Next Article in Journal
Antioxidant, Anti-Inflammatory and Anti-Diabetic Activities of Tectona grandis Methanolic Extracts, Fractions, and Isolated Compounds
Next Article in Special Issue
Preparation and Evaluation of 6-Gingerol Derivatives as Novel Antioxidants and Antiplatelet Agents
Previous Article in Journal
GC-MS Profiling and Biomedical Applications of Essential Oil of Euphorbia larica Boiss.: A New Report
Previous Article in Special Issue
Generation of Highly Antioxidant Submicron Particles from Myrtus communis Leaf Extract by Supercritical Antisolvent Extraction Process
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Potential of Flavonoids and Flavonoid Metabolites in the Treatment of Neurodegenerative Pathology in Disorders of Cognitive Decline

1
Raymond Purves Laboratory, Institute of Bone and Joint Research, Kolling Institute of Medical Research, Faculty of Health and Science, University of Sydney at Royal North Shore Hospital, St. Leonards, NSW 2065, Australia
2
Graduate School of Biomedical Engineering, University of NSW, Sydney, NSW 2052, Australia
3
Sydney Medical School, Northern Campus, University of Sydney at Royal North Shore Hospital, St. Leonards, NSW 2065, Australia
Antioxidants 2023, 12(3), 663; https://doi.org/10.3390/antiox12030663
Submission received: 11 January 2023 / Revised: 23 February 2023 / Accepted: 1 March 2023 / Published: 7 March 2023

Abstract

:
Flavonoids are a biodiverse family of dietary compounds that have antioxidant, anti-inflammatory, antiviral, and antibacterial cell protective profiles. They have received considerable attention as potential therapeutic agents in biomedicine and have been widely used in traditional complimentary medicine for generations. Such complimentary medical herbal formulations are extremely complex mixtures of many pharmacologically active compounds that provide a therapeutic outcome through a network pharmacological effects of considerable complexity. Methods are emerging to determine the active components used in complimentary medicine and their therapeutic targets and to decipher the complexities of how network pharmacology provides such therapeutic effects. The gut microbiome has important roles to play in the generation of bioactive flavonoid metabolites retaining or exceeding the antioxidative and anti-inflammatory properties of the intact flavonoid and, in some cases, new antitumor and antineurodegenerative bioactivities. Certain food items have been identified with high prebiotic profiles suggesting that neutraceutical supplementation may be beneficially employed to preserve a healthy population of bacterial symbiont species and minimize the establishment of harmful pathogenic organisms. Gut health is an important consideration effecting the overall health and wellbeing of linked organ systems. Bioconversion of dietary flavonoid components in the gut generates therapeutic metabolites that can also be transported by the vagus nerve and systemic circulation to brain cell populations to exert a beneficial effect. This is particularly important in a number of neurological disorders (autism, bipolar disorder, AD, PD) characterized by effects on moods, resulting in depression and anxiety, impaired motor function, and long-term cognitive decline. Native flavonoids have many beneficial properties in the alleviation of inflammation in tissues, however, concerns have been raised that therapeutic levels of flavonoids may not be achieved, thus allowing them to display optimal therapeutic effects. Dietary manipulation and vagal stimulation have both yielded beneficial responses in the treatment of autism spectrum disorders, depression, and anxiety, establishing the vagal nerve as a route of communication in the gut-brain axis with established roles in disease intervention. While a number of native flavonoids are beneficial in the treatment of neurological disorders and are known to penetrate the blood–brain barrier, microbiome-generated flavonoid metabolites (e.g., protocatechuic acid, urolithins, γ-valerolactones), which retain the antioxidant and anti-inflammatory potency of the native flavonoid in addition to bioactive properties that promote mitochondrial health and cerebrovascular microcapillary function, should also be considered as potential biotherapeutic agents. Studies are warranted to experimentally examine the efficacy of flavonoid metabolites directly, as they emerge as novel therapeutic options.

Graphical Abstract

1. Introduction

The vagus nerve is the longest and most complex of the 12 cranial nerves in the human body and a major component of the parasympathetic nervous system. It provides autonomic control and functional regulation of internal organs, controlling such fundamental processes as digestion, the pulsatile behaviour of the heart (which controls blood circulation), and the behavior of the muscle systems that control the respiratory system [1,2,3]. The autonomic nervous system also controls reflex actions, such as coughing, sneezing, and swallowing, and coordinates with the sympathetic nervous system to achieve organ homeostasis. The vagus nerve has important roles as a line of communication between the gut microbiome and linked organ systems [4]. In infancy, the microbiome contributes to the education of the immune system by exposing it to a range of epitopes, leading to a diverse recognition system that can identify self- from non-self-preventing sensitivities to food epitopes in adulthood, and also to the development of auto-immune disorders and life-threatening allergies [5]. The microbiota, gut, and brain communicate through the vagus nerve In a bidirectional communication system in what has been termed the gut-brain axis (Figure 1) [5]. The vagus nerve is a mixed nerve containing 80% afferent and 20% efferent fibres that deliver important instructive information in the form of vesicular neurotransmitters using a sophisticated transport system [6]. Activated nerves transport neurotransmitters to the synaptic gap where neurotransmitters release transducer signals and motor functions mediated by neural networks. Bioactive compounds are transported to the brain by efferent vagal fibres to stimulate specific brain regions. Vagal stimulation has been used to treat neuropsychiatric disorders [7].
Although controversial, psychedelic drugs belong to a general class of compounds known as psychoplastogens, which robustly promote structural and functional neural plasticity in key neural circuits that in practice have been shown to be beneficial to brain health [8]. Progress in this branch of medicine has historically been hampered by legislation banning the use of such psychotropic medications [9]. Psychedelics are serotonin 2A receptor agonists that can lead to profound changes in perception, cognition, and mood, and display a potential in the treatment of mental health brain disorders that is unlike any other treatments currently available [10]. Psychedelics can produce sustained therapeutic benefit following a single administration, and also have broad therapeutic value and efficacy in the treatment of disorders, such as depression, post-traumatic stress, anxiety disorders, and addictive substance and alcohol abuse disorders [11]. A number of flavonoids have been identified with an ability to regulate neural functional properties of potential therapeutic value in the treatment of neurological disorders [10,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35]. One class of flavonoid metabolite (urolithins) shows particular promise in the treatment of neurodegenerative disorders and in the provision of general health and wellbeing [20,36,37,38]. Elligatannins are degraded to ellagic acid, which is further processed to the urolithins by gut bacteria. Human intestinal bacteria capable of producing isourolithin A from ellagic acid have been isolated [33].
Traditional complimentary medical herbal infusions have been used for centuries to treat pain symptoms in the treatment of headaches [39,40], as antipyretics in the treatment of fevers [41], and have also been shown to be beneficial in the treatment of neurological symptoms in disorders of functional cognitive decline [42,43,44]. In Chinese traditional medicine, the liver is considered to be of central importance in the regulation of the Qi vital life force, which in therapeutic procedures is re-directed through the meridians to organ systems to re-balance vital life forces [45,46] Chinese medicinal herbal preparations are considered to re-balance the harmony of the opposing life elements of the yin and yang [47,48]. The Qi represents the functional activities of the body classified as yin, while the vital control of these bodily functions is provided by the yang component [49]. The gut-liver-brain axis thus has a central role in Chinese medicinal doctrine and the benefits provided by Chinese herbal formulations. While there are no equivalent or plausible explanations of this abstract theory in Western medicine [50], the existence of a gut-liver, gut-lung, and gut-brain regulatory connection that exerts some measure of control over linked organ systems has received considerable attention [51,52,53,54,55,56,57,58]. Many studies have proposed gut-liver, gut-lung, and gut brain axes as potential routes of intervention in disease resolution. The vagus nerve provides communication between the brain and gut, facilitating cross-talk between the brain and gut microbiota, and is a major parasympathetic heart regulatory nerve. In the intestines, the vagus nerve regulates the contraction of smooth muscles and glandular secretions. The vagus nerve thus oversees crucial bodily functions, such as mood control, immune response, digestion, and heart rate [59,60].

2. Therapeutic Vagus Nerve Stimulation

The microbiota, gut, and brain bidirectionally communicate via the microbiota-gut-brain axis [5]. The vagus nerve transports microbiota metabolites through its efferent fibres to the CNS where a number of responses in neuronal cell populations occur. A cholinergic anti-inflammatory pathway in the vagus efferants dampens peripheral inflammation; flavonoid metabolites have antioxidant and anti-inflammatory properties and are bioavailable to neural cells thus positive outcomes can also be expected on brain cell populations. The vagus nerve of the parasympathetic nervous system oversees a vast array of crucial bodily functions, including control of mood, immune response, digestion, and heart rate [4]. It establishes a crucial connection between the brain and the gastrointestinal tract and sends information about the state of the inner organs to the brain, which aids in homeostasis of bodily functions. Vagus nerve stimulation is a promising supportive treatment for refractory depression, posttraumatic stress disorder, and inflammatory bowel disease, inhibits cytokine production, and positively effects beneficial monoaminergic brain signaling in psychiatric conditions, such as mood and anxiety [61,62,63], and also for the treatment of traumatically injured brain tissues [64,65,66,67,68,69]. Gut bacteria have beneficial effects on mood and anxiety through the bioactive factors they produce, which are transported by the vagus nerve to the brain. The transfer of information between the gut and the brain via the vagus nerve is a two-way communicative highway with afferent vagal fibers actually outnumbering efferent fibres to a significant degree [70,71]. Vagus nerve stimulation has been used to treat epilepsy [6,72,73] and depression [6,72,73,74,75], and to improve learning and memory [76]. Thus positive functional outcomes are achievable when the vagus nerve is used as a conduit to stimulate brain tissue [77].
Preclinical evidence firmly establishes bidirectional communication between the brain, gut, and the gut microbiome through at least three nerve communication channels [78]. The vagus nerve has a cholinergic anti-inflammatory pathway that dampens peripheral inflammation, decreases intestinal permeability, and may also modulate the microbiota cell populations [5]. A large number of studies highlight potential roles for microbial dysbiosis as a contributing factor in many chronic disorders [79]. The gut microbiota and brain communicate through the gut-brain axis [80], when disturbed this may contribute to the pathophysiology of neurodegenerative disorders [81,82,83]. Methylation of ingested dietary flavonoids increases their lipophilic character, facilitating transport by the vagal cholinergic pathway from the gut to the brain. Flavonoids have anti-inflammatory and antioxidant properties that inhibit neuroinflammation and improve brain health. Microbiome dysbiosis, including a low abundance of Faecalibacterium and Bacteroides sp. and decreased production of butyrate in the gut, may foster inflammation and may contribute to the underlying pathophysiology of bipolar disorder [84]. A disturbance in the autonomic nervous system may provoke and maintain gastrointestinal dysbiosis in autism spectrum disorder [85]. Emerging data has identified a link between gut microbiota dysbiosis and neurodegenerative disorders, such as PD, AD, and ALS [86]. Neuroinflammation is, therefore, now being increasingly recognized as a driver of neurodegenerative disease pathology [87,88]. Gut bacteria also have crucial roles to play in the maintenance and regulation of the immune system, thus alterations in gut microbial cell populations may detrimentally affect neuro-immune interactions, synaptic plasticity, and regulation of skeletal muscle activity. This opens up the possibility of translational interventional studies in the treatment of neurodegenerative disorders and the emergence of psychobiotic programmes [89,90,91,92].

3. Transporter Proteins in the Afferent Fibres of the Vagus Nerve

Gastrointestinal vagal afferent fibres outnumber efferent fibres in the vagus nerve, however, these convey sensory signals from the gastrointestinal tract to the brain. Numerous subtypes of gastrointestinal vagal afferents have been identified [93]. Stimulation of the vagus nerve has been used in the treatment of epilepsy and seizures, but also shows therapeutic potential in a range of other serious neurodegenerative disorders [4] and has found application in the treatment of inflammation [94], and also to combat the cytokine storm of ARDS in the COVID-19 disease [95]. Vagus nerve stimulation limits cytokine production and dampens systemic inflammation and inflammation-induced lung tissue damage [1]. Neurotransmitters are synthesized in the cytoplasm of nerves and by the gut microbiota and are transported in secretory vesicles in nerves for regulated release at synaptic membrane interfaces with communicating neural networks (Figure 2) [96]. Neurotransmission depends on the efficient regulation of the transport and release of chemical transmitter molecules. Neurotransmitters are packaged into specialized secretory vesicles in neurons and neuroendocrine cells, and these are transported by specific vesicular transporter proteins [97]. The vagus nerve contains transporter proteins that send amino acids and sugar nutrients to the brain generating satiety responses to hunger [98] and signals that regulate hunger responses/food intake and the production of gastric and pancreatic secretions [99]. Relatively little is published, however, on transport systems for flavonoid or flavonoid metabolites generated in the gut to the brain. Most of the flavonoid metabolites generated by the gut microbiota are of a similar size and chemical composition to neurotransmitter compounds and nutrient derived components, and thus may also be shuttled by these transporter systems in the efferent fibres of the vagus nerve to the brain (Figure 3). In vitro experiments show that many of the flavonoid and flavonoid metabolites have antioxidant and anti-inflammatory effects on neurons, stimulate the biogenesis of mitochondrial components, and have vasodilatory properties beneficial to the brain microvasculature. Transport of these gut components to the brain may thus be the therapeutic basis of vagal stimulation and its beneficial properties in the treatment of neurodegenerative conditions, and a link between the diet and autism, bipolar and other neurological disorders [4,51,57,59,80,81,83,84,88,92,100].
Neural transmitters are small bioactive molecules that are carried in synaptic vesicles in nerves. When a nerve is activated (Figure 2) a number of proteins transport and release neurotransmitters at the synaptic gap. Figure 3 shows some bioactive flavonoid metabolites that we propose are transported by nerves and have stimulatory effects when delivered to neural cell populations in the brain.

4. Neuroregulatory Properties of Flavonoids

Figure 2 demonstrates the crucial role of Ca2+ entry into neurons in their activation and transport of neurotransmitters that transduce signals in neural networks. Phenolic compounds (numbering in excess of 8000 compounds) have long been known to have medicinal properties. In this review, we concentrated on a sub-category of the phenolics flavonoids), which have been categorized into six sub-categories (Figure 4a,b). These have a generic 3 ring structure, as shown in Figure 4c.
Some flavones and flavone metabolites have been observed to modulate neural processes. Catechin and procyanidin flavan-3-ols are transported by nerves and effect adipose tissue mediated nerve activation [101]. The flavonoid isoliquiritigenin activates GABAB receptors, decreasing entry of Ca2+ into rat cerebrocortical nerves through voltage-gated Ca2+ channels, affecting glutamate transport and release from synaptic vesicles, and the transduction of neurotransmitters in neural networks [102]. Genistein isoflavone, a tyrosine kinase inhibitor, reduces Ca2+ influx through T-type CaV3.3 voltage-gated ion channels affecting nerve activation [103]. Dysfunction of T-type calcium channels is associated with epilepsy, neuropathic pain, cardiac problems, and major depressive disorders. Molecular agents that modulate the T channel function may thus be therapeutic. Baicalin ameliorates neuropathic pain by suppressing TRPV1 up-regulation and ERK phosphorylation in DRGs [104] and modulates the dopamine system, thus modulating behavior seen in attention deficit hyperactivity disorder [105]. The pharmacological properties of baicalin are associated with the synthesis, vesicular localization, transport, and release of dopamine from synaptic vesicles. Naringenin has antinociceptive analgesic effects through its ability to inhibit NaV1.8 voltage-gated sodium channels preventing nerve activation and the generation of neuropathic pain signals [106]. Green tea EGCG has vasodilatory effects, reduces blood pressure, and activates zebrafish TRPA1 channels in sensory neurons triggering CGRP release, a potent vasodilator [107]. Diabetic peripheral neuropathy and neuropathic pain are major public health issues impacting on quality of life. TRPV1 has a crucial role in nociceptive transmission of pathological pain. Baicalin is an antioxidant flavonoid whose analgesic effects on spinal neuropathic pain are apparently mediated through TRPV1 [108]. The excessive release of glutamate critically effects the neuropathology of acute and chronic brain disorders. Apigenin reduces presynaptic Ca2+ entry mediated by the Cav2.2 (N-type) and Cav2.1 (P/Q-type) channels, thereby inhibiting glutamate release from the rat hippocampal nerve terminals [109]. Myricetin, a natural neuroprotective flavonoid, also inhibits the release of glutamate from nerve terminals (synaptosomes) of the rat cerebral cortex through effects on Cav2.2 (N-type) and Cav2.1 (P/Q-type) channels by attenuating voltage-dependent Ca2+ entry and activation of nerves that generate pain responses [110]. Kaempferol-3-rhamnoside and rosmarinic acid also inhibit synaptic glutamate release, inhibiting nerve activation and generation of pain responses [111,112].
Neuroinflammation has detrimental effects on neurons and contributes to the pathology of neurodegeneration. The beneficial antioxidant properties of flavonoids [113,114] is attributable to their ability to inhibit lipoxygenase (LOX), cyclooxygenase (COX), myeloperoxidase (MPO), NADPH oxidase, and xanthine oxidase (XO). Flavonoids also stimulate free radical scavenging enzymes, such as superoxide dismutase (SOD) and catalase (CAT), which reduce levels of free radical oxygen species (ROS), including superoxide radical, hydroxyl radical, and singlet oxygen. This involves conversion of the superoxide ion into hydrogen peroxide by SOD, and this is converted into water and oxygen by CAT. The nuclear factor erythroid 2-related factor 2 (Nrf2)/antioxidant response element (ARE) pathway is an important cell signaling pathway responsible for the maintenance of redox homeostasis in humans [115,116]. Nrf2 is a master regulatory pleiotropic transcription factor that controls hundreds of genes in the phase II antioxidant response, controlling a multitude of cytoprotective genes responsive to oxidative stress and inflammation. Activation of Nrf2 produces antioxidant, anti-inflammatory, and neuroprotective effects, and is a critical component in the regulation of oxidative stress and anti-inflammatory responses in the CNS [117]. Luteolin, apigenin, quercetin, myricetin, rutin, naringenin, epicatechin, and genistein are all capable of activating the Nrf2/ARE pathway contributing to neuroprotection and the homeostasis of the CNS [118]. Table 1 illustrates further examples of flavonoids that also induce Nrf2 expression and its protective effects.
Quercetin occurs in plants as a glycosylated compound called rutin (Figure 5a), however, when ingested, rutin is converted to the aglycone form (Figure 5b) and modified by glucuronidation (Figure 5c) or sulfation (Figure 5d).

5. Natural Flavonoids Used in the Treatment of Neurodegenerative Conditions

Traditional Chinese medicine using herbal preparations have been used for centuries in complementary alternative medical practices [155,156]. Traditional Chinese herbal preparations are extremely complex mixtures of pharmacological agents often derived from up to seven different herbs. With the modern analytical techniques now available, attempts have been made to demystify these preparations to identify specific compounds and their mechanisms of action and to better understand their operational pharmacologic networks. The aim is to put these traditional medical practices on a more scientific basis to determine if they can be applied in Western medical practices. Network pharmacology, molecular docking, and in vitro cell-based investigations have identified a number of active components in these herbal preparations that could potentially provide a therapeutic effect [157,158,159,160,161].

