Next Article in Journal
Inokosterone from Gentiana rigescens Franch Extends the Longevity of Yeast and Mammalian Cells via Antioxidative Stress and Mitophagy Induction
Next Article in Special Issue
Ginkgo biloba in the Aging Process: A Narrative Review
Previous Article in Journal
Bioavailability Study of Isothiocyanates and Other Bioactive Compounds of Brassica oleracea L. var. Italica Boiled or Steamed: Functional Food or Dietary Supplement?
Previous Article in Special Issue
Antioxidant Properties and Cytoprotective Effect of Pistacia lentiscus L. Seed Oil against 7β-Hydroxycholesterol-Induced Toxicity in C2C12 Myoblasts: Reduction in Oxidative Stress, Mitochondrial and Peroxisomal Dysfunctions and Attenuation of Cell Death
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Naturally Occurring Antioxidant Therapy in Alzheimer’s Disease

Department of Biomedical Sciences and Collaborative Specialization in Neuroscience Program, University of Guelph, Guelph, ON N1G 2W1, Canada
*
Author to whom correspondence should be addressed.
Antioxidants 2022, 11(2), 213; https://doi.org/10.3390/antiox11020213
Submission received: 30 December 2021 / Revised: 17 January 2022 / Accepted: 19 January 2022 / Published: 23 January 2022
(This article belongs to the Special Issue Antioxidants in Age-Related Diseases and Anti-Aging Strategies)

Abstract

:
It is estimated that the prevalence rate of Alzheimer’s disease (AD) will double by the year 2040. Although currently available treatments help with symptom management, they do not prevent, delay the progression of, or cure the disease. Interestingly, a shared characteristic of AD and other neurodegenerative diseases and disorders is oxidative stress. Despite profound evidence supporting the role of oxidative stress in the pathogenesis and progression of AD, none of the currently available treatment options address oxidative stress. Recently, attention has been placed on the use of antioxidants to mitigate the effects of oxidative stress in the central nervous system. In preclinical studies utilizing cellular and animal models, natural antioxidants showed therapeutic promise when administered alone or in combination with other compounds. More recently, the concept of combination antioxidant therapy has been explored as a novel approach to preventing and treating neurodegenerative conditions that present with oxidative stress as a contributing factor. In this review, the relationship between oxidative stress and AD pathology and the neuroprotective role of natural antioxidants from natural sources are discussed. Additionally, the therapeutic potential of natural antioxidants as preventatives and/or treatment for AD is examined, with special attention paid to natural antioxidant combinations and conjugates that are currently being investigated in human clinical trials.

1. Introduction

Mild declines in cognitive and motor abilities are common aspects of human aging. However, the development of neurodegenerative diseases and neurological conditions is not. Interestingly, although the brain is arguably the most essential organ in the human body, it remains susceptible to its own degradation. As a highly metabolically active organ, the brain’s oxygen demand is high [1]. As a result, free radicals are produced as the brain’s requirement for oxygen increases [2]. The brain contains high amounts of polyunsaturated fatty acids, which are quickly oxidized by reactive oxygen species (ROS) but lacks essential enzymes that metabolize several toxic oxygen-containing reactants into harmless compounds [3]. This susceptibility to oxidative damage is observed in several neurodegenerative diseases. Alzheimer’s disease (AD), Huntington’s disease (HD), Parkinson’s disease (PD), amyotrophic lateral sclerosis (ALS) and stroke, among many others (see Table 1), possess this shared pathology of oxidative stress. Recently, considerable attention has been placed on the use of naturally occurring (non-synthetic) antioxidants to mitigate the effects of oxidative stress in the central nervous system (CNS). More recently, the concept of combination antioxidant therapy has been explored as a novel approach to preventing and treating neurodegenerative conditions that present oxidative stress as a contributing factor to the pathogenesis and/or progression of the disease. This review explores the relationship between AD pathology and oxidative stress and the therapeutic potential of natural antioxidants as preventatives and/or treatments for AD, with an emphasis on natural antioxidant combinations and conjugates that are currently being investigated in human clinical trials.

2. Antioxidants and Oxidative Stress

Antioxidants are compounds that protect the body from damage due to oxidative stress. Oxidative stress occurs when there is an imbalance between antioxidants and the production and accumulation of ROS [reviewed in 4]. ROS can be defined as oxygen-containing reactive molecules that are endogenously generated through mitochondrial oxygen metabolism [36]. ROS can also be produced through interplay with exogenous substances such as xenobiotic compounds [36]. This collective term includes compounds such as hydrogen peroxide (H2O2), superoxide (O2•−), hydroxyl radical (OH), nitric oxide (NO) and singlet oxygen (1O2) [37]. During oxidative stress, an insufficient or dysfunctional antioxidant defence system permits damage to important cellular structures such as proteins, lipids, and nucleic acids [37,38], and is implicated in several pathologies of neurodegeneration [38] and aging [39]. Although ROS are damaging in excess, ROS maintain several endogenous functions at low levels. Under normal circumstances, low levels of ROS are produced through ordinary aerobic metabolism and any damage to cells is rapidly repaired through deployment of the antioxidant defence system [40]. ROS perform a critical role in cellular signalling processes, also known as redox signalling [41]. Therefore, to maintain adequate cellular homeostasis, a balance must be established between the production and depletion of ROS. This occurs through the protective mechanisms of antioxidants, which limit the damage induced by ROS and the eventual development of diseases and accelerated aging [39].

2.1. Antioxidant Classification and Mechanisms of Action

Antioxidants can be classified into two main categories: natural antioxidants and synthetic antioxidants. Synthetic antioxidants are artificially generated using a variety of chemical synthesis techniques [40,41]. Natural antioxidants are found in plants and animals and perform various biological roles, including but not limited to anti-inflammatory, anticancer and antiaging effects [42,43,44,45]. Natural antioxidants can be further divided into enzymatic and non-enzymatic antioxidants.

2.1.1. Enzymatic Antioxidants

Enzymatic antioxidants are enzymes produced within the body that possess free radical scavenging abilities and perform antioxidant functions. This group includes primary enzymes such as superoxide dismutase (SOD), catalase (CAT), and glutathione peroxidase (GPx), and secondary enzymes including glutathione reductase (GR) and glucose-6-phosphate dehydrogenase (G6PDH) [46]. SOD eliminates O2•− radicals by catalyzing the reduction of O2•− anions to H2O2 [15]. SOD is present in its many forms in several cellular locations, including the cytosol (Cu-Zn SOD1 or SOD1), mitochondria (Mn SOD or SOD2) and the extracellular space (SOD3) [47]. In turn, CAT is responsible for the decomposition of H2O2 into water and oxygen molecules [48]. In the absence of CAT, H2O2 can react with metal ions to form toxic OH radicals that further perpetuate the effects of oxidative stress [49]. GPx refers to a family of selenocysteine-containing enzymes, including GPx-1, which is predominantly found in the cytoplasm, that utilize glutathione (GSH) as a co-substrate to catalyze the reduction of H2O2 to water and oxidized glutathione (GSSG), and reduce other hydroperoxide substrates to alcohols [50,51]. It was suggested that a cooperative activity between CAT and GPx is required to achieve cellular protection against harmful peroxides [52]. However, differences in the antioxidant capacity of these compounds regarding the rate of removal and the capacity to ravage H2O2 were also identified [53]. GSSG is then reduced to GSH by nicotinamide adenine dinucleotide phosphate (NADPH). G6PDH, which produces NADPH, and GR, which recycles GSSG using NADPH, are considered secondary enzymatic antioxidants (Figure 1). The processes involved in the breakdown and elimination of elevated/toxic oxidative compounds may also include the presence of essential cofactors such as zinc (Zn), copper (Cu), iron (Fe), selenium (Se) and manganese (Mn) [54].

2.1.2. Non-Enzymatic Antioxidants

Non-enzymatic antioxidants can further be divided into groups of vitamins, carotenoids, polyphenols, and minerals [55]. Vitamins are a group of micronutrients that cannot be produced within the body; hence, they require supplementation through the diet [56]. Vitamins perform various functions within the body that are vital for human health and metabolism and are categorized into two groups based on solubility: fat-soluble vitamins and water-soluble vitamins [56]. Fat-soluble vitamins include vitamins A, D, E and K, which are dissolved in fat prior to their absorption into the bloodstream [56]. Water-soluble vitamins include the group of B-complex vitamins and vitamin C, which are dissolved in water [56].

Vitamin A

Sources of vitamin A include dietary supplements, animal products such as fish, meat, poultry, and dairy products, as well as plant products including fruits and vegetables that contain the provitamin A carotenoids (described below) such as beta-carotene, which is endogenously converted to vitamin A [57]. Vitamin A maintains essential roles in vision and synaptic function [58], bone growth and development [59], gene expression [60], cell division [61], reproduction [62], the maintenance of epithelial cells in respiratory, intestinal and urinary tracts, and a healthy immune system [63,64].

Vitamin D

Vitamin D, also known as vitamin D3 (cholecalciferol) and vitamin D2 (ergocalciferol), is predominantly produced endogenously within the skin from the provitamins 7-dehydrocholesterol (7-DHC) and ergosterol [65]. Skin exposure to ultraviolet-B stimulates the synthesis of pre-vitamin D3, followed by thermal isomerization, producing vitamin D3 [65]. Vitamins D3 and D2 can be obtained from the diet in supplements and fortified foods. Dietary vitamin D is predominantly absorbed in the small intestine by chylomicrons, before entering the lymphatic system and then the bloodstream [66]. Once in the bloodstream, from skin or intestinal absorption, vitamin D is converted into 25-hydroxyvitamin D in the liver, which undergoes further conversion to its active form 1,25-dihydrovitamin D in the kidneys [67,68,69]. Both compounds circulate in the blood bound to the vitamin D-binding protein. Once released from vitamin D-binding protein at tissues sites, 1,25-dihydrovitamin D binds to intracellular vitamin D receptors to elicit various metabolic functions throughout the body such as cell differentiation and proliferation, and calcium and phosphorus homeostasis [70,71].

Vitamin E

Vitamin E (α-tocopherol) acts as an antioxidant by protecting membrane components such as polyunsaturated fatty acids from lipid peroxidation by free radicals. Notably, vitamin E has been found to protect low-density lipoproteins from oxidation [72] and is present in high levels within the membranes of red blood cells, mitochondria, and endoplasmic reticulum [73,74,75,76]. Since vitamin E is predominantly synthesized in plants, it can be found in plant products such as nuts, seeds, vegetable oils and leafy green vegetables [77]. Vitamin E maintains various other biological functions, in addition to its role as an antioxidant, including its impact on signal transduction and gene expression [78], and its capacity to regulate enzymatic activity such as protein kinase C, which is important for regulating processes such as cell proliferation and inflammatory responses [79,80,81,82]. The effects of vitamin E may also differ between the four isoforms (α, β, γ and δ), with some studies reporting contradictory effects between these isoforms [82,83]. The ability of vitamin E to donate protons, and thereby saturate and detoxify unpaired electrons on highly reactive radicals such as OH, support its recognition as a potent antioxidant [84]. The importance of vitamin E in brain health is observed in its ability to inhibit the production and progression of chain reactions that lead to lipid peroxidation by hindering the oxidation of unsaturated side chains present within lipid membranes [85]. This role highlights vitamin E as a potential therapeutic agent for neurological conditions, particularly those characterized by oxidative damage. For example, cerebral ischemia and subsequent infarction are consequences of oxidative stress in which vitamin E provides protection [86]. In addition to scavenging free radicals, studies report that vitamin E reduces the toxic effects of NO by converting it to a less toxic nitrite ester in vitro and decreases the production of NO and O2•− within the brain [87,88].

Vitamin K

Vitamin K is bio-actively found in two forms, vitamins K1 and K2. Vitamin K1 (phylloquinone) is predominately found in green leafy plants and can also be found in animals and further converted to vitamin K2 (menaquinone) by anaerobic gut bacteria in animals [89,90]. Vitamin K2 is primarily known for its role in blood-clotting by synthesizing coagulation proteins. It exerts its primary function by creating gamma-carboxyglutamate residues during the production of clotting factors by combining glutamate residues with carboxylic acids [91]. The addition of two carboxylic acid groups to an individual carbon present within a gamma-carboxyglutamate residue permits it to chelate calcium ions. Calcium ion binding is essential for vitamin K-dependent clotting factors, resulting in continued clotting cascades [92]. In a process called the vitamin K cycle, vitamin K is reduced to its metabolic form, vitamin K hydroquinone, by vitamin K epoxide reductase (VKOR), within the cell [93]. In turn, vitamin K hydroquinone is oxidized by vitamin K-dependent carboxylase. This enzyme then carboxylates glutamate residues to gamma-carboxyglutamate residues, ultimately producing vitamin K epoxide [94]. Both carboxylation and epoxidation reactions occur simultaneously. Vitamin K epoxide is then converted to vitamin K by VKOR [95,96,97]. Since vitamin K is continuously recycled within cells, vitamin K deficiency is uncommon in humans. The health benefits of vitamin K extend beyond coagulation to include hepatic functions [95]. More recently, vitamin K’s role in preventing and treating cancer [96] has been explored, as well as its implication in age-related diseases such as osteoporosis and osteoarthrosis, cardiovascular diseases, and neurodegenerative diseases [97,98,99].

B Vitamins

B vitamins constitute a cluster of seven essential water-soluble vitamins; B1 (thiamine), B2 (riboflavin), B3 (niacin), B5 (pantothenic acid), B6 (pyridoxine), B9 (folate) and B12 (cobalamin). B vitamins are primarily produced within the mitochondria, chloroplasts, and the cytosol of plants, and play essential roles in energy production, and the composition and alteration of bioactive compounds through catabolic and anabolic processes, respectively [100]. In a significant portion of enzymatic processes, B vitamins carry out physiological functions by acting as coenzymes. As coenzymes, the biologically active forms of B vitamins tightly bind to the apoenzyme of a protein, producing a holoenzyme that is complete and catalytically active [101]. Through this binding activity, B vitamins play various ubiquitous roles in cellular functioning. The ubiquitous role of B vitamins is demonstrated by vitamin B6. Vitamin B6 functions primarily through its bioactive form, pyridoxal 5′-phosphate. Pyridoxal 5′-phosphate is an important cofactor that influences the functionality of several enzymes that are necessary for the production, degeneration, and conversion of amino acids in all organisms [101]. The essential requirement of B vitamins is also observed with coenzyme A (CoA), the bioactive coenzyme of vitamin B5. CoA is a compulsory co-factor for approximately 4% of mammalian enzymes functioning as a carbonyl-activating group and acyl carrier in numerous biochemical transformation reactions [102]. B vitamins may also act as precursors of metabolic substances, although this is less frequent. CoA also provides a good example of this. CoA can be further acetylated by acetyltransferase enzymes to produce acetyl-CoA, which participates in the biochemical metabolism of proteins, carbohydrates, and lipids, as well as energy production [103]. Notably, B vitamins possess various brain-specific functions. Vitamin B1 plays a critical role in the synthesis of amino acid precursors for neurotransmitters such as acetylcholine (Ach) and was reported to have neuromodulatory functions, including alterations to cholinergic transmission [104]. Vitamin B1 plays structural and functional roles within cell membranes, including in neurons and neuroglia [105] and vitamin B1 deficiency resulted in AD-like abnormalities, such as dysregulation of the cholinergic system, reduced neurotransmitter levels and memory deficits, in preclinical mouse models [106,107,108].

Vitamin C

Vitamin C (ascorbic acid) is predominately found within cells in its redox state, ascorbate [109]. Since humans are unable to endogenously produce vitamin C, the nutrient is obtained from fruit and vegetable sources such as citrus fruits (orange, berries, tomatoes) and leafy green vegetables (broccoli, brussels sprouts). Vitamin C primarily functions as a cofactor for many enzymes, such as hydroxylases, that are implicated in collagen synthesis [110]. More importantly, vitamin C acts as a potent antioxidant through attenuating lipid peroxidation. Lipid peroxidation is a form of radical chain reaction that is initiated by ROS-mediated dissociation of hydrogen atoms from C-H bonds producing lipid radicals [111]. ROS are often entrenched within lipid bilayers [111]. This renders lipids susceptible to the harmful effects of free radicals. Vitamin C can prevent lipid peroxidation by scavenging ROS and working synergistically with other antioxidant compounds, such as vitamin E, to reduce radical formation through the vitamin E redox cycle [112]. Vitamin C is also an exceptional source of electrons. Vitamin C can donate electrons to free radicals that seek out electrons from cellular components such as lipids, proteins, and DNA [111]. By donating an electron to free radicals, vitamin C stabilizes these volatile compounds, reducing their reactivity and the subsequent cellular damage. Vitamin C is constantly recycled within the cell, which increases its antioxidant activity [113]. As a donor of high-energy electrons, vitamin C is oxidized to dehydroascorbic acid. Dehydroascorbic acid can be converted back to vitamin C by dehydroascorbate reductase to be reused or further metabolized, which releases more electrons for ROS stabilization [110,111]. However, the ability of vitamin C to act as a reducing agent for metals such as Cu and Fe heightens the pro-oxidant composition and activity of these metals [114]. Therefore, vitamin C can behave as an antioxidant and a pro-oxidant, and this may depend on its concentration. The pro-oxidant role of vitamin C was initially reported to occur at low concentrations and the antioxidant activities at high concentrations [114]. However, more recent reports contradict these findings and suggest a switch-like behaviour of vitamin C, where it possesses bimodal activity as an antioxidant under normal conditions and switches over to a pro-oxidant at high concentrations and/or under pathophysiological conditions [115,116]. This raises questions regarding the therapeutic use of vitamin C when there is uncertainty of its antioxidant status within the literature [113,114]. Nonetheless, the potential health benefits of vitamin C continue to be explored for their therapeutic effectiveness in several diseases such as cancer, cardiovascular disease, diabetes, immunity, and neurodegenerative disorders [116,117,118,119].
Vitamin C assists in maintaining the function and integrity of various processes within the CNS including antioxidant protection, neuronal development and differentiation, myelination, catecholamine synthesis and regulation of neurotransmission [120]. Studies using animal models have reported detrimental impacts on the brain when vitamin C is decreased, such as enhanced oxidative stress, increased mortality, and the acceleration of amyloid plaque development and aggregation [121,122,123]. Vitamin C deficiency also reduced blood glucose levels and caused oxidative damage to proteins and lipids in the cortex of mice [122,124]. Dopamine and serotonin metabolites in the cortex and striatum of mice and physical strength and locomotor activity were decreased, and treatment with vitamin C was able to restore these deficits [125]. Since the highest concentrations of vitamin C are found in the brain [126] and several neurodegenerative diseases are characterized by oxidative stress, the protective role that vitamin C may play in altering the development and progression of neurological disease such as AD is being investigated.

Carotenoids

Carotenoids are a group of natural pigments that exist ubiquitously across all organisms [127]. Carotenoids perform active roles in photosynthesis in plants and function primarily through photoprotection in non-photosynthetic organisms [127]. In humans, carotenoids, which are found in blood and tissues, are essential precursors of vitamin A. Carotenoids maintain their status as antioxidants through their efficiency in chemically and physically quenching singlet oxygen and scavenging other ROS [127,128]. The structure of carotenoids is the most significant characteristic, contributing to their protective effects. Carotenoids are comprised of several conjugated double bonds, which are essential for photoprotection in all living organisms and light absorption in photosynthetic organisms [128]. Additionally, carotenoids are lipophilic compounds; therefore, they primarily reside within cell membranes.
The most commonly described carotenoids include β-carotene, α-carotene, lutein, lycopene, and zeaxanthin. β-carotene and lycopene are examples of rigid hydrocarbons that are entirely organized within the inner portion of the lipid bilayer [129]. Lutein and zeaxanthin are polar compounds that contain oxygen atoms and exist horizontally to the membrane surface, exposing their hydrophilic segments to the aqueous surroundings [129,130]. It is suggested that the inclusion of carotenoids may impact membrane properties such as permeability, thickness, fluidity, and rigidity, all of which are essential for maintaining membrane integrity [131]. Membrane modifications by carotenoids may enhance resistance to ROS, thereby reducing susceptibility to ROS. Several reports have described the participation of carotenoids in various biological systems and general physiology. These include modulating gap junction communication via intracellular signaling pathways [131] and regulating cell cycle, differentiation, and growth factors [132]. Carotenoids and their metabolites have been implicated as having functions in human health and providing protection in various ROS-induced disorders. These roles include but are not limited to cognitive functions [133,134,135], cancer prevention [131,136], immune stimulation/modulation [137], fertility [138,139] and genomic impacts on transcription and translation [140].

Polyphenols

Polyphenols are naturally occurring compounds found in food sources such as beverages, cereals, fruits, and vegetables and are classified based on chemical structure and resulting activity. The primary classes of polyphenols are phenolic acids, flavonoids, lignans and stilbenes [141]. Phenolic acids are further divided into two classes: hydroxybenzoic acids and hydroxycinnamic acids. Hydroxybenzoic acids are less common than hydroxycinnamic acids and consist of compounds such as gallic and vanillic acid [142]. Hydroxycinnamic acids include caffeic, ferulic, ρ-coumaric and sinapic acids [142]. Flavonoids are the most highly studied cluster of polyphenols. Compounds within this group share a primary structure consisting of two aromatic rings, held by three carbon atoms creating an oxygenated heterocycle [143]. Flavonoids are separated into six subdivisions: anthocyanins, flavan-3-ols, flavonols, flavones, flavanones and isoflavones (Figure 2). Variations within these clusters are derived from differences in the composition and number of hydroxyl groups and their degree of glycosylation and/or alkylation [143]. The most widely studied flavonoids Are catechins, commonly found in green tea, quercetin, found in red wine and foods, and myricetin, which is also commonly found in medicinal plants [144,145,146]. Lignans are referred to as diphenolic compounds, formed by two cinnamic acid residues dimerizing to create a 2,3-dibenxylbutane structure. Many lignans, such as secoisolariciresinol, are regarded as phytoestrogens, possessing antioxidation and antitumor bioactivities [144]. Stilbenes are primarily found in grape skins and berries, and the most prevalent of this subdivision are resveratrol and its derivative pterostilbene [145]. Stilbenes, although low in the human diet, have been implicated for their potential for treating human disease due to their antioxidant and anti-inflammatory activities [145,146]. Other polyphenols include curcumin and gingerol, which have both been reported to provide health benefits and restoration to normal physiology in diseased states [146,147,148]. Overall, there is evidence to support the protective role of polyphenols in multiple disease conditions, such as cancer [149,150,151], cardiovascular disease [152], type 2 diabetes and obesity [149,153], inflammation [154] and neurodegenerative diseases [155,156,157]. Notable polyphenolic compounds studied for their neuroprotective effects include resveratrol, curcumin, quercetin, and epigallocatechin-3-gallate (EGCG) [158]. Recent studies examining cognitive deficits in transgenic AD mice report improvements in AD-like cognitive deficits through anti-amyloidogenic, anti-inflammatory and anti-apoptotic effects of polyphenolic compounds [159,160,161,162].

Minerals

Recently, minerals have been studied for their participation in the antioxidant defence system [163,164,165,166,167,168,169,170,171]. Cu is one of many trace elements that are essential to the biochemistry of active organisms due to its activity as a cofactor and a constituent of metalloenzymes [163]. Cu participates in electron transfer catalysis due to its ability to maintain two oxidation states [164]. Modest amounts of Cu are necessary and beneficial in maintaining metal homeostatic levels; however, the accumulation of redox transition metals such as excess Cu within tissues is cytotoxic [164]. Disturbances in metal homeostasis stimulate the development of oxidative stress and free radical formation, targeting membranes and essential molecules. At high levels, Cu has been implicated in the pathology of neurodegenerative conditions [163]. A study examining the effects of excess Cu showed significantly lower SOD and GSH activity in the brain tissue of Cu-intoxicated rats [163]. Similarly, Fe can result in comparable toxicity due to its ability to donate and accept electrons. Redox-active Fe is a significant contributor to oxidative damage in cellular compartments through the generation of free radicals from ROS by reducing H2O2 to produce OH radicals [165]. This eventually results in damage to several cellular structures, as the OH radical formed by Fe and even Cu can react with H2O2 to stimulate lipid peroxidation [166]. For this reason, Fe is predominantly bound to other molecules for transport and storage, leaving minuscule amounts of redox-active Fe in the labile pool. Even then, Fe does not remain unbound, as it forms complexes with carboxylates, phosphates, and peptides within the labile pool [166].
Mn is an essential element in the synthesis and activation of several enzymes. Its main antioxidant activity occurs through its role as a metalloenzyme via SOD2 to reduce mitochondrial oxidative stress [167]. SOD2 is the principal antioxidant that probes for O2•−, produced within the mitochondria to protect against oxidative stress [167]. SOD2 has also been suggested to provide protection for several disease states such as atherosclerosis, metabolic syndrome, and obesity [167]. Se is another natural trace element. Se is necessary for the composition of selenoproteins, which have been reported to play a critical role in the antioxidant defence system. The activity of GPx, one of the most efficient enzymatic antioxidants, is relatively dependent on Se [168]. Se is inorganically present as selenates, selenides, and selenite, which are more toxic than the organic states selenomethionine and selenocysteine [168]. Se present with GPx has been implicated in the repair of damaged DNA, and the ability to increase GPx activity contributes to the potential benefits of increased Se intake [169].
Zn has been considered an essential metal since its deficiency in humans was first recognized over 50 years ago [170,171]. Zn acts as an antioxidant through multiple mechanisms. Zn can compete with Cu and Fe for binding to proteins and cell membranes, which displaces redox-active Cu and Fe, resulting in the generation of OH from H2O2. Zn can also protect biomolecules such as sulfhydryl groups by binding to them, preventing oxidation [172]. Zn is also capable of enhancing the activation of antioxidant enzymes CAT, SOD and GSH, and diminishing the activity of pro-oxidant enzymes such as inducible NO synthase and NADPH oxide, while blocking the production of lipid peroxidation products [172]. Interestingly, Zn has been reported to upregulate nuclear erythroid 2-related factor 2 (Nrf2) activity, resulting in reduced oxidative stress [173,174]. Nrf2 is a member of the “cap’n’collar” subfamily of basic region leucine zipper transcription factors [175]. Nrf2 regulates the gene expression of many antioxidant and detoxifying enzymes such as SOD and GSH, and glutathione-S-transferase-1 and heme oxygenase-1, respectively. The binding activity of Nrf2 to the antioxidant response element found in the promoter of these target genes leads to the production of enzymes that function as part of the antioxidant defence system [176].

