Next Article in Journal
Estrogen Receptor Beta (ERβ) Maintains Mitochondrial Network Regulating Invasiveness in an Obesity-Related Inflammation Condition in Breast Cancer
Next Article in Special Issue
The Emerging Scenario of the Gut–Brain Axis: The Therapeutic Actions of the New Actor Kefir against Neurodegenerative Diseases
Previous Article in Journal
Metabolic Alterations Identified in Urine, Plasma and Aortic Smooth Muscle Cells Reflect Cardiovascular Risk in Patients with Programmed Coronary Artery Bypass Grafting
Previous Article in Special Issue
Failure of the Brain Glucagon-Like Peptide-1-Mediated Control of Intestinal Redox Homeostasis in a Rat Model of Sporadic Alzheimer’s Disease
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Potential Role of Polyphenols in Oxidative Stress and Inflammation Induced by Gut Microbiota in Alzheimer’s Disease

Department of Food Science and Biotechnology, College of Agriculture and Life Sciences, Kangwon National University, Chuncheon 200-701, Korea
*
Author to whom correspondence should be addressed.
Antioxidants 2021, 10(9), 1370; https://doi.org/10.3390/antiox10091370
Submission received: 12 August 2021 / Revised: 23 August 2021 / Accepted: 25 August 2021 / Published: 27 August 2021

Abstract

:
Gut microbiota (GM) play a role in the metabolic health, gut eubiosis, nutrition, and physiology of humans. They are also involved in the regulation of inflammation, oxidative stress, immune responses, central and peripheral neurotransmission. Aging and unhealthy dietary patterns, along with oxidative and inflammatory responses due to gut dysbiosis, can lead to the pathogenesis of neurodegenerative diseases, especially Alzheimer’s disease (AD). Although the exact mechanism between AD and GM dysbiosis is still unknown, recent studies claim that secretions from the gut can enhance hallmarks of AD by disturbing the intestinal permeability and blood–brain barrier via the microbiota–gut–brain axis. Dietary polyphenols are the secondary metabolites of plants that possess anti-oxidative and anti-inflammatory properties and can ameliorate gut dysbiosis by enhancing the abundance of beneficial bacteria. Thus, modulation of gut by polyphenols can prevent and treat AD and other neurodegenerative diseases. This review summarizes the role of oxidative stress, inflammation, and GM in AD. Further, it provides an overview on the ability of polyphenols to modulate gut dysbiosis, oxidative stress, and inflammation against AD.

1. Introduction

The imbalance between oxidants and antioxidants in living organisms that occurs due to the inappropriate functioning of the antioxidant system or excess level of reactive oxygen species (ROS)/reactive nitrogen species (RNS) is known as oxidative stress [1]. On the other hand, inflammation is a complex set of interactions between cells and soluble factors. It arises in any tissue as a protective and adaptive response of the innate immune system during injury to re-establish the homeostasis of damaged tissues [2,3]. The proper regulation of the inflammation mechanism is necessary to avoid uncontrolled amplification and prevent the change from the normal tissue repair toward diseases onset and collateral damage [4]. An uncontrolled generation of reactive species triggers the production of more highly reactive species (a condition of oxidative stress) and ensuing perpetuation of inflammation. The excessive reactive species can damage the structure of DNA, lipids, and protein and can lead to aging [5]. In addition, it can promote cell death that activates necrosis, apoptosis, and extracellular matrix breakdown and releases various intracellular and extracellular factors to hyperactivate the inflammatory cascade, resulting in increased oxidative stress and free radical production in a vicious circle [6]. Both oxidative stress and inflammation give rise to the etiopathogenesis of many chronic disorders including cancer, diabetes, metabolic syndromes, and cardiovascular and neurodegenerative diseases [7]. However, under normal physiological conditions, free radicals and inflammation are important for the prevention of chronic degenerative diseases and the maintenance of human well-being. In addition, ROS and RNS take part in the regulation of many molecular pathways such as differentiation, metabolism, survival, proliferation, and iron homeostasis [4,5].
Human intestines comprise an intricate ecological colony of microorganisms known as gut microbiota (GM), and approximately 100 trillion microorganisms live in the human gut [8]. Recently, research has been shifted from the diversity and abundance of GM to their functional aspects. Shabbir et al. [9] stated that GM take part in metabolic, neural, immune, and defence mechanisms and have an impact on the host in health and disease. Moreover, GM ferment resistant starch and dietary fibres, releasing short-chain fatty acids (SCFA: acetate, butyrate, and propionate), gamma-aminobutyric acid, serotonin, dopamine, glutamate, and acetylcholine. The impairment of GM composition, known as dysbiosis, can lead to several metabolic disorders such as ulcerative colitis, type 2 diabetes, obesity, colorectal cancer, and cardiometabolic and metabolic liver disease. Other than metabolic diseases, dysbiosis also contributes to neurological disorders, including anxiety, bipolar disorder, depression, obsessive-compulsive disorder, epilepsy, and Parkinson’s and Alzheimer’s disease (AD) through the microbiota–gut–brain axis (MGBX). However, the actual relationship and mechanism between gut dysbiosis and neurodegeneration is elusive [10].
Studies are revealing that complex interactions between GM influence oxidative stress and inflammation and resists counter-regulatory mechanisms of antioxidants. Nutritional interventions can be an effective approach to treat gut dysbiosis. Additionally, dietary polyphenols and their metabolites (via metabolism of GM) may regulate the oxidative and inflammation state of the central nervous system and can be potent agents against AD and other neurodegenerative diseases. The current paper offers an overview on the role of oxidative stress, inflammation, and GM in the pathogenesis of neurodegenerative diseases, especially AD. Additionally, the potential implications of polyphenols on GM modulation to reduce oxidative stress and inflammation to prevent and treat AD are addressed.

2. Inflammation and Oxidative Stress

Inflammation and oxidative stress are closely associated in the pathophysiological events where redox homeostasis (endogenous capacity of cells to deal with challenges that generate electrophiles [10] perpetually) is disrupted due to the imbalance of oxidants and reductants [11]. The leading factors that enhance chronic inflammation are the uncontrolled production of pro-inflammatory cytokines, oxidative stress, chronic infections, and alterations in the metabolism of adipose tissues. The NADPH oxidases (NOXs) and mitochondria are the primary cellular sources of ROS throughout the mitochondrial electron transport chain. Moreover, Complexes I and III of the electron transport chain are the main source of ROS production in mitochondria [12]. Parra-Ortiz et al. [13] stated that oxidative stress instigates several modifications in lipids that generate oxidized- specific products (e.g., oxidized low-density lipoprotein or cholesteryl-esters that stimulate macrophages via toll-like receptor-4 (TLR4) and spleen tyrosine kinase) that excite inflammation and induce immune responses [14]. Further, generation of ROS in adipocytes perpetuates chronic inflammation and stimulates pro-inflammatory adipokines in the target tissue [13]. Additionally, modulation of macrophages activities due to the bioenergetics and metabolic alteration increase phospholipid oxidation in tissues that leads to the modification of membrane properties and stimulate inflammation [15]. Moreover, secretions from activated macrophages such as interleukin (IL)-6, tumor necrosis factor-α (TNF-α), and other pro-inflammatory molecules (such as NO, NO synthase, cyclooxygenase-2, and ROS) can damage DNA via oxidation [16].
Regarding molecular mechanisms, Battino et al. [7] reported that ROS activate redox-sensitive transcription factors, activator protein-1, their up-regulating kinases (especially posphoinositide 3-Kinase, extracellular signal-regulated kinases, c-Jun N-terminal kinase, and mitogen-activated protein kinases), and nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) have a significant contribution in the pro-inflammatory responses. Studies have been reported that NF-κB, chemokines, and pro-inflammatory cytokines (e.g., IL-1β, IL-6, IL-2, IL-12, and TNF-α) recruit the neutrophils and the macrophages to the inflammation site, reinforcing the formation of oxidative species by neutrophils and macrophages which lead to the inflammation [17]. Moreover, Raucci et al. [6] and Shah et al. [18] stated that endoplasmic reticulum and mitochondrial dysfunction might be due to the excessive production of ROS that activates necrosis and apoptosis. Due to the release of high mobility group box-1 (HMGB1) through various receptors of the TLR4-dependent pathway, the necrotic tissues are responsible for the inflammation. HMGB1 as a representative damage associated molecular pattern (DAMP) protein has been documented to be involved in inflammatory diseases related to brain including stroke, epilepsy, traumatic brain injury, and hypoxic-ischemic brain injury [19,20]. Hatayama and Stonestreet et al. [19] revealed that HMGB1 translocate in damaged neurons from nucleus to cytoplasm and is released to extracellular as DAMP. They bind to receptors such as TLR4 or receptors for advanced glycation end products on astrocytes and activate them, releasing ROS, pro-inflammatory cytokines, matrix metalloproteinases, and chemokines, resulting in endothelial activation neutrophil attraction and damaging the blood-brain barrier (BBB). Neutrophils in blood vessels of the brain are activated and migrate to brain parenchyma through the damaged BBB. The activated neurological cells and migrated neutrophils release ROS, NO, and pro-inflammatory cytokines that lead to neural cell death. Figure 1 represents the role of oxidative stress is neuroinflammation and neurodegenration.
Furthermore, HMGB1 with extracellular ATP, phagocytosis, NOX, Cathepsin B, and phagolysosomes disruption activate nucleotide-binding oligomerization domain leucine-rich repeat containing protein 3 (NLRP3) inflammasome, which enhances ROS production, thioredoxin, thioredoxin-interacting protein, and spark inflammasome activating signals that increase the agglomeration of inflammasome. The NLRP3 inflammasome also contributes to systematic inflammation and increases age-related diseases (especially neurodegenerative diseases) [21,22].

3. GM, Oxidative Stress and Inflammation

GM play several roles in the host, such as immune responses (as independent production of IgA antibodies, induction of T cell-dependent, promotion of IL-10, and mucosal Th17 cell response from intestinal macrophages), protection against pathogen colonization, and intestinal epithelial barrier protection. Among other functions of the GM, the production/regulation of oxidative stress is the most interesting one. It has been reported that the epithelial lining of the gut and other cell types in the presence of microbiota generate ROS. Additionally, intestinal tissues, commensal anaerobes, and leukocytes are a rich source of NO (the neurotransmitter of the non-cholinergic and non-adrenergic nervous system that exerts a neuroprotective function). Gut bifidobacteria and lactobacilli convert nitrite and nitrate in NO and increase the release of NO by host epithelial cells [23]. In addition, gut bacilli and streptomycetes produce NO via NO synthetase from L-arginine. Aberrant production of NO generates ROS associated with cellular damage, neuroinflammation, neurodegenerative disorders, and axonal degeneration [24]. Moreover, Salmonella, E. coli, and other bacteria break sulphur amino acids and produce hydrogen sulphide in the gastrointestinal tract (GIT). Higher levels of hydrogen sulphide inhibit cyclooxygenase activity, shift the metabolism towards glycolysis, increase lactate, decrease ATP production, and decrease mitochondrial oxygen consumption and overexpression of pro-inflammatory effects [25]. Other than that, He et al. [26] disclosed that trimethylamine N-oxide (TMAO; GM metabolite) is involved in oxidative stress and associated with aging and exhibited increased plasma levels of monocyte chemoattractant protein-1, IL-1β, and TNF-α, along with higher and lower plasma concentration of malondialdehyde and glutathione peroxidase/superoxide dismutase activities (implying oxidative stress). Loffredo et al. [27] and Kesika et al. [28] revealed that gram-negative bacteria (such as E. coli and Shigella) increase the production of amyloids and lipopolysaccharides (LPS) that induce local systematic inflammation and give rise to dysfunction in the permeability of GIT and BBB function during dysbiosis. GM dysbiosis increases pro-inflammatory bacteria such as Escerchia/Shigella, Verrucomicrobia, Pseudomonas aeruginosa and Proteobacteria and decrease the anti-inflammatory bacteria such as Bifidobacterium, Bacteroides fragilis, Eubacterium hallii, Eubacterium rectale, Bacillus fragilis, and Faecalibacterium prausnitzii that promote inflammation and contribute to neurodegenration [29]. Table 1 represents the neurodegenerative diseases induced by oxidative stress and inflammation due to gut dysbiosis.

