [CrossRef] [PubMed]
  196. Mandour, D.A.; Bendary, M.A.; Alsemeh, A.E. Histological and imunohistochemical alterations of hippocampus and prefrontal cortex in a rat model of Alzheimer like-disease with a preferential role of the flavonoid “hesperidin” . J. Mol. Histol. 2021, 52, 1043–1065. [Google Scholar] [CrossRef]
  197. Wang, D.; Liu, L.; Zhu, X.; Wu, W.; Wang, Y. Hesperidin Alleviates Cognitive Impairment, Mitochondrial Dysfunction and Oxidative Stress in a Mouse Model of Alzheimer’s Disease. Cell. Mol. Neurobiol. 2014, 34, 1209–1221. [Google Scholar] [CrossRef]
  198. Hong, Y.; An, Z. Hesperidin attenuates learning and memory deficits in APP/PS1 mice through activation of Akt/Nrf2 signaling and inhibition of RAGE/NF-κB signaling. Arch. Pharm. Res. 2018, 41, 655–663. [Google Scholar] [CrossRef]
  199. Javed, H.; Vaibhav, K.; Ahmed, M.E.; Khan, A.; Tabassum, R.; Islam, F.; Safhi, M.M.; Islam, F. Effect of hesperidin on neurobehavioral, neuroinflammation, oxidative stress and lipid alteration in intracerebroventricular streptozotocin induced cognitive impairment in mice. J. Neurol. Sci. 2015, 348, 51–59. [Google Scholar] [CrossRef]
  200. Kean, R.J.; Lamport, D.J.; Dodd, G.F.; Freeman, J.E.; Williams, C.M.; Ellis, J.A.; Butler, L.T.; Spencer, J.P. Chronic consumption of flavanone-rich orange juice is associated with cognitive benefits: An 8-wk, randomized, double-blind, placebo-controlled trial in healthy older adults. Am. J. Clin. Nutr. 2015, 101, 506–514. [Google Scholar] [CrossRef] [Green Version]
  201. Zhang, S.; Tomata, Y.; Sugiyama, K.; Sugawara, Y.; Tsuji, I. Citrus consumption and incident dementia in elderly Japanese: The Ohsaki Cohort 2006 Study. Br. J. Nutr. 2017, 117, 1174–1180. [Google Scholar] [CrossRef] [Green Version]
  202. Ahsan, A.U.; Sharma, V.L.; Wani, A.; Chopra, M. Naringenin Upregulates AMPK-Mediated Autophagy to Rescue Neuronal Cells From β-Amyloid (1–42) Evoked Neurotoxicity. Mol. Neurobiol. 2020, 57, 3589–3602. [Google Scholar] [CrossRef]
  203. Wu, L.-H.; Lin, C.; Lin, H.-Y.; Liu, Y.-S.; Wu, C.Y.-J.; Tsai, C.-F.; Chang, P.-C.; Yeh, W.-L.; Lu, D.-Y. Naringenin Suppresses Neuroinflammatory Responses Through Inducing Suppressor of Cytokine Signaling 3 Expression. Mol. Neurobiol. 2016, 53, 1080–1091. [Google Scholar] [CrossRef]
  204. Yang, Z.; Kuboyama, T.; Tohda, C. Naringenin promotes microglial M2 polarization and degradation enzyme expression. Phytother. Res. 2019, 33, 1114–1121. [Google Scholar] [CrossRef]
  205. Md, S.; Gan, S.Y.; Haw, Y.H.; Ho, C.L.; Wong, S.; Choudhury, H. In vitro neuroprotective effects of naringenin nanoemulsion against β-amyloid toxicity through the regulation of amyloidogenesis and tau phosphorylation. Int. J. Biol. Macromol. 2018, 118, 1211–1219. [Google Scholar] [CrossRef]
  206. Khan, M.B.; Khan, M.M.; Khan, A.; Ahmed, M.E.; Ishrat, T.; Tabassum, R.; Vaibhav, K.; Ahmad, A.; Islam, F. Naringenin ameliorates Alzheimer’s disease (AD)-type neurodegeneration with cognitive impairment (AD-TNDCI) caused by the intracerebroventricular-streptozotocin in rat model. Neurochem. Int. 2012, 61, 1081–1093. [Google Scholar] [CrossRef]
  207. Ghofrani, S.; Joghataei, M.-T.; Mohseni, S.; Baluchnejadmojarad, T.; Bagheri, M.; Khamse, S.; Roghani, M. Naringenin improves learning and memory in an Alzheimer’s disease rat model: Insights into the underlying mechanisms. Eur. J. Pharmacol. 2015, 764, 195–201. [Google Scholar] [CrossRef] [Green Version]
  208. Haider, S.; Liaquat, L.; Ahmad, S.; Batool, Z.; Siddiqui, R.A.; Tabassum, S.; Shahzad, S.; Rafiq, S.; Naz, N. Naringenin protects AlCl3/D-galactose induced neurotoxicity in rat model of AD via attenuation of acetylcholinesterase levels and inhibition of oxidative stress. PLoS ONE 2020, 15, e0227631. [Google Scholar] [CrossRef] [Green Version]
  209. Zhou, T.; Liu, L.; Wang, Q.; Gao, Y. Naringenin alleviates cognition deficits in high-fat diet-fed SAMP8 mice. J. Food Biochem. 2020, 44, e13375. [Google Scholar] [CrossRef]
  210. Rebello, C.J.; Beyl, R.A.; Lertora, J.J.L.; Greenway, F.L.; Ravussin, E.; Ribnicky, D.M.; Poulev, A.; Kennedy, B.J.; Castro, H.F.; Campagna, S.R.; et al. Safety and pharmacokinetics of naringenin: A randomized, controlled, single-ascending-dose clinical trial. Diabetes Obes. Metab. 2020, 22, 91–98. [Google Scholar] [CrossRef]
  211. Nakajima, A.; Ohizumi, Y. Potential Benefits of Nobiletin, A Citrus Flavonoid, against Alzheimer’s Disease and Parkinson’s Disease. Int. J. Mol. Sci. 2019, 20, 3380. [Google Scholar] [CrossRef] [Green Version]
  212. Nakajima, A.; Aoyama, Y.; Nguyen, T.-T.L.; Shin, E.-J.; Kim, H.-C.; Yamada, S.; Nakai, T.; Nagai, T.; Yokosuka, A.; Mimaki, Y.; et al. Nobiletin, a citrus flavonoid, ameliorates cognitive impairment, oxidative burden, and hyperphosphorylation of tau in senescence-accelerated mouse. Behav. Brain Res. 2013, 250, 351–360. [Google Scholar] [CrossRef] [PubMed]
  213. Qi, G.; Mi, Y.; Fan, R.; Li, R.; Liu, Z.; Liu, X. Nobiletin Protects against Systemic Inflammation-Stimulated Memory Impairment via MAPK and NF-κB Signaling Pathways. J. Agric. Food Chem. 2019, 67, 5122–5134. [Google Scholar] [CrossRef] [PubMed]
  214. Youn, K.; Lee, S.; Jun, M. Discovery of Nobiletin from Citrus Peel as a Potent Inhibitor of β-Amyloid Peptide Toxicity. Nutrients 2019, 11, 2648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Chai, W.; Zhang, J.; Xiang, Z.; Zhang, H.; Mei, Z.; Nie, H.; Xu, R.; Zhang, P. Potential of nobiletin against Alzheimer’s disease through inhibiting neuroinflammation. Metab. Brain Dis. 2022, 37, 1145–1154. [Google Scholar] [CrossRef] [PubMed]
  216. Ghasemi-Tarie, R.; Kiasalari, Z.; Fakour, M.; Khorasani, M.; Keshtkar, S.; Baluchnejadmojarad, T.; Roghani, M. Nobiletin prevents amyloid β1-40-induced cognitive impairment via inhibition of neuroinflammation and oxidative/nitrosative stress. Metab. Brain Dis. 2022, 37, 1337–1349. [Google Scholar] [CrossRef]
  217. Wirianto, M.; Wang, C.; Kim, E.; Koike, N.; Gomez-Gutierrez, R.; Nohara, K.; Escobedo, G.; Choi, J.M.; Han, C.; Yagita, K.; et al. The clock modulator Nobiletin mitigates astrogliosis-associated neuroinflammation and disease hallmarks in an Alzheimer’s disease model. FASEB J. 2022, 36, e22186. [Google Scholar] [CrossRef]
  218. Matsuzaki, K.; Yamakuni, T.; Hashimoto, M.; Haque, A.M.; Shido, O.; Mimaki, Y.; Sashida, Y.; Ohizumi, Y. Nobiletin restoring β-amyloid-impaired CREB phosphorylation rescues memory deterioration in Alzheimer’s disease model rats. Neurosci. Lett. 2006, 400, 230–234. [Google Scholar] [CrossRef]
  219. Onozuka, H.; Nakajima, A.; Matsuzaki, K.; Shin, R.-W.; Ogino, K.; Saigusa, D.; Tetsu, N.; Yokosuka, A.; Sashida, Y.; Mimaki, Y.; et al. Nobiletin, a Citrus Flavonoid, Improves Memory Impairment and Aβ Pathology in a Transgenic Mouse Model of Alzheimer’s Disease. J. Pharmacol. Exp. Ther. 2008, 326, 739–744. [Google Scholar] [CrossRef]
  220. Yamada, S.; Shirai, M.; Ono, K.; Teruya, T.; Yamano, A.; Tae Woo, J. Beneficial effects of a nobiletin-rich formulated supplement of Sikwasa (C. depressa) peel on cognitive function in elderly Japanese subjects; A multicenter, randomized, double-blind, placebo-controlled study. Food Sci. Nutr. 2021, 9, 6844–6853. [Google Scholar] [CrossRef]
  221. Meguro, K.; Yamaguchi, S. Decreased Behavioral Abnormalities After Treatment with Combined Donepezil and Yokukansankachimpihange in Alzheimer Disease: An Observational Study. The Osaki-Tajiri Project. Neurol. Ther. 2018, 7, 333–340. [Google Scholar] [CrossRef] [Green Version]
  222. Seki, T.; Kamiya, T.; Furukawa, K.; Azumi, M.; Ishizuka, S.; Takayama, S.; Nagase, S.; Arai, H.; Yamakuni, T.; Yaegashi, N. Nobiletin-rich Citrus reticulata peels, a kampo medicine for Alzheimer’s disease: A case series: Letters to the Editor. Geriatr. Gerontol. Int. 2013, 13, 236–238. [Google Scholar] [CrossRef]
  223. Nabavi, S.F.; Khan, H.; D’onofrio, G.; Šamec, D.; Shirooie, S.; Dehpour, A.R.; Argüelles, S.; Habtemariam, S.; Sobarzo-Sanchez, E. Apigenin as neuroprotective agent: Of mice and men. Pharmacol. Res. 2018, 128, 359–365. [Google Scholar] [CrossRef]
  224. Salehi, B.; Venditti, A.; Sharifi-Rad, M.; Kręgiel, D.; Sharifi-Rad, J.; Durazzo, A.; Lucarini, M.; Santini, A.; Souto, E.; Novellino, E.; et al. The Therapeutic Potential of Apigenin. Int. J. Mol. Sci. 2019, 20, 1305. [Google Scholar] [CrossRef] [Green Version]
  225. Zhao, F.; Dang, Y.; Zhang, R.; Jing, G.; Liang, W.; Xie, L.; Li, Z. Apigenin attenuates acrylonitrile-induced neuro-inflammation in rats: Involved of inactivation of the TLR4/NF-κB signaling pathway. Int. Immunopharmacol. 2019, 75, 105697. [Google Scholar] [CrossRef]
  226. 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.d.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]
  227. Zhao, L.; Wang, J.-L.; Liu, R.; Li, X.-X.; Li, J.-F.; Zhang, L. Neuroprotective, Anti-Amyloidogenic and Neurotrophic Effects of Apigenin in an Alzheimer’s Disease Mouse Model. Molecules 2013, 18, 9949–9965. [Google Scholar] [CrossRef] [Green Version]
  228. Alsadat, A.M.; Nikbakht, F.; Hossein Nia, 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]
  229. Malar, D.S.; Prasanth, M.I.; Jeyakumar, M.; Balamurugan, K.; Devi, K.P. Vitexin prevents Aβ proteotoxicity in transgenic Caenorhabditis elegans model of Alzheimer’s disease by modulating unfolded protein response. J. Biochem. Mol. Toxicol. 2021, 35, e22632. [Google Scholar] [CrossRef]
  230. Balez, R.; Steiner, N.; Engel, M.; Muñoz, S.S.; Lum, J.S.; Wu, Y.; Wang, D.; Vallotton, P.; Sachdev, P.; O’Connor, M.; et al. Neuroprotective effects of apigenin against inflammation, neuronal excitability and apoptosis in an induced pluripotent stem cell model of Alzheimer’s disease. Sci. Rep. 2016, 6, 31450. [Google Scholar] [CrossRef] [Green Version]
  231. Panche, A.N.; Diwan, A.D.; Chandra, S.R. Flavonoids: An overview. J. Nutr. Sci. 2016, 5, e47. [Google Scholar] [CrossRef] [Green Version]
  232. Anand David, A.; Arulmoli, R.; Parasuraman, S. Overviews of biological importance of quercetin: A bioactive flavonoid. Pharmacogn. Rev. 2016, 10, 84. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Wang, D.-M.; Li, S.-Q.; Wu, W.-L.; Zhu, X.-Y.; Wang, Y.; Yuan, H.-Y. Effects of Long-Term Treatment with Quercetin on Cognition and Mitochondrial Function in a Mouse Model of Alzheimer’s Disease. Neurochem. Res. 2014, 39, 1533–1543. [Google Scholar] [CrossRef] [PubMed]
  234. Xu, M.; Huang, H.; Mo, X.; Zhu, Y.; Chen, X.; Li, X.; Peng, X.; Xu, Z.; Chen, L.; Rong, S.; et al. Quercetin-3-O-Glucuronide Alleviates Cognitive Deficit and Toxicity in Aβ 1-42-Induced AD-Like Mice and SH-SY5Y Cells. Mol. Nutr. Food Res. 2021, 65, 2000660. [Google Scholar] [CrossRef] [PubMed]
  235. Paula, P.-C.; Angelica Maria, S.-G.; Luis, C.-H.; Gloria Patricia, C.-G. Preventive Effect of Quercetin in a Triple Transgenic Alzheimer’s Disease Mice Model. Molecules 2019, 24, 2287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Zhang, X.-W.; Chen, J.-Y.; Ouyang, D.; Lu, J.-H. Quercetin in Animal Models of Alzheimer’s Disease: A Systematic Review of Preclinical Studies. Int. J. Mol. Sci. 2020, 21, 493. [Google Scholar] [CrossRef] [Green Version]
  237. Nakagawa, T.; Itoh, M.; Ohta, K.; Hayashi, Y.; Hayakawa, M.; Yamada, Y.; Akanabe, H.; Chikaishi, T.; Nakagawa, K.; Itoh, Y.; et al. Improvement of memory recall by quercetin in rodent contextual fear conditioning and human early-stage Alzheimer’s disease patients. NeuroReport 2016, 27, 671–676. [Google Scholar] [CrossRef]
  238. Holland, T.M.; Agarwal, P.; Wang, Y.; Leurgans, S.E.; Bennett, D.A.; Booth, S.L.; Morris, M.C. Dietary flavonols and risk of Alzheimer dementia. Neurology 2020, 94, e1749–e1756. [Google Scholar] [CrossRef]
  239. Hou, Y.; Aboukhatwa, M.A.; Lei, D.-L.; Manaye, K.; Khan, I.; Luo, Y. Anti-depressant natural flavonols modulate BDNF and beta amyloid in neurons and hippocampus of double TgAD mice. Neuropharmacology 2010, 58, 911–920. [Google Scholar] [CrossRef] [Green Version]
  240. Choi, R.C.Y.; Zhu, J.T.T.; Leung, K.W.; Chu, G.K.Y.; Xie, H.Q.; Chen, V.P.; Zheng, K.Y.Z.; Lau, D.T.W.; Dong, T.T.X.; Chow, P.C.Y.; et al. A Flavonol Glycoside, Isolated from Roots of Panax notoginseng, Reduces Amyloid-β-Induced Neurotoxicity in Cultured Neurons: Signaling Transduction and Drug Development for Alzheimer’s Disease. J. Alzheimers Dis. 2010, 19, 795–811. [Google Scholar] [CrossRef]
  241. Bhullar, K.S.; Rupasinghe, H.P.V. Partridgeberry polyphenols protect primary cortical and hippocampal neurons against β-amyloid toxicity. Food Res. Int. 2015, 74, 237–249. [Google Scholar] [CrossRef]
  242. Huebbe, P.; Wagner, A.E.; Boesch-Saadatmandi, C.; Sellmer, F.; Wolffram, S.; Rimbach, G. Effect of dietary quercetin on brain quercetin levels and the expression of antioxidant and Alzheimer’s disease relevant genes in mice. Pharmacol. Res. 2010, 61, 242–246. [Google Scholar] [CrossRef]
  243. Pinheiro, R.G.R.; Granja, A.; Loureiro, J.A.; Pereira, M.C.; Pinheiro, M.; Neves, A.R.; Reis, S. RVG29-Functionalized Lipid Nanoparticles for Quercetin Brain Delivery and Alzheimer’s Disease. Pharm. Res. 2020, 37, 139. [Google Scholar] [CrossRef]
  244. Pinheiro, R.G.R.; Granja, A.; Loureiro, J.A.; Pereira, M.C.; Pinheiro, M.; Neves, A.R.; Reis, S. Quercetin lipid nanoparticles functionalized with transferrin for Alzheimer’s disease. Eur. J. Pharm. Sci. 2020, 148, 105314. [Google Scholar] [CrossRef]
  245. Moreno, L.C.G.e.I.; Puerta, E.; Suárez-Santiago, J.E.; Santos-Magalhães, N.S.; Ramirez, M.J.; Irache, J.M. Effect of the oral administration of nanoencapsulated quercetin on a mouse model of Alzheimer’s disease. Int. J. Pharm. 2017, 517, 50–57. [Google Scholar] [CrossRef]
  246. Maher, P. Preventing and Treating Neurological Disorders with the Flavonol Fisetin. Brain Plast. 2021, 6, 155–166. [Google Scholar] [CrossRef]
  247. Ahmad, A.; Ali, T.; Park, H.Y.; Badshah, H.; Rehman, S.U.; Kim, M.O. Neuroprotective Effect of Fisetin Against Amyloid-Beta-Induced Cognitive/Synaptic Dysfunction, Neuroinflammation, and Neurodegeneration in Adult Mice. Mol. Neurobiol. 2017, 54, 2269–2285. [Google Scholar] [CrossRef]
  248. Kim, S.; Choi, K.J.; Cho, S.-J.; Yun, S.-M.; Jeon, J.-P.; Koh, Y.H.; Song, J.; Johnson, G.V.W.; Jo, C. Fisetin stimulates autophagic degradation of phosphorylated tau via the activation of TFEB and Nrf2 transcription factors. Sci. Rep. 2016, 6, 24933. [Google Scholar] [CrossRef] [Green Version]
  249. Currais, A.; Prior, M.; Dargusch, R.; Armando, A.; Ehren, J.; Schubert, D.; Quehenberger, O.; Maher, P. Modulation of p25 and inflammatory pathways by fisetin maintains cognitive function in A lzheimer’s disease transgenic mice. Aging Cell 2014, 13, 379–390. [Google Scholar] [CrossRef]
  250. Currais, A.; Farrokhi, C.; Dargusch, R.; Armando, A.; Quehenberger, O.; Schubert, D.; Maher, P. Fisetin Reduces the Impact of Aging on Behavior and Physiology in the Rapidly Aging SAMP8 Mouse. J. Gerontol. Ser. A 2018, 73, 299–307. [Google Scholar] [CrossRef] [Green Version]
  251. Jung, H.A.; Jin, S.E.; Park, J.-S.; Choi, J.S. Antidiabetic complications and anti-alzheimer activities of sophoflavescenol, a prenylated flavonol from Sophora flavescens, and its structure-activity relationship: Diabetic complications inhibitory activity of sophoflavescenol. Phytother. Res. 2011, 25, 709–715. [Google Scholar] [CrossRef]
  252. Yu, H.; Shi, J.; Lin, Y.; Zhang, Y.; Luo, Q.; Huang, S.; Wang, S.; Wei, J.; Huang, J.; Li, C.; et al. Icariin Ameliorates Alzheimer’s Disease Pathology by Alleviating Myelin Injury in 3 × Tg-AD Mice. Neurochem. Res. 2022, 47, 1049–1059. [Google Scholar] [CrossRef] [PubMed]
  253. Calderón-Garcidueñas, L.; Mora-Tiscareño, A.; Franco-Lira, M.; Cross, J.V.; Engle, R.; Aragón-Flores, M.; Gómez-Garza, G.; Jewells, V.; Medina-Cortina, H.; Solorio, E.; et al. Flavonol-rich dark cocoa significantly decreases plasma endothelin-1 and improves cognition in urban children. Front. Pharmacol. 2013, 4, 104. [Google Scholar] [CrossRef] [PubMed]
  254. Augustin, S.; Huebbe, P.; Matzner, N.; Augustin, K.; Schliebs, R.; Cermak, R.; Wolffram, S.; Rimbach, G. Ginkgo biloba Extract and its Flavonol and Terpenelactone Fractions do not Affect β-Secretase mRNA and Enzyme Activity Levels in Cultured Neurons and in Mice. Planta Med. 2008, 74, 6–13. [Google Scholar] [CrossRef] [PubMed]
  255. Augustin, S.; Rimbach, G.; Augustin, K.; Cermak, R.; Wolffram, S. Gene Regulatory Effects of Ginkgo biloba Extract and Its Flavonol and Terpenelactone Fractions in Mouse Brain. J. Clin. Biochem. Nutr. 2009, 45, 315–321. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Zhang, X.; Wang, J.; Hong, C.; Luo, W.; Wang, C. Design, synthesis and evaluation of genistein-polyamine conjugates as multi-functional anti-Alzheimer agents. Acta Pharm. Sin. B 2015, 5, 67–73. [Google Scholar] [CrossRef] [Green Version]