Estrogens

Neuroprotective antioxidants also include compounds such as hormones that are capable of exerting antioxidant effects within the body. Estrogens have been of particular interest as they are well-documented for their neuroprotective roles as steroid hormone antioxidants [177]. The neuroprotective effects of estrogen are proposed to occur through several mechanisms, including estrogen receptor (ER)-dependent and ER-independent actions, and several studies have explored the effects of estrogens in the aging brain and cognition. Estrogens are reportedly involved in learning and memory [178,179] and protect against neurodegenerative conditions such as AD [180,181]. Studies have explored the potential of estrogen replacement therapy (ERT) on improvements in cognitive function. These reports support the postulation of estrogen-mediated enhancements of cognitive function in women across age ranges and defer the onset of AD [182,183,184]. While estrogens have been reported to provide benefits against disease onset and progression, several findings contradict the role of estrogens in neurodegenerative diseases such as AD, reporting minimal differences in cognitive function between placebo and estrogen-treated groups [185,186,187]. These discrepancies leave room for further clarification and investigations into the therapeutic impacts of estrogens by means of ERT. Therefore, although it appears that estrogens may provide protective effects against the onset of AD, further clarification is required to determine whether estrogens are effective once neurodegenerative conditions have already developed.
In summary, antioxidants function through three main mechanisms of action: they (1) act as scavengers to terminate or prevent the production of free radicals; (2) inhibit chain initiation and proliferation reactions; and (3) repair damaged DNA, proteins, and lipid biomolecules. A schematic representation of the classification of natural antioxidants is depicted in Figure 3.

3. Oxidative Stress in Alzheimer’s Disease

The role of oxidative stress in the pathogenesis of AD is well-established in the literature [188,189,190,191,192,193,194,195,196,197,198,199,200,201,202]. To this effect, the oxidative stress hypothesis of AD development postulates potential mechanisms by which oxidative damage causes and/or contributes to the development and progression of AD [203]. This hypothesis is supported by findings from molecular, genetic, and biochemical studies and highlights the detrimental role of ROS in AD onset and progression. Heightened levels of biomarkers of oxidative stress, impairments in the antioxidant defence system, gene mutations and mitochondrial dysfunction have all been implicated [194,196]. As previously mentioned, highly reactive molecules such as ROS target biomolecules such as DNA, proteins, and lipids and, in the instance of AD, mitochondrial dysfunction is proposed to underlie the increase in ROS [191,192,193,194,195,196,197,198,199,200,201,202,203]. Neuronal mitochondria consume high amounts of intracellular oxygen to perform essential functions including energy metabolism, the metabolism of amino acids, fatty acids and lipids, intracellular calcium homeostasis, ROS generation and regulation and more [204]. During mitochondrial respiration, O2•−, a by-product of adenosine triphosphate production, is created. In large amounts, O2•− contributes to oxidative stress by oxidizing cellular targets directly or indirectly by reacting with other molecules and oxidants to form additional ROS and reactive nitrogen species [204]. Mitochondria also produce H2O2, which can further exacerbate oxidative stress by the endogenous conversion reaction of H2O2 to OH by Fe2+ via the Fenton reaction [190,204]. Although the primary generation sites of O2•− are mitochondrial respiratory transport chains I and III, additional cellular sources that could contribute to neuronal oxidative stress include xanthine oxidase, NADPH oxidase and cytochrome P450 enzymes. To prevent a cascade of ROS production, O2•− is neutralized by SOD (Figure 1). Oxidative stress does not exist alone as a potential cause or contributing factor to AD [205,206,207,208,209]. Oxidative stress is reported to contribute to other hypotheses of AD, which implicate the aggregation of intracellular tau, elastin degradation, N-methyl-d-aspartate receptor (NMDAR)-mediated cell stress and abnormal extracellular amyloid accumulation as the primary cause [210,211,212,213,214,215,216,217,218,219,220,221].
The tau hypothesis of AD is widely described in the literature and explains the role of tau-induced neurotoxicity via abnormal hyperphosphorylation of the microtubule-associated protein, tau [222,223,224,225,226,227,228,229,230,231]. Under normal physiological conditions, tau proteins stabilize microtubules within healthy neurons, which maintains neuron morphology and facilitates the transport of enzymes and organelles along the cytoskeleton [224,225,226]. This action is regulated by the level of tau phosphorylation, which primarily depends on the balance between phosphorylation and dephosphorylation, resulting in the activity of various tau kinases and phosphatases, respectively [223,224,225]. Tau hyperphosphorylation reduces the tau microtubule-binding affinity, resulting in the destabilization of microtubules and oligomerization of hyperphosphorylated tau monomers [226,227,228]. In turn, tau oligomers aggregate to form neurofibrillary tangles that induce neurotoxicity and eventual cell death [229,230,231]. The link between oxidative stress and tauopathies is also well-described and attributes tauopathy to oxidative stress-induced aggregate formation that results in the degradation of the microtubule network [232,233,234,235,236].
A recent review by Atlante et al. [237] explains the active and reciprocal relationship between oxidative stress and tau pathology in AD. The researchers report both oxidative stress-induced tau phosphorylation and tau-induced oxidative stress as contributors to the development of AD due to factors, including reductions in cytoplasmic SOD1 and mitochondrial SOD2, which increases the profile of tau phosphorylation and the induction of mitochondrial dysfunction, resulting in H2O2 production by hippocampal tau phosphorylation, respectively [237,238,239,240]. In vitro, the inhibition of glutathione, which triggered mild oxidative stress, increased the levels of phosphorylated tau [234] and the oxidation of fatty acids stimulated tau polymerization [241]. Several reports demonstrate that oxidative stress-induced increases in metal ion redox potential also stimulate the upregulation of tau kinases [242,243,244,245,246,247,248,249,250]. Amyloid-beta (Aβ) is also implicated as a contributor to the cascade of molecular events that result in tau hyperphosphorylation and the inhibition of tau binding to microtubules by promoting glycogen synthase kinase 3 (GSK3) activation [251,252,253,254,255,256,257].
Recently, researchers have identified elastin degradation as a potential contributor to aging, oxidative stress and AD pathology [258,259,260,261,262,263,264,265,266,267]. Elastin is an essential protein that maintains the structural matrix of organs and tissues, including the skin, lungs, cartilage, and blood vessels [258,259,260]. Although elastin is structurally stable, it readily undergoes proteolytic degradation, producing elastin-derived peptides (EDPs) [261,262,263]. Utilizing in vivo and in vitro models, researchers have found that EDPs enhance Aβ formation, which could contribute to subsequent AD development [257,264]. Additionally, EDPs released from elastin due to proteolytic degradation gradually develop into amyloid-like structures [265,266]. Interestingly, like tau, described above, a recent review by Szychowski and Skóra examined the reciprocal relationship between the production of ROS and EDPs [268]. An essential factor in the mechanism of action of EDPs is the peroxisome proliferator-activated receptor gamma (PPARγ) pathway, which is reported to increase the production of ROS by increasing calcium (Ca2+) influx and disrupting the activity and expression of antioxidant enzymes [268,269,270]. PPARγ is reported to enhance SOD, CAT and GPx activity and increase lipid peroxidation [271,272]. Resveratrol, a PPAR agonist, is an example of an antioxidant compound that exerts these neuroprotective properties and was examined for its therapeutic potential in AD [273,274]. Interestingly, in other forms of neurodegeneration, such as ischemia, EDPs are formed in the brain post-injury, which is also when Aβ formation is induced [275]. Some reports have demonstrated EDP-induced increases in ROS in the brain in neuronal stem cells as well as astrocytes [270,272,276]. In turn, EDPs are reported to induce Aβ formation by inducing the overexpression of γ-secretase, which results in excess cleavage activities and the overproduction of Aβ [265]. A few studies suggest that proteoglycans, which are present in the extracellular matrix, may contribute to AD pathogenesis by promoting the fibrilization of Aβ and tau and protecting Aβ from proteolytic degradation [277,278]. The presence and accumulation of EDPs in the brain are also associated with age and correspond with the incidence of AD [275,279]. Increased levels of EDPs have also been detected in CSF patients following a stroke [275,280]. Taken together, these findings suggest the presence of EDPs as potential biomarkers of neurodegenerative disease and that therapies directed at elastin degradation may be useful in the treatment of AD.
The NMDAR hypothesis suggests that excess NMDAR activation leads to the neurodegeneration that occurs in AD [281,282]. Under normal conditions, excitatory neurotransmission by glutamate through the NMDAR is essential for synaptic plasticity and the survival of neurons [281,282]. However, the superfluous activity of the NMDAR induces excitotoxicity causing neuron death, which is a foundational mechanism of the neurodegeneration observed in AD [281,282]. Since the NMDAR mediates Ca2+ regulation and influx [283,284,285,286], dysregulation and overactivity are reported to induce oxidative stress by enhancing the production of ROS within the brain through mechanisms involving EDPs [287,288]. Ca2+ channel blockers such as nifedipine and verapamil are also reported to attenuate ROS production induced by EDP fragments, influence EDP levels in the brain and possibly delay the progression of AD [276,289,290].
Several reports support the involvement of oxidative stress in Aβ toxicity involving metal ions [291,292,293,294,295,296,297,298,299,300,301,302,303,304,305]. Extracellular senile plaques/fibrils comprised of aggregated Aβ peptides exist with metal ions such as Fe, Cu and Zn [293,294,295]. These redox-active metal ions can catalyze the production of ROS when bound to Aβ [296,297,298,299,300,301,302,303,304,305]. Subsequently, newly generated ROS can oxidize both Aβ peptides as well as surrounding biomolecules such as lipids, nucleic acids, and proteins [306,307,308,309,310,311,312,313,314,315]. The oxidation of lipids such as cholesterol within neuronal membranes obstructs membrane integrity [313,314]. In addition, the oxidation of Aβ by ROS and redox-active metal ions impairs its clearance by low-density lipoprotein receptor-related proteins, which could contribute to the accumulation of Aβ in AD [315,316].
Oxidative stress, caused by the production of ROS, creates a favourable environment for Aβ synthesis and accumulation through transcriptional, translational, and epigenetic mechanisms [317,318,319,320,321,322,323,324,325,326,327,328]. Researchers have found that the activation of stress-related signalling pathways stimulates the transcription of amyloid precursor protein (APP) and beta-secretase 1 (BACE1), an essential enzyme for Aβ production [318,319,320]. Additionally, an enhanced protein expression of BACE1 due to ROS such as H2O2 has been reported and is proposed to be regulated by eukaryotic translational initiation factor-2alpha (eIF2α), which was implicated in AD when activated via phosphorylation [321,322]. Several studies have established the role of epigenetic modifications, such as DNA methylation, histone acetylation and chromatin remodelling, in changes to Aβ and AD progression [323,324,325,326,327]. More recently, researchers have demonstrated a link between oxidative stress and epigenetic changes in Aβ production. Gu et al. showed that when neuroblastoma cells were treated with H2O2, a significant decrease in DNA methylation and an increase in histone acetylation occurred [328]. This resulted in increased APP and BACE1 transcription, which was followed by enhanced Aβ production and plaque accumulation [328].
In turn, Aβ exerts its toxic effects through several mechanisms, including oxidative stress. Aβ has been reported to alter mitochondrial function by localizing within the mitochondrial membrane, where it blocks the transport of nuclear-encoded mitochondrial proteins [236,237,238,239,240,241,242,243,244,245,246,247,248,249,250,251,252,253,254,255,256,257,258,259,260,261,262,263,264,265,266,267,268,269,270,271,272,273,274,275,276,277,278,279,280,281,282,283,284,285,286,287,288,289,290,291,292,293,294,295,296,297,298,299,300,301,302,303,304,305,306,307,308,309,310,311,312,313,314,315,316,317,318,319,320,321,322,323,324,325,326,327,328,329]. Additionally, Aβ prevents normal neuronal functions by interacting with mitochondrial proteins, dysregulating the electron transport chain, and stimulating the production of ROS [237,238,239,240,241,242,243,244,245,246,247,248,249,250,251,252,253,254,255,256,257,258,259,260,261,262,263,264,265,266,267,268,269,270,271,272,273,274,275,276,277,278,279,280,281,282,283,284,285,286,287,288,289,290,291,292,293,294,295,296,297,298,299,300,301,302,303,304,305,306,307,308,309,310,311,312,313,314,315,316,317,318,319,320,321,322,323,324,325,326,327,328,329,330,331,332,333,334,335,336,337,338,339]. Additional actions include Aβ-mediated dysregulation of Ca2+ homeostasis, ion leakage through pore formation and depletion of membrane potential [340,341,342]. As a result, this disrupts the cytoskeleton, causes synaptic dysfunction, and stimulates neuronal apoptosis [343]. The examination of human brains from patients diagnosed with AD showed a high degree of membrane damage due to oxidation within the cerebral cortex [344]. Evidence supports the validation of the oxidation of proteins as biomarkers of AD, as enhanced levels of carboxylate proteins have been reported in the hippocampus and parietal cortex [345,346,347].
Considering that ROS production can be a product of tissue injury [348,349,350], it is currently unclear whether oxidative stress is a primary or secondary cause of AD. Even as a secondary cause, oxidative stress perpetuates a detrimental cascade of toxic events that ultimately result in neuron loss. Despite profound evidence supporting the role of oxidative stress in the pathogenesis and progression of AD, none of the currently available treatment options are designed to address oxidative stress. The development of innovative therapies that target the pathological contributors of the disease, such as ROS, could substantially improve the care of patients with AD.

4. Current Treatments for Alzheimer’s Disease

Currently, the only United States Food and Drug Administration (FDA)- and Health Canada-approved medications for AD fall under the classifications of acetylcholinesterase (AChE) inhibitors and NMDAR antagonists [351,352,353]. Donepezil, galantamine and rivastigmine fall under the category of cholinesterase inhibitors. AChE is found predominantly in neuromuscular junctions and synapses of cholinergic neurons in the periphery and CNS, where it rapidly degrades ACh. This neurotransmitter is reported to be involved in learning and memory [354], and the loss of cholinergic neurons projecting from the basal forebrain to the hippocampus and cortex increases as AD progresses [355]. Donepezil and galantamine act by reversibly binding to AChE, which inhibits the hydrolysis (breakdown) of ACh, increasing its levels at synapses throughout the CNS [356]. Rivastigmine also acts to enhance cholinergic communication by binding to and inhibiting AChE, as well as butyrylcholinesterase [357]. These drugs are indicated as long-term symptomatic treatments of AD; however, these drugs lose their efficacy as fewer cholinergic neurons remain in the brain as AD progresses [355]. Donepezil is approved for all stages of AD, while rivastigmine and galantamine are recommended for patients exhibiting mild to moderate symptoms [358]. Memantine, an NMDAR antagonist, acts by blocking the flow of ions through the NMDAR ion channel [359]. Memantine is indicated for moderate to severe AD [360]. Manufactured conjugate (combination) drugs, comprised of donepezil and memantine as extended-release capsules, also exist to alleviate the pill burden of taking multiple medications and increase patient compliance, while mitigating challenges with swallowing that are often associated with AD [361].
Non-pharmacological treatment options include identifying any potential harmful supplements and medications and removing them from the patient’s regimen [362]. First-line treatments for the neuropsychiatric symptoms and behavioural issues associated with the disease include repetitive evaluations, identifying triggers, providing psychoeducation, and modifying both behavioural and environmental interventions [362,363].
The currently available treatments for AD are ineffective in preventing, delaying progression, or curing disease [351]. This necessitates the development of novel disease-modifying therapies that target the pathological hallmarks of the disease, such as tau protein hyperphosphorylation, the development and accumulation of Aβ, inflammation, and oxidative stress [351]. Recently, aducanumab, the only potential disease-modifying therapy, was approved by the FDA through the FDA’s accelerated approval program. Aducanumab is a human monoclonal antibody that significantly reduced the formation and increased the clearance of existing Aβ plaques in mouse models of AD [364,365,366]. However, there is controversy regarding whether the drug slows disease progression in humans, as findings from currently available clinical trial data indicate strategies that reduce amyloid levels do not significantly improve cognition [367]. Biogen, the drug company that created aducanumab, is conducting additional studies to assess the clinical benefit of aducanumab post-approval. If the additional studies fail to show evidence of a clinical benefit, the FDA can withdraw drug approval. Phase 4 clinical trial results for aducanumab are expected to be accessible as early as 2030.
Despite tremendous efforts to find a cure or an effective treatment, AD remains progressive and incurable. Studies utilizing animal models to depict AD show improvements in AD-like phenotypes when utilizing novel therapies such as antioxidants [368,369,370]. Some epidemiological studies also show a reduced risk of AD due to the dietary intake of antioxidants [371,372,373]. Recently, researchers have begun exploring the use of antioxidants in combination with other antioxidant compounds, as well as drugs that are currently being used to treat neurogenerative diseases. This is referred to as combination or conjugate therapy.

5. Conjugate Therapies and the Blood Brain Barrier

The concept of conjugate drug therapy was initially developed as a novel avenue for cancer treatment and has now produced treatment strategies such as antibody–drug conjugates (ADCs). ADCs are designed to target and destroy cancer cells while preserving healthy cells by chemically linking two or more distinct substances [374]. ADCs utilize monoclonal antibodies to deliver cytotoxic agents to antigen-expressing target cells. This approach to treatment has been applied to various types of cancer, such as breast cancer, non-small-cell lung cancer and ovarian cancer, to name a few [375,376,377]. The impact of oxidative stress as a contributing factor to the development of various cancers is well-studied [378,379,380,381,382,383,384], which makes it a primary target in the development of anti-cancer drugs. Interestingly, researchers also utilize the harmful effects of ROS as a tool to target cancer cells [385,386,387,388,389,390]. This includes activating ROS-specific cell death mechanisms such as apoptosis, autophagy, ferroptosis (Fe-dependent) and necrotic cell death in tumour targeted therapy [385,386,387,388,389,390]. An example of the application of ROS in cancer therapy is through targeted tyrosine therapies, which include monoclonal antibodies and small-molecule inhibitors that have been shown to elicit anticancer ROS-mediated effects [391,392,393,394,395,396]. More recently, ADCs have been applied to neurological cancers and neurodegenerative diseases. One example is glioblastoma, an aggressive form of brain cancer that can develop within the brain and spinal cord [397]. However, the efficacy and applications of ADCs within the CNS have been reported to be limited due to the inability of these large drug conjugates to cross the blood–brain barrier (BBB) [398].
When developing novel therapies for neurodegenerative disorders, several factors must be considered for effective drug delivery. One of the most significant is BBB permeability. Unfortunately, the effectiveness of various antioxidants, alone or in conjugated form, are limited by their inability to cross the BBB [399]. The BBB functions as the brain’s endogenous defence system, by excluding non-lipophilic and high-molecular-weight compounds. For a drug or compound to elicit its desired effects, it must first permeate the BBB to reach its drug targets. BBB permeation can occur through several mechanisms, including transmembrane diffusion, saturable transporters, absorption via endocytosis and other extracellular pathways [399]. Several drugs cross the BBB through transmembrane diffusion. This mechanism largely depends on the drug or compound’s ability to cross the cell membrane, which depends on the exogenous compound’s molecular weight, charge, and degree of lipid solubility [400]. Once a drug or compound has diffused through the lipid membranes of the BBB, it will enter the brain’s aqueous environment before reaching its therapeutic target. Therefore, the substance must possess a desirable level of lipid solubility but not be “too lipid soluble”, so that it does not get trapped within the BBB [399]. Saturable transport systems are also an advantageous mechanism of drug delivery and transport across the BBB. Transporters increase the rate of uptake across the BBB compared to what a drug would achieve through transmembrane diffusion alone [401]. However, uptake is limited, as transport occurs via saturable transport systems [401]. The BBB also contains transporters that remove compounds from the brain. These transporters assist with removing toxins from the brain but can also reduce the effectiveness of some therapeutics by increasing their efflux [402]. Under normal conditions, BBB uptake and efflux transporters adapt to meet the needs of the CNS; however, during diseased states, dysregulation can occur. This is observed in AD, for example. Deposition of Aβ damages the BBB and, inversely, reduces Aβ efflux, which contributes to the disease cyclically, as disturbances in BBB function further provoke Aβ deposition [403,404]. BBB dysregulation can be further exacerbated by oxidative stress, either directly or by stimulating the damaging effects of the Aβ peptide (discussed above).
Several antioxidants, including non-traditional antioxidants such as ebselen, have been explored for their ability to cross the BBB and exert neuroprotective roles within the brain when administered alone, in combination, or conjugated with other compounds. Ebselen, a Se-containing compound, has also been assessed for its GPx-like effects [405,406]. As previously mentioned, Se is an essential trace element that maintains antioxidant activity within the brain through oxidative stress resistance [407]. Ebselen has been shown to mitigate the impacts of AD pathology in cell line and primary culture models, as well as triple transgenic AD mouse models. A study conducted by Xie et al. demonstrated the ability of ebselen to inhibit oxidative stress in both cellular and mouse models of AD through enhancing GPx and SOD activity while reducing the activity of p38 mitogen-activated protein kinases [408]. Additionally, ebselen was able to reduce oligomeric Aβ levels within the brains of AD mice by diminishing the expression of APP and BACE-1, both of which are involved in the amyloidogenic pathway of Aβ synthesis [408]. Similar mechanisms have also been reported for other antioxidant compounds that have shown promise in studies utilizing animal subjects to model the onset and course of progression of AD, and the impacts of novel drugs and/or compounds as therapeutic options. Antioxidants that have been explored in combination with other compounds in both cellular and animal models of AD including but not limited to ebselen and donepezil [409], lipoic acid and donepezil [410], ferulic acid and tacrine (the first AChE inhibitor approved for AD, but now discontinued) [411], and polyphenolic hybrids [412,413].
Researchers are currently exploring the neuroprotective roles of antioxidants in humans when these drugs are administered alone, in combination with other antioxidants or drugs but not chemically linked, or in conjugated form with other antioxidants or drugs. To be effective, potential antioxidant drug compounds must be lipid-soluble, small molecule, and/or be chauffeured by other mobilizing non-toxic substances from the bloodstream, through the BBB and into the brain. Instead of utilizing only one antioxidant compound, a combination of antioxidant compounds would increase the overall antioxidant capacity of the drug therapy, heighten the bioavailability to various cellular locations and increase the functionality of antioxidant molecules, such as through facilitating redox cycling [414].

6. Clinical Trials

Various studies have explored the role of compounds with antioxidant activity for the prevention and treatment of cognitive decline and dementia caused by AD. Table 2 summarizes the data from human clinical trials investigating antioxidants in AD. This summary includes results published within the last two decades and those available from on-going clinical trials. Data from clinical trials were collected from the NIH U.S National Library of Medicine site: ClinicalTrials.gov. The inclusion criteria for the clinical trials for this review required that the study (1) includes participants diagnosed with a neurodegenerative disease with evidence of oxidative stress such as AD, (2) utilized at least one natural antioxidant as a form of treatment or preventative therapy, (3) utilized more than one natural antioxidant as combination treatment, and/or (4) utilized a natural antioxidant in combination with a drug currently used to treat AD.

6.1. Vitamins

Studies have reported that vitamins may delay the progression of AD in patients with moderate to severe AD [435,436,437]. Notably, due to the findings from preclinical data supporting the potent effects of vitamin E [438,439,440,441,442], it has been explored as a suitable antioxidant treatment in humans. In addition, vitamin E has been tested in combination with other vitamins and drugs that are currently being used to treat AD for its neuroprotective effects. A randomized, controlled, double-blind study compared the effects of daily administration of 2000 IU vitamin E or 10 mg selegiline, administered alone or in combination, compared to a placebo over 2 years in 341 patients with moderate AD [415]. Selegiline is a selective irreversible monoamine oxidase B (MAO-B) inhibitor that increases the level of dopamine within the synapse by inhibiting dopamine metabolism and is primarily indicated for the treatment of PD [443]. However, earlier trials showed promise for its role in treating AD [444,445], which led researchers to explore the potential benefit of selegiline when combined with other promising compounds such as vitamin E. Based on the primary outcome measures from this study, including the time until death, institutionalization, loss of ability to perform activities of daily living, or severe dementia, this study reported that treatment with both vitamin E and selegiline slowed the progression of disease in patients with moderately severe impairment from AD [444]. The effects of vitamin E have also been examined compared to the AD drug donepezil [416]. However, this study, which included 790 patients with mild cognitive impairment (MCI) and probable AD, showed findings that conflict with other clinical trials involving vitamin E. Patients received 2000 IU vitamin E, 10 mg donepezil or placebo, daily for 3 years. The main findings from this double-blind study showed that vitamin E had no benefit, while donepezil was associated with a lower rate of progression in the first 12 months [416]. In contrast, patients with mild to moderate AD receiving 2000 IU vitamin E, 20 mg memantine or both daily for 5 years showed improvements compared to placebo in another randomized clinical trial [417]. Although a daily dose of 2000 IU vitamin E alone slowed functional decline, there was no difference between the groups receiving memantine alone or with vitamin E [417]. Vitamin E was also tested in combination with other vitamins and minerals [434]. A randomized control trial assessed the changes in cerebrospinal fluid (CSF) biomarkers related to AD and oxidative stress, cognition, and function after antioxidant administration in 75 patients with mild to moderate AD who received 800 IU vitamin E in combination with 500 mg vitamin C, 900 mg ALA and 400 mg coenzyme Q10 3 times/day or placebo daily for 16 weeks. The researchers found that these antioxidants did not influence CSF biomarkers related to amyloid or tau pathology. Although markers of oxidative stress in the brain were reduced, the researchers raised concerns that this antioxidant combination may promote cognitive decline, which would have to be assessed on a long-term basis [418].
B vitamins have also been investigated for their potential protective role in AD in human clinical trials. Although several studies are still ongoing, B vitamins are being explored for their impacts on factors such as changes in phosphorylated tau, brain energy metabolism, oxidative stress, and cognitive function [446,447]. One randomized clinical trial assessed the role of high-dose vitamin B supplementation on homocysteine levels among 340 patients with mild to moderate AD [419]. Elevated homocysteine levels are reported to be a risk factor for dementias such as AD and are attenuated by B vitamin supplementation [448,449,450,451,452]. In this study, patients received a combination of 5 mg folate, 25 mg vitamin B6 and 1 mg vitamin B12 or a placebo, daily for 18 months with the objective of assessing changes in the cognitive subscale of the Alzheimer’s Disease Assessment Scale (ADAS-Cog) [419]. The results indicated that although high-dose B vitamin supplementation reduced homocysteine levels, it did not slow cognitive decline in individuals with mild to moderate AD. The researchers note that several factors could have influenced this negative result. One of them is a difference in the reduction of homocysteine levels observed in participants with milder AD symptoms compared to those with moderate AD, which may indicate a need for further studies that separate the cohorts based on stage, as supplementation may be more beneficial in older individuals with higher homocysteine levels [419]. Additionally, factors such as mental health, diet, supplements, and the activity of the patients should be considered when monitoring or assessing cognitive decline in patients with AD as they may contribute to the worsening of symptoms over time.
Vitamin D has also been explored, notably in combination with memantine [420]. Although this study is ongoing, researchers have set the criteria to assess the effects of vitamin D and memantine in 90 patients with moderate AD. Patients will receive 100,000 IU vitamin D3 (every 4 weeks), in combination with 20 mg memantine or placebo, daily for 24 weeks. The primary objective is to measure changes in cognitive performance measured with the ADAS-Cog and Mini-Mental State Examination (MMSE). Additional measures include changes in functional performance, posture and gait, and comparisons of compliance and tolerance to treatment.
A multivitamin approach was also explored as a treatment for AD [421,453]. In one study including 135 patients with AD or MCI, patients received a multivitamin in the form of a nutraceutical formulation (NF) of 400 ug folic acid, 6 ug vitamin B12, 30 UI vitamin E, 400 mg S-adenosyl methionine, 600 mg N-acetyl cysteine, 500 mg, and acetyl-l-carnitine, daily for 1 year [421]. The primary outcome measures were cognitive improvement or maintenance of cognitive performance, rated by caregivers using the Dementia Rating Scale (DMS), CLOX-1 clock drawing test and the Neuropsychiatric Inventory Questionnaire (NPI-Q). The phase II study reported that patients who received the NF showed improvements compared to the placebo cohort, demonstrating a maintained or improved cognitive performance and mood/behaviour based on the DMS and CLOX-1 measurements. However, no significant improvements were reported in NPI-Q scores. These findings support the conclusions of the phase I study that reported maintenance and/or improvements in cognitive performance and mood/behaviour [453].