6. Traditional Chinese Medicinal Formulations Used to Treat Alzheimer’s Disease

6.1. LeZhe

The Menispermaceae are small woody flowering climbing shrubs that contain a wide range of pharmacologically active benzylisoquinoline alkaloids, lignans, flavones, flavonols and pro-anthocyanidins. Tinospora sinensis is a member of the Menispermaceae family used in traditional Chinese medicine to treat AD. The formulation used, LeZhe, is a nerve calmative detoxifying antipyretic. Network pharmacology and molecular docking studies have identified LeZhe’s active compounds and molecular targets. Screening of DrugBank, Therapeutic Target Database and published AD studies have been used to identify pharmacological agents of interest. Kyoto Encyclopedia of Genes and Genomes (KEGG) target pathway enrichment analyses using Database for Annotation, and Visualization and Integrated Discovery (DAVID) have been undertaken and the neuroprotective properties of T. sinensis bioactive compounds have been evaluated in PC12 primary hippocampal neural cultures where injury had been induced using Aβ25-35. A total of 105 T. sinensis compounds and 38 molecular target proteins were identified. The main bioactive compounds of LeZhe include alkaloids such as berberine, a tetracyclic isoquinoline alkaloid derived from tyrosine, the aromatic amide aurantiomide, a quinazoline alkaloid; coumaroyl tyramine, hydroxycinnamic acid; trans-syringin, a β-D-glucoside derivative; and 3-dimethyl phillyrin phenylpropanoids. Phillyrin is a lignan produced by the endophytic fungus Paraconiothyrium sp. associated with the Chinese medicinal plant Forsythia suspensa with reported anti-pyretic detoxifier, antioxidant, anti-infective, anti-inflammatory, and antiviral properties [162] (Figure 6a–f). Many of these compounds can penetrate the blood brain barrier. Molecular targets of T. sinensis herbal compounds include Protein kinase B (AKT), Phosphoinositide 3-kinase (PI3K), Tyrosine-protein kinase JAK1 (JAK1), mammalian target of rapamycin (mTOR), TNF-α, Neuronal NOS, and the cholinergic function-related proteins, α4-Nicotinic acetylcholine receptor (α4 nAChR) and Muscarinic acetylcholine receptor M1 (Muscarinic M1). Inflammation and cholinergic dysfunction are targeted through PI3K/Akt, neurotrophic factor (NTF), Hypoxia-inducible factor 1 (HIF-1), mTOR, TNF and insulin resistance (IR) signalling pathways [160]. Significant improvement in PC12 cell survival and inhibition of apoptosis of Aβ25-35 injured primary hippocampal neuron cell cultures demonstrates the therapeutic potential of T. sinensis preparations in AD through a complex multi-compound-multi-target regulatory network however details still need to be unraveled of the mode of action of specific bioactive compounds [161]. Several bioactive flavonoid components have been identified in LeZhe preparations (Figure 6a–f) Berberine has anti-diabetic, anti-inflammatory properties, lowers blood sugar levels, causes weight loss and lowers blood pressure [163]. Berberine protects against TNF α induced inflammation in adipocytes [164] and is neuroprotective suppressing NF-κB-mediated neuroinflammation and pyroptosis [165]. Aurantiomides A-C isolated from the sponge-derived fungus Penicillium aurantiogriseum await detailed characterisation [166]. Syringin is a natural anti-inflammatory glucoside that attenuates NO production in LPS-stimulated RAW264.7 cells and has anti-oxidant and anti-cancer properties [167]. Phillyrin is a heterocyclic lignan glycoside flavonoid that attenuates TNF α-mediated insulin resistance and accelerated lipolysis by adipocytes [168,169].

6.2. Shuang-Huang-Lian Herbal Preparations

Shuang-Huang-Lian is listed in the Chinese pharmacopeia for the treatment of respiratory infections and has purported antiviral SARS-CoV-2 activity [170]. Baicalin and its aglycone form, baicalein, are two ingredients of Shuang-Huang-Lian herbal preparations (Figure 6g,h), and have been identified as BBB penetrating noncovalent, nonpeptidomimetic inhibitors of SARS-CoV-2. 3CLpro and may also be beneficial in the treatment of attention deficit hyperactivity disorder. Baicilin and baicalein are positive allosteric modulators of the benzodiazepine/non-benzodiazepine sites of the GABAA receptor, [171,172] providing anxiolytic [173,174,175] and anticonvulsant properties [176,177,178] and are neuroprotective prolyl endopeptidase inhibitors [179]. Prolyl endopeptidase/oligopeptidase (PEO) is implicated in a number of neurological disorders of the CNS, such as amnesia and stages of depression, and has roles in lithium sensitive signal transduction and depression [180,181]. PEO is implicated in neurodegeneration and neuroinflammation, and is considered a drug target for the enhancement of memory in dementia [182]. Inhibition of PEO reduces α-synuclein aggregation in PD [183]. increases α-synuclein degradation by neural cells [184]. and reduces α-synuclein toxicity [181]. In silico approaches inspired by the natural flavonoid baicilin, baicalein, and wogonin PEO inhibitors, are being used to produce synthetic PEO inhibitors of improved efficacy to reduce α-synuclein expression [185] (Figure 6i). A deficiency of PEO in mice reduces anxiety-like behavior and improves cognitive function [186], thus this approach is likely to be successful in the treatment of human neurological disorders. Such PEO inhibitors are of a small molecular weight similar to that of the neurotransmitters that are known to be transported by the vesicular transport system of nerves and are also expected to be transported by a similar mechanism.

6.3. Chaihu-Shugan-San

Chaihu-Shugan-San (CSS) is another well-known herbal antidepressant Chinese medicine that may also be beneficial in the treatment of cognitive dysfunction in AD [159]. Active compounds in CSS have been screened using the Traditional Chinese Medicine Systems Pharmacology database. Compound-related targets retrieved using the SwissTarget Prediction database identified major depressive disorder (MDD)-related targets using the DisGeNET Therapeutic Target and DrugBank databases. The identification of the active compounds in CSS affecting MDD targets has permitted the construction of a MDD target network in chronic unpredictable mild stress (CUMS) mice. Molecular docking established the binding affinities of these bioactive CSS compounds. Multi-target mechanisms of action of CSS compounds in network pharmacology identified a total of 152 active compounds, 520 predicted biological and 160 AD-specific targets. Sixty key targets providing beneficial effect in AD treatment were nuclear or cytoplasmic proteins with regulatory roles in PI3K-Akt, MAPK, and HIF signaling pathways in GO function and KEGG pathway enrichment analysis. Pre-treatment of PC12 neural cell cultures with CSS reduced Aβ-induced neural cell death and apoptosis. Increased phosphorylation of Akt and decreased pGSK3β/GSK3β levels in the hippocampus of CUMS mice established effects on PI3K/Akt signalling, and improved depressive-like behavior and neurogenesis of CSS in CUMS mice. Flavonoids identified in CSS include quercetin, luteolin and kaempferol; these warrant further examination in the treatment of AD.

6.4. Qingfei Paidu and Ma Xing Shi Gan

Qingfei Paidu and Ma Xing Shi Gan antiviral decoctions used to treat COVID-19 and AD in traditional Chinese medicine are also of considerable complexity. Molecular networking of mass spectrometry data has identified a number of bio-active flavone and chalcones present in these formulations [187] (Figure 7). Hesperidin, glycyrrhizic acid, baicalin, baicalein, naringin, phillyrin, quercetin, luteolin, kaempferol, licochalcone B, and mangiferin have all been identified in these formulations. Further studies are required to fully decipher all therapeutic bioactive component combinations and their interactions in the pharmacological networks. This initial study, however, made significant inroads into better understanding the complex therapeutic basis of these traditional Chinese herbal medications, but further work is required to fully understand how these components provide their therapeutic effect.

7. Complex Heterocyclic Polyphenolic Precursor Compounds That Are Processed by the Gut Microbiome Releasing Bioactive Metabolites

7.1. Eligatannins

Ellagitannins (ETs) are polyphenol compounds that are abundant in some fruits (blackberries, raspberries, strawberries), nuts (walnuts and almonds), and pomegranatesm, and have been used in complimentary medicine for centuries. ETs represent one of the most diverse groups of plant phenolics encompassing over 1000 natural bioactive compounds [188,189]. The gut microbiome converts ETs to ellagic acid (EA). EA has a variety of health benefits related to the protection it provides from oxidative stress [20,188,190,191,192]. EA is reported to have a low water solubility and bioavailability, however, when it is converted to urolithin A (UA) by the gut microbiome, UA retains the biological activities of EA and has high solubility and bioavailability (Figure 8). Urolithins are biologically active compounds exhibiting strong antioxidant effects [193,194,195,196] and anti-inflammatory [196,197] and neuroprotective properties [100,198]. Punicalagin, chebulinic acid, and chebulagic acid are complex polyphenolic ellagitannins that occur in pomegranate and are degraded to form EA by the gut microbiome (Figure 8). EA is further degraded to the urolithins; these are not synthesised or generated by mammalian cells and have antioxidant and anti-inflammatory properties [20,100,190,191,192,193,194,195,196,197,198]. Panduratin A is an antioxidant polycyclic chalcone phenolic that has also been identified in pomegranate and in the Thai medicinal plant Boesenbergia rotunda that has reported antiviral properties against SARS-CoV-2 [199].
EA is a candidate drug for the treatment of traumatic brain injury and neurodegenerative disorders, due to its neuroprotective properties mediated by inhibition of the PI3K/Akt/mTOR and Akt/IKK/NFκB signaling pathways, reducing neuroinflammation and enhancing autophagy [20,200,201,202,203]. The urolithins may be the bioactive metabolites that provide these beneficial therapeutic properties for EA [100,204,205,206]. EA can modulate the expression of the proinflammatory cytokines IL-1β,TNF-α, and IL-17 [20,193,196,197,206]. EA down regulates IL and lipid peroxidation, improves cognitive functions, and is provides neuroprotective benefits by scavenging free radicals and regulating antioxidant enzymes [100,198]. Urolithin species have neuroprotective properties through their antioxidant properties and ability to inhibit Aβ25-35-induced neurotoxicity and monoamine oxidase [20,201,207,208].

7.2. The Urolithins

Urolithin A is a benzocoumarin metabolite produced by the gut microbiome by digestion of ellagic acid and ellagitannins found in dietary pomegranates, strawberries, raspberries, and walnuts. Urolithin A does not occur freely in dietary foods, nor is it produced by mammalian enzyme systems [33,195]. Urolithin A is a natural prebiotic that promotes mitophagy, mitochondrial biogenesis, and metabolic function, impacting on muscle health in preclinical models of aging and in the elderly and middle-aged. Urolithin A improves mitochondrial function in the articular chondrocytes of diarthrodial joints, reducing disease progression in a mouse OA model, and inhibits cartilage degeneration, synovial inflammation, and the pain associated with this condition [36].

7.3. Hydroxybenzoic Acids

Protocatechuic acid has potent anti-inflammatory properties [209] and activates the master transcription factor, nuclear factor erythroid 2-related factor 2 (Nrf2), [210] through Jun kinase (JNK) modification of the Nrf2 signalling system [210]. Nrf2 binds to antioxidant response elements in the promoter regions of a large number of genes encoding cytoprotective proteins [211]. Activation of Nrf2 results in the induction of a large range of phase II detoxifying antioxidant enzyme systems [212] and also inhibits the NLR family pyrin domain containing 3 (NLRP3) inflammasome [213]. NLRP3 operates as part of the innate immune response as a pattern recognition receptor recognizing pathogen associated molecular patterns (PAMPs). Inflammasomes are multiprotein complexes in the innate immune system that induce inflammation in response to pathogenic organisms and stress. Activation of proinflammatory caspases, such as caspase-1, leads to an upregulation in proinflammatory cytokine levels, such as IL-1, -18, and -33, which promote neuroinflammation and pathological changes in brain tissues [214]. The NLRP3 inflammasome has important roles in the pathology of neurodegeneration and is a logical therapeutic target to alleviate the damaging aspects of neuroinflammation. Protocatechuic acid has significant potential in the inhibition of the NLRP3 inflammasome. Urolithin A is reported to improve mitochondrial and neuronal cell health [36,200,215,216,217].

8. Vasodilatory Flavonoids

Flavonoids exert positive beneficial effects on the cardiovascular system through their vasodilatory properties and ability to regulate apoptotic processes in the endothelium [218]. Hesperidin has been used for decades to treat vascular insufficiency in tissues [219].
The potential use of flavonoids and flavonoid metabolites to improve cerebrovascular circulation could prove to be useful to improve the treatment of neurodegenerative conditions.
Quercetin displays useful cardiovascular properties, however, its low bioavailability may limit its therapeutic application. The bioavailability of quercetin in the systemic circulation is low, with maximum plasma concentrations rarely exceeding 1 μM after consumption of 80–100 mg quercetin equivalents [220,221]. However when non-absorbed quercetin reaches the colon, it is subjected to further processing by the gut microbiome [218]. This includes C-ring cleavage, dihydroxylation, and decarboxylation, generating quercetin metabolites, such as 3,3-dihydroxyphenyl propionic acid and 3,4-dihydroxyphenyl acetic acid, which display vasodilatory properties in animal models and decrease arterial blood pressure [222,223].
EGCG catechins and epicatechins have beneficial effects on vascular function [224], cardioprotective effects through the reduction of systolic and diastolic blood pressure, and positive effects on the cerebrovascular circulation, which improves therapeutic treatment of neurodegenerative disorders [225,226]. Studies have also shown that flavonoid metabolites can have different biological and antioxidant properties and efficacy than the parent flavonoid. Modifications of the ingested flavonoid by methylation, glucuronidation, or sulphation also influences the biological activity of the flavonoid and has significant effects on their antioxidant and anti-inflammatory properties, and how they affect expression of cell-adhesive proteins [227]. NF-κB activation is the main transcription factor mediating TNFα-induced expression of inflammatory genes [228]. Pharmacological inhibitors of NF-κB activity, however, may also act through the stimulation of the Nrf2 pathway [229]. (-)-Epicatechin (EC) is metabolized by microbiota in the large intestine producing a major metabolite 5-(3′,4′-dihydroxyphenyl)-γ-valerolactone (3,4-diHPV). EC and 3,4-diHPV both activate Nrf2-mediated gene expression, however, 3,4-diHPV shows higher potency in the upregulation of Nrf2 gene expression than EC. Conversion of EC to 3,4-diHPV by the gut microbiota improves the overall health-promoting effects of EC consumption due to this ability to selectively promote Nrf2 pathway activation [230].

The Anthocyanidins

Anthocyanins are antioxidant plant flavonoids with reported beneficial health-promoting effects in a number of chronic diseases [231]. Studies investigating anthocyanin absorption by Caco-2 intestinal cells report very low absorption of these compounds. The gut microbiome, however, converts the anthocyanins to protocatechuic acid and phlorglucinaldehyde, and these may be the pharmacologic bioforms that exert the purported therapeutic effects of the anthocyanins [232] (Figure 9).
Alzheimer’s disease (AD) is a serious and progressive neurodegenerative disorder of the elderly. Genetic, environmental, and lifestyle factors are associated with the pathogenesis of AD, leading to deleterious effects on the brain’s neuronal cell population manifested as cognitive dysfunctions, behavioural disability, and psychological impairment. Accumulation of amyloid beta (Aβ) peptides and neurofibrillary tangles in AD-affected brains are hallmarks of this disease. Several reports indicate flavonoids improve cognitive functions, inhibit or delay the formation of pathological amyloid beta aggregates and neurofibrillary tangles, thus improving neural function.
Current research has uncovered that dietary use of flavonoid-rich food sources essentially increases intellectual abilities and postpones or hinders the senescence cycle and related neurodegenerative problems, including AD [233]. During AD pathogenesis, multiple signalling pathways are involved, and to target a single pathway may relieve the symptoms but not provide a permanent cure [233,234]. Flavonoids scavenge free radical species (ROS), however, upon reaction with ROS, the antioxidant capacity of flavonoids can become compromised. Recent evidence for at least some flavonoids shows that the oxidation of reactive phenolic residues can in fact enhance their antioxidant properties. This antioxidant activity arises from generation of metabolites that activate the Nrf2-Keap1 pathway [233,234], upregulating the cell’s endogenous antioxidant capacity, by the prevention of activation of prooxidant and proinflammatory NF-κB pathways [235]. Flavonoid metabolites, such as protocatechuic acid [236,237] and urolithin A [33,36,195] generated by the gut microbiome, also have potent direct antioxidant activities or provide mitochondrial protection by promoting mitochondrial biogenesis and metabolic activity, enhancing neural cell activity in the CNS in neurodegenerative conditions [203,238].

9. Natural Flavonoids and Multifunctional Analog Derivatives Used in Western Medicine to Treat Neurodegenerative Conditions

9.1. Hesperidin/Hesperitin

Hesperidin’s antioxidant, anti-inflammatory, and neuroprotective properties are useful in the treatment of neurodegenerative conditions [239] and have inspired the development of therapeutic multifunctional flavone and chalcone analogs of improved efficacy. Hesperitin also has considerable potential in the treatment of neurological disorders and has inspired the development of multifunctional agents of improved efficacy [26,233,234]. The central position of chalcones in medicinal chemistry, and its amenability to chemical modification, facilitates its use as a template for the development of multifunctional analog chalcone/flavone forms of improved efficacy in a number of depressive neurodegenerative disorders, including PD and AD. A multi-tier flavone screening protocol employing molecular docking for BACE1 inhibitory, and antiamyloidogenic and antioxidant activities, demonstrates hesperidin as a multi-potent phytochemical in AD therapeutics [240,241].
Hesperidin is a high affinity BACE1 inhibitor providing complete inhibition of amyloid fibril formation, moderate ABTS(+) radical scavenging, and strong hydroxyl radical scavenging activity [240]. Inhibition of BACE1 and Aβ aggregation occurs by binding close to the catalytic aspartate dyad-constraining BACE1, precluding APP recognition and inhibiting amyloid fibril formation, Aβ25-35 induced ROS generation, and mitochondrial dysfunction [242]. Mitochondrial dysfunction and oxidative stress both induce pathological neurodegenerative changes contributing to the development of AD [242]. Hesperidin inhibits Aβ-induced cognitive dysfunction, oxidative damage, and mitochondrial dysfunction in mice, reduces learning and memory deficits, and improves locomotor activity. Increased phosphorylation of GSK-3β by hesperidin, reduced mitochondrial dysfunction, and increased antioxidative defence improve cognitive function in the APPswe/PS1dE9 transgenic mouse model of AD [242]. Hesperidin also inhibits the development of neurodegenerative disease by elevating expression of neural growth factors and endogenous antioxidant defence, reducing the impact of neuroinflammatory and apoptotic pathways. A limited number of human clinical trials have shown that hesperidin-enriched dietary supplements significantly improved cerebral blood flow, cognition, and memory performance [243]. Cerebral ischaemic injury and degenerative pathology in AD are linked, hesperidin downregulates Bcl-2 and Akt/PI3K, protecting against Aβ25-35-induced apoptotic neurotoxic effects [243]. Oxidative stress and inflammation have pivotal roles in the pathophysiology of AD and are attenuated by hesperidin in APP/PS1 mice, resulting in a reduction in ROS, LPO, and increased activity of HO-1, SOD, catalase, and GSH. This inhibits neuroinflammation by decreasing TNF-α, C-reactive protein, MCP-1 levels, and NF-κB activity [244]. Phosphorylation of Akt and GSK-3β are decreased by hesperidin and RAGE expression is inhibited, while the enhanced phosphorylation of IκBα and the nuclear translocation of NF-κB/p65 in APP/PS1 mice evidences neuroprotective properties by suppressing neuroinflammation [245].

9.2. Kaempferol and Luteolin

Plant secondary metabolite inhibitors that target monoamine oxidases may be useful in the treatment of depressive neurodegenerative disorders such as PD and AD [246]. Kaempferol and luteolin are selective human MAO-A inhibitors [247,248].