  257. Mas-Bargues, C.; Borrás, C.; Viña, J. The multimodal action of genistein in Alzheimer’s and other age-related diseases. Free Radic. Biol. Med. 2022, 183, 127–137. [Google Scholar] [CrossRef]
  258. Castillo, W.O.; Palomino, N.V.; Takahashi, C.S.; Giuliatti, S. Genistein and Galantamine Combinations Decrease β-Amyloid Peptide (1–42)–Induced Genotoxicity and Cell Death in SH-SY5Y Cell Line: An In Vitro and In Silico Approach for Mimic of Alzheimer’s Disease. Neurotox. Res. 2020, 38, 691–706. [Google Scholar] [CrossRef]
  259. Petry, F.d.S.; Coelho, B.P.; Gaelzer, M.M.; Kreutz, F.; Guma, F.T.C.R.; Salbego, C.G.; Trindade, V.M.T. Genistein protects against amyloid-beta-induced toxicity in SH-SY5Y cells by regulation of Akt and Tau phosphorylation. Phytother. Res. 2020, 34, 796–807. [Google Scholar] [CrossRef]
  260. Shentu, Y.-P.; Hu, W.-T.; Liang, J.-W.; Liuyang, Z.-Y.; Wei, H.; Qun, W.; Wang, X.-C.; Wang, J.-Z.; Westermarck, J.; Liu, R. Genistein Decreases APP/tau Phosphorylation and Ameliorates Aβ Overproduction Through Inhibiting CIP2A. Curr. Alzheimer Res. 2019, 16, 732–740. [Google Scholar] [CrossRef]
  261. Chatterjee, G.; Roy, D.; Kumar Khemka, V.; Chattopadhyay, M.; Chakrabarti, S. Genistein, the Isoflavone in Soybean, Causes Amyloid Beta Peptide Accumulation in Human Neuroblastoma Cell Line: Implications in Alzheimer’s Disease. Aging Dis. 2015, 6, 456. [Google Scholar] [CrossRef] [Green Version]
  262. Mirahmadi, S.-M.-S.; Shahmohammadi, A.; Rousta, A.-M.; Azadi, M.-R.; Fahanik-Babaei, J.; Baluchnejadmojarad, T.; Roghani, M. Soy isoflavone genistein attenuates lipopolysaccharide-induced cognitive impairments in the rat via exerting anti-oxidative and anti-inflammatory effects. Cytokine 2018, 104, 151–159. [Google Scholar] [CrossRef]
  263. Wang, Y.; Xia, Z.; Jiang, X.; Li, L.; Wang, H.; An, D.; Liu, Y. Genistein inhibits amyloid peptide 25-35-induced neuronal death by modulating estrogen receptors, choline acetyltransferase and glutamate receptors. Arch. Biochem. Biophys. 2020, 693, 108561. [Google Scholar] [CrossRef]
  264. Bagheri, M.; Roghani, M.; Joghataei, M.-T.; Mohseni, S. Genistein inhibits aggregation of exogenous amyloid-beta1–40 and alleviates astrogliosis in the hippocampus of rats. Brain Res. 2012, 1429, 145–154. [Google Scholar] [CrossRef] [Green Version]
  265. Pierzynowska, K.; Podlacha, M.; Gaffke, L.; Majkutewicz, I.; Mantej, J.; Węgrzyn, A.; Osiadły, M.; Myślińska, D.; Węgrzyn, G. Autophagy-dependent mechanism of genistein-mediated elimination of behavioral and biochemical defects in the rat model of sporadic Alzheimer’s disease. Neuropharmacology 2019, 148, 332–346. [Google Scholar] [CrossRef]
  266. Liu, J.Y.H.; Sun, M.Y.Y.; Sommerville, N.; Ngan, M.P.; Ponomarev, E.D.; Lin, G.; Rudd, J.A. Soy flavonoids prevent cognitive deficits induced by intra-gastrointestinal administration of beta-amyloid. Food Chem. Toxicol. 2020, 141, 111396. [Google Scholar] [CrossRef]
  267. Bonet-Costa, V.; Herranz-Pérez, V.; Blanco-Gandía, M.; Mas-Bargues, C.; Inglés, M.; Garcia-Tarraga, P.; Rodriguez-Arias, M.; Miñarro, J.; Borras, C.; Garcia-Verdugo, J.M.; et al. Clearing Amyloid-β through PPARγ/ApoE Activation by Genistein is a Treatment of Experimental Alzheimer’s Disease. J. Alzheimers Dis. 2016, 51, 701–711. [Google Scholar] [CrossRef]
  268. 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]
  269. 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]
  270. Gleason, C.E.; Fischer, B.L.; Dowling, N.M.; Setchell, K.D.R.; Atwood, C.S.; Carlsson, C.M.; Asthana, S. Cognitive Effects of Soy Isoflavones in Patients with Alzheimer’s Disease. J. Alzheimers Dis. 2015, 47, 1009–1019. [Google Scholar] [CrossRef] [Green Version]
  271. Guo, J.; Yang, G.; He, Y.; Xu, H.; Fan, H.; An, J.; Zhang, L.; Zhang, R.; Cao, G.; Hao, D.; et al. Involvement of α7nAChR in the Protective Effects of Genistein Against β-Amyloid-Induced Oxidative Stress in Neurons via a PI3K/Akt/Nrf2 Pathway-Related Mechanism. Cell. Mol. Neurobiol. 2021, 41, 377–393. [Google Scholar] [CrossRef]
  272. Zhou, X.; Yuan, L.; Zhao, X.; Hou, C.; Ma, W.; Yu, H.; Xiao, R. Genistein antagonizes inflammatory damage induced by β-amyloid peptide in microglia through TLR4 and NF-κB. Nutrition 2014, 30, 90–95. [Google Scholar] [CrossRef] [PubMed]
  273. Duro-Castano, A.; Borrás, C.; Herranz-Pérez, V.; Blanco-Gandía, M.C.; Conejos-Sánchez, I.; Armiñán, A.; Mas-Bargues, C.; Inglés, M.; Miñarro, J.; Rodríguez-Arias, M.; et al. Targeting Alzheimer’s disease with multimodal polypeptide-based nanoconjugates. Sci. Adv. 2021, 7, eabf9180. [Google Scholar] [CrossRef] [PubMed]
  274. Hong, C.; Guo, H.; Chen, S.; Lv, J.; Zhang, X.; Yang, Y.; Huang, K.; Zhang, Y.; Tian, Z.; Luo, W.; et al. Synthesis and biological evaluation of genistein-O-alkylamine derivatives as potential multifunctional anti-Alzheimer agents. Chem. Biol. Drug Des. 2019, 93, 188–200. [Google Scholar] [CrossRef]
  275. Sang, Z.; Shi, J.; Zhou, Y.; Wang, K.; Zhao, Y.; Li, Q.; Qiao, Z.; Wu, A.; Tan, Z.; Liu, W. Development of genistein-O-alkylamines derivatives as multifunctional agents for the treatment of Alzheimer’s disease. Bioorganic Chem. 2021, 107, 104602. [Google Scholar] [CrossRef] [PubMed]
  276. Afzal, M.; Redha, A.; AlHasan, R. Anthocyanins Potentially Contribute to Defense against Alzheimer’s Disease. Molecules 2019, 24, 4255. [Google Scholar] [CrossRef] [Green Version]
  277. Zhang, J.; Wu, J.; Liu, F.; Tong, L.; Chen, Z.; Chen, J.; He, H.; Xu, R.; Ma, Y.; Huang, C. Neuroprotective effects of anthocyanins and its major component cyanidin-3-O-glucoside (C3G) in the central nervous system: An outlined review. Eur. J. Pharmacol. 2019, 858, 172500. [Google Scholar] [CrossRef]
  278. Suresh, S.; Begum, R.F.; Singh, S.A.; Chitra, V. Anthocyanin as a therapeutic in Alzheimer’s disease: A systematic review of preclinical evidences. Ageing Res. Rev. 2022, 76, 101595. [Google Scholar] [CrossRef]
  279. 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]
  280. Rosli, H.; Shahar, S.; Rajab, N.F.; Che Din, N.; Haron, H. The effects of polyphenols-rich tropical fruit juice on cognitive function and metabolomics profile—A randomized controlled trial in middle-aged women. Nutr. Neurosci. 2022, 25, 1577–1593. [Google Scholar] [CrossRef]
  281. Li, J.; Zhao, R.; Jiang, Y.; Xu, Y.; Zhao, H.; Lyu, X.; Wu, T. Bilberry anthocyanins improve neuroinflammation and cognitive dysfunction in APP/PSEN1 mice via the CD33/TREM2/TYROBP signaling pathway in microglia. Food Funct. 2020, 11, 1572–1584. [Google Scholar] [CrossRef]
  282. Khan, M.S.; Ali, T.; Kim, M.W.; Jo, M.H.; Chung, J.I.; Kim, M.O. Anthocyanins Improve Hippocampus-Dependent Memory Function and Prevent Neurodegeneration via JNK/Akt/GSK3β Signaling in LPS-Treated Adult Mice. Mol. Neurobiol. 2019, 56, 671–687. [Google Scholar] [CrossRef]
  283. Wen, H.; Cui, H.; Tian, H.; Zhang, X.; Ma, L.; Ramassamy, C.; Li, J. Isolation of Neuroprotective Anthocyanins from Black Chokeberry (Aronia melanocarpa) against Amyloid-β-Induced Cognitive Impairment. Foods 2020, 10, 63. [Google Scholar] [CrossRef]
  284. Rehman, S.U.; Shah, S.A.; Ali, T.; Chung, J.I.; Kim, M.O. Anthocyanins Reversed D-Galactose-Induced Oxidative Stress and Neuroinflammation Mediated Cognitive Impairment in Adult Rats. Mol. Neurobiol. 2017, 54, 255–271. [Google Scholar] [CrossRef]
  285. Fan, Z.; Wen, H.; Zhang, X.; Li, J.; Zang, J. Cyanidin 3-O-β-Galactoside Alleviated Cognitive Impairment in Mice by Regulating Brain Energy Metabolism During Aging. J. Agric. Food Chem. 2022, 70, 1111–1121. [Google Scholar] [CrossRef]
  286. El-Shiekh, R.A.; Ashour, R.M.; Abd El-Haleim, E.A.; Ahmed, K.A.; Abdel-Sattar, E. Hibiscus sabdariffa L.: A potent natural neuroprotective agent for the prevention of streptozotocin-induced Alzheimer’s disease in mice. Biomed. Pharmacother. 2020, 128, 110303. [Google Scholar] [CrossRef]
  287. Gutierres, J.M.; Carvalho, F.B.; Schetinger, M.R.C.; Marisco, P.; Agostinho, P.; Rodrigues, M.; Rubin, M.A.; Schmatz, R.; da Silva, C.R.; Cognato, G.D.P.; et al. Anthocyanins restore behavioral and biochemical changes caused by streptozotocin-induced sporadic dementia of Alzheimer’s type. Life Sci. 2014, 96, 7–17. [Google Scholar] [CrossRef] [Green Version]
  288. Shimada, M.; Maeda, H.; Nanashima, N.; Yamada, K.; Nakajima, A. Anthocyanin-rich blackcurrant extract improves long-term memory impairment and emotional abnormality in senescence-accelerated mice. J. Food Biochem. 2022, 46, e14295. [Google Scholar] [CrossRef]
  289. Ma, H.; Johnson, S.; Liu, W.; DaSilva, N.; Meschwitz, S.; Dain, J.; Seeram, N. Evaluation of Polyphenol Anthocyanin-Enriched Extracts of Blackberry, Black Raspberry, Blueberry, Cranberry, Red Raspberry, and Strawberry for Free Radical Scavenging, Reactive Carbonyl Species Trapping, Anti-Glycation, Anti-β-Amyloid Aggregation, and Microglial Neuroprotective Effects. Int. J. Mol. Sci. 2018, 19, 461. [Google Scholar] [CrossRef] [Green Version]
  290. Navarro-Hortal, M.D.; Romero-Márquez, J.M.; Esteban-Muñoz, A.; Sánchez-González, C.; Rivas-García, L.; Llopis, J.; Cianciosi, D.; Giampieri, F.; Sumalla-Cano, S.; Battino, M.; et al. Strawberry (Fragaria × ananassa cv. Romina) methanolic extract attenuates Alzheimer’s beta amyloid production and oxidative stress by SKN-1/NRF and DAF-16/FOXO mediated mechanisms in C. elegans. Food Chem. 2022, 372, 131272. [Google Scholar] [CrossRef]
  291. Sun, C.; Zhang, S.; Ba, S.; Dang, J.; Ren, Q.; Zhu, Y.; Liu, K.; Jin, M. Eucommia ulmoides Olive Male Flower Extracts Ameliorate Alzheimer’s Disease-Like Pathology in Zebrafish via Regulating Autophagy, Acetylcholinesterase, and the Dopamine Transporter. Front. Mol. Neurosci. 2022, 15, 901953. [Google Scholar] [CrossRef]
  292. Badshah, H.; Kim, T.H.; Kim, M.O. Protective effects of Anthocyanins against Amyloid beta-induced neurotoxicity in vivo and in vitro. Neurochem. Int. 2015, 80, 51–59. [Google Scholar] [CrossRef] [PubMed]
  293. Kim, M.J.; Rehman, S.U.; Amin, F.U.; Kim, M.O. Enhanced neuroprotection of anthocyanin-loaded PEG-gold nanoparticles against Aβ1-42-induced neuroinflammation and neurodegeneration via the NF-KB/JNK/GSK3β signaling pathway. Nanomed. Nanotechnol. Biol. Med. 2017, 13, 2533–2544. [Google Scholar] [CrossRef] [PubMed]
  294. Ali, T.; Kim, T.; Rehman, S.U.; Khan, M.S.; Amin, F.U.; Khan, M.; Ikram, M.; Kim, M.O. Natural Dietary Supplementation of Anthocyanins via PI3K/Akt/Nrf2/HO-1 Pathways Mitigate Oxidative Stress, Neurodegeneration, and Memory Impairment in a Mouse Model of Alzheimer’s Disease. Mol. Neurobiol. 2018, 55, 6076–6093. [Google Scholar] [CrossRef] [PubMed]
  295. Pacheco, S.M.; Soares, M.S.P.; Gutierres, J.M.; Gerzson, M.F.B.; Carvalho, F.B.; Azambuja, J.H.; Schetinger, M.R.C.; Stefanello, F.M.; Spanevello, R.M. Anthocyanins as a potential pharmacological agent to manage memory deficit, oxidative stress and alterations in ion pump activity induced by experimental sporadic dementia of Alzheimer’s type. J. Nutr. Biochem. 2018, 56, 193–204. [Google Scholar] [CrossRef] [PubMed]
  296. Ochiishi, T.; Kaku, M.; Kajsongkram, T.; Thisayakorn, K. Mulberry fruit extract alleviates the intracellular amyloid-β oligomer-induced cognitive disturbance and oxidative stress in Alzheimer’s disease model mice. Genes Cells 2021, 26, 861–873. [Google Scholar] [CrossRef]
  297. Li, H.; Zheng, T.; Lian, F.; Xu, T.; Yin, W.; Jiang, Y. Anthocyanin-rich blueberry extracts and anthocyanin metabolite protocatechuic acid promote autophagy-lysosomal pathway and alleviate neurons damage in in vivo and in vitro models of Alzheimer’s disease. Nutrition 2022, 93, 111473. [Google Scholar] [CrossRef]
  298. Ali, T.; Kim, M.J.; Rehman, S.U.; Ahmad, A.; Kim, M.O. Anthocyanin-Loaded PEG-Gold Nanoparticles Enhanced the Neuroprotection of Anthocyanins in an Aβ1–42 Mouse Model of Alzheimer’s Disease. Mol. Neurobiol. 2017, 54, 6490–6506. [Google Scholar] [CrossRef]
  299. Singla, R.K.; Dubey, A.K.; Garg, A.; Sharma, R.K.; Fiorino, M.; Ameen, S.M.; Haddad, M.A.; Al-Hiary, M. Natural Polyphenols: Chemical Classification, Definition of Classes, Subcategories, and Structures. J. AOAC Int. 2019, 102, 1397–1400. [Google Scholar] [CrossRef]
  300. Varesi, A.; Chirumbolo, S.; Campagnoli, L.I.M.; Pierella, E.; Piccini, G.B.; Carrara, A.; Ricevuti, G.; Scassellati, C.; Bonvicini, C.; Pascale, A. The Role of Antioxidants in the Interplay between Oxidative Stress and Senescence. Antioxidants 2022, 11, 1224. [Google Scholar] [CrossRef]
  301. Marino, A.; Battaglini, M.; Moles, N.; Ciofani, G. Natural Antioxidant Compounds as Potential Pharmaceutical Tools against Neurodegenerative Diseases. ACS Omega 2022, 7, 25974–25990. [Google Scholar] [CrossRef]
  302. Mori, T.; Koyama, N.; Guillot-Sestier, M.-V.; Tan, J.; Town, T. Ferulic Acid Is a Nutraceutical β-Secretase Modulator That Improves Behavioral Impairment and Alzheimer-like Pathology in Transgenic Mice. PLoS ONE 2013, 8, e55774. [Google Scholar] [CrossRef] [Green Version]
  303. Pi, R.; Mao, X.; Chao, X.; Cheng, Z.; Liu, M.; Duan, X.; Ye, M.; Chen, X.; Mei, Z.; Liu, P.; et al. Tacrine-6-Ferulic Acid, a Novel Multifunctional Dimer, Inhibits Amyloid-β-Mediated Alzheimer’s Disease-Associated Pathogenesis In Vitro and In Vivo. PLoS ONE 2012, 7, e31921. [Google Scholar] [CrossRef]
  304. Sul, D.; Kim, H.-S.; Lee, D.; Joo, S.S.; Hwang, K.W.; Park, S.-Y. Protective effect of caffeic acid against beta-amyloid-induced neurotoxicity by the inhibition of calcium influx and tau phosphorylation. Life Sci. 2009, 84, 257–262. [Google Scholar] [CrossRef]
  305. Kim, J.H.; Wang, Q.; Choi, J.M.; Lee, S.; Cho, E.J. Protective role of caffeic acid in an Aβ 25-35-induced Alzheimer’s disease model. Nutr. Res. Pract. 2015, 9, 480. [Google Scholar] [CrossRef] [Green Version]
  306. Morroni, F.; Sita, G.; Graziosi, A.; Turrini, E.; Fimognari, C.; Tarozzi, A.; Hrelia, P. Neuroprotective Effect of Caffeic Acid Phenethyl Ester in A Mouse Model of Alzheimer’s Disease Involves Nrf2/HO-1 Pathway. Aging Dis. 2018, 9, 605. [Google Scholar] [CrossRef] [Green Version]
  307. Yoon, J.-H.; Youn, K.; Ho, C.-T.; Karwe, M.V.; Jeong, W.-S.; Jun, M. p-Coumaric Acid and Ursolic Acid from Corni fructus Attenuated β-Amyloid 25–35-Induced Toxicity through Regulation of the NF-κB Signaling Pathway in PC12 Cells. J. Agric. Food Chem. 2014, 62, 4911–4916. [Google Scholar] [CrossRef]
  308. Daroi, P.A.; Dhage, S.N.; Juvekar, A.R. p-Coumaric acid protects against D-galactose induced neurotoxicity by attenuating neuroinflammation and apoptosis in mice brain. Metab. Brain Dis. 2022, 37, 2569–2579. [Google Scholar] [CrossRef]
  309. Shin, K.Y.; Lee, G.H.; Park, C.H.; Kim, H.J.; Park, S.-H.; Kim, S.; Kim, H.-S.; Lee, K.-S.; Won, B.Y.; Lee, H.G.; et al. A novel compound, maltolylp-coumarate, attenuates cognitive deficits and shows neuroprotective effects in vitro and in vivo dementia models. J. Neurosci. Res. 2007, 85, 2500–2511. [Google Scholar] [CrossRef]
  310. Ogunlade, B.; Adelakun, S.A.; Agie, J.A. Nutritional supplementation of gallic acid ameliorates Alzheimer-type hippocampal neurodegeneration and cognitive impairment induced by aluminum chloride exposure in adult Wistar rats. Drug Chem. Toxicol. 2022, 45, 651–662. [Google Scholar] [CrossRef]
  311. Yu, M.; Chen, X.; Liu, J.; Ma, Q.; Zhuo, Z.; Chen, H.; Zhou, L.; Yang, S.; Zheng, L.; Ning, C.; et al. Gallic acid disruption of Aβ1–42 aggregation rescues cognitive decline of APP/PS1 double transgenic mouse. Neurobiol. Dis. 2019, 124, 67–80. [Google Scholar] [CrossRef]
  312. Zhong, L.; Liu, H.; Zhang, W.; Liu, X.; Jiang, B.; Fei, H.; Sun, Z. Ellagic acid ameliorates learning and memory impairment in APP/PS1 transgenic mice via inhibition of β-amyloid production and tau hyperphosphorylation. Exp. Ther. Med. 2018, 16, 4951–4958. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  313. Kiasalari, Z.; Heydarifard, R.; Khalili, M.; Afshin-Majd, S.; Baluchnejadmojarad, T.; Zahedi, E.; Sanaierad, A.; Roghani, M. Ellagic acid ameliorates learning and memory deficits in a rat model of Alzheimer’s disease: An exploration of underlying mechanisms. Psychopharmacology 2017, 234, 1841–1852. [Google Scholar] [CrossRef] [PubMed]
  314. Ramadan, W.S.; Alkarim, S. Ellagic Acid Modulates the Amyloid Precursor Protein Gene via Superoxide Dismutase Regulation in the Entorhinal Cortex in an Experimental Alzheimer’s Model. Cells 2021, 10, 3511. [Google Scholar] [CrossRef] [PubMed]
  315. Porquet, D.; Casadesús, G.; Bayod, S.; Vicente, A.; Canudas, A.M.; Vilaplana, J.; Pelegrí, C.; Sanfeliu, C.; Camins, A.; Pallàs, M.; et al. Dietary resveratrol prevents Alzheimer’s markers and increases life span in SAMP8. AGE 2013, 35, 1851–1865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  316. Gu, J.; Li, Z.; Chen, H.; Xu, X.; Li, Y.; Gui, Y. Neuroprotective Effect of Trans-Resveratrol in Mild to Moderate Alzheimer Disease: A Randomized, Double-Blind Trial. Neurol. Ther. 2021, 10, 905–917. [Google Scholar] [CrossRef]