6.2. Polyphenols

Polyphenolic compounds are also being explored for their potential as antioxidant treatments for AD. Resveratrol, a potent stilbene antioxidant, has been assessed for its safety, tolerability and efficacy and its role in impacting biomarkers associated with AD. A pilot study involving 39 patients with mild to moderate AD showed that low-dose resveratrol was as safe and well-tolerated as a placebo when administered at 5 mg resveratrol in combination with 5 mg dextrose and 5 mg malate twice daily for 1 year [422]. However, a larger study was necessary to evaluate its beneficial effects. Another study that included 119 patients with mild to moderate AD receiving up to 1 mg resveratrol twice daily or placebo for 52 weeks demonstrated that resveratrol decreased CSF biomarkers, modulated neuro-inflammation, and induced adaptive immunity, which are linked to the progression of AD [423].
Curcumin is another polyphenol that showed promise in preclinical studies. However, these data lack translation in human clinical trials. Although human clinical trials are ongoing, some studies report that curcumin may not be as beneficial for the treatment of AD as some researchers had hoped. One study that included 36 patients with mild to moderate AD who were given 2 g or 4 g curcumin or placebo, daily for 24 weeks showed that although curcumin was well-tolerated, there was no clinical or biochemical evidence of efficacy [424]. In addition, the data suggested the limited bioavailability of curcumin [424]. This is supported by other reports describing fewer adverse events but a concern of elevated serum Aβ in patients receiving either 1 g curcumin or 4 g curcumin in combination with 120 mg ginkgo leaf extract daily for 6 months when compared to a placebo [425]. However, it is possible that this finding was the result of the other compounds consumed in combination with curcumin.
Quercetin and EGCG are also popular antioxidants that are currently under investigation in ongoing trials recruiting patients with early AD and/or patients that are carriers of the ApoE4 allele [426,427,428,429,454]. Carriers of the ApoE4 allele present an increased susceptibility and risk of developing AD [449]. One ongoing study is exploring the role of quercetin on changes in cellular senescence blood markers in patients with early AD [426]. Patients will receive a combination of 1000 mg quercetin and 100 mg dasatinib, a tyrosine receptor inhibitor, or placebo for 2 consecutive days followed by a 13-day +/− no drug period for 12 weeks. Another ongoing study is examining the cognitive impacts of EGCG treatment in patients with early AD [428]. Patients will receive daily treatments of tri-monthly increasing doses of 200 mg, 400 mg, 600 mg, and 800 mg EGCG or placebo for 18 months. Cognitive improvements will be assessed using the ADAS-Cog scale.
Genistein, an isoflavone, has also been investigated both alone and in combination with other compounds for its potential neuroprotective effects in AD [430,431]. Ongoing trials are evaluating genistein-induced changes in Aβ concentrations in the CSF of patients with mild AD [430]. Genistein has also been tested for its effects on cognition in combination with other soy isoflavones such as daidzein [431]. A study of 72 patients with mild AD who received 100 mg of soy isoflavones or placebo daily for 6 months was tested for cognitive outcomes [431]. The 100 mg soy isoflavones combination consisted of genistein and daidzein in equal 50 mg capsules. Cognitive outcomes included language execution function, verbal memory and recall, attention, visual memory, and planning. However, the main findings from this study demonstrated that, after 6 months, the combination treatment of isoflavones genistein and daidzein did not benefit cognition in older men and women with AD [431]. This study was one of the first to examine the function of soy isoflavones in older adults with cognitive decline and AD. The researchers propose that these findings are likely influenced by individual differences in isoflavone metabolism.
ALA has been explored in combination with omega fatty acids [432]. In a pilot trial, 39 patients with mild AD received 600 mg ALA and 3 g fish oil, 3 g fish oil only or placebo daily for 12 months. The results showed that the combination of ALA with omega-3 fatty acids slowed cognitive and functional decline compared to placebo. However, since the size of study participants is relatively small, future studies that include larger sample sizes are essential to determine the neuroprotective benefits of this antioxidant combination as a treatment for AD.

6.3. Minerals

Minerals such as Cu and Se have also been explored for their role in AD and potential to act as a form of therapy. An ongoing trial is evaluating the role of Cu on cognitive function in patients with mild to moderate AD [433]. Patients will receive 8 mg of Cu or placebo daily for 1 year. Changes in cognitive function will be measured using the ADAS-Cog scale. Se has also been studied for its potential role in AD when administered in combination with vitamin E [434]. The Prevention of Alzheimer’s Disease by Vitamin E and Selenium (PREADVISE) trial recruited 7540 men, of which 3786 participated and received 200 μg Se in combination with 400 IU Vitamin E, 200 μg Se + placebo, or 400 IU vitamin E + placebo or placebo + placebo, daily for 7–12 years [434]. The primary outcome was to assess the incidence of dementia (including AD); however, findings showed that neither supplement prevented the development of dementia or the progression from mild cognitive impairment to AD.
Conflicting clinical trial data perpetuate the ongoing disposition on the benefits of antioxidants in AD treatment. Researchers have postulated that these disparities may be due to a variety of factors. Firstly, the equilibrium status between the production of oxidants and the presence of antioxidants is relatively unknown, and this creates a greater challenge when testing human subjects that may present remarkably different profiles of adequacy in endogenous antioxidant defence [455]. Secondly, factors such as the insufficiency of utilizing only one antioxidant compound in a treatment plan should be considered in addition to correcting for the dosages that would provide the most desirable effects specific to the patient [456,457]. Another factor, and arguably the most significant, is BBB permeability. There may be individual differences in BBB permeability; however, results from animal studies indicate that several antioxidants, alone and in combination, can permeate the BBB to some degree [458,459,460,461,462,463]. More recently, novel avenues for drug delivery have emerged to tackle this challenge. These include the utilization of nanoparticles as a strategy to deliver drugs into the CNS [464], as well as synthesizing antioxidant compounds that are chemically linked and developed to meet the criteria for BBB permeation [464,465,466]. Therefore, it is reasonable to conclude that the use of conjugated antioxidants, establishing a reliable profile of biomarkers for each patient and addressing BBB permeation may result in more conclusive findings regarding the effects of antioxidant therapy in AD.

7. Conclusions

AD is currently the leading cause of dementia worldwide, with a prevalence of more than 20 million, which is expected to double by 2040 [286]. Although research has progressed in investigating the etiology and pathogenesis of the disease, much remains unknown about AD. This review describes the well-documented role of oxidative stress in AD; however, the ability of antioxidants to prevent and/or mitigate the impacts of oxidative stress in AD remains uncertain. Compounds such as vitamins, carotenoids, polyphenols, and minerals have shown promise in cellular and animal-based models of AD, prompting their investigation in human clinical trials for their neuroprotective effects, both alone and in combination with other antioxidants or drugs that are currently approved for the treatment of AD. In general, results from previous and ongoing clinical trials remain inconclusive.
Although antioxidants show promise as potential therapies for AD, limitations exist regarding their capacity to treat AD. These limitations include challenges with dosing and determining appropriate timepoints and intervals for intervention, the probability that factors other than oxidative stress may be the predominant cause or propagator of neurodegeneration or that one antioxidant compound may not sufficiently combat oxidative stress to have an impact on disease development or progression. The latter consideration supports the need to explore the use of combination and/or conjugate antioxidant therapy where more than one antioxidant is utilized as a novel approach to treating AD and other neurogenerative conditions that include oxidative stress as a contributing factor. Other considerations for the development of therapies that target ROS-mediated harm in AD, include employing strategies that enhance the activity of molecular targets such as Nrf2 to increase the production of antioxidant enzymes and strengthen the endogenous antioxidant defence system. These approaches will enhance the understanding and application of antioxidant therapies in ROS-mediated neurodegenerative disease.

Author Contributions

Conceptualization, A.E.C.; data curation, A.E.C.; writing—original draft preparation, A.E.C.; writing—review and editing, B.E.K. and T.M.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