10. The Potential of Flavonoids in Tissue Repair Processes in a Biodiverse Range of Diseases in Linked Organ Systems

Literally hundreds of in vitro studies have demonstrated the antioxidant and anti-inflammatory tissue protective properties of flavonoids. A number of flavonoids have also been shown to have neuroprotective and neuroregenerative properties. Preclinical studies in rodent, pig, and monkey models of AD, PD, HD, and ALS [34,249,250,251,252,253,254,255] have also demonstrated that flavonoids have properties that counter neuroinflammation, prevent the neurotoxic effects of pathological protein aggregates of amyloid and hyperphosphorylated tau protein, free radical generation from peroxidation of lipids, and that brain tissues are rich in phospholipids that are susceptible to oxygen radical release during neuroinflammation. Some flavonoids act as monoamine oxidase inhibitors, combat apoptosis, are neuroprotective, promote neurogenesis and memory, and reduce cognitive decline [256,257,258,259,260,261,262,263,264,265,266,267]. Examination of the clinical trials that have been conducted on flavonoids demonstrates their diverse areas of action and therapeutic potential.
Despite their low bioavailability, positive responses have nevertheless been observed with several flavonoids in many clinical trials indicating their therapeutic potential.

11. Flavonoid Clinical Trials

11.1. Hesperidin/Hesperitin Trials

Hesperidin has antioxidant tissue protective properties [268] and may improve cerebrovascular circulation, cognitive function, and the clinical manifestations associated with ocular disorders [269]. Hesperidin can improve vascular health and treat hypertension, improves cardiovascular function, has tissue protective properties in type 2 diabetes [270], and may be useful in the control of obesity, acute hemorrhoidal disease [271], muscle metabolism [272,273], and has skin antiaging properties [274].
Hesperidin, alone or in combination with other citrus flavonoids, such as diosmin, has been used to treat vascular defects, such as hemorrhoids, varicose veins, and poor circulation (venous stasis). Preclinical studies have also demonstrated its beneficial effects in the treatment of neurodegenerative disorders [275]. A review of preclinical trial data showed the beneficial neuropharmacological potential of hesperidin, including anticonvulsant, antidepressant, antioxidant, anti-inflammatory, memory, and locomotor enhancing activities [275].

11.2. Epigallocatechin Gallate Clinical Trials

A significant number of clinical trials have been conducted on the EGCG polyphenolic flavonoid component of green tea. An examination of a selection of these trials [265,276,277,278,279,280,281,282,283,284,285] amply demonstrate the diverse biological properties of EGCG and its potential therapeutic applications. EGCG can potentially improve cognition in children with Down syndrome [276]. EGCG prevents skin dermatitis in skin cancer patients receiving radiotherapy [277], and topical application of EGCG improves the treatment of vitilago [278]. EGCG improves surgical skin scarring reducing mast cell numbers, improving blood flow, angiogenesis and the elastin content of skin samples (ISRCTN70155584) [282]. In an international standard randomized controlled trial (registration number ISRCTN 18643079), EGCG improved scar repair, scar thickness, hydration, and elasticity [280]. EGCG acutely enhances muscle microvascular blood flow in healthy young adults [281]. Combination therapy of EGCG with hesperidin prevents obesity [282]. EGCG supplementation improves blood pressure, lipid profiles, plasma atherogenic index, and oxidative status in type 2 diabetes [283] when used in a controlled clinical trial on subjects receiving a high fat diet improved lipid profiles. EGCG has been described as a potent natural inhibitor of fatty acid synthase [284]. EGCG shows promise in an animal model of AD in the regulation of α-, β-, γ-secretase activity, inhibiting tau phosphorylation, has antioxidative, anti-inflammatory, antiapoptotic activity, and inhibits AChE activity, all contributing to EGCG’s neuroprotective properties [265]. A double-blind placebo-controlled phase I clinical trial of the cognitive effect of EGCG on Fragile X syndrome (TESXF; NCT01855971) also showed improved memory and cognition [285].
The gut microbiome generates metabolites from EGCG that improve cerebrovascular function and have therapeutic utility in the treatment of neurodegenerative disorders.
Epicatechin is known to improve cognitive functions, lowering the risk of AD or stroke, however, the biologically active molecular forms of epicatechin that are responsible are poorly understood. γ-Valerolactone metabolites of EGCG are biologically active and can simultaneously modulate the expression of protein-coding and non-coding genes to effect cellular regulation, effecting cell adhesion, cytoskeleton organization, focal adhesion, cell signaling pathways, regulation of endothelial cell permeability, and their interactions with immune cells [225]. Two major EGCG metabolites generated by the gut microbiome and detected in plasma are 5-(4′-hydroxyphenyl)-γ-valerolactone-3′-sulfate and 5-(4′-hydroxyphenyl)-γ-valerolactone-3′-O-glucuronide [286]. γ-valerolactones have high bioavailability and anti-inflammatory properties, decrease blood pressure [287], and improve cerebrovascular blood flow, improving cognitive impairment in neurological disorders [288]. Cerebrovascular dysfunction can accelerate brain atrophy with ageing, reduce cognitive capability, and lead to an increased risk of stroke and neurodegenerative diseases, such as AD and dementia. Flavonoids, including EGCG, have been shown in animal models [286,289,290] to maintain neurocognitive function in aging rats, decrease the risk of development of AD and stroke in humans, and exert beneficial effects on cerebrovascular blood flow in dementia [291,292,293].

11.3. Anthocyanin Clinical Trials

A randomised placebo-controlled trial of the effect of purified anthocyanins on cognition in individuals at increased risk for dementia has been undertaken [294]. A phase II clinical trial (ClinicalTrials.gov, NCT0341903) also assessed intervention strategies to prevent or delay the onset of dementia, and a further phase III trial of anthocyanins is also planned [295]. A trial has also examined the effect of dietary anthocyanins on endothelial function and arterial stiffness in individuals of excess body weight [296]. The consumption of anthocyanin is reported to improve memory in older adults with mild cognitive impairment [297,298]. The effects of anthocyanins on inflammatory and metabolic responses in a high-fat diet with cyanidin and delphinidin reported to exert beneficial effects in unhealthy diets [299]. Cyanadins were also reported to improve lipid profiles and lowered systemic inflammation in subjects with cardiovascular risk factors (ClinicalTrials.gov, NCT number: NCT04084847) [300]. The effects of anthocyanin supplementation on platelet function in subjects with dyslipidemia are shown to attenuate platelet function dyslipidemia [301]. The beneficial effects of berry anthocyanin consumption on cognitive performance, vascular function, and cardiometabolic risk markers uncovered in clinical trials has recently been reviewed [302]. A randomized controlled trial of consumption of tropical fruits rich in anthocyanins has also shown improvement in cognitive function, learning, memory, mental acuity, flexibility, and visual-motor skills in middle-aged women [303]. A clinical trial of anthocyanins has been shown to decrease concentrations of TNF-α in older adults with mild cognitive impairment (Australian New Zealand Clinical Trials Registry: ACTRN12618001184268) [304]. A cross-over, randomized, double-blind clinical trial (Australian New Zealand Clinical Trials Registry, identifier no. ACTRN12620000437965) showed anthocyanins attenuated vascular and inflammatory responses in overweight older adults [305].

11.4. Quercetin Clinical Trials

Quercetin exhibits many beneficial properties in cell and tissue protection in disease processes and optimal tissue function. A significant number of clinical trials have been undertaken examining the therapeutic efficacy of quercetin. These include potential roles in the treatment of cognitive function and cerebral blood flow [306] and modulation of the progression of AD [307] and CNS viral infection [308]. The efficacy of quercetin in muscle physiology has been evaluated [309] and its roles in the modulation of IGF-I and IGF-II levels following muscle damage [310]. Quercetin antioxidant effects have been examined in metabolic syndrome [311], and its efficacy in the treatment of blood pressure and endothelial dysfunction and regulation of lipid profiles and inflammatory biomarkers in metabolic syndrome [312,313]. A meta-analysis has been conducted on randomised controlled human trials assessing the impact of quercetin on systemic levels of inflammation [314]. Quercetin has been used to target IL-1β and suppress apoptosis in vascular endothelial cells in the treatment of atherosclerosis [315,316], in the treatment of cardiovascular disease [317], and in inflammatory processes effecting quality of life in post-myocardial infarction [318]. Quercetin has also been examined in inflammatory processes in polycystic ovary syndrome [319], combination therapies in the treatment of endometriosis [320], and in antiviral applications in COVID-19 [321]. Quercetin has been examined in combination therapy with green tea polyphenols in the treatment of prostate cancer [322], and the safety of quercetin supplementation assessed in the treatment of chronic obstructive pulmonary disease [323].

12. Bioactive Quercetin Metabolites

The gut microbiome members Escherichia coli, Streptococcus lutetiensis, Lactobacillus acidophilus, Weissella confusa, Enterococcus gilvus, Clostridium perfringens and Bacteroides fragilis have all been shown to degrade quercetin into a number of metabolites with C. perfringens and B. fragilis having the strongest degradative capability in vitro [324]. Peng et al. also demonstrated the presence of 3,4-dihydroxyphenylacetic acid production by C. perfringens and B. fragilis demonstrating their quercetin degradative capacity [325]. Fecal gut bacteria also degrade rutin as a substrate in-vitro releasing 3,4-dihydroxyphenylacetic acid as a metabolite [326].
Quercetin metabolites produced by C perfringens and B fragilis display a strong statistically significant inhibitory effect on HCT-116 human colorectal carcinoma cells. Weissella confusa produces quercetin metabolites with strong cytostatic tumor inhibitory activity over the growth of both A549 human lung carcinoma cells and HeLa cells comparable to or stronger than the tumor inhibitory activity displayed by intact quercetin but are more readily bio-available [327]. Eubacterium ramulus isolated from human feces is a strictly anaerobic bacterium of the gastrointestinal tract. E. ramulus cleaves the ring system of several flavonols and flavones giving rise to the corresponding hydroxyphenylacetic and hydroxyphenylpropionic acids, respectively, as well as acetate and butyrate. E ramulus generates 3,4-dihydroxyphenyl acetic acid from the biotransformation of quercetin in vitro and in vivo [328].
Neurodegeneration induced by the pesticide rotenone can be countered by quercetin in an animal model of PD [329,330], and in a transgenic model of AD [331], it reduced the neurotoxic effects of β-amyloidosis and decreased tauopathy in the hippocampus and amygdala, improving cognitive functional recovery. Quercetin is a multifunctional therapeutic in the treatment of neurodegenerative disorders [12,329,330,332]. Bioconversion of quercetin by gut bacteria generates bioactive quercetin metabolites, such as 3,4-dihydroxyphenylacetic acid and protocatechuic acid [333]. Protocatechuic acid is also a major metabolite of complex polyphenols, such as the anthocyanins and proanthocyanins [333,334,335,336]. Polyphenolic metabolites that also arise during flavonoid metabolism, such as 3,4-dihydroxyphenylacetic acid, can positively influence beneficial gut bacterial populations, such as Bifidobacterium spp., Lactobacillus spp. and Bacteroides spp., and inhibit colonization by the pathogenic bacteria Fusobacterium varium, Bilophila, and Enterobacteriaceae, thus promoting gut health [334] and enhancing the expression of several phase II drug-metabolizing enzymes that lower oxidative species in tissues.

13. Catechin Metabolites

Human phenyl-γ-valerolactone is a major metabolite of flavan-3-ols produced by gut bacteria. Phenyl-γ-valerolactone has neuroprotective properties and inhibits neurotoxic protein aggregate deposition, such as amyloid and tau in brain tissues [337], and promotes memory retention, preventing cognitive decline in an AD mouse model [338]. Phenyl-γ-valerolactone also improves endothelial cell function and cerebral blood flow [339,340,341]. Cerebral blood vessels are lined with endothelial cells and these form the blood–brain barrier (BBB). Endothelial dysfunction constitutes a crucial event in the pathophysiology of neurodegenerative disorders and cognitive impairment. Neuroinflammation can lead to neurodegeneration, endothelial cell dysfunction, defective cerebral blood flow, and deleterious effects on the permeability of the BBB. Phenyl-γ-valerolactone penetrates the BBB [342] and has functional attributes akin to other catechin catabolites that counter many of the earlier mentioned deleterious effects on brain tissues. Genomic and proteomic studies show that catechin metabolites have multimodal properties, modulating cellular pathways affecting cell adhesion, cytoskeletal organization, focal adhesion, endothelial permeability, and interaction with immune cells [225,226,343].
Flavonoids have therapeutic properties through their antioxidant and anti-inflammatory properties demonstrated in vitro. Some of the flavonoids can penetrate the blood–brain barrier from the systemic circulation to enter the brain directly, however, in general the bioavailability of intact flavonoids is limited due to poor absorption [29,344,345,346,347,348,349]. However, fortuitous generation of flavonoid metabolites by gut microbes that retain antioxidant and anti-inflammatory activity also needs to be considered in the overall therapeutic utility of these compounds [18,341,350,351,352,353,354]. Of the flavonoids, the isoflavones are the most bioavailable, however, anthocyanins and galloylated catechins are very poorly absorbed but can be converted into bioactive metabolites with therapeutic potential by gut microbes [345,346,347]. Gut microbes thus have important roles to play in the transformation and utilization of natural dietary flavonoids through the diverse enzyme systems that process these components in the gut [351,353,354]. Flavonoids generally cannot be metabolized effectively by human digestive enzyme systems but they can be transformed by enzymes produced by gut microorganisms into bioactive metabolites that can be transported by the gut-brain axis, or they can enter the systemic circulation from the gut and be transported to the brain where they more effectively penetrate the blood–brain barrier and thus have improved bioavailability and therapeutic utility [29,351,353,354]. Flavonoid metabolites that retain or exceed the antioxidant and anti-inflammatory capacity of the intact flavonoid indicate these have potent therapeutic potential. Furthermore, some flavonoid metabolites display biological activities not evident in the native flavonoid, which can be of therapeutic utility in the treatment of pathological neurodegenerative features in AD, PD, and HD by inhibiting the assembly and promoting the disassembly of protein aggregates in these disorders, reducing apoptosis of neurons and improving memory reducing cognitive decline in these neurodegenerative diseases [20,36,37,38,190,194,197,207,355,356,357].

14. Bioactive Flavonoid Metabolites and Regulation of Microbiome Bacterial Populations

The gut microbiome is a community of symbiotic microorganisms that inhabit the large intestine. These microbes have important roles to play in the maintenance of gut barrier integrity, inflammation, lipid and carbohydrate metabolism, immunity, and protection from pathogenic organisms. Colonization of the gut by pathogenic bacteria can lead to gut dysbiosis, significant alterations in gut bacterial populations, and an increase in the development of several diseases.

15. Metabolites Generated from Ellagic Acid with Bioactive Properties

Urolithin A has recently been approved as a functional food ingredient. Urolithin A and B have both been shown to improve metabolic functions and the maintenance of a healthy gut microbiome [358,359]. Uro-A and B also improve liver and kidney functions and induce the growth of Akkermansia muciniphila, a human mucin-degrading bacterium with health-promoting properties. Strategies have been developed to increase levels of Akkermansia muciniphila in the gut that counter obesity, diabetes, inflammation, and metabolic disorders [360,361,362]. A number of human and animal studies have shown that the abundance of A. muciniphila in the gut can be enhanced through dietary intervention. A muciniphila is available as a probiotic supplement [106].
Gordonibacter urolithinfaciens and Ellagibacter isourolithinifaciens are two human gut bacterial species that convert ellagic acid into urolithins [363].

16. Bioactive Quercetin Metabolites

Quercetin is a flavonoid that has been extensively examined in many studies that have demonstrated its antioxidant and anti-inflammatory properties in therapeutic applications. It is only relatively recently that interest has focused on quercetin metabolites and their biological properties [364]. The use of quercetin to treat OA rats has also been shown to influence gut bacterial populations with an elevation in the numbers of members of the Clostridia, Bacteroidia, and Bacilli families [365]. An increase in the number of quercetin metabolite species was also noted. A total of 94 human gut bacterial species have been examined for their ability to biotransform quercetin into different metabolites. Bacillus glycinifermentans, Flavonifractor plautii, Bacteroides eggerthii, Olsenella scatoligenes, and Eubacterium eligens were all shown to be capable of transforming quercetin into a number of metabolites displaying antiproliferative anticancer properties [366].

17. Flavonoids can Induce Neuroinflammation

While flavonoids have been shown to have many favourable properties in the alleviation of neuroinflammation and neurodegenerative processes [350,367,368], in a few cases gut bacteria have also been shown to detrimentally impact on inflammation in the brain, thus any prospective procedures conducted with flavonoids as therapeutics need to be carefully evaluated. In these cases of flavonoid-induced neuroinflammation, the balance of inflammatory cytokines in the gut and changes in intestinal and blood–brain barrier permeability can all produce detrimental impacts promoting the neuroinflammatory process. Having a healthy gut microbiome helps to prevent such detrimental effects on brain health [369,370,371].

18. Cellular Transport of Flavonoids by ATP-Binding Cassette (ABC)

Transporter Proteins and their Potential Roles in the Modulation of Cellular Influx/Efflux in Disease Processes

A total of 30 lactic acid bacterial strains transform punicalagin in pomegranate extracts into ellagic acid and urolithins [372]. Proteomic analysis showed that this resulted in an increase in transglycosylases with potential hydrolytic roles in the target phenolic compound. An increase in levels of ATP-binding cassette (ABC) transporters was also observed and these may be relevant cellular transporters for flavonoid metabolites. Nine ABC transporter genes have been identified with proposed roles in the transport of flavonoid metabolites in Salvia miltiorrhiza. It is proposed that these control the distribution of pigmentation in this plant genus, and ABC transporter proteins may have similar roles to play in mammalian tissues [373]. An increase in ABC transporters has been observed in sows fed a diet rich in fermented Chinese medicine herbal additives, which also resulted in an increase in the antibacterial and anti-inflammatory properties of the milk these animals produced, improving milk quality [374]. ABC transporter proteins have proposed roles in flavonoid transport in the suppression of colitis and its transformation into colon cancer induced by kaempferol, increasing its bioavailability and efficacy [375]. Streptococcus suis (S. suis) is a highly virulent zoonotic pathogen that causes severe economic losses in the swine industry and is a public health concern with the rise of antibacterial antibiotic resistant strains that may transfer up the food chain to humans [376]. EGCG has antibacterial and other health benefits, significantly reducing the hemolytic activity of S. suis, and has been suggested as a potential treatment of S. suis infection. Laboratory investigations have shown ABC transporters have active roles to play in the mechanism of action of EGCG [376]. Kaempferol displays antibacterial activity against H. pylori with an action comparable to that of clarithromycin and amoxicillin. ATP-binding cassette transporters, flagellar assembly, and fatty acid metabolism are the major pathways in which H. pylori cells are responsive to kaempferol treatment. ABC transporters thus have key roles to play in the antibacterial action of kaempferol [377]. Inhibition of drug-efflux membrane transporters by prenylated flavonoids and their interactions with azole antifungals has been suggested as an approach to chemosensitize multidrug-resistant C. albicans strains that otherwise can be difficult to treat clinically [378]. The antibacterial basis of flavonoids has been shown to be due to their disruptive effects on bacterial efflux pumps [379]. Furthermore, the multi-drug resistance conferred by the P-glycoprotein efflux pump is a major cause of failure of cancer chemotherapy treatments. The multi-drug resistance-reversing activity of isobavachalcone through inhibition of the action of P-glycoprotein thus holds promise in the development of more effective anticancer treatment strategies using specific flavonoids [380].