  317. Kong, D.; Yan, Y.; He, X.-Y.; Yang, H.; Liang, B.; Wang, J.; He, Y.; Ding, Y.; Yu, H. Effects of Resveratrol on the Mechanisms of Antioxidants and Estrogen in Alzheimer’s Disease. BioMed Res. Int. 2019, 2019, 8983752. [Google Scholar] [CrossRef] [Green Version]
  318. Lim, G.P.; Chu, T.; Yang, F.; Beech, W.; Frautschy, S.A.; Cole, G.M. The Curry Spice Curcumin Reduces Oxidative Damage and Amyloid Pathology in an Alzheimer Transgenic Mouse. J. Neurosci. 2001, 21, 8370–8377. [Google Scholar] [CrossRef] [Green Version]
  319. Sadegh Malvajerd, S.; Izadi, Z.; Azadi, A.; Kurd, M.; Derakhshankhah, H.; Sharifzadeh, M.; Akbari Javar, H.; Hamidi, M. Neuroprotective Potential of Curcumin-Loaded Nanostructured Lipid Carrier in an Animal Model of Alzheimer’s Disease: Behavioral and Biochemical Evidence. J. Alzheimers Dis. 2019, 69, 671–686. [Google Scholar] [CrossRef]
  320. Ringman, J.M.; Frautschy, S.A.; Teng, E.; Begum, A.N.; Bardens, J.; Beigi, M.; Gylys, K.H.; Badmaev, V.; Heath, D.D.; Apostolova, L.G.; et al. Oral curcumin for Alzheimer’s disease: Tolerability and efficacy in a 24-week randomized, double blind, placebo-controlled study. Alzheimers Res. Ther. 2012, 4, 43. [Google Scholar] [CrossRef] [Green Version]
  321. Mostafa, N.M.; Mostafa, A.M.; Ashour, M.L.; Elhady, S.S. Neuroprotective Effects of Black Pepper Cold-Pressed Oil on Scopolamine-Induced Oxidative Stress and Memory Impairment in Rats. Antioxidants 2021, 10, 1993. [Google Scholar] [CrossRef]
  322. Mao, X.; Liao, Z.; Guo, L.; Xu, X.; Wu, B.; Xu, M.; Zhao, X.; Bi, K.; Jia, Y. Schisandrin C Ameliorates Learning and Memory Deficits by Aβ 1-42-induced Oxidative Stress and Neurotoxicity in Mice: AD.; SCH-C.; AΒ1-42; Neuroprotective; Intracerebroventricular Injection. Phytother. Res. 2015, 29, 1373–1380. [Google Scholar] [CrossRef]
  323. Jeong, E.J.; Lee, H.K.; Lee, K.Y.; Jeon, B.J.; Kim, D.H.; Park, J.-H.; Song, J.-H.; Huh, J.; Lee, J.-H.; Sung, S.H. The effects of lignan-riched extract of Shisandra chinensis on amyloid-β-induced cognitive impairment and neurotoxicity in the cortex and hippocampus of mouse. J. Ethnopharmacol. 2013, 146, 347–354. [Google Scholar] [CrossRef]
  324. Kiokias, S.; Proestos, C.; Oreopoulou, V. Phenolic Acids of Plant Origin—A Review on Their Antioxidant Activity In Vitro (O/W Emulsion Systems) Along with Their in Vivo Health Biochemical Properties. Foods 2020, 9, 534. [Google Scholar] [CrossRef] [Green Version]
  325. Kiokias, S.; Oreopoulou, V. A Review of the Health Protective Effects of Phenolic Acids against a Range of Severe Pathologic Conditions (Including Coronavirus-Based Infections). Molecules 2021, 26, 5405. [Google Scholar] [CrossRef]
  326. Chandrasekara, A. Phenolic Acids. In Encyclopedia of Food Chemistry; Elsevier: Amsterdam, The Netherlands, 2019; pp. 535–545. ISBN 978-0-12-814045-1. [Google Scholar]
  327. Szwajgier, D.; Borowiec, K.; Pustelniak, K. The Neuroprotective Effects of Phenolic Acids: Molecular Mechanism of Action. Nutrients 2017, 9, 477. [Google Scholar] [CrossRef] [Green Version]
  328. Stevenson, D.E.; Hurst, R.D. Polyphenolic phytochemicals—Just antioxidants or much more? Cell. Mol. Life Sci. 2007, 64, 2900–2916. [Google Scholar] [CrossRef]
  329. Phadke, A.V.; Tayade, A.A.; Khambete, M.P. Therapeutic potential of ferulic acid and its derivatives in Alzheimer’s disease—A systematic review. Chem. Biol. Drug Des. 2021, 98, 713–721. [Google Scholar] [CrossRef]
  330. Zduńska, K.; Dana, A.; Kolodziejczak, A.; Rotsztejn, H. Antioxidant Properties of Ferulic Acid and Its Possible Application. Skin Pharmacol. Physiol. 2018, 31, 332–336. [Google Scholar] [CrossRef]
  331. Liu, Y.; Shi, L.; Qiu, W.; Shi, Y. Ferulic acid exhibits anti-inflammatory effects by inducing autophagy and blocking NLRP3 inflammasome activation. Mol. Cell. Toxicol. 2022, 18, 509–519. [Google Scholar] [CrossRef]
  332. Mishra, T.; Nagarajan, K.; Dixit, P.K.; Kumar, V. Neuroprotective potential of ferulic acid against cyclophosphamide-induced neuroinflammation and behavioral changes. J. Food Biochem. 2022, 46, e14436. [Google Scholar] [CrossRef]
  333. Ren, Z.; Zhang, R.; Li, Y.; Li, Y.; Yang, Z.; Yang, H. Ferulic acid exerts neuroprotective effects against cerebral ischemia/reperfusion-induced injury via antioxidant and anti-apoptotic mechanisms in vitro and in vivo. Int. J. Mol. Med. 2017, 40, 1444–1456. [Google Scholar] [CrossRef] [Green Version]
  334. Long, T.; Wu, Q.; Wei, J.; Tang, Y.; He, Y.-N.; He, C.-L.; Chen, X.; Yu, L.; Yu, C.-L.; Law, B.Y.-K.; et al. Ferulic Acid Exerts Neuroprotective Effects via Autophagy Induction in C. elegans and Cellular Models of Parkinson’s Disease. Oxid. Med. Cell. Longev. 2022, 2022, 3723567. [Google Scholar] [CrossRef] [PubMed]
  335. Haque, E.; Javed, H.; Azimullah, S.; Abul Khair, S.B.; Ojha, S. Neuroprotective potential of ferulic acid in the rotenone model of Parkinson’s disease. Drug Des. Dev. Ther. 2015, 2015, 5499–5518. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  336. Ghobadi, M.; Arji Roodsari, B.; Yadegari, M.; Homayouni-Moghadam, F.; Rezvani, M.E. The Effect of Ferulic Acid on Cognitive Disorder and Brain Structural Deficit in Mice in Cuprizone Induced Animal Model of Multiple Sclerosis. J. Rafsanjan Univ. Med. Sci. 2017, 16, 541–552. [Google Scholar]
  337. Wang, E.-J.; Wu, M.-Y.; Lu, J.-H. Ferulic Acid in Animal Models of Alzheimer’s Disease: A Systematic Review of Preclinical Studies. Cells 2021, 10, 2653. [Google Scholar] [CrossRef]
  338. Yan, J.-J.; Jung, J.-S.; Kim, T.-K.; Hasan, M.A.; Hong, C.-W.; Nam, J.-S.; Song, D.-K. Protective Effects of Ferulic Acid in Amyloid Precursor Protein Plus Presenilin-1 Transgenic Mouse Model of Alzheimer Disease. Biol. Pharm. Bull. 2013, 36, 140–143. [Google Scholar] [CrossRef] [Green Version]
  339. Ono, K.; Hirohata, M.; Yamada, M. Ferulic acid destabilizes preformed β-amyloid fibrils in vitro. Biochem. Biophys. Res. Commun. 2005, 336, 444–449. [Google Scholar] [CrossRef]
  340. Wang, Q. Effect of Ferulic Acid on Learning and Memory Impairment by the Repairing of Mitochondrial Fission-Fusion Imbalance in AD Mice. Chin. Pharm. J. 2019, 24, 703–710. [Google Scholar]
  341. Picone, P.; Bondi, M.L.; Picone, P.; Bondi, M.L.; Montana, G.; Bruno, A.; Pitarresi, G.; Giammona, G.; Di Carlo, M. Ferulic acid inhibits oxidative stress and cell death induced by Ab oligomers: Improved delivery by solid lipid nanoparticles. Free Radic. Res. 2009, 43, 1133–1145. [Google Scholar] [CrossRef]
  342. Zafeer, M.F.; Firdaus, F.; Anis, E.; Mobarak Hossain, M. Prolong treatment with Trans-ferulic acid mitigates bioenergetics loss and restores mitochondrial dynamics in streptozotocin-induced sporadic dementia of Alzheimer’s type. NeuroToxicology 2019, 73, 246–257. [Google Scholar] [CrossRef]
  343. Wang, Y.; Wang, C.; Wang, X.; Yu, S. Effects of Ferulic Acid on Oxidative Stress and Apoptosis Related Proteins in Alzheimer’s Disease Transgenic Mice. Nat. Prod. Res. Dev. 2017, 29, 762. [Google Scholar] [CrossRef]
  344. Wu, Y.; Shi, Y.; Zheng, X.; Dang, Y.; Zhu, C.; Zhang, R.; Fu, Y.; Zhou, T.; Li, J. Lipophilic ferulic acid derivatives protect PC12 cells against oxidative damage via modulating β-amyloid aggregation and activating Nrf2 enzymes. Food Funct. 2020, 11, 4707–4718. [Google Scholar] [CrossRef]
  345. Rui, M.; Yi-qing, C.; Qin, C. Effects of ferulic acid on glial activation and inflammatory cytokines expression in the cerebral cortex of Alzheimer’s disease like model mice. Chin. Hosp. Pharm. J. 2018, 38, 50–53. [Google Scholar]
  346. Wang, Y.; Wang, Y.; Li, J.; Hua, L.; Han, B.; Zhang, Y.; Yang, X.; Zeng, Z.; Bai, H.; Yin, H.; et al. Effects of caffeic acid on learning deficits in a model of Alzheimer’s disease. Int. J. Mol. Med. 2016, 38, 869–875. [Google Scholar] [CrossRef] [Green Version]
  347. Paciello, F.; Di Pino, A.; Rolesi, R.; Troiani, D.; Paludetti, G.; Grassi, C.; Fetoni, A.R. Anti-oxidant and anti-inflammatory effects of caffeic acid: In vivo evidences in a model of noise-induced hearing loss. Food Chem. Toxicol. 2020, 143, 111555. [Google Scholar] [CrossRef]
  348. Kanimozhi, G.; Prasad, N.R. Anticancer Effect of Caffeic Acid on Human Cervical Cancer Cells. In Coffee in Health and Disease Prevention; Elsevier: Amsterdam, The Netherlands, 2015; pp. 655–661. ISBN 978-0-12-409517-5. [Google Scholar]
  349. Kilani-Jaziri, S.; Mokdad-Bzeouich, I.; Krifa, M.; Nasr, N.; Ghedira, K.; Chekir-Ghedira, L. Immunomodulatory and cellular anti-oxidant activities of caffeic, ferulic, and p -coumaric phenolic acids: A structure–activity relationship study. Drug Chem. Toxicol. 2017, 40, 416–424. [Google Scholar] [CrossRef]
  350. Damasceno, S.S.; Dantas, B.B.; Ribeiro-Filho, J.; Araújo, D.A.M.; Costa, J.G.M. Chemical Properties of Caffeic and Ferulic Acids in Biological System: Implications in Cancer Therapy. A Review. Curr. Pharm. Des. 2017, 23, 3015–3023. [Google Scholar] [CrossRef]
  351. Gulcin, I. Antioxidant activity of caffeic acid (3,4-dihydroxycinnamic acid). Toxicology 2006, 217, 213–220. [Google Scholar] [CrossRef]
  352. Liang, G.; Shi, B.; Luo, W.; Yang, J. The protective effect of caffeic acid on global cerebral ischemia-reperfusion injury in rats. Behav. Brain Funct. 2015, 11, 18. [Google Scholar] [CrossRef] [Green Version]
  353. Zhang, L.; Zhang, W.-P.; Chen, K.-D.; Qian, X.-D.; Fang, S.-H.; Wei, E.-Q. Caffeic acid attenuates neuronal damage, astrogliosis and glial scar formation in mouse brain with cryoinjury. Life Sci. 2007, 80, 530–537. [Google Scholar] [CrossRef]
  354. Zhang, Y.; Wu, Q.; Zhang, L.; Wang, Q.; Yang, Z.; Liu, J.; Feng, L. Caffeic acid reduces A53T α-synuclein by activating JNK/Bcl-2-mediated autophagy in vitro and improves behaviour and protects dopaminergic neurons in a mouse model of Parkinson’s disease. Pharmacol. Res. 2019, 150, 104538. [Google Scholar] [CrossRef] [PubMed]
  355. Zhou, Y.; Wang, J.; Chang, Y.; Li, R.; Sun, X.; Peng, L.; Zheng, W.; Qiu, W. Caffeic Acid Phenethyl Ester Protects against Experimental Autoimmune Encephalomyelitis by Regulating T Cell Activities. Oxid. Med. Cell. Longev. 2020, 2020, 7274342. [Google Scholar] [CrossRef] [PubMed]
  356. Wan, T.; Wang, Z.; Luo, Y.; Zhang, Y.; He, W.; Mei, Y.; Xue, J.; Li, M.; Pan, H.; Li, W.; et al. FA-97, a New Synthetic Caffeic Acid Phenethyl Ester Derivative, Protects against Oxidative Stress-Mediated Neuronal Cell Apoptosis and Scopolamine-Induced Cognitive Impairment by Activating Nrf2/HO-1 Signaling. Oxid. Med. Cell. Longev. 2019, 2019, 8239642. [Google Scholar] [CrossRef] [PubMed]
  357. Deshmukh, R.; Kaundal, M.; Bansal, V. Samardeep Caffeic acid attenuates oxidative stress, learning and memory deficit in intra-cerebroventricular streptozotocin induced experimental dementia in rats. Biomed. Pharmacother. 2016, 81, 56–62. [Google Scholar] [CrossRef] [PubMed]
  358. Haam, J.; Yakel, J.L. Cholinergic modulation of the hippocampal region and memory function. J. Neurochem. 2017, 142, 111–121. [Google Scholar] [CrossRef]
  359. Pei, K.; Ou, J.; Huang, J.; Ou, S. p -Coumaric acid and its conjugates: Dietary sources, pharmacokinetic properties and biological activities: P -Coumaric acid and its conjugates. J. Sci. Food Agric. 2016, 96, 2952–2962. [Google Scholar] [CrossRef]
  360. Boz, H. p -Coumaric acid in cereals: Presence, antioxidant and antimicrobial effects. Int. J. Food Sci. Technol. 2015, 50, 2323–2328. [Google Scholar] [CrossRef]
  361. Abdel-Wahab, M.; El-Mahdy, M.A.; Abd-Ellah, M.F.; Helal, G.K.; Khalifa, F.; Hamada, F.M.A. Influence of p-coumaric acid on doxorubicin-induced oxidative stress in rat’s heart. Pharmacol. Res. 2003, 48, 461–465. [Google Scholar] [CrossRef]
  362. Pragasam, S.J.; Venkatesan, V.; Rasool, M. Immunomodulatory and Anti-inflammatory Effect of p-Coumaric Acid, a Common Dietary Polyphenol on Experimental Inflammation in Rats. Inflammation 2013, 36, 169–176. [Google Scholar] [CrossRef]
  363. Kong, C.-S.; Jeong, C.-H.; Choi, J.-S.; Kim, K.-J.; Jeong, J.-W. Antiangiogenic Effects of P -Coumaric Acid in Human Endothelial Cells: P -COUMARIC ACID INHIBITS TUMOR ANGIOGENESIS. Phytother. Res. 2013, 27, 317–323. [Google Scholar] [CrossRef]
  364. Hudson, E.A.; Dinh, P.A.; Kokubun, T.; Simmonds, M.S.; Gescher, A. Characterization of potentially chemopreventive phenols in extracts of brown rice that inhibit the growth of human breast and colon cancer cells. Cancer Epidemiol. Biomark. Prev. Publ. Am. Assoc. Cancer Res. Cosponsored Am. Soc. Prev. Oncol. 2000, 9, 1163–1170. [Google Scholar]
  365. Kiliç, I.; Yeşiloğlu, Y. Spectroscopic studies on the antioxidant activity of p-coumaric acid. Spectrochim. Acta. A. Mol. Biomol. Spectrosc. 2013, 115, 719–724. [Google Scholar] [CrossRef]
  366. Roychoudhury, S.; Sinha, B.; Choudhury, B.P.; Jha, N.K.; Palit, P.; Kundu, S.; Mandal, S.C.; Kolesarova, A.; Yousef, M.I.; Ruokolainen, J.; et al. Scavenging Properties of Plant-Derived Natural Biomolecule Para-Coumaric Acid in the Prevention of Oxidative Stress-Induced Diseases. Antioxidants 2021, 10, 1205. [Google Scholar] [CrossRef]
  367. Guven, M.; Bozkurt Aras, A.; Akman, T.; Murat Sen, H.; Ozkan, A.; Salis, O.; Sehitoglu, I.; Kalkan, Y.; Silan, C.; Deniz, M.; et al. Neuroprotective effect of p-coumaric acid in rat model of embolic cerebral ischemia. Iran. J. Basic Med. Sci. 2015, 18, 356–363. [Google Scholar] [CrossRef]
  368. Sakamula, R.; Thong-asa, W. Neuroprotective effect of p-coumaric acid in mice with cerebral ischemia reperfusion injuries. Metab. Brain Dis. 2018, 33, 765–773. [Google Scholar] [CrossRef]
  369. Daroi, P.A.; Dhage, S.N.; Juvekar, A.R. p-Coumaric acid mitigates lipopolysaccharide induced brain damage via alleviating oxidative stress, inflammation and apoptosis. J. Pharm. Pharmacol. 2022, 74, 556–564. [Google Scholar] [CrossRef]
  370. Khan, M.S.; Althobaiti, M.S.; Almutairi, G.S.; Alokail, M.S.; Altwaijry, N.; Alenad, A.M.; Al-Bagmi, M.S.; Alafaleq, N.O. Elucidating the binding and inhibitory potential of p-coumaric acid against amyloid fibrillation and their cytotoxicity: Biophysical and docking analysis. Biophys. Chem. 2022, 291, 106823. [Google Scholar] [CrossRef]
  371. Ghaderi, S.; Gholipour, P.; Komaki, A.; Salehi, I.; Rashidi, K.; Esmaeil Khoshnam, S.; Rashno, M. p-Coumaric acid ameliorates cognitive and non-cognitive disturbances in a rat model of Alzheimer’s disease: The role of oxidative stress and inflammation. Int. Immunopharmacol. 2022, 112, 109295. [Google Scholar] [CrossRef]
  372. Bai, J.; Zhang, Y.; Tang, C.; Hou, Y.; Ai, X.; Chen, X.; Zhang, Y.; Wang, X.; Meng, X. Gallic acid: Pharmacological activities and molecular mechanisms involved in inflammation-related diseases. Biomed. Pharmacother. 2021, 133, 110985. [Google Scholar] [CrossRef]
  373. Daglia, M.; Lorenzo, A.; Nabavi, S.; Talas, Z.; Nabavi, S. Polyphenols: Well Beyond The Antioxidant Capacity: Gallic Acid and Related Compounds as Neuroprotective Agents: You are What You Eat! Curr. Pharm. Biotechnol. 2014, 15, 362–372. [Google Scholar] [CrossRef]
  374. Kahkeshani, N.; Farzaei, F.; Fotouhi, M.; Alavi, S.S.; Bahramsoltani, R.; Naseri, R.; Momtaz, S.; Abbasabadi, Z.; Rahimi, R.; Farzaei, M.H.; et al. Pharmacological effects of gallic acid in health and disease: A mechanistic review. Iran. J. Basic Med. Sci. 2019, 22, 225–237. [Google Scholar] [CrossRef] [PubMed]
  375. Punithavathi, V.R.; Prince, P.S.M.; Kumar, R.; Selvakumari, J. Antihyperglycaemic, antilipid peroxidative and antioxidant effects of gallic acid on streptozotocin induced diabetic Wistar rats. Eur. J. Pharmacol. 2011, 650, 465–471. [Google Scholar] [CrossRef] [PubMed]
  376. Kim, Y.-J. Antimelanogenic and Antioxidant Properties of Gallic Acid. Biol. Pharm. Bull. 2007, 30, 1052–1055. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  377. Kim, M.-J.; Seong, A.-R.; Yoo, J.-Y.; Jin, C.-H.; Lee, Y.-H.; Kim, Y.J.; Lee, J.; Jun, W.J.; Yoon, H.-G. Gallic acid, a histone acetyltransferase inhibitor, suppresses β-amyloid neurotoxicity by inhibiting microglial-mediated neuroinflammation. Mol. Nutr. Food Res. 2011, 55, 1798–1808. [Google Scholar] [CrossRef]
  378. Verma, S.; Singh, A.; Mishra, A. Gallic acid: Molecular rival of cancer. Environ. Toxicol. Pharmacol. 2013, 35, 473–485. [Google Scholar] [CrossRef]
  379. Kratz, J.M.; Andrighetti-Fröhner, C.R.; Kolling, D.J.; Leal, P.C.; Cirne-Santos, C.C.; Yunes, R.A.; Nunes, R.J.; Trybala, E.; Bergström, T.; Frugulhetti, I.C.; et al. Anti-HSV-1 and anti-HIV-1 activity of gallic acid and pentyl gallate. Mem. Inst. Oswaldo Cruz 2008, 103, 437–442. [Google Scholar] [CrossRef]
  380. Mirshekar, M.A.; Sarkaki, A.; Farbood, Y.; Gharib Naseri, M.K.; Badavi, M.; Mansouri, M.T.; Haghparast, A. Neuroprotective effects of gallic acid in a rat model of traumatic brain injury: Behavioral, electrophysiological and molecular studies. Iran. J. Basic Med. Sci. 2018, 21, 1056–1063. [Google Scholar] [CrossRef]