4. Alzheimer’s Disease

Dementia is a general term for loss of memory, thinking ability, language, judgement, and behaviour that can deteriorate daily life activities [44]. According to Alzheimer’s Disease International, someone in the world develops dementia in every 3 s. About 50 million people have dementia globally, and this figure is expected to double in the next 20 years. Low and middle-income countries suffer the most and have around 60% of cases, which is supposed to increase (71%) by 2050 [45]. AD is the most progressive disease of the brain, comprising about 60–80% of cases of dementia and posing difficulties for families and society and a severe burden on the economy [46]. People suffering from AD may have difficulty in remembering names and recent conversions, and can have anxiety or depression in the early stages. The conditions continue to worsen over the years, leading to confusion, behavioral changes, disorientation, and ultimately facing problems in speaking, walking, swallowing, and needing extensive care [47]. The development of amyloid-beta (Aβ) plaques (Aβ-oligomers and Aβ peptides), neurofibrillary tangles, oxidative stress, neuroinflammation in the nerve cells, mitochondrial dysfunction, and insulin resistance are the hallmarks of AD [46,48,49].

4.1. Microbiota-Gut-Brain Axis and AD

More than 2000 clinical trials have targeted Aβ plaques, neurofibrillary tangles, and other biomarkers but have been failed to treat AD [50]. Thereby, recent findings claim that MGBX is the bidirectional pathway that communicates through vagal and spinal nerves between gut and brain via endocrine, immune, metabolic, and neural pathways (Figure 2) and take part in the pathophysiology of AD [51]. The pro-inflammatory cytokines and bacterial metabolites (TMAO, SCFA, amyloids, LPS, and peptidoglycans) can enter into circulation via leaky gut, and can reach the brain and contribute to brain aging and cognitive decline [52,53]. Furthermore, they can interfere with Aβ1-40 and Aβ1-42 peptide interactions and hyperphosphorylation of tau, and activate glial cells leading to neurotoxic Aβ plaque formation, neuroinflammation, and neuronal degradation [54,55].

4.2. Oxidative Stress, Inflammation and AD: The Role of GM

Although AD is a neurodegenerative disease, preclinical and clinical studies evidently suggest altering GM is linked with AD development. The involvement of oxidative stress in the key events to initiate neural loss is clear, but determination of the immediate role of oxidative stress in the neurodegeneration process is still elusive. Markers of lipid peroxidation and high levels of protein oxidation markers (e.g., carbonyl) have been detected in both AD animal and human studies [27,56]. In this context, eubiosis in GM composition can exhibit a positive role in the reduction of reactive species through SCFA such as butyrate, while dysbiosis may contribute to systematic inflammation, activation of microglia, and BBB damage [46]. Moreover, trimethylamine is metabolized by GM, then conveyed to the liver and broken into TMAO upon oxygenation, and has been found in the cerebrospinal fluid of AD and mildly cognitively impaired (MCI) patients [53]. Additionally, Botchway and colleagues suspected that increased circulatory levels of TMAO can instigate overexpression of cytokines to elevate oxidative stress and endothelial function that results in AD and other neurodegenerative diseases [56]. A recent study on ADLPAPT mice (carry amyloid precursor protein (APP), tau, and presenilin-1, with six mutations) disclosed that daily transfer of fecal microbiota alleviated a myriad of AD-related pathological signs and features, including gliosis, Aβ accumulation, tau-pathology, and MCI. In addition to that, alteration in GM aggravated the gut permeability which resulted in systematic and intestinal inflammation [57]. Another study disclosed that inflammation-related taxa such as Blautia, Desulfovibrio, Escherichia-Shigella, and Akkermansia were distinctly changed in the APP/PSI transgenic mice (a chimeric mouse with human APP and a mutant human presenilin 1 [58]) [59]. Further, a clinical study by Wu et al. [40] revealed that alteration in GM composition is linked with pre-onset amnestic MCI and dementia AD. Saji et al. also found that higher levels of Enterotype I and III bacteria are associated with the occurrence of dementia [60]. Additionly, the abundance of Bifidobacterium, Blautia, Lactobacillus, and Sphingomonas was found higher than Anaerobacterium, Papillibacter, and Odoribacter in AD patients [61]. Furthermore, the abundance of Firmicutes, Proteobacteria, and Tenericutes at phylum, Enterobacteriaceae, Coriobacteriaceae, and Mogibacteriaceae at at family, Phascolarctobacterium and Coprococcus at genus levels was observed higher in the patients with MCI [62]. Nagpal et al. [63] revealed that not only gut bacteria contribute to AD markers but fungal-bacterial co-regulation networks also. The higher proportion of Phaffomyceteceae, Sclerotiniaceae, Cystofilobasidiaceae, Togniniaceae and Trichocomaceae families, Botrytis, Cladosporium, Kazachstania, and Phaeoacremonium genera and lower abundance of Meyerozyma were observed in the patients with MCI. Above-mentioned studies are corroborating GM as a unique factor that has the potential to affect cognitive health and can contribute to AD. Thus, diet or specific bioactive components that have the ability to modulate GM can act as potential therapeutics in MCI and AD.

5. Polyphenols

Dietary polyphenols) are a group of phytochemicals that are naturally present in fruits and vegetables with potential health-promoting effects (e.g., anti-inflammatory, antioxidant, and anti-mutagenic [64,65]). Polyphenols exist in the following forms, (1) free form (such as aglycones), (2) polymers or oligomers (i.e., macromolecules), and (3) derivatives (e.g., glycosylated aglycones, acylated, or esterified). They are classified as either flavonoid (anthocyanins, flavones, flavonols, flavanones, flavanols, and isoflavones) or non-flavonoid (stilbenes, lignans, tannins, phenolic acids, and hydroxycinnamic acids) [66]. Polyphenols exert positive effects and have a broad spectrum of biological activities against many human diseases such as type-2 diabetes mellitus, cancer, cardio-metabolic diseases, and neurodegenerative diseases, as well as having potential to modulate gut dysbiosis (Figure 3) [67]. However, due to the extensive metabolism by phase-I and II enzymatic reactions and poor absorption, their bioactivity on targeted organs is significantly less. Moreover, their transformation into another chemical structure before reaching the site of action may affect their health benefits [68]. Additionally, structural stability, the impact of food matrices, solubility, interaction with GM, etc., also affect the bioavailability of polyphenols. To overcome the bioavailability issues and utilization of the beneficial properties of polyphenols, different techniques can be used, such as nanoencapsulation, microencapsulation, fermentation, or germination [8,69].

5.1. Anti-Oxidative Properties of Polyphenols

In the human diet, the most abundant antioxidants are dietary polyphenols. They can neutralize free radicals through transferring electrons/hydrogen atoms and decrease cell apoptosis via modulation of mitochondrial dysfunction. Further, they can reduce the production of hydroxyl radicals (metal-dependent) along with the chelation mechanism and instigate the nuclear factor erythroid 2-related factor 2 by inducing endogenous antioxidant enzymes [70]. They can scavenge expression of genes, ROS and RNS and activate redox-responsible transcription factors to modulate coding antioxidants, pro-survival neurotrophic factors, and anti-apoptotic Bcl-2 protein family. In addition, they can modulate the mitochondrial apoptosis system in promoting or preventing ways and can regulate mitochondrial biogenesis, autophagic degradation, and dynamics (fission and fusion) [71]. The ability of polyphenols to scavenge radicals primarily depends on the position and number of the OH groups connected with the aromatic rings [72]. In addition to OH groups, polyphenols with two or more groups –NR2, –PO3H2, –COOH, –O–, –SH, C=O, and –S– groups can enhance the chelation of metal ions [73]. For instance, the SH-SY5Y cells were pre-treated with butein, scopoletin, and isoliquiritigenin that protected the cell death induced by H2O2 and decreased ROS and apoptotic cells [74]. Table 2 summarizes the effects of different polyphenols on oxidative stress and inflammation (biomarkers).