The authors would like to thank the Ontario Veterinary College and the Department of Biomedical Sciences at the University of Guelph for supporting this work. A.E.C. is the recipient of an Ontario Veterinary College Graduate Scholarship.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Halliwell, B. Reactive Oxygen Species and the Central Nervous System. J. Neurochem. 1992, 59, 1609–1623. [Google Scholar] [CrossRef] [PubMed]
  2. Shukla, V.; Mishra, S.K.; Pant, H.C. Oxidative Stress in Neurodegeneration. Adv. Pharmacol. Sci. 2011, 2011, 572634. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Perry, G.; Nunomura, A.; Hirai, K.; Zhu, X.; Pérez, M.; Avila, J.; Castellani, R.J.; Atwood, C.S.; Aliev, G.; Sayre, L.M.; et al. Is Oxidative Damage the Fundamental Pathogenic Mechanism of Alzheimer’s and Other Neurodegenerative Diseases? Free Radic. Biol. Med. 2002, 33, 1475–1479. [Google Scholar] [CrossRef]
  4. 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] [PubMed] [Green Version]
  5. Frontiers|Mitochondrial Dysfunction and Oxidative Stress in Alzheimer’s Disease|Aging Neuroscience. Available online: https://www.frontiersin.org/articles/10.3389/fnagi.2021.617588/full (accessed on 28 December 2021).
  6. Barber, S.C.; Shaw, P.J. Oxidative Stress in ALS: Key Role in Motor Neuron Injury and Therapeutic Target. Free Radic. Biol. Med. 2010, 48, 629–641. [Google Scholar] [CrossRef]
  7. Oxidative Stress in ALS: A Mechanism of Neurodegeneration and a Therapeutic Target—ScienceDirect. Available online: https://www.sciencedirect.com/science/article/pii/S0925443906000524 (accessed on 28 December 2021).
  8. Castellani, R.; Smith, M.A.; Richey, P.L.; Kalaria, R.; Gambetti, P.; Perry, G. Evidence for Oxidative Stress in Pick Disease and Corticobasal Degeneration. Brain Res. 1995, 696, 268–271. [Google Scholar] [CrossRef]
  9. Prasad, K.N.; Bondy, S.C. Oxidative and Inflammatory Events in Prion Diseases: Can They Be Therapeutic Targets? Curr. Aging Sci. 2018, 11, 216–225. [Google Scholar] [CrossRef] [Green Version]
  10. Muchová, J.; Žitňanová, I.; Ďuračková, Z. Oxidative Stress and Down Syndrome. Do Antioxidants Play a Role in Therapy? Physiol. Res. 2014, 63, 535–542. [Google Scholar] [CrossRef]
  11. Pop-Busui, R.; Sima, A.; Stevens, M. Diabetic Neuropathy and Oxidative Stress. Diabetes Metab. Res. Rev. 2006, 22, 257–273. [Google Scholar] [CrossRef] [Green Version]
  12. Hosseini, A.; Abdollahi, M. Diabetic Neuropathy and Oxidative Stress: Therapeutic Perspectives. Oxid. Med. Cell. Longev. 2013, 2013, 168039. [Google Scholar] [CrossRef]
  13. Lupoli, F.; Vannocci, T.; Longo, G.; Niccolai, N.; Pastore, A. The Role of Oxidative Stress in Friedreich’s Ataxia. FEBS Lett. 2018, 592, 718–727. [Google Scholar] [CrossRef]
  14. Frontiers|Impaired Redox Signaling in Huntington’s Disease: Therapeutic Implications|Molecular Neuroscience. Available online: https://www.frontiersin.org/articles/10.3389/fnmol.2019.00068/full (accessed on 28 December 2021).
  15. Velusamy, T.; Panneerselvam, A.S.; Purushottam, M.; Anusuyadevi, M.; Pal, P.K.; Jain, S.; Essa, M.M.; Guillemin, G.J.; Kandasamy, M. Protective Effect of Antioxidants on Neuronal Dysfunction and Plasticity in Huntington’s Disease. Oxidative Med. Cell. Longev. 2017, 2017, e3279061. [Google Scholar] [CrossRef]
  16. Túnez, I.; Sánchez-López, F.; Agüera, E.; Fernández-Bolaños, R.; Sánchez, F.M.; Tasset-Cuevas, I. Important Role of Oxidative Stress Biomarkers in Huntington’s Disease. J. Med. Chem. 2011, 54, 5602–5606. [Google Scholar] [CrossRef] [PubMed]
  17. Dalfó, E.; Portero-Otín, M.; Ayala, V.; Martínez, A.; Pamplona, R.; Ferrer, I. Evidence of Oxidative Stress in the Neocortex in Incidental Lewy Body Disease. J. Neuropathol. Exp. Neurol. 2005, 64, 816–830. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Ohl, K.; Tenbrock, K.; Kipp, M. Oxidative Stress in Multiple Sclerosis: Central and Peripheral Mode of Action. Exp. Neurol. 2016, 277, 58–67. [Google Scholar] [CrossRef] [PubMed]
  19. Adamczyk, B.; Adamczyk-Sowa, M. New Insights into the Role of Oxidative Stress Mechanisms in the Pathophysiology and Treatment of Multiple Sclerosis. Oxid. Med. Cell. Longev. 2016, 2016, 1973834. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Fu, R.; Yanjanin, N.M.; Bianconi, S.; Pavan, W.J.; Porter, F.D. Oxidative Stress in Niemann-Pick Disease, Type C. Mol. Genet. Metab. 2010, 101, 214–218. [Google Scholar] [CrossRef] [Green Version]
  21. Vázquez, M.C.; Balboa, E.; Alvarez, A.R.; Zanlungo, S. Oxidative Stress: A Pathogenic Mechanism for Niemann-Pick Type C Disease. Oxidative Med. Cell. Longev. 2012, 2012, e205713. [Google Scholar] [CrossRef] [Green Version]
  22. Pentón-Rol, G.; Cervantes-Llanos, M.; Martínez-Sánchez, G.; Cabrera-Gómez, J.A.; Valenzuela-Silva, C.M.; Ramírez-Nuñez, O.; Casanova-Orta, M.; Robinson-Agramonte, M.A.; Lopategui-Cabezas, I.; López-Saura, P.A. TNF-α and IL-10 Downregulation and Marked Oxidative Stress in Neuromyelitis Optica. J. Inflamm. 2009, 6, 18. [Google Scholar] [CrossRef] [Green Version]
  23. Blesa, J.; Trigo-Damas, I.; Quiroga-Varela, A.; Jackson-Lewis, V.R. Oxidative Stress and Parkinson’s Disease. Front. Neuroanat. 2015, 9, 91. [Google Scholar] [CrossRef] [Green Version]
  24. Wei, Z.; Li, X.; Li, X.; Liu, Q.; Cheng, Y. Oxidative Stress in Parkinson’s Disease: A Systematic Review and Meta-Analysis. Front. Mol. Neurosci. 2018, 11, 236. [Google Scholar] [CrossRef] [PubMed]
  25. Puspita, L.; Chung, S.Y.; Shim, J. Oxidative Stress and Cellular Pathologies in Parkinson’s Disease. Mol. Brain 2017, 10, 53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Aoyama, K.; Matsubara, K.; Kobayashi, S. Aging and Oxidative Stress in Progressive Supranuclear Palsy. Eur. J. Neurol. 2006, 13, 89–92. [Google Scholar] [CrossRef] [PubMed]
  27. Dennis, A.-G.; Almaguer-Mederos, L.E.; Raúl, R.-A.; Roberto, R.-L.; Luis, V.-P.; Dany, C.-A.; Yanetza, G.-Z.; Yaimeé, V.-M.; Annelié, E.-D.; Arnoy, P.-A.; et al. Redox Imbalance Associates with Clinical Worsening in Spinocerebellar Ataxia Type 2. Oxidative Med. Cell. Longev. 2021, 2021, e9875639. [Google Scholar] [CrossRef] [PubMed]
  28. Wang, Y.-C.; Lee, C.-M.; Lee, L.-C.; Tung, L.-C.; Hsieh-Li, H.-M.; Lee-Chen, G.-J.; Su, M.-T. Mitochondrial Dysfunction and Oxidative Stress Contribute to the Pathogenesis of Spinocerebellar Ataxia Type 12 (SCA12). J. Biol. Chem. 2011, 286, 21742–21754. [Google Scholar] [CrossRef] [Green Version]
  29. Torres-Ramos, Y.; Montoya-Estrada, A.; Cisneros, B.; Tercero-Pérez, K.; León-Reyes, G.; Leyva-García, N.; Hernández-Hernández, O.; Magaña, J.J. Oxidative Stress in Spinocerebellar Ataxia Type 7 Is Associated with Disease Severity. Cerebellum 2018, 17, 601–609. [Google Scholar] [CrossRef]
  30. Allen, C.L.; Bayraktutan, U. Oxidative Stress and Its Role in the Pathogenesis of Ischaemic Stroke. Int. J. Stroke. 2009, 4, 461–470. [Google Scholar] [CrossRef]
  31. Žitňanová, I.; Šiarnik, P.; Kollár, B.; Chomová, M.; Pazderová, P.; Andrezálová, L.; Ježovičová, M.; Koňariková, K.; Laubertová, L.; Krivošíková, Z.; et al. Oxidative Stress Markers and Their Dynamic Changes in Patients after Acute Ischemic Stroke. Oxidative Med. Cell. Longev. 2016, 2016, e9761697. [Google Scholar] [CrossRef] [Green Version]
  32. Komsiiska, D. Oxidative Stress and Stroke: A Review of Upstream and Downstream Antioxidant Therapeutic Options. Comp. Clin. Pathol. 2019, 28, 915–926. [Google Scholar] [CrossRef]
  33. Cornelius, C.; Crupi, R.; Calabrese, V.; Graziano, A.; Milone, P.; Pennisi, G.; Radak, Z.; Calabrese, E.J.; Cuzzocrea, S. Traumatic Brain Injury: Oxidative Stress and Neuroprotection. Antioxid. Redox Signal. 2013, 19, 836–853. [Google Scholar] [CrossRef]
  34. Ismail, H.; Shakkour, Z.; Tabet, M.; Abdelhady, S.; Kobaisi, A.; Abedi, R.; Nasrallah, L.; Pintus, G.; Al-Dhaheri, Y.; Mondello, S.; et al. Traumatic Brain Injury: Oxidative Stress and Novel Anti-Oxidants Such as Mitoquinone and Edaravone. Antioxidants 2020, 9, 943. [Google Scholar] [CrossRef] [PubMed]
  35. Mendes Arent, A.; de Souza, L.F.; Walz, R.; Dafre, A.L. Perspectives on Molecular Biomarkers of Oxidative Stress and Antioxidant Strategies in Traumatic Brain Injury. Biomed. Res. Int. 2014, 2014, e723060. [Google Scholar] [CrossRef] [PubMed]
  36. Ray, P.D.; Huang, B.-W.; Tsuji, Y. Reactive Oxygen Species (ROS) Homeostasis and Redox Regulation in Cellular Signaling. Cell. Signal. 2012, 24, 981–990. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Li, R.; Jia, Z.; Trush, M.A. Defining ROS in Biology and Medicine. React. Oxyg. Species 2016, 1, 9–21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Haigis, M.C.; Yankner, B.A. The Aging Stress Response. Mol. Cell 2010, 40, 333–344. [Google Scholar] [CrossRef]
  39. Yu, B.P. Cellular Defenses against Damage from Reactive Oxygen Species. Physiol. Rev. 1994, 74, 139–162. [Google Scholar] [CrossRef]
  40. Khajeh Dangolani, S.; Panahi, F.; Tavaf, Z.; Nourisefat, M.; Yousefi, R.; Khalafi-Nezhad, A. Synthesis and Antioxidant Activity Evaluation of Some Novel Aminocarbonitrile Derivatives Incorporating Carbohydrate Moieties. ACS Omega 2018, 3, 10341–10350. [Google Scholar] [CrossRef]
  41. Mohana, K.N.; Kumar, C.B.P. Synthesis and Antioxidant Activity of 2-Amino-5-Methylthiazol Derivatives Containing 1,3,4-Oxadiazole-2-Thiol Moiety. ISRN Org. Chem. 2013, 2013, e620718. [Google Scholar] [CrossRef]
  42. Li, A.-N.; Li, S.; Zhang, Y.-J.; Xu, X.-R.; Chen, Y.-M.; Li, H.-B. Resources and Biological Activities of Natural Polyphenols. Nutrients 2014, 6, 6020–6047. [Google Scholar] [CrossRef]
  43. Zhou, Y.; Li, Y.; Zhou, T.; Zheng, J.; Li, S.; Li, H.-B. Dietary Natural Products for Prevention and Treatment of Liver Cancer. Nutrients 2016, 8, 156. [Google Scholar] [CrossRef] [Green Version]
  44. Peng, C.; Wang, X.; Chen, J.; Jiao, R.; Wang, L.; Li, Y.M.; Zuo, Y.; Liu, Y.; Lei, L.; Ma, K.Y.; et al. Biology of Ageing and Role of Dietary Antioxidants. Biomed. Res. Int. 2014, 2014, 831841. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Arulselvan, P.; Fard, M.T.; Tan, W.S.; Gothai, S.; Fakurazi, S.; Norhaizan, M.E.; Kumar, S.S. Role of Antioxidants and Natural Products in Inflammation. Oxid. Med. Cell. Longev. 2016, 2016, 5276130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Biotransformation of Waste Biomass into High Value Biochemicals. Available online: https://www.springerprofessional.de/en/biotransformation-of-waste-biomass-into-high-value-biochemicals/1889608 (accessed on 28 December 2021).
  47. Fukai, T.; Ushio-Fukai, M. Superoxide Dismutases: Role in Redox Signaling, Vascular Function, and Diseases. Antioxid. Redox Signal. 2011, 15, 1583–1606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Marín-García, J. Chapter 14—Oxidative Stress and Cell Death in Cardiovascular Disease: A Post-Genomic Appraisal. In Post-Genomic Cardiology, 2nd ed.; Marín-García, J., Ed.; Academic Press: Boston, MA, USA, 2014; pp. 471–498. [Google Scholar] [CrossRef]
  49. van Lith, R.; Ameer, G.A. Chapter Ten—Antioxidant Polymers as Biomaterial. In Oxidative Stress and Biomaterials; Dziubla, T., Butterfield, D.A., Eds.; Academic Press: Cambridge, MA, USA, 2016; pp. 251–296. [Google Scholar] [CrossRef]
  50. Lubos, E.; Loscalzo, J.; Handy, D.E. Glutathione Peroxidase-1 in Health and Disease: From Molecular Mechanisms to Therapeutic Opportunities. Antioxid. Redox Signal. 2011, 15, 1957–1997. [Google Scholar] [CrossRef] [Green Version]
  51. Dalvi, S.M.; Patil, V.W.; Ramraje, N.N. The Roles of Glutathione, Glutathione Peroxidase, Glutathione Reductase and the Carbonyl Protein in Pulmonary and Extra Pulmonary Tuberculosis. J. Clin. Diagn. Res. 2012, 6, 1462–1465. [Google Scholar] [CrossRef]
  52. Baud, O.; Greene, A.E.; Li, J.; Wang, H.; Volpe, J.J.; Rosenberg, P.A. Glutathione Peroxidase-Catalase Cooperativity Is Required for Resistance to Hydrogen Peroxide by Mature Rat Oligodendrocytes. J. Neurosci. 2004, 24, 1531–1540. [Google Scholar] [CrossRef] [Green Version]
  53. Mulholland, C.W.; Elwood, P.C.; Davis, A.; Thurnham, D.I.; Kennedy, O.; Coulter, J.; Fehily, A.; Strain, J.J. Antioxidant Enzymes, Inflammatory Indices and Lifestyle Factors in Older Men: A Cohort Analysis. QJM Int. J. Med. 1999, 92, 579–585. [Google Scholar] [CrossRef] [Green Version]
  54. Nimse, S.B.; Pal, D. Free Radicals, Natural Antioxidants, and Their Reaction Mechanisms. RSC Adv. 2015, 5, 27986–28006. [Google Scholar] [CrossRef] [Green Version]
  55. Shahidi, F.; Zhong, Y. Novel Antioxidants in Food Quality Preservation and Health Promotion. Eur. J. Lipid Sci. Technol. 2010, 112, 930–940. [Google Scholar] [CrossRef]
  56. Reddy, P.; Jialal, I. Biochemistry, Fat Soluble Vitamins. In StatPearls; StatPearls Publishing: Treasure Island, FL, USA, 2021. [Google Scholar]
  57. Micronutrients, I. Vitamin A; National Academies Press: Cambridge, MA, USA, 2001. [Google Scholar]
  58. Dowling, J.E. Vitamin A: Its Many Roles-from Vision and Synaptic Plasticity to Infant Mortality. J. Comp. Physiol. A Neuroethol. Sens. Neural. Behav. Physiol. 2020, 206, 389–399. [Google Scholar] [CrossRef]
  59. Tanumihardjo, S.A. Vitamin A and Bone Health: The Balancing Act. J. Clin. Densitom. 2013, 16, 414–419. [Google Scholar] [CrossRef] [PubMed]
  60. McGrane, M.M. Vitamin A Regulation of Gene Expression: Molecular Mechanism of a Prototype Gene. J. Nutr. Biochem. 2007, 18, 497–508. [Google Scholar] [CrossRef] [PubMed]
  61. Pozniakov, S.P. Mechanism of action of vitamin A on cell differentiation and function. Ontogenez 1986, 17, 578–586. [Google Scholar] [PubMed]
  62. Clagett-Dame, M.; Knutson, D. Vitamin A in Reproduction and Development. Nutrients 2011, 3, 385–428. [Google Scholar] [CrossRef] [Green Version]
  63. Bar-El Dadon, S.; Reifen, R. Vitamin A and the Epigenome. Crit. Rev. Food Sci. Nutr. 2017, 57, 2404–2411. [Google Scholar] [CrossRef]
  64. Cantorna, M.T.; Snyder, L.; Arora, J. Vitamin A and Vitamin D Regulate the Microbial Complexity, Barrier Function, and the Mucosal Immune Responses to Ensure Intestinal Homeostasis. Crit. Rev. Biochem. Mol. Biol. 2019, 54, 184–192. [Google Scholar] [CrossRef]
  65. Holick, M.F. Chapter 4—Photobiology of Vitamin D. In Vitamin D, 4th ed.; Feldman, D., Ed.; Academic Press: Cambridge, MA, USA, 2018; pp. 45–55. [Google Scholar] [CrossRef]
  66. Jones, G. Metabolism and Biomarkers of Vitamin D. Scand. J. Clin. Lab. Investig. Suppl. 2012, 243, 7–13. [Google Scholar] [CrossRef]
  67. Jones, G.; Prosser, D.E.; Kaufmann, M. Cytochrome P450-Mediated Metabolism of Vitamin D. J. Lipid. Res. 2014, 55, 13–31. [Google Scholar] [CrossRef] [Green Version]
  68. Jones, G.; Prosser, D.E.; Kaufmann, M. 25-Hydroxyvitamin D-24-Hydroxylase (CYP24A1): Its Important Role in the Degradation of Vitamin D. Arch. Biochem. Biophys. 2012, 523, 9–18. [Google Scholar] [CrossRef]
  69. Norman, A.W. From Vitamin D to Hormone D: Fundamentals of the Vitamin D Endocrine System Essential for Good Health. Am. J. Clin. Nutr. 2008, 88, 491S–499S. [Google Scholar] [CrossRef] [Green Version]
  70. Samuel, S.; Sitrin, M.D. Vitamin D’s role in cell proliferation and differentiation. Nutr. Rev. 2008, 66, S116–S124. [Google Scholar] [CrossRef]
  71. Shaker, J.L.; Deftos, L. Calcium and Phosphate Homeostasis. In Endotext; Feingold, K.R., Anawalt, B., Boyce, A., Chrousos, G., de Herder, W.W., Dhatariya, K., Dungan, K., Hershman, J.M., Hofland, J., Kalra, S., et al., Eds.; MDText.com, Inc.: South Dartmouth, MA, USA, 2000. [Google Scholar]
  72. Bowry, V.W.; Ingold, K.U.; Stocker, R. Vitamin E in Human Low-Density Lipoprotein. When and How This Antioxidant Becomes a pro-Oxidant. Biochem. J. 1992, 288 Pt 2, 341–344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Jilani, T.; Iqbal, M.P. Does Vitamin E Have a Role in Treatment and Prevention of Anemia? Pak. J. Pharm. Sci. 2011, 24, 237–242. [Google Scholar] [PubMed]
  74. Napolitano, G.; Fasciolo, G.; Di Meo, S.; Venditti, P. Vitamin E Supplementation and Mitochondria in Experimental and Functional Hyperthyroidism: A Mini-Review. Nutrients 2019, 11, 2900. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Chow, C.K. Vitamin E Regulation of Mitochondrial Superoxide Generation. Biol. Signals Recept. 2001, 10, 112–124. [Google Scholar] [CrossRef]
  76. Bozaykut, P.; Ekren, R.; Sezerman, O.U.; Gladyshev, V.N.; Ozer, N.K. High-Throughput Profiling Reveals Perturbation of Endoplasmic Reticulum Stress-Related Genes in Atherosclerosis Induced by High-Cholesterol Diet and the Protective Role of Vitamin E. Biofactors 2020, 46, 653–664. [Google Scholar] [CrossRef]
  77. Office of Dietary Supplements—Vitamin E. Available online: https://ods.od.nih.gov/factsheets/VitaminE-HealthProfessional/ (accessed on 28 December 2021).
  78. Zingg, J.-M. Vitamin E: A Role in Signal Transduction. Annu. Rev. Nutr. 2015, 35, 135–173. [Google Scholar] [CrossRef]
  79. Boscoboinik, D.; Szewczyk, A.; Azzi, A. Alpha-Tocopherol (Vitamin E) Regulates Vascular Smooth Muscle Cell Proliferation and Protein Kinase C Activity. Arch. Biochem. Biophys. 1991, 286, 264–269. [Google Scholar] [CrossRef]
  80. Boscoboinik, D.; Szewczyk, A.; Hensey, C.; Azzi, A. Inhibition of Cell Proliferation by Alpha-Tocopherol. Role of Protein Kinase C. J. Biol. Chem. 1991, 266, 6188–6194. [Google Scholar] [CrossRef]
  81. Lloret, A.; Esteve, D.; Monllor, P.; Cervera-Ferri, A.; Lloret, A. The Effectiveness of Vitamin E Treatment in Alzheimer’s Disease. Int. J. Mol. Sci. 2019, 20, 879. [Google Scholar] [CrossRef] [Green Version]
  82. Berdnikovs, S.; Abdala-Valencia, H.; McCary, C.; Somand, M.; Cole, R.; Garcia, A.; Bryce, P.; Cook-Mills, J.M. Isoforms of Vitamin E Have Opposing Immunoregulatory Functions during Inflammation by Regulating Leukocyte Recruitment. J. Immunol. 2009, 182, 4395–4405. [Google Scholar] [CrossRef] [Green Version]
  83. Cook-Mills, J.M. Isoforms of Vitamin E Differentially Regulate PKC α and Inflammation: A Review. J. Clin. Cell. Immunol. 2013, 4, 1000137. [Google Scholar] [CrossRef]
  84. Njus, D.; Kelley, P.M. Vitamins C and E donate single hydrogen atoms in vivo. FEBS Lett. 1991, 284, 147–151. [Google Scholar] [CrossRef] [Green Version]
  85. Sultana, R.; Perluigi, M.; Butterfield, D.A. Lipid Peroxidation Triggers Neurodegeneration: A Redox Proteomics View into the Alzheimer Disease Brain. Free Radic. Biol. Med. 2013, 62, 157–169. [Google Scholar] [CrossRef] [Green Version]
  86. Neelamegam, M.; Looi, I.; Ng, K.S.; Malavade, S.S. Vitamin E Supplementation for Preventing Recurrent Stroke and Other Vascular Events in Patients with Stroke or Transient Ischaemic Attack. Cochrane Database Syst. Rev. 2017, 2017, CD010797. [Google Scholar] [CrossRef]
  87. Chow, C.K.; Hong, C.B. Dietary Vitamin E and Selenium and Toxicity of Nitrite and Nitrate. Toxicology 2002, 180, 195–207. [Google Scholar] [CrossRef]
  88. d’Ischia, M.; Novellino, L. Nitric Oxide-Induced Oxidation of Alpha-Tocopherol. Bioorg. Med. Chem. 1996, 4, 1747–1753. [Google Scholar] [CrossRef]
  89. Booth, S.L. Vitamin K: Food Composition and Dietary Intakes. Food Nutr. Res. 2012, 56, 5505. [Google Scholar] [CrossRef] [Green Version]
  90. Imbrescia, K.; Moszczynski, Z. Vitamin K. In StatPearls; StatPearls Publishing: Treasure Island, FL, USA, 2021. [Google Scholar]
  91. Tie, J.-K.; Stafford, D.W. Chapter Fourteen—Functional Study of the Vitamin K Cycle Enzymes in Live Cells. In Methods in Enzymology; Gelb, M.H., Ed.; Enzymology at the Membrane Interface: Intramembrane Proteases; Academic Press: Cambridge, MA, USA, 2017; Volume 584, pp. 349–394. [Google Scholar] [CrossRef] [Green Version]
  92. Tie, J.-K.; Stafford, D.W. Structural and Functional Insights into Enzymes of the Vitamin K Cycle. J. Thromb. Haemost. 2016, 14, 236–247. [Google Scholar] [CrossRef] [Green Version]
  93. Oldenburg, J.; Bevans, C.G.; Müller, C.R.; Watzka, M. Vitamin K Epoxide Reductase Complex Subunit 1 (VKORC1): The Key Protein of the Vitamin K Cycle. Antioxid. Redox Signal. 2006, 8, 347–353. [Google Scholar] [CrossRef] [PubMed]
  94. Suttie, J.W. Vitamin K-Dependent Carboxylase. Annu. Rev. Biochem. 1985, 54, 459–477. [Google Scholar] [CrossRef] [PubMed]
  95. The Vitamin K Cycle—STAFFORD—2005—Journal of Thrombosis and Haemostasis—Wiley Online Library. Available online: https://onlinelibrary.wiley.com/doi/full/10.1111/j.1538-7836.2005.01419.x (accessed on 28 December 2021).
  96. Lanzkowsky, P. Chapter 7—Megaloblastic Anemia. In Lanzkowsky’s Manual of Pediatric Hematology and Oncology, 6th ed.; Lanzkowsky, P., Lipton, J.M., Fish, J.D., Eds.; Academic Press: San Diego, CA, USA, 2016; pp. 84–101. [Google Scholar] [CrossRef]
  97. Xv, F.; Chen, J.; Duan, L.; Li, S. Research Progress on the Anticancer Effects of Vitamin K2 (Review). Oncol. Lett. 2018, 15, 8926–8934. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Harshman, S.G.; Shea, M.K. The Role of Vitamin K in Chronic Aging Diseases: Inflammation, Cardiovascular Disease, and Osteoarthritis. Curr. Nutr. Rep. 2016, 5, 90–98. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Ferland, G. Vitamin K and the Nervous System: An Overview of Its Actions. Adv. Nutr. 2012, 3, 204–212. [Google Scholar] [CrossRef] [Green Version]
  100. Kennedy, D.O. B Vitamins and the Brain: Mechanisms, Dose and Efficacy—A Review. Nutrients 2016, 8, 68. [Google Scholar] [CrossRef] [Green Version]
  101. Wilson, D.; Branda, N.R. Turning “On” and “Off” a Pyridoxal 5′-Phosphate Mimic Using Light. Angew. Chem. Int. Ed. 2012, 51, 5431–5434. [Google Scholar] [CrossRef]
  102. Daugherty, M.; Polanuyer, B.; Farrell, M.; Scholle, M.; Lykidis, A.; de Crécy-Lagard, V.; Osterman, A. Complete Reconstitution of the Human Coenzyme A Biosynthetic Pathway via Comparative Genomics. J. Biol. Chem. 2002, 277, 21431–21439. [Google Scholar] [CrossRef] [Green Version]
  103. Shi, L.; Tu, B.P. Acetyl-CoA and the Regulation of Metabolism: Mechanisms and Consequences. Curr. Opin. Cell. Biol. 2015, 33, 125–131. [Google Scholar] [CrossRef] [Green Version]
  104. New Considerations on the Neuromodulatory Role of Thiamine—Abstract—Pharmacology 2012. Volume 89, No. 1–2—Karger Publishers. Available online: https://www.karger.com/Article/Abstract/336339 (accessed on 28 December 2021).
  105. Dias, C.; Nylandsted, J. Plasma Membrane Integrity in Health and Disease: Significance and Therapeutic Potential. Cell Discov. 2021, 7, 4. [Google Scholar] [CrossRef]
  106. Sharma, A.; Bist, R. Alteration in Cholinesterases, γ-Aminobutyric Acid and Serotonin Level with Respect to Thiamine Deficiency in Swiss Mice. Turk. J. Biochem. 2019, 44, 218–223. [Google Scholar] [CrossRef]
  107. Zhao, N.; Zhong, C.; Wang, Y.; Zhao, Y.; Gong, N.; Zhou, G.; Xu, T.; Hong, Z. Impaired Hippocampal Neurogenesis Is Involved in Cognitive Dysfunction Induced by Thiamine Deficiency at Early Pre-Pathological Lesion Stage. Neurobiol. Dis. 2008, 29, 176–185. [Google Scholar] [CrossRef] [PubMed]
  108. Barclay, L.L.; Gibson, G.E.; Blass, J.P. Impairment of Behavior and Acetylcholine Metabolism in Thiamine Deficiency. J. Pharm. Exp. 1981, 217, 537–543. [Google Scholar]
  109. Pinnell, S.R. Regulation of Collagen Biosynthesis by Ascorbic Acid: A Review. Yale J. Biol. Med. 1985, 58, 553–559. [Google Scholar] [PubMed]