  381. Mansouri, M.T.; Farbood, Y.; Sameri, M.J.; Sarkaki, A.; Naghizadeh, B.; Rafeirad, M. Neuroprotective effects of oral gallic acid against oxidative stress induced by 6-hydroxydopamine in rats. Food Chem. 2013, 138, 1028–1033. [Google Scholar] [CrossRef]
  382. Maya, S.; Prakash, T.; Goli, D. Evaluation of neuroprotective effects of wedelolactone and gallic acid on aluminium-induced neurodegeneration: Relevance to sporadic amyotrophic lateral sclerosis. Eur. J. Pharmacol. 2018, 835, 41–51. [Google Scholar] [CrossRef]
  383. Shabani, S.; Rabiei, Z.; Amini-Khoei, H. Exploring the multifaceted neuroprotective actions of gallic acid: A review. Int. J. Food Prop. 2020, 23, 736–752. [Google Scholar] [CrossRef]
  384. Mori, T.; Koyama, N.; Yokoo, T.; Segawa, T.; Maeda, M.; Sawmiller, D.; Tan, J.; Town, T. Gallic acid is a dual α/β-secretase modulator that reverses cognitive impairment and remediates pathology in Alzheimer mice. J. Biol. Chem. 2020, 295, 16251–16266. [Google Scholar] [CrossRef]
  385. Ban, J.Y.; Nguyen, H.T.T.; Lee, H.-J.; Cho, S.O.; Ju, H.S.; Kim, J.Y.; Bae, K.; Song, K.-S.; Seong, Y.H. Neuroprotective Properties of Gallic Acid from Sanguisorbae Radix on Amyloid. BETA. Protein (25-35)-Induced Toxicity in Cultured Rat Cortical Neurons. Biol. Pharm. Bull. 2008, 31, 149–153. [Google Scholar] [CrossRef] [Green Version]
  386. Gerzson, M.F.B.; Bona, N.P.; Soares, M.S.P.; Teixeira, F.C.; Rahmeier, F.L.; Carvalho, F.B.; da Cruz Fernandes, M.; Onzi, G.; Lenz, G.; Gonçales, R.A.; et al. Tannic Acid Ameliorates STZ-Induced Alzheimer’s Disease-Like Impairment of Memory, Neuroinflammation, Neuronal Death and Modulates Akt Expression. Neurotox. Res. 2020, 37, 1009–1017. [Google Scholar] [CrossRef]
  387. Ardah, M.T.; Bharathan, G.; Kitada, T.; Haque, M.E. Ellagic Acid Prevents Dopamine Neuron Degeneration from Oxidative Stress and Neuroinflammation in MPTP Model of Parkinson’s Disease. Biomolecules 2020, 10, 1519. [Google Scholar] [CrossRef]
  388. Ahmed, T.; Setzer, W.N.; Fazel Nabavi, S.; Erdogan Orhan, I.; Braidy, N.; Sobarzo-Sanchez, E.; Mohammad Nabavi, S. Insights into Effects of Ellagic Acid on the Nervous System: A Mini Review. Curr. Pharm. Des. 2016, 22, 1350–1360. [Google Scholar] [CrossRef]
  389. Wang, Y.; Wu, Y.; Liang, C.; Tan, R.; Tan, L.; Tan, R. Pharmacodynamic Effect of Ellagic Acid on Ameliorating Cerebral Ischemia/Reperfusion Injury. Pharmacology 2019, 104, 320–331. [Google Scholar] [CrossRef]
  390. Aslan, A.; Gok, O.; Beyaz, S.; Arslan, E.; Erman, O.; Ağca, C.A. The preventive effect of ellagic acid on brain damage in rats via regulating of Nrf-2, NF-kB and apoptotic pathway. J. Food Biochem. 2020, 44, e13217. [Google Scholar] [CrossRef]
  391. Lee, W.-J.; Ou, H.-C.; Hsu, W.-C.; Chou, M.-M.; Tseng, J.-J.; Hsu, S.-L.; Tsai, K.-L.; Sheu, W.H.-H. Ellagic acid inhibits oxidized LDL-mediated LOX-1 expression, ROS generation, and inflammation in human endothelial cells. J. Vasc. Surg. 2010, 52, 1290–1300. [Google Scholar] [CrossRef] [Green Version]
  392. Dalvi, L.T.; Moreira, D.C.; Andrade, R.; Ginani, J.; Alonso, A.; Hermes-Lima, M. Ellagic acid inhibits iron-mediated free radical formation. Spectrochim. Acta. A Mol. Biomol. Spectrosc. 2017, 173, 910–917. [Google Scholar] [CrossRef]
  393. Ríos, J.-L.; Giner, R.; Marín, M.; Recio, M. A Pharmacological Update of Ellagic Acid. Planta Med. 2018, 84, 1068–1093. [Google Scholar] [CrossRef] [Green Version]
  394. Gomes, B.A.Q.; Silva, J.P.B.; Romeiro, C.F.R.; dos Santos, S.M.; Rodrigues, C.A.; Gonçalves, P.R.; Sakai, J.T.; Mendes, P.F.S.; Varela, E.L.P.; Monteiro, M.C. Neuroprotective Mechanisms of Resveratrol in Alzheimer’s Disease: Role of SIRT1. Oxid. Med. Cell. Longev. 2018, 2018, 8152373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  395. Rocha-González, H.I.; Ambriz-Tututi, M.; Granados-Soto, V. Resveratrol: A Natural Compound with Pharmacological Potential in Neurodegenerative Diseases. CNS Neurosci. Ther. 2008, 14, 234–247. [Google Scholar] [CrossRef] [PubMed]
  396. Rossi, M.; Caruso, F.; Antonioletti, R.; Viglianti, A.; Traversi, G.; Leone, S.; Basso, E.; Cozzi, R. Scavenging of hydroxyl radical by resveratrol and related natural stilbenes after hydrogen peroxide attack on DNA. Chem. Biol. Interact. 2013, 206, 175–185. [Google Scholar] [CrossRef] [PubMed]
  397. Sadi, G.; Konat, D. Resveratrol regulates oxidative biomarkers and antioxidant enzymes in the brain of streptozotocin-induced diabetic rats. Pharm. Biol. 2015, 54, 1156–1163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  398. Tong, W.; Chen, X.; Song, X.; Chen, Y.; Jia, R.; Zou, Y.; Li, L.; Yin, L.; He, C.; Liang, X.; et al. Resveratrol inhibits LPS-induced inflammation through suppressing the signaling cascades of TLR4-NF-κB/MAPKs/IRF3. Exp. Ther. Med. 2019, 19, 1824–1834. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  399. Moussa, C.; Hebron, M.; Huang, X.; Ahn, J.; Rissman, R.A.; Aisen, P.S.; Turner, R.S. Resveratrol regulates neuro-inflammation and induces adaptive immunity in Alzheimer’s disease. J. Neuroinflamm. 2017, 14, 1. [Google Scholar] [CrossRef] [PubMed]
  400. Turner, R.S.; Thomas, R.G.; Craft, S.; van Dyck, C.H.; Mintzer, J.; Reynolds, B.A.; Brewer, J.B.; Rissman, R.A.; Raman, R.; Aisen, P.S.; et al. A randomized, double-blind, placebo-controlled trial of resveratrol for Alzheimer disease. Neurology 2015, 85, 1383–1391. [Google Scholar] [CrossRef] [PubMed]
  401. Frozza, R.L.; Bernardi, A.; Hoppe, J.B.; Meneghetti, A.B.; Matté, A.; Battastini, A.M.O.; Pohlmann, A.R.; Guterres, S.S.; Salbego, C. Neuroprotective Effects of Resveratrol Against Aβ Administration in Rats are Improved by Lipid-Core Nanocapsules. Mol. Neurobiol. 2013, 47, 1066–1080. [Google Scholar] [CrossRef]
  402. Zhang, W.; Kung, M.-P.; Oya, S.; Hou, C.; Kung, H.F. 18F-labeled styrylpyridines as PET agents for amyloid plaque imaging. Nucl. Med. Biol. 2007, 34, 89–97. [Google Scholar] [CrossRef]
  403. Zhu, L.-N.; Mei, X.; Zhang, Z.-G.; Xie, Y.; Lang, F. Curcumin intervention for cognitive function in different types of people: A systematic review and meta-analysis: Curcumin intervention for cognitive function. Phytother. Res. 2019, 33, 524–533. [Google Scholar] [CrossRef] [Green Version]
  404. Hanai, H.; Sugimoto, K. Curcumin has Bright Prospects for the Treatment of Inflammatory Bowel Disease. Curr. Pharm. Des. 2009, 15, 2087–2094. [Google Scholar] [CrossRef]
  405. Jackson, J.K.; Higo, T.; Hunter, W.L.; Burt, H.M. The antioxidants curcumin and quercetin inhibit inflammatory processes associated with arthritis. Inflamm. Res. 2006, 55, 168–175. [Google Scholar] [CrossRef]
  406. Nebrisi, E.E. Neuroprotective Activities of Curcumin in Parkinson’s Disease: A Review of the Literature. Int. J. Mol. Sci. 2021, 22, 11248. [Google Scholar] [CrossRef]
  407. Wu, S.; Guo, T.; Qi, W.; Li, Y.; Gu, J.; Liu, C.; Sha, Y.; Yang, B.; Hu, S.; Zong, X. Curcumin ameliorates ischemic stroke injury in rats by protecting the integrity of the blood-brain barrier. Exp. Ther. Med. 2021, 22, 783. [Google Scholar] [CrossRef]
  408. Labanca, F.; Ullah, H.; Khan, H.; Milella, L.; Xiao, J.; Dajic-Stevanovic, Z.; Jeandet, P. Therapeutic and Mechanistic Effects of Curcumin in Huntington’s Disease. Curr. Neuropharmacol. 2021, 19, 1007–1018. [Google Scholar] [CrossRef]
  409. Alisi, I.O.; Uzairu, A.; Abechi, S.E.; Idris, S.O. Evaluation of the antioxidant properties of curcumin derivatives by genetic function algorithm. J. Adv. Res. 2018, 12, 47–54. [Google Scholar] [CrossRef]
  410. Jurenka, J.S. Anti-inflammatory properties of curcumin, a major constituent of Curcuma longa: A review of preclinical and clinical research. Altern. Med. Rev. J. Clin. Ther. 2009, 14, 141–153. [Google Scholar]
  411. Zheng, K.; Dai, X.; Xiao, N.; Wu, X.; Wei, Z.; Fang, W.; Zhu, Y.; Zhang, J.; Chen, X. Curcumin Ameliorates Memory Decline via Inhibiting BACE1 Expression and β-Amyloid Pathology in 5×FAD Transgenic Mice. Mol. Neurobiol. 2017, 54, 1967–1977. [Google Scholar] [CrossRef]
  412. Xiong, Z.; Hongmei, Z.; Lu, S.; Yu, L. Curcumin mediates presenilin-1 activity to reduce β-amyloid production in a model of Alzheimer’s disease. Pharmacol. Rep. 2011, 63, 1101–1108. [Google Scholar] [CrossRef]
  413. Ono, K.; Hasegawa, K.; Naiki, H.; Yamada, M. Curcumin has potent anti-amyloidogenic effects for Alzheimer’s?-amyloid fibrils in vitro. J. Neurosci. Res. 2004, 75, 742–750. [Google Scholar] [CrossRef]
  414. Huang, H.-C.; Tang, D.; Xu, K.; Jiang, Z.-F. Curcumin attenuates amyloid-β-induced tau hyperphosphorylation in human neuroblastoma SH-SY5Y cells involving PTEN/Akt/GSK-3β signaling pathway. J. Recept. Signal Transduct. 2014, 34, 26–37. [Google Scholar] [CrossRef] [PubMed]
  415. Cianciulli, A.; Calvello, R.; Porro, C.; Trotta, T.; Salvatore, R.; Panaro, M.A. PI3k/Akt signalling pathway plays a crucial role in the anti-inflammatory effects of curcumin in LPS-activated microglia. Int. Immunopharmacol. 2016, 36, 282–290. [Google Scholar] [CrossRef] [PubMed]
  416. Shi, X.; Zheng, Z.; Li, J.; Xiao, Z.; Qi, W.; Zhang, A.; Wu, Q.; Fang, Y. Curcumin inhibits Aβ-induced microglial inflammatory responses in vitro: Involvement of ERK1/2 and p38 signaling pathways. Neurosci. Lett. 2015, 594, 105–110. [Google Scholar] [CrossRef] [PubMed]
  417. Voulgaropoulou, S.D.; van Amelsvoort, T.A.M.J.; Prickaerts, J.; Vingerhoets, C. The effect of curcumin on cognition in Alzheimer’s disease and healthy aging: A systematic review of pre-clinical and clinical studies. Brain Res. 2019, 1725, 146476. [Google Scholar] [CrossRef] [PubMed]
  418. Baum, L.; Lam, C.W.K.; Cheung, S.K.-K.; Kwok, T.; Lui, V.; Tsoh, J.; Lam, L.; Leung, V.; Hui, E.; Ng, C.; et al. Six-Month Randomized, Placebo-Controlled, Double-Blind, Pilot Clinical Trial of Curcumin in Patients with Alzheimer Disease. J. Clin. Psychopharmacol. 2008, 28, 110–113. [Google Scholar] [CrossRef] [Green Version]
  419. Tang, M.; Taghibiglou, C. The Mechanisms of Action of Curcumin in Alzheimer’s Disease. J. Alzheimers Dis. 2017, 58, 1003–1016. [Google Scholar] [CrossRef]
  420. Ruan, Y.; Xiong, Y.; Fang, W.; Yu, Q.; Mai, Y.; Cao, Z.; Wang, K.; Lei, M.; Xu, J.; Liu, Y.; et al. Highly sensitive Curcumin-conjugated nanotheranostic platform for detecting amyloid-beta plaques by magnetic resonance imaging and reversing cognitive deficits of Alzheimer’s disease via NLRP3-inhibition. J. Nanobiotechnol. 2022, 20, 322. [Google Scholar] [CrossRef]
  421. Lazar, A.N.; Mourtas, S.; Youssef, I.; Parizot, C.; Dauphin, A.; Delatour, B.; Antimisiaris, S.G.; Duyckaerts, C. Curcumin-conjugated nanoliposomes with high affinity for Aβ deposits: Possible applications to Alzheimer disease. Nanomedicine Nanotechnol. Biol. Med. 2013, 9, 712–721. [Google Scholar] [CrossRef]
  422. Goozee, K.G.; Shah, T.M.; Sohrabi, H.R.; Rainey-Smith, S.R.; Brown, B.; Verdile, G.; Martins, R.N. Examining the potential clinical value of curcumin in the prevention and diagnosis of Alzheimer’s disease. Br. J. Nutr. 2016, 115, 449–465. [Google Scholar] [CrossRef]
  423. Szopa, A.; Dziurka, M.; Warzecha, A.; Kubica, P.; Klimek-Szczykutowicz, M.; Ekiert, H. Targeted Lignan Profiling and Anti-Inflammatory Properties of Schisandra rubriflora and Schisandra chinensis Extracts. Molecules 2018, 23, 3103. [Google Scholar] [CrossRef] [Green Version]
  424. Eklund, P.C.; Långvik, O.K.; Wärnå, J.P.; Salmi, T.O.; Willför, S.M.; Sjöholm, R.E. Chemical studies on antioxidant mechanisms and free radical scavenging properties of lignans. Org. Biomol. Chem. 2005, 3, 3336. [Google Scholar] [CrossRef]
  425. Laura Esquivel-Campos, A.; Pérez-Gutiérrez, S.; Sánchez-Pérez, L.; Campos-Xolalpa, N.; Pérez-Ramos, J. Cytotoxicity and Antitumor Action of Lignans and Neolignans. In Secondary Metabolites—Trends and Reviews; Vijayakumar, R., Selvapuram Sudalaimuthu Raja, S., Eds.; IntechOpen: London, UK, 2022; ISBN 978-1-80355-207-1. [Google Scholar]
  426. Sowndhararajan, K.; Deepa, P.; Kim, M.; Park, S.J.; Kim, S. An overview of neuroprotective and cognitive enhancement properties of lignans from Schisandra chinensis. Biomed. Pharmacother. 2018, 97, 958–968. [Google Scholar] [CrossRef]
  427. Hsieh, M.-T.; Wu, C.-R.; Wang, W.-H.; Lin, L.-W. The ameliorating effect of the water layer of fructusschisandrae on cycloheximide-induced amnesia in rats: Interaction with drugs acting at neurotransmitter receptors. Pharmacol. Res. 2001, 43, 17–22. [Google Scholar] [CrossRef]
  428. Yu, J.; Kwon, H.; Cho, E.; Jeon, J.; Kang, R.H.; Youn, K.; Jun, M.; Lee, Y.C.; Ryu, J.H.; Kim, D.H. The effects of pinoresinol on cholinergic dysfunction-induced memory impairments and synaptic plasticity in mice. Food Chem. Toxicol. 2019, 125, 376–382. [Google Scholar] [CrossRef]
  429. Xu, M.; Dong, Y.; Wan, S.; Yan, T.; Cao, J.; Wu, L.; Bi, K.; Jia, Y. Schisantherin B ameliorates Aβ 1–42 -induced cognitive decline via restoration of GLT-1 in a mouse model of Alzheimer’s disease. Physiol. Behav. 2016, 167, 265–273. [Google Scholar] [CrossRef]
  430. Song, J.-X.; Lin, X.; Wong, R.N.-S.; Sze, S.C.-W.; Tong, Y.; Shaw, P.-C.; Zhang, Y.-B. Protective effects of dibenzocyclooctadiene lignans from Schisandra chinensis against beta-amyloid and homocysteine neurotoxicity in PC12 cells. Phytother. Res. 2010, 25, 435–443. [Google Scholar] [CrossRef]
  431. Zhao, X.; Liu, C.; Xu, M.; Li, X.; Bi, K.; Jia, Y. Total Lignans of Schisandra chinensis Ameliorates Aβ1-42-Induced Neurodegeneration with Cognitive Impairment in Mice and Primary Mouse Neuronal Cells. PLoS ONE 2016, 11, e0152772. [Google Scholar] [CrossRef]
  432. Hu, D.; Cao, Y.; He, R.; Han, N.; Liu, Z.; Miao, L.; Yin, J. Schizandrin, an Antioxidant Lignan from Schisandra chinensis, Ameliorates A β1–42 -Induced Memory Impairment in Mice. Oxid. Med. Cell. Longev. 2012, 2012, 721721. [Google Scholar] [CrossRef] [Green Version]
  433. Liu, Q.; Xie, T.; Xi, Y.; Li, L.; Mo, F.; Liu, X.; Liu, Z.; Gao, J.-M.; Yuan, T. Sesamol Attenuates Amyloid Peptide Accumulation and Cognitive Deficits in APP/PS1 Mice: The Mediating Role of the Gut–Brain Axis. J. Agric. Food Chem. 2021, 69, 12717–12729. [Google Scholar] [CrossRef]
  434. Liu, Q.; Wang, J.; Lin, B.; Cheng, Z.-Y.; Bai, M.; Shi, S.; Huang, X.-X.; Song, S.-J. Phenylpropanoids and lignans from Prunus tomentosa seeds as efficient β-amyloid (Aβ) aggregation inhibitors. Bioorganic Chem. 2019, 84, 269–275. [Google Scholar] [CrossRef]
  435. Gaucher, C.; Boudier, A.; Bonetti, J.; Clarot, I.; Leroy, P.; Parent, M. Glutathione: Antioxidant Properties Dedicated to Nanotechnologies. Antioxidants 2018, 7, 62. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  436. Lv, H.; Zhen, C.; Liu, J.; Yang, P.; Hu, L.; Shang, P. Unraveling the Potential Role of Glutathione in Multiple Forms of Cell Death in Cancer Therapy. Oxid. Med. Cell. Longev. 2019, 2019, 3150145. [Google Scholar] [CrossRef] [Green Version]
  437. Traverso, N.; Ricciarelli, R.; Nitti, M.; Marengo, B.; Furfaro, A.L.; Pronzato, M.A.; Marinari, U.M.; Domenicotti, C. Role of Glutathione in Cancer Progression and Chemoresistance. Oxid. Med. Cell. Longev. 2013, 2013, 972913. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  438. Wu, G.; Fang, Y.-Z.; Yang, S.; Lupton, J.R.; Turner, N.D. Glutathione Metabolism and Its Implications for Health. J. Nutr. 2004, 134, 489–492. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  439. Chen, J.J.; Thiyagarajah, M.; Song, J.; Chen, C.; Herrmann, N.; Gallagher, D.; Rapoport, M.J.; Black, S.E.; Ramirez, J.; Andreazza, A.C.; et al. Altered central and blood glutathione in Alzheimer’s disease and mild cognitive impairment: A meta-analysis. Alzheimers Res. Ther. 2022, 14, 23. [Google Scholar] [CrossRef]
  440. Lin, C.-H.; Lane, H.-Y. Plasma Glutathione Levels Decreased with Cognitive Decline among People with Mild Cognitive Impairment (MCI): A Two-Year Prospective Study. Antioxidants 2021, 10, 1839. [Google Scholar] [CrossRef]
  441. Mandal, P.K.; Tripathi, M.; Sugunan, S. Brain oxidative stress: Detection and mapping of anti-oxidant marker ‘Glutathione’ in different brain regions of healthy male/female, MCI and Alzheimer patients using non-invasive magnetic resonance spectroscopy. Biochem. Biophys. Res. Commun. 2012, 417, 43–48. [Google Scholar] [CrossRef]
  442. Duffy, S.L.; Lagopoulos, J.; Hickie, I.B.; Diamond, K.; Graeber, M.B.; Lewis, S.J.G.; Naismith, S.L. Glutathione relates to neuropsychological functioning in mild cognitive impairment. Alzheimers Dement. 2014, 10, 67–75. [Google Scholar] [CrossRef]
  443. Mandal, P.K.; Saharan, S.; Tripathi, M.; Murari, G. Brain Glutathione Levels—A Novel Biomarker for Mild Cognitive Impairment and Alzheimer’s Disease. Biol. Psychiatry 2015, 78, 702–710. [Google Scholar] [CrossRef]
  444. Shukla, D.; Mandal, P.K.; Mishra, R.; Punjabi, K.; Dwivedi, D.; Tripathi, M.; Badhautia, V. Hippocampal Glutathione Depletion and pH Increment in Alzheimer’s Disease: An in vivo MRS Study. J. Alzheimers Dis. 2021, 84, 1139–1152. [Google Scholar] [CrossRef]