5.2. Anti-Inflammatory Properties of Polyphenols

Many studies have stated that an increased level of pro-inflammatory molecules can act as aging indicators. However, the particular mechanisms to relate age-related diseases with inflammation and the reason why old age people are vulnerable to inflammation are still elusive [70]. As an anti-inflammatory agent, polyphenols (such as galangin, luteolin, quercetin, and epigallocatechin-3-gallate) can modulate or suppress the NF-κB activation pathway at different steps that entirely depends on the chemical structure of polyphenols [75]. Further, sirtuins 1 (family of mono-ADP-ribosyltransferase and NAD+-dependent deacylase) can inhibit and deacetylate transcription of p65 subunit of NF-κB at lysine 310 and as a result, attenuate NF-κB induced inflammatory signalling transductions. Polyphenols including quercetin [76], caffeic acid phenylethyl ester [77], hydroxycinnamic acids [78], and ferulic acid [79] activated sirtuins 1 in different study models and identified to protect against the senescence-associated secretory phenotype via NF-κB pathway inhibition. Other than NF-κB, polyphenols can also modulate the NLRP3 inflammasome, e.g., apigenin (flavone class) decreased the LPS-induced IL-6 and IL-1β production via inhibition of caspase-1 activation by interfering with the NLRP3 inflammasome assembly in mouse J774A.1 macrophages [80]. Rutin [81], quercetin [82], and anthocyanins [83] in rats or cultured cells suppressed NLRP3 inflammasome activation that restricted the related inflammatory pathways (Table 2).
Table 2. Summary of the effects of polyphenols on oxidative stress and inflammation (biomarkers).
Table 2. Summary of the effects of polyphenols on oxidative stress and inflammation (biomarkers).
PolyphenolsStudyFindingsReference
AnthocyaninsMouse microglial cellsIL-1β, TNF-α, and NO release, NF-κB nuclear translocation, COX-2 and iNOS expressions.[84]
HumanIL-6, IL-18, and TNF-α[85]
QuercetinMouse BV2 microglial cells and miceOxygen glucose deprivation induced expression of inflammatory factors and TLR4/MyD88/NF-κB signalling. Ameliorated cognitive, cerebral infarct volume and motor function in mice.[86]
Wistar ratsActivity of enzymatic antioxidants and sirtuin 1, NF-κB and IL-1β levels, increased IL-10 and modulated AMPK/SIRT1/NF-κB signaling pathway.[87]
ResveratrolSH-SY5Y neuronal cellsTNF-α, IL-1β, mitochondrial, and cytosolic ROS, improved the intracellular Ca2+ responses and mitochondrial function.[88]
CurcuminSprague-Dawley ratsiNOS, COX-2 expression and inflammatory factor[89]
Epigallocatechin GallateSPF Wistar ratsAcetyl-CoA carboxylase, NF-κB, and free fatty acid synthase and fatty acid binding protein-1, carnitine palmitoyltransferase II and sirtuin 1.[90]
WI-38 cellsAntioxidant enzymes, superoxide dismutase 1 and 2 and IL-32 and TNF-α expression.[91]
LuteolinWistar rats Oxidative stress parameters, levels of NF-κB, malondialdehyde, and hydrogen peroxide and glutathione S-transferase.[92]
KaempferolC57 BL/6J miceTNF-α and IL-6, and the activation of NF-κB and NRF2/HO-1 signaling pathway and level antioxidants[93]
MyricetinWistar ratsMarkers of inflammation such as NF-κB, IL-6, TNF-α, and NRF2, xanthine oxidase activity and phase-II detoxifying enzyme activity and ameliorated lipid peroxidation[94]
Green Tea polyphenolsC57BL/6 miceNLRP3 inflammasome expression, NRF2 pathways, hepatic inflammatory damage and immunological reaction[95]
Grape Seed ExtractHuman colorectal adenocarcinoma cell line Caco-2Pro-inflammatory cytokine gene expression, intracellular ROS and mitochondrial superoxide production, anti-inflammatory cytokines, and mitochondrial membrane potential.[96]
: Higher/increased, : Lower/decreased, NRF2: Nuclear factor-erythroid factor 2-related factor 2, NF-κB: nuclear factor kappa-light-chain-enhancer of activated B cells, IL: interleukin, TNF-α: tumor necrosis factor-α, NO: nitric oxide, iNOS: inducible nitric oxide synthase, ROS: reactive oxygen species, COX-2: cyclooxygenase-2, NLRP3: Nod-like receptor family, pyrin domain containing 3.

5.3. GM and Polyphenols

The relationship between GM and polyphenols is bidirectional as GM bio-transform polyphenols and polyphenols modulate GM. Very low (5–10%) absorption of polyphenols takes place in the small intestine while 90–95% absorption occurs in the large intestine, but bio-transformation of polyphenols in the body is dependent on the GM composition and the structure of polyphenols [8]. Lactase-phlorizin hydrolase hydrolyses the free and simple polyphenols in the small intestine, and the resulting aglycones enter the enterocyte by passive diffusion. Recycled aglycones and the polyphenols gather in the colon, where GM degrade them and facilitate absorptivity [97]. For instance, the sugar moiety of quercetin that intestinal β-glucosidases cannot hydrolyse, but GM (e.g., Enterococcus, Blautia, and Bacteroides) deglycosylation can, yields quercetin aglycon. Strains of Bacteroides, Clostridium perfringens, fragilis, Escherichia coli, Enterococcus gilvus, Lactobacillus acidophilus, Streptococcus S-2, and Weissella confusa can transform quercetin and other polyphenols into bioavailable metabolites [98]. Furthermore, Clostridium saccbarogumia and Eubacterium ramulus can catalyse cyaniding-3-O-glucoside into DHBA, THBAld and other products [99]. On the other hand, Wu and colleagues [70] stated that polyphenols not only act as classic prebiotics to enhance beneficial bacteria (for example Akkermansia, Bifidobacterium, Christensenellaceae, Lactobacillus, and Verrucomicrobia) but also inhibit pathogenic bacteria. Further, Peng et al. [100] documented that long-term consumption of anthocyanins can increase the growth of SCFA-producing bacteria such as Barnesiella, Faecalibacterium, Odoribacter, Prausnitzii, Ruminococcaceae, and Roseburia. Moreover, the consumption of neohesperidin, resveratrol combined with curcumin, green, oolong, and black tea can significantly restrain the growth of pathogenic bacteria (e.g., Clostridiumm, Prevotella, Proteobacteria, and Desulfovibrionaceae) [101,102,103]. Another study by Li et al. [104] documented that the ratio of Firmicutes to Bacteroidetes (positively correlated with many diseases) was increased after the feeding of tea polyphenols in canines. Liu et al. [105] declared that epigallocatechin-3-gallate treatment stimulated the abundance of beneficial bacteria such as Bacteroides, Bifidobacterium, and Christensenellaceae and inhibited pathogenic bacteria including Bilophila, Enterobacteriaceae, and Fusobacterium varium. Above-mentioned studies are suggesting that GM modulation by dietary polyphenols can affect MGBX and can be used as nutraceutical to treat AD and other neurodegenerative diseases.

5.4. Polyphenols and AD

Polyphenols have anti-oxidative and anti-inflammatory properties that abrogate ROS and RNS and sequester the production of Aβ plaques (Aβ-oligomers and Aβ peptides) and tau protein hyperphophorylation to prevent the development of neurofibrillary tangles [106]. Moreover, polyphenols (e.g., hesperidin, neohesperidin, hesperetin, and citrus flavanones) restrict neuronal disintegration by interacting with major signal transduction pathways, cerebral vasculature, and the BBB [106,107]. In addition, quercetin (RVG29-nanoparticles) showed permeability across the BBB and inhibited Aβ aggregation in thioflavin T binding assay [108]. Anthocyanins were found in the cerebellum and cortex of the mice that significantly reduced the loss of neuronal cells and memory impairment [109]. In addition, curcumin and its derivatives can pass the BBB and have neuroprotective effects against mitochondrial dysfunction, damage, and nitrosative stress [46]. Moreover, resveratrol significantly decreased the Aβ-42 peptide toxicity toward SH-SY5Y cells that resulted in the cleavage of Aβ1-42 peptides into smaller fragments [110]. Metabolism of flavan-3-ols by GM result in various arylvaleric acid, and aryl-γ-valerolactone derivatives that can selectively detoxify Aβ oligomers and prevent AD symptoms in mice [111]. Secondary metabolites of valerolactones such as hydroxybenzoic acid, (hydroxyaryl)valeric acid, (hydroxyaryl)propanoic acid, (hydroxyaryl)cinnamic acid, and (hydroxyaryl)acetic acid derivatives have more permeability across the BBB to reduce neuroinflammation and are comparatively more bioavailable than the dietary flavonoids or flavanoids [112]. Moreover, metabolites of epicatechin, 5-(4′-Hydroxyphenyl)-γ-valerolactone-3′-O-glucuronide and 5-(4′-Hydroxyphenyl)-γ-valerolactone-3′-sulfate modulate cellular pathways such as focal adhesion, cell adhesion, signalling pathways, and cytoskeleton organization to preserve brain vascular endothelial cell integrity [113]. Therefore, polyphenols, including those derived from GM metabolism, can be effective therapeutics to treat neurodegenerative diseases such as AD. Several in vivo and in vitro examples are illustrated in Table 3.

6. Research Limitations

Although polyphenols have demonstrated anti-oxidative and anti-inflammatory properties in vitro and in vivo animal studies, there is still inconclusive evidence regarding their effects in human studies. Moreover, the possible interaction between cognitive function, GM composition, and polyphenols has been studied well in animal studies. However, clinical studies have been carried out with a small number of samples that are lacking the comprehensive profiling in GM composition and functionality. Besides, the inconsistency in the results of GM has been observed, which may be due to the difference in the species and nutrients status of animals, treatment time and method, and composition and concentration of polyphenols in the diet. The health benefits of the polyphenols are derived from the GM metabolites that are bioavailable to the host and the interplay between the reshaping of GM, whereas, the mechanism of the GM reshaping is still elusive and may occur either by the parent compounds alone or microbial-derived polyphenolic metabolites. Thus, accurate microbiome studies are needed for future clinical diet interventions [10]. Additionally, more studies are required to check the appropriate concentration of polyphenols for their beneficial and adverse effects [126]. Furthermore, the implementation of artificial intelligence, machine learning algorithms, and use of large datasets are required to understand the complex network of interactions amongst the polyphenols, GM, and host metabolome.

7. Conclusions

There is a close interrelationship between oxidative stress and inflammatory pathways. Each may appear before or after the other and take part in the progression of several chronic diseases [14]. Moreover, gut dysbiosis has been reported to exert regulatory functions on oxidative stress and inflammation and play a role in neurodegenerative diseases, especially AD via MGBX. A balanced diet enriched in antioxidants such as polyphenols can be helpful in maintaining gut homeostasis (eubiosis) by counteracting oxidative stress and inflammation. In this study, we explored the possible role of polyphenols in scavenging free radical species, inhibiting the formation of pro-inflammatory cytokines, increasing anti-inflammatory cytokines, and maintaining gut dysbiosis. Despite the mentioned benefits, they have low bioavailability due to their complex absorption and metabolic process to enter into the bloodstream and succeeding to the target location. Thus, further studies are required to develop methods to improve the stability, permeability, and solubility of dietary polyphenols for their usage in nutraceutical and pharmaceutical applications to develop an efficient approach for preventing and treating neurodegenerative diseases.

Author Contributions

Writing—original draft, writing—review and editing, conceptualization—U.S.; writing—review and editing—A.T., F.E., S.O.A.; supervision, conceptualization—D.-H.O. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Korean Government under the National Institute of International Education (NIIED). Scholarship number—GKS-G-2019-470.

Conflicts of Interest

No author has a direct or indirect conflict of interest in the publication of this review.

Abbreviations

GMGut microbiota
ADAlzheimer’s disease
ROSReactive oxygen species
RNSReactive nitrogen species
SCFAShort-chain fatty acids
MGBXMicrobiota-gut-brain axis
NOXsNADPH oxidases
NONitric oxide
TLR4Toll-like receptor-4
IL-6Interleukin
GITGastrointestinal tract
APPAmyloid precursor protein
TNF-αTumor necrosis factor-α
HMGB1High mobility group box-1
DAMPDamage associated molecular pattern
NF-κBNuclear factor kappa-light-chain-enhancer of activated B cells
NLRP3Nucleotide binding oligomerization domain leucine rich repeat containing protein 3
TMAOTrimethylamine N-oxide
LPSLipopolysaccharides
Amyloid-beta
MCIMild cognitive impaired
MDDMajor depressive disorder
PDParkinson’s disease
BBBBlood-brain barrier
NRF2Nuclear factor-erythroid factor 2-related factor 2
iNOSInducible nitric oxide synthase
COX-2Cyclooxygenase-2
CREBCyclic AMP response element binding protein.