  110. Buettner, G.R. The Pecking Order of Free Radicals and Antioxidants: Lipid Peroxidation, Alpha-Tocopherol, and Ascorbate. Arch. Biochem. Biophys. 1993, 300, 535–543. [Google Scholar] [CrossRef] [PubMed]
  111. Traber, M.G.; Stevens, J.F. Vitamins C and E: Beneficial Effects from a Mechanistic Perspective. Free Radic. Biol. Med. 2011, 51, 1000–1013. [Google Scholar] [CrossRef] [Green Version]
  112. Wilson, J.X. Regulation of Vitamin C Transport. Annu. Rev. Nutr. 2005, 25, 105–125. [Google Scholar] [CrossRef]
  113. Buettner, G.R.; Jurkiewicz, B.A. Catalytic Metals, Ascorbate and Free Radicals: Combinations to Avoid. Radiat. Res. 1996, 145, 532–541. [Google Scholar] [CrossRef] [Green Version]
  114. Kaźmierczak-Barańska, J.; Boguszewska, K.; Adamus-Grabicka, A.; Karwowski, B.T. Two Faces of Vitamin C—Antioxidative and Pro-Oxidative Agent. Nutrients 2020, 12, 1501. [Google Scholar] [CrossRef]
  115. Chakraborthy, A.; Ramani, P.; Sherlin, H.J.; Premkumar, P.; Natesan, A. Antioxidant and Pro-Oxidant Activity of Vitamin C in Oral Environment. Indian J. Dent. Res. 2014, 25, 499–504. [Google Scholar] [CrossRef]
  116. Cameron, E.; Pauling, L. Supplemental Ascorbate in the Supportive Treatment of Cancer: Prolongation of Survival Times in Terminal Human Cancer. Proc. Natl. Acad. Sci. USA 1976, 73, 3685–3689. [Google Scholar] [CrossRef] [Green Version]
  117. Pfister, R.; Sharp, S.J.; Luben, R.; Wareham, N.J.; Khaw, K.-T. Plasma Vitamin C Predicts Incident Heart Failure in Men and Women in European Prospective Investigation into Cancer and Nutrition-Norfolk Prospective Study. Am. Heart. J. 2011, 162, 246–253. [Google Scholar] [CrossRef] [PubMed]
  118. Osganian, S.K.; Stampfer, M.J.; Rimm, E.; Spiegelman, D.; Hu, F.B.; Manson, J.E.; Willett, W.C. Vitamin C and Risk of Coronary Heart Disease in Women. J. Am. Coll. Cardiol. 2003, 42, 246–252. [Google Scholar] [CrossRef] [Green Version]
  119. Moretti, M.; Fraga, D.B.; Rodrigues, A.L.S. Preventive and Therapeutic Potential of Ascorbic Acid in Neurodegenerative Diseases. CNS Neurosci. 2017, 23, 921–929. [Google Scholar] [CrossRef]
  120. Hansen, S.N.; Tveden-Nyborg, P.; Lykkesfeldt, J. Does Vitamin C Deficiency Affect Cognitive Development and Function? Nutrients 2014, 6, 3818–3846. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Dixit, S.; Bernardo, A.; Walker, M.J.; Kennard, J.A.; Kim, G.Y.; Kessler, E.S.; Harrison, F.E. Vitamin C Deficiency in the Brain Impairs Cognition, Increases Amyloid Accumulation and Deposition, and Oxidative Stress in APP/PSEN1 and Normally-Aging Mice. ACS Chem. Neurosci. 2015, 6, 570–581. [Google Scholar] [CrossRef] [Green Version]
  122. Harrison, F.E.; Green, R.J.; Dawes, S.M.; May, J.M. Vitamin C Distribution and Retention in the Mouse Brain. Brain Res. 2010, 1348, 181–186. [Google Scholar] [CrossRef] [Green Version]
  123. Harrison, F.E.; May, J.M.; McDonald, M.P. Vitamin C Deficiency Increases Basal Exploratory Activity but Decreases Scopolamine-Induced Activity in APP/PSEN1 Transgenic Mice. Pharm. Biochem. Behav. 2010, 94, 543–552. [Google Scholar] [CrossRef] [Green Version]
  124. Ghosh, M.K.; Chattopadhyay, D.J.; Chatterjee, I.B. Vitamin C Prevents Oxidative Damage. Free Radic. Res. 1996, 25, 173–179. [Google Scholar] [CrossRef]
  125. Ward, M.S.; Lamb, J.; May, J.M.; Harrison, F.E. Behavioral and Monoamine Changes Following Severe Vitamin C Deficiency. J. Neurochem. 2013, 124, 363–375. [Google Scholar] [CrossRef] [Green Version]
  126. Harrison, F.E.; May, J.M. Vitamin C Function in the Brain: Vital Role of the Ascorbate Transporter (SVCT2). Free Radic. Biol. Med. 2009, 46, 719–730. [Google Scholar] [CrossRef] [Green Version]
  127. Landrum, J.T. (Ed.) Carotenoids: Physical, Chemical, and Biological Functions and Properties; CRC Press: Boca Raton, FL, USA, 2009. [Google Scholar] [CrossRef]
  128. Fiedor, J.; Fiedor, L.; Haessner, R.; Scheer, H. Cyclic Endoperoxides of Beta-Carotene, Potential pro-Oxidants, as Products of Chemical Quenching of Singlet Oxygen. Biochim. Biophys. Acta 2005, 1709, 1–4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Wisniewska, A.; Subczynski, W.K. Effects of Polar Carotenoids on the Shape of the Hydrophobic Barrier of Phospholipid Bilayers. Biochim. Biophys. Acta 1998, 1368, 235–246. [Google Scholar] [CrossRef] [Green Version]
  130. Accumulation of Macular Xanthophylls in Unsaturated Membrane Domains—Abstract—Europe PMC. Available online: https://europepmc.org/article/med/16678020 (accessed on 28 December 2021).
  131. Bertram, J.S. Cancer Prevention by Carotenoids. Mechanistic Studies in Cultured Cells. Ann. N. Y. Acad. Sci. 1993, 691, 177–191. [Google Scholar] [CrossRef]
  132. Krinsky, N.I. Micronutrients and Their Influence on Mutagenicity and Malignant Transformation. Ann. N. Y. Acad. Sci. 1993, 686, 229–242. [Google Scholar] [CrossRef]
  133. Walk, A.M.; Khan, N.A.; Barnett, S.M.; Raine, L.B.; Kramer, A.F.; Cohen, N.J.; Moulton, C.J.; Renzi-Hammond, L.M.; Hammond, B.R.; Hillman, C.H. From Neuro-Pigments to Neural Efficiency: The Relationship between Retinal Carotenoids and Behavioral and Neuroelectric Indices of Cognitive Control in Childhood. Int. J. Psychophysiol. 2017, 118, 1–8. [Google Scholar] [CrossRef]
  134. Grodstein, F.; Kang, J.H.; Glynn, R.J.; Cook, N.R.; Gaziano, J.M. A Randomized Trial of Beta Carotene Supplementation and Cognitive Function in Men: The Physicians’ Health Study II. Arch. Intern. Med. 2007, 167, 2184–2190. [Google Scholar] [CrossRef] [Green Version]
  135. Johnson, E.J.; McDonald, K.; Caldarella, S.M.; Chung, H.-Y.; Troen, A.M.; Snodderly, D.M. Cognitive Findings of an Exploratory Trial of Docosahexaenoic Acid and Lutein Supplementation in Older Women. Nutr. Neurosci. 2008, 11, 75–83. [Google Scholar] [CrossRef]
  136. Rowles, J.L.; Erdman, J.W. Carotenoids and Their Role in Cancer Prevention. Biochim. Biophys. Acta BBA Mol. Cell Biol. Lipids 2020, 1865, 158613. [Google Scholar] [CrossRef]
  137. Jonasson, L.; Wikby, A.; Olsson, A.G. Low Serum Beta-Carotene Reflects Immune Activation in Patients with Coronary Artery Disease. Nutr. Metab. Cardiovasc. Dis. 2003, 13, 120–125. [Google Scholar] [CrossRef]
  138. Adewoyin, M.; Ibrahim, M.; Roszaman, R.; Md Isa, M.L.; Mat Alewi, N.A.; Abdul Rafa, A.A.; Anuar, M.N.N. Male Infertility: The Effect of Natural Antioxidants and Phytocompounds on Seminal Oxidative Stress. Diseases 2017, 5, 9. [Google Scholar] [CrossRef]
  139. Pike, T.W.; Blount, J.D.; Lindström, J.; Metcalfe, N.B. Dietary Carotenoid Availability, Sexual Signalling and Functional Fertility in Sticklebacks. Biol. Lett. 2010, 6, 191–193. [Google Scholar] [CrossRef] [PubMed]
  140. Sharoni, Y.; Linnewiel-Hermoni, K.; Zango, G.; Khanin, M.; Salman, H.; Veprik, A.; Danilenko, M.; Levy, J. The Role of Lycopene and Its Derivatives in the Regulation of Transcription Systems: Implications for Cancer Prevention. Am. J. Clin. Nutr. 2012, 96, 1173S–1178S. [Google Scholar] [CrossRef] [PubMed]
  141. Kumar, N.; Goel, N. Phenolic Acids: Natural Versatile Molecules with Promising Therapeutic Applications. Biotechnol. Rep. Amst. 2019, 24, e00370. [Google Scholar] [CrossRef] [PubMed]
  142. Fernandes, I.; Pérez-Gregorio, R.; Soares, S.; Mateus, N.; de Freitas, V. Wine Flavonoids in Health and Disease Prevention. Molecules 2017, 22, 292. [Google Scholar] [CrossRef] [PubMed]
  143. Musial, C.; Kuban-Jankowska, A.; Gorska-Ponikowska, M. Beneficial Properties of Green Tea Catechins. Int. J. Mol. Sci. 2020, 21, 1744. [Google Scholar] [CrossRef] [Green Version]
  144. Sultana, B.; Anwar, F. Flavonols (Kaempeferol, Quercetin, Myricetin) Contents of Selected Fruits, Vegetables and Medicinal Plants. Food Chem. 2008, 108, 879–884. [Google Scholar] [CrossRef]
  145. Reinisalo, M.; Kårlund, A.; Koskela, A.; Kaarniranta, K.; Karjalainen, R.O. Polyphenol Stilbenes: Molecular Mechanisms of Defence against Oxidative Stress and Aging-Related Diseases. Oxidative Med. Cell. Longev. 2015, 2015, 340520. [Google Scholar] [CrossRef] [Green Version]
  146. Al-Suhaimi, E.A.; Al-Riziza, N.A.; Al-Essa, R.A. Physiological and Therapeutical Roles of Ginger and Turmeric on Endocrine Functions. Am. J. Chin. Med. 2011, 39, 215–231. [Google Scholar] [CrossRef]
  147. Eren, D.; Betul, Y.M. Revealing the Effect of 6-Gingerol, 6-Shogaol and Curcumin on MPGES-1, GSK-3β and β-Catenin Pathway in A549 Cell Line. Chem. Biol. Interact. 2016, 258, 257–265. [Google Scholar] [CrossRef] [PubMed]
  148. Zhou, Y.; Zheng, J.; Li, Y.; Xu, D.-P.; Li, S.; Chen, Y.-M.; Li, H.-B. Natural Polyphenols for Prevention and Treatment of Cancer. Nutrients 2016, 8, E515. [Google Scholar] [CrossRef]
  149. Fujiki, H.; Sueoka, E.; Watanabe, T.; Suganuma, M. Primary Cancer Prevention by Green Tea, and Tertiary Cancer Prevention by the Combination of Green Tea Catechins and Anticancer Compounds. J. Cancer Prev. 2015, 20, 1–4. [Google Scholar] [CrossRef] [PubMed]
  150. Sirerol, J.A.; Rodríguez, M.L.; Mena, S.; Asensi, M.A.; Estrela, J.M.; Ortega, A.L. Role of Natural Stilbenes in the Prevention of Cancer. Oxidative Med. Cell. Longev. 2015, 2016, e3128951. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Guo, R.; Li, W.; Liu, B.; Li, S.; Zhang, B.; Xu, Y. Resveratrol Protects Vascular Smooth Muscle Cells against High Glucose-Induced Oxidative Stress and Cell Proliferation in Vitro. Med. Sci. Monit. Basic Res. 2014, 20, 82–92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Yamagata, K.; Tagami, M.; Yamori, Y. Dietary Polyphenols Regulate Endothelial Function and Prevent Cardiovascular Disease. Nutrition 2015, 31, 28–37. [Google Scholar] [CrossRef]
  153. Xiao, J.B.; Högger, P. Dietary Polyphenols and Type 2 Diabetes: Current Insights and Future Perspectives. Curr. Med. Chem. 2015, 22, 23–38. [Google Scholar] [CrossRef]
  154. Zhang, H.; Tsao, R. Dietary Polyphenols, Oxidative Stress and Antioxidant and Anti-Inflammatory Effects. Curr. Opin. Food Sci. 2016, 8, 33–42. [Google Scholar] [CrossRef]
  155. Rossi, L.; Mazzitelli, S.; Arciello, M.; Capo, C.R.; Rotilio, G. Benefits from Dietary Polyphenols for Brain Aging and Alzheimer’s Disease. Neurochem. Res. 2008, 33, 2390–2400. [Google Scholar] [CrossRef]
  156. Noguchi-Shinohara, M.; Yuki, S.; Dohmoto, C.; Ikeda, Y.; Samuraki, M.; Iwasa, K.; Yokogawa, M.; Asai, K.; Komai, K.; Nakamura, H.; et al. Consumption of Green Tea, but Not Black Tea or Coffee, Is Associated with Reduced Risk of Cognitive Decline. PLoS ONE 2014, 9, e96013. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Aquilano, K.; Baldelli, S.; Rotilio, G.; Ciriolo, M.R. Role of Nitric Oxide Synthases in Parkinson’s Disease: A Review on the Antioxidant and Anti-Inflammatory Activity of Polyphenols. Neurochem. Res. 2008, 33, 2416–2426. [Google Scholar] [CrossRef]
  158. Bao, J.; Liu, W.; Zhou, H.-Y.; Gui, Y.-R.; Yang, Y.-H.; Wu, M.-J.; Xiao, Y.-F.; Shang, J.-T.; Long, G.-F.; Shu, X.-J. Epigallocatechin-3-Gallate Alleviates Cognitive Deficits in APP/PS1 Mice. Curr. Med. Sci. 2020, 40, 18–27. [Google Scholar] [CrossRef]
  159. Broderick, T.L.; Rasool, S.; Li, R.; Zhang, Y.; Anderson, M.; Al-Nakkash, L.; Plochocki, J.H.; Geetha, T.; Babu, J.R. Neuroprotective Effects of Chronic Resveratrol Treatment and Exercise Training in the 3xTg-AD Mouse Model of Alzheimer’s Disease. Int. J. Mol. Sci. 2020, 21, 7337. [Google Scholar] [CrossRef]
  160. 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] [PubMed]
  161. 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] [PubMed]
  162. Spagnuolo, C.; Napolitano, M.; Tedesco, I.; Moccia, S.; Milito, A.; Russo, G.L. Neuroprotective Role of Natural Polyphenols. Curr. Top. Med. Chem. 2016, 16, 1943–1950. [Google Scholar] [CrossRef]
  163. Waggoner, D.J.; Bartnikas, T.B.; Gitlin, J.D. The Role of Copper in Neurodegenerative Disease. Neurobiol. Dis. 1999, 6, 221–230. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Ozcelik, D.; Uzun, H. Copper Intoxication; Antioxidant Defenses and Oxidative Damage in Rat Brain. Biol. Trace Elem. Res. 2009, 127, 45–52. [Google Scholar] [CrossRef] [PubMed]
  165. Hirayama, K.; Yasutake, A.; Inoue, M. Free Radicals and Trace Elements. Prog. Clin. Biol. Res. 1993, 380, 257–268. [Google Scholar] [CrossRef]
  166. Hentze, M.W.; Muckenthaler, M.U.; Galy, B.; Camaschella, C. Two to Tango: Regulation of Mammalian Iron Metabolism. Cell 2010, 142, 24–38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Li, L.; Yang, X. The Essential Element Manganese, Oxidative Stress, and Metabolic Diseases: Links and Interactions. Oxidative Med. Cell. Longev. 2018, 2018, e7580707. [Google Scholar] [CrossRef] [Green Version]
  168. Coskun, M.; Kayis, T.; Gulsu, E.; Alp, E. Effects of Selenium and Vitamin E on Enzymatic, Biochemical, and Immunological Biomarkers in Galleria mellonella L. Sci. Rep. 2020, 10, 9953. [Google Scholar] [CrossRef]
  169. Jerome-Morais, A.; Bera, S.; Rachidi, W.; Gann, P.H.; Diamond, A.M. The Effects of Selenium and the GPx-1 Selenoprotein on the Phosphorylation of H2AX. Biochim. Biophys. Acta 2013, 1830, 3399–3406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Prasad, A.S.; Miale, A.; Farid, Z.; Sandstead, H.H.; Schulert, A.R. Zinc Metabolism in Patients with the Syndrome of Iron Deficiency Anemia, Hepatosplenomegaly, Dwarfism, and Hypognadism. J. Lab. Clin. Med. 1963, 61, 537–549. [Google Scholar] [PubMed]
  171. Bao, B.; Ahmad, A.; Azmi, A.; Li, Y.; Prasad, A.; Sarkar, F.H. Chapter 2—The Biological Significance of Zinc in Inflammation and Aging. In Inflammation, Advancing Age and Nutrition; Rahman, I., Bagchi, D., Eds.; Academic Press: San Diego, FL, USA, 2014; pp. 15–27. [Google Scholar] [CrossRef]
  172. Ha, K.-N.; Chen, Y.; Cai, J.; Sternberg, P., Jr. Increased Glutathione Synthesis through an ARE-Nrf2–Dependent Pathway by Zinc in the RPE: Implication for Protection against Oxidative Stress. Investig. Ophthalmol. Vis. Sci. 2006, 47, 2709–2715. [Google Scholar] [CrossRef] [Green Version]
  173. Kaufman, Z.; Salvador, G.A.; Liu, X.; Oteiza, P.I. Zinc and the Modulation of Nrf2 in Human Neuroblastoma Cells. Free Radic. Biol. Med. 2020, 155, 1–9. [Google Scholar] [CrossRef]
  174. Li, B.; Cui, W.; Tan, Y.; Luo, P.; Chen, Q.; Zhang, C.; Qu, W.; Miao, L.; Cai, L. Zinc Is Essential for the Transcription Function of Nrf2 in Human Renal Tubule Cells in Vitro and Mouse Kidney in Vivo under the Diabetic Condition. J. Cell. Mol. Med. 2014, 18, 895–906. [Google Scholar] [CrossRef]
  175. Kocot, J.; Luchowska-Kocot, D.; Kiełczykowska, M.; Musik, I.; Kurzepa, J. Does Vitamin C Influence Neurodegenerative Diseases and Psychiatric Disorders? Nutrients 2017, 9, 659. [Google Scholar] [CrossRef] [Green Version]
  176. Ma, Q. Role of Nrf2 in Oxidative Stress and Toxicity. Annu. Rev. Pharm. Toxicol. 2013, 53, 401–426. [Google Scholar] [CrossRef] [Green Version]
  177. Norbury, R.; Cutter, W.; Compton, J.; Robertson, D.; Craig, M.; Whitehead, M.; Murphy, D. The Neuroprotective Effects of Estrogen on the Aging Brain. Exp. Gerontol. 2003, 38, 109–117. [Google Scholar] [CrossRef]
  178. Sherwin, B.B. Estrogen Effects on Cognition in Menopausal Women. Neurology 1997, 48, S21–S26. [Google Scholar] [CrossRef]
  179. Sherwin, B.B. Estrogenic Effects on Memory in Women. Ann. N. Y. Acad. Sci. 1994, 743, 213–230; discussion 230–231. [Google Scholar] [CrossRef]
  180. Henderson, V.W.; Watt, L.; Buckwalter, J.G. Cognitive Skills Associated with Estrogen Replacement in Women with Alzheimer’s Disease. Psychoneuroendocrinology 1996, 21, 421–430. [Google Scholar] [CrossRef]
  181. Birge, S.J. The Role of Estrogen in the Treatment of Alzheimer’s Disease. Neurology 1997, 48, S36–S41. [Google Scholar] [CrossRef] [PubMed]
  182. Fillit, H. Estrogens in the Pathogenesis and Treatment of Alzheimer’s Disease in Postmenopausal Women. Ann. N. Y. Acad. Sci. 1994, 743, 233–238; discussion 238–239. [Google Scholar] [CrossRef] [PubMed]
  183. Paganini-Hill, A.; Henderson, V.W. Estrogen Deficiency and Risk of Alzheimer’s Disease in Women. Am. J. Epidemiol. 1994, 140, 256–261. [Google Scholar] [CrossRef] [PubMed]
  184. Fillit, H.; Weinreb, H.; Cholst, I.; Luine, V.; McEwen, B.; Amador, R.; Zabriskie, J. Observations in a Preliminary Open Trial of Estradiol Therapy for Senile Dementia-Alzheimer’s Type. Psychoneuroendocrinology 1986, 11, 337–345. [Google Scholar] [CrossRef]
  185. Barrett-Connor, E.; Kritz-Silverstein, D. Estrogen Replacement Therapy and Cognitive Function in Older Women. JAMA 1993, 269, 2637–2641. [Google Scholar] [CrossRef]
  186. Mulnard, R.A.; Cotman, C.W.; Kawas, C.; van Dyck, C.H.; Sano, M.; Doody, R.; Koss, E.; Pfeiffer, E.; Jin, S.; Gamst, A.; et al. Estrogen Replacement Therapy for Treatment of Mild to Moderate Alzheimer Disease: A Randomized Controlled Trial. Alzheimer’s Disease Cooperative Study. JAMA 2000, 283, 1007–1015. [Google Scholar] [CrossRef]
  187. Brenner, D.E.; Kukull, W.A.; Stergachis, A.; van Belle, G.; Bowen, J.D.; McCormick, W.C.; Teri, L.; Larson, E.B. Postmenopausal Estrogen Replacement Therapy and the Risk of Alzheimer’s Disease: A Population-Based Case-Control Study. Am. J. Epidemiol. 1994, 140, 262–267. [Google Scholar] [CrossRef]
  188. Sultana, R.; Butterfield, D.A. Role of Oxidative Stress in the Progression of Alzheimer’s Disease. J. Alzheimer’s Dis. 2010, 19, 341–353. [Google Scholar] [CrossRef] [Green Version]
  189. Chen, Z.; Zhong, C. Oxidative Stress in Alzheimer’s Disease. Neurosci. Bull. 2014, 30, 271–281. [Google Scholar] [CrossRef]
  190. Wang, X.; Wang, W.; Li, L.; Perry, G.; Lee, H.; Zhu, X. Oxidative Stress and Mitochondrial Dysfunction in Alzheimer’s Disease. Biochim. Biophys. Acta BBA Mol. Basis Dis. 2014, 1842, 1240–1247. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Padurariu, M.; Ciobica, A.; Hritcu, L.; Stoica, B.; Bild, W.; Stefanescu, C. Changes of Some Oxidative Stress Markers in the Serum of Patients with Mild Cognitive Impairment and Alzheimer’s Disease. Neurosci. Lett. 2010, 469, 6–10. [Google Scholar] [CrossRef] [PubMed]
  192. Guglielmotto, M.; Giliberto, L.; Tamagno, E.; Tabaton, M. Oxidative Stress Mediates the Pathogenic Effect of Different Alzheimer’s Disease Risk Factors. Front. Aging Neurosci. 2010, 2, 3. [Google Scholar] [CrossRef] [PubMed]
  193. Darvesh, A.S.; Carroll, R.T.; Bishayee, A.; Geldenhuys, W.J.; Van der Schyf, C.J. Oxidative Stress and Alzheimer’s Disease: Dietary Polyphenols as Potential Therapeutic Agents. Exp. Rev. Neurother. 2010, 10, 729–745. [Google Scholar] [CrossRef] [PubMed]
  194. Torres, L.L.; Quaglio, N.B.; de Souza, G.T.; Garcia, R.T.; Dati, L.M.M.; Moreira, W.L.; de Melo Loureiro, A.P.; de souza-Talarico, J.N.; Smid, J.; Porto, C.S.; et al. Peripheral Oxidative Stress Biomarkers in Mild Cognitive Impairment and Alzheimer’s Disease. J. Alzheimer’s Dis. 2011, 26, 59–68. [Google Scholar] [CrossRef] [PubMed]
  195. Sutherland, G.T.; Chami, B.; Youssef, P.; Witting, P.K. Oxidative Stress in Alzheimer’s Disease: Primary Villain or Physiological by-Product? Redox Rep. 2013, 18, 134–141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Chang, Y.-T.; Chang, W.-N.; Tsai, N.-W.; Huang, C.-C.; Kung, C.-T.; Su, Y.-J.; Lin, W.-C.; Cheng, B.-C.; Su, C.-M.; Chiang, Y.-F.; et al. The Roles of Biomarkers of Oxidative Stress and Antioxidant in Alzheimer’s Disease: A Systematic Review. BioMed. Res. Int. 2014, 2014, e182303. [Google Scholar] [CrossRef]
  197. Meraz-Ríos, M.A.; Franco-Bocanegra, D.; Toral Rios, D.; Campos-Peña, V. Early Onset Alzheimer’s Disease and Oxidative Stress. Oxidative Med. Cell. Longev. 2014, 2014, e375968. [Google Scholar] [CrossRef] [Green Version]
  198. Bonda, D.J.; Lee, H.; Blair, J.A.; Zhu, X.; Perry, G.; Smith, M.A. Role of Metal Dyshomeostasis in Alzheimer’s Disease. Metallomics 2011, 3, 267–270. [Google Scholar] [CrossRef] [Green Version]
  199. Pohanka, M. Alzheimer´s Disease and Oxidative Stress: A Review. Curr. Med. Chem. 2014, 21, 356–364. [Google Scholar] [CrossRef]
  200. Du, X.; Wang, X.; Geng, M. Alzheimer’s Disease Hypothesis and Related Therapies. Transl. Neurodegener. 2018, 7, 2. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  201. Luca, M.; Luca, A.; Calandra, C. The Role of Oxidative Damage in the Pathogenesis and Progression of Alzheimer’s Disease and Vascular Dementia. Oxidative Med. Cell. Longev. 2015, 2015, e504678. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Wojtunik-Kulesza, K.A.; Oniszczuk, A.; Oniszczuk, T.; Waksmundzka-Hajnos, M. The Influence of Common Free Radicals and Antioxidants on Development of Alzheimer’s Disease. Biomed. Pharmacother. 2016, 78, 39–49. [Google Scholar] [CrossRef] [PubMed]
  203. Padurariu, M.; Ciobica, A.; Lefter, R.; Serban, I.L.; Stefanescu, C.; Chirita, R. The Oxidative Stress Hypothesis in Alzheimer’s Disease. Psychiatr. Danub. 2013, 25, 401–409. [Google Scholar]
  204. Hung, C.H.-L.; Cheng, S.S.-Y.; Cheung, Y.-T.; Wuwongse, S.; Zhang, N.Q.; Ho, Y.-S.; Lee, S.M.-Y.; Chang, R.C.-C. A Reciprocal Relationship between Reactive Oxygen Species and Mitochondrial Dynamics in Neurodegeneration. Redox Biol. 2018, 14, 7–19. [Google Scholar] [CrossRef] [PubMed]
  205. Agostinho, P.; Cunha, R.A.; Oliveira, C. Neuroinflammation, Oxidative Stress and the Pathogenesis of Alzheimer’s Disease. Curr. Pharm. Des. 2010, 16, 2766–2778. [Google Scholar] [CrossRef]
  206. Ganguly, G.; Chakrabarti, S.; Chatterjee, U.; Saso, L. Proteinopathy, Oxidative Stress and Mitochondrial Dysfunction: Cross Talk in Alzheimer’s Disease and Parkinson’s Disease. Drug. Des. Devel. 2017, 11, 797–810. [Google Scholar] [CrossRef] [Green Version]
  207. González-Reyes, R.E.; Nava-Mesa, M.O.; Vargas-Sánchez, K.; Ariza-Salamanca, D.; Mora-Muñoz, L. Involvement of Astrocytes in Alzheimer’s Disease from a Neuroinflammatory and Oxidative Stress Perspective. Front. Mol. Neurosci. 2017, 10, 427. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Verri, M.; Pastoris, O.; Dossena, M.; Aquilani, R.; Guerriero, F.; Cuzzoni, G.; Venturini, L.; Ricevuti, G.; Bongiorno, A.I. Mitochondrial Alterations, Oxidative Stress and Neuroinflammation in Alzheimer’s Disease. Int. J. Immunopathol. Pharm. 2012, 25, 345–353. [Google Scholar] [CrossRef] [Green Version]
  209. Butterfield, D.A.; Boyd-Kimball, D. Oxidative Stress, Amyloid-β Peptide, and Altered Key Molecular Pathways in the Pathogenesis and Progression of Alzheimer’s Disease. J. Alzheimer’s Dis. 2018, 62, 1345–1367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Takahashi, R.H.; Nagao, T.; Gouras, G.K. Plaque Formation and the Intraneuronal Accumulation of β-Amyloid in Alzheimer’s Disease. Pathol. Int. 2017, 67, 185–193. [Google Scholar] [CrossRef] [PubMed]
  211. Tillement, L.; Lecanu, L.; Papadopoulos, V. Alzheimer’s Disease: Effects of β-Amyloid on Mitochondria. Mitochondrion 2011, 11, 13–21. [Google Scholar] [CrossRef] [PubMed]
  212. Viola, K.L.; Klein, W.L. Amyloid β Oligomers in Alzheimer’s Disease Pathogenesis, Treatment, and Diagnosis. Acta Neuropathol. 2015, 129, 183–206. [Google Scholar] [CrossRef] [PubMed]
  213. Cheng, Y.; Bai, F. The Association of Tau With Mitochondrial Dysfunction in Alzheimer’s Disease. Front. Neurosci. 2018, 12, 163. [Google Scholar] [CrossRef] [PubMed]