  445. Izumi, H.; Sato, K.; Kojima, K.; Saito, T.; Saido, T.C.; Fukunaga, K. Oral glutathione administration inhibits the oxidative stress and the inflammatory responses in AppNL−G-F/NL−G-F knock-in mice. Neuropharmacology 2020, 168, 108026. [Google Scholar] [CrossRef]
  446. Christopher Kwon, Y.I.; Xie, W.; Zhu, H.; Xie, J.; Shinn, K.; Juckel, N.; Vince, R.; More, S.S.; Lee, M.K. γ-Glutamyl-Transpeptidase-Resistant Glutathione Analog Attenuates Progression of Alzheimer’s Disease-like Pathology and Neurodegeneration in a Mouse Model. Antioxidants 2021, 10, 1796. [Google Scholar] [CrossRef]
  447. Li, Q.; Cui, J.; Fang, C.; Liu, M.; Min, G.; Li, L. S-Adenosylmethionine Attenuates Oxidative Stress and Neuroinflammation Induced by Amyloid-β Through Modulation of Glutathione Metabolism. J. Alzheimers Dis. 2017, 58, 549–558. [Google Scholar] [CrossRef]
  448. Yang, H.; Xie, Z.; Wei, L.; Ding, M.; Wang, P.; Bi, J. Glutathione-mimetic D609 alleviates memory deficits and reduces amyloid-β deposition in an AβPP/PS1 transgenic mouse model. NeuroReport 2018, 29, 833–838. [Google Scholar] [CrossRef]
  449. Braidy, N.; Zarka, M.; Jugder, B.-E.; Welch, J.; Jayasena, T.; Chan, D.K.Y.; Sachdev, P.; Bridge, W. The Precursor to Glutathione (GSH), γ-Glutamylcysteine (GGC), Can Ameliorate Oxidative Damage and Neuroinflammation Induced by Aβ40 Oligomers in Human Astrocytes. Front. Aging Neurosci. 2019, 11, 177. [Google Scholar] [CrossRef] [Green Version]
  450. Barkats, M.; Millecamps, S.; Abrioux, P.; Geoffroy, M.-C.; Mallet, J. Overexpression of Glutathione Peroxidase Increases the Resistance of Neuronal Cells to Aβ-Mediated Neurotoxicity. J. Neurochem. 2002, 75, 1438–1446. [Google Scholar] [CrossRef]
  451. Shin, E.-J.; Chung, Y.H.; Sharma, N.; Nguyen, B.T.; Lee, S.H.; Kang, S.W.; Nah, S.-Y.; Wie, M.B.; Nabeshima, T.; Jeong, J.H.; et al. Glutathione Peroxidase-1 Knockout Facilitates Memory Impairment Induced by β-Amyloid (1–42) in Mice via Inhibition of PKC βII-Mediated ERK Signaling; Application with Glutathione Peroxidase-1 Gene-Encoded Adenovirus Vector. Neurochem. Res. 2020, 45, 2991–3002. [Google Scholar] [CrossRef]
  452. Song, J.; Kang, S.M.; Lee, W.T.; Park, K.A.; Lee, K.M.; Lee, J.E. Glutathione Protects Brain Endothelial Cells from Hydrogen Peroxide-Induced Oxidative Stress by Increasing Nrf2 Expression. Exp. Neurobiol. 2014, 23, 93–103. [Google Scholar] [CrossRef] [Green Version]
  453. Ansary, J.; Forbes-Hernández, T.Y.; Gil, E.; Cianciosi, D.; Zhang, J.; Elexpuru-Zabaleta, M.; Simal-Gandara, J.; Giampieri, F.; Battino, M. Potential Health Benefit of Garlic Based on Human Intervention Studies: A Brief Overview. Antioxidants 2020, 9, 619. [Google Scholar] [CrossRef]
  454. Qu, Z.; Mossine, V.V.; Cui, J.; Sun, G.Y.; Gu, Z. Protective Effects of AGE and Its Components on Neuroinflammation and Neurodegeneration. NeuroMolecular Med. 2016, 18, 474–482. [Google Scholar] [CrossRef]
  455. Ray, B.; Chauhan, N.B.; Lahiri, D.K. The “aged garlic extract:” (AGE) and one of its active ingredients S-allyl-L-cysteine (SAC) as potential preventive and therapeutic agents for Alzheimer’s disease (AD). Curr. Med. Chem. 2011, 18, 3306–3313. [Google Scholar] [CrossRef] [PubMed]
  456. Tedeschi, P.; Nigro, M.; Travagli, A.; Catani, M.; Cavazzini, A.; Merighi, S.; Gessi, S. Therapeutic Potential of Allicin and Aged Garlic Extract in Alzheimer’s Disease. Int. J. Mol. Sci. 2022, 23, 6950. [Google Scholar] [CrossRef] [PubMed]
  457. Griffin, B.; Selassie, M.; Gwebu, E.T. Effect of Aged Garlic Extract on the Cytotoxicity of Alzheimer β-Amyloid Peptide in Neuronal PC12 Cells. Nutr. Neurosci. 2000, 3, 139–142. [Google Scholar] [CrossRef] [PubMed]
  458. Gupta, V.B.; Indi, S.S.; Rao, K.S.J. Garlic extract exhibits antiamyloidogenic activity on amyloid-beta fibrillogenesis: Relevance to Alzheimer’s disease. Phytother. Res. 2009, 23, 111–115. [Google Scholar] [CrossRef] [PubMed]
  459. Gupta, V.B.; Rao, K.S.J. Anti-amyloidogenic activity of S-allyl-l-cysteine and its activity to destabilize Alzheimer’s β-amyloid fibrils in vitro. Neurosci. Lett. 2007, 429, 75–80. [Google Scholar] [CrossRef]
  460. Jeong, J.H.; Jeong, H.R.; Jo, Y.N.; Kim, H.J.; Shin, J.H.; Heo, H.J. Ameliorating effects of aged garlic extracts against Aβ-induced neurotoxicity and cognitive impairment. BMC Complement. Altern. Med. 2013, 13, 268. [Google Scholar] [CrossRef] [Green Version]
  461. Luo, J.-F.; Dong, Y.; Chen, J.-Y.; Lu, J.-H. The effect and underlying mechanisms of garlic extract against cognitive impairment and Alzheimer’s disease: A systematic review and meta-analysis of experimental animal studies. J. Ethnopharmacol. 2021, 280, 114423. [Google Scholar] [CrossRef]
  462. Chauhan, N.B.; Sandoval, J. Amelioration of early cognitive deficits by aged garlic extract in Alzheimer’s transgenic mice. Phytother. Res. 2007, 21, 629–640. [Google Scholar] [CrossRef]
  463. Elosta, A.; Slevin, M.; Rahman, K.; Ahmed, N. Aged garlic has more potent antiglycation and antioxidant properties compared to fresh garlic extract in vitro. Sci. Rep. 2017, 7, 39613. [Google Scholar] [CrossRef]
  464. Nillert, N.; Pannangrong, W.; Welbat, J.; Chaijaroonkhanarak, W.; Sripanidkulchai, K.; Sripanidkulchai, B. Neuroprotective Effects of Aged Garlic Extract on Cognitive Dysfunction and Neuroinflammation Induced by β-Amyloid in Rats. Nutrients 2017, 9, 24. [Google Scholar] [CrossRef] [Green Version]
  465. Thorajak, P.; Pannangrong, W.; Welbat, J.U.; Chaijaroonkhanarak, W.; Sripanidkulchai, K.; Sripanidkulchai, B. Effects of Aged Garlic Extract on Cholinergic, Glutamatergic and GABAergic Systems with Regard to Cognitive Impairment in Aβ-Induced Rats. Nutrients 2017, 9, 686. [Google Scholar] [CrossRef] [Green Version]
  466. Burbaeva, G.S.; Boksha, I.S.; Tereshkina, E.B.; Savushkina, O.K.; Prokhorova, T.A.; Vorobyeva, E.A. Glutamate and GABA-Metabolizing Enzymes in Post-mortem Cerebellum in Alzheimer’s Disease: Phosphate-Activated Glutaminase and Glutamic Acid Decarboxylase. The Cerebellum 2014, 13, 607–615. [Google Scholar] [CrossRef]
  467. Rodriguez-Perdigon, M.; Tordera, R.M.; Gil-Bea, F.J.; Gerenu, G.; Ramirez, M.J.; Solas, M. Down-regulation of glutamatergic terminals (VGLUT1) driven by Aβ in Alzheimer’s disease: Aβ and VGLUT1 in Alzheimer’s Disease. Hippocampus 2016, 26, 1303–1312. [Google Scholar] [CrossRef]
  468. Ray, B.; Chauhan, N.B.; Lahiri, D.K. Oxidative insults to neurons and synapse are prevented by aged garlic extract and S-allyl-l-cysteine treatment in the neuronal culture and APP-Tg mouse model: Oxidative insults to neurons and synapse: Role of AGE and SAC. J. Neurochem. 2011, 117, 388–402. [Google Scholar] [CrossRef] [Green Version]
  469. Javed, H.; Khan, M.M.; Khan, A.; Vaibhav, K.; Ahmad, A.; Khuwaja, G.; Ahmed, M.E.; Raza, S.S.; Ashafaq, M.; Tabassum, R.; et al. S-allyl cysteine attenuates oxidative stress associated cognitive impairment and neurodegeneration in mouse model of streptozotocin-induced experimental dementia of Alzheimer’s type. Brain Res. 2011, 1389, 133–142. [Google Scholar] [CrossRef]
  470. Chauhan, N.B. Effect of aged garlic extract on APP processing and tau phosphorylation in Alzheimer’s transgenic model Tg2576. J. Ethnopharmacol. 2006, 108, 385–394. [Google Scholar] [CrossRef]
  471. Nadeem, M.S.; Kazmi, I.; Ullah, I.; Muhammad, K.; Anwar, F. Allicin, an Antioxidant and Neuroprotective Agent, Ameliorates Cognitive Impairment. Antioxidants 2021, 11, 87. [Google Scholar] [CrossRef]
  472. Li, X.-H.; Li, C.-Y.; Lu, J.-M.; Tian, R.-B.; Wei, J. Allicin ameliorates cognitive deficits ageing-induced learning and memory deficits through enhancing of Nrf2 antioxidant signaling pathways. Neurosci. Lett. 2012, 514, 46–50. [Google Scholar] [CrossRef]
  473. Kaur, S.; Raj, K.; Gupta, Y.K.; Singh, S. Allicin ameliorates aluminium- and copper-induced cognitive dysfunction in Wistar rats: Relevance to neuro-inflammation, neurotransmitters and Aβ(1–42) analysis. JBIC J. Biol. Inorg. Chem. 2021, 26, 495–510. [Google Scholar] [CrossRef]
  474. Li, X.-H.; Li, C.-Y.; Xiang, Z.-G.; Zhong, F.; Chen, Z.-Y.; Lu, J.-M. Allicin can reduce neuronal death and ameliorate the spatial memory impairment in Alzheimer’s disease models. Neurosci. Riyadh Saudi Arab. 2010, 15, 237–243. [Google Scholar]
  475. Zhang, H.; Wang, P.; Xue, Y.; Liu, L.; Li, Z.; Liu, Y. Allicin ameliorates cognitive impairment in APP/PS1 mice via Suppressing oxidative stress by Blocking JNK Signaling Pathways. Tissue Cell 2018, 50, 89–95. [Google Scholar] [CrossRef] [PubMed]
  476. Zhu, Y.-F.; Li, X.-H.; Yuan, Z.-P.; Li, C.-Y.; Tian, R.-B.; Jia, W.; Xiao, Z.-P. Allicin improves endoplasmic reticulum stress-related cognitive deficits via PERK/Nrf2 antioxidative signaling pathway. Eur. J. Pharmacol. 2015, 762, 239–246. [Google Scholar] [CrossRef] [PubMed]
  477. de Sousa, C.N.S.; da Silva Leite, C.M.G.; da Silva Medeiros, I.; Vasconcelos, L.C.; Cabral, L.M.; Patrocínio, C.F.V.; Patrocínio, M.L.V.; Mouaffak, F.; Kebir, O.; Macedo, D.; et al. Alpha-lipoic acid in the treatment of psychiatric and neurological disorders: A systematic review. Metab. Brain Dis. 2019, 34, 39–52. [Google Scholar] [CrossRef] [PubMed]
  478. Lyu, M.; Liu, H.; Ye, Y.; Yin, Z. Inhibition effect of thiol-type antioxidants on protein oxidative aggregation caused by free radicals. Biophys. Chem. 2020, 260, 106367. [Google Scholar] [CrossRef] [PubMed]
  479. Zhang, Y.-H.; Wang, D.-W.; Xu, S.-F.; Zhang, S.; Fan, Y.-G.; Yang, Y.-Y.; Guo, S.-Q.; Wang, S.; Guo, T.; Wang, Z.-Y.; et al. α-Lipoic acid improves abnormal behavior by mitigation of oxidative stress, inflammation, ferroptosis, and tauopathy in P301S Tau transgenic mice. Redox Biol. 2018, 14, 535–548. [Google Scholar] [CrossRef]
  480. Memudu, A.E.; Adewumi, A.E. Alpha lipoic acid ameliorates scopolamine induced memory deficit and neurodegeneration in the cerebello-hippocampal cortex. Metab. Brain Dis. 2021, 36, 1729–1745. [Google Scholar] [CrossRef]
  481. Farr, S.A.; Price, T.O.; Banks, W.A.; Ercal, N.; Morley, J.E. Effect of Alpha-Lipoic Acid on Memory, Oxidation, and Lifespan in SAMP8 Mice. J. Alzheimers Dis. 2012, 32, 447–455. [Google Scholar] [CrossRef]
  482. Cho, J.Y.; Um, H.S.; Kang, H.S.; Cho, I.H.; Kim, C.H.; Cho, J.S.; Hwang, D.Y. The combination of exercise training and α-lipoic acid treatment has therapeutic effects on the pathogenic phenotypes of Alzheimer’s disease in NSE/APPsw-transgenic mice. Int. J. Mol. Med. 2010, 25, 337–346. [Google Scholar] [CrossRef]
  483. Quinn, J.F.; Bussiere, J.R.; Hammond, R.S.; Montine, T.J.; Henson, E.; Jones, R.E.; Stackman, R.W. Chronic dietary α-lipoic acid reduces deficits in hippocampal memory of aged Tg2576 mice. Neurobiol. Aging 2007, 28, 213–225. [Google Scholar] [CrossRef]
  484. Sharma, M.; Gupta, Y.K. Effect of alpha lipoic acid on intracerebroventricular streptozotocin model of cognitive impairment in rats. Eur. Neuropsychopharmacol. 2003, 13, 241–247. [Google Scholar] [CrossRef]
  485. Di Stefano, A.; Sozio, P.; Cerasa, L.S.; Iannitelli, A.; Cataldi, A.; Zara, S.; Giorgioni, G.; Nasuti, C. Ibuprofen and Lipoic Acid Diamide as Co-Drug with Neuroprotective Activity: Pharmacological Properties and Effects in β-Amyloid (1–40) Infused Alzheimer’s Disease Rat Model. Int. J. Immunopathol. Pharmacol. 2010, 23, 589–599. [Google Scholar] [CrossRef] [Green Version]
  486. Hager, K.; Kenklies, M.; McAfoose, J.; Engel, J.; Münch, G. α-Lipoic acid as a new treatment option for Alzheimer’s disease—A 48 months follow-up analysis. In Neuropsychiatric Disorders An Integrative Approach; Gerlach, M., Deckert, J., Double, K., Koutsilieri, E., Eds.; Springer: Vienna, Austria, 2007; pp. 189–193. ISBN 978-3-211-73573-2. [Google Scholar]
  487. Shinto, L.; Quinn, J.; Montine, T.; Dodge, H.H.; Woodward, W.; Baldauf-Wagner, S.; Waichunas, D.; Bumgarner, L.; Bourdette, D.; Silbert, L.; et al. A Randomized Placebo-Controlled Pilot Trial of Omega-3 Fatty Acids and Alpha Lipoic Acid in Alzheimer’s Disease. J. Alzheimers Dis. 2013, 38, 111–120. [Google Scholar] [CrossRef] [Green Version]
  488. López-Otín, C.; Blasco, M.A.; Partridge, L.; Serrano, M.; Kroemer, G. Hallmarks of aging: An expanding universe. Cell 2023, S0092867422013770. [Google Scholar] [CrossRef]
  489. Calkins, M.; Manczak, M.; Reddy, P. Mitochondria-Targeted Antioxidant SS31 Prevents Amyloid Beta-Induced Mitochondrial Abnormalities and Synaptic Degeneration in Alzheimer’s Disease. Pharmaceuticals 2012, 5, 1103–1119. [Google Scholar] [CrossRef] [Green Version]
  490. Du, F.; Yu, Q.; Kanaan, N.M.; Yan, S.S. Mitochondrial oxidative stress contributes to the pathological aggregation and accumulation of tau oligomers in Alzheimer’s disease. Hum. Mol. Genet. 2022, 31, 2498–2507. [Google Scholar] [CrossRef]
  491. Lin, M.T.; Beal, M.F. Mitochondrial dysfunction and oxidative stress in neurodegenerative diseases. Nature 2006, 443, 787–795. [Google Scholar] [CrossRef]
  492. Lanzillotta, C.; Di Domenico, F.; Perluigi, M.; Butterfield, D.A. Targeting Mitochondria in Alzheimer Disease: Rationale and Perspectives. CNS Drugs 2019, 33, 957–969. [Google Scholar] [CrossRef]
  493. Reddy, P.H.; Manczak, M.; Yin, X.; Reddy, A.P. Synergistic Protective Effects of Mitochondrial Division Inhibitor 1 and Mitochondria-Targeted Small Peptide SS31 in Alzheimer’s Disease. J. Alzheimers Dis. 2018, 62, 1549–1565. [Google Scholar] [CrossRef]
  494. Soeda, Y.; Saito, M.; Maeda, S.; Ishida, K.; Nakamura, A.; Kojima, S.; Takashima, A. Methylene Blue Inhibits Formation of Tau Fibrils but not of Granular Tau Oligomers: A Plausible Key to Understanding Failure of a Clinical Trial for Alzheimer’s Disease. J. Alzheimers Dis. 2019, 68, 1677–1686. [Google Scholar] [CrossRef]
  495. Young, M.L.; Franklin, J.L. The mitochondria-targeted antioxidant MitoQ inhibits memory loss, neuropathology, and extends lifespan in aged 3xTg-AD mice. Mol. Cell. Neurosci. 2019, 101, 103409. [Google Scholar] [CrossRef]
  496. McManus, M.J.; Murphy, M.P.; Franklin, J.L. The Mitochondria-Targeted Antioxidant MitoQ Prevents Loss of Spatial Memory Retention and Early Neuropathology in a Transgenic Mouse Model of Alzheimer’s Disease. J. Neurosci. 2011, 31, 15703–15715. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  497. Ng, L.F.; Gruber, J.; Cheah, I.K.; Goo, C.K.; Cheong, W.F.; Shui, G.; Sit, K.P.; Wenk, M.R.; Halliwell, B. The mitochondria-targeted antioxidant MitoQ extends lifespan and improves healthspan of a transgenic Caenorhabditis elegans model of Alzheimer disease. Free Radic. Biol. Med. 2014, 71, 390–401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  498. Rudnitskaya, E.A.; Burnyasheva, A.O.; Kozlova, T.A.; Peunov, D.A.; Kolosova, N.G.; Stefanova, N.A. Changes in Glial Support of the Hippocampus during the Development of an Alzheimer’s Disease-like Pathology and Their Correction by Mitochondria-Targeted Antioxidant SkQ1. Int. J. Mol. Sci. 2022, 23, 1134. [Google Scholar] [CrossRef] [PubMed]
  499. Stefanova, N.A.; Muraleva, N.A.; Maksimova, K.Y.; Rudnitskaya, E.A.; Kiseleva, E.; Telegina, D.V.; Kolosova, N.G. An antioxidant specifically targeting mitochondria delays progression of Alzheimer’s disease-like pathology. Aging 2016, 8, 2713–2733. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  500. Muraleva, N.A.; Stefanova, N.A.; Kolosova, N.G. SkQ1 Suppresses the p38 MAPK Signaling Pathway Involved in Alzheimer’s Disease-Like Pathology in OXYS Rats. Antioxidants 2020, 9, 676. [Google Scholar] [CrossRef]
  501. Stefanova, N.A.; Fursova, A.Z.; Kolosova, N.G. Behavioral Effects Induced by Mitochondria-Targeted Antioxidant SkQ1 in Wistar and Senescence-Accelerated OXYS Rats. J. Alzheimers Dis. 2010, 21, 479–491. [Google Scholar] [CrossRef] [Green Version]
  502. Kapay, N.A.; Popova, O.V.; Isaev, N.K.; Stelmashook, E.V.; Kondratenko, R.V.; Zorov, D.B.; Skrebitsky, V.G.; Skulachev, V.P. Mitochondria-Targeted Plastoquinone Antioxidant SkQ1 Prevents Amyloid-β-Induced Impairment of Long-Term Potentiation in Rat Hippocampal Slices. J. Alzheimers Dis. 2013, 36, 377–383. [Google Scholar] [CrossRef]
  503. Attia, H.; Albuhayri, S.; Alaraidh, S.; Alotaibi, A.; Yacoub, H.; Mohamad, R.; Al-Amin, M. Biotin, coenzyme Q10, and their combination ameliorate aluminium chloride-induced Alzheimer’s disease via attenuating neuroinflammation and improving brain insulin signaling. J. Biochem. Mol. Toxicol. 2020, 34, e22519. [Google Scholar] [CrossRef]
  504. Ibrahim Fouad, G. Combination of Omega 3 and Coenzyme Q10 Exerts Neuroprotective Potential Against Hypercholesterolemia-Induced Alzheimer’s-Like Disease in Rats. Neurochem. Res. 2020, 45, 1142–1155. [Google Scholar] [CrossRef]
  505. Wadsworth, T.L.; Bishop, J.A.; Pappu, A.S.; Woltjer, R.L.; Quinn, J.F. Evaluation of Coenzyme Q as an Antioxidant Strategy for Alzheimer’s Disease. J. Alzheimers Dis. 2008, 14, 225–234. [Google Scholar] [CrossRef]