References

  1. Kumar Singh, A.; Cabral, C.; Kumar, R.; Ganguly, R.; Kumar Rana, H.; Gupta, A.; Rosaria Lauro, M.; Carbone, C.; Reis, F.; Pandey, A.K. Beneficial effects of dietary polyphenols on gut microbiota and strategies to improve delivery efficiency. Nutrients 2019, 11, 2216. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Rossi, J.-F.; Lu, Z.Y.; Massart, C.; Levon, K. Dynamic immune/inflammation precision medicine: The good and the bad inflammation in infection and cancer. Front. Immunol. 2021, 12, 97. [Google Scholar] [CrossRef]
  3. Mihai, S.; Codrici, E.; Popescu, I.D.; Enciu, A.-M.; Albulescu, L.; Necula, L.G.; Mambet, C.; Anton, G.; Tanase, C. Inflammation-related mechanisms in chronic kidney disease prediction, progression, and outcome. J. Immunol. Res. 2018, 2018, 2180373. [Google Scholar] [CrossRef]
  4. 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] [Green Version]
  5. Vona, R.; Pallotta, L.; Cappelletti, M.; Severi, C.; Matarrese, P. The Impact of Oxidative Stress in Human Pathology: Focus on Gastrointestinal Disorders. Antioxidants 2021, 10, 201. [Google Scholar] [CrossRef]
  6. Raucci, A.; Di Maggio, S.; Scavello, F.; D’Ambrosio, A.; Bianchi, M.E.; Capogrossi, M.C. The Janus face of HMGB1 in heart disease: A necessary update. Cell. Mol. Life Sci. 2019, 76, 211–229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Battino, M.; Giampieri, F.; Cianciosi, D.; Ansary, J.; Chen, X.; Zhang, D.; Gil, E.; Forbes-Hernández, T. The roles of strawberry and honey phytochemicals on human health: A possible clue on the molecular mechanisms involved in the prevention of oxidative stress and inflammation. Phytomedicine 2021, 86, 153170. [Google Scholar] [CrossRef] [PubMed]
  8. Shabbir, U.; Rubab, M.; Daliri, E.B.-M.; Chelliah, R.; Javed, A.; Oh, D.-H. Curcumin, quercetin, catechins and metabolic diseases: The role of gut microbiota. Nutrients 2021, 13, 206. [Google Scholar] [CrossRef]
  9. Shabbir, U.; Arshad, M.S.; Sameen, A.; Oh, D.-H. Crosstalk between Gut and Brain in Alzheimer’s Disease: The Role of Gut Microbiota Modulation Strategies. Nutrients 2021, 13, 690. [Google Scholar] [CrossRef] [PubMed]
  10. Łuc, M.; Misiak, B.; Pawłowski, M.; Stańczykiewicz, B.; Zabłocka, A.; Szcześniak, D.; Pałęga, A.; Rymaszewska, J. Gut microbiota in dementia. Critical review of novel findings and their potential application. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2021, 104, 110039. [Google Scholar] [CrossRef]
  11. Bhatt, S.; Nagappa, A.N.; Patil, C.R. Role of oxidative stress in depression. Drug Discov. Today 2020, 25, 1270–1276. [Google Scholar] [CrossRef]
  12. Nolfi-Donegan, D.; Braganza, A.; Shiva, S. Mitochondrial Electron Transport: Oxidative Phosphorylation, Mitochondrial Oxidant Production, and Methods of Measurement. Redox Biol. 2020, 37, 101674. [Google Scholar] [CrossRef] [PubMed]
  13. Parra-Ortiz, E.; Browning, K.L.; Damgaard, L.S.; Nordström, R.; Micciulla, S.; Bucciarelli, S.; Malmsten, M. Effects of oxidation on the physicochemical properties of polyunsaturated lipid membranes. J. Colloid Interface Sci. 2019, 538, 404–419. [Google Scholar] [CrossRef]
  14. Chen, T.; Luo, W.; Wu, G.; Wu, L.; Huang, S.; Li, J.; Wang, J.; Hu, X.; Huang, W.; Liang, G. A novel MyD88 inhibitor LM9 prevents atherosclerosis by regulating inflammatory responses and oxidative stress in macrophages. Toxicol. Appl. Pharmacol. 2019, 370, 44–55. [Google Scholar] [CrossRef]
  15. Serbulea, V.; Upchurch, C.M.; Ahern, K.W.; Bories, G.; Voigt, P.; DeWeese, D.E.; Meher, A.K.; Harris, T.E.; Leitinger, N. Macrophages sensing oxidized DAMPs reprogram their metabolism to support redox homeostasis and inflammation through a TLR2-Syk-ceramide dependent mechanism. Mol. Metab. 2018, 7, 23–34. [Google Scholar] [CrossRef] [PubMed]
  16. Włodarczyk, M.; Nowicka, G. Obesity, DNA damage, and development of obesity-related diseases. Int. J. Mol. Sci. 2019, 20, 1146. [Google Scholar] [CrossRef]
  17. Cezar, T.L.; Martinez, R.M.; da Rocha, C.; Melo, C.P.; Vale, D.L.; Borghi, S.M.; Fattori, V.; Vignoli, J.A.; Camilios-Neto, D.; Baracat, M.M. Treatment with maresin 1, a docosahexaenoic acid-derived pro-resolution lipid, protects skin from inflammation and oxidative stress caused by UVB irradiation. Sci. Rep. 2019, 9, 3062. [Google Scholar] [CrossRef] [PubMed]
  18. Shah, B.S.; Burt, K.G.; Jacobsen, T.; Fernandes, T.D.; Alipui, D.O.; Weber, K.T.; Levine, M.; Chavan, S.S.; Yang, H.; Tracey, K.J. High mobility group box-1 induces pro-inflammatory signaling in human nucleus pulposus cells via toll-like receptor 4-dependent pathway. J. Orthop. Res. 2019, 37, 220–231. [Google Scholar] [CrossRef] [Green Version]
  19. Hatayama, K.; Stonestreet, B.S. High mobility group box-1 protein as a therapeutic target in perinatal hypoxic-ischemic brain injury. Neural Regen. Res. 2021, 16, 2006–2007. [Google Scholar]
  20. Nishibori, M.; Mori, S.; Takahashi, H.K. Anti-HMGB1 monoclonal antibody therapy for a wide range of CNS and PNS diseases. J. Pharmacol. Sci. 2019, 140, 94–101. [Google Scholar] [CrossRef]
  21. Duan, J.; Zhang, Q.; Hu, X.; Lu, D.; Yu, W.; Bai, H. N4-acetylcytidine is required for sustained NLRP3 inflammasome activation via HMGB1 pathway in microglia. Cell. Signal. 2019, 58, 44–52. [Google Scholar] [CrossRef]
  22. Martinon, F. Signaling by ROS drives inflammasome activation. Eur. J. Immunol. 2010, 40, 616–619. [Google Scholar] [CrossRef]
  23. Oleskin, A.V.; Shenderov, B.A. Neuromodulatory effects and targets of the SCFAs and gasotransmitters produced by the human symbiotic microbiota. Microb. Ecol. Health Dis. 2016, 27, 30971. [Google Scholar] [CrossRef]
  24. Luca, M.; Di Mauro, M.; Perry, G. Neuropsychiatric Disturbances and Diabetes Mellitus: The Role of Oxidative Stress; Hindawi: London, UK, 2019. [Google Scholar]
  25. Beaumont, M.; Andriamihaja, M.; Lan, A.; Khodorova, N.; Audebert, M.; Blouin, J.-M.; Grauso, M.; Lancha, L.; Benetti, P.-H.; Benamouzig, R. Detrimental effects for colonocytes of an increased exposure to luminal hydrogen sulfide: The adaptive response. Free Radic. Biol. Med. 2016, 93, 155–164. [Google Scholar] [CrossRef]
  26. He, Z.; Kwek, E.; Hao, W.; Zhu, H.; Liu, J.; Ma, K.Y.; Chen, Z.-Y. Hawthorn fruit extract reduced trimethylamine-N-oxide (TMAO)-exacerbated atherogenesis in mice via anti-inflammation and anti-oxidation. Nutr. Metab. 2021, 18, 1–15. [Google Scholar] [CrossRef]
  27. Loffredo, L.; Ettorre, E.; Zicari, A.M.; Inghilleri, M.; Nocella, C.; Perri, L.; Spalice, A.; Fossati, C.; De Lucia, M.C.; Pigozzi, F. Oxidative stress and gut-derived lipopolysaccharides in neurodegenerative disease: Role of NOX2. Oxidative Med. Cell. Longev. 2020, 2020, 8630275. [Google Scholar] [CrossRef] [Green Version]
  28. Kesika, P.; Suganthy, N.; Sivamaruthi, B.S.; Chaiyasut, C. Role of gut-brain axis, gut microbial composition, and probiotic intervention in Alzheimer’s disease. Life Sci. 2020, 264, 118627. [Google Scholar] [CrossRef]
  29. Cattaneo, A.; Cattane, N.; Galluzzi, S.; Provasi, S.; Lopizzo, N.; Festari, C.; Ferrari, C.; Guerra, U.P.; Paghera, B.; Muscio, C. Association of brain amyloidosis with pro-inflammatory gut bacterial taxa and peripheral inflammation markers in cognitively impaired elderly. Neurobiol. Aging 2017, 49, 60–68. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Chung, Y.-C.E.; Chen, H.-C.; Chou, H.-C.L.; Chen, I.-M.; Lee, M.-S.; Chuang, L.-C.; Liu, Y.-W.; Lu, M.-L.; Chen, C.-H.; Wu, C.-S. Exploration of microbiota targets for major depressive disorder and mood related traits. J. Psychiatr. Res. 2019, 111, 74–82. [Google Scholar] [CrossRef]
  31. Liśkiewicz, P.; Kaczmarczyk, M.; Misiak, B.; Wroński, M.; Bąba-Kubiś, A.; Skonieczna-Żydecka, K.; Marlicz, W.; Bieńkowski, P.; Misera, A.; Pełka-Wysiecka, J. Analysis of gut microbiota and intestinal integrity markers of inpatients with major depressive disorder. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2021, 106, 110076. [Google Scholar] [CrossRef]
  32. Jiang, H.-y.; Zhang, X.; Yu, Z.-h.; Zhang, Z.; Deng, M.; Zhao, J.-H.; Ruan, B. Altered gut microbiota profile in patients with generalized anxiety disorder. J. Psychiatr. Res. 2018, 104, 130–136. [Google Scholar] [CrossRef]
  33. Chen, Y.-h.; Bai, J.; Wu, D.; Yu, S.-F.; Qiang, X.-L.; Bai, H.; Wang, H.-N.; Peng, Z.-W. Association between fecal microbiota and generalized anxiety disorder: Severity and early treatment response. J. Affect. Disord. 2019, 259, 56–66. [Google Scholar] [CrossRef]
  34. Turna, J.; Grosman Kaplan, K.; Anglin, R.; Patterson, B.; Soreni, N.; Bercik, P.; Surette, M.; Van Ameringen, M. The gut microbiome and inflammation in obsessive-compulsive disorder patients compared to age-and sex-matched controls: A pilot study. Acta Psychiatr. Scand. 2020, 142, 337–347. [Google Scholar] [CrossRef]
  35. Ren, T.; Gao, Y.; Qiu, Y.; Jiang, S.; Zhang, Q.; Zhang, J.; Wang, L.; Zhang, Y.; Wang, L.; Nie, K. Gut microbiota altered in mild cognitive impairment compared with normal cognition in sporadic Parkinson’s disease. Front. Neurol. 2020, 11, 137. [Google Scholar] [CrossRef] [PubMed]