19. Bioavailability of Flavonoids

Many native flavonoids display beneficial neuro-therapeutic effects and are capable of penetrating the BBB, however, concerns have been raised on the poor bioavailability of some flavonoid members. In general, polyphenolic compounds show a low bioavailability due to their interaction with other dietary components, and with phase I and II metabolic processes mediated by the liver, intestine, and microbiota. However, bioactive flavonoid metabolites generated by the gut microbiome that retain antioxidant and antiphlogistic properties of the native flavonoid occur, and these are of therapeutic value [381]. Methylation of flavonoid aglycones upon ingestion may improve their bioavailability to cells compared to the native glycosylated form [382]. Native anthocyanidins of low bioavailability may be converted to metabolites with improved bioavailability and are more easily absorbed by the gut epithelium microcapilleries and by the stomach, kidney and liver [383]. Citrus flavonoids, such as hesperidin, naringin and nobiletin, display a number of health benefits, including antioxidative, anti-inflammatory, and neuroprotective properties, however, they also have limited bioavailability with a large proportion of dietary flavonoids remaining unabsorbed in the colon [384,385]. Fortunately the gut microbiota convert these into bioactive fragments that are more readily absorbed.
Protocatechuic acid, a simple phenolic acid, is one such example of a bioactive flavonoid metabolite (Figure 10). Protocatechuic acid remains in the circulation for significantly longer periods and at higher concentrations than parent flavonoids, and it easily crosses the blood–brain barrier. Experimental studies strongly support the role of protocatechuic acid in the prevention of neurodegenerative processes affecting AD and PD [386,387]. Protocatechuic acid inhibits detrimental processes leading to cognitive and behavioral impairment, including accumulation of β-amyloid plaques in brain tissue, hyperphosphorylation of tau protein in neurons, excessive ROS generation, and neuroinflammation. Growing evidence shows protocatechuic acid is efficacious in the treatment of neurodegeneration and a safe substance with antineurodegenerative compound warrant further investigation [237]. Protocatechuic acid is a widely distributed, naturally occurring flavonoid metabolite active pharmacological component with antioxidant and anti-inflammatory properties. Protocatechuic acid can be generated from a number of flavonoids. Over the past two decades, there have been an increasing numbers of publications demonstrating the importance of flavonoids and their metabolites in biomedical applications [236]. Protocatechuic acid can be produced from many flavonoid metabolites in hibiscus but has not been specifically examined as a biotherapeutic from this tissue source. Phenolic hibiscus extracts possess inhibitory activities against acetylcholinesterase, butyrylcholinesterase, monoamine oxidase, and ecto-5′ nucleotidase memory-enhancing, antineuroinflammatory, antioxidative properties [388].

20. Conversion of Quercetin to Bioactive Metabolites by the Gut Microbiome

Quercetin is processed by gut microbiome members to a number of metabolites with improved bioavailability which retain the antioxidant properties of quercetin (Figure 10).

21. Processing of Epigallocatechin Gallate by the Gut Microbiome

As already discussed, the gut microbiome can generate EGCG metabolites that improve cerebrovascular function and have therapeutic value in the treatment of neurodegenerative disorders (Figure 11). Epicatechin improves cognitive functions, lowering the risk of AD or stroke, however, the biologically active molecular forms of epicatechin responsible are poorly understood [389]. γ-Valerolactone metabolites of EGCG modulate cellular regulation, cytoskeleton organization, focal adhesion, cell signaling, endothelial cell permeability, and interactions with immune cells [225]. Two major EGCG metabolites generated by the gut microbiome, 5-(4′-hydroxyphenyl)-γ-valerolactone-3′-sulfate and 5-(4′-hydroxyphenyl)-γ-valerolactone-3′-O-glucuronide [286] have high bioavailability and anti-inflammatory potency, decrease blood pressure and improve cerebrovascular blood flow [288]. Cerebrovascular dysfunction accelerates brain atrophy, reduces cognitive capability and increases risk of stroke, AD and dementia. EGCG in animal models [286,289] maintains neurocognitive function in aging rats, decreases risk of AD, and improves cerebrovascular blood flow [291,292,293]. Gut microbiome members with EGCG transforming properties include Enterobacter aerogenes, Raoultella planticola, Klebsiella pneumoniae susp. pneumoniae, and Bifidobacterium longum subsp. Infantis [311].
Table 2 summarises the major beneficial properties of plant flavonoids and gut microbiome generated flavonoid metabolites clearly showing their therapeutic potential. particularly in the alleviation of symptoms generated in neurological disorders of cognitive decline. Studies examining these compounds as biomedicines is warranted based on the data uncovered in this review. There is some urgency in undertaking these studies, with the ever increasing incidence of neurological disorders in the ageing global general population. A global burden of disease study showed the overall burden of global neurological disorders has increased significantly in the last decade from 1990 to 2019 with 10 million deaths and 349 million disability-adjusted life years due to neurological disorders reported [390], and very significant projected increases in the incidence of AD and dementia calculated to increase from 57.4 million cases globally in 2019 to 152.8 million cases in 2050 [391].

22. Conclusions

The complexity of phenolic and flavonoid compounds of therapeutic utility in the prevention of tissue degeneration or infection is huge. The problems of bioavailability limiting the therapeutic utility of flavonoids may be overcome by potent flavonoid metabolites that retain the antioxidant and anti-inflammatory potency of the native flavonoid. Furthermore, new biological activities displayed by the flavonoid metabolite not evident in the native flavonoid may extend the therapeutic utility of this class of compound. Studies are warranted to examine this aspect of gut microbiome-generated flavonoid metabolites and may be particularly useful in the treatment of neurological disorders. Therapeutic probiotics may be a means of engineering microbiome members that produce the beneficial flavonoid metabolites outlined in this review as a means of selectively treating neurological disorders. Thus mood, anxiety, and neurological disorders that result in cognitive deficits and motor dysfunction may potentially be targeted using such an approach. Therapeutic nutraceuticals that enhance the levels of these beneficial flavonoid metabolites may also be an approach worth investigation to improve overall health and wellbeing.
The importance of maintaining a dominant population of beneficial gut symbionts to prevent establishment of pathogenic organisms in the gut microbiome [423,424] has recently become apparent in a study which showed Klebsiella aerogenes producing 3β-hydroxysteroid dehydrogenase degraded estradiol leading to depression in menopausal female mice [425].

Funding

This study was funded by Melrose Personal Research Fund, Sydney, Australia.

Acknowledgments

The author has received consultancy fees from Arthropharm and Sylvan Pharmaceutical companies. These companies had no input into the writing, design of experimental data or interpretation of this information or the decision to publish.

Conflicts of Interest

The author has no conflict to report.