  506. Dumont, M.; Kipiani, K.; Yu, F.; Wille, E.; Katz, M.; Calingasan, N.Y.; Gouras, G.K.; Lin, M.T.; Beal, M.F. Coenzyme Q10 Decreases Amyloid Pathology and Improves Behavior in a Transgenic Mouse Model of Alzheimer’s Disease. J. Alzheimers Dis. 2011, 27, 211–223. [Google Scholar] [CrossRef]
  507. Muthukumaran, K.; Kanwar, A.; Vegh, C.; Marginean, A.; Elliott, A.; Guilbeault, N.; Badour, A.; Sikorska, M.; Cohen, J.; Pandey, S. Ubisol-Q10 (a Nanomicellar Water-Soluble Formulation of CoQ10) Treatment Inhibits Alzheimer-Type Behavioral and Pathological Symptoms in a Double Transgenic Mouse (TgAPEswe, PSEN1dE9) Model of Alzheimer’s Disease. J. Alzheimers Dis. 2017, 61, 221–236. [Google Scholar] [CrossRef]
  508. Lee, H.; Jeong, H.-R.; Park, J.-H.; Hoe, H.-S. Idebenone Decreases Aβ Pathology by Modulating RAGE/Caspase-3 Signaling and the Aβ Degradation Enzyme NEP in a Mouse Model of AD. Biology 2021, 10, 938. [Google Scholar] [CrossRef]
  509. Komaki, H.; Faraji, N.; Komaki, A.; Shahidi, S.; Etaee, F.; Raoufi, S.; Mirzaei, F. Investigation of protective effects of coenzyme Q10 on impaired synaptic plasticity in a male rat model of Alzheimer’s disease. Brain Res. Bull. 2019, 147, 14–21. [Google Scholar] [CrossRef]
  510. Vegh, C.; Pupulin, S.; Wear, D.; Culmone, L.; Huggard, R.; Ma, D.; Pandey, S. Resumption of Autophagy by Ubisol-Q 10 in Presenilin-1 Mutated Fibroblasts and Transgenic AD Mice: Implications for Inhibition of Senescence and Neuroprotection. Oxid. Med. Cell. Longev. 2019, 2019, 7404815. [Google Scholar] [CrossRef] [Green Version]
  511. Yang, X.; Yang, Y.; Li, G.; Wang, J.; Yang, E.S. Coenzyme Q10 Attenuates β-Amyloid Pathology in the Aged Transgenic Mice with Alzheimer Presenilin 1 Mutation. J. Mol. Neurosci. 2008, 34, 165–171. [Google Scholar] [CrossRef]
  512. Singh, A.; Kumar, A. Microglial Inhibitory Mechanism of Coenzyme Q10 Against Aβ (1-42) Induced Cognitive Dysfunctions: Possible Behavioral, Biochemical, Cellular, and Histopathological Alterations. Front. Pharmacol. 2015, 6, 268. [Google Scholar] [CrossRef] [Green Version]
  513. Sheykhhasan, M.; Amini, R.; Soleimani Asl, S.; Saidijam, M.; Hashemi, S.M.; Najafi, R. Neuroprotective effects of coenzyme Q10-loaded exosomes obtained from adipose-derived stem cells in a rat model of Alzheimer’s disease. Biomed. Pharmacother. 2022, 152, 113224. [Google Scholar] [CrossRef]
  514. Yang, X.; Dai, G.; Li, G.; Yang, E.S. Coenzyme Q10 Reduces β-Amyloid Plaque in an APP/PS1 Transgenic Mouse Model of Alzheimer’s Disease. J. Mol. Neurosci. 2010, 41, 110–113. [Google Scholar] [CrossRef]
  515. Hargreaves, I.; Heaton, R.A.; Mantle, D. Disorders of Human Coenzyme Q10 Metabolism: An Overview. Int. J. Mol. Sci. 2020, 21, 6695. [Google Scholar] [CrossRef]
  516. Garrido-Maraver, J.; Cordero, M.D.; Oropesa-Avila, M.; Vega, A.F.; de la Mata, M.; Delgado Pavon, A.; Alcocer-Gomez, E.; Perez Calero, C.; Villanueva Paz, M.; Alanis, M.; et al. Clinical applications of coenzyme Q10. Front. Biosci. 2014, 19, 619. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  517. Isobe, C.; Abe, T.; Terayama, Y. Levels of reduced and oxidized coenzyme Q-10 and 8-hydroxy-2′-deoxyguanosine in the CSF of patients with Alzheimer’s disease demonstrate that mitochondrial oxidative damage and/or oxidative DNA damage contributes to the neurodegenerative process. J. Neurol. 2010, 257, 399–404. [Google Scholar] [CrossRef] [PubMed]
  518. the ActiFE Ulm study group; von Arnim, C.A.F.; Herbolsheimer, F.; Nikolaus, T.; Peter, R.; Biesalski, H.K.; Ludolph, A.C.; Riepe, M.; Nagel, G. Dietary Antioxidants and Dementia in a Population-Based Case-Control Study among Older People in South Germany. J. Alzheimers Dis. 2012, 31, 717–724. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  519. McCarthy, S.; Somayajulu, M.; Sikorska, M.; Borowy-Borowski, H.; Pandey, S. Paraquat induces oxidative stress and neuronal cell death; neuroprotection by water-soluble Coenzyme Q10. Toxicol. Appl. Pharmacol. 2004, 201, 21–31. [Google Scholar] [CrossRef] [PubMed]
  520. Choi, H.; Park, H.-H.; Lee, K.-Y.; Choi, N.-Y.; Yu, H.-J.; Lee, Y.J.; Park, J.; Huh, Y.-M.; Lee, S.-H.; Koh, S.-H. Coenzyme Q10 Restores Amyloid Beta-Inhibited Proliferation of Neural Stem Cells by Activating the PI3K Pathway. Stem Cells Dev. 2013, 22, 2112–2120. [Google Scholar] [CrossRef]
  521. Li, L.; Xu, D.; Lin, J.; Zhang, D.; Wang, G.; Sui, L.; Ding, H.; Du, J. Coenzyme Q10 attenuated β-amyloid 25–35 –induced inflammatory responses in PC12 cells through regulation of the NF–κB signaling pathway. Brain Res. Bull. 2017, 131, 192–198. [Google Scholar] [CrossRef]
  522. Yang, M.; Lian, N.; Yu, Y.; Wang, Y.; Xie, K.; Yu, Y. Coenzyme Q10 alleviates sevoflurane-induced neuroinflammation by regulating the levels of apolipoprotein E and phosphorylated tau protein in mouse hippocampal neurons. Mol. Med. Rep. 2020, 22, 445–453. [Google Scholar] [CrossRef]
  523. Ali, A.A.; Abo El-Ella, D.M.; El-Emam, S.Z.; Shahat, A.S.; El-Sayed, R.M. Physical & mental activities enhance the neuroprotective effect of vinpocetine & coenzyme Q10 combination against Alzheimer & bone remodeling in rats. Life Sci. 2019, 229, 21–35. [Google Scholar] [CrossRef]
  524. Wainwright, L.; Hargreaves, I.P.; Georgian, A.R.; Turner, C.; Dalton, R.N.; Abbott, N.J.; Heales, S.J.R.; Preston, J.E. CoQ10 Deficient Endothelial Cell Culture Model for the Investigation of CoQ10 Blood–Brain Barrier Transport. J. Clin. Med. 2020, 9, 3236. [Google Scholar] [CrossRef]
  525. Wear, D.; Vegh, C.; Sandhu, J.K.; Sikorska, M.; Cohen, J.; Pandey, S. Ubisol-Q10, a Nanomicellar and Water-Dispersible Formulation of Coenzyme-Q10 as a Potential Treatment for Alzheimer’s and Parkinson’s Disease. Antioxidants 2021, 10, 764. [Google Scholar] [CrossRef]
  526. Wani, W.Y.; Gudup, S.; Sunkaria, A.; Bal, A.; Singh, P.P.; Kandimalla, R.J.L.; Sharma, D.R.; Gill, K.D. Protective efficacy of mitochondrial targeted antioxidant MitoQ against dichlorvos induced oxidative stress and cell death in rat brain. Neuropharmacology 2011, 61, 1193–1201. [Google Scholar] [CrossRef]
  527. Xiao, L.; Xu, X.; Zhang, F.; Wang, M.; Xu, Y.; Tang, D.; Wang, J.; Qin, Y.; Liu, Y.; Tang, C.; et al. The mitochondria-targeted antioxidant MitoQ ameliorated tubular injury mediated by mitophagy in diabetic kidney disease via Nrf2/PINK1. Redox Biol. 2017, 11, 297–311. [Google Scholar] [CrossRef]
  528. Manczak, M.; Mao, P.; Calkins, M.J.; Cornea, A.; Reddy, A.P.; Murphy, M.P.; Szeto, H.H.; Park, B.; Reddy, P.H. Mitochondria-Targeted Antioxidants Protect Against Amyloid-β Toxicity in Alzheimer’s Disease Neurons. J. Alzheimers Dis. 2010, 20, S609–S631. [Google Scholar] [CrossRef] [Green Version]
  529. Samluk, L.; Ostapczuk, P.; Dziembowska, M. Long-term mitochondrial stress induces early steps of Tau aggregation by increasing reactive oxygen species levels and affecting cellular proteostasis. Mol. Biol. Cell 2022, 33, ar67. [Google Scholar] [CrossRef]
  530. Dudek, J. Role of Cardiolipin in Mitochondrial Signaling Pathways. Front. Cell Dev. Biol. 2017, 5, 90. [Google Scholar] [CrossRef] [Green Version]
  531. Ma, T.; Hoeffer, C.A.; Wong, H.; Massaad, C.A.; Zhou, P.; Iadecola, C.; Murphy, M.P.; Pautler, R.G.; Klann, E. Amyloid -Induced Impairments in Hippocampal Synaptic Plasticity Are Rescued by Decreasing Mitochondrial Superoxide. J. Neurosci. 2011, 31, 5589–5595. [Google Scholar] [CrossRef] [Green Version]
  532. Skulachev, V.P. Mitochondria-Targeted Antioxidants as Promising Drugs for Treatment of Age-Related Brain Diseases. J. Alzheimers Dis. 2012, 28, 283–289. [Google Scholar] [CrossRef] [Green Version]
  533. Kolosova, N.G.; Tyumentsev, M.A.; Muraleva, N.A.; Kiseleva, E.; Vitovtov, A.O.; Stefanova, N.A. Antioxidant SkQ1 Alleviates Signs of Alzheimer’s Disease-like Pathology in Old OXYS Rats by Reversing Mitochondrial Deterioration. Curr. Alzheimer Res. 2017, 14, 1283–1292. [Google Scholar] [CrossRef]
  534. Muraleva, N.A.; Kozhevnikova, O.S.; Fursova, A.Z.; Kolosova, N.G. Suppression of AMD-Like Pathology by Mitochondria-Targeted Antioxidant SkQ1 Is Associated with a Decrease in the Accumulation of Amyloid β and in mTOR Activity. Antioxidants 2019, 8, 177. [Google Scholar] [CrossRef] [Green Version]
  535. Adlard, P.A.; Bush, A.I. Metals and Alzheimer’s Disease: How Far Have We Come in the Clinic? J. Alzheimers Dis. 2018, 62, 1369–1379. [Google Scholar] [CrossRef] [Green Version]
  536. Kim, G.H.; Kim, J.E.; Rhie, S.J.; Yoon, S. The Role of Oxidative Stress in Neurodegenerative Diseases. Exp. Neurobiol. 2015, 24, 325–340. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  537. Socha, K.; Klimiuk, K.; Naliwajko, S.K.; Soroczyńska, J.; Puścion-Jakubik, A.; Markiewicz-Żukowska, R.; Kochanowicz, J. Dietary Habits, Selenium, Copper, Zinc and Total Antioxidant Status in Serum in Relation to Cognitive Functions of Patients with Alzheimer’s Disease. Nutrients 2021, 13, 287. [Google Scholar] [CrossRef] [PubMed]
  538. Vural, H.; Demirin, H.; Kara, Y.; Eren, I.; Delibas, N. Alterations of plasma magnesium, copper, zinc, iron and selenium concentrations and some related erythrocyte antioxidant enzyme activities in patients with Alzheimer’s disease. J. Trace Elem. Med. Biol. 2010, 24, 169–173. [Google Scholar] [CrossRef] [PubMed]
  539. Cardoso, B.R.; Roberts, B.R.; Malpas, C.B.; Vivash, L.; Genc, S.; Saling, M.M.; Desmond, P.; Steward, C.; Hicks, R.J.; Callahan, J.; et al. Supranutritional Sodium Selenate Supplementation Delivers Selenium to the Central Nervous System: Results from a Randomized Controlled Pilot Trial in Alzheimer’s Disease. Neurotherapeutics 2019, 16, 192–202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  540. Gholamigeravand, B.; Shahidi, S.; Amiri, I.; Samzadeh-kermani, A.; Abbasalipourkabir, R.; Soleimani Asl, S. Administration of Selenium Nanoparticles Reverses Streptozotocin-Induced Neurotoxicity in the male rats. Metab. Brain Dis. 2021, 36, 1259–1266. [Google Scholar] [CrossRef]
  541. Ishrat, T.; Parveen, K.; Khan, M.M.; Khuwaja, G.; Khan, M.B.; Yousuf, S.; Ahmad, A.; Shrivastav, P.; Islam, F. Selenium prevents cognitive decline and oxidative damage in rat model of streptozotocin-induced experimental dementia of Alzheimer’s type. Brain Res. 2009, 1281, 117–127. [Google Scholar] [CrossRef]
  542. Balaban, H.; Nazıroğlu, M.; Demirci, K.; Övey, İ.S. The Protective Role of Selenium on Scopolamine-Induced Memory Impairment, Oxidative Stress, and Apoptosis in Aged Rats: The Involvement of TRPM2 and TRPV1 Channels. Mol. Neurobiol. 2017, 54, 2852–2868. [Google Scholar] [CrossRef]
  543. Cansev, M.; Turkyilmaz, M.; Sijben, J.W.C.; Sevinc, C.; Broersen, L.M.; van Wijk, N. Synaptic Membrane Synthesis in Rats Depends on Dietary Sufficiency of Vitamin C, Vitamin E, and Selenium: Relevance for Alzheimer’s Disease. J. Alzheimers Dis. 2017, 59, 301–311. [Google Scholar] [CrossRef]
  544. Zhang, Z.-H.; Chen, C.; Jia, S.-Z.; Cao, X.-C.; Liu, M.; Tian, J.; Hoffmann, P.R.; Xu, H.-X.; Ni, J.-Z.; Song, G.-L. Selenium Restores Synaptic Deficits by Modulating NMDA Receptors and Selenoprotein K in an Alzheimer’s Disease Model. Antioxid. Redox Signal. 2021, 35, 863–884. [Google Scholar] [CrossRef]
  545. Song, G.; Zhang, Z.; Wen, L.; Chen, C.; Shi, Q.; Zhang, Y.; Ni, J.; Liu, Q. Selenomethionine Ameliorates Cognitive Decline, Reduces Tau Hyperphosphorylation, and Reverses Synaptic Deficit in the Triple Transgenic Mouse Model of Alzheimer’s Disease. J. Alzheimers Dis. 2014, 41, 85–99. [Google Scholar] [CrossRef]
  546. Zhang, Z.-H.; Wu, Q.-Y.; Zheng, R.; Chen, C.; Chen, Y.; Liu, Q.; Hoffmann, P.R.; Ni, J.-Z.; Song, G.-L. Selenomethionine Mitigates Cognitive Decline by Targeting Both Tau Hyperphosphorylation and Autophagic Clearance in an Alzheimer’s Disease Mouse Model. J. Neurosci. 2017, 37, 2449–2462. [Google Scholar] [CrossRef] [Green Version]
  547. Zhang, Z.-H.; Wu, Q.-Y.; Chen, C.; Zheng, R.; Chen, Y.; Ni, J.-Z.; Song, G.-L. Comparison of the effects of selenomethionine and selenium-enriched yeast in the triple-transgenic mouse model of Alzheimer’s disease. Food Funct. 2018, 9, 3965–3973. [Google Scholar] [CrossRef]
  548. Van der Jeugd, A.; Parra-Damas, A.; Baeta-Corral, R.; Soto-Faguás, C.M.; Ahmed, T.; LaFerla, F.M.; Giménez-Llort, L.; D’Hooge, R.; Saura, C.A. Reversal of memory and neuropsychiatric symptoms and reduced tau pathology by selenium in 3xTg-AD mice. Sci. Rep. 2018, 8, 6431. [Google Scholar] [CrossRef] [Green Version]
  549. Iqbal, J.; Zhang, K.; Jin, N.; Zhao, Y.; Liu, Q.; Ni, J.; Shen, L. Selenium positively affects the proteome of 3 × Tg-AD mice cortex by altering the expression of various key proteins: Unveiling the mechanistic role of selenium in AD prevention. J. Neurosci. Res. 2018, 96, 1798–1815. [Google Scholar] [CrossRef]
  550. Abozaid, O.A.R.; Sallam, M.W.; El-Sonbaty, S.; Aziza, S.; Emad, B.; Ahmed, E.S.A. Resveratrol-Selenium Nanoparticles Alleviate Neuroinflammation and Neurotoxicity in a Rat Model of Alzheimer’s Disease by Regulating Sirt1/miRNA-134/GSK3β Expression. Biol. Trace Elem. Res. 2022, 200, 5104–5114. [Google Scholar] [CrossRef]
  551. Cosín-Tomàs, M.; Senserrich, J.; Arumí-Planas, M.; Alquézar, C.; Pallàs, M.; Martín-Requero, A.; Suñol, C.; Kaliman, P.; Sanfeliu, C. Role of Resveratrol and Selenium on Oxidative Stress and Expression of Antioxidant and Anti-Aging Genes in Immortalized Lymphocytes from Alzheimer’s Disease Patients. Nutrients 2019, 11, 1764. [Google Scholar] [CrossRef] [Green Version]
  552. Zhang, Z.H.; Cao, X.-C.; Peng, J.-Y.; Huang, S.-L.; Chen, C.; Jia, S.-Z.; Ni, J.-Z.; Song, G.-L. Reversal of Lipid Metabolism Dysregulation by Selenium and Folic Acid Co-Supplementation to Mitigate Pathology in Alzheimer’s Disease. Antioxidants 2022, 11, 829. [Google Scholar] [CrossRef]
  553. Tamtaji, O.R.; Heidari-soureshjani, R.; Mirhosseini, N.; Kouchaki, E.; Bahmani, F.; Aghadavod, E.; Tajabadi-Ebrahimi, M.; Asemi, Z. Probiotic and selenium co-supplementation, and the effects on clinical, metabolic and genetic status in Alzheimer’s disease: A randomized, double-blind, controlled trial. Clin. Nutr. 2019, 38, 2569–2575. [Google Scholar] [CrossRef]
  554. Cuajungco, M.P.; Goldstein, L.E.; Nunomura, A.; Smith, M.A.; Lim, J.T.; Atwood, C.S.; Huang, X.; Farrag, Y.W.; Perry, G.; Bush, A.I. Evidence that the β-Amyloid Plaques of Alzheimer’s Disease Represent the Redox-silencing and Entombment of Aβ by Zinc. J. Biol. Chem. 2000, 275, 19439–19442. [Google Scholar] [CrossRef] [Green Version]
  555. Donnelly, P.S.; Caragounis, A.; Du, T.; Laughton, K.M.; Volitakis, I.; Cherny, R.A.; Sharples, R.A.; Hill, A.F.; Li, Q.-X.; Masters, C.L.; et al. Selective Intracellular Release of Copper and Zinc Ions from Bis(thiosemicarbazonato) Complexes Reduces Levels of Alzheimer Disease Amyloid-β Peptide. J. Biol. Chem. 2008, 283, 4568–4577. [Google Scholar] [CrossRef] [Green Version]
  556. Zafar, R.; Zubair, M.; Ali, S.; Shahid, K.; Waseem, W.; Naureen, H.; Haider, A.; Jan, M.S.; Ullah, F.; Sirajuddin, M.; et al. Zinc metal carboxylates as potential anti-Alzheimer’s candidate: In vitro anticholinesterase, antioxidant and molecular docking studies. J. Biomol. Struct. Dyn. 2021, 39, 1044–1054. [Google Scholar] [CrossRef] [PubMed]
  557. Vilella, A.; Belletti, D.; Sauer, A.K.; Hagmeyer, S.; Sarowar, T.; Masoni, M.; Stasiak, N.; Mulvihill, J.J.E.; Ruozi, B.; Forni, F.; et al. Reduced plaque size and inflammation in the APP23 mouse model for Alzheimer’s disease after chronic application of polymeric nanoparticles for CNS targeted zinc delivery. J. Trace Elem. Med. Biol. 2018, 49, 210–221. [Google Scholar] [CrossRef] [PubMed]
  558. Singla, N.; Dhawan, D.K. Zinc Improves Cognitive and Neuronal Dysfunction During Aluminium-Induced Neurodegeneration. Mol. Neurobiol. 2017, 54, 406–422. [Google Scholar] [CrossRef] [PubMed]
  559. Harris, C.J.; Voss, K.; Murchison, C.; Ralle, M.; Frahler, K.; Carter, R.; Rhoads, A.; Lind, B.; Robinson, E.; Quinn, J.F. Oral Zinc Reduces Amyloid Burden in Tg2576 Mice. J. Alzheimers Dis. 2014, 41, 179–192. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  560. Brewer, G.J. Alzheimer’s disease causation by copper toxicity and treatment with zinc. Front. Aging Neurosci. 2014, 6, 92. [Google Scholar] [CrossRef]
  561. Farbood, Y.; Sarkaki, A.; Mahdavinia, M.; Ghadiri, A.; Teimoori, A.; Seif, F.; Dehghani, M.A.; Navabi, S.P. Protective Effects of Co-administration of Zinc and Selenium Against Streptozotocin-Induced Alzheimer’s Disease: Behavioral, Mitochondrial Oxidative Stress, and GPR39 Expression Alterations in Rats. Neurotox. Res. 2020, 38, 398–407. [Google Scholar] [CrossRef]
  562. Fu, C.; Dai, L.; Yuan, X.; Xu, Y. Effects of Fish Oil Combined with Selenium and Zinc on Learning and Memory Impairment in Aging Mice and Amyloid Precursor Protein Processing. Biol. Trace Elem. Res. 2021, 199, 1855–1863. [Google Scholar] [CrossRef]