  214. Pooler, A.M.; Polydoro, M.; Maury, E.A.; Nicholls, S.B.; Reddy, S.M.; Wegmann, S.; William, C.; Saqran, L.; Cagsal-Getkin, O.; Pitstick, R.; et al. Amyloid Accelerates Tau Propagation and Toxicity in a Model of Early Alzheimer’s Disease. Acta Neuropathol. Commun. 2015, 3, 14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Medina, M.; Avila, J. New Perspectives on the Role of Tau in Alzheimer’s Disease. Implications for Therapy. Biochem. Pharmacol. 2014, 88, 540–547. [Google Scholar] [CrossRef] [PubMed]
  216. Malinow, R. New Developments on the Role of NMDA Receptors in Alzheimer’s Disease. Curr. Opin. Neurobiol. 2012, 22, 559–563. [Google Scholar] [CrossRef] [Green Version]
  217. Zhang, Y.; Li, P.; Feng, J.; Wu, M. Dysfunction of NMDA Receptors in Alzheimer’s Disease. Neurol. Sci. 2016, 37, 1039–1047. [Google Scholar] [CrossRef] [Green Version]
  218. Bordji, K.; Becerril-Ortega, J.; Buisson, A. Synapses, NMDA Receptor Activity and Neuronal Aβ Production in Alzheimer’s Disease. Gruyter 2011, 22, 285–294. [Google Scholar] [CrossRef]
  219. Mota, S.I.; Ferreira, I.L.; Rego, A.C. Dysfunctional Synapse in Alzheimer’s Disease—A Focus on NMDA Receptors. Neuropharmacology 2014, 76, 16–26. [Google Scholar] [CrossRef]
  220. Foster, T.C.; Kyritsopoulos, C.; Kumar, A. Central Role for NMDA Receptors in Redox Mediated Impairment of Synaptic Function during Aging and Alzheimer’s Disease. Behav. Brain Res. 2017, 322, 223–232. [Google Scholar] [CrossRef] [PubMed]
  221. Kamat, P.K.; Kalani, A.; Rai, S.; Swarnkar, S.; Tota, S.; Nath, C.; Tyagi, N. Mechanism of Oxidative Stress and Synapse Dysfunction in the Pathogenesis of Alzheimer’s Disease: Understanding the Therapeutics Strategies. Mol. Neurobiol. 2016, 53, 648–661. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Beyrent, E.; Gomez, G. Oxidative Stress Differentially Induces Tau Dissociation from Neuronal Microtubules in Neurites of Neurons Cultured from Different Regions of the Embryonic Gallus Domesticus Brain. J. Neurosci. Res. 2020, 98, 734–747. [Google Scholar] [CrossRef] [PubMed]
  223. Cassidy, L.; Fernandez, F.; Johnson, J.B.; Naiker, M.; Owoola, A.G.; Broszczak, D.A. Oxidative Stress in Alzheimer’s Disease: A Review on Emergent Natural Polyphenolic Therapeutics. Complement. Ther. Med. 2020, 49, 102294. [Google Scholar] [CrossRef] [PubMed]
  224. Šerý, O.; Povová, J.; Míšek, I.; Pešák, L.; Janout, V. Molecular Mechanisms of Neuropathological Changes in Alzheimer’s Disease: A Review. Folia Neuropathol. 2013, 51, 1–9. [Google Scholar] [CrossRef]
  225. Lasagna-Reeves, C.A.; Castillo-Carranza, D.L.; Sengupta, U.; Sarmiento, J.; Troncoso, J.; Jackson, G.R.; Kayed, R. Identification of Oligomers at Early Stages of Tau Aggregation in Alzheimer’s Disease. FASEB J. 2012, 26, 1946–1959. [Google Scholar] [CrossRef] [Green Version]
  226. Evans, D.B.; Rank, K.B.; Bhattacharya, K.; Thomsen, D.R.; Gurney, M.E.; Sharma, S.K. Tau Phosphorylation at Serine 396 and Serine 404 by Human Recombinant Tau Protein Kinase II Inhibits Tau’s Ability to Promote Microtubule Assembly. J. Biol. Chem. 2000, 275, 24977–24983. [Google Scholar] [CrossRef] [Green Version]
  227. Andorfer, C. Cell-Cycle Reentry and Cell Death in Transgenic Mice Expressing Nonmutant Human Tau Isoforms. J. Neurosci. 2005, 25, 5446–5454. [Google Scholar] [CrossRef] [Green Version]
  228. Steinhilb, M.L.; Dias-Santagata, D.; Fulga, T.A.; Felch, D.L.; Feany, M.B. Tau Phosphorylation Sites Work in Concert to Promote Neurotoxicity In Vivo. MBoC 2007, 18, 5060–5068. [Google Scholar] [CrossRef] [Green Version]
  229. Alonso, A.; Grundke-Iqbal, I.; Iqbal, K. Alzheimer’s Disease Hyperphosphorylated Tau Sequesters Normal Tau into Tangles of Filaments and Disassembles Microtubules. Nat. Med. 1996, 2, 783–787. [Google Scholar] [CrossRef]
  230. Chung, C.-W.; Song, Y.-H.; Kim, I.-K.; Yoon, W.-J.; Ryu, B.-R.; Jo, D.-G.; Woo, H.-N.; Kwon, Y.-K.; Kim, H.-H.; Gwag, B.-J.; et al. Proapoptotic Effects of Tau Cleavage Product Generated by Caspase-3. Neurobiol. Dis. 2001, 8, 162–172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  231. Braak, H.; Braak, E. Neuropathological Stageing of Alzheimer-Related Changes. Acta Neuropathol. 1991, 82, 239–259. [Google Scholar] [CrossRef] [PubMed]
  232. Melov, S.; Adlard, P.A.; Morten, K.; Johnson, F.; Golden, T.R.; Hinerfeld, D.; Schilling, B.; Mavros, C.; Masters, C.L.; Volitakis, I.; et al. Mitochondrial Oxidative Stress Causes Hyperphosphorylation of Tau. PLoS ONE 2007, 2, e536. [Google Scholar] [CrossRef] [PubMed]
  233. 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. Alzheimer’s Dis. 2017, 57, 475–482. [Google Scholar] [CrossRef] [PubMed]
  234. Su, B.; Wang, X.; Lee, H.; Tabaton, M.; Perry, G.; Smith, M.A.; Zhu, X. Chronic Oxidative Stress Causes Increased Tau Phosphorylation in M17 Neuroblastoma Cells. Neurosci. Lett. 2010, 468, 267–271. [Google Scholar] [CrossRef]
  235. Egaña, J.T.; Zambrano, C.; Nuñez, M.T.; Gonzalez-Billault, C.; Maccioni, R.B. Iron-Induced Oxidative Stress Modify Tau Phosphorylation Patterns in Hippocampal Cell Cultures. Biometals 2003, 16, 215–223. [Google Scholar] [CrossRef]
  236. Wang, D.-L.; Ling, Z.-Q.; Cao, F.-Y.; Zhu, L.-Q.; Wang, J.-Z. Melatonin Attenuates Isoproterenol-Induced Protein Kinase A Overactivation and Tau Hyperphosphorylation in Rat Brain. J. Pineal Res. 2004, 37, 11–16. [Google Scholar] [CrossRef]
  237. Atlante, A.; Valenti, D.; Latina, V.; Amadoro, G. Role of Oxygen Radicals in Alzheimer’s Disease: Focus on Tau Protein. Oxygen 2021, 1, 96–120. [Google Scholar] [CrossRef]
  238. Kandimalla, R.; Manczak, M.; Yin, X.; Wang, R.; Reddy, P.H. Hippocampal Phosphorylated Tau Induced Cognitive Decline, Dendritic Spine Loss and Mitochondrial Abnormalities in a Mouse Model of Alzheimer’s Disease. Hum. Mol. Genet. 2018, 27, 30–40. [Google Scholar] [CrossRef] [Green Version]
  239. Horiguchi, T.; Uryu, K.; Giasson, B.I.; Ischiropoulos, H.; LightFoot, R.; Bellmann, C.; Richter-Landsberg, C.; Lee, V.M.-Y.; Trojanowski, J.Q. Nitration of Tau Protein Is Linked to Neurodegeneration in Tauopathies. Am. J. Pathol. 2003, 163, 1021–1031. [Google Scholar] [CrossRef] [Green Version]
  240. Torres, A.K.; Jara, C.; Olesen, M.A.; Tapia-Rojas, C. Pathologically Phosphorylated Tau at S396/404 (PHF-1) Is Accumulated inside of Hippocampal Synaptic Mitochondria of Aged Wild-Type Mice. Sci. Rep. 2021, 11, 4448. [Google Scholar] [CrossRef] [PubMed]
  241. Gamblin, T.C.; King, M.E.; Kuret, J.; Berry, R.W.; Binder, L.I. Oxidative Regulation of Fatty Acid-Induced Tau Polymerization. Biochemistry 2000, 39, 14203–14210. [Google Scholar] [CrossRef] [PubMed]
  242. Zhou, F.; Chen, S.; Xiong, J.; Li, Y.; Qu, L. Luteolin Reduces Zinc-Induced Tau Phosphorylation at Ser262/356 in an ROS-Dependent Manner in SH-SY5Y Cells. Biol. Trace Elem. Res. 2012, 149, 273–279. [Google Scholar] [CrossRef]
  243. Yamamoto, A.; Shin, R.-W.; Hasegawa, K.; Naiki, H.; Sato, H.; Yoshimasu, F.; Kitamoto, T. Iron (III) Induces Aggregation of Hyperphosphorylated τ and Its Reduction to Iron (II) Reverses the Aggregation: Implications in the Formation of Neurofibrillary Tangles of Alzheimer’s Disease. J. Neurochem. 2002, 82, 1137–1147. [Google Scholar] [CrossRef]
  244. Yang, L.; Ksiezak-Reding, H. Ca2+ and Mg2+ Selectively Induce Aggregates of PHF-Tau but Not Normal Human Tau. J. Neurosci. Res. 1999, 55, 36–43. [Google Scholar] [CrossRef]
  245. Bihaqi, S.W.; Bahmani, A.; Adem, A.; Zawia, N.H. Infantile Postnatal Exposure to Lead (Pb) Enhances Tau Expression in the Cerebral Cortex of Aged Mice: Relevance to AD. Neurotoxicology 2014, 44, 114–120. [Google Scholar] [CrossRef] [Green Version]
  246. Jiang, L.-F.; Yao, T.-M.; Zhu, Z.-L.; Wang, C.; Ji, L.-N. Impacts of Cd(II) on the Conformation and Self-Aggregation of Alzheimer’s Tau Fragment Corresponding to the Third Repeat of Microtubule-Binding Domain. Biochim. Biophys. Acta BBA Proteins Proteom. 2007, 1774, 1414–1421. [Google Scholar] [CrossRef]
  247. Olivieri, G.; Brack, C.; Müller-Spahn, F.; Stähelin, H.B.; Herrmann, M.; Renard, P.; Brockhaus, M.; Hock, C. Mercury Induces Cell Cytotoxicity and Oxidative Stress and Increases β-Amyloid Secretion and Tau Phosphorylation in SHSY5Y Neuroblastoma Cells. J. Neurochem. 2000, 74, 231–236. [Google Scholar] [CrossRef]
  248. Walton, J.R. An Aluminum-Based Rat Model for Alzheimer’s Disease Exhibits Oxidative Damage, Inhibition of PP2A Activity, Hyperphosphorylated Tau, and Granulovacuolar Degeneration. J. Inorg. Biochem. 2007, 101, 1275–1284. [Google Scholar] [CrossRef]
  249. Prema, A.; Justin Thenmozhi, A.; Manivasagam, T.; Mohamed Essa, M.; Guillemin, G.J. Fenugreek Seed Powder Attenuated Aluminum Chloride-Induced Tau Pathology, Oxidative Stress, and Inflammation in a Rat Model of Alzheimer’s Disease. J. Alzheimer’s Dis. 2017, 60, S209–S220. [Google Scholar] [CrossRef]
  250. Kim, A.C.; Lim, S.; Kim, Y.K. Metal Ion Effects on Aβ and Tau Aggregation. Int. J. Mol. Sci. 2018, 19, 128. [Google Scholar] [CrossRef] [Green Version]
  251. Yao, K.; Zhao, Y.-F.; Zu, H.-B. Melatonin Receptor Stimulation by Agomelatine Prevents Aβ-Induced Tau Phosphorylation and Oxidative Damage in PC12 Cells. Drug Des. Dev. 2019, 13, 387–396. [Google Scholar] [CrossRef] [Green Version]
  252. Busciglio, J.; Lorenzo, A.; Yeh, J. β-Amyloid fibrils induce tau phosphorylation and loss of microtubule binding. Neuron 1995, 14, 879–888. [Google Scholar] [CrossRef] [Green Version]
  253. Hernández, F.; Gómez de Barreda, E.; Fuster-Matanzo, A.; Lucas, J.J.; Avila, J. GSK3: A Possible Link between Beta Amyloid Peptide and Tau Protein. Exp. Neurol. 2010, 223, 322–325. [Google Scholar] [CrossRef]
  254. Hanger, D.P.; Anderton, B.H.; Noble, W. Tau Phosphorylation: The Therapeutic Challenge for Neurodegenerative Disease. Trends Mol. Med. 2009, 15, 112–119. [Google Scholar] [CrossRef]
  255. Noble, W.; Planel, E.; Zehr, C.; Olm, V.; Meyerson, J.; Suleman, F.; Gaynor, K.; Wang, L.; LaFrancois, J.; Feinstein, B.; et al. Inhibition of Glycogen Synthase Kinase-3 by Lithium Correlates with Reduced Tauopathy and Degeneration in Vivo. Proc. Natl. Acad. Sci. USA 2005, 102, 6990–6995. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Takashima, A.; Honda, T.; Yasutake, K.; Michel, G.; Murayama, O.; Murayama, M.; Ishiguro, K.; Yamaguchi, H. Activation of Tau Protein Kinase I/Glycogen Synthase Kinase-3β by Amyloid β Peptide (25–35) Enhances Phosphorylation of Tau in Hippocampal Neurons. Neurosci. Res. 1998, 31, 317–323. [Google Scholar] [CrossRef]
  257. Terwel, D.; Muyllaert, D.; Dewachter, I.; Borghgraef, P.; Croes, S.; Devijver, H.; Van Leuven, F. Amyloid Activates GSK-3β to Aggravate Neuronal Tauopathy in Bigenic Mice. Am. J. Pathol. 2008, 172, 786–798. [Google Scholar] [CrossRef] [Green Version]
  258. Ma, J.; Ma, C.; Li, J.; Sun, Y.; Ye, F.; Liu, K.; Zhang, H. Extracellular Matrix Proteins Involved in Alzheimer’s Disease. Chem. A Eur. J. 2020, 26, 12101–12110. [Google Scholar] [CrossRef]
  259. Debelle, L.; Tamburro, A.M. Elastin: Molecular Description and Function. Int. J. Biochem. Cell Biol. 1999, 31, 261–272. [Google Scholar] [CrossRef]
  260. Sandberg, L.B.; Soskel, N.T.; Leslie, J.G. Elastin Structure, Biosynthesis, and Relation to Disease States. Available online: https://www.nejm.org/doi/pdf/10.1056/NEJM198103053041004 (accessed on 29 December 2021). [CrossRef]
  261. Powell, J.T.; Vine, N.; Crossman, M. On the Accumulation of D-Aspartate in Elastin and Other Proteins of the Ageing Aorta. Atherosclerosis 1992, 97, 201–208. [Google Scholar] [CrossRef]
  262. Robert, L.; Molinari, J.; Ravelojaona, V.; Andrès, E.; Robert, A.M. Age- and Passage-Dependent Upregulation of Fibroblast Elastase-Type Endopeptidase Activity. Role of Advanced Glycation Endproducts, Inhibition by Fucose- and Rhamnose-Rich Oligosaccharides. Arch. Gerontol. Geriatr. 2010, 50, 327–331. [Google Scholar] [CrossRef]
  263. Fulop, T.; Khalil, A.; Larbi, A. The Role of Elastin Peptides in Modulating the Immune Response in Aging and Age-Related Diseases. Pathol. Biol. 2012, 60, 28–33. [Google Scholar] [CrossRef]
  264. Edgar, S.; Hopley, B.; Genovese, L.; Sibilla, S.; Laight, D.; Shute, J. Effects of Collagen-Derived Bioactive Peptides and Natural Antioxidant Compounds on Proliferation and Matrix Protein Synthesis by Cultured Normal Human Dermal Fibroblasts. Sci. Rep. 2018, 8, 10474. [Google Scholar] [CrossRef]
  265. Ma, C.; Su, J.; Sun, Y.; Feng, Y.; Shen, N.; Li, B.; Liang, Y.; Yang, X.; Wu, H.; Zhang, H.; et al. Significant Upregulation of Alzheimer’s β-Amyloid Levels in a Living System Induced by Extracellular Elastin Polypeptides. Angew. Chem. Int. Ed. 2019, 58, 18703–18709. [Google Scholar] [CrossRef] [Green Version]
  266. Bochicchio, B.; Lorusso, M.; Pepe, A.; Tamburro, A.M. On Enhancers and Inhibitors of Elastin-Derived Amyloidogenesis. Available online: https://www.futuremedicine.com/doi/abs/10.2217/17435889.4.1.31 (accessed on 29 December 2021).
  267. Bochicchio, B.; Pepe, A.; Delaunay, F.; Lorusso, M.; Baud, S.; Dauchez, M. Amyloidogenesis of Proteolytic Fragments of Human Elastin. RSC Adv. 2013, 3, 13273–13285. [Google Scholar] [CrossRef] [Green Version]
  268. Szychowski, K.A.; Skóra, B. Review of the Relationship between Reactive Oxygen Species (ROS) and Elastin-Derived Peptides (EDPs). Appl. Sci. 2021, 11, 8732. [Google Scholar] [CrossRef]
  269. Lehrke, M.; Lazar, M.A. The Many Faces of PPARγ. Cell 2005, 123, 993–999. [Google Scholar] [CrossRef] [Green Version]
  270. Szychowski, K.A.; Gmiński, J. Impact of Elastin-Derived VGVAPG Peptide on Bidirectional Interaction between Peroxisome Proliferator-Activated Receptor Gamma (Pparγ) and Beta-Galactosidase (β-Gal) Expression in Mouse Cortical Astrocytes in Vitro. Naunyn-Schmiedeberg’s Arch Pharm. 2019, 392, 405–413. [Google Scholar] [CrossRef] [Green Version]
  271. Gmiński, J.; Wȩglarz, L.; Dróżdż, M.; Goss, M. Pharmacological Modulation of the Antioxidant Enzymes Activities and the Concentration of Peroxidation Products in Fibroblasts Stimulated with Elastin Peptides. Gen. Pharmacol. Vasc. Syst. 1991, 22, 495–497. [Google Scholar] [CrossRef]
  272. Szychowski, K.A.; Skóra, B.; Wójtowicz, A.K. Elastin-Derived Peptides in the Central Nervous System: Friend or Foe. Cell Mol. Neurobiol. 2021, 1–15. [Google Scholar] [CrossRef]
  273. Calleri, E.; Pochetti, G.; Dossou, K.S.S.; Laghezza, A.; Montanari, R.; Capelli, D.; Prada, E.; Loiodice, F.; Massolini, G.; Bernier, M.; et al. Resveratrol and Its Metabolites Bind to PPARs. Chembiochem 2014, 15, 1154–1160. [Google Scholar] [CrossRef] [Green Version]
  274. Ma, T.; Tan, M.-S.; Yu, J.-T.; Tan, L. Resveratrol as a Therapeutic Agent for Alzheimer’s Disease. BioMed. Res. Int. 2014, 2014, e350516. [Google Scholar] [CrossRef]
  275. Nicoloff, G.; Tzvetanov, P.; Christova, P.; Baydanoff, S. Detection of Elastin Derived Peptides in Cerebrospinal Fluid of Patients with First Ever Ischaemic Stroke. Neuropeptides 2008, 42, 277–282. [Google Scholar] [CrossRef]
  276. Szychowski, K.A.; Gmiński, J. The VGVAPG Peptide Regulates the Production of Nitric Oxide Synthases and Reactive Oxygen Species in Mouse Astrocyte Cells In Vitro. Neurochem. Res. 2019, 44, 1127–1137. [Google Scholar] [CrossRef] [Green Version]
  277. van Horssen, J.; Wesseling, P.; van den Heuvel, L.P.; de Waal, R.M.; Verbeek, M.M. Heparan Sulphate Proteoglycans in Alzheimer’s Disease and Amyloid-related Disorders. Lancet Neurol. 2003, 2, 482–492. [Google Scholar] [CrossRef]
  278. Karamanos, N.K.; Piperigkou, Z.; Theocharis, A.D.; Watanabe, H.; Franchi, M.; Baud, S.; Brézillon, S.; Götte, M.; Passi, A.; Vigetti, D.; et al. Proteoglycan Chemical Diversity Drives Multifunctional Cell Regulation and Therapeutics. Chem. Rev. 2018, 118, 9152–9232. [Google Scholar] [CrossRef]
  279. Kawecki, C.; Hézard, N.; Bocquet, O.; Poitevin, G.; Rabenoelina, F.; Kauskot, A.; Duca, L.; Blaise, S.; Romier, B.; Martiny, L.; et al. Elastin-Derived Peptides Are New Regulators of Thrombosis. Arter. Thromb. Vasc. Biol. 2014, 34, 2570–2578. [Google Scholar] [CrossRef] [Green Version]
  280. Tzvetanov, P.; Nicoloff, G.; Rousseff, R.; Christova, P. Increased Levels of Elastin-Derived Peptides in Cerebrospinal Fluid of Patients with Lacunar Stroke. Clin. Neurol. Neurosurg. 2008, 110, 239–244. [Google Scholar] [CrossRef]
  281. Wang, R.; Reddy, P.H. Role of Glutamate and NMDA Receptors in Alzheimer’s Disease. J. Alzheimers Dis. 2017, 57, 1041–1048. [Google Scholar] [CrossRef] [Green Version]
  282. Liu, J.; Chang, L.; Song, Y.; Li, H.; Wu, Y. The Role of NMDA Receptors in Alzheimer’s Disease. Front. Neurosci. 2019, 13, 43. [Google Scholar] [CrossRef] [Green Version]
  283. Xia, Z.; Dudek, H.; Miranti, C.K.; Greenberg, M.E. Calcium Influx via the NMDA Receptor Induces Immediate Early Gene Transcription by a MAP Kinase/ERK-Dependent Mechanism. J. Neurosci. 1996, 16, 5425–5436. [Google Scholar] [CrossRef] [Green Version]
  284. Evans, R.C.; Morera-Herreras, T.; Cui, Y.; Du, K.; Sheehan, T.; Kotaleski, J.H.; Venance, L.; Blackwell, K.T. The Effects of NMDA Subunit Composition on Calcium Influx and Spike Timing-Dependent Plasticity in Striatal Medium Spiny Neurons. PLoS Comput. Biol. 2012, 8, e1002493. [Google Scholar] [CrossRef]
  285. Paoletti, P.; Neyton, J. NMDA Receptor Subunits: Function and Pharmacology. Curr. Opin. Pharmacol. 2007, 7, 39–47. [Google Scholar] [CrossRef] [Green Version]
  286. Paoletti, P.; Bellone, C.; Zhou, Q. NMDA Receptor Subunit Diversity: Impact on Receptor Properties, Synaptic Plasticity and Disease. Nat. Rev. Neurosci. 2013, 14, 383–400. [Google Scholar] [CrossRef]
  287. Kamat, P.K.; Rai, S.; Swarnkar, S.; Shukla, R.; Ali, S.; Najmi, A.K.; Nath, C. Okadaic Acid-Induced Tau Phosphorylation in Rat Brain: Role of NMDA Receptor. Neuroscience 2013, 238, 97–113. [Google Scholar] [CrossRef]
  288. Bezprozvanny, I.; Mattson, M.P. Neuronal Calcium Mishandling and the Pathogenesis of Alzheimer’s Disease. Trends Neurosci. 2008, 31, 454–463. [Google Scholar] [CrossRef] [Green Version]
  289. Ahmed, H.A.; Ismael, S.; Mirzahosseini, G.; Ishrat, T. Verapamil Prevents Development of Cognitive Impairment in an Aged Mouse Model of Sporadic Alzheimer’s Disease. Mol. Neurobiol. 2021, 58, 3374–3387. [Google Scholar] [CrossRef]
  290. Jomsky, M.; Kerna, N. Nifedipine: Can This Calcium Channel Blocker Be Used “Off Label” to Inhibit the Development and Symptoms of Alzheimer’s Disease and Related Beta-Amyloid-Producing Syndromes? EC Pharmacol. Toxicol. 2019. [Google Scholar] [CrossRef]
  291. Pithadia, A.S.; Lim, M.H. Metal-Associated Amyloid-β Species in Alzheimer’s Disease. Curr. Opin. Chem. Biol. 2012, 16, 67–73. [Google Scholar] [CrossRef]
  292. Greenough, M.A.; Camakaris, J.; Bush, A.I. Metal Dyshomeostasis and Oxidative Stress in Alzheimer’s Disease. Neurochem. Int. 2013, 62, 540–555. [Google Scholar] [CrossRef] [PubMed]
  293. Das, N.; Raymick, J.; Sarkar, S. Role of Metals in Alzheimer’s Disease. Metab. Brain Dis. 2021, 36, 1627–1639. [Google Scholar] [CrossRef]
  294. Tiiman, A.; Palumaa, P.; Tõugu, V. The Missing Link in the Amyloid Cascade of Alzheimer’s Disease—Metal Ions. Neurochem. Int. 2013, 62, 367–378. [Google Scholar] [CrossRef]
  295. Wang, L.; Yin, Y.-L.; Liu, X.-Z.; Shen, P.; Zheng, Y.-G.; Lan, X.-R.; Lu, C.-B.; Wang, J.-Z. Current Understanding of Metal Ions in the Pathogenesis of Alzheimer’s Disease. Transl. Neurodegener. 2020, 9, 10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  296. Strodel, B.; Coskuner-Weber, O. Transition Metal Ion Interactions with Disordered Amyloid-β Peptides in the Pathogenesis of Alzheimer’s Disease: Insights from Computational Chemistry Studies. J. Chem. Inf. Model. 2019, 59, 1782–1805. [Google Scholar] [CrossRef]
  297. Bagheri, S.; Squitti, R.; Haertlé, T.; Siotto, M.; Saboury, A.A. Role of Copper in the Onset of Alzheimer’s Disease Compared to Other Metals. Front. Aging Neurosci. 2018, 9, 446. [Google Scholar] [CrossRef]
  298. Kepp, K.P. Alzheimer’s Disease: How Metal Ions Define β-Amyloid Function. Coord. Chem. Rev. 2017, 351, 127–159. [Google Scholar] [CrossRef] [Green Version]
  299. Watt, A.D.; Villemagne, V.L.; Barnham, K.J. Metals, Membranes, and Amyloid-β Oligomers: Key Pieces in the Alzheimer’s Disease Puzzle? J. Alzheimer’s Dis. 2013, 33, S283–S293. [Google Scholar] [CrossRef]
  300. Roberts, B.R.; Ryan, T.M.; Bush, A.I.; Masters, C.L.; Duce, J.A. The Role of Metallobiology and Amyloid-β Peptides in Alzheimer’s Disease. J. Neurochem. 2012, 120, 149–166. [Google Scholar] [CrossRef]
  301. Tõugu, V.; Tiiman, A.; Palumaa, P. Interactions of Zn(Ii) and Cu(Ii) Ions with Alzheimer’s Amyloid-Beta Peptide. Metal Ion Binding, Contribution to Fibrillization and Toxicity. Metallomics 2011, 3, 250–261. [Google Scholar] [CrossRef]
  302. Dahms, S.O.; Könnig, I.; Roeser, D.; Gührs, K.-H.; Mayer, M.C.; Kaden, D.; Multhaup, G.; Than, M.E. Metal Binding Dictates Conformation and Function of the Amyloid Precursor Protein (APP) E2 Domain. J. Mol. Biol. 2012, 416, 438–452. [Google Scholar] [CrossRef] [Green Version]
  303. Kozin, S.A.; Mezentsev, Y.V.; Kulikova, A.A.; Indeykina, M.I.; Golovin, A.V.; Ivanov, A.S.; Tsvetkov, P.O.; Makarov, A.A. Zinc-Induced Dimerization of the Amyloid-β Metal-Binding Domain 1–16 Is Mediated by Residues 11–14. Mol. Biosyst. 2011, 7, 1053–1055. [Google Scholar] [CrossRef]
  304. 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]
  305. Girvan, P.; Teng, X.; Brooks, N.J.; Baldwin, G.S.; Ying, L. Redox Kinetics of the Amyloid-β-Cu Complex and Its Biological Implications. Biochemistry 2018, 57, 6228–6233. [Google Scholar] [CrossRef]
  306. Ayala, A.; Muñoz, M.F.; Argüelles, S. Lipid Peroxidation: Production, Metabolism, and Signaling Mechanisms of Malondialdehyde and 4-Hydroxy-2-Nonenal. Oxid. Med. Cell Longev. 2014, 2014, 360438. [Google Scholar] [CrossRef]
  307. Lipinski, M.M.; Zheng, B.; Lu, T.; Yan, Z.; Py, B.F.; Ng, A.; Xavier, R.J.; Li, C.; Yankner, B.A.; Scherzer, C.R.; et al. Genome-Wide Analysis Reveals Mechanisms Modulating Autophagy in Normal Brain Aging and in Alzheimer’s Disease. Proc. Natl. Acad. Sci. USA 2010, 107, 14164–14169. [Google Scholar] [CrossRef] [Green Version]
  308. Kaur, U.; Banerjee, P.; Bir, A.; Sinha, M.; Biswas, A.; Chakrabarti, S. Reactive Oxygen Species, Redox Signaling and Neuroinflammation in Alzheimer’s Disease: The NF-ΚB Connection. Curr. Top. Med. Chem. 2015, 15, 446–457. [Google Scholar] [CrossRef]
  309. Mota, S.I.; Costa, R.O.; Ferreira, I.L.; Santana, I.; Caldeira, G.L.; Padovano, C.; Fonseca, A.C.; Baldeiras, I.; Cunha, C.; Letra, L.; et al. Oxidative Stress Involving Changes in Nrf2 and ER Stress in Early Stages of Alzheimer’s Disease. Biochim. Biophys. Acta BBA Mol. Basis Dis. 2015, 1852, 1428–1441. [Google Scholar] [CrossRef] [Green Version]
  310. Patten, D.A.; Germain, M.; Kelly, M.A.; Slack, R.S. Reactive Oxygen Species: Stuck in the Middle of Neurodegeneration. J. Alzheimer’s Dis. 2010, 20, S357–S367. [Google Scholar] [CrossRef] [Green Version]
  311. Caldeira, G.L.; Ferreira, I.L.; Rego, A.C. Impaired Transcription in Alzheimer’s Disease: Key Role in Mitochondrial Dysfunction and Oxidative Stress. J. Alzheimer’s Dis. 2013, 34, 115–131. [Google Scholar] [CrossRef]
  312. Nesi, G.; Sestito, S.; Digiacomo, M.; Rapposelli, S. Oxidative Stress, Mitochondrial Abnormalities and Proteins Deposition: Multitarget Approaches in Alzheimer’s Disease. Curr. Top. Med. Chem. 2017, 17, 3062–3079. [Google Scholar] [CrossRef] [PubMed]