References

  1. Yuan, H.; Silberstein, S.D. Vagus Nerve and Vagus Nerve Stimulation, a Comprehensive Review: Part III. Headache 2016, 56, 479–490. [Google Scholar] [CrossRef] [PubMed]
  2. Yuan, H.; Silberstein, S.D. Vagus Nerve and Vagus Nerve Stimulation, a Comprehensive Review: Part I. Headache 2016, 56, 71–78. [Google Scholar] [CrossRef]
  3. Yuan, H.; Silberstein, S.D. Vagus Nerve and Vagus Nerve Stimulation, a Comprehensive Review: Part II. Headache 2016, 56, 259–266. [Google Scholar] [CrossRef] [PubMed]
  4. Breit, S.; Kupferberg, A.; Rogler, G.; Hasler, G. Vagus Nerve as Modulator of the Brain-Gut Axis in Psychiatric and Inflammatory Disorders. Front. Psychiatry 2018, 9, 44. [Google Scholar] [CrossRef] [PubMed]
  5. Bonaz, B.; Bazin, T.; Pellissier, S. The Vagus Nerve at the Interface of the Microbiota-Gut-Brain Axis. Front. Neurosci. 2018, 12, 49. [Google Scholar] [CrossRef]
  6. George, M.; Sackeim, H.A.; Rush, A.J.; Marangell, L.B.; Nahas, Z.; Husain, M.M.; Lisanby, S.; Burt, T.; Goldman, J.; Ballenger, J.C. Vagus nerve stimulation: A new tool for brain research and therapy. Biol. Psychiatry 2000, 47, 287–295. [Google Scholar] [CrossRef] [PubMed]
  7. George, M.; Nahas, Z.; Borckardt, J.J.; Anderson, B.; Burns, C.; Kose, S.; Short, E.B. Vagus nerve stimulation for the treatment of depression and other neuropsychiatric disorders. Expert Rev. Neurother. 2007, 7, 63–74. [Google Scholar]
  8. Vargas, M.; Meyer, R.; Avanes, A.A.; Rus, M.; Olson, D.E. Psychedelics and Other Psychoplastogens for Treating Mental Illness. Front. Psychiatry 2021, 12, 727117. [Google Scholar] [CrossRef]
  9. Calder, A.; Hasler, G. Towards an understanding of psychedelic-induced neuroplasticity. Neuropsychopharmacology 2023, 48, 104–112. [Google Scholar] [CrossRef]
  10. Kwan, A.; Olson, D.E.; Preller, K.H.; Roth, B.L. The neural basis of psychedelic action. Nat. Neurosci. 2022, 25, 1407–1419. [Google Scholar] [CrossRef]
  11. Olson, D. Psychoplastogens: A Promising Class of Plasticity-Promoting Neurotherapeutics. J. Exp. Neurosci. 2018, 12, 1179069518800508. [Google Scholar] [CrossRef] [PubMed]
  12. Amanzadeh, E.; Esmaeili, A.; Rahgozar, S.; Nourbakhshnia, M. Application of quercetin in neurological disorders: From nutrition to nanomedicine. Rev. Neurosci. 2019, 30, 555–572. [Google Scholar] [CrossRef]
  13. Antoniolli, G.; Almeida, W.P.; Frias, C.C.; de Oliveira, T.B. Chalcones Acting as Inhibitors of Cholinesterases, β-Secretase and β- Amyloid Aggregation and other Targets for Alzheimer’s Disease: A Critical Review. Curr. Med. Chem. 2021, 28, 4259–4282. [Google Scholar] [CrossRef]
  14. Banc, R.; Rusu, M.E.; Filip, L.; Popa, D.S. The Impact of Ellagitannins and Their Metabolites through Gut Microbiome on the Gut Health and Brain Wellness within the Gut-Brain Axis. Foods 2023, 12, 270. [Google Scholar] [CrossRef] [PubMed]
  15. Calderaro, A.; Patanè, G.T.; Tellone, E.; Barreca, D.; Ficarra, S.; Misiti, F.; Laganà, G. The Neuroprotective Potentiality of Flavonoids on Alzheimer’s Disease. Int. J. Mol. Sci. 2022, 23, 14835. [Google Scholar] [CrossRef] [PubMed]
  16. Calis, Z.; Mogulkoc, R.; Baltaci, A.K. The Roles of Flavonols/Flavonoids in Neurodegeneration and Neuroinflammation. Mini. Rev. Med. Chem. 2020, 20, 1475–1488. [Google Scholar] [CrossRef] [PubMed]
  17. Carlessi, A.; Borba, L.A.; Zugno, A.I.; Quevedo, J.; Réus, G.Z. Gut microbiota-brain axis in depression: The role of neuroinflammation. Eur. J. Neurosci. 2021, 53, 222–235. [Google Scholar] [CrossRef]
  18. Evans, J.; Mendonca, P.; Soliman, K.F.A. Neuroprotective Effects and Therapeutic Potential of the Citrus Flavonoid Hesperetin in Neurodegenerative Diseases. Nutrients 2022, 14, 2228. [Google Scholar] [CrossRef]
  19. Flanagan, E.; Müller, M.; Hornberger, M.; Vauzour, D. Impact of Flavonoids on Cellular and Molecular Mechanisms Underlying Age-Related Cognitive Decline and Neurodegeneration. Curr. Nutr. Rep. 2018, 7, 49–57. [Google Scholar] [CrossRef]
  20. García-Villalba, R.; Giménez-Bastida, J.A.; Cortés-Martín, A.; Ávila-Gálvez, M.Á.; Tomás-Barberán, F.A.; Selma, M.V.; Espín, J.C.; González-Sarrías, A. Urolithins: A Comprehensive Update on their Metabolism, Bioactivity, and Associated Gut Microbiota. Mol. Nutr. Food Res. 2022, 66, e2101019. [Google Scholar] [CrossRef]
  21. Hole, K.; Williams, R.J. Flavonoids as an Intervention for Alzheimer’s Disease: Progress and Hurdles Towards Defining a Mechanism of Action. Brain Plast. 2021, 6, 167–192. [Google Scholar] [CrossRef] [PubMed]
  22. Islam, F.; Islam, M.M.; Khan Meem, A.F.; Nafady, M.H.; Islam, M.R.; Akter, A.; Mitra, S.; Alhumaydhi, F.A.; Emran, T.B.; Khusro, A.; et al. Multifaceted role of polyphenols in the treatment and management of neurodegenerative diseases. Chemosphere 2022, 307 Pt 3, 136020. [Google Scholar] [CrossRef]
  23. Islam, M.; Quispe, C.; Hossain, R.; Islam, M.T.; Al-Harrasi, A.; Al-Rawahi, A.; Martorell, M.; Mamurova, A.; Seilkhan, A.; Altybaeva, N.; et al. Neuropharmacological Effects of Quercetin: A Literature-Based Review. Front. Pharmacol. 2021, 12, 665031. [Google Scholar] [CrossRef] [PubMed]
  24. Jung, U.; Kim, S.R. Beneficial Effects of Flavonoids Against Parkinson’s Disease. J. Med. Food 2018, 21, 421–432. [Google Scholar] [CrossRef]
  25. Kempuraj, D.; Thangavel, R.; Kempuraj, D.D.; Ahmed, M.E.; Selvakumar, G.P.; Raikwar, S.P.; Zaheer, S.A.; Iyer, S.S.; Govindarajan, R.; Chandrasekaran, P.N.; et al. Neuroprotective effects of flavone luteolin in neuroinflammation and neurotrauma. Biofactors 2021, 47, 190–197. [Google Scholar] [CrossRef] [PubMed]
  26. Khan, A.; Ikram, M.; Hahm, J.R.; Kim, M.O. Antioxidant and Anti-Inflammatory Effects of Citrus Flavonoid Hesperetin: Special Focus on Neurological Disorders. Antioxidants 2020, 9, 609. [Google Scholar] [CrossRef] [PubMed]
  27. Khan, H.; Tundis, R.; Ullah, H.; Aschner, M.; Belwal, T.; Mirzaei, H.; Akkol, E.K. Flavonoids targeting NRF2 in neurodegenerative disorders. Food Chem. Toxicol. 2020, 146, 111817. [Google Scholar] [CrossRef]
  28. Koklesova, L.; Liskova, A.; Samec, M.; Zhai, K.; Al-Ishaq, R.K.; Bugos, O.; Šudomová, M.; Biringer, K.; Pec, M.; Adamkov, M.; et al. Protective Effects of Flavonoids Against Mitochondriopathies and Associated Pathologies: Focus on the Predictive Approach and Personalized Prevention. Int. J. Mol. Sci. 2021, 22, 8649. [Google Scholar] [CrossRef]
  29. Magni, G.; Riboldi, B.; Petroni, K.; Ceruti, S. Flavonoids bridging the gut and the brain: Intestinal metabolic fate, and direct or indirect effects of natural supporters against neuroinflammation and neurodegeneration. Biochem. Pharmacol. 2022, 205, 115257. [Google Scholar] [CrossRef]
  30. Meng-Zhen, S.; Ju, L.; Lan-Chun, Z.; Cai-Feng, D.; Shu-da, Y.; Hao-Fei, Y.; Wei-Yan, H. Potential therapeutic use of plant flavonoids in AD and PD. Heliyon 2022, 8, e11440. [Google Scholar] [CrossRef]
  31. Nouri, Z.; Fakhri, S.; El-Senduny, F.F.; Sanadgol, N.; Abd-ElGhani, G.E.; Farzaei, M.H.; Chen, J.T. On the Neuroprotective Effects of Naringenin: Pharmacological Targets, Signaling Pathways, Molecular Mechanisms, and Clinical Perspective. Biomolecules 2019, 9, 690. [Google Scholar] [CrossRef] [PubMed]
  32. Pan, L.; Cho, K.S.; Yi, I.; To, C.H.; Chen, D.F.; Do, C.W. Baicalein, Baicalin, and Wogonin: Protective Effects against Ischemia-Induced Neurodegeneration in the Brain and Retina. Oxid. Med. Cell. Longev. 2021, 2021, 8377362. [Google Scholar] [CrossRef] [PubMed]
  33. Selma, M.; Beltrán, D.; Luna, M.C.; Romo-Vaquero, M.; García-Villalba, R.; Mira, A.; Espín, J.C.; Tomás-Barberán, F.A. Isolation of Human Intestinal Bacteria Capable of Producing the Bioactive Metabolite Isourolithin A from Ellagic Acid. Front. Microbiol. 2017, 8, 1521. [Google Scholar] [CrossRef] [PubMed]
  34. Sergi, C. Epigallocatechin gallate for Parkinson’s disease. Clin. Exp. Pharmacol. Physiol. 2022, 49, 1029–1041. [Google Scholar] [CrossRef]
  35. Winter, A.; Bickford, P.C. Anthocyanins and Their Metabolites as Therapeutic Agents for Neurodegenerative Disease. Antioxidants 2019, 8, 333. [Google Scholar] [CrossRef]
  36. D’Amico, D.; Olmer, M.; Fouassier, A.M.; Valdés, P.; Andreux, P.A.; Rinsch, C.; Lotz, M. Urolithin A improves mitochondrial health, reduces cartilage degeneration, and alleviates pain in osteoarthritis. Aging Cell. 2022, 21, e13662. [Google Scholar] [CrossRef]
  37. D’Amico, D.; Andreux, P.A.; Valdes, P.; Singh, A.; Rinsch, C.; Auwerx, J. Impact of the Natural Compound Urolithin A on Health, Disease, and Aging. Trends Mol. Med. 2021, 27, 687–699. [Google Scholar] [CrossRef]
  38. Garcia-VIllalba, R.; Tomas-Berberan, F.A.; Iglesias-Aguire, C.E.; Gimenez-Bastida, J.A.; Gonzales-Sarrias, A.; Selma, M.V.; Espin, J.C. Ellagitannins, urolithins, and neuroprotection: Human evidence and the possible link to the gut microbiota. Mol. Asp. Med. 2023, 89, 101109. [Google Scholar] [CrossRef]
  39. Millstine, D.; Chen, C.Y.; Bauer, B. Complementary and integrative medicine in the management of headache. BMJ 2017, 357, j1805. [Google Scholar] [CrossRef]
  40. Wang, T.; Chen, H.; Xia, S.; Chen, X.; Sun, H.; Xu, Z. Ameliorative Effect of Parishin C Against Cerebral Ischemia-Induced Brain Tissue Injury by Reducing Oxidative Stress and Inflammatory Responses in Rat Model. Neuropsychiatr. Dis. Treat. 2021, 17, 1811–1823. [Google Scholar] [CrossRef]
  41. Ma, L.; Liu, H.M.; Luo, C.H.; He, Y.N.; Wang, F.; Huang, H.Z.; Han, L.; Yang, M.; Xu, R.C.; Zhang, D.K. Fever and Antipyretic Supported by Traditional Chinese Medicine: A Multi-Pathway Regulation. Front. Pharmacol. 2021, 12, 583279. [Google Scholar] [CrossRef]
  42. Friedemann, T.; Schumacher, U.; Tao, Y.; Leung, A.K.; Schröder, S. Neuroprotective Activity of Coptisine from Coptis chinensis (Franch). Evid. Based Complement. Altern. Med. 2015, 2015, 827308. [Google Scholar] [CrossRef]
  43. Srivastava, V.; Mathur, D.; Rout, S.; Mishra, B.K.; Pannu, V.; Rao, R.; Anand, A. Ayurvedic Herbal Therapies: A Review of Treatment and Management of Dementia. Curr. Alzheimer Res. 2022, 19, 568–584. [Google Scholar] [CrossRef]
  44. Yang, X.; Li, C.S.; Chen, C.; Tang, X.Y.; Cheng, G.Q.; Li, X. Protective effect of Shouwu Yizhi decoction against vascular dementia by promoting angiogenesis. Chin. J. Nat. Med. 2017, 15, 740–750. [Google Scholar] [CrossRef]
  45. Chang, S. The meridian system and mechanism of acupuncture: A comparative review. Part 3: Mechanisms of acupuncture therapies. Taiwan. J. Obstet. Gynecol. 2013, 52, 171–184. [Google Scholar] [CrossRef]
  46. Inanç, B. A New Theory on the Evaluation of Traditional Chinese Acupuncture Mechanisms from the Latest Medical Scientific Point of View. Acupunct. Electrother. Res. 2015, 40, 189–204. [Google Scholar] [CrossRef]
  47. Guo, Z. Chinese Confucian culture and the medical ethical tradition. J. Med. Ethics 1995, 21, 239–246. [Google Scholar] [CrossRef] [PubMed]
  48. Chen, Y. Chinese values, health and nursing. J. Adv. Nurs. 2001, 36, 270–273. [Google Scholar] [CrossRef]
  49. Wang, Q. Individualized medicine, health medicine, and constitutional theory in Chinese medicine. Front. Med. 2012, 6, 1–7. [Google Scholar] [CrossRef]
  50. Ye, X.; Dong, M.H. A review on different English versions of an ancient classic of Chinese medicine: Huang Di Nei Jing. J. Integr. Med. 2017, 15, 11–18. [Google Scholar] [CrossRef]
  51. Fairbrass, K.; Lovatt, J.; Barberio, B.; Yuan, Y.; Gracie, D.J.; Ford, A.C. Bidirectional brain-gut axis effects influence mood and prognosis in IBD: A systematic review and meta-analysis. Gut 2021, 71, 325985. [Google Scholar] [CrossRef] [PubMed]
  52. Mogilevski, T. The bi-directional role of the gut-brain axis in inflammatory and other gastrointestinal diseases. Curr. Opin. Gastroenterol. 2021, 37, 572–577. [Google Scholar] [CrossRef] [PubMed]
  53. Frausto, D.; Forsyth, C.B.; Keshavarzian, A.; Voigt, R.M. Dietary Regulation of Gut-Brain Axis in Alzheimer’s Disease: Importance of Microbiota Metabolites. Front. Neurosci. 2021, 15, 736814. [Google Scholar] [CrossRef] [PubMed]
  54. Menozzi, E.; Macnaughtan, J.; Schapira, A.H.V. The gut-brain axis and Parkinson disease: Clinical and pathogenetic relevance. Ann. Med. 2021, 53, 611–625. [Google Scholar] [CrossRef]
  55. Tang, Y.; Wang, Q.; Liu, J. Microbiota-gut-brain axis: A novel potential target of ketogenic diet for epilepsy. Curr. Opin. Pharmacol. 2021, 61, 36–41. [Google Scholar] [CrossRef]
  56. Matsubara, Y.; Kiyohara, H.; Teratani, T.; Mikami, Y.; Kanai, T. Organ and brain crosstalk: The liver-brain axis in gastrointestinal, liver, and pancreatic diseases. Neuropharmacology 2021, 205, 108915. [Google Scholar] [CrossRef]
  57. Zheng, Z.; Wang, B. The Gut-Liver Axis in Health and Disease: The Role of Gut Microbiota-Derived Signals in Liver Injury and Regeneration. Front. Immunol. 2021, 12, 775526. [Google Scholar] [CrossRef]
  58. Gokulan, K.; Joshi, M.; Khare, S.; Bartter, T. Lung microbiome, gut-lung axis and chronic obstructive pulmonary disease. Curr. Opin. Pulm. Med. 2021, 28, 134–138. [Google Scholar] [CrossRef]
  59. Zhang, J.; Ma, L.; Chang, L.; Pu, Y.; Qu, Y.; Hashimoto, K. A key role of the subdiaphragmatic vagus nerve in the depression-like phenotype and abnormal composition of gut microbiota in mice after lipopolysaccharide administration. Transl. Psychiatry 2020, 10, 186. [Google Scholar] [CrossRef]
  60. Kaelberer, M.; Buchanan, K.L.; Klein, M.E.; Barth, B.B.; Montoya, M.M.; Shen, X.; Bohórquez, D.V. A gut-brain neural circuit for nutrient sensory transduction. Science 2018, 361, eaat5236. [Google Scholar] [CrossRef]
  61. Ko, D. Transcutaneous vagus nerve stimulation (tVNS) as a potential therapeutic application for neurodegenerative disorders—A focus on dysautonomia in Parkinson’s disease. Auton. Neurosci. 2021, 235, 102858. [Google Scholar] [CrossRef] [PubMed]
  62. Sigurdsson, H.; Raw, R.; Hunter, H.; Baker, M.R.; Taylor, J.P.; Rochester, L.; Yarnall, A.J. Noninvasive vagus nerve stimulation in Parkinson’s disease: Current status and future prospects. Expert Rev. Med. Devices 2021, 18, 971–984. [Google Scholar] [CrossRef] [PubMed]
  63. Wang, Y.; Zhan, G.; Cai, Z.; Jiao, B.; Zhao, Y.; Li, S.; Luo, A. Vagus nerve stimulation in brain diseases: Therapeutic applications and biological mechanisms. Neurosci. Biobehav. Rev. 2021, 127, 37–53. [Google Scholar] [CrossRef] [PubMed]
  64. Abiega-Franyutti, P.; Freyre-Fonseca, V. Chronic consumption of food-additives lead to changes via microbiota gut-brain axis. Toxicology 2021, 464, 153001. [Google Scholar] [CrossRef]
  65. Srihagulang, C.; Vongsfak, J.; Vaniyapong, T.; Chattipakorn, N.; Chattipakorn, S.C. Potential roles of vagus nerve stimulation on traumatic brain injury: Evidence from in vivo and clinical studies. Exp. Neurol. 2022, 247, 113887. [Google Scholar] [CrossRef]
  66. Tang, Y.; Dong, X.; Chen, G.; Ye, W.; Kang, J.; Tang, Y.; Feng, Z. Vagus Nerve Stimulation Attenuates Early Traumatic Brain Injury by Regulating the NF-κB/NLRP3 Signaling Pathway. Neurorehabil Neural Repair 2020, 34, 831–843. [Google Scholar] [CrossRef]
  67. Neren, D.; Johnson, M.D.; Legon, W.; Bachour, S.P.; Ling, G.; Divani, A.A. Vagus Nerve Stimulation and Other Neuromodulation Methods for Treatment of Traumatic Brain Injury. Neurocrit Care 2016, 24, 308–319. [Google Scholar] [CrossRef]
  68. Lopez, N.; Krzyzaniak, M.J.; Costantini, T.W.; Putnam, J.; Hageny, A.M.; Eliceiri, B.; Coimbra, R.; Bansal, V. Vagal nerve stimulation decreases blood-brain barrier disruption after traumatic brain injury. J. Trauma Acute Care Surg. 2012, 72, 1562–1566. [Google Scholar] [CrossRef]
  69. Tang, H.; Li, J.; Zhou, Q.; Li, S.; Xie, C.; Niu, L.; Ma, J.; Li, C. Vagus nerve stimulation alleviated cerebral ischemia and reperfusion injury in rats by inhibiting pyroptosis via α7 nicotinic acetylcholine receptor. Cell. Death Discov. 2022, 8, 54. [Google Scholar] [CrossRef]
  70. Pitra, S.; Smith, B.N. Musings on the wanderer: What’s new in our understanding of vago-vagal reflexes? VI. Central vagal circuits that control glucose metabolism. Am. J. Physiol. Gastrointest. Liver Physiol. 2021, 320, G175–G182. [Google Scholar] [CrossRef]
  71. Travagli, R.; Hermann, G.E.; Browning, K.N.; Rogers, R.C. Musings on the wanderer: What’s new in our understanding of vago-vagal reflexes? III. Activity-dependent plasticity in vago-vagal reflexes controlling the stomach. Am. J. Physiol. Gastrointest. Liver Physiol. 2003, 284, G180. [Google Scholar] [CrossRef]
  72. Binnie, C. Vagus nerve stimulation for epilepsy: A review. Seizure 2000, 9, 161–169. [Google Scholar] [CrossRef]
  73. Van Laere, K.; Vonck, K.; Boon, P.; Brans, B.; Vandekerckhove, T.; Dierckx, R. Vagus nerve stimulation in refractory epilepsy: SPECT activation study. J. Nucl. Med. 2000, 41, 1145–1154. [Google Scholar]
  74. Aaronson, S.; Sears, P.; Ruvuna, F.; Bunker, M.; Conway, C.R.; Dougherty, D.D.; Reimherr, F.W.; Schwartz, T.L.; Zajecka, J.M. A 5-Year Observational Study of Patients with Treatment-Resistant Depression Treated With Vagus Nerve Stimulation or Treatment as Usual: Comparison of Response, Remission, and Suicidality. Am. J. Psychiatry 2017, 174, 640–648. [Google Scholar] [CrossRef]
  75. Rush, A.; George, M.S.; Sackeim, H.A.; Marangell, L.B.; Husain, M.M.; Giller, C.; Nahas, Z.; Haines, S.; Simpson, R.K., Jr.; Goodman, R. Vagus nerve stimulation (VNS) for treatment-resistant depressions: A multicenter study. Biol. Psychiatry 2000, 47, 276–286. [Google Scholar] [CrossRef]