  36. Aho, V.T.; Houser, M.C.; Pereira, P.A.; Chang, J.; Rudi, K.; Paulin, L.; Hertzberg, V.; Auvinen, P.; Tansey, M.G.; Scheperjans, F. Relationships of gut microbiota, short-chain fatty acids, inflammation, and the gut barrier in Parkinson’s disease. Mol. Neurodegener. 2021, 16, 1–14. [Google Scholar] [CrossRef] [PubMed]
  37. Li, X.; Yuan, X.; Pang, L.; Miao, Y.; Wang, S.; Zhang, X.; Hu, S.; Wang, Y.; Andreassen, O.A.; Song, X. Gut Microbiota Markers for Antipsychotics Induced Metabolic Disturbance in Drug Naïve Patients with First Episode Schizophrenia—A 24 Weeks Follow-up Study. medRxiv 2021, 2020-12. [Google Scholar] [CrossRef]
  38. McIntyre, R.S.; Subramaniapillai, M.; Shekotikhina, M.; Carmona, N.E.; Lee, Y.; Mansur, R.B.; Brietzke, E.; Fus, D.; Coles, A.S.; Iacobucci, M. Characterizing the gut microbiota in adults with bipolar disorder: A pilot study. Nutr. Neurosci. 2021, 24, 173–180. [Google Scholar] [CrossRef]
  39. Lai, W.-T.; Zhao, J.; Xu, S.-X.; Deng, W.-F.; Xu, D.; Wang, M.-B.; He, F.-S.; Liu, Y.-H.; Guo, Y.-Y.; Ye, S.-W. Shotgun metagenomics reveals both taxonomic and tryptophan pathway differences of gut microbiota in bipolar disorder with current major depressive episode patients. J. Affect. Disord. 2021, 278, 311–319. [Google Scholar] [CrossRef]
  40. Wu, L.; Han, Y.; Zheng, Z.; Peng, G.; Liu, P.; Yue, S.; Zhu, S.; Chen, J.; Lv, H.; Shao, L. Altered gut microbial metabolites in amnestic mild cognitive impairment and Alzheimer’s disease: Signals in host–microbe interplay. Nutrients 2021, 13, 228. [Google Scholar] [CrossRef]
  41. Şafak, B.; Altunan, B.; Topçu, B.; Topkaya, A.E. The gut microbiome in epilepsy. Microb. Pathog. 2020, 139, 103853. [Google Scholar] [CrossRef]
  42. Lee, H.; Lee, S.; Lee, D.-H.; Kim, D.W. A comparison of the gut microbiota among adult patients with drug-responsive and drug-resistant epilepsy: An exploratory study. Epilepsy Res. 2021, 172, 106601. [Google Scholar] [CrossRef]
  43. Stan, T.L.; Soylu-Kucharz, R.; Burleigh, S.; Prykhodko, O.; Cao, L.; Franke, N.; Sjögren, M.; Haikal, C.; Hållenius, F.; Björkqvist, M. Increased intestinal permeability and gut dysbiosis in the R6/2 mouse model of Huntington’s disease. Sci. Rep. 2020, 10, 18270. [Google Scholar] [CrossRef]
  44. WHO. W.H.O. Dementia Fact Sheets. 2020. Available online: https://www.who.int/news-room/fact-sheets/detail/dementia (accessed on 10 July 2021).
  45. Alzheimer’s Disease International. Dementia Statistics. 2020. Available online: https://www.alzint.org/about/dementia-facts-figures/dementia-statistics/ (accessed on 10 July 2021).
  46. Shabbir, U.; Rubab, M.; Tyagi, A.; Oh, D.-H. Curcumin and its derivatives as theranostic agents in Alzheimer’s disease: The implication of nanotechnology. Int. J. Mol. Sci. 2021, 22, 196. [Google Scholar] [CrossRef]
  47. Alzheimer’s Association. Stages of Alzheimer’s. 2021. Available online: https://www.alz.org/alzheimers-dementia/stages (accessed on 7 July 2021).
  48. He, Y.; Li, B.; Sun, D.; Chen, S. Gut microbiota: Implications in Alzheimer’s disease. J. Clin. Med. 2020, 9, 2042. [Google Scholar] [CrossRef]
  49. Zetterberg, H.; Bendlin, B.B. Biomarkers for Alzheimer’s disease—Preparing for a new era of disease-modifying therapies. Mol. Psychiatry 2021, 26, 296–308. [Google Scholar] [CrossRef]
  50. Liu, P.-P.; Xie, Y.; Meng, X.-Y.; Kang, J.-S. History and progress of hypotheses and clinical trials for Alzheimer’s disease. Signal Transduct. Target. Ther. 2019, 4, 1–22. [Google Scholar] [CrossRef]
  51. Sun, M.; Ma, K.; Wen, J.; Wang, G.; Zhang, C.; Li, Q.; Bao, X.; Wang, H. A review of the brain-gut-microbiome axis and the potential role of microbiota in Alzheimer’s disease. J. Alzheimer’s Dis. 2020, 73, 849–865. [Google Scholar] [CrossRef]
  52. Giau, V.V.; Wu, S.Y.; Jamerlan, A.; An, S.S.A.; Kim, S.; Hulme, J. Gut microbiota and their neuroinflammatory implications in Alzheimer’s disease. Nutrients 2018, 10, 1765. [Google Scholar] [CrossRef] [Green Version]
  53. Vogt, N.M.; Romano, K.A.; Darst, B.F.; Engelman, C.D.; Johnson, S.C.; Carlsson, C.M.; Asthana, S.; Blennow, K.; Zetterberg, H.; Bendlin, B.B. The gut microbiota-derived metabolite trimethylamine N-oxide is elevated in Alzheimer’s disease. Alzheimer’s Res. Ther. 2018, 10, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Ho, L.; Ono, K.; Tsuji, M.; Mazzola, P.; Singh, R.; Pasinetti, G.M. Protective roles of intestinal microbiota derived short chain fatty acids in Alzheimer’s disease-type beta-amyloid neuropathological mechanisms. Expert Rev. Neurother. 2018, 18, 83–90. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Borsom, E.M.; Lee, K.; Cope, E.K. Do the bugs in your gut eat your memories? Relationship between gut microbiota and Alzheimer’s disease. Brain Sci. 2020, 10, 814. [Google Scholar] [CrossRef]
  56. Botchway, B.O.; Okoye, F.C.; Chen, Y.; Arthur, W.E.; Fang, M. Alzheimer Disease: Recent Updates on Apolipoprotein E and Gut Microbiome Mediation of Oxidative Stress, and Prospective Interventional Agents. Aging Dis. 2021, 13, 55–85. [Google Scholar]
  57. Kim, M.-S.; Kim, Y.; Choi, H.; Kim, W.; Park, S.; Lee, D.; Kim, D.K.; Kim, H.J.; Choi, H.; Hyun, D.-W. Transfer of a healthy microbiota reduces amyloid and tau pathology in an Alzheimer’s disease animal model. Gut 2020, 69, 283–294. [Google Scholar] [CrossRef]
  58. Zhang, A.H.; Ma, Z.m.; Kong, L.; Gao, H.L.; Sun, H.; Wang, X.Q.; Yu, J.B.; Han, Y.; Yan, G.L.; Wang, X.J. High-throughput lipidomics analysis to discover lipid biomarkers and profiles as potential targets for evaluating efficacy of Kai-Xin-San against APP/PS1 transgenic mice based on UPLC–Q/TOF–MS. Biomed. Chromatogr. 2020, 34, e4724. [Google Scholar] [CrossRef] [PubMed]
  59. Chen, Y.; Fang, L.; Chen, S.; Zhou, H.; Fan, Y.; Lin, L.; Li, J.; Xu, J.; Chen, Y.; Ma, Y. Gut microbiome alterations precede cerebral amyloidosis and microglial pathology in a mouse model of Alzheimer’s disease. Biomed. Res. Int. 2020, 2020, 8456596. [Google Scholar] [CrossRef] [PubMed]
  60. Saji, N.; Murotani, K.; Hisada, T.; Kunihiro, T.; Tsuduki, T.; Sugimoto, T.; Kimura, A.; Niida, S.; Toba, K.; Sakurai, T. Relationship between dementia and gut microbiome-associated metabolites: A cross-sectional study in Japan. Sci. Rep. 2020, 10, 8088. [Google Scholar] [CrossRef]
  61. Zhou, Y.; Wang, Y.; Quan, M.; Zhao, H.; Jia, J. Gut Microbiota Changes and Their Correlation with Cognitive and Neuropsychiatric Symptoms in Alzheimer’s Disease. J. Alzheimer’s Dis. 2021, 81, 583–595. [Google Scholar] [CrossRef] [PubMed]
  62. Nagpal, R.; Neth, B.J.; Wang, S.; Craft, S.; Yadav, H. Modified Mediterranean-ketogenic diet modulates gut microbiome and short-chain fatty acids in association with Alzheimer’s disease markers in subjects with mild cognitive impairment. EBioMedicine 2019, 47, 529–542. [Google Scholar] [CrossRef] [Green Version]
  63. Nagpal, R.; Neth, B.J.; Wang, S.; Mishra, S.P.; Craft, S.; Yadav, H. Gut mycobiome and its interaction with diet, gut bacteria and Alzheimer’s disease markers in subjects with mild cognitive impairment: A pilot study. EBioMedicine 2020, 59, 102950. [Google Scholar] [CrossRef]
  64. Aravind, S.M.; Wichienchot, S.; Tsao, R.; Ramakrishnan, S.; Chakkaravarthi, S. Role of dietary polyphenols on gut microbiota, their metabolites and health benefits. Food Res. Int. 2021, 142, 110189. [Google Scholar] [CrossRef]
  65. Shabbir, U.; Khalid, S.; Abbas, M.; Suleria, H.A.R. Natural carotenoids: Weapon against life-style-related disorders. In Phytochemicals from Medicinal Plants; Apple Academic Press: Palm Bay, FL, USA, 2019; pp. 159–178. [Google Scholar]
  66. Di Lorenzo, C.; Colombo, F.; Biella, S.; Stockley, C.; Restani, P. Polyphenols and human health: The role of bioavailability. Nutrients 2021, 13, 273. [Google Scholar] [CrossRef]
  67. Bucciantini, M.; Leri, M.; Nardiello, P.; Casamenti, F.; Stefani, M. Olive Polyphenols: Antioxidant and Anti-Inflammatory Properties. Antioxidants 2021, 10, 1044. [Google Scholar] [CrossRef]
  68. Zhao, D.; Simon, J.E.; Wu, Q. A critical review on grape polyphenols for neuroprotection: Strategies to enhance bioefficacy. Crit. Rev. Food Sci. Nutr. 2020, 60, 597–625. [Google Scholar] [CrossRef]
  69. Tyagi, A.; Shabbir, U.; Chelliah, R.; Daliri, E.B.-M.; Chen, X.; Oh, D.-H. Limosilactobacillus reuteri Fermented Brown Rice: A Product with Enhanced Bioactive Compounds and Antioxidant Potential. Antioxidants 2021, 10, 1077. [Google Scholar] [CrossRef]
  70. Wu, M.; Luo, Q.; Nie, R.; Yang, X.; Tang, Z.; Chen, H. Potential implications of polyphenols on aging considering oxidative stress, inflammation, autophagy, and gut microbiota. Crit. Rev. Food Sci. Nutr. 2020, 61, 2175–2193. [Google Scholar] [CrossRef] [PubMed]
  71. Naoi, M.; Wu, Y.; Shamoto-Nagai, M.; Maruyama, W. Mitochondria in neuroprotection by phytochemicals: Bioactive polyphenols modulate mitochondrial apoptosis system, function and structure. Int. J. Mol. Sci. 2019, 20, 2451. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Wojtunik-Kulesza, K.; Oniszczuk, A.; Oniszczuk, T.; Combrzyński, M.; Nowakowska, D.; Matwijczuk, A. Influence of in vitro digestion on composition, bioaccessibility and antioxidant activity of food polyphenols—A non-systematic review. Nutrients 2020, 12, 1401. [Google Scholar] [CrossRef] [PubMed]