  563. Cherny, R.A.; Atwood, C.S.; Xilinas, M.E.; Gray, D.N.; Jones, W.D.; McLean, C.A.; Barnham, K.J.; Volitakis, I.; Fraser, F.W.; Kim, Y.-S.; et al. Treatment with a Copper-Zinc Chelator Markedly and Rapidly Inhibits β-Amyloid Accumulation in Alzheimer’s Disease Transgenic Mice. Neuron 2001, 30, 665–676. [Google Scholar] [CrossRef]
  564. Tian, Z.-Y.; Wang, C.-Y.; Wang, T.; Li, Y.-C.; Wang, Z.-Y. Glial S100A6 Degrades β-amyloid Aggregation through Targeting Competition with Zinc Ions. Aging Dis. 2019, 10, 756. [Google Scholar] [CrossRef] [Green Version]
  565. Guo, C.; Wang, T.; Zheng, W.; Shan, Z.-Y.; Teng, W.-P.; Wang, Z.-Y. Intranasal deferoxamine reverses iron-induced memory deficits and inhibits amyloidogenic APP processing in a transgenic mouse model of Alzheimer’s disease. Neurobiol. Aging 2013, 34, 562–575. [Google Scholar] [CrossRef]
  566. Zhaba, W.-D.; Deji, Q.-Z.; Gao, S.-Q.; Han, Y.-L.; Gao, C.-C.; Deng, H.-J.; Liu, X.-L.; Li, T.; Zhou, M.-L. Deferoxamine reduces amyloid-beta peptides genesis and alleviates neural apoptosis after traumatic brain injury. NeuroReport 2021, 32, 472–478. [Google Scholar] [CrossRef]
  567. Fawzi, S.F.; Menze, E.T.; Tadros, M.G. Deferiprone ameliorates memory impairment in Scopolamine-treated rats: The impact of its iron-chelating effect on β-amyloid disposition. Behav. Brain Res. 2020, 378, 112314. [Google Scholar] [CrossRef]
  568. Rao, S.S.; Portbury, S.D.; Lago, L.; Bush, A.I.; Adlard, P.A. The Iron Chelator Deferiprone Improves the Phenotype in a Mouse Model of Tauopathy1. J. Alzheimers Dis. 2020, 77, 753–771. [Google Scholar] [CrossRef]
  569. Aalikhani, M.; Safdari, Y.; Jahanshahi, M.; Alikhani, M.; Khalili, M. Comparison Between Hesperidin, Coumarin, and Deferoxamine Iron Chelation and Antioxidant Activity Against Excessive Iron in the Iron Overloaded Mice. Front. Neurosci. 2022, 15, 811080. [Google Scholar] [CrossRef]
  570. Jahanshahi, M.; Khalili, M.; Margedari, A. Naringin Chelates Excessive Iron and Prevents the Formation of Amyloid-Beta Plaques in the Hippocampus of Iron-Overloaded Mice. Front. Pharmacol. 2021, 12, 651156. [Google Scholar] [CrossRef]
  571. Dairam, A.; Fogel, R.; Daya, S.; Limson, J.L. Antioxidant and Iron-Binding Properties of Curcumin, Capsaicin, and S -Allylcysteine Reduce Oxidative Stress in Rat Brain Homogenate. J. Agric. Food Chem. 2008, 56, 3350–3356. [Google Scholar] [CrossRef]
  572. Kupershmidt, L.; Amit, T.; Bar-Am, O.; Youdim, M.B.H.; Weinreb, O. The Novel Multi-Target Iron Chelating-Radical Scavenging Compound M30 Possesses Beneficial Effects on Major Hallmarks of Alzheimer’s Disease. Antioxid. Redox Signal. 2012, 17, 860–877. [Google Scholar] [CrossRef]
  573. Xiao, Y.; Gong, X.; Deng, R.; Liu, W.; Yang, Y.; Wang, X.; Wang, J.; Bao, J.; Shu, X. Iron Chelation Remits Memory Deficits Caused by the High-Fat Diet in a Mouse Model of Alzheimer’s Disease. J. Alzheimers Dis. 2022, 86, 1959–1971. [Google Scholar] [CrossRef]
  574. Wu, Z.; Palanimuthu, D.; Braidy, N.; Salikin, N.H.; Egan, S.; Huang, M.L.H.; Richardson, D.R. Novel multifunctional iron chelators of the aroyl nicotinoyl hydrazone class that markedly enhance cellular NAD+/NADH ratios. Br. J. Pharmacol. 2020, 177, 1967–1987. [Google Scholar] [CrossRef]
  575. Bailey, D.K.; Clark, W.; Kosman, D.J. The iron chelator, PBT434, modulates transcellular iron trafficking in brain microvascular endothelial cells. PLoS ONE 2021, 16, e0254794. [Google Scholar] [CrossRef]
  576. Liu, G.; Men, P.; Harris, P.L.R.; Rolston, R.K.; Perry, G.; Smith, M.A. Nanoparticle iron chelators: A new therapeutic approach in Alzheimer disease and other neurologic disorders associated with trace metal imbalance. Neurosci. Lett. 2006, 406, 189–193. [Google Scholar] [CrossRef] [PubMed]
  577. Shen, X.; Liu, J.; Fujita, Y.; Liu, S.; Maeda, T.; Kikuchi, K.; Obara, T.; Takebe, A.; Sayama, R.; Takahashi, T.; et al. Iron treatment inhibits Aβ42 deposition in vivo and reduces Aβ42/Aβ40 ratio. Biochem. Biophys. Res. Commun. 2019, 512, 653–658. [Google Scholar] [CrossRef] [PubMed]
  578. Ceccom, J.; Coslédan, F.; Halley, H.; Francès, B.; Lassalle, J.M.; Meunier, B. Copper Chelator Induced Efficient Episodic Memory Recovery in a Non-Transgenic Alzheimer’s Mouse Model. PLoS ONE 2012, 7, e43105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  579. Zhao, J.; Shi, Q.; Tian, H.; Li, Y.; Liu, Y.; Xu, Z.; Robert, A.; Liu, Q.; Meunier, B. TDMQ20, a Specific Copper Chelator, Reduces Memory Impairments in Alzheimer’s Disease Mouse Models. ACS Chem. Neurosci. 2021, 12, 140–149. [Google Scholar] [CrossRef] [PubMed]
  580. Maher, P. Modulation of the Neuroprotective and Anti-inflammatory Activities of the Flavonol Fisetin by the Transition Metals Iron and Copper. Antioxidants 2020, 9, 1113. [Google Scholar] [CrossRef]
  581. Harris, C.J.; Gray, N.E.; Caruso, M.; Hunter, M.; Ralle, M.; Quinn, J.F. Copper Modulation and Memory Impairment due to Hippocampal Tau Pathology. J. Alzheimers Dis. 2020, 78, 49–60. [Google Scholar] [CrossRef] [Green Version]
  582. Sestito, S.; Wang, S.; Chen, Q.; Lu, J.; Bertini, S.; Pomelli, C.; Chiellini, G.; He, X.; Pi, R.; Rapposelli, S. Multi-targeted ChEI-copper chelating molecules as neuroprotective agents. Eur. J. Med. Chem. 2019, 174, 216–225. [Google Scholar] [CrossRef]
  583. Hamulakova, S.; Poprac, P.; Jomova, K.; Brezova, V.; Lauro, P.; Drostinova, L.; Jun, D.; Sepsova, V.; Hrabinova, M.; Soukup, O.; et al. Targeting copper(II)-induced oxidative stress and the acetylcholinesterase system in Alzheimer’s disease using multifunctional tacrine-coumarin hybrid molecules. J. Inorg. Biochem. 2016, 161, 52–62. [Google Scholar] [CrossRef]
  584. Solovyev, N.D. Importance of selenium and selenoprotein for brain function: From antioxidant protection to neuronal signalling. J. Inorg. Biochem. 2015, 153, 1–12. [Google Scholar] [CrossRef]
  585. Reddy, V.S.; Bukke, S.; Dutt, N.; Rana, P.; Pandey, A.K. A systematic review and meta-analysis of the circulatory, erythrocellular and CSF selenium levels in Alzheimer’s disease: A metal meta-analysis (AMMA Study-I). J. Trace Elem. Med. Biol. 2017, 42, 68–75. [Google Scholar] [CrossRef]
  586. Chmatalova, Z.; Vyhnalek, M.; Laczo, J.; Hort, J.; Pospisilova, R.; Pechova, M.; Skoumalova, A. Relation of Plasma Selenium and Lipid Peroxidation End Products in Patients with Alzheimer’s Disease. Physiol. Res. 2017, 66, 1049–1056. [Google Scholar] [CrossRef]
  587. Varikasuvu, S.R.; Prasad, V.S.; Kothapalli, J.; Manne, M. Brain Selenium in Alzheimer’s Disease (BRAIN SEAD Study): A Systematic Review and Meta-Analysis. Biol. Trace Elem. Res. 2019, 189, 361–369. [Google Scholar] [CrossRef]
  588. Yan, X.; Liu, K.; Sun, X.; Qin, S.; Wu, M.; Qin, L.; Wang, Y.; Li, Z.; Zhong, X.; Wei, X. A cross-sectional study of blood selenium concentration and cognitive function in elderly Americans: National Health and Nutrition Examination Survey 2011–2014. Ann. Hum. Biol. 2020, 47, 610–619. [Google Scholar] [CrossRef]
  589. Nazıroğlu, M.; Muhamad, S.; Pecze, L. Nanoparticles as potential clinical therapeutic agents in Alzheimer’s disease: Focus on selenium nanoparticles. Expert Rev. Clin. Pharmacol. 2017, 10, 773–782. [Google Scholar] [CrossRef]
  590. Iqbal, J.; Zhang, K.; Jin, N.; Zhao, Y.; Liu, Q.; Ni, J.; Shen, L. Effect of Sodium Selenate on Hippocampal Proteome of 3×Tg-AD Mice—Exploring the Antioxidant Dogma of Selenium against Alzheimer’s Disease. ACS Chem. Neurosci. 2018, 9, 1637–1651. [Google Scholar] [CrossRef]
  591. Tamtaji, O.R.; Taghizadeh, M.; Daneshvar Kakhaki, R.; Kouchaki, E.; Bahmani, F.; Borzabadi, S.; Oryan, S.; Mafi, A.; Asemi, Z. Clinical and metabolic response to probiotic administration in people with Parkinson’s disease: A randomized, double-blind, placebo-controlled trial. Clin. Nutr. 2019, 38, 1031–1035. [Google Scholar] [CrossRef]
  592. Kiyohara, A.C.P.; Torres, D.J.; Hagiwara, A.; Pak, J.; Rueli, R.H.L.H.; Shuttleworth, C.W.R.; Bellinger, F.P. Selenoprotein P Regulates Synaptic Zinc and Reduces Tau Phosphorylation. Front. Nutr. 2021, 8, 683154. [Google Scholar] [CrossRef]
  593. Squitti, R.; Pal, A.; Picozza, M.; Avan, A.; Ventriglia, M.; Rongioletti, M.C.; Hoogenraad, T. Zinc Therapy in Early Alzheimer’s Disease: Safety and Potential Therapeutic Efficacy. Biomolecules 2020, 10, 1164. [Google Scholar] [CrossRef]
  594. Craven, K.M.; Kochen, W.R.; Hernandez, C.M.; Flinn, J.M. Zinc Exacerbates Tau Pathology in a Tau Mouse Model. J. Alzheimers Dis. 2018, 64, 617–630. [Google Scholar] [CrossRef]
  595. Wang, C.-Y.; Wang, T.; Zheng, W.; Zhao, B.-L.; Danscher, G.; Chen, Y.-H.; Wang, Z.-Y. Zinc Overload Enhances APP Cleavage and Aβ Deposition in the Alzheimer Mouse Brain. PLoS ONE 2010, 5, e15349. [Google Scholar] [CrossRef]
  596. Fasae, K.D.; Abolaji, A.O.; Faloye, T.R.; Odunsi, A.Y.; Oyetayo, B.O.; Enya, J.I.; Rotimi, J.A.; Akinyemi, R.O.; Whitworth, A.J.; Aschner, M. Metallobiology and therapeutic chelation of biometals (copper, zinc and iron) in Alzheimer’s disease: Limitations, and current and future perspectives. J. Trace Elem. Med. Biol. 2021, 67, 126779. [Google Scholar] [CrossRef] [PubMed]
  597. Akiyama, H.; Hosokawa, M.; Kametani, F.; Kondo, H.; Chiba, M.; Fukushima, M.; Tabira, T. Long-term oral intake of aluminium or zinc does not accelerate Alzheimer pathology in AβPP and AβPP/tau transgenic mice: Aluminium, zinc and Alzheimer’s disease. Neuropathology 2012, 32, 390–397. [Google Scholar] [CrossRef] [PubMed]
  598. Joppe, K.; Roser, A.-E.; Maass, F.; Lingor, P. The Contribution of Iron to Protein Aggregation Disorders in the Central Nervous System. Front. Neurosci. 2019, 13, 15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  599. Li, D.; Liu, Y.; Zeng, X.; Xiong, Z.; Yao, Y.; Liang, D.; Qu, H.; Xiang, H.; Yang, Z.; Nie, L.; et al. Quantitative Study of the Changes in Cerebral Blood Flow and Iron Deposition During Progression of Alzheimer’s Disease. J. Alzheimers Dis. 2020, 78, 439–452. [Google Scholar] [CrossRef] [PubMed]
  600. Meadowcroft, M.D.; Connor, J.R.; Yang, Q.X. Cortical iron regulation and inflammatory response in Alzheimer’s disease and APPSWE/PS1ΔE9 mice: A histological perspective. Front. Neurosci. 2015, 9, 255. [Google Scholar] [CrossRef] [Green Version]
  601. Zhang, Y.; Bai, X.; Zhang, Y.; Yao, S.; Cui, Y.; You, L.-H.; Yu, P.; Chang, Y.-Z.; Gao, G. Hippocampal Iron Accumulation Impairs Synapses and Memory via Suppressing Furin Expression and Downregulating BDNF Maturation. Mol. Neurobiol. 2022, 59, 5574–5590. [Google Scholar] [CrossRef]
  602. Singh, S.K.; Balendra, V.; Obaid, A.A.; Esposto, J.; Tikhonova, M.A.; Gautam, N.K.; Poeggeler, B. Copper-mediated β-amyloid toxicity and its chelation therapy in Alzheimer’s disease. Metallomics 2022, 14, mfac018. [Google Scholar] [CrossRef]
  603. Zubčić, K.; Hof, P.R.; Šimić, G.; Jazvinšćak Jembrek, M. The Role of Copper in Tau-Related Pathology in Alzheimer’s Disease. Front. Mol. Neurosci. 2020, 13, 572308. [Google Scholar] [CrossRef]
  604. Chen, C.; Jiang, X.; Li, Y.; Yu, H.; Li, S.; Zhang, Z.; Xu, H.; Yang, Y.; Liu, G.; Zhu, F.; et al. Low-dose oral copper treatment changes the hippocampal phosphoproteomic profile and perturbs mitochondrial function in a mouse model of Alzheimer’s disease. Free Radic. Biol. Med. 2019, 135, 144–156. [Google Scholar] [CrossRef]
  605. Redman, L.M.; Smith, S.R.; Burton, J.H.; Martin, C.K.; Il’yasova, D.; Ravussin, E. Metabolic Slowing and Reduced Oxidative Damage with Sustained Caloric Restriction Support the Rate of Living and Oxidative Damage Theories of Aging. Cell Metab. 2018, 27, 805–815.e4. [Google Scholar] [CrossRef] [Green Version]
  606. Mazzotti, D.R.; Guindalini, C.; Moraes, W.A.d.S.; Andersen, M.L.; Cendoroglo, M.S.; Ramos, L.R.; Tufik, S. Human longevity is associated with regular sleep patterns, maintenance of slow wave sleep, and favorable lipid profile. Front. Aging Neurosci. 2014, 6, 134. [Google Scholar] [CrossRef] [Green Version]
  607. Epel, E.S.; Lithgow, G.J. Stress Biology and Aging Mechanisms: Toward Understanding the Deep Connection Between Adaptation to Stress and Longevity. J. Gerontol. A. Biol. Sci. Med. Sci. 2014, 69, S10–S16. [Google Scholar] [CrossRef] [Green Version]
  608. Jenkinson, C.E.; Dickens, A.P.; Jones, K.; Thompson-Coon, J.; Taylor, R.S.; Rogers, M.; Bambra, C.L.; Lang, I.; Richards, S.H. Is volunteering a public health intervention? A systematic review and meta-analysis of the health and survival of volunteers. BMC Public Health 2013, 13, 773. [Google Scholar] [CrossRef] [Green Version]
  609. Arab, L.; Sabbagh, M.N. Are Certain Lifestyle Habits Associated with Lower Alzheimer’s Disease Risk? J. Alzheimers Dis. 2010, 20, 785–794. [Google Scholar] [CrossRef] [Green Version]
  610. Bartolotti, N.; Lazarov, O. Lifestyle and Alzheimer’s Disease. In Genes, Environment and Alzheimer’s Disease; Elsevier: Amsterdam, The Netherlands, 2016; pp. 197–237. ISBN 978-0-12-802851-3. [Google Scholar]
  611. Dhana, K.; Evans, D.A.; Rajan, K.B.; Bennett, D.A.; Morris, M.C. Healthy lifestyle and the risk of Alzheimer dementia: Findings from 2 longitudinal studies. Neurology 2020, 95, e374–e383. [Google Scholar] [CrossRef]
  612. Kivipelto, M.; Mangialasche, F.; Ngandu, T. Lifestyle interventions to prevent cognitive impairment, dementia and Alzheimer disease. Nat. Rev. Neurol. 2018, 14, 653–666. [Google Scholar] [CrossRef]
  613. Ngandu, T.; Lehtisalo, J.; Solomon, A.; Levälahti, E.; Ahtiluoto, S.; Antikainen, R.; Bäckman, L.; Hänninen, T.; Jula, A.; Laatikainen, T.; et al. A 2 year multidomain intervention of diet, exercise, cognitive training, and vascular risk monitoring versus control to prevent cognitive decline in at-risk elderly people (FINGER): A randomised controlled trial. Lancet 2015, 385, 2255–2263. [Google Scholar] [CrossRef]
  614. Ionescu-Tucker, A.; Cotman, C.W. Emerging roles of oxidative stress in brain aging and Alzheimer’s disease. Neurobiol. Aging 2021, 107, 86–95. [Google Scholar] [CrossRef]
  615. Zhao, Y.; Zhao, B. Oxidative Stress and the Pathogenesis of Alzheimer’s Disease. Oxid. Med. Cell. Longev. 2013, 2013, 316523. [Google Scholar] [CrossRef] [Green Version]
  616. Ferruzzi, M.G.; Lobo, J.K.; Janle, E.M.; Cooper, B.; Simon, J.E.; Wu, Q.-L.; Welch, C.; Ho, L.; Weaver, C.; Pasinetti, G.M. Bioavailability of Gallic Acid and Catechins from Grape Seed Polyphenol Extract is Improved by Repeated Dosing in Rats: Implications for Treatment in Alzheimer’s Disease. J. Alzheimers Dis. 2009, 18, 113–124. [Google Scholar] [CrossRef] [Green Version]
  617. Mangoni, A.A.; Jackson, S.H.D. Age-related changes in pharmacokinetics and pharmacodynamics: Basic principles and practical applications: Age-related changes in pharmacokinetics and pharmacodynamics. Br. J. Clin. Pharmacol. 2003, 57, 6–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  618. Erdoğan, M.E.; Aydın, S.; Yanar, K.; Mengi, M.; Kansu, A.D.; Cebe, T.; Belce, A.; Çelikten, M.; Çakatay, U. The effects of lipoic acid on redox status in brain regions and systemic circulation in streptozotocin-induced sporadic Alzheimer’s disease model. Metab. Brain Dis. 2017, 32, 1017–1031. [Google Scholar] [CrossRef] [PubMed]
  619. Sharman, M.J.; Gyengesi, E.; Liang, H.; Chatterjee, P.; Karl, T.; Li, Q.-X.; Wenk, M.R.; Halliwell, B.; Martins, R.N.; Münch, G. Assessment of diets containing curcumin, epigallocatechin-3-gallate, docosahexaenoic acid and α-lipoic acid on amyloid load and inflammation in a male transgenic mouse model of Alzheimer’s disease: Are combinations more effective? Neurobiol. Dis. 2019, 124, 505–519. [Google Scholar] [CrossRef] [PubMed]
  620. Sneideris, T.; Sakalauskas, A.; Sternke-Hoffmann, R.; Peduzzo, A.; Ziaunys, M.; Buell, A.K.; Smirnovas, V. The Environment Is a Key Factor in Determining the Anti-Amyloid Efficacy of EGCG. Biomolecules 2019, 9, 855. [Google Scholar] [CrossRef]
Figure 1. Antioxidant classification. The figure depicts the main classes of non-enzymatic antioxidants with potential diagnostic, preventative, and therapeutic applicability in the context of Alzheimer’s disease: mitochondria-targeted antioxidants, vitamins, carotenoids, organosulfur compounds, flavonoids, minerals, and non-flavonoids. The center depicts an iconic representation of a molecule with antioxidant properties.
Figure 1. Antioxidant classification. The figure depicts the main classes of non-enzymatic antioxidants with potential diagnostic, preventative, and therapeutic applicability in the context of Alzheimer’s disease: mitochondria-targeted antioxidants, vitamins, carotenoids, organosulfur compounds, flavonoids, minerals, and non-flavonoids. The center depicts an iconic representation of a molecule with antioxidant properties.
Antioxidants 12 00180 g001
Table 1. Effects of non-flavonoid treatment in several experimental studies.
Table 1. Effects of non-flavonoid treatment in several experimental studies.