  76. Clark, K.; Naritoku, D.K.; Smith, D.C.; Browning, R.A.; Jensen, R.A. Enhanced recognition memory following vagus nerve stimulation in human subjects. Nat. Neurosci. 1999, 2, 94–98. [Google Scholar] [CrossRef]
  77. Gershon, M.; Ratcliffe, E.M. Developmental biology of the enteric nervous system: Pathogenesis of Hirschsprung’s disease and other congenital dysmotilities. Semin. Pediatr. Surg. 2004, 13, 224–235. [Google Scholar] [CrossRef]
  78. Mayer, E.; Nance, K.; Chen, S. The Gut-Brain Axis. Annu. Rev. Med. 2022, 73, 439–453. [Google Scholar] [CrossRef]
  79. Lynch, S.; Pedersen, O. The human intestinal microbiome in health and disease. N. Engl. J. Med. 2016, 375, 2369–2379. [Google Scholar] [CrossRef]
  80. Cerovic, M.; Forloni, G.; Balducci, C. Neuroinflammation and the Gut Microbiota: Possible Alternative Therapeutic Targets to Counteract Alzheimer’s Disease? Front. Aging Neurosci. 2019, 11, 284. [Google Scholar] [CrossRef]
  81. Cenit, M.; Sanz, Y.; Codoñer-Franch, P. Influence of gut microbiota on neuropsychiatric disorders. World J. Gastroenterol. 2017, 23, 5486–5498. [Google Scholar] [CrossRef]
  82. Kobayashi, Y.; Sugahara, H.; Shimada, K.; Mitsuyama, E.; Kuhara, T.; Yasuoka, A.; Kondo, T.; Abe, K.; Xiao, J.-Z. Therapeutic potential of bifidobacterium breve strain A1 for preventing cognitive impairment in Alzheimer’s disease. Sci. Rep. 2017, 7, 13510. [Google Scholar] [CrossRef]
  83. Quigley, E. Microbiota-brain-gut axis and neurodegenerative diseases. Curr. Neurol. Neurosci. Rep. 2017, 17, 94. [Google Scholar] [CrossRef]
  84. Sublette, M.; Cheung, S.; Lieberman, E.; Hu, S.; Mann, J.J.; Uhlemann, A.C.; Miller, J.M. Bipolar disorder and the gut microbiome: A systematic review. Bipolar Disord. 2021, 23, 544–564. [Google Scholar] [CrossRef]
  85. Beopoulos, A.; Gea, M.; Fasano, A.; Iris, F. Autonomic Nervous System Neuroanatomical Alterations Could Provoke and Maintain Gastrointestinal Dysbiosis in Autism Spectrum Disorder (ASD): A Novel Microbiome-Host Interaction Mechanistic Hypothesis. Nutrients 2021, 14, 65. [Google Scholar] [CrossRef]
  86. Generoso, J.; Giridharan, V.V.; Lee, J.; Macedo, D.; Barichello, T. The role of the microbiota-gut-brain axis in neuropsychiatric disorders. Braz. J. Psychiatry 2021, 43, 293–305. [Google Scholar] [CrossRef]
  87. Cryan, J.; O’Riordan, K.J.; Cowan, C.S.M.; Sandhu, K.V.; Bastiaanssen, T.F.S.; Boehme, M.; Codagnone, M.G.; Cussotto, S.; Fulling, C.; Golubeva, A.V.; et al. The Microbiota-Gut-Brain Axis. Physiol. Rev. 2019, 99, 1877–2013. [Google Scholar] [CrossRef]
  88. Iannone, L.F.; Preda, A.; Blottière, H.M.; Clarke, G.; Albani, D.; Belcastro, V.; Carotenuto, M.; Cattaneo, A.; Citraro, R.; Ferraris, C.; et al. Microbiota-gut brain axis involvement in neuropsychiatric disorders. Expert Rev. Neurother. 2019, 19, 1037–1050. [Google Scholar] [CrossRef]
  89. Schächtle, M.; Rosshart, S.P. The Microbiota-Gut-Brain Axis in Health and Disease and Its Implications for Translational Research. Front. Cell. Neurosci. 2021, 15, 698172. [Google Scholar] [CrossRef]
  90. Larroya, A.; Pantoja, J.; Codoñer-Franch, P.; Cenit, M.C. Towards Tailored Gut Microbiome-Based and Dietary Interventions for Promoting the Development and Maintenance of a Healthy Brain. Front. Pediatr. 2021, 9, 705859. [Google Scholar] [CrossRef]
  91. Dicks, L.; Hurn, D.; Hermanus, D. Gut Bacteria and Neuropsychiatric Disorders. Microorganisms 2021, 9, 2583. [Google Scholar] [CrossRef] [PubMed]
  92. Barrio, C.; Arias-Sánchez, S.; Martín-Monzón, I. The gut microbiota-brain axis, psychobiotics and its influence on brain and behaviour: A systematic review. Psychoneuroendocrinology 2021, 137, 105640. [Google Scholar] [CrossRef]
  93. Wang, Y.; de Lartigue, G.; Page, A.J. Dissecting the Role of Subtypes of Gastrointestinal Vagal Afferents. Front. Physiol. 2020, 11, 643. [Google Scholar] [CrossRef] [PubMed]
  94. Caravaca, A.; Gallina, A.L.; Tarnawski, L.; Tracey, K.J.; Pavlov, V.A.; Levine, Y.A.; Olofsson, P.S. An Effective Method for Acute Vagus Nerve Stimulation in Experimental Inflammation. Front. Neurosci. 2019, 13, 877. [Google Scholar] [CrossRef] [PubMed]
  95. Mastitskaya, S.; Thompson, N.; Holder, D. Selective Vagus Nerve Stimulation as a Therapeutic Approach for the Treatment of ARDS: A Rationale for Neuro-Immunomodulation in COVID-19 Disease. Front. Neurosci. 2021, 15, 667036. [Google Scholar] [CrossRef] [PubMed]
  96. Liu, Y.; Edwards, R.H. The role of vesicular transport proteins in synaptic transmission and neural degeneration. Annu. Rev. Neurosci. 1997, 20, 125–156. [Google Scholar] [CrossRef]
  97. Varoqui, H.; Erickson, J.D. Vesicular neurotransmitter transporters. Potential sites for the regulation of synaptic function. Mol. Neurobiol. 1997, 15, 165–191. [Google Scholar] [CrossRef]
  98. Jordi, J.; Herzog, B.; Camargo, S.M.; Boyle, C.N.; Lutz, T.A.; Verrey, F. Specific amino acids inhibit food intake via the area postrema or vagal afferents. J. Physiol. 2013, 59, 5611–5621. [Google Scholar] [CrossRef]
  99. Tomé, D.; Schwarz, J.; Darcel, N.; Fromentin, G. Protein, amino acids, vagus nerve signaling, and the brain. Am. J. Clin. Nutr. 2009, 90, 838S–843S. [Google Scholar] [CrossRef]
  100. Yuan, T.; Ma, H.; Liu, W.; Niesen, D.B.; Shah, N.; Crews, R.; Rose, K.N.; Vattem, D.A.; Seeram, N.P. Pomegranate’s Neuroprotective Effects against Alzheimer’s Disease Are Mediated by Urolithins, Its Ellagitannin-Gut Microbial Derived Metabolites. ACS Chem. Neurosci. 2016, 7, 26–33. [Google Scholar] [CrossRef]
  101. Ishii, Y.; Muta, O.; Teshima, T.; Hirasima, N.; Odaka, M.; Fushimi, T.; Fujii, Y.; Osakabe, N. Repeated Oral Administration of Flavan-3-ols Induces Browning in Mice Adipose Tissues through Sympathetic Nerve Activation. Nutrients 2021, 13, 4214. [Google Scholar] [CrossRef] [PubMed]
  102. Lin, T.; Lu, C.W.; Hsieh, P.W.; Chiu, K.M.; Lee, M.Y.; Wang, S.J. Natural Product Isoliquiritigenin Activates GABAB Receptors to Decrease Voltage-Gate Ca2+ Channels and Glutamate Release in Rat Cerebrocortical Nerve Terminals. Biomolecules 2021, 11, 1537. [Google Scholar] [CrossRef] [PubMed]
  103. Rangel-Galván, M.; Rangel, A.; Romero-Méndez, C.; Dávila, E.M.; Castro, M.E.; Caballero, N.A.; Meléndez Bustamante, F.J.; Sanchez-Gaytan, B.L.; Meza, U.; Perez-Aguilar, J.M. Inhibitory Mechanism of the Isoflavone Derivative Genistein in the Human CaV3.3 Channel. ACS Chem. Neurosci. 2021, 12, 651–659. [Google Scholar] [CrossRef]
  104. Wang, Z.; Ling, D.; Wu, C.; Han, J.; Zhao, Y. Baicalin prevents the up-regulation of TRPV1 in dorsal root ganglion and attenuates chronic neuropathic pain. Vet. Med. Sci. 2020, 6, 1034–1040. [Google Scholar] [CrossRef] [PubMed]
  105. Zhou, R.; Wang, J.; Han, X.; Ma, B.; Yuan, H.; Song, Y. Baicalin regulates the dopamine system to control the core symptoms of ADHD. Mol. Brain 2019, 12, 11. [Google Scholar] [CrossRef] [PubMed]
  106. Zhou, Y.; Cai, S.; Moutal, A.; Yu, J.; Gómez, K.; Madura, C.L.; Shan, Z.; Pham, N.Y.N.; Serafini, M.J.; Dorame, A.; et al. The Natural Flavonoid Naringenin Elicits Analgesia through Inhibition of NaV1.8 Voltage-Gated Sodium Channels. ACS Chem. Neurosci. 2019, 10, 4834–4846. [Google Scholar] [CrossRef]
  107. Peixoto-Neves, D.; Soni, H.; Adebiyi, A. CGRPergic Nerve TRPA1 Channels Contribute to Epigallocatechin Gallate-Induced Neurogenic Vasodilation. ACS Chem. Neurosci. 2019, 10, 216–220. [Google Scholar] [CrossRef]
  108. Li, P.; Xiong, D.L.; Sun, W.P.; Xu, S.Y. Effects of baicalin on diabetic neuropathic pain involving transient receptor potential vanilloid 1 in the dorsal root ganglia of rats. Neuroreport 2018, 29, 1492–1498. [Google Scholar] [CrossRef]
  109. Chang, C.; Lin, T.Y.; Lu, C.W.; Wang, C.C.; Wang, Y.C.; Chou, S.S.; Wang, S.J. Apigenin, a natural flavonoid, inhibits glutamate release in the rat hippocampus. Eur. J. Pharmacol. 2015, 762, 72–81. [Google Scholar] [CrossRef]
  110. Chang, Y.; Chang, C.Y.; Wang, S.J.; Huang, S.K. Myricetin inhibits the release of glutamate in rat cerebrocortical nerve terminals. J. Med. Food 2015, 18, 516–523. [Google Scholar] [CrossRef]
  111. Lin, T.; Hung, C.F.; Weng, J.R.; Hsieh, T.Y.; Wang, S.J. Kaempferol 3-Rhamnoside on Glutamate Release from Rat Cerebrocortical Nerve Terminals Involves P/Q-Type Ca2+ Channel and Ca2+/Calmodulin-Dependent Protein Kinase II-Dependent Pathway Suppression. Molecules 2022, 27, 1342. [Google Scholar] [CrossRef] [PubMed]
  112. Wang, C.; Hsieh, P.W.; Kuo, J.R.; Wang, S.J. Rosmarinic Acid, a Bioactive Phenolic Compound, Inhibits Glutamate Release from Rat Cerebrocortical Synaptosomes through GABAA Receptor Activation. Biomolecules 2021, 11, 1029. [Google Scholar] [CrossRef] [PubMed]
  113. Murota, K.; Nakamura, Y.; Uehara, M. Flavonoid metabolism: The interaction of metabolites and gut microbiota. Biosci. Biotechnol. Biochem. 2018, 82, 600–610. [Google Scholar] [CrossRef] [PubMed]
  114. Suraweera, T.; Rupasinghe, H.P.V.; Dellaire, G.; Xu, Z. Regulation of Nrf2/ARE Pathway by Dietary Flavonoids: A Friend or Foe for Cancer Management? Antioxidants 2020, 9, 973. [Google Scholar] [CrossRef] [PubMed]
  115. Hiebert, P. The Nrf2 transcription factor: A multifaceted regulator of the extracellular matrix. Matrix Biol Plus 2021, 10, 100057. [Google Scholar] [CrossRef]
  116. Vomund, S.; Schäfer, A.; Parnham, M.J.; Brüne, B.; von Knethen, A. Nrf2, the Master Regulator of Anti-Oxidative Responses. Int. J. Mol. Sci. 2017, 18, 2772. [Google Scholar] [CrossRef]
  117. Heurtaux, T.; Bouvier, D.S.; Benani, A.; Helgueta Romero, S.; Frauenknecht, K.B.M.; Mittelbronn, M.; Sinkkonen, L. Normal and Pathological NRF2 Signalling in the Central Nervous System. Antioxidants 2022, 11, 1426. [Google Scholar] [CrossRef]
  118. Cores, Á.; Piquero, M.; Villacampa, M.; León, R.; Menéndez, J.C. NRF2 Regulation Processes as a Source of Potential Drug Targets against Neurodegenerative Diseases. Biomolecules 2020, 10, 904. [Google Scholar] [CrossRef]
  119. Guan, Y.; Wang, J.; Wu, X.; Song, L.; Wang, Y.; Gong, M.; Li, B. Quercetin reverses chronic unpredictable mild stress-induced depression-like behavior in vivo by involving nuclear factor-E2-related factor 2. Brain Res. 2021, 1772, 147661. [Google Scholar] [CrossRef]
  120. Lei, L.; Chai, Y.; Lin, H.; Chen, C.; Zhao, M.; Xiong, W.; Zhuang, J.; Fan, X. Dihydroquercetin Activates AMPK/Nrf2/HO-1 Signaling in Macrophages and Attenuates Inflammation in LPS-Induced Endotoxemic Mice. Front. Pharmacol. 2020, 11, 662. [Google Scholar] [CrossRef]
  121. Sun, L.; Xu, G.; Dong, Y.; Li, M.; Yang, L.; Lu, W. Quercetin Protects Against Lipopolysaccharide-Induced Intestinal Oxidative Stress in Broiler Chickens through Activation of Nrf2 Pathway. Molecules 2020, 25, 1053. [Google Scholar] [CrossRef] [PubMed]
  122. Song, X.; Tan, L.; Wang, M.; Ren, C.; Guo, C.; Yang, B.; Ren, Y.; Cao, Z.; Li, Y.; Pei, J. Myricetin: A review of the most recent research. Biomed. Pharmacother. 2021, 134, 111017. [Google Scholar] [CrossRef]
  123. Alshehri, A.; El-kott, A.F.; El-Gerbed, M.S.A.; El-Kenawy, A.E.; Albadrani, G.M.; Khalifa, H.S. Kaempferol prevents cadmium chloride-induced liver damage by upregulating Nrf2 and suppressing NF-κB and keap1. Environ. Sci. Pollut. Res. 2022, 29, 13917–13929. [Google Scholar] [CrossRef] [PubMed]
  124. Feng, Z.; Wang, C.; Yue Jin Meng, Q.; Wu, J.; Sun, H. Kaempferol induced GPER upregulation attenuates atherosclerosis via the PI3K/AKT/Nrf2 pathway. Pharm. Biol. 2021, 59, 1104–1114. [Google Scholar] [CrossRef] [PubMed]
  125. Talebi, M.; Talebi, M.; Farkhondeh, T.; Mishra, G.; İlgün, S.; Samarghandian, S. New insights into the role of the Nrf2 signaling pathway in green tea catechin applications. Phytother. Res. 2021, 35, 3078–3112. [Google Scholar] [CrossRef]
  126. Vendidandala, N.; Yin, T.P.; Nelli, G.; Pasupuleti, V.R.; Nyamathulla, S.; Mokhtar, S.I. Gallocatechin-silver nanoparticle impregnated cotton gauze patches enhance wound healing in diabetic rats by suppressing oxidative stress and inflammation via modulating the Nrf2/HO-1 and TLR4/NF-κB pathways. Life Sci. 2021, 286, 120019. [Google Scholar] [CrossRef]
  127. Zhang, Y.; Yang, Y.; Yu, H.; Li, M.; Hang, L.; Xu, X. Apigenin Protects Mouse Retina against Oxidative Damage by Regulating the Nrf2 Pathway and Autophagy. Oxidative Med. Cell. Longev. 2020, 2020, 9420704. [Google Scholar] [CrossRef]
  128. Albarakati, A.; Baty, R.S.; Aljoudi, A.M.; Habotta, O.A.; Elmahallawy, E.K.; Kassab, R.B.; Moneim, A.E.A. Luteolin protects against lead acetate-induced nephrotoxicity through antioxidant, antiinflammatory, anti-apoptotic, and Nrf2/HO-1 signaling pathways. Mol. Biol. Rep. 2020, 47, 2591–2603. [Google Scholar] [CrossRef]
  129. Alekhya Sita, G.; Gowthami, M.; Srikanth, G.; Krishna, M.M.; Rama Sireesha, K.; Sajjarao, M.; Nagarjuna, K.; Nagarjuna, M.; Chinnaboina, G.K.; Mishra, A. Protective role of luteolin against bisphenol A-induced renal toxicity through suppressing oxidative stress, inflammation, and upregulating Nrf2/ARE/ HO-1 pathway. IUBMB Life 2019, 71, 1041–1047. [Google Scholar] [CrossRef]
  130. Kang, K.; Piao, M.J.; Hyun, Y.J.; Zhen, A.X.; Cho, S.J.; Ahn, M.J.; Yi, J.M.; Hyun, J.W. Luteolin promotes apoptotic cell death via upregulation of Nrf2 expression by DNA demethylase and the interaction of Nrf2 with p53 in human colon cancer cells. Exp. Mol. Med. 2019, 51, 1–14. [Google Scholar] [CrossRef]
  131. Dai, J.; Guo, W.; Tan, Y.; Niu, K.; Zhang, J.; Liu, C.; Yang, X.-M.; Tao, K.-S.; Chen, Z.-N.; Dai, J.-Y. Wogonin alleviates liver injury in sepsis through Nrf2-mediated NF-κB signalling suppression. J. Cell. Mol. Med. 2021, 25, 5782–5798. [Google Scholar] [CrossRef] [PubMed]
  132. Yu, W.; Xu, Z.; Gao, Q.; Xu, Y.; Wang, B.; Dai, Y. Protective role of wogonin against cadmium induced testicular toxicity: Involvement of antioxidant, anti-inflammatory and anti-apoptotic pathways. Life Sci. 2020, 258, 118192. [Google Scholar] [CrossRef] [PubMed]
  133. Xingyue, L.; Shuang, L.; Qiang, W.; Jinjuan, F.; Yongjian, Y. Chrysin Ameliorates Sepsis-Induced Cardiac Dysfunction through Upregulating Nfr2/Heme Oxygenase 1 Pathway. J. Cardiovasc. Pharmacol. 2021, 77, 491–500. [Google Scholar] [CrossRef]
  134. Yuvaraj, S.; Ramprasath, T.; Saravanan, B.; Vasudevan, V.; Sasikumar, S.; Selvam, G.S. Chrysin attenuates high-fat-diet-induced myocardial oxidative stress via upregulating eNOS and Nrf2 target genes in rats. Mol. Cell. Biochem. 2021, 476, 2719–2727. [Google Scholar] [CrossRef]
  135. Yu, Z.; Yang, L.; Deng, S.; Liang, M. Daidzein ameliorates LPSinduced hepatocyte injury by inhibiting inflammation and oxidative stress. Eur. J. Pharmacol. 2020, 885, 173399. [Google Scholar] [CrossRef] [PubMed]
  136. Guo, J.; Yang, G.; He, Y.; Xu, H.; Fan, H.; An, J.; Zhang, L.; Zhang, R.; Cao, G.; Hao, D.; et al. Involvement of α7nAChR in the Protective Effects of Genistein against β-Amyloid-Induced Oxidative Stress in Neurons via a PI3K/Akt/Nrf2 Pathway-Related Mechanism. Cell. Mol. Neurobiol. 2021, 41, 377–393. [Google Scholar] [CrossRef] [PubMed]
  137. Wang, L.; Li, A.; Liu, Y.; Zhan, S.; Zhong, L.; Du, Y.; Xu, D.; Wang, W.; Huang, W. Genistein protects against acetaminophen-induced liver toxicity through augmentation of SIRT1 with induction of Nrf2 signalling. Biochem. Biophys. Res. Commun. 2020, 527, 90–97. [Google Scholar] [CrossRef]
  138. Yi, S.; Chen, S.; Xiang, J.; Tan, J.; Huang, K.; Zhang, H.; Wang, Y.; Wu, H. Genistein exerts a cell-protective effect via Nrf2/HO-1/ /PI3K signaling in Ab25-35-induced Alzheimer’s disease models in vitro. Folia Histochem. Cytobiol. 2021, 59, 49–56. [Google Scholar] [CrossRef]
  139. Guo, K.; Ren, J.; Gu, G.; Wang, G.; Gong, W.; Wu, X.; Ren, H.; Hong, Z.; Li, J. Hesperidin Protects against Intestinal Inflammation by Restoring Intestinal Barrier Function and up-Regulating Treg Cells. Mol. Nutr. Food Res. 2019, 63, 1800975. [Google Scholar] [CrossRef]
  140. Xin, X.; Li, Y.; Liu, H. Hesperidin ameliorates hypobaric hypoxiainduced retinal impairment through activation of Nrf2/HO-1 pathway and inhibition of apoptosis. Sci. Rep. 2020, 10, 19426. [Google Scholar] [CrossRef]
  141. Chen, Y.; Kong, L.; Tang, Z.; Zhang, Y.; Liu, Y.; Wang, T.; Liu, Y. Hesperetin ameliorates diabetic nephropathy in rats by activating Nrf2/are/glyoxalase 1 pathway. Biomed. Pharmacother. 2019, 111, 1166–1175. [Google Scholar] [CrossRef] [PubMed]
  142. Li, J.; Wang, T.; Liu, P.; Yang, F.; Wang, X.; Zheng, W.; Sun, W. Hesperetin ameliorates hepatic oxidative stress and inflammation via the PI3K/AKT-Nrf2-ARE pathway in oleic acid-induced HepG2 cells and a rat model of high-fat diet-induced NAFLD. Food Funct. 2021, 12, 3898–3918. [Google Scholar] [CrossRef]
  143. Lin, Z.; Fu, C.; Yan, Z.; Wu, Y.; Zhan, J.; Lou, Z.; Liao, X.; Pan, J. The protective effect of hesperetin in osteoarthritis: An in vitro and in vivo study. Food Funct. 2020, 11, 2654–2666. [Google Scholar] [CrossRef]
  144. Chen, W.; Ye, Y.; Wu, Z.; Lin, J.; Wang, Y.; Ding, Q.; Yang, X.; Yang, W.; Lin, B.; Lin, B. Temporary Upregulation of Nrf2 by Naringenin Alleviates Oxidative Damage in the Retina and ARPE-19 Cells. Oxidative Med. Cell. Longev. 2021, 2021, 4053276. [Google Scholar] [CrossRef] [PubMed]
  145. Tseng, Y.; Hsu, H.; Lee, T.; Chang, W.; Lo, Y. Naringenin, a dietary flavanone, enhances insulin-like growth factor 1 receptormediated antioxidant defense and attenuates methylglyoxalinduced neurite damage and apoptotic death. Nutr. Neurosci. 2021, 24, 71–81. [Google Scholar] [CrossRef] [PubMed]
  146. Sharath Babu, G.; Anand, T.; Ilaiyaraja, N.; Khanum, F.; Gopalan, N. Pelargonidin Modulates Keap1/Nrf2 Pathway Gene Expression and Ameliorates Citrinin-Induced Oxidative Stress in HepG2 Cells. Front. Pharmacol. 2017, 8, 868. [Google Scholar] [CrossRef] [PubMed]