  73. Russo, G.L.; Spagnuolo, C.; Russo, M.; Tedesco, I.; Moccia, S.; Cervellera, C. Mechanisms of aging and potential role of selected polyphenols in extending healthspan. Biochem. Pharmacol. 2020, 173, 113719. [Google Scholar] [CrossRef]
  74. Gay, N.H.; Suwanjang, W.; Ruankham, W.; Songtawee, N.; Wongchitrat, P.; Prachayasittikul, V.; Prachayasittikul, S.; Phopin, K. Butein, isoliquiritigenin, and scopoletin attenuate neurodegeneration via antioxidant enzymes and SIRT1/ADAM10 signaling pathway. RSC Adv. 2020, 10, 16593–16606. [Google Scholar] [CrossRef]
  75. 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]
  76. Sarubbo, F.; Ramis, M.; Kienzer, C.; Aparicio, S.; Esteban, S.; Miralles, A.; Moranta, D. Chronic silymarin, quercetin and naringenin treatments increase monoamines synthesis and hippocampal Sirt1 levels improving cognition in aged rats. J. Neuroimmune Pharmacol. 2018, 13, 24–38. [Google Scholar] [CrossRef]
  77. Li, D.; Wang, X.; Huang, Q.; Li, S.; Zhou, Y.; Li, Z. Cardioprotection of CAPE-oNO2 against myocardial ischemia/reperfusion induced ROS generation via regulating the SIRT1/eNOS/NF-κB pathway in vivo and in vitro. Redox Biol. 2018, 15, 62–73. [Google Scholar] [CrossRef]
  78. Gerardi, G.; Cavia-Saiz, M.; Rivero-Pérez, M.D.; González-SanJosé, M.L.; Muñiz, P. Modulation of Akt-p38-MAPK/Nrf2/SIRT1 and NF-κB pathways by wine pomace product in hyperglycemic endothelial cell line. J. Funct. Foods 2019, 58, 255–265. [Google Scholar] [CrossRef]
  79. Hou, T.; Zhang, L.; Yang, X. Ferulic acid, a natural polyphenol, protects against osteoporosis by activating SIRT1 and NF-κB in neonatal rats with glucocorticoid-induced osteoporosis. Biomed. Pharmacother. 2019, 120, 109205. [Google Scholar] [CrossRef]
  80. Zhang, X.; Wang, G.; Gurley, E.C.; Zhou, H. Flavonoid apigenin inhibits lipopolysaccharide-induced inflammatory response through multiple mechanisms in macrophages. PLoS ONE 2014, 9, e107072. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Wang, W.; Wu, Q.-H.; Sui, Y.; Wang, Y.; Qiu, X. Rutin protects endothelial dysfunction by disturbing Nox4 and ROS-sensitive NLRP3 inflammasome. Biomed. Pharmacother. 2017, 86, 32–40. [Google Scholar] [CrossRef]
  82. Li, H.; Xiao, L.; He, H.; Zeng, H.; Liu, J.; Jiang, C.; Mei, G.; Yu, J.; Chen, H.; Yao, P. Quercetin Attenuates Atherosclerotic Inflammation by Inhibiting Gal-3-NLRP3 Signaling Pathway. Mol. Nutr. Food Res. 2021, 65, 2000746. [Google Scholar] [CrossRef] [PubMed]
  83. Molagoda, I.M.N.; Lee, K.T.; Choi, Y.H.; Jayasingha, J.A.C.C.; Kim, G.-Y. Anthocyanins from Hibiscus syriacus L. Inhibit NLRP3 Inflammasome in BV2 Microglia Cells by Alleviating NF-κB-and ER Stress-Induced Ca2+ Accumulation and Mitochondrial ROS Production. Oxidative Med. Cell. Longev. 2021, 2021, 1246491. [Google Scholar] [CrossRef] [PubMed]
  84. Zhao, L.; Chen, S.; Liu, T.; Wang, X.; Huang, H.; Liu, W. Callistephin enhances the protective effects of isoflurane on microglial injury through downregulation of inflammation and apoptosis. Mol. Med. Rep. 2019, 20, 802–812. [Google Scholar] [CrossRef]
  85. Nikbakht, E.; Singh, I.; Vider, J.; Williams, L.T.; Vugic, L.; Gaiz, A.; Kundur, A.R.; Colson, N. Potential of anthocyanin as an anti-inflammatory agent: A human clinical trial on type 2 diabetic, diabetic at-risk and healthy adults. Inflamm. Res. 2021, 70, 275–284. [Google Scholar] [CrossRef] [PubMed]
  86. Le, K.; Song, Z.; Deng, J.; Peng, X.; Zhang, J.; Wang, L.; Zhou, L.; Bi, H.; Liao, Z.; Feng, Z. Quercetin alleviates neonatal hypoxic-ischemic brain injury by inhibiting microglia-derived oxidative stress and TLR4-mediated inflammation. Inflamm. Res. 2020, 69, 1201–1213. [Google Scholar] [CrossRef]
  87. Zhang, F.; Feng, J.; Zhang, J.; Kang, X.; Qian, D. Quercetin modulates AMPK/SIRT1/NF-κB signaling to inhibit inflammatory/oxidative stress responses in diabetic high fat diet-induced atherosclerosis in the rat carotid artery. Exp. Ther. Med. 2020, 20, 280. [Google Scholar] [CrossRef]
  88. Akyuva, Y.; Nazıroğlu, M. Resveratrol attenuates hypoxia-induced neuronal cell death, inflammation and mitochondrial oxidative stress by modulation of TRPM2 channel. Sci. Rep. 2020, 10, 6449. [Google Scholar] [CrossRef] [Green Version]
  89. Zhao, Y.H.; Shen, C.F.; Wang, G.J.; Kang, Y.; Song, Y.H.; Liu, J.W. Curcumin alleviates acute kidney injury in a dry-heat environment by reducing oxidative stress and inflammation in a rat model. J. Biochem. Mol. Toxicol. 2021, 35, e22630. [Google Scholar] [CrossRef]
  90. Yuan, H.; Li, Y.; Ling, F.; Guan, Y.; Zhang, D.; Zhu, Q.; Liu, J.; Wu, Y.; Niu, Y. The phytochemical epigallocatechin gallate prolongs the lifespan by improving lipid metabolism, reducing inflammation and oxidative stress in high-fat diet-fed obese rats. Aging Cell 2020, 19, e13199. [Google Scholar] [CrossRef]
  91. Zhang, Q.; Wu, Y.; Guan, Y.; Ling, F.; Li, Y.; Niu, Y. Epigallocatechin gallate prevents senescence by alleviating oxidative stress and inflammation in WI-38 human embryonic fibroblasts. RSC Adv. 2019, 9, 26787–26798. [Google Scholar] [CrossRef] [Green Version]
  92. Oyagbemi, A.A.; Akinrinde, A.S.; Adebiyi, O.E.; Jarikre, T.A.; Omobowale, T.O.; Ola-Davies, O.E.; Saba, A.B.; Emikpe, B.O.; Adedapo, A.A. Luteolin supplementation ameliorates cobalt-induced oxidative stress and inflammation by suppressing NF-kB/Kim-1 signaling in the heart and kidney of rats. Environ. Toxicol. Pharmacol. 2020, 80, 103488. [Google Scholar] [CrossRef]
  93. Yao, H.; Sun, J.; Wei, J.; Zhang, X.; Chen, B.; Lin, Y. Kaempferol protects blood vessels from damage induced by oxidative stress and inflammation in association with the Nrf2/HO-1 signaling pathway. Front. Pharmacol. 2020, 11, 1118. [Google Scholar] [CrossRef]
  94. Rehman, M.U.; Rather, I.A. Myricetin abrogates cisplatin-induced oxidative stress, inflammatory response, and goblet cell disintegration in colon of wistar rats. Plants 2020, 9, 28. [Google Scholar] [CrossRef] [Green Version]
  95. Wang, D.; Zhang, M.; Wang, T.; Liu, T.; Guo, Y.; Granato, D. Green tea polyphenols mitigate the plant lectins-induced liver inflammation and immunological reaction in C57BL/6 mice via NLRP3 and Nrf2 signaling pathways. Food Chem. Toxicol. 2020, 144, 111576. [Google Scholar] [CrossRef]
  96. Nallathambi, R.; Poulev, A.; Zuk, J.B.; Raskin, I. Proanthocyanidin-rich grape seed extract reduces inflammation and oxidative stress and restores tight junction barrier function in Caco-2 colon cells. Nutrients 2020, 12, 1623. [Google Scholar] [CrossRef]
  97. Dey, P. Gut microbiota in phytopharmacology: A comprehensive overview of concepts, reciprocal interactions, biotransformations and mode of actions. Pharmacol. Res. 2019, 147, 104367. [Google Scholar] [CrossRef]
  98. Santangelo, R.; Silvestrini, A.; Mancuso, C. Ginsenosides, catechins, quercetin and gut microbiota: Current evidence of challenging interactions. Food Chem. Toxicol. 2019, 123, 42–49. [Google Scholar] [CrossRef]
  99. Hanske, L.; Engst, W.; Loh, G.; Sczesny, S.; Blaut, M.; Braune, A. Contribution of gut bacteria to the metabolism of cyanidin 3-glucoside in human microbiota-associated rats. Br. J. Nutr. 2013, 109, 1433–1441. [Google Scholar] [CrossRef]
  100. Peng, Y.; Yan, Y.; Wan, P.; Dong, W.; Huang, K.; Ran, L.; Mi, J.; Lu, L.; Zeng, X.; Cao, Y. Effects of long-term intake of anthocyanins from Lycium ruthenicum Murray on the organism health and gut microbiota in vivo. Food Res. Int. 2020, 130, 108952. [Google Scholar] [CrossRef]
  101. Sun, H.; Chen, Y.; Cheng, M.; Zhang, X.; Zheng, X.; Zhang, Z. The modulatory effect of polyphenols from green tea, oolong tea and black tea on human intestinal microbiota in vitro. J. Food Sci. Technol. 2018, 55, 399–407. [Google Scholar] [CrossRef]
  102. Sreng, N.; Champion, S.; Martin, J.-C.; Khelaifia, S.; Christensen, J.E.; Padmanabhan, R.; Azalbert, V.; Blasco-Baque, V.; Loubieres, P.; Pechere, L. Resveratrol-mediated glycemic regulation is blunted by curcumin and is associated to modulation of gut microbiota. J. Nutr. Biochem. 2019, 72, 108218. [Google Scholar] [CrossRef]
  103. Gong, Y.; Dong, R.; Gao, X.; Li, J.; Jiang, L.; Zheng, J.; Cui, S.; Ying, M.; Yang, B.; Cao, J. Neohesperidin prevents colorectal tumorigenesis by altering the gut microbiota. Pharmacol. Res. 2019, 148, 104460. [Google Scholar] [CrossRef]
  104. Li, Y.; Rahman, S.U.; Huang, Y.; Zhang, Y.; Ming, P.; Zhu, L.; Chu, X.; Li, J.; Feng, S.; Wang, X. Green tea polyphenols decrease weight gain, ameliorate alteration of gut microbiota, and mitigate intestinal inflammation in canines with high-fat-diet-induced obesity. J. Nutr. Biochem. 2020, 78, 108324. [Google Scholar] [CrossRef]
  105. Liu, Z.; de Bruijn, W.J.; Bruins, M.E.; Vincken, J.-P. Reciprocal interactions between epigallocatechin-3-gallate (EGCG) and human gut microbiota in vitro. J. Agric. Food Chem. 2020, 68, 9804–9815. [Google Scholar] [CrossRef]