Non-FlavonoidExperimental ModelTreatment DurationTreatment EffectsReference
Ferulic acidIn vivo—PSAPP transgenic mouse model6 monthsAmelioration of behavioral performance
Reduction in Aβ deposits, as well as in Aβ1–42 and Aβ1–40 abundance by the inhibition of BACE1 activity
Attenuation of neuroinflammation (↓GFAP and Iba1 levels, as well as ↓TNFα and IL-1β mRNA expression) and OS (↑SOD1, CAT and GPx1 mRNA expression)
[302]
In vitro—Aβ1–40 -damaged PC12 cells30 minPrevention of cell death
Reduction in intracellular ROS
Inhibition of Aβ1–40 aggregation
[303]
In vivo—Aβ1–40 -induced mouse model21 daysImprovement in cognitive abilities by increasing SOD and ChAT activity and by decreasing AChE activity
Attenuation of lipid peroxidation (↓MDA levels)
Caffeic acidIn vitro—Aβ25–35 -damaged PC12 cells1 hProtection against Aβ-induced toxicity by inhibiting OS, calcium influx, and tau hyperphosphorylation[304]
In vivo—Aβ25–35 -induced mouse model2 weeksImprovement in spatial cognitive and memory functions
Inhibition of lipid peroxidation (↓MDA levels) and NO formation
[305]
In vivo—Aβ1–42 -induced mouse model10 daysDecreasing in neuronal apoptosis (↓caspase 9) and neuroinflammation (↓GFAP and Iba1 expression)
Improvement in learning and memory
Attenuation of OS by inducing the Nrf2/HO-1 signaling pathway
[306]
p-Coumaric acidIn vitro—Aβ25–35 -damaged PC12 cells1 hAttenuation of Aβ25–35 -induced toxicity, through reduction of neuroinflammation (↓iNOS and COX-2), via downregulation of NF-kB and MAPKs pathways[307]
In vivo—D-galactose mouse model42 daysAmelioration of cognitive performance by decreasing AChE levels
Attenuation of OS (↑SOD and GSH) and neuronal apoptosis (↓Caspase3)
Reduction in NF-kB and BACE1 levels
[308]
In vitro—SH-SY5Y cells4 hAttenuation of apoptotic cell death (↓Caspase3)
Reduction in ROS accumulation, as well as in cytochrome c release into the cytosol
[309]
In vivo—Aβ1–42 -induced rat model2 weeksAmelioration of learning and memory deficits and neuronal apoptosis
Gallic acidIn vivo—AlCl3 -induced rat model60 daysAmelioration of spatial memory and learning deficits
Reduction in neurofibrillary tangles and amyloid plaques
Increase in CAT, GSH, and SOD activity, and a decrease in MDA and NO contents
[310]
In vivo—APP/PS1 transgenic mouse model30 daysAmelioration of spatial memory and learning impairments
Inhibition of Aβ aggregation
Reduction in neuroinflammation (↓GFAP) and increase in synaptic strength
[311]
Ellagic acidIn vivo—APP/PS1 transgenic mouse model60 daysImprovement in learning and memory abilities
Decrease in Aβ production by reduction of BACE1 levels
Inhibition of tau phosphorylation by modulation of Akt/GSK-3β signaling pathway
[312]
In vivo—Aβ25–35 -induced rat model1 weekAmelioration of learning and memory deficits
Attenuation of OS, by increasing antioxidant defense (CAT, GSH), and neuroinflammation (↓NF-kB)
Reduction in AChE activity
[313]
In vivo—AlCl3 -induced rat model4 weeksIncrease in SOD and GSH levels
Attenuation of lipid peroxidation
Reduction in neurofibrillary tangles and neuritic plaques
[314]
ResveratrolIn vivo—SAMP8 mouse model7 monthsIncrease in lifespan
Activation of AMPK signaling pathways and increase in SIRT1 levels
Reduction in cognitive impairments as well as in Aβ burden and tau phosphorylation
[315]
In vivo—SAMP8 mouse model15 daysImprovement in learning abilities
Increase in the activity of SOD, GSH-Px, and CAT through the Nrf2/HO-1 signaling pathway
Decrease in MDA content
[316]
In vivo—30 AD patients52 weeksNot statistically significant changes in Aβ1–40 levels in blood and CSF
Reduction in brain volume at 52 weeks
Decrease by 46% in MMP-9 levels in CSF
[317]
CurcuminIn vivo—Transgenic mouse model APPSw6 monthsSuppression of inflammation (↓GFAP, and IL-1β) and oxidative damage
Decrease in insoluble and soluble amyloid as well as in plaque burden
[318]
In vivo—Aβ-induced rat model4 daysImprovement in learning and memory performance
Reduction in OS parameters (ROS formation, lipid peroxidation, and ADP/ATP ratio) as well as in amyloid plaques
[319]
In vivo—48 AD patients24 weeksNot effective on CSF and plasma AD markers, including Aβ1–42, tau, and p-tau[320]
LignansIn vivo—Scopolamine-induced rat model2 weeksImprovement in rat behaviors
Alleviation of OS (↑CAT and SOD) and lipid peroxidation (↓MDA)
Decrease in AChE levels
[321]
In vivo—Aβ1–42 -induced mouse model5 daysImprovement in learning and memory abilities by reduction of ChE total levels and increase of SOD, GSH-Px activity as well as GSH content
Attenuation of memory impairment
[322]
In vivo—Aβ1–42 -induced mouse model4 daysReduction in Aβ1–42 levels by inhibition of β-secretase activity
Inhibition of AChE activity and reduction of GSH levels
[323]
Abbreviations: Aβ: amyloid beta peptide; AChE: acetylcholinesterase; AD: Alzheimer’s disease; ADP/ATP: adenosine diphosphate/adenosine triphosphate; Akt/GSK-3β: protein kinase B/glycogen synthase kinase-3β; AlCl3: aluminum chloride; AMPK: AMP-activated protein kinase; BACE1: beta-site APP cleaving enzyme 1; CAT: catalase; ChAT: choline acetyltransferase; ChE: cholinesterase; COX-2: cyclo-oxygenase-2; CSF: cerebrospinal fluid; GFAP: glial fibrillary acidic protein; GPx1: glutathione Peroxidase 1; GSH: glutathione; GSH-Px: plasma glutathione peroxidase; Iba1: ionized calcium-binding adapter molecule 1; IL-1β: interleukin 1 beta; iNOS: inducible nitric oxide synthase; MAPK: mitogen-activated protein kinase; MDA: malondialdehyde; MMP-9: matrix metallopeptidase 9; NF-kB: nuclear factor kappa B; NO: nitric oxide; Nrf2/HO-1: nuclear factor erythroid 2-related factor/Heme Oxygenase-1; OS: oxidative stress; ROS: reactive oxygen species; SAMP8: senescence-accelerated mouse prone 8; SIRT1: sirtuin 1; SOD: superoxide dismutase; SOD1: superoxide dismutase 1; TNFα: tumor necrosis factor alpha; ↑: increase; ↓: decrease.
Table 2. Preclinical and clinical studies on mitochondria-targeted compounds in AD.
Table 2. Preclinical and clinical studies on mitochondria-targeted compounds in AD.
CompoundExperimental ModelTreatmentResultsRef.
MitoQ3xTg-AD female mice1: 4 mix of 100 μM MitoQ + β-cyclodextrin for 5 months in drinking water↑ memory, lifespan
↓ brain OS, astrogliosis, Aβ accumulation, tau hyperphosphorylation, microglial proliferation, caspase activation
[495,496]
Caenorhabditis elegans overexpressing human Aβ1 µM MitoQ in NGM agar and Escherichia coli OP50-1↑ lifespan, healthspan, electron transport chain function[497]
SkQ1OXYS male rats (12-month-old)250 nM SkQ1/kg daily for 6 months↑ resting/activated microglia ratio, learning, memory, synaptic function, neurotrophic supply, locomotor, and exploratory functions
↓ inflammation, neurodegeneration, neuronal loss, synaptic damage, p38 MAPK signaling, AD progression, tau hyperphosphorylation, Aβ1–42
[498,499,500,501]
Male Wistar ratsOne i.p. injection of 250 nM SkQ1/kg↑ neuroprotection
↓ Aβ-induced OS
[502]
CoQ10AlCl3 treated ratsBiotin (2 mg/kg), CoQ10 (10 mg/kg) for 60 days↑ insulin signaling
↓ inflammation
[503]
Hypercholesterolemic rats10 mg/kg for 30 days (oral)↑ memory, cholinergic function
↓ brain OS and inflammation, amyloidosis
[504]
C65/Bl6 mice10 g/kg for one month↓ brain OS measured by protein carbonyls[505]
Tg19959 mice3 months of 0.4% CoQ10 in chow or 5 months of 2.4% CoQ10 in chow↑ cognitive function (Morris water maze test)
↓ amyloid pathology, brain OS measured by protein carbonyls
[506]
Male Wistar rats50 mg/kg
of CoQ10 daily for 6 weeks (3 before and 3 after AD induction)
↑ EPSP slope and
population spike amplitude
↓ serum malondialdehyde, OS
[509]
Female mice overexpressing presenilin 1-L235P1200 mg/kg of CoQ10 daily for 60 days↑ SOD activity
↓ MDA levels, cortex Aβ burden
[511]
Male Sprague–Dawley rats20 and 40 mg/kg for 21 days↑ SOD, CAT, GSH, mitochondrial respiration
↓ transfer latency, AChE activity, TNFα, LPO, nitrite
[512]
Wistar ratsCoQ10-loaded ADSCs-exosomes↑ Cognition, memory, hippocampal BDNF and SOX2[513]
APP/PS1 transgenic mice1200 mg/kg CoQ10 daily for 60 days↓ Aβ plaque burden[514]
Ubisol-Q10TgAPEswe, PSEN1dE9 mouse6 mg/kg Ubisol-Q10 daily for 18 months↑ long-term and working spatial memory
↓ circulating Aβ, Aβ plaque formation
[507]
Male APP/PS-1 mice200 μg/mL of Ubisol-Q10 in drinking water for 18 months↑ cortical beclin-1 and JNK1, autophagy[510]
Idebenone5xFAD micei.p. injection of 100 mg/kg/day for 14 days↑ NEP, α-secretase ADAM17, tau hyperphosphorylation, total tau
↓ Aβ plaque number, RAGE/caspase 3 signaling
[508]
Abbreviations: 3xTg-AD mice: triple transgenic Alzheimer’s disease mice; AChE: acetylcholinesterase; ADAM17: ADAM metallopeptidase domain 17; ADSCs: adipose-derived stem cells; APP/PS1: amyloid precursor protein/presenilin 1; BDNF: brain-derived neurotrophic factor; CAT: catalase; CoQ10: coenzyme Q10; EPSP: excitatory postsynaptic potential; GSH: glutathione; i.p.: intraperitoneal; JNK1: Jun N-terminal kinase 1; LPO: lipid peroxidation; MDA: malondialdehyde; MitoQ: mitoquinone mesylate; NEP: neprilysin; NGM: nematode growth medium; OS: oxidative stress; OXYS rats: an experimentally renowned model of inbred strains of rats for a range of degenerative diseases in man; RAGE: receptor for advanced glycation end-products; SkQ1: plastoquinonyl-decyltriphenylphosphonium; SOD: superoxide dismutase; SOX2: SRY-Box Transcription Factor 2; TgAPEswe, PSEN1dE9: double transgenic mouse model of Alzheimer’s disease; TNFα: tumor necrosis factor alpha; ↑: increase; ↓: decrease.
Table 3. Preclinical and clinical studies on the therapeutic use of minerals or ion chelators for AD.
Table 3. Preclinical and clinical studies on the therapeutic use of minerals or ion chelators for AD.
Study DesignTreatmentResultsConclusionReference
CSF of AD patients0.32 or 10 mg sodium selenate oral supplementation 3 times daily↑ Se CSF levelsSodium selenate as a possible therapeutic tool against AD[539]
STZ-induced male rats0.4 mg/kg Se nanoparticles oral administration daily for one month↓ ROS
↓ Aβ deposition
Selenium nanoparticles to contrast AD pathogenesis[540]
ICV-STZ rats0.1 mg/kg intraperitoneal sodium selenite for 7 days↓ reduction of GPxSodium selenite as a possible supportive approach to treat SDAT[541]
Hippocampal and dorsal root ganglion neuronal cultures from 1 or 1.5 mg/kg/day scopolamine-treated aged rats1.5 mg/kg intraperitoneal Se supplementation for 14 days↓ membrane permeability to Ca2+
↑ membrane phospholipids
↓ reduction of GPx
Se as a neuroprotective factor[542]
Brain tissue from rats treated with DHA + EPA + uridine (fish oil)1600 mg/kg vitamin C, 1600 mg/kg E and 1.2 mg/kg Se diet for 6 weeks↑ membrane phospholipidsCo-supplementation of Se, vitamins, and fish oil promotes synaptogenesis[543]
Triple transgenic AD mice6 μg/mL selenomethionine supplementation through drinking water for 12 weeks↓ extrasynaptic NMDARs activity
↑ synaptic NMDARs activity
↓ membrane permeability to Ca2+
Selenomethionine improves synaptic plasticity and cognitive functioning[544]
Triple transgenic AD mice6 μg/mL selenomethionine supplementation through drinking water for 12 weeks↓ total tau and phosphorylated tau
↓ synaptic protein loss
Selenomethionine to restore synapses[545]
Triple transgenic AD mice6 μg/mL selenomethionine supplementation through drinking water for 12 weeks↓ total tau and hyperphosphorylated tau
↓ autophagic dysfunction
Selenomethionine to restore synapses[546]
Triple transgenic AD mice6 μg/mL selenomethionine supplementation through drinking water for 12 weeks↓ tau pathologiesSe supplementation, as a potential tool to improve cognitive deficits related to AD[547]
Triple transgenic AD mice12 μg/mL sodium selenate chronic dietary supplementation↓ tau aggregationSodium selenate, as a promising supportive therapy against AD[548]
iTRAQ proteomics technology in hippocampus of triple transgenic AD mice9–12 μg sodium selenate per day in drinking water for 4 months↓ expression of cortical proteins involved in AD pathogenesisSodium selenate, as a potential supportive therapeutic agent for AD[549]
Wistar rats intoxicated with aluminum chloride to mimic AD neurodegeneration100 mg/kg/day oral resveratrol-Se nanoparticles for 60 days↑ antioxidant effect compared to Se alone administrationSe as a promising supplementation against AD when combined with resveratrol[550]
Lymphoblastoid cell lines from AD patientsResveratrol and Se exposure↑ antioxidant effectSe as a protective agent against AD when combined with resveratrol[551]
APP/Tau/PSEN and APP/PS1 transgenic mouse models3 or 1.5 μg/g Se and 36 or 18 μg/g folic acid oral co-supplementation↓ Aβ generation
↓ tau hyperphosphorylation
Se as a potential therapy against AD when combined with folic acid[552]
AD patientsCombined probiotics (Lactobacillus acidophilus, Bifidobacterium bifidum, and Bifidobacterium longum, 2 × 109 CFU/day each) and 200 mg/day selenium oral supplementation for 12 weeks↑ antioxidant effectSe as a potential supportive therapy against AD when combined with probiotics[553]
Primary cortical and human embryonic kidney cells exposed to Aβ1–42Zn2+ incubation↑ antioxidant effect toward H2O2 formationZn as a neuroprotective factor against AD[554]
Chinese hamster ovary cells overexpressing amyloid precursor proteinExposure to Cu or Zn bis (thiosemicarbazonato) therapy↓ monomeric Aβ peptideZn bis (thiosemicarbazonato) as a potential AD supplementation[555]
Acetylcholinesterase enzymes from Electric eelZinc carboxylate derivatives exposure↓ acetylcholinesterase enzyme activityZinc carboxylate derivatives to treat AD[556]
APP23 miceZn nanoparticle injection for 14 days↓ plaques deposition↑ brain Zn levels to counteract AD[557]
Male Sprague Dawley rats receiving aluminum chloride227 mg/L Zinc sulfate in drinking water for 8 weeks↓ tau levels
↓ APP levels
↓ α-synuclein levels
↓ alterations in histological architecture
Zn to reverse the effects of aluminum-induced neurodegeneration, which is correlated with AD[558]
Tg2576 mice treated with Cu2 g/L Zinc acetate in drinking water for 6 months↓ ROS formation
↓ amyloid burden
↓ Cu absorption
Zn to reverse the Cu toxic effects, which are correlated with AD[559]
AD patients150 mg Zn supplementation for 6 months↑ Zn levels
↓ Cu levels
Zn therapy to lower Cu absorption and to restore Zn levels, with the aim to protect from cognitive impairment[560]
Male Wistar rats treated with STZCo-administration of 10 mg/kg Zn and 0.1 mg/kg Se intraperitoneally for 1 week↓ oxidative stress
↓ mitochondrial membranes collapse
↑ GPx
↑ superoxide dismutase
Zn and Se co-administration to improve cognitive functions and prevent the development of AD[561]
Male Kunming miceZn, Se, and fish-oil (EPA + DHA) co-administration for 7 weeks↓ APP cleavageZn, Se and fish-oil co-administration to improve cognitive functions in AD models[562]
APP2576 transgenic miceOral clioquinol (Cu/Zn chelator) administration for 9 weeks↓ Aβ plaquesCu/Zn chelators as a supportive therapeutic strategy[563]
APP/PS1 mouse brain sections300 μg/mL recombinant human S100A6 protein (Zn chelator) incubation for 12 h or culture with human S100A6-expressing cells↓ Aβ plaquesZn sequestration as a supportive therapeutic strategy[564]
APP/PS1 transgenic mice watered with high quantities of iron200 mg/kg intranasal deferoxamine (Fe chelator) once every other day for 3 months↓ Aβ plaquesDeferoxamine as a supplemental treatment for AD[565]
Traumatic brain injury murine modelDeferoxamine (Fe chelator) intraperitoneal treatment↓ Aβ plaques
↓ brain ferritin
Deferoxamine as a potential preventive treatment to avoid neurodegeneration in AD patients[566]
1.14 mg/kg/day scopolamine-treated rats for 7 days5, 10, 20 mg/kg oral deferiprone (Fe chelator) for 14 days↓ Aβ plaquesDeferiprone as a potential preventive treatment in AD patients[567]
rTg(tauP301L)4510 tauopathy murine model100  mg/kg oral deferiprone (Fe chelator) for 16 weeks↓ cognitive deficitDeferiprone as a potential supportive therapy for tauopathies[568]
Blood samples and brain tissues from NMRI male mice treated with 100 mg/kg/day iron dextran for 4 times a week for 6 weeksHesperidin/coumarin/desferal (all Fe chelators) treatment for 4 times a week for 4 weeks↓ Fe levels
↑ antioxidant enzymatic activity
Hesperidin and coumarin to enhance antioxidant enzymatic activity[569]
Brain sections from NMRI male mice following treatment with 100 mg/kg/day iron dextran injections for 4 times a week for 4 weeks30/60 mg/kg/day naringin (Fe chelator) administration for a month↓ Fe levels
↓ Aβ plaques
Naringin as a preventive supportive treatment for AD[570]
Brain homogenates from ratsCurcumin, capsaicin, and S-allylcysteine (Fe chelators) exposure↓ Fe levels
↑ antioxidant effect
Curcumin, capsaicin, and S-allylcysteine as possible tools for the prevention and treatment of AD[571]
APP/PS1 double transgenic AD miceM30 (Fe chelator) oral administration 4 times a week for 9 months↓ Fe levels
↓ APP levels
↓ APP and tau phosphorylation
↓ Aβ plaques
M30 as a potential supportive therapy for AD[572]
AD murine model under a high-fat diet0.5 mg/kg M30 oral administration once every 2 days for 1 month↓ Fe levels
↓ Aβ burden
↓ neuroinflammation
↓ synaptic impairment
High-fat diet as a risk factor for AD and M30 as a potential therapeutic compound[573]
Human cells and Caenorhabditis elegans nematode20 multifunctional synthetic compounds based on the nicotinoyl hydrazone scaffold, in particular, SNH6 (Fe chelator and NAD+ donor)↓ Fe levels
↓ Aβ burden
↓ oxidative stress
↑ sirtuin
nicotinoyl hydrazone-based compounds, especially SNH6 as a promising supportive therapy for AD[574]
Human brain micro-vascular endothelial cellsPBT434 (Fe chelator) exposure↓ Fe reuptake by blood-brain barrier endothelial cellsPBT434, used to prevent Fe-induced cytotoxic effects[575]
Human plasmaChelator–nanoparticles systems↑ blood-brain barrier permeability to Fe chelatorsNanoparticles as a tool to improve chelation treatment for AD[576]
AD murine modelFe-enriched water administration for 8 months↓ Aβ42 burdenFe as a supplemental treatment for AD[577]
Murine AD model by an ICV injection of Aβ1–42 peptideOral administration of 25 mg/kg of bis-8-aminoquinoline PA1637 (Cu chelator) three times per week (8 doses in total)↑ functioning of the episodic memoryPA1637 as a possible supportive treatment for AD[578]
AD mouse modelsOral administration of 10 mg kg−1 TDMQ20 (Cu chelator) in 100 μL of solvent every 2 days for 3 months↑ cognitive and behavioral performanceTDMQ20 as a possible supplemental treatment for AD[579]
Mouse brain cellsFlavonoid fisetin (Cu and Fe chelator) exposure↓ cell deathFisetin as a neuroprotective compound against AD[580]
PS19 transgenic murine modelOral zinc acetate (Cu chelator) treatment↓ spatial memory deficit in female mice, but not in male ones
No significant differences in tau pathology
Cu chelation may improve cognitive symptoms[581]
HT22 cellsMTDLs (Cu chelator) with a rivastigmine skeleton (inhibitor of AChE)↓ AChE and BuChE activity
↓ Cu quantities
MTDLs as promising protective compounds for neurons[582]
In vitro assaysMultifunctional tacrine-7-hydroxycoumarin hybrids chelating Cu and inhibiting AChE and BuChE↓ AChE and BuChE activity
↓ Cu quantities
Multifunctional agents as promising compounds to treat AD[583]
Abbreviations: Aβ: amyloid beta peptide; AChE: acetylcholinesterase; AD: Alzheimer’s disease; APP: amyloid precursor protein; BuChE: butyrylcholinesterase; CSF: cerebrospinal fluid; Cu: copper; DHA: docosahexaenoic acid; EPA: eicosapentaenoic acid; Fe: iron; GPx: glutathione peroxidase; ICV: intracerebroventricular; iTRAQ: isobaric tags for relative and absolute quantitation; MTDLs: multi-target-directed ligands; NMDAR: N-methyl-D-aspartate acid receptors; NMRI mice: female Naval Medical Research Institute (NMRI) outbred mice; PS1: presenilin 1; ROS: reactive oxygen species; SDAT: sporadic dementia of Alzheimer’s type; Se: selenium; STZ: streptozotocin; Zn: zinc; 3D-ASL: three-dimensional arterial spin labeling; ↑: increase; ↓: decrease.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Varesi, A.; Campagnoli, L.I.M.; Carrara, A.; Pola, I.; Floris, E.; Ricevuti, G.; Chirumbolo, S.; Pascale, A. Non-Enzymatic Antioxidants against Alzheimer’s Disease: Prevention, Diagnosis and Therapy. Antioxidants 2023, 12, 180. https://doi.org/10.3390/antiox12010180

AMA Style

Varesi A, Campagnoli LIM, Carrara A, Pola I, Floris E, Ricevuti G, Chirumbolo S, Pascale A. Non-Enzymatic Antioxidants against Alzheimer’s Disease: Prevention, Diagnosis and Therapy. Antioxidants. 2023; 12(1):180. https://doi.org/10.3390/antiox12010180

Chicago/Turabian Style

Varesi, Angelica, Lucrezia Irene Maria Campagnoli, Adelaide Carrara, Ilaria Pola, Elena Floris, Giovanni Ricevuti, Salvatore Chirumbolo, and Alessia Pascale. 2023. "Non-Enzymatic Antioxidants against Alzheimer’s Disease: Prevention, Diagnosis and Therapy" Antioxidants 12, no. 1: 180. https://doi.org/10.3390/antiox12010180

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