  313. 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] [PubMed]
  314. Uddin, M.S.; Tewari, D.; Sharma, G.; Kabir, M.T.; Barreto, G.E.; Bin-Jumah, M.N.; Perveen, A.; Abdel-Daim, M.M.; Ashraf, G.M. Molecular Mechanisms of ER Stress and UPR in the Pathogenesis of Alzheimer’s Disease. Mol. Neurobiol. 2020, 57, 2902–2919. [Google Scholar] [CrossRef] [PubMed]
  315. Lane, D.J.R.; Ayton, S.; Bush, A.I. Iron and Alzheimer’s Disease: An Update on Emerging Mechanisms. J. Alzheimer’s Dis. 2018, 64, S379–S395. [Google Scholar] [CrossRef] [PubMed]
  316. Sharma, C.; Kim, S.R. Linking Oxidative Stress and Proteinopathy in Alzheimer’s Disease. Antioxidants 2021, 10, 1231. [Google Scholar] [CrossRef] [PubMed]
  317. Zuo, L.; Hemmelgarn, B.T.; Chuang, C.-C.; Best, T.M. The Role of Oxidative Stress-Induced Epigenetic Alterations in Amyloid-β Production in Alzheimer’s Disease. Oxid. Med. Cell Longev. 2015, 2015, 604658. [Google Scholar] [CrossRef] [Green Version]
  318. Lin, H.-C.; Hsieh, H.-M.; Chen, Y.-H.; Hu, M.-L. S-Adenosylhomocysteine Increases Beta-Amyloid Formation in BV-2 Microglial Cells by Increased Expressions of Beta-Amyloid Precursor Protein and Presenilin 1 and by Hypomethylation of These Gene Promoters. Neurotoxicology 2009, 30, 622–627. [Google Scholar] [CrossRef]
  319. Sung, H.Y.; Choi, E.N.; Ahn Jo, S.; Oh, S.; Ahn, J.-H. Amyloid Protein-Mediated Differential DNA Methylation Status Regulates Gene Expression in Alzheimer’s Disease Model Cell Line. Biochem. Biophys. Res. Commun. 2011, 414, 700–705. [Google Scholar] [CrossRef]
  320. Mouton-Liger, F.; Paquet, C.; Dumurgier, J.; Bouras, C.; Pradier, L.; Gray, F.; Hugon, J. Oxidative Stress Increases BACE1 Protein Levels through Activation of the PKR-EIF2α Pathway. Biochim. Biophys. Acta 2012, 1822, 885–896. [Google Scholar] [CrossRef] [Green Version]
  321. Ma, T.; Trinh, M.A.; Wexler, A.J.; Bourbon, C.; Gatti, E.; Pierre, P.; Cavener, D.R.; Klann, E. Suppression of EIF2α Kinases Alleviates AD-Related Synaptic Plasticity and Spatial Memory Deficits. Nat. Neurosci. 2013, 16, 1299–1305. [Google Scholar] [CrossRef] [Green Version]
  322. Oliveira, M.M.; Klann, E. EIF2-Dependent Translation Initiation: Memory Consolidation and Disruption in Alzheimer’s Disease. Semin. Cell Dev. Biol. 2021. [Google Scholar] [CrossRef] [PubMed]
  323. Wang, S.-C.; Oelze, B.; Schumacher, A. Age-Specific Epigenetic Drift in Late-Onset Alzheimer’s Disease. PLoS ONE 2008, 3, e2698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  324. Scarpa, S.; Cavallaro, R.A.; D’Anselmi, F.; Fuso, A. Gene Silencing through Methylation: An Epigenetic Intervention on Alzheimer Disease. J. Alzheimers Dis. 2006, 9, 407–414. [Google Scholar] [CrossRef] [PubMed]
  325. Lithner, C.U.; Hernandez, C.; Sweatt, J.D.; Nordberg, A. O3-05-05: Epigenetic Effects of Aβ and the Implication on the Pathophysiology in Alzheimer’s Disease. Alzheimer’s Dement. 2011, 7, S508. [Google Scholar] [CrossRef]
  326. Marques, S.C.F.; Lemos, R.; Ferreiro, E.; Martins, M.; de Mendonça, A.; Santana, I.; Outeiro, T.F.; Pereira, C.M.F. Epigenetic Regulation of BACE1 in Alzheimer’s Disease Patients and in Transgenic Mice. Neuroscience 2012, 220, 256–266. [Google Scholar] [CrossRef]
  327. Chouliaras, L.; Mastroeni, D.; Delvaux, E.; Grover, A.; Kenis, G.; Hof, P.R.; Steinbusch, H.W.M.; Coleman, P.D.; Rutten, B.P.F.; van den Hove, D.L.A. Consistent Decrease in Global DNA Methylation and Hydroxymethylation in the Hippocampus of Alzheimer’s Disease Patients. Neurobiol. Aging 2013, 34, 2091–2099. [Google Scholar] [CrossRef] [Green Version]
  328. Gu, X.; Sun, J.; Li, S.; Wu, X.; Li, L. Oxidative Stress Induces DNA Demethylation and Histone Acetylation in SH-SY5Y Cells: Potential Epigenetic Mechanisms in Gene Transcription in Aβ Production. Neurobiol. Aging 2013, 34, 1069–1079. [Google Scholar] [CrossRef]
  329. Luque-Contreras, D.; Carvajal, K.; Toral-Rios, D.; Franco-Bocanegra, D.; Campos-Peña, V. Oxidative Stress and Metabolic Syndrome: Cause or Consequence of Alzheimer’s Disease? Oxidative Med. Cell. Longev. 2014, 2014, e497802. [Google Scholar] [CrossRef]
  330. Massaad, C.A. Neuronal and Vascular Oxidative Stress in Alzheimer’s Disease. Curr. Neuropharmacol. 2011, 9, 662–673. [Google Scholar] [CrossRef] [Green Version]
  331. Hamilton, A.; Holscher, C. The Effect of Ageing on Neurogenesis and Oxidative Stress in the APPswe/PS1deltaE9 Mouse Model of Alzheimer’s Disease. Brain Res. 2012, 1449, 83–93. [Google Scholar] [CrossRef]
  332. Sultana, R.; Mecocci, P.; Mangialasche, F.; Cecchetti, R.; Baglioni, M.; Butterfield, D.A. Increased Protein and Lipid Oxidative Damage in Mitochondria Isolated from Lymphocytes from Patients with Alzheimer’s Disease: Insights into the Role of Oxidative Stress in Alzheimer’s Disease and Initial Investigations into a Potential Biomarker for This Dementing Disorder. J. Alzheimer’s Dis. 2011, 24, 77–84. [Google Scholar] [CrossRef] [Green Version]
  333. Readnower, R.D.; Sauerbeck, A.D.; Sullivan, P.G. Mitochondria, Amyloid β, and Alzheimer’s Disease. Int. J. Alzheimer’s Dis. 2011, 2011, e104545. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  334. Sun, X.; Chen, W.-D.; Wang, Y.-D. β-Amyloid: The Key Peptide in the Pathogenesis of Alzheimer’s Disease. Front. Pharmacol. 2015, 6, 221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  335. Reddy, P.H.; Tripathi, R.; Troung, Q.; Tirumala, K.; Reddy, T.P.; Anekonda, V.; Shirendeb, U.P.; Calkins, M.J.; Reddy, A.P.; Mao, P.; et al. Abnormal Mitochondrial Dynamics and Synaptic Degeneration as Early Events in Alzheimer’s Disease: Implications to Mitochondria-Targeted Antioxidant Therapeutics. Biochim. Biophys. Acta BBA Mol. Basis Dis. 2012, 1822, 639–649. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  336. Sirk, D.; Zhu, Z.; Wadia, J.S.; Shulyakova, N.; Phan, N.; Fong, J.; Mills, L.R. Chronic Exposure to Sub-Lethal Beta-Amyloid (Abeta) Inhibits the Import of Nuclear-Encoded Proteins to Mitochondria in Differentiated PC12 Cells. J. Neurochem. 2007, 103, 1989–2003. [Google Scholar] [CrossRef]
  337. Crouch, P.J.; Blake, R.; Duce, J.A.; Ciccotosto, G.D.; Li, Q.X.; Barnham, K.J.; Curtain, C.C.; Cherny, R.A.; Cappai, R.; Dyrks, T.; et al. Copper-Dependent Inhibition of Human Cytochrome c Oxidase by a Dimeric Conformer of Amyloid-Beta(1-42). J. Neurosci. 2005, 25, 672–679. [Google Scholar] [CrossRef] [Green Version]
  338. Chen, J.X.; Yan, S.D. Amyloid-β-Induced Mitochondrial Dysfunction. J. Alzheimers Dis. 2007, 12, 177–184. [Google Scholar] [CrossRef]
  339. Reddy, P.H.; Beal, M.F. Amyloid Beta, Mitochondrial Dysfunction and Synaptic Damage: Implications for Cognitive Decline in Aging and Alzheimer’s Disease. Trends Mol. Med. 2008, 14, 45–53. [Google Scholar] [CrossRef] [Green Version]
  340. Ferreira, I.L.; Bajouco, L.M.; Mota, S.I.; Auberson, Y.P.; Oliveira, C.R.; Rego, A.C. Amyloid Beta Peptide 1-42 Disturbs Intracellular Calcium Homeostasis through Activation of GluN2B-Containing N-Methyl-d-Aspartate Receptors in Cortical Cultures. Cell Calcium. 2012, 51, 95–106. [Google Scholar] [CrossRef]
  341. Shirwany, N.A.; Payette, D.; Xie, J.; Guo, Q. The Amyloid Beta Ion Channel Hypothesis of Alzheimer’s Disease. Neuropsychiatr. Dis. Treat. 2007, 3, 597–612. [Google Scholar]
  342. Magi, S.; Castaldo, P.; Macrì, M.L.; Maiolino, M.; Matteucci, A.; Bastioli, G.; Gratteri, S.; Amoroso, S.; Lariccia, V. Intracellular Calcium Dysregulation: Implications for Alzheimer’s Disease. Biomed. Res. Int. 2016, 2016, 6701324. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  343. Parihar, M.S.; Brewer, G.J. Amyloid Beta as a Modulator of Synaptic Plasticity. J. Alzheimers Dis. 2010, 22, 741–763. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  344. Granold, M.; Moosmann, B.; Staib-Lasarzik, I.; Arendt, T.; Del Rey, A.; Engelhard, K.; Behl, C.; Hajieva, P. High Membrane Protein Oxidation in the Human Cerebral Cortex. Redox Biol. 2015, 4, 200–207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  345. Butterfield, D.A.; Reed, T.; Newman, S.F.; Sultana, R. Roles of Amyloid β-Peptide-Associated Oxidative Stress and Brain Protein Modifications in the Pathogenesis of Alzheimer’s Disease and Mild Cognitive Impairment. Free Radic. Biol. Med. 2007, 43, 658–677. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  346. Rothman, S.M.; Mattson, M.P. Adverse Stress, Hippocampal Networks, and Alzheimer’s Disease. Neuromol. Med. 2010, 12, 56–70. [Google Scholar] [CrossRef] [Green Version]
  347. Venkateshappa, C.; Harish, G.; Mahadevan, A.; Srinivas Bharath, M.M.; Shankar, S.K. Elevated Oxidative Stress and Decreased Antioxidant Function in the Human Hippocampus and Frontal Cortex with Increasing Age: Implications for Neurodegeneration in Alzheimer’s Disease. Neurochem. Res. 2012, 37, 1601–1614. [Google Scholar] [CrossRef] [PubMed]
  348. Müller, W.E.; Eckert, A.; Kurz, C.; Eckert, G.P.; Leuner, K. Mitochondrial Dysfunction: Common Final Pathway in Brain Aging and Alzheimer’s Disease—Therapeutic Aspects. Mol. Neurobiol. 2010, 41, 159–171. [Google Scholar] [CrossRef]
  349. Sivanandam, T.M.; Thakur, M.K. Traumatic Brain Injury: A Risk Factor for Alzheimer’s Disease. Neurosci. Biobehav. Rev. 2012, 36, 1376–1381. [Google Scholar] [CrossRef]
  350. Johnson, V.E.; Stewart, W.; Smith, D.H. Traumatic Brain Injury and Amyloid-β Pathology: A Link to Alzheimer’s Disease? Nat. Rev. Neurosci. 2010, 11, 361–370. [Google Scholar] [CrossRef] [Green Version]
  351. Yiannopoulou, K.G.; Papageorgiou, S.G. Current and Future Treatments in Alzheimer Disease: An Update. J. Cent. Nerv. Syst. Dis. 2020, 12, 1179573520907397. [Google Scholar] [CrossRef] [Green Version]
  352. Government of Canada, H.C. Drug Product Database Online Query. Available online: https://health-products.canada.ca/dpd-bdpp/info.do?lang=en&code=93258 (accessed on 28 December 2021).
  353. Government of Canada, H.C. Drug Product Database Online Query. Available online: https://health-products.canada.ca/dpd-bdpp/info.do?lang=en&code=80954 (accessed on 28 December 2021).
  354. Hasselmo, M.E. The Role of Acetylcholine in Learning and Memory. Curr. Opin. Neurobiol. 2006, 16, 710–715. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  355. Ferreira-Vieira, T.H.; Guimaraes, I.M.; Silva, F.R.; Ribeiro, F.M. Alzheimer’s Disease: Targeting the Cholinergic System. Curr. Neuropharmacol. 2016, 14, 101–115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  356. Kumar, A.; Gupta, V.; Sharma, S. Donepezil. In StatPearls; StatPearls Publishing: Treasure Island, FL, USA, 2021. [Google Scholar]
  357. Müller, T. Rivastigmine in the Treatment of Patients with Alzheimer’s Disease. Neuropsychiatr. Dis. Treat. 2007, 3, 211–218. [Google Scholar] [CrossRef] [PubMed]
  358. Razay, G.; Wilcock, G.K. Galantamine in Alzheimer’s Disease. Exp. Rev. Neurother. 2008, 8, 9–17. [Google Scholar] [CrossRef]
  359. Johnson, J.W.; Kotermanski, S.E. Mechanism of Action of Memantine. Curr. Opin. Pharm. 2006, 6, 61–67. [Google Scholar] [CrossRef]
  360. Reisberg, B.; Doody, R.; Stöffler, A.; Schmitt, F.; Ferris, S.; Möbius, H.J.; Memantine Study Group. Memantine in Moderate-to-Severe Alzheimer’s Disease. N. Engl. J. Med. 2003, 348, 1333–1341. [Google Scholar] [CrossRef]
  361. Deardorff, W.J.; Grossberg, G.T. A Fixed-Dose Combination of Memantine Extended-Release and Donepezil in the Treatment of Moderate-to-Severe Alzheimer’s Disease. Drug Des. Dev. 2016, 10, 3267–3279. [Google Scholar] [CrossRef] [Green Version]
  362. InformedHealth.org. Non-Drug Interventions for Alzheimer’s Disease; Institute for Quality and Efficiency in Health Care (IQWiG): Cologne, Germany, 2017. [Google Scholar]
  363. Non-pharmacological Treatment of Alzheimer’s|IntechOpen. Available online: https://www.intechopen.com/chapters/65852 (accessed on 28 December 2021).
  364. Bastrup, J.; Hansen, K.H.; Poulsen, T.B.G.; Kastaniegaard, K.; Asuni, A.A.; Christensen, S.; Belling, D.; Helboe, L.; Stensballe, A.; Volbracht, C. Anti-Aβ Antibody Aducanumab Regulates the Proteome of Senile Plaques and Closely Surrounding Tissue in a Transgenic Mouse Model of Alzheimer’s Disease. J. Alzheimer’s Dis. 2021, 79, 249–265. [Google Scholar] [CrossRef]
  365. Sevigny, J.; Chiao, P.; Bussière, T.; Weinreb, P.H.; Williams, L.; Maier, M.; Dunstan, R.; Salloway, S.; Chen, T.; Ling, Y.; et al. The Antibody Aducanumab Reduces Aβ Plaques in Alzheimer’s Disease. Nature 2016, 537, 50–56. [Google Scholar] [CrossRef]
  366. Leinenga, G.; Koh, W.K.; Götz, J. A Comparative Study of the Effects of Aducanumab and Scanning Ultrasound on Amyloid Plaques and Behavior in the APP23 Mouse Model of Alzheimer Disease. Alzheimers Res. 2021, 13, 76. [Google Scholar] [CrossRef]
  367. Knopman, D.S.; Jones, D.T.; Greicius, M.D. Failure to Demonstrate Efficacy of Aducanumab: An Analysis of the EMERGE and ENGAGE Trials as Reported by Biogen, December 2019. Alzheimers Dement. 2021, 17, 696–701. [Google Scholar] [CrossRef] [PubMed]
  368. Boonruamkaew, P.; Chonpathompikunlert, P.; Vong, L.B.; Sakaue, S.; Tomidokoro, Y.; Ishii, K.; Tamaoka, A.; Nagasaki, Y. Chronic Treatment with a Smart Antioxidative Nanoparticle for Inhibition of Amyloid Plaque Propagation in Tg2576 Mouse Model of Alzheimer’s Disease. Sci. Rep. 2017, 7, 3785. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  369. El Sayed, N.S.; Ghoneum, M.H. Antia, a Natural Antioxidant Product, Attenuates Cognitive Dysfunction in Streptozotocin-Induced Mouse Model of Sporadic Alzheimer’s Disease by Targeting the Amyloidogenic, Inflammatory, Autophagy, and Oxidative Stress Pathways. Oxid. Med. Cell Longev. 2020, 2020, 4386562. [Google Scholar] [CrossRef] [PubMed]
  370. Nicolakakis, N.; Aboulkassim, T.; Ongali, B.; Lecrux, C.; Fernandes, P.; Rosa-Neto, P.; Tong, X.-K.; Hamel, E. Complete Rescue of Cerebrovascular Function in Aged Alzheimer’s Disease Transgenic Mice by Antioxidants and Pioglitazone, a Peroxisome Proliferator-Activated Receptor Gamma Agonist. J. Neurosci. 2008, 28, 9287–9296. [Google Scholar] [CrossRef] [PubMed]
  371. Veurink, G.; Perry, G.; Singh, S.K. Role of Antioxidants and a Nutrient Rich Diet in Alzheimer’s Disease. Open Biol. 2020, 10, 200084. [Google Scholar] [CrossRef] [PubMed]
  372. 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]
  373. Fenech, M. Vitamins Associated with Brain Aging, Mild Cognitive Impairment, and Alzheimer Disease: Biomarkers, Epidemiological and Experimental Evidence, Plausible Mechanisms, and Knowledge Gaps. Adv. Nutr. 2017, 8, 958–970. [Google Scholar] [CrossRef] [Green Version]
  374. Antibody–Drug Conjugates: A Comprehensive Review|Molecular Cancer Research. Available online: https://mcr.aacrjournals.org/content/18/1/3 (accessed on 28 December 2021).
  375. Barroso-Sousa, R.; Tolaney, S.M. Clinical Development of New Antibody–Drug Conjugates in Breast Cancer: To Infinity and Beyond. BioDrugs 2021, 35, 159–174. [Google Scholar] [CrossRef]
  376. Li, B.T.; Smit, E.F.; Goto, Y.; Nakagawa, K.; Udagawa, H.; Mazières, J.; Nagasaka, M.; Bazhenova, L.; Saltos, A.N.; Felip, E.; et al. DESTINY-Lung01 Trial Investigators. Trastuzumab Deruxtecan in HER2-Mutant Non-Small-Cell Lung Cancer. N. Engl. J. Med. 2021, 386, 241–251. [Google Scholar] [CrossRef]
  377. Calo, C.A.; O’Malley, D.M. Antibody-Drug Conjugates for the Treatment of Ovarian Cancer. Exp. Opin. Biol. 2021, 21, 875–887. [Google Scholar] [CrossRef]
  378. Prasad, S.; Gupta, S.C.; Pandey, M.K.; Tyagi, A.K.; Deb, L. Oxidative Stress and Cancer: Advances and Challenges. Oxidative Med. Cell. Longev. 2016, 2016, e5010423. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  379. Sosa, V.; Moliné, T.; Somoza, R.; Paciucci, R.; Kondoh, H.; LLeonart, M.E. Oxidative Stress and Cancer: An Overview. Ageing Res. Rev. 2013, 12, 376–390. [Google Scholar] [CrossRef] [PubMed]
  380. Reuter, S.; Gupta, S.C.; Chaturvedi, M.M.; Aggarwal, B.B. Oxidative Stress, Inflammation, and Cancer: How Are They Linked? Free Radic. Biol. Med. 2010, 49, 1603–1616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  381. Andrisic, L.; Dudzik, D.; Barbas, C.; Milkovic, L.; Grune, T.; Zarkovic, N. Short Overview on Metabolomics Approach to Study Pathophysiology of Oxidative Stress in Cancer. Redox Biol. 2018, 14, 47–58. [Google Scholar] [CrossRef]
  382. Barrera, G. Oxidative Stress and Lipid Peroxidation Products in Cancer Progression and Therapy. ISRN Oncol. 2012, 2012, 1–21. [Google Scholar] [CrossRef] [Green Version]
  383. Thanan, R.; Oikawa, S.; Hiraku, Y.; Ohnishi, S.; Ma, N.; Pinlaor, S.; Yongvanit, P.; Kawanishi, S.; Murata, M. Oxidative Stress and Its Significant Roles in Neurodegenerative Diseases and Cancer. Int. J. Mol. Sci. 2015, 16, 193–217. [Google Scholar] [CrossRef] [Green Version]
  384. Hayes, J.D.; Dinkova-Kostova, A.T.; Tew, K.D. Oxidative Stress in Cancer. Cancer Cell 2020, 38, 167–197. [Google Scholar] [CrossRef]
  385. Zou, Z.; Chang, H.; Li, H.; Wang, S. Induction of Reactive Oxygen Species: An Emerging Approach for Cancer Therapy. Apoptosis 2017, 22, 1321–1335. [Google Scholar] [CrossRef]
  386. Chen, P.; Luo, X.; Nie, P.; Wu, B.; Xu, W.; Shi, X.; Chang, H.; Li, B.; Yu, X.; Zou, Z. CQ Synergistically Sensitizes Human Colorectal Cancer Cells to SN-38/CPT-11 through Lysosomal and Mitochondrial Apoptotic Pathway via P53-ROS Cross-Talk. Free Radic. Biol. Med. 2017, 104, 280–297. [Google Scholar] [CrossRef]
  387. Dewangan, J.; Tandon, D.; Srivastava, S.; Verma, A.K.; Yapuri, A.; Rath, S.K. Novel Combination of Salinomycin and Resveratrol Synergistically Enhances the Anti-Proliferative and pro-Apoptotic Effects on Human Breast Cancer Cells. Apoptosis 2017, 22, 1246–1259. [Google Scholar] [CrossRef]
  388. Zhao, Y.; Qu, T.; Wang, P.; Li, X.; Qiang, J.; Xia, Z.; Duan, H.; Huang, J.; Zhu, L. Unravelling the Relationship between Macroautophagy and Mitochondrial ROS in Cancer Therapy. Apoptosis 2016, 21, 517–531. [Google Scholar] [CrossRef] [PubMed]
  389. Zhang, Y.; Su, S.S.; Zhao, S.; Yang, Z.; Zhong, C.-Q.; Chen, X.; Cai, Q.; Yang, Z.-H.; Huang, D.; Wu, R.; et al. RIP1 Autophosphorylation Is Promoted by Mitochondrial ROS and Is Essential for RIP3 Recruitment into Necrosome. Nat. Commun. 2017, 8, 14329. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  390. Dixon, S.J.; Lemberg, K.M.; Lamprecht, M.R.; Skouta, R.; Zaitsev, E.M.; Gleason, C.E.; Patel, D.N.; Bauer, A.J.; Cantley, A.M.; Yang, W.S.; et al. Ferroptosis: An Iron-Dependent Form of Nonapoptotic Cell Death. Cell 2012, 149, 1060–1072. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  391. Yang, Y.; Guo, R.; Tian, X.; Zhang, Z.; Zhang, P.; Li, C.; Feng, Z. Synergistic Anti-Tumor Activity of Nimotuzumab in Combination with Trastuzumab in HER2-Positive Breast Cancer. Biochem. Biophys. Res. Commun. 2017, 489, 523–527. [Google Scholar] [CrossRef] [PubMed]
  392. Santoro, V.; Jia, R.; Thompson, H.; Nijhuis, A.; Jeffery, R.; Kiakos, K.; Silver, A.R.; Hartley, J.A.; Hochhauser, D. Role of Reactive Oxygen Species in the Abrogation of Oxaliplatin Activity by Cetuximab in Colorectal Cancer. JNCI J. Natl. Cancer Inst. 2016, 108, djv394. [Google Scholar] [CrossRef] [Green Version]
  393. Combined Oridonin with Cetuximab Treatment Shows Synergistic Anticancer Effects on Laryngeal Squamous Cell Carcinoma: Involvement of Inhibition of EGFR and Activation of Reactive Oxygen Species-Mediated JNK Pathway. Available online: https://www.spandidos-publications.com/10.3892/ijo.2016.3696 (accessed on 29 December 2021).
  394. Fack, F.; Espedal, H.; Keunen, O.; Golebiewska, A.; Obad, N.; Harter, P.N.; Mittelbronn, M.; Bähr, O.; Weyerbrock, A.; Stuhr, L.; et al. Bevacizumab Treatment Induces Metabolic Adaptation toward Anaerobic Metabolism in Glioblastomas. Acta Neuropathol. 2015, 129, 115–131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  395. Guo, X.; Li, D.; Sun, K.; Wang, J.; Liu, Y.; Song, J.; Zhao, Q.; Zhang, S.; Deng, W.; Zhao, X.; et al. Inhibition of Autophagy Enhances Anticancer Effects of Bevacizumab in Hepatocarcinoma. J. Mol. Med. 2013, 91, 473–483. [Google Scholar] [CrossRef] [Green Version]
  396. Leone, A.; Roca, M.S.; Ciardiello, C.; Terranova-Barberio, M.; Vitagliano, C.; Ciliberto, G.; Mancini, R.; Di Gennaro, E.; Bruzzese, F.; Budillon, A. Vorinostat Synergizes with EGFR Inhibitors in NSCLC Cells by Increasing ROS via Up-Regulation of the Major Mitochondrial Porin VDAC1 and Modulation of the c-Myc-NRF2-KEAP1 Pathway. Free Radic. Biol. Med. 2015, 89, 287–299. [Google Scholar] [CrossRef]
  397. Abounader, R.; Schiff, D. The Blood-Brain Barrier Limits the Therapeutic Efficacy of Antibody-Drug Conjugates in Glioblastoma. Neuro. Oncol. 2021, 23, 1993–1994. [Google Scholar] [CrossRef]
  398. Frontiers|Antibody Drug Conjugates in Glioblastoma—Is There a Future for Them?|Oncology. Available online: https://www.frontiersin.org/articles/10.3389/fonc.2021.718590/full (accessed on 28 December 2021).
  399. Banks, W.A. Characteristics of Compounds That Cross the Blood-Brain Barrier. BMC Neurol. 2009, 9, S3. [Google Scholar] [CrossRef] [Green Version]
  400. Cavaco, M.; Frutos, S.; Oliete, P.; Valle, J.; Andreu, D.; Castanho, M.A.R.B.; Vila-Perelló, M.; Neves, V. Conjugation of a Blood Brain Barrier Peptide Shuttle to an Fc Domain for Brain Delivery of Therapeutic Biomolecules. ACS Med. Chem. Lett. 2021, 12, 1663–1668. [Google Scholar] [CrossRef]
  401. Oldendorf, W.H. Brain Uptake of Radiolabeled Amino Acids, Amines, and Hexoses after Arterial Injection. Am. J. Physiol. 1971, 221, 1629–1639. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  402. Taylor, E.M. The Impact of Efflux Transporters in the Brain on the Development of Drugs for CNS Disorders. Clin. Pharm. 2002, 41, 81–92. [Google Scholar] [CrossRef] [PubMed]
  403. Deane, R.; Sagare, A.; Zlokovic, B. The Role of the Cell Surface LRP and Soluble LRP in Blood-Brain Barrier Abeta Clearance in Alzheimers Disease. Curr. Pharm. Des. 2008, 14, 1601–1605. [Google Scholar] [CrossRef]
  404. Frontiers|Relationship Between Amyloid-β Deposition and Blood–Brain Barrier Dysfunction in Alzheimer’s Disease|Cellular Neuroscience. Available online: https://www.frontiersin.org/articles/10.3389/fncel.2021.695479/full (accessed on 28 December 2021).
  405. Sarma, B.K.; Mugesh, G. Glutathione Peroxidase (GPx)-like Antioxidant Activity of the Organoselenium Drug Ebselen: Unexpected Complications with Thiol Exchange Reactions. J. Am. Chem. Soc. 2005, 127, 11477–11485. [Google Scholar] [CrossRef] [PubMed]
  406. Pearson, J.K.; Boyd, R.J. Effect of Substituents on the GPx-like Activity of Ebselen: Steric versus Electronic. J. Phys. Chem. A 2008, 112, 1013–1017. [Google Scholar] [CrossRef]
  407. Asbaghi, O.; Zakeri, N.; Rezaei Kelishadi, M.; Naeini, F.; Mirzadeh, E. Selenium Supplementation and Oxidative Stress: A Review. PharmaNutrition 2021, 17, 100263. [Google Scholar] [CrossRef]
  408. Xie, Y.; Tan, Y.; Zheng, Y.; Du, X.; Liu, Q. Ebselen Ameliorates β-Amyloid Pathology, Tau Pathology, and Cognitive Impairment in Triple-Transgenic Alzheimer’s Disease Mice. J. Biol. Inorg. Chem. 2017, 22, 851–865. [Google Scholar] [CrossRef]