  147. Rahman, S.; Mathew, S.; Nair, P.; Ramadan, W.S.; Vazhappilly, C.G. Health benefits of cyanidin-3-glucoside as a potent modulator of Nrf2-mediated oxidative stress. Inflammopharmacology 2021, 29, 907–923. [Google Scholar] [CrossRef]
  148. Kuo, H.; Wu, R.; Li, S.; Yang, A.Y.; Kong, A.N. Anthocyanin Delphinidin Prevents Neoplastic Transformation of Mouse Skin JB6 P+ Cells: Epigenetic Re-activation of Nrf2-ARE Pathway. AAPS J. 2019, 21, 83. [Google Scholar] [CrossRef]
  149. Zheng, S.; Deng, Z.; Chen, F.; Zheng, L.; Pan, Y.; Xing, Q.; Tsao, R.; Li, H. Synergistic antioxidant effects of petunidin and lycopene in H9c2 cells submitted to hydrogen peroxide: Role of Akt/Nrf2 pathway. J. Food Sci. 2020, 85, 1752–1763. [Google Scholar] [CrossRef]
  150. Qi, W.; Boliang, W.; Xiaoxi, T.; Guoqiang, F.; Jianbo, X.; Gang, W. Cardamonin protects against doxorubicin-induced cardiotoxicity in mice by restraining oxidative stress and inflammation associated with Nrf2 signaling. Biomed. Pharmacother. 2020, 122, 109547. [Google Scholar] [CrossRef]
  151. Xie, C.; Ma, H.; Shi, Y.; Li, J.; Wu, H.; Wang, B.; Shao, Z.; Huang, C.; Chen, J.; Sun, L.; et al. Cardamonin protects nucleus pulposus cells against IL-1β-induced inflammation and catabolism via Nrf2/NF-κB axis. Food Funct. J. 2021, 12, 2703–2714. [Google Scholar] [CrossRef] [PubMed]
  152. Chen, X.; Li, Z.; Hong, H.; Wang, N.; Chen, J.; Lu, S.; Zhang, H.; Zhang, X.; Bei, C. Xanthohumol suppresses inflammation in chondrocytes and ameliorates osteoarthritis in mice. Biomed. Pharmacother. 2021, 137, 111238. [Google Scholar] [CrossRef] [PubMed]
  153. Wang, W.; Chen, Z.; Zheng, T.; Zhang, M. Xanthohumol alleviates T2DM-induced liver steatosis and fibrosis by mediating the NRF2/RAGE/NF-κB signaling pathway. Future Med. Chem. 2021, 13, 2069–2081. [Google Scholar] [CrossRef]
  154. Liu, X.; Zhu, Q.; Zhang, M.; Yin, T.; Xu, R.; Xiao, W.; Wu, J.; Deng, B.; Gao, X.; Gong, W.; et al. Isoliquiritigenin Ameliorates Acute Pancreatitis in Mice via Inhibition of Oxidative Stress and Modulation of the Nrf2/HO-1 Pathway. Oxid. Med. Cell. Longev. 2018, 2018, 7161592. [Google Scholar] [CrossRef] [PubMed]
  155. Lu, A.; Jia, H.W.; Xiao, C.; Lu, Q.P. Theory of traditional Chinese medicine and therapeutic method of diseases. World J. Gastroenterol. 2004, 10, 1854–1856. [Google Scholar] [CrossRef]
  156. Jiang, W. Therapeutic wisdom in traditional Chinese medicine: A perspective from modern science. Trends Pharmacol. Sci. 2005, 26, 558–563. [Google Scholar] [CrossRef]
  157. Choudhry, N.; Zhao, X.; Xu, D.; Zanin, M.; Chen, W.; Yang, Z.; Chen, J. Chinese Therapeutic Strategy for Fighting COVID-19 and Potential Small-Molecule Inhibitors against Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2). J. Med. Chem. 2020, 63, 13205–13227. [Google Scholar] [CrossRef]
  158. Li, Y.; Chu, F.; Li, P.; Johnson, N.; Li, T.; Wang, Y.; An, R.; Wu, D.; Chen, J.; Su, Z.; et al. Potential effect of Maxing Shigan decoction against coronavirus disease 2019 (COVID-19) revealed by network pharmacology and experimental verification. J. Ethnopharmacol. 2021, 271, 113854. [Google Scholar] [CrossRef]
  159. Zeng, Q.; Li, L.; Siu, W.; Jin, Y.; Cao, M.; Li, W.; Chen, J.; Cong, W.; Ma, M.; Chen, K.; et al. A combined molecular biology and network pharmacology approach to investigate the multi-target mechanisms of Chaihu Shugan San on Alzheimer’s disease. Biomed. Pharmacother. 2019, 120, 109370. [Google Scholar] [CrossRef]
  160. Zhang, S.; Lu, Y.; Chen, W.; Shi, W.; Zhao, Q.; Zhao, J.; Li, L. Network Pharmacology and Experimental Evidence: PI3K/AKT Signaling Pathway is Involved in the Antidepressive Roles of Chaihu Shugan San. Drug. Des. Devel. Ther. 2021, 15, 3425–34441. [Google Scholar] [CrossRef]
  161. Zhou, F.; He, K.; Guan, Y.; Yang, X.; Chen, Y.; Sun, M.; Qiu, X.; Yan, F.; Huang, H.; Yao, L.; et al. Network pharmacology-based strategy to investigate pharmacological mechanisms of Tinospora sinensis for treatment of Alzheimer’s disease. J. Ethnopharmacol. 2020, 259, 112940. [Google Scholar] [CrossRef] [PubMed]
  162. Xu, X.; Saadeldeen, F.S.A.; Xu, L.; Zhao, Y.; Wei, J.; Wang, H.D.; Liu, Z.; Kang, W. The Mechanism of Phillyrin from the Leaves of Forsythia suspensa for Improving Insulin Resistance. Biomed. Res. Int. 2019, 2019, 3176483. [Google Scholar] [CrossRef] [PubMed]
  163. Singh, S.; Pathak, N.; Fatima, E.; Negi, A.S. Plant isoquinoline alkaloids: Advances in the chemistry and biology of berberine. Eur. J. Med. Chem. 2021, 226, 113839. [Google Scholar] [CrossRef] [PubMed]
  164. Cai, Z.; Chen, Y. Synergetic protective effect of berberine and ginsenoside Rb1 against tumor necrosis factor alpha-induced inflammation in adipocytes. Bioengineered 2021, 12, 11784–11796. [Google Scholar] [CrossRef]
  165. Zhao, Y.; Li, Z.; Lu, E.; Sheng, Q.; Zhao, Y. Berberine exerts neuroprotective activities against cerebral ischemia/reperfusion injury through up-regulating PPAR-γ to suppress NF-κB-mediated pyroptosis. Brain Res. Bull. 2021, 177, 22–30. [Google Scholar] [CrossRef]
  166. Xin, Z.; Fang, Y.; Du, L.; Zhu, T.; Duan, L.; Chen, J.; Gu, Q.Q.; Zhu, W.M. Aurantiomides A-C, quinazoline alkaloids from the sponge-derived fungus Penicillium aurantiogriseum SP0-19. J. Nat. Prod. 2007, 70, 853–855. [Google Scholar] [CrossRef]
  167. Dong, H.; Wu, M.; Wang, Y.; Du, W.; He, Y.; Shi, Z. Total Syntheses and Anti-inflammatory Activities of Syringin and Its Natural Analogues. J. Nat. Prod. 2021, 84, 2866–2874. [Google Scholar] [CrossRef] [PubMed]
  168. Kong, P.; Zhang, L.; Guo, Y.; Lu, Y.; Lin, D. Phillyrin, a natural lignan, attenuates tumor necrosis factor α-mediated insulin resistance and lipolytic acceleration in 3T3-L1 adipocytes. Planta Med. 2014, 80, 880–886. [Google Scholar] [CrossRef]
  169. Fang, Z.; Wei, L.; Lv, Y.; Wang, T.; Hamezah, H.S.; Han, R.; Tong, X. Phillyrin restores metabolic disorders in mice fed with high-fat diet through inhibition of interleukin-6-mediated basal lipolysis. Front Nutr. 2022, 9, 956218. [Google Scholar] [CrossRef]
  170. Su, H.-x.; Yao, S.; Zhao, W.-f.; Li, M.-j.; Liu, J.; Shang, W.-j.; Xie, H.; Ke, C.-q.; Hu, H.-c.; Gao, M.-n.; et al. Anti-SARS-CoV-2 activities in vitro of Shuanghuanglian preparations and bioactive ingredients. Acta Pharmacol. Sin. 2020, 41, 1167–1177. [Google Scholar] [CrossRef]
  171. Wang, H.; Hui, K.M.; Xu, S.; Chen, Y.; Wong, J.T.; Xue, H. Two flavones from Scutellaria baicalensis Georgi and their binding affinities to the benzodiazepine site of the GABAA receptor complex. Pharmazie 2002, 57, 857–858. [Google Scholar] [CrossRef] [PubMed]
  172. Hui, K.; Wang, X.H.; Xue, H. Interaction of flavones from the roots of Scutellaria baicalensis with the benzodiazepine site. Planta Med. 2000, 66, 91–93. [Google Scholar] [CrossRef] [PubMed]
  173. Xu, Z.; Wang, F.; Tsang, S.Y.; Ho, K.H.; Zheng, H.; Yuen, C.T.; Chow, C.Y.; Xue, H. Anxiolytic-Like Effect of baicalin and its additivity with other anxiolytics. Planta Med. 2006, 72, 189–192. [Google Scholar] [CrossRef] [PubMed]
  174. Liao, J.; Hung, W.Y.; Chen, C.F. Anxiolytic-like effects of baicalein and baicalin in the Vogel conflict test in mice. Eur. J. Pharmacol. 2003, 464, 141–146. [Google Scholar] [CrossRef]
  175. Awad, R.; Arnason, J.T.; Trudeau, V.; Bergeron, C.; Budzinski, J.W.; Foster, B.C.; Merali, Z. Phytochemical and biological analysis of skullcap (Scutellaria lateriflora L.): A medicinal plant with anxiolytic properties. Phytomedicine 2003, 10, 640–649. [Google Scholar] [CrossRef]
  176. de Carvalho, R.; Duarte, F.S.; de Lima, T.C. Involvement of GABAergic non-benzodiazepine sites in the anxiolytic-like and sedative effects of the flavonoid baicalein in mice. Behav. Brain Res. 2011, 221, 75–82. [Google Scholar] [CrossRef]
  177. Park, H.; Yoon, S.Y.; Choi, J.Y.; Lee, G.S.; Choi, J.H.; Shin, C.Y.; Son, K.H.; Lee, Y.S.; Kim, W.K.; Ryu, J.H.; et al. Anticonvulsant effect of wogonin isolated from Scutellaria baicalensis. Eur. J. Pharmacol. 2007, 574, 112–119. [Google Scholar] [CrossRef]
  178. Hu, Z.; Guan, Y.; Hu, W.; Xu, Z.; Ishfaq, M. An overview of pharmacological activities of baicalin and its aglycone baicalein: New insights into molecular mechanisms and signaling pathways. Iran. J. Basic Med. Sci. 2022, 25, 14–26. [Google Scholar]
  179. Tarragó, T.; Kichik, N.; Claasen, B.; Prades, R.; Teixidó, M.; Giralt, E. Baicalin, a prodrug able to reach the CNS, is a prolyl oligopeptidase inhibitor. Bioorganic Med. Chem. 2008, 16, 7516–7524. [Google Scholar] [CrossRef]
  180. Svarcbahs, R.; Julku, U.; Kilpeläinen, T.; Kyyrö, M.; Jäntti, M.; Myöhänen, T.T. New tricks of prolyl oligopeptidase inhibitors—A common drug therapy for several neurodegenerative diseases. Biochem. Pharmacol. 2019, 161, 113–120. [Google Scholar] [CrossRef]
  181. Szeltner, Z.; Polgár, L. Structure, function and biological relevance of prolyl oligopeptidase. Curr. Protein. Pept. Sci. 2008, 9, 96–107. [Google Scholar] [PubMed]
  182. Dethe, S.; Deepak, M.; Agarwal, A. Elucidation of Molecular Mechanism(s) of Cognition Enhancing Activity of Bacomind®: A Standardized Extract of Bacopa Monnieri. Pharmacogn. Mag. 2016, 12 (Suppl. S4), S482–S487. [Google Scholar] [CrossRef] [PubMed]
  183. Cui, H.; Kilpeläinen, T.; Zouzoula, L.; Auno, S.; Trontti, K.; Kurvonen, S.; Norrbacka, S.; Hovatta, I.; Jensen, P.H.; Myöhänen, T.T. Prolyl oligopeptidase inhibition reduces alpha-synuclein aggregation in a cellular model of multiple system atrophy. J. Cell. Mol. Med. 2021, 25, 9634–9646. [Google Scholar] [CrossRef] [PubMed]
  184. Rostami, J.; Jäntti, M.; Cui, H.; Rinne, M.K.; Kukkonen, J.P.; Falk, A.; Erlandsson, A.; Myöhänen, T. Prolyl oligopeptidase inhibition by KYP-2407 increases alpha-synuclein fibril degradation in neuron-like cells. Biomed. Pharmacother. 2020, 131, 110788. [Google Scholar] [CrossRef]
  185. Kumar, R.; Bavi, R.; Jo, M.G.; Arulalapperumal, V.; Baek, A.; Rampogu, S.; Kim, M.O.; Lee, K.W. New compounds identified through in silico approaches reduce the α-synuclein expression by inhibiting prolyl oligopeptidase in vitro. Sci. Rep. 2017, 7, 10827. [Google Scholar] [CrossRef]
  186. Höfling, C.; Kulesskaya, N.; Jaako, K.; Peltonen, I.; Männistö, P.T.; Nurmi, A.; Vartiainen, N.; Morawski, M.; Zharkovsky, A.; Võikar, V.; et al. Deficiency of prolyl oligopeptidase in mice disturbs synaptic plasticity and reduces anxiety-like behaviour, body weight, and brain volume. Eur. Neuropsychopharmacol. 2016, 26, 1048–1061. [Google Scholar] [CrossRef]
  187. Yang, R.; Liu, H.; Bai, C.; Wang, Y.; Zhang, X.; Guo, R.; Wu, S.; Wang, J.; Leung, E.; Chang, H.; et al. Chemical composition and pharmacological mechanism of Qingfei Paidu Decoction and Ma Xing Shi Gan Decoction against Coronavirus Disease 2019 (COVID-19): In silico and experimental study. Pharmacol. Res. 2020, 157, 104820. [Google Scholar] [CrossRef]
  188. Sharifi-Rad, J.; Quispe, C.; Castillo, C.M.S.; Caroca, R.; Lazo-Vélez, M.A.; Antonyak, H.; Polishchuk, A.; Lysiuk, R.; Oliinyk, P.; De Masi, L.; et al. Ellagic Acid: A Review on Its Natural Sources, Chemical Stability, and Therapeutic Potential. Oxid. Med. Cell. Longev. 2022, 2022, 3848084. [Google Scholar] [CrossRef]
  189. Yamada, H.; Wakamori SHirokane, T.; Ikeuchi, K.; Matsumoto, S. Structural revisions in natural ellagitannins. Molecules 2018, 23, 1901. [Google Scholar] [CrossRef]
  190. Gong, Q.; Cai, L.; Jing, Y.; Wang, W.; Yang, D.X.; Chen, S.W.; Tian, H.L. Urolithin A alleviates blood-brain barrier disruption and attenuates neuronal apoptosis following traumatic brain injury in mice. Neural Regen. Res. 2022, 17, 2007–2013. [Google Scholar]
  191. Larrosa, M.; González-Sarrías, A.; Yanez-Gascon, M.J.; Selma, M.V.; Azorin-Ortuno, M.; Toti, S.; Tomas-Barberan, F.; Dolara, P.; Espin, J.C. Anti-inflammatory properties of a pomegranate extract and its metabolite urolithin-A in a colitis rat model and the effect of colon inflammation on phenolic metabolism. J. Nutr. Biochem. 2010, 21, 717–725. [Google Scholar] [CrossRef]
  192. Lipińska, L.; Klewicka, E.; Sójka, M. The structure, occurrence and biological activity of ellagitannins: A general review. Acta Sci. Pol. Technol. Aliment. 2014, 13, 289–299. [Google Scholar] [CrossRef] [PubMed]
  193. Alfei, S.; Marengo, B.; Zuccari, G. Oxidative Stress, Antioxidant Capabilities, and Bioavailability: Ellagic Acid or Urolithins? Antioxidants 2020, 9, 707. [Google Scholar] [CrossRef] [PubMed]
  194. Bialonska, D.; Kasimsetty, S.G.; Khan, S.I.; Ferreira, D. Urolithins, intestinal microbial metabolites of Pomegranate ellagitannins, exhibit potent antioxidant activity in a cell-based assay. J. Agric. Food. Chem. 2009, 57, 10181–10186. [Google Scholar] [CrossRef]
  195. Cerdá, B.; Tomás-Barberán, F.A.; Espín, J.C. Metabolism of antioxidant and chemopreventive ellagitannins from strawberries, raspberries, walnuts, and oak-aged wine in humans: Identification of biomarkers and individual variability. J. Agr. Food. Chem. 2005, 53, 227–235. [Google Scholar] [CrossRef] [PubMed]
  196. Djedjibegovic, J.; Marjanovic, A.; Panieri, E.; Saso, L. Ellagic Acid-Derived Urolithins as Modulators of Oxidative Stress. Oxid. Med. Cell. Longev. 2020, 2020, 5194508. [Google Scholar] [CrossRef]
  197. Giménez-Bastida, J.; González-Sarrías, A.; Larrosa, M.; Tomás-Barberán, F.; Espín, J.C.; García-Conesa, M.T. Ellagitannin metabolites, urolithin A glucuronide and its aglycone urolithin A, ameliorate TNF-α-induced inflammation and associated molecular markers in human aortic endothelial cells. Mol. Nutr. Food Res. 2012, 56, 784–796. [Google Scholar] [CrossRef] [PubMed]
  198. Toney, A.; Albusharif, M.; Works, D.; Polenz, L.; Schlange, S.; Chaidez, V.; Ramer-Tait, A.E.; Chung, S. Differential Effects of Whole Red Raspberry Polyphenols and Their Gut Metabolite Urolithin A on Neuroinflammation in BV-2 Microglia. Int. J. Environ. Res. Public Health 2020, 18, 68. [Google Scholar] [CrossRef]
  199. Kanjanasirirat, P.; Suksatu, A.; Manopwisedjaroen, S.; Munyoo, B.; Tuchinda, P.; Jearawuttanakul, K.; Seemakhan, S.; Charoensutthivarakul, S.; Wongtrakoongate, P.; Rangkasenee, N.; et al. High-content screening of Thai medicinal plants reveals Boesenbergia rotunda extract and its component Panduratin A as anti-SARS-CoV-2 agents. Sci. Rep. 2020, 10, 19963. [Google Scholar] [CrossRef]
  200. Jayatunga, D.; Hone, E.; Khaira, H.; Lunelli, T.; Singh, H.; Guillemin, G.J.; Fernando, B.; Garg, M.L.; Verdile, G.; Martins, R.N. Therapeutic Potential of Mitophagy-Inducing Microflora Metabolite, Urolithin A for Alzheimer’s Disease. Nutrients 2021, 13, 3744. [Google Scholar] [CrossRef]
  201. Li, H.; Zhang, S.Y.; Ren, Y.S.; Zhou, J.C.; Zhou, Y.X.; Huang, W.Z.; Piao, X.H.; Yang, Z.Y.; Wang, S.M.; Ge, Y.W. Identification of ellagic acid and urolithins as natural inhibitors of Aβ25-35-induced neurotoxicity and the mechanism predication using network pharmacology analysis and molecular docking. Front. Nutr. 2022, 9, 966276. [Google Scholar] [CrossRef]
  202. Liu, J.; Jiang, J.; Qiu, J.; Wang, L.; Zhuo, J.; Wang, B.; Sun, D.; Yu, S.; Lou, H. Urolithin A protects dopaminergic neurons in experimental models of Parkinson’s disease by promoting mitochondrial biogenesis through the SIRT1/PGC-1α signaling pathway. Food Funct. 2022, 13, 375–385. [Google Scholar] [CrossRef]
  203. Qiu, J.; Chen, Y.; Zhuo, J.; Zhang, L.; Liu, J.; Wang, B.; Sun, D.; Yu, S.; Lou, H. Urolithin A promotes mitophagy and suppresses NLRP3 inflammasome activation in lipopolysaccharide-induced BV2 microglial cells and MPTP-induced Parkinson’s disease model. Neuropharmacology 2022, 207, 108963. [Google Scholar] [CrossRef] [PubMed]
  204. DaSilva, N.; Nahar, P.P.; Ma, H.; Eid, A.; Wei, Z.; Meschwitz, S.; Zawia, N.H.; Slitt, A.L.; Seeram, N.P. Pomegranate ellagitannin-gut microbial-derived metabolites, urolithins, inhibit neuroinflammation in vitro. Nutr. Neurosci. 2019, 22, 185–195. [Google Scholar] [CrossRef] [PubMed]
  205. Kujawska, M.; Jourdes, M.; Kurpik, M.; Szulc, M.; Szaefer, H.; Chmielarz, P.; Kreiner, G.; Krajka-Kuźniak, V.; Mikołajczak, P.Ł.; Teissedre, P.L.; et al. Neuroprotective Effects of Pomegranate Juice against Parkinson’s Disease and Presence of Ellagitannins-Derived Metabolite-Urolithin A-In the Brain. Int. J. Mol. Sci. 2019, 21, 202. [Google Scholar] [CrossRef] [PubMed]
  206. Xu, J.; Yuan, C.; Wang, G.; Luo, J.; Ma, H.; Xu, L.; Mu, Y.; Li, Y.; Seeram, N.P.; Huang, X.; et al. Urolithins Attenuate LPS-Induced Neuroinflammation in BV2Microglia via MAPK, Akt, and NF-κB Signaling Pathways. J. Agric. Food Chem. 2018, 66, 571–580. [Google Scholar] [CrossRef]
  207. Bobowska, A.; Granica, S.; Filipek, A.; Melzig, M.F.; Moeslinger, T.; Zentek, J.; Kruk, A.; Piwowarski, J.P. Comparative studies of urolithins and their phase II metabolites on macrophage and neutrophil functions. Eur. J. Nutr. 2021, 60, 1957–1972. [Google Scholar] [CrossRef]
  208. Singh, R.; Chandrashekharappa, S.; Vemula, P.K.; Haribabu, B.; Jala, V.R. Microbial Metabolite Urolithin B Inhibits Recombinant Human Monoamine Oxidase A Enzyme. Metabolites 2020, 10, 258. [Google Scholar] [CrossRef]
  209. Masella, R.; Santangelo, C.; D’Archivio, M.; Li Volti, G.; Giovannini, C.; Galvano, F. Protocatechuic acid and human disease prevention: Biological activities and molecular mechanisms. Curr. Med. Chem. 2012, 19, 2901–2917. [Google Scholar] [CrossRef]