  106. Reddy, V.P.; Aryal, P.; Robinson, S.; Rafiu, R.; Obrenovich, M.; Perry, G. Polyphenols in Alzheimer’s Disease and in the Gut–Brain Axis. Microorganisms 2020, 8, 199. [Google Scholar] [CrossRef] [Green Version]
  107. Jaeger, B.N.; Parylak, S.L.; Gage, F.H. Mechanisms of dietary flavonoid action in neuronal function and neuroinflammation. Mol. Asp. Med. 2018, 61, 50–62. [Google Scholar] [CrossRef]
  108. Pinheiro, R.; Granja, A.; Loureiro, J.; Pereira, M.; Pinheiro, M.; Neves, A.; Reis, S. RVG29-functionalized lipid nanoparticles for quercetin brain delivery and Alzheimer’s disease. Pharm. Res. 2020, 37, 1–12. [Google Scholar] [CrossRef]
  109. Khan, M.S.; Khan, A.; Ahmad, S.; Ahmad, R.; Rehman, I.U.; Ikram, M.; Kim, M.O. Inhibition of JNK alleviates chronic hypoperfusion-related ischemia induces oxidative stress and brain degeneration via Nrf2/HO-1 and NF-κB signaling. Oxidative Med. Cell. Longev. 2020, 2020, 5291852. [Google Scholar] [CrossRef]
  110. Al-Edresi, S.; Alsalahat, I.; Freeman, S.; Aojula, H.; Penny, J. Resveratrol-mediated cleavage of amyloid β1–42 peptide: Potential relevance to Alzheimer’s disease. Neurobiol. Aging 2020, 94, 24–33. [Google Scholar] [CrossRef]
  111. Ruotolo, R.; Minato, I.; La Vitola, P.; Artioli, L.; Curti, C.; Franceschi, V.; Brindani, N.; Amidani, D.; Colombo, L.; Salmona, M. Flavonoid-Derived Human Phenyl-γ-Valerolactone Metabolites Selectively Detoxify Amyloid-β Oligomers and Prevent Memory Impairment in a Mouse Model of Alzheimer’s Disease. Mol. Nutr. Food Res. 2020, 64, 1900890. [Google Scholar] [CrossRef]
  112. Carregosa, D.; Carecho, R.; Figueira, I.S.; Santos, C.U. Low-molecular weight metabolites from polyphenols as effectors for attenuating neuroinflammation. J. Agric. Food Chem. 2019, 68, 1790–1807. [Google Scholar] [CrossRef] [Green Version]
  113. Corral-Jara, K.F.; Nuthikattu, S.; Rutledge, J.; Villablanca, A.; Morand, C.; Schroeter, H.; Milenkovic, D. Integrated Multi-Omic Analyses of the Genomic Modifications by Gut Microbiome-Derived Metabolites of Epicatechin, 5-(4′-Hydroxyphenyl)-γ-Valerolactone, in TNFα-Stimulated Primary Human Brain Microvascular Endothelial Cells. Front. Neurosci. 2021, 15, 622640. [Google Scholar] [CrossRef]
  114. Sun, Z.-Z.; Li, X.-Y.; Wang, S.; Shen, L.; Ji, H.-F. Bidirectional interactions between curcumin and gut microbiota in transgenic mice with Alzheimer’s disease. Appl. Microbiol. Biotechnol. 2020, 104, 3507–3515. [Google Scholar] [CrossRef]
  115. Xu, M.; Huang, H.; Mo, X.; Zhu, Y.; Chen, X.; Li, X.; Peng, X.; Xu, Z.; Chen, L.; Rong, S. Quercetin-3-O-Glucuronide Alleviates Cognitive Deficit and Toxicity in Aβ1-42-Induced AD-Like Mice and SH-SY5Y Cells. Mol. Nutr. Food Res. 2021, 65, 2000660. [Google Scholar] [CrossRef]
  116. Selvakumar, K.; Bavithra, S.; Krishnamoorthy, G.; Arunakaran, J. Impact of quercetin on tight junctional proteins and BDNF signaling molecules in hippocampus of PCBs-exposed rats. Interdiscip. Toxicol. 2018, 11, 294. [Google Scholar] [CrossRef] [Green Version]
  117. Yammine, A.; Zarrouk, A.; Nury, T.; Vejux, A.; Latruffe, N.; Vervandier-Fasseur, D.; Samadi, M.; Mackrill, J.J.; Greige-Gerges, H.; Auezova, L. Prevention by dietary polyphenols (resveratrol, quercetin, apigenin) against 7-ketocholesterol-induced oxiapoptophagy in neuronal N2a cells: Potential interest for the treatment of neurodegenerative and age-related diseases. Cells 2020, 9, 2346. [Google Scholar] [CrossRef]
  118. Fu, X.; Zhang, J.; Guo, L.; Xu, Y.; Sun, L.; Wang, S.; Feng, Y.; Gou, L.; Zhang, L.; Liu, Y. Protective role of luteolin against cognitive dysfunction induced by chronic cerebral hypoperfusion in rats. Pharmacol. Biochem. Behav. 2014, 126, 122–130. [Google Scholar] [CrossRef]
  119. Facchinetti, R.; Valenza, M.; Bronzuoli, M.R.; Menegoni, G.; Ratano, P.; Steardo, L.; Campolongo, P.; Scuderi, C. Looking for a Treatment for the Early Stage of Alzheimer’s Disease: Preclinical Evidence with Co-Ultramicronized Palmitoylethanolamide and Luteolin. Int. J. Mol. Sci. 2020, 21, 3802. [Google Scholar] [CrossRef] [PubMed]
  120. Li, J.; Wu, T.; Li, N.; Wang, X.; Chen, G.; Lyu, X. Bilberry anthocyanin extract promotes intestinal barrier function and inhibits digestive enzyme activity by regulating the gut microbiota in aging rats. Food Funct. 2019, 10, 333–343. [Google Scholar] [CrossRef]
  121. Li, J.; Zhao, R.; Jiang, Y.; Xu, Y.; Zhao, H.; Lyu, X.; Wu, T. Bilberry anthocyanins improve neuroinflammation and cognitive dysfunction in APP/PSEN1 mice via the CD33/TREM2/TYROBP signaling pathway in microglia. Food Funct. 2020, 11, 1572–1584. [Google Scholar] [CrossRef] [PubMed]
  122. Song, C.; Zhang, Y.; Cheng, L.; Shi, M.; Li, X.; Zhang, L.; Zhao, H. Tea polyphenols ameliorates memory decline in aging model rats by inhibiting brain TLR4/NF-κB inflammatory signaling pathway caused by intestinal flora dysbiosis. Exp. Gerontol. 2021, 111476. [Google Scholar] [CrossRef] [PubMed]
  123. Nan, S.; Wang, P.; Zhang, Y.; Fan, J. Epigallocatechin-3-Gallate Provides Protection Against Alzheimer’s Disease-Induced Learning and Memory Impairments in Rats. Drug Des. Dev. Ther. 2021, 15, 2013–2024. [Google Scholar] [CrossRef] [PubMed]
  124. Sarroca, S.; Gatius, A.; Rodríguez-Farré, E.; Vilchez, D.; Pallàs, M.; Griñán-Ferré, C.; Sanfeliu, C.; Corpas, R. Resveratrol confers neuroprotection against high-fat diet in a mouse model of Alzheimer’s disease via modulation of proteolytic mechanisms. J. Nutr. Biochem. 2021, 89, 108569. [Google Scholar] [CrossRef]
  125. Hussien, H.M.; Abd-Elmegied, A.; Ghareeb, D.A.; Hafez, H.S.; Ahmed, H.E.; Abd El-moneam, N. Neuroprotective effect of berberine against environmental heavy metals-induced neurotoxicity and Alzheimer’s-like disease in rats. Food Chem. Toxicol. 2018, 111, 432–444. [Google Scholar] [CrossRef]
  126. Singh, A.; Yau, Y.F.; Leung, K.S.; El-Nezami, H.; Lee, J.C.-Y. Interaction of polyphenols as antioxidant and anti-inflammatory compounds in brain–liver–gut axis. Antioxidants 2020, 9, 669. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The oxidative stress leads to neuroinflammation and neurodegenration. In homeostatic conditions, astrocytes release antioxidants by degrading reactive oxygen species, uptake and metabolism of neurotransmitters, provide energy and neurotrophin. In pathological conditions, astrocytes could be activated via stimulation from activated microglia. Therefore, high levels of oxidative stress activate signalling pathways that activate microglia and astrocyte (major glial inflammatory characters). Pro-inflammatory factors secreted by glial cells induce a neuroinflammatory response that disrupts the blood brain barrier’s integrity and infiltrates into the brain, secreting factors that lead to neurodegeneration, in which the most characteristic feature is neuron injury and death. iNOS: inducible nitric oxide synthase, COX-2: cyclooxigenase-2, NOX: NADPH oxidase, IL: interleukin, TNF-α: tumor necrosis factor alpha.
Figure 1. The oxidative stress leads to neuroinflammation and neurodegenration. In homeostatic conditions, astrocytes release antioxidants by degrading reactive oxygen species, uptake and metabolism of neurotransmitters, provide energy and neurotrophin. In pathological conditions, astrocytes could be activated via stimulation from activated microglia. Therefore, high levels of oxidative stress activate signalling pathways that activate microglia and astrocyte (major glial inflammatory characters). Pro-inflammatory factors secreted by glial cells induce a neuroinflammatory response that disrupts the blood brain barrier’s integrity and infiltrates into the brain, secreting factors that lead to neurodegeneration, in which the most characteristic feature is neuron injury and death. iNOS: inducible nitric oxide synthase, COX-2: cyclooxigenase-2, NOX: NADPH oxidase, IL: interleukin, TNF-α: tumor necrosis factor alpha.
Antioxidants 10 01370 g001
Figure 2. The microbiota-gut-brain axis is the bidirectional pathway between intestinal microbiota, the gut, and the central nervous system. It can be modulated by gut microbiota through endocrine (cortisol), neural (enteric and vagus nervous system), and immune (cytokines) systems. Microbial metabolites (LPS, GABA, SCFA, and PPG) and other neurotransmitters also participate in GM modulation. Gut dysbiosis can alter the tryptophan levels, hormones, SCFA, immune system, and gut permeability. Furthermore, release of cytokines and chemokines contribute to neuroinflammation and activate HPA axis (affecting gut permeability, barrier function, and immune cells through the secretion of cortisol). HPA axis: hypothalamic-pituitary-adrenal axis, ACTH: adrenocorticotropic hormone, CRF: corticotropin-releasing factor, LPS: lipopolysaccharides, GABA: y-aminobutyric acid, PPG: peptidoglycans, SCFA: short-chain fatty acids.