  409. Luo, Z.; Sheng, J.; Sun, Y.; Lu, C.; Yan, J.; Liu, A.; Luo, H.-B.; Huang, L.; Li, X. Synthesis and Evaluation of Multi-Target-Directed Ligands against Alzheimer’s Disease Based on the Fusion of Donepezil and Ebselen. J. Med. Chem. 2013, 56, 9089–9099. [Google Scholar] [CrossRef]
  410. Terra, B.; da Silva, P.; Tramarin, A.; Franco, L.; da Cunha, E.; Macedo Junior, F.; Ramalho, T.; Bartolini, M.; Bolognesi, M.; de Fátima, Â. Two Novel Donepezil-Lipoic Acid Hybrids: Synthesis, Anticholinesterase and Antioxidant Activities and Theoretical Studies. J. Braz. Chem. Soc. 2017, 29, 738–747. [Google Scholar] [CrossRef]
  411. 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] [PubMed] [Green Version]
  412. Pérez-Cruz, K.; Moncada-Basualto, M.; Morales-Valenzuela, J.; Barriga-González, G.; Navarrete-Encina, P.; Núñez-Vergara, L.; Squella, J.A.; Olea-Azar, C. Synthesis and Antioxidant Study of New Polyphenolic Hybrid-Coumarins. Arab. J. Chem. 2018, 11, 525–537. [Google Scholar] [CrossRef]
  413. Trang, N.V.; Thuy, P.T.; Thanh, D.T.M.; Son, N.T. Benzofuran–Stilbene Hybrid Compounds: An Antioxidant Assessment—A DFT Study. RSC Adv. 2021, 11, 12971–12980. [Google Scholar] [CrossRef]
  414. Zhao, Y.; Zhao, B. Oxidative Stress and the Pathogenesis of Alzheimer’s Disease. Oxidative Med. Cell. Longev. 2013, 2013, e316523. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  415. Sano, M.; Ernesto, C.; Thomas, R.G.; Klauber, M.R.; Schafer, K.; Grundman, M.; Woodbury, P.; Growdon, J.; Cotman, C.W.; Pfeiffer, E.; et al. A Controlled Trial of Selegiline, Alpha-Tocopherol, or Both as Treatment for Alzheimer’s Disease. The Alzheimer’s Disease Cooperative Study. N. Engl. J. Med. 1997, 336, 1216–1222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  416. Petersen, R.C.; Thomas, R.G.; Grundman, M.; Bennett, D.; Doody, R.; Ferris, S.; Galasko, D.; Jin, S.; Kaye, J.; Levey, A.; et al. Alzheimer’s Disease Cooperative Study Group. Vitamin E and Donepezil for the Treatment of Mild Cognitive Impairment. N. Engl. J. Med. 2005, 352, 2379–2388. [Google Scholar] [CrossRef] [Green Version]
  417. 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, 33–44. [Google Scholar] [CrossRef]
  418. 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. Alzheimer’s Disease Cooperative Study. Antioxidants for Alzheimer Disease: A Randomized Clinical Trial with Cerebrospinal Fluid Biomarker Measures. Arch. Neurol. 2012, 69, 836–841. [Google Scholar] [CrossRef] [Green Version]
  419. Aisen, P.S.; Schneider, L.S.; Sano, M.; Diaz-Arrastia, R.; van Dyck, C.H.; Weiner, M.F.; Bottiglieri, T.; Jin, S.; Stokes, K.T.; Thomas, R.G.; et al. Alzheimer Disease Cooperative Study. High-Dose B Vitamin Supplementation and Cognitive Decline in Alzheimer Disease: A Randomized Controlled Trial. JAMA 2008, 300, 1774–1783. [Google Scholar] [CrossRef]
  420. Annweiler, C.; Fantino, B.; Parot-Schinkel, E.; Thiery, S.; Gautier, J.; Beauchet, O. Alzheimer’s Disease--Input of Vitamin D with MEmantine Assay (AD-IDEA Trial): Study Protocol for a Randomized Controlled Trial. Trials 2011, 12, 230. [Google Scholar] [CrossRef] [Green Version]
  421. Remington, R.; Bechtel, C.; Larsen, D.; Samar, A.; Doshanjh, L.; Fishman, P.; Luo, Y.; Smyers, K.; Page, R.; Morrell, C.; et al. A Phase II Randomized Clinical Trial of a Nutritional Formulation for Cognition and Mood in Alzheimer’s Disease. J. Alzheimers Dis. 2015, 45, 395–405. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  422. Zhu, C.W.; Grossman, H.; Neugroschl, J.; Parker, S.; Burden, A.; Luo, X.; Sano, M. A Randomized, Double-Blind, Placebo-Controlled Trial of Resveratrol with Glucose and Malate (RGM) to Slow the Progression of Alzheimer’s Disease: A Pilot Study. Alzheimers Dement. 2018, 4, 609–616. [Google Scholar] [CrossRef] [PubMed]
  423. 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] [Green Version]
  424. 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. 2012, 4, 43. [Google Scholar] [CrossRef] [Green Version]
  425. 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]
  426. Pilot Study to Investigate the Safety and Feasibility of Senolytic Therapy to Modulate Progression of Alzheimer’s Disease (SToMP-AD). ClinicalTrials.gov Identifier: NCT04063124. Available online: https://www.clinicaltrials.gov/ct2/show/NCT04063124 (accessed on 18 December 2021).
  427. Phase II Clinical Trial to Evaluate the Safety and Feasibility of Senolytic Therapy in Alzheimer’s Disease. ClinicalTrials.gov Identifier: NCT04685590. Available online: https://www.clinicaltrials.gov/ct2/show/NCT04685590 (accessed on 18 December 2021).
  428. Sunphenon EGCg (Epigallocatechin-Gallate) in the Early Stage of Alzheimer´s Disease. ClinicalTrials.gov Identifier: NCT00951834. Available online: https://www.clinicaltrials.gov/ct2/show/NCT00951834 (accessed on 18 December 2021).
  429. Prevention of Cognitive Decline in ApoE4 Carriers with Subjective Cognitive Decline After EGCG and a Multimodal Intervention. ClinicalTrials.gov Identifier: NCT03978052. Available online: https://www.clinicaltrials.gov/ct2/show/NCT03978052 (accessed on 18 December 2021).
  430. Vina, J. Effect of Activation of the Receptor PPARg/RXR as a Possible Treatment for Alzheimer’s Disease. Role of Genistein. Clinical Trial Registration NCT01982578; clinicaltrials.gov, 2021. Available online: https://www.clinicaltrials.gov/ct2/show/NCT01982578 (accessed on 18 December 2021).
  431. 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]
  432. 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. 2014, 38, 111–120. [Google Scholar] [CrossRef] [Green Version]
  433. Treatment with Copper in Patients with Mild Alzheimer´s Dementia—Full Text View—ClinicalTrials.gov. Available online: https://clinicaltrials.gov/ct2/show/NCT00608946 (accessed on 28 December 2021).
  434. Kryscio, R.J.; Abner, E.L.; Caban-Holt, A.; Lovell, M.; Goodman, P.; Darke, A.K.; Yee, M.; Crowley, J.; Schmitt, F.A. Association of Antioxidant Supplement Use and Dementia in the Prevention of Alzheimer’s Disease by Vitamin E and Selenium Trial (PREADViSE). JAMA Neurol. 2017, 74, 567–573. [Google Scholar] [CrossRef]
  435. Morris, M.C.; Evans, D.A.; Bienias, J.L.; Tangney, C.C.; Bennett, D.A.; Aggarwal, N.; Wilson, R.S.; Scherr, P.A. Dietary Intake of Antioxidant Nutrients and the Risk of Incident Alzheimer Disease in a Biracial Community Study. JAMA 2002, 287, 3230–3237. [Google Scholar] [CrossRef]
  436. Grundman, M.; Grundman, M.; Delaney, P. Antioxidant Strategies for Alzheimer’s Disease. Proc. Nutr. Soc. 2002, 61, 191–202. [Google Scholar] [CrossRef] [Green Version]
  437. Zandi, P.P.; Anthony, J.C.; Khachaturian, A.S.; Stone, S.V.; Gustafson, D.; Tschanz, J.T.; Norton, M.C.; Welsh-Bohmer, K.A.; Breitner, J.C.S.; Cache County Study Group. Reduced Risk of Alzheimer Disease in Users of Antioxidant Vitamin Supplements: The Cache County Study. Arch. Neurol. 2004, 61, 82–88. [Google Scholar] [CrossRef] [Green Version]
  438. 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 Abeta Levels and Amyloid Deposition in a Transgenic Model of Alzheimer’s Disease. FASEB J. 2004, 18, 323–325. [Google Scholar] [CrossRef]
  439. Nakashima, H.; Ishihara, T.; Yokota, O.; Terada, S.; Trojanowski, J.Q.; Lee, V.M.-Y.; Kuroda, S. Effects of Alpha-Tocopherol on an Animal Model of Tauopathies. Free Radic. Biol. Med. 2004, 37, 176–186. [Google Scholar] [CrossRef] [Green Version]
  440. 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]
  441. Veinbergs, I.; Mallory, M.; Sagara, Y.; Masliah, E. Vitamin E Supplementation Prevents Spatial Learning Deficits and Dendritic Alterations in Aged Apolipoprotein E-Deficient Mice. Eur. J. Neurosci. 2000, 12, 4541–4546. [Google Scholar]
  442. Ishihara, Y.; Itoh, K.; Mitsuda, Y.; Shimada, T.; Kubota, T.; Kato, C.; Song, S.Y.; Kobayashi, Y.; Mori-Yasumoto, K.; Sekita, S.; et al. Involvement of Brain Oxidation in the Cognitive Impairment in a Triple Transgenic Mouse Model of Alzheimer’s Disease: Noninvasive Measurement of the Brain Redox State by Magnetic Resonance Imaging. Free Radic. Res. 2013, 47, 731–739. [Google Scholar] [CrossRef]
  443. Moore, J.J.; Saadabadi, A. Selegiline. In StatPearls; StatPearls Publishing: Treasure Island, FL, USA, 2021. [Google Scholar]
  444. Filip, V.; Kolibás, E. Selegiline in the Treatment of Alzheimer’s Disease: A Long-Term Randomized Placebo-Controlled Trial. Czech and Slovak Senile Dementia of Alzheimer Type Study Group. J. Psychiatry Neurosci. 1999, 24, 234–243. [Google Scholar]
  445. Birks, J.; Flicker, L. Selegiline for Alzheimer’s Disease. Cochrane Database Syst. Rev. 2003, CD000442. [Google Scholar] [CrossRef]
  446. Calderón-Ospina, C.A.; Nava-Mesa, M.O. B Vitamins in the Nervous System: Current Knowledge of the Biochemical Modes of Action and Synergies of Thiamine, Pyridoxine, and Cobalamin. CNS Neurosci. 2019, 26, 5–13. [Google Scholar] [CrossRef] [Green Version]
  447. Kim, H.; Kim, G.; Jang, W.; Kim, S.Y.; Chang, N. Association between Intake of B Vitamins and Cognitive Function in Elderly Koreans with Cognitive Impairment. Nutr. J. 2014, 13, 118. [Google Scholar] [CrossRef] [Green Version]
  448. Morris, M.S. Homocysteine and Alzheimer’s Disease. Lancet Neurol. 2003, 2, 425–428. [Google Scholar] [CrossRef]
  449. Yazdi, D.S.; Bar-Yosef, D.L.; Adsi, H.; Kreiser, T.; Sigal, S.; Bera, S.; Zaguri, D.; Shaham-Niv, S.; Oluwatoba, D.S.; Levy, D.; et al. Homocysteine Fibrillar Assemblies Display Cross-Talk with Alzheimer’s Disease β-Amyloid Polypeptide. Proc. Natl. Acad. Sci. USA 2021, 118, e2017575118C. [Google Scholar] [CrossRef]
  450. Smith, A.D.; Refsum, H.; Bottiglieri, T.; Fenech, M.; Hooshmand, B.; McCaddon, A.; Miller, J.W.; Rosenberg, I.H.; Obeid, R. Homocysteine and Dementia: An International Consensus Statement1. J. Alzheimers Dis. 2018, 62, 561–570. [Google Scholar] [CrossRef] [Green Version]
  451. Seshadri, S.; Beiser, A.; Selhub, J.; Jacques, P.F.; Rosenberg, I.H.; D’Agostino, R.B.; Wilson, P.W.F.; Wolf, P.A. Plasma Homocysteine as a Risk Factor for Dementia and Alzheimer’s Disease. N. Engl. J. Med. 2002, 346, 476–483. [Google Scholar] [CrossRef]
  452. Strain, J.J.; Dowey, L.; Ward, M.; Pentieva, K.; McNulty, H. B-Vitamins, Homocysteine Metabolism and CVD. Proc. Nutr. Soc. 2004, 63, 597–603. [Google Scholar] [CrossRef] [Green Version]
  453. Remington, R.; Chan, A.; Paskavitz, J.; Shea, T.B. Efficacy of a Vitamin/Nutriceutical Formulation for Moderate-Stage to Later-Stage Alzheimer’s Disease: A Placebo-Controlled Pilot Study. Am. J. Alzheimers Dis. Other Demen. 2009, 24, 27–33. [Google Scholar] [CrossRef]
  454. Liu, C.-C.; Kanekiyo, T.; Xu, H.; Bu, G. Apolipoprotein E and Alzheimer Disease: Risk, Mechanisms, and Therapy. Nat. Rev. Neurol. 2013, 9, 106–118. [Google Scholar] [CrossRef] [Green Version]
  455. Lobo, V.; Patil, A.; Phatak, A.; Chandra, N. Free Radicals, Antioxidants and Functional Foods: Impact on Human Health. Pharm. Rev. 2010, 4, 118–126. [Google Scholar] [CrossRef] [Green Version]
  456. Prasad, K.N.; Hovland, A.R.; Cole, W.C.; Prasad, K.C.; Nahreini, P.; Edwards-Prasad, J.; Andreatta, C.P. Multiple Antioxidants in the Prevention and Treatment of Alzheimer Disease: Analysis of Biologic Rationale. Clin. Neuropharmacol. 2000, 23, 2–13. [Google Scholar] [CrossRef]
  457. Salehi, B.; Martorell, M.; Arbiser, J.L.; Sureda, A.; Martins, N.; Maurya, P.K.; Sharifi-Rad, M.; Kumar, P.; Sharifi-Rad, J. Antioxidants: Positive or Negative Actors? Biomolecules 2018, 8, 124. [Google Scholar] [CrossRef] [Green Version]
  458. Figueira, I.; Garcia, G.; Pimpão, R.C.; Terrasso, A.P.; Costa, I.; Almeida, A.F.; Tavares, L.; Pais, T.F.; Pinto, P.; Ventura, M.R.; et al. Polyphenols Journey through Blood-Brain Barrier towards Neuronal Protection. Sci. Rep. 2017, 7, 11456. [Google Scholar] [CrossRef] [PubMed]
  459. Hall, E.D.; Andrus, P.K.; Smith, S.L.; Fleck, T.J.; Scherch, H.M.; Lutzke, B.S.; Sawada, G.A.; Althaus, J.S.; Vonvoigtlander, P.F.; Padbury, G.E.; et al. Pyrrolopyrimidines: Novel Brain-Penetrating Antioxidants with Neuroprotective Activity in Brain Injury and Ischemia Models. J. Pharm. Exp. 1997, 281, 895–904. [Google Scholar]
  460. Agus, D.B.; Gambhir, S.S.; Pardridge, W.M.; Spielholz, C.; Baselga, J.; Vera, J.C.; Golde, D.W. Vitamin C Crosses the Blood-Brain Barrier in the Oxidized Form through the Glucose Transporters. J. Clin. Investig. 1997, 100, 2842–2848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  461. Spector, R.; Johanson, C.E. Vitamin Transport and Homeostasis in Mammalian Brain: Focus on Vitamins B and E. J. Neurochem. 2007, 103, 425–438. [Google Scholar] [CrossRef]
  462. Figueira, I.; Tavares, L.; Jardim, C.; Costa, I.; Terrasso, A.P.; Almeida, A.F.; Govers, C.; Mes, J.J.; Gardner, R.; Becker, J.D.; et al. Blood–Brain Barrier Transport and Neuroprotective Potential of Blackberry-Digested Polyphenols: An in Vitro Study. Eur. J. Nutr. 2019, 58, 113–130. [Google Scholar] [CrossRef] [PubMed]
  463. Milbury, P.E.; Kalt, W. Xenobiotic Metabolism and Berry Flavonoid Transport across the Blood-Brain Barrier. J. Agric. Food Chem. 2010, 58, 3950–3956. [Google Scholar] [CrossRef]
  464. Teleanu, D.M.; Negut, I.; Grumezescu, V.; Grumezescu, A.M.; Teleanu, R.I. Nanomaterials for Drug Delivery to the Central Nervous System. Nanomaterials 2019, 9, 371. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  465. Klyachko, N.L.; Manickam, D.S.; Brynskikh, A.M.; Uglanova, S.V.; Li, S.; Higginbotham, S.M.; Bronich, T.K.; Batrakova, E.V.; Kabanov, A.V. Cross-Linked Antioxidant Nanozymes for Improved Delivery to CNS. Nanomedicine 2012, 8, 119–129. [Google Scholar] [CrossRef] [Green Version]
  466. Khalil, I.; Yehye, W.A.; Etxeberria, A.E.; Alhadi, A.A.; Dezfooli, S.M.; Julkapli, N.B.M.; Basirun, W.J.; Seyfoddin, A. Nanoantioxidants: Recent Trends in Antioxidant Delivery Applications. Antioxidants 2019, 9, 24. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic representation of enzymatic antioxidant mechanisms of action. Created with BioRender.com. SOD, superoxide dismutase; CAT, catalase; GPx, glutathione peroxidase; GR, glutathione reductase; G6PDH, glucose-6-phosphate dehydrogenase.
Figure 1. Schematic representation of enzymatic antioxidant mechanisms of action. Created with BioRender.com. SOD, superoxide dismutase; CAT, catalase; GPx, glutathione peroxidase; GR, glutathione reductase; G6PDH, glucose-6-phosphate dehydrogenase.
Antioxidants 11 00213 g001
Figure 2. Chemical structure of flavonoid polyphenols. Created with BioRender.com. The flavonoid subclass of polyphenols is further classified into six main groups: anthocyanins, flavan-3-ols, flavonols, flavone, flavanones and isoflavones. Differences in chemical structure arise based on variations in the placement and number of hydroxyl groups and unsaturated bonds.
Figure 2. Chemical structure of flavonoid polyphenols. Created with BioRender.com. The flavonoid subclass of polyphenols is further classified into six main groups: anthocyanins, flavan-3-ols, flavonols, flavone, flavanones and isoflavones. Differences in chemical structure arise based on variations in the placement and number of hydroxyl groups and unsaturated bonds.
Antioxidants 11 00213 g002
Figure 3. Schematic diagram of classification of natural antioxidants. Created with BioRender.com. CAT, catalase; SOD, superoxide dismutase; GPx, glutathione peroxidase; GR, glutathione reductase; G6PDH, glucose-6-phosphate dehydrogenase.
Figure 3. Schematic diagram of classification of natural antioxidants. Created with BioRender.com. CAT, catalase; SOD, superoxide dismutase; GPx, glutathione peroxidase; GR, glutathione reductase; G6PDH, glucose-6-phosphate dehydrogenase.
Antioxidants 11 00213 g003
Table 1. Some neurodegenerative diseases and conditions associated with oxidative stress.
Table 1. Some neurodegenerative diseases and conditions associated with oxidative stress.
Disease or Condition
Alzheimer’s disease [4,5]
Amyotrophic lateral sclerosis [6,7]
Corticobasal degeneration [8]
Creutzfeldt-Jakob disease (Prion disease) [9]
Down syndrome [10]
Diabetic neuropathy [11,12]
Friedreich’s ataxia [13]
Huntington’s disease [14,15,16]
Lewy body disease [17]
Multiple sclerosis [18,19]
Neiman-Pick C disease [20,21]
Neuromyelitis optica [22]
Parkinson’s disease [23,24,25]
Progressive supranuclear palsy [26]
Spinocerebellar ataxia [27,28,29]
Stroke [30,31,32]
Traumatic brain injury [33,34,35]
Table 2. Summary of clinical trials utilizing combination/conjugate antioxidants as preventative therapy or treatment in AD. ADCS-ADL: Alzheimer’s Disease Cooperative Study—Activity of Daily Living; ADAS-Cog: Alzheimer’s Disease Assessment Scale (Cog: Cognitive score).
Table 2. Summary of clinical trials utilizing combination/conjugate antioxidants as preventative therapy or treatment in AD. ADCS-ADL: Alzheimer’s Disease Cooperative Study—Activity of Daily Living; ADAS-Cog: Alzheimer’s Disease Assessment Scale (Cog: Cognitive score).
ClassificationCompound(s)ParticipantsInterventionPrimary Outcome
Measures
Main ResultsIn-Text
Reference
VitaminsVitamin E + Selegiline341 patients with moderate AD2000 IU vitamin E, 10 mg selegiline, both or placebo daily for 2 yearsTime until occurrence of death, institutionalization, loss of ability to perform activities of daily living, or severe dementiaTreatment with vitamin E or selegiline slowed the progression of disease in patients with moderately severe impairment from AD[415]
Vitamin E + Donepezil790 patients with mild cognitive impairment (MCI)2000 IU vitamin E, 10 mg donepezil or placebo, daily for 3 yearsClinically possible or
probable AD
Vitamin E had no benefit. Donepezil was associated with a lower rate of progression in first 12 months[416]
Vitamin E + Memantine613 patients with mild to moderate AD2000 IU vita-min E, 20 mg memantine, both or placebo daily for 5 yearsADCS-ADL2000 IU/day of vitamin E compared to placebo slowed functional decline. No difference between groups receiving memantine alone or memantine + vitamin E[417]
Vitamin E + Vitamin C +
Alpha-Lipoic Acid
75 patients with mild to moderate AD800 IU vitamin E + 500 mg vitamin C + 900 mg alpha-lipoic acid, 400 mg coenzyme Q10 3 times/day or placebo daily for 16 weeksChanges in cerebral spinal fluid (CSF) biomarkers related to AD and oxidative stress, cognition and functionAntioxidants did not influence CSF biomarkers related to amyloid or tau pathology[418]
B Vitamins340 patients with mild to moderate AD5 mg folate + 25 mg vitamin B6 + 1 mg vitamin B12 or placebo daily for 18 monthsChanges in the cognitive subscale of the ADAS-CogRegimen of high-dose B vitamin supplements does not slow cognitive decline in individuals with mild to moderate AD[419]
Vitamin D + Memantine90 patients with moderate AD100,000 IU vitamin D3 (every 4 weeks) + 20 mg memantine or placebo daily for 24 weeksChange of cognitive performanceOngoing[420]
Multivitamin135 patients with AD or MCINutraceutical formulation (NF) of: 400 ug folic acid, 6 ug vitamin B12, 30 IU vitamin E, 400 mg S-adenosylmethionine, 600 mg N-acetyl cysteine, 500 mg acetyl-l-carnitine daily for 1 yearCognitive improvement or maintenance of cognitive performanceNF maintained or improved cognitive performance and mood/behaviour[421]
PolyphenolsResveratrol39 patients with mild to moderate AD5 mg resveratrol + 5 mg dextrose + 5 mg malate or placebo twice daily for 1 yearEvaluate the safety, tolerability and efficacy of resveratrol, glucose and malate in slowing the progression of ADLow-dose resveratrol is safe and well-tolerated[422]
Resveratrol119 patients with mild to moderate ADUp to 1 mg resveratrol twice daily or placebo for 52 weeksSafety and tolerability of treatment with resveratrol and change in ADLResveratrol decreases CSF biomarkers, modulates neuro-inflammation and induces adaptive immunity[423]
Curcumin36 patients with mild to moderate AD2 g or 4 g curcumin or placebo daily for 24 weeksExamine safety and tolerability of curcumin and determine its side effects on patientsCurcumin well-tolerated. Unable to demonstrate clinical or biochemical evidence of efficacy of curcumin C3 complex. Data suggest limited bioavailability[424]
Curcumin36 patients with dementia, presumed AD1 g curcumin + 120 mg ginkgo leaf extract, 4 g curcumin + ginkgo leaf extract or placebo daily for 6 monthsChange in isoprostane levels in plasma and change in beta-amyloid levels in serumSerum beta-amyloid rose on curcumin. Fewer adverse events reported[425]
Quercetin48 patients with MCI or early AD1000 mg quercetin or 100 mg dasatinib or placebo daily for 2 daysSerious adverse events and adverse events, and change in cellular senescence blood markersOngoing[426]
QuercetinRecruiting
patients with early AD
Quercetin + dasatinib for 2 days on, 14 days off for 12 weeks (6 cycles)Brain penetrance of dasatinib and quercetinOngoing[427]
EGCG21 patients with early AD200 mg, 400 mg, 600 mg and 800 mg EGCG tri-monthly or placebo for 18 monthsADAS-CogOngoing[428]
EGCG200 patients with AD carrying ApoE4
allele
260–520 mg EGCG + personalized intervention or placebo + non personalized intervention or 260–520 mg EGCG + non personalized intervention or placebo to personalized intervention, daily for 15 monthsEvaluate the efficacy of multimodal intervention
(dietary, physical and
cognition) combined with EGCG in slowing down
cognitive decline
Ongoing[429]
Genistein27 patients with mild AD60 mg genistein or placebo daily for 360 daysChanges in amyloid beta concentration of CSFOngoing[430]
Genistein + Daidzein72 patients with AD100 mg of soy isoflavones or placebo daily for 6 monthsCognitive outcomes: language execution function, verbal memory and recall, attention, visual memory and planningSix months of 100 mg/day isoflavones did not benefit cognition in older men and women with AD[431]
Alpha-Lipoic Acid + Omega-3 Fatty Acids39 patients with mild AD600 mg alpha-lipoic acid + 3 g fish oil, 3 g fish oil alone or placebo daily for 12 monthsPeripheral F2-isoprostane levels (oxidative stress measure)Combination of
alpha lipoic acid with omega-3 fatty acids slowed cognitive and functional decline
[432]
MineralsCopper68 patients with mild to moderate AD8 mg copper or placebo daily for 1 yearChange in cognitive function, measured by ADAS-CogResults not
yet published
[433]
Selenium +
Vitamin E
7540 participants with dementia200 μg Selenium + 400 IU Vitamin E, 200 μg selenium + placebo or 400 IU vitamin E + placebo or placebo + placebo daily for 7–12 yearsIncidence of dementia
(including AD)
Neither supplement prevented dementia[434]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Collins, A.E.; Saleh, T.M.; Kalisch, B.E. Naturally Occurring Antioxidant Therapy in Alzheimer’s Disease. Antioxidants 2022, 11, 213. https://doi.org/10.3390/antiox11020213

AMA Style

Collins AE, Saleh TM, Kalisch BE. Naturally Occurring Antioxidant Therapy in Alzheimer’s Disease. Antioxidants. 2022; 11(2):213. https://doi.org/10.3390/antiox11020213

Chicago/Turabian Style

Collins, Andrila E., Tarek M. Saleh, and Bettina E. Kalisch. 2022. "Naturally Occurring Antioxidant Therapy in Alzheimer’s Disease" Antioxidants 11, no. 2: 213. https://doi.org/10.3390/antiox11020213

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