Figure 2. The microbiota-gut-brain axis is the bidirectional pathway between intestinal microbiota, the gut, and the central nervous system. It can be modulated by gut microbiota through endocrine (cortisol), neural (enteric and vagus nervous system), and immune (cytokines) systems. Microbial metabolites (LPS, GABA, SCFA, and PPG) and other neurotransmitters also participate in GM modulation. Gut dysbiosis can alter the tryptophan levels, hormones, SCFA, immune system, and gut permeability. Furthermore, release of cytokines and chemokines contribute to neuroinflammation and activate HPA axis (affecting gut permeability, barrier function, and immune cells through the secretion of cortisol). HPA axis: hypothalamic-pituitary-adrenal axis, ACTH: adrenocorticotropic hormone, CRF: corticotropin-releasing factor, LPS: lipopolysaccharides, GABA: y-aminobutyric acid, PPG: peptidoglycans, SCFA: short-chain fatty acids.
Antioxidants 10 01370 g002
Figure 3. Potential health-promoting effects of dietary polyphenols and their role in gut microbiota modulation. increase, decrease.
Figure 3. Potential health-promoting effects of dietary polyphenols and their role in gut microbiota modulation. increase, decrease.
Antioxidants 10 01370 g003
Table 1. Neurodegenerative diseases influenced by the change of GM (induced by oxidative stress and inflammation).
Table 1. Neurodegenerative diseases influenced by the change of GM (induced by oxidative stress and inflammation).
DiseaseStudyChange in GMFindingsReference
Major Depressive DisorderHuman (n = 36)Phylum Firmicutes and Actinobacteria were overrepresented, Bifidobacterium and Blautia at the genus level.Sucrose, starch and pentose phosphate metabolism were important pathways for depression via GM functions.[30]
Human (n = 90)Paraprevotella showed positive correlation while Clostridia, Clostridiales, Firmicutes, and the RF32 order negatively correlated with depression.Integrity intestinal and inflammation markers were linked with the response to treat the MDD.[31]
AnxietyHuman (n = 9)Fusobacterium, Ruminococcus gnavus, and Escherichia/ShigellaMicrobial richness and diversity.Enhanced gut permeability and the abundance of pro-inflammatory bacteria linked with neuroinflammation.[32]
Human (n = 36)Bacteroidaceae, Bacteroides, Betaproteobacteriales, Burkholderiaceae, Tyzzerella 3, Escherichia/Shigella, Hungatella,
Enterobacteriales, and Enterobacteriaceae.
The abundance of Ruminococcaceae_UCG-014, Eubacterium_coprostanoligenes group, and Prevotella_9 was negatively associated with anxiety severity and positively with anxiety reduction, whereas Escherichia/Shigella and Bacteroides was positively correlated with anxiety severity.[33]
Obsessive-Compulsive DisorderHuman (n = 43)species richness, evenness, and abundance of Anaerostipes, Odoribacter, and Oscillospira.C-reactive protein was increased that demonstrated mild to strong linkage with psychiatric symptomatology.[34]
Parkinson’s DiseaseHuman (n = 40)relative abundance of Ruminococcaceae and Rikenellaceae family and Barnesiella, Alistipes, Odoribacter, and Butyricimonas genera.Significant enhancement in genera from the Porphyromonadaceae family and decrease in the abundance of genera Blautia and Ruminococcus was observed in PD patients with compromised cognitive ability.[35]
Human (n = 111)Firmicutes enterotype Prevotella enterotype Increased intestinal inflammatory responses, reduced SCFA level, and shifts in microbiota-host interactions between earlier PD onset.[36]
SchizophreniaHuman (n = 194)Bacteroidetes, Firmicutes and ActinobacteriaMetabolic disturbance (levels of glucose, low-density lipid-cholesterol, high-density lipid-cholesterol, triglyceride, and homeostasis model assessment of insulin resistance) was observed in the patients.[37]
Bipolar DisorderHuman (n = 46)microbiota diversity, Clostridiaceae and CollinsellaDifferences in GM colonization may modulate metabolic and metabolomic alterations and other biological processes such as inflammation.[38]
Human (n = 53)Bacteroidetes, Actinobacteria and FirmicutesChange in GM can be a potential biomarker. [39]
DementiaHuman (n = 77)Clostridia, Clostridiales Ruminococcaceae, Firmicutes, and RuminococcusDecrease in indole-3-pyruvic acid and SCFA producing bacteria as a signature for discrimination and prediction of dementia.[40]
EpilepsyHuman (n = 40)Delftia, Campylobacter, Lautropia, Haemophilus, and Neisseria genera among Proteobacteria phylum and Leptotrichia and Fusobacterium genera among Fusobacteria phylumInflammation and autoimmune mechanisms due to the taxonomic drift and differences in the intestinal microbiota have a role in the etiology of epilepsy.[41]
Human (n = 44)Ruminococcus_g2 and Bacteroides finegoldii in drug-resistant group, Negativicutes from Firmicutes in drug-resistant group and Bifidobacterium in all patients.Alteration in GM can be a biomarker to evaluate and diagnose the treatment response in patients. [42]
Huntington’s DiseaseR6/2 HD miceabundance of Bacteroidetes and FirmicutesDifferent compositions of Bacteroides, Coprobacillus, Enterobacteriaceae, Lactobacillus, and Parabacteroides were found in diseased animals. [43]
: Higher/increased, : Lower/decreased, PD: Parkinson’s disease, SCFA: short-chain fatty acids, MDD: Major depressive disorder, GM: gut microbiota, R6/2 HD mice: expressing exon 1 of the Huntington’s disease gene, n: number of total patients taken part in the study but the columns of change in GM and findings are only representing the data of diseased ones.
Table 3. Potential role of polyphenols in AD and related findings.
Table 3. Potential role of polyphenols in AD and related findings.
PolyphenolsStudyFindingsReference
CurcuminAPP/PS1 double transgenic miceChange in Lactobacillaceae, Rikenellaceae, Prevotellaceae, and Bacteroidaceae at family level, and Bacteroides, Prevotella, and Parabacteroides at genus level. Curcumin reduced the Aβ plaques burden and improved the cognitive abilities.[114]
Quercetin-3-O-GlucuronideMice and SH-SY5Y CellsAmeliorated tau phosphorylation, and Aβ plaques. Restored CREB and brain-derived neurotrophic factor levels in the hippocampus, and gut dysbiosis.[115]
QuercetinAdult male albino ratsProtected and prevented neuronal damage in the hippocampus.[116]
RSV, QCT and APIHuman SK-N-BE and SH-SY5Y cellsReduced mitochondrial and peroxisomal dysfunction, 7KC-induced toxicity and cell death.[117]
LuteolinSprague-Dawley ratsDown-regulated the expression of BASE1 and NF-κB and reduced Aβ levels in the hippocampus and cortex. Moreover, increased antioxidant potential, and suppressed inflammation and lipid peroxide production.[118]
Palmitoylethanolamide and LuteolinSprague-Dawley ratsUp-regulated the gene expression of enzymes, pro-inflammatory cytokines, and reduction of mRNA levels.Moreover, inhibited the Aβ-induced astrogliosis and microgliosis. [119]
Bilberry AnthocyaninsSprague-Dawley ratsEnhanced the growth of Aspergillus oryzae, Bacteroidales-S24-7-group, Bacteroides, Clostridiaceae-1, Lactobacillus, and Lachnospiraceae_NK4A136_group and inhibited the growth of Verrucomicrobia and Euryarchaeota in aging rats.[120]
APP/PSEN1 transgenic AD miceDown-regulated the expression of inflammatory factors, chemokine receptor CX3CR1, serum and brain LPS. Reversed the brain, kidney, and liver injury caused by AD.[121]
Tea PolyphenolsAging model ratsPrevented memory decline and TLR4/NF-κB inflammatory signal pathway. Besides, significantly improved the composition and diversity of intestinal microflora, shape and function of epithelium, and brain inflammation.[122]
Epigallocatechin-3-GallateSprague-Dawley ratsDecreased the tau hyperphosphorylation in hippocampus and expression of BACE1 and Aβ1-42 by improving the antioxidant system, learning and memory function.[123]
ResveratrolAD transgenic 5XFADPrevented memory loss and reduced the amyloid burden and tau pathology.[124]
BerberineSprague-Dawley ratsProduction of COX-2, TNF-α, IL-12, IL-6 and IL-1β was normalized, inhibited the production of Aβ42 and evoked the formation of antioxidant Aβ40.[125]
AD: Alzheimer’s disease, RSV, QCT and API: Resveratrol, Quercetin, and Apigenin, BACE1: β-site amyloid precursor protein cleaving enzyme, NF-κB: nuclear factor-κB, Aβ: amyloid beta, TNF-α: tumor necrosis factor-α, IL: interleukin, LPS: lipopolysaccharide, COX-2: cyclooxygenase-2, CREB: cyclic AMP response element binding protein.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shabbir, U.; Tyagi, A.; Elahi, F.; Aloo, S.O.; Oh, D.-H. The Potential Role of Polyphenols in Oxidative Stress and Inflammation Induced by Gut Microbiota in Alzheimer’s Disease. Antioxidants 2021, 10, 1370. https://doi.org/10.3390/antiox10091370

AMA Style

Shabbir U, Tyagi A, Elahi F, Aloo SO, Oh D-H. The Potential Role of Polyphenols in Oxidative Stress and Inflammation Induced by Gut Microbiota in Alzheimer’s Disease. Antioxidants. 2021; 10(9):1370. https://doi.org/10.3390/antiox10091370

Chicago/Turabian Style

Shabbir, Umair, Akanksha Tyagi, Fazle Elahi, Simon Okomo Aloo, and Deog-Hwan Oh. 2021. "The Potential Role of Polyphenols in Oxidative Stress and Inflammation Induced by Gut Microbiota in Alzheimer’s Disease" Antioxidants 10, no. 9: 1370. https://doi.org/10.3390/antiox10091370

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