  210. Li, J.; Deng, Y.; Yuan, C.; Pan, L.; Chai, H.; Keller, W.J.; Kinghorn, A.D. Antioxidant and quinone reductase-inducing constituents of black chokeberry (Aronia melanocarpa) fruits. J. Agric. Food Chem. 2012, 60, 11551–11559. [Google Scholar] [CrossRef]
  211. Gureev, A.; Popov, V.N.; Starkov, A.A. Crosstalk between the mTOR and Nrf2/ARE signaling pathways as a target in the improvement of long-term potentiation. Exp. Neurol. 2020, 328, 113285. [Google Scholar] [CrossRef]
  212. Zhu, Y.; Yang, Q.; Liu, H.; Song, Z.; Chen, W. Phytochemical compounds targeting on Nrf2 for chemoprevention in colorectal cancer. Eur. J. Pharmacol. 2020, 887, 173588. [Google Scholar] [CrossRef] [PubMed]
  213. Ahmed, S.; Luo, L.; Namani, A.; Wang, X.J.; Tang, X. Nrf2 signaling pathway: Pivotal roles in inflammation. Biochim. Biophys. Acta Mol. Basis Dis. 2017, 1863, 585–597. [Google Scholar] [CrossRef] [PubMed]
  214. Pennisi, M.; Crupi, R.; Di Paola, R.; Ontario, M.L.; Bella, R.; Calabrese, E.J.; Crea, R.; Cuzzocrea, S.; Calabrese, V. Inflammasomes, hormesis, and antioxidants in neuroinflammation: Role of NRLP3 in Alzheimer disease. J. Neurosci. Res. 2017, 95, 1360–1372. [Google Scholar] [CrossRef]
  215. Andreux, P.; Blanco-Bose, W.; Ryu, D.; Burdet, F.; Ibberson, M.; Aebischer, P.; Auwerx, J.; Singh, A.; Rinsch, C. The mitophagy activator urolithin A is safe and induces a molecular signature of improved mitochondrial and cellular health in humans. Nat. Metab. 2019, 1, 595–603. [Google Scholar] [CrossRef] [PubMed]
  216. Liu, S.; D’Amico, D.; Shankland, E.; Bhayana, S.; Garcia, J.M.; Aebischer, P.; Rinsch, C.; Singh, A.; Marcinek, D.J. Effect of Urolithin A Supplementation on Muscle Endurance and and Mitochondrial Health in Older Adults: A Randomized Clinical Trial. JAMA Netw. Open. 2022, 5, e2144279. [Google Scholar] [CrossRef] [PubMed]
  217. Singh, A.; D’Amico, D.; Andreux, P.A.; Fouassier, A.M.; Blanco-Bose, W.; Evans, M.; Aebischer, P.; Auwerx, J.; Rinsch, C. Urolithin A improves muscle strength, exercise performance, and biomarkers of mitochondrial health in a randomized trial in middle-aged adults. Cell. Rep. Med. 2022, 3, 100633. [Google Scholar] [CrossRef]
  218. Ciumărnean, L.; Milaciu, M.V.; Runcan, O.; Vesa, Ș.C.; Răchișan, A.L.; Negrean, V.; Perné, M.G.; Donca, V.I.; Alexescu, T.G.; Para, I.; et al. The Effects of Flavonoids in Cardiovascular Diseases. Molecules 2020, 25, 4320. [Google Scholar] [CrossRef]
  219. Aguilar Peralta, G.; Arévalo Gardoqui, J.; Llamas Macías, F.J.; Navarro Ceja, V.H.; Mendoza Cisneros, S.A.; Martínez Macías, C.G. Clinical and capillaroscopic evaluation in the treatment of chronic venous insufficiency with Ruscus aculeatus, hesperidin methylchalcone and ascorbic acid in venous insufficiency treatment of ambulatory patients. Int. Angiol. 2007, 26, 378–384. [Google Scholar]
  220. Manach, C.; Scalbert, A.; Morand, C.; Rémésy, C.; Jiménez, L. Polyphenols: Food sources and bioavailability. Am. J. Clin. Nutr. 2004, 79, 727–747. [Google Scholar] [CrossRef]
  221. Scalbert, A.; Williamson, G. Dietary intake and bioavailability of polyphenols. J. Nutr. 2000, 130, 2073S–2085S. [Google Scholar] [CrossRef] [PubMed]
  222. Najmanová, I.; Pourová, J.; Vopršalová, M.; Pilařová, V.; Semecký, V.; Nováková, L.; Mladěnka, P. Flavonoid metabolite 3-(3-hydroxyphenyl)propionic acid formed by human microflora decreases arterial blood pressure in rats. Mol. Nutr. Food Res. 2016, 60, 981–991. [Google Scholar] [CrossRef] [PubMed]
  223. Pourová, J.; Najmanová, I.; Vopršalová, M.; Migkos, T.; Pilařová, V.; Applová, L.; Nováková, L.; Mladěnka, P. Two flavonoid metabolites, 3,4-dihydroxyphenylacetic acid and 4-methylcatechol, relax arteries ex vivo and decrease blood pressure in vivo. Vascul. Pharmacol. 2018, 111, 36–43. [Google Scholar] [CrossRef] [PubMed]
  224. Bhardwaj, P.; Khanna, D. Green tea catechins: Defensive role in cardiovascular disorders. Chin. J. Nat. Med. 2013, 11, 345–353. [Google Scholar] [CrossRef]
  225. Corral-Jara, K.; 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 TNFalpha-Stimulated Primary Human Brain Microvascular Endothelial Cells. Front. Neurosci. 2021, 15, 622640. [Google Scholar] [CrossRef]
  226. Corral-Jara, K.; Nuthikattu, S.; Rutledge, J.; Villablanca, A.; Fong, R.; Heiss, C.; Ottaviani, J.I.; Milenkovic, D. Structurally related (-)-epicatechin metabolites and gut microbiota derived metabolites exert genomic modifications via VEGF signaling pathways in brain microvascular endothelial cells under lipotoxic conditions: Integrated multi-omic study. J. Proteomics 2022, 263, 104603. [Google Scholar] [CrossRef]
  227. Lotito, S.; Zhang, W.J.; Yang, C.S.; Crozier, A.; Frei, B. Metabolic conversion of dietary flavonoids alters their anti-inflammatory and antioxidant properties. Free. Radic. Biol. Med. 2011, 51, 454–463. [Google Scholar] [CrossRef]
  228. Parvez, S.; Long, M.J.; Poganik, J.R.; Aye, Y. Redox signaling by reactive electrophiles and oxidants. Chem. Rev. 2018, 118, 8798–8888. [Google Scholar] [CrossRef]
  229. Na, H.-K.; Surh, Y.-J. Modulation of Nrf2-mediated antioxidant and detoxifying enzyme induction by the green tea polyphenol EGCG. Food Chem. Toxicol. 2008, 46, 1271–1278. [Google Scholar] [CrossRef]
  230. Liu, C.; Boeren, S.; Rietjens, I.M.C.M. Intra- and Inter-individual Differences in the Human Intestinal Microbial Conversion of (-)-Epicatechin and Bioactivity of Its Major Colonic Metabolite 5-(3’,4’-Dihydroxy-Phenyl)-γ-Valerolactone in Regulating Nrf2-Mediated Gene Expression. Front. Nutr. 2022, 9, 910785. [Google Scholar] [CrossRef]
  231. De Rosso, V.; Morán Vieyra, F.E.; Mercadante, A.Z.; Borsarelli, C.D. Singlet oxygen quenching by anthocyanin’s flavylium cations. Free. Radic. Res. 2008, 42, 885–891. [Google Scholar] [CrossRef] [PubMed]
  232. Kamiloglu, S.; Capanoglu, E.; Grootaert, C.; Van Camp, J. Anthocyanin Absorption and Metabolism by Human Intestinal Caco-2 Cells--A Review. Int. J. Mol. Sci. 2015, 16, 1555–1574. [Google Scholar] [CrossRef]
  233. Minocha, T.; Birla, H.; Obaid, A.A.; Rai, V.; Sushma, P.; Shivamallu, C.; Moustafa, M.; Al-Shehri, M.; Al-Emam, A.; Tikhonova, M.A.; et al. Flavonoids as Promising Neuroprotectants and Their Therapeutic Potential against Alzheimer’s Disease. Oxid. Med. Cell. Longev. 2022, 2022, 6038996. [Google Scholar] [CrossRef]
  234. Melrose, J.; Smith, M.M. Natural and Semi-Synthetic Flavonoid Anti-SARS-CoV-2 Agents for the Treatment of Long COVID-19 Disease and Neurodegenerative Disorders of Cognitive Decline. Front. Biosci. Elite 2022, 14, 27. [Google Scholar] [CrossRef] [PubMed]
  235. Speisky, H.; Shahidi, F.; Costa de Camargo, A.; Fuentes, J. Revisiting the Oxidation of Flavonoids: Loss, Conservation or Enhancement of Their Antioxidant Properties. Antioxidants 2022, 11, 133. [Google Scholar] [CrossRef]
  236. Kakkar, S.; Bais, S. A review on protocatechuic Acid and its pharmacological potential. ISRN Pharmacol. 2014, 2014, 952943. [Google Scholar] [CrossRef]
  237. Krzysztoforska, K.; Mirowska-Guzel, D.; Widy-Tyszkiewicz, E. Pharmacological effects of protocatechuic acid and its therapeutic potential in neurodegenerative diseases: Review on the basis of in vitro and in vivo studies in rodents and humans. Nutr. Neurosci. 2019, 22, 72–82. [Google Scholar] [CrossRef]
  238. He, X.; Zhou, Y.Z.; Sheng, S.; Li, J.J.; Wang, G.Q.; Zhang, F. Ellagic Acid Protects Dopamine Neurons via Inhibition of NLRP3 Inflammasome Activation in Microglia. Oxid. Med. Cell. Longev. 2020, 2020, 2963540. [Google Scholar] [CrossRef] [PubMed]
  239. Nones, J.; Spohr, T.C.; Gomes, F.C. Hesperidin, a flavone glycoside, as mediator of neuronal survival. Neurochem. Res. 2011, 36, 1776–1784. [Google Scholar] [CrossRef] [PubMed]
  240. Chakraborty, S.; Bandyopadhyay, J.; Chakraborty, S.; Basu, S. Multi-target screening mines hesperidin as a multi-potent inhibitor: Implication in Alzheimer’s disease therapeutics. Eur. J. Med. Chem. 2016, 121, 810–822. [Google Scholar] [CrossRef]
  241. Chakraborty, S.; Rakshit, J.; Bandyopadhyay, J.; Basu, S. Multi-target inhibition ability of neohesperidin dictates its neuroprotective activity: Implication in Alzheimer’s disease therapeutics. Int. J. Biol. Macromol. 2021, 176, 315–324. [Google Scholar] [CrossRef] [PubMed]
  242. Wang, D.; Liu, L.; Zhu, X.; Wu, W.; Wang, Y. Hesperidin alleviates cognitive impairment, mitochondrial dysfunction and oxidative stress in a mouse model of Alzheimer’s disease. Cell. Mol. Neurobiol. 2014, 34, 1209–1221. [Google Scholar] [CrossRef] [PubMed]
  243. Hajialyani, M.; Hosein Farzaei, M.; Echeverría, J.; Nabavi, S.M.; Uriarte, E.; Sobarzo-Sánchez, E. Hesperidin as a Neuroprotective Agent: A Review of Animal and Clinical Evidence. Molecules 2019, 24, 648. [Google Scholar] [CrossRef] [PubMed]
  244. Wang, J.; Yuan, Y.; Zhang, P.; Zhang, H.; Liu, X.; Zhang, Y. Neohesperidin Prevents Aβ25-35-Induced Apoptosis in Primary Cultured Hippocampal Neurons by Blocking the S-Nitrosylation of Protein-Disulphide Isomerase. Neurochem. Res. 2018, 43, 1736–1744. [Google Scholar] [CrossRef] [PubMed]
  245. Hong, Y.; An, Z. Hesperidin attenuates learning and memory deficits in APP/PS1 mice through activation of Akt/Nrf2 signaling and inhibition of RAGE/NF-κB signaling. Arch. Pharm. Res. 2018, 41, 655–663. [Google Scholar] [CrossRef]
  246. Mathew, B.; Suresh, J.; Mathew, G.E.; Parasuraman, R.; Abdulla, N. Plant secondary metabolites- potent inhibitors of monoamine oxidase isoforms. Cent. Nerv. Syst. Agents Med. Chem. 2014, 14, 28–33. [Google Scholar] [CrossRef]
  247. Gidaro, M.; Astorino, C.; Petzer, A.; Carradori, S.; Alcaro, F.; Costa, G.; Artese, A.; Rafele, G.; Russo, F.M.; Petzer, J.P.; et al. Kaempferol as Selective Human MAO-A Inhibitor: Analytical Detection in Calabrian Red Wines, Biological and Molecular Modeling Studies. J. Agric. Food Chem. 2016, 64, 1394–1400. [Google Scholar] [CrossRef]
  248. Park, S.; Paudel, P.; Wagle, A.; Seong, S.H.; Kim, H.R.; Fauzi, F.M.; Jung, H.A.; Choi, J.S. Luteolin, a Potent Human Monoamine Oxidase-A Inhibitor and Dopamine D4 and Vasopressin V1A Receptor Antagonist. J. Agric. Food Chem. 2020, 68, 10719–10729. [Google Scholar] [CrossRef]
  249. Andersen, O.M.B.N.; Landau, A.M.; Pløen, G.G.; Jensen, A.M.G.; Monti, G.; Ulhøi, B.P.; Nyengaard, J.R.; Jacobsen, K.R.; Jørgensen, M.M.; Holm, I.E.; et al. A genetically modified minipig model for Alzheimer’s disease with SORL1 haploinsufficiency. Cell. Rep. Med. 2022, 3, 100740. [Google Scholar] [CrossRef]
  250. Dawson, T.; Golde, T.E.; Lagier-Tourenne, C. Animal models of neurodegenerative diseases. Nat. Neurosci. 2018, 21, 1370–1379. [Google Scholar] [CrossRef]
  251. Hassan, H.; Elnagar, M.R.; Abdelrazik, E.; Mahdi, M.R.; Hamza, E.; Elattar, E.M.; ElNashar, E.M.; Alghamdi, M.A.; Al-Qahtani, Z.; Al-Khater, K.M.; et al. Neuroprotective effect of naringin against cerebellar changes in Alzheimer’s disease through modulation of autophagy, oxidative stress and tau expression: An experimental study. Front. Neuroanat. 2022, 16, 1012422. [Google Scholar] [CrossRef]
  252. Holm, I.E.; Alstrup, A.K.O.; Luo, Y. Genetically modified pig models for neurodegenerative disorders. J. Pathol. 2016, 238, 267–287. [Google Scholar] [CrossRef]
  253. Rockenstein, E.C.L.; Masliah, E. Transgenic animal models of neurodegenerative diseases and their application to treatment development. Adv. Drug. Deliv. Rev. 2007, 59, 1093–1102. [Google Scholar] [CrossRef] [PubMed]
  254. Yin, P.; Li, S.; Li, X.J.; Yang, W. New pathogenic insights from large animal models of neurodegenerative diseases. Protein Cell. 2022, 13, 707–720. [Google Scholar] [CrossRef] [PubMed]
  255. Yang, W.; Chen, X.; Li, S.; Li, X.J. Genetically modified large animal models for investigating neurodegenerative diseases. Cell. Biosci. 2021, 11, 218. [Google Scholar] [CrossRef] [PubMed]
  256. Lee, J.; Lee, Y.K.; Ban, J.O.; Ha, T.Y.; Yun, Y.P.; Han, S.B.; Oh, K.W.; Hong, J.T. Green tea (-)-epigallocatechin-3-gallate inhibits beta-amyloid-induced cognitive dysfunction through modification of secretase activity via inhibition of ERK and NF-kappaB pathways in mice. J. Nutr. 2009, 139, 1987–1993. [Google Scholar] [CrossRef]
  257. Rezai-Zadeh, K.; Arendash, G.W.; Hou, H.; Fernandez, F.; Jensen, M.; Runfeldt, M.; Shytle, R.D.; Tan, J. Green tea epigallocatechin-3-gallate (EGCG) reduces beta-amyloid mediated cognitive impairment and modulates tau pathology in Alzheimer transgenic mice. Brain Res. 2008, 1214, 177–187. [Google Scholar] [CrossRef]
  258. Rendeiro, C.; Spencer, J.P.; Vauzour, D.; Butler, L.T.; Ellis, J.A.; Williams, C.M. The impact of flavonoids on spatial memory in rodents: From behaviour to underlying hippocampal mechanisms. Genes. Nutr. 2009, 4, 251–270. [Google Scholar] [CrossRef]
  259. Vauzour, D. Effect of flavonoids on learning, memory and neurocognitive performance: Relevance and potential implications for Alzheimer’s disease pathophysiology. J. Sci. Food Agric. 2014, 94, 1042–1056. [Google Scholar] [CrossRef]
  260. Zhang, L.; Yao, J.Z.; Li, Y.; Li, K.; Chen, H.X.; Zhang, Y.Z.; Li, Y.F. Anxiolytic effects of flavonoids in animal models of posttraumatic stress disorder. Evid. Based Complement. Altern. Med. 2012, 2012, 623753. [Google Scholar] [CrossRef]
  261. Wang, Q.; Dong, X.; Zhang, R.; Zhao, C. Flavonoids with Potential Anti-Amyloidogenic Effects as Therapeutic Drugs for Treating Alzheimer’s Disease. J. Alzheimers Dis. 2021, 84, 505–533. [Google Scholar] [CrossRef]
  262. Gandhi, G.; Neta, M.T.S.L.; Sathiyabama, R.G.; Quintans, J.S.S.; de Oliveira ESilva, A.M.; Araújo, A.A.S.; Narain, N.; Júnior, L.J.Q.; Gurgel, R.Q. Flavonoids as Th1/Th2 cytokines immunomodulators: A systematic review of studies on animal models. Phytomedicine 2018, 44, 74–84. [Google Scholar] [CrossRef]
  263. Guo, Y.; Zhao, Y.; Nan, Y.; Wang, X.; Chen, Y.; Wang, S. (-)-Epigallocatechin-3-gallate ameliorates memory impairment and rescues the abnormal synaptic protein levels in the frontal cortex and hippocampus in a mouse model of Alzheimer’s disease. Neuroreport 2017, 28, 590–597. [Google Scholar] [CrossRef] [PubMed]
  264. 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. Devel. Ther. 2021, 15, 2013–2024. [Google Scholar] [CrossRef]
  265. Zhang, S.; Zhu, Q.; Chen, J.Y.; OuYang, D.; Lu, J.H. The pharmacological activity of epigallocatechin-3-gallate (EGCG) on Alzheimer’s disease animal model: A systematic review. Phytomedicine 2020, 79, 153316. [Google Scholar] [CrossRef]
  266. Zhang, X.; Chen, J.Y.; Ouyang, D.; Lu, J.H. Quercetin in Animal Models of Alzheimer’s Disease: A Systematic Review of Preclinical Studies. Int. J. Mol. Sci. 2020, 21, 493. [Google Scholar] [CrossRef]
  267. Fakhri, S.; Abdian, S.; Zarneshan, S.N.; Akkol, E.K.; Farzaei, M.H.; Sobarzo-Sánchez, E. Targeting Mitochondria by Plant Secondary Metabolites: A Promising Strategy in Combating Parkinson’s Disease. Int. J. Mol. Sci. 2021, 22, 12570. [Google Scholar] [CrossRef] [PubMed]
  268. Lorzadeh, E.; Ramezani-Jolfaie, N.; Mohammadi, M.; Khoshbakht, Y.; Salehi-Abargouei, A. The effect of hesperidin supplementation on inflammatory markers in human adults: A systematic review and meta-analysis of randomized controlled clinical trials. Chem. Biol. Interact. 2019, 307, 8–15. [Google Scholar] [CrossRef] [PubMed]
  269. Davinelli, S.; Ali, S.; Scapagnini, G.; Costagliola, C. Effects of Flavonoid Supplementation on Common Eye Disorders: A Systematic Review and Meta-Analysis of Clinical Trials. Front. Nutr. 2021, 8, 651441. [Google Scholar] [CrossRef] [PubMed]
  270. Homayouni, F.; Haidari, F.; Hedayati, M.; Zakerkish, M.; Ahmadi, K. Blood pressure lowering and anti-inflammatory effects of hesperidin in type 2 diabetes; a randomized double-blind controlled clinical trial. Phytother. Res. 2018, 32, 1073–1079. [Google Scholar] [CrossRef]
  271. Giannini, I.; Amato, A.; Basso, L.; Tricomi, N.; Marranci, M.; Pecorella, G.; Tafuri, S.; Pennisi, D.; Altomare, D.F. Flavonoids mixture (diosmin, troxerutin, hesperidin) in the treatment of acute hemorrhoidal disease: A prospective, randomized, triple-blind, controlled trial. Tech. Coloproctol. 2015, 19, 339–345. [Google Scholar] [CrossRef] [PubMed]
  272. Martínez Noguera, F.; Alcaraz, P.E.; Carlos Vivas, J.; Chung, L.H.; Marín Cascales, E.; Marín Pagán, C. 8 weeks of 2S-Hesperidin supplementation improves muscle mass and reduces fat in amateur competitive cyclists: Randomized controlled trial. Food Funct. 2021, 12, 3872–3882. [Google Scholar] [CrossRef] [PubMed]
  273. van Iersel, L.; Stevens, Y.R.; Conchillo, J.M.; Troost, F.J. The effect of citrus flavonoid extract supplementation on anaerobic capacity in moderately trained athletes: A randomized controlled trial. J. Int. Soc. Sports Nutr. 2021, 18, 2. [Google Scholar] [CrossRef] [PubMed]
  274. Sheen, Y.; Huang, H.Y.; Liao, Y.H. The efficacy and safety of an antiaging topical serum containing hesperetin and sodium cyclic lysophosphatidic acid: A single-center clinical trial. J. Cosmet. Dermatol. 2021, 20, 3960–3967. [Google Scholar] [CrossRef]
  275. Joshi, S.; Dhingra, A.K.; Chopra, B.; Guarve, K.; Bhateja, D. Therapeutic Potential and Clinical Evidence of Hesperidin as Neuroprotective Agent. Cent. Nerv. Syst. Agents Med. Chem. 2022, 22, 5–14. [Google Scholar] [CrossRef]
  276. Cieuta-Walti, C.; Cuenca-Royo, A.; Langohr, K.; Rakic, C.; López-Vílchez, M.Á.; Lirio, J.; González-Lamuño Leguina, D.; González, T.B.; García, J.G.; Roure, M.R.; et al. Safety and preliminary efficacy on cognitive performance and adaptive functionality of epigallocatechin gallate (EGCG) in children with Down syndrome. A randomized phase Ib clinical trial (PERSEUS study). Genet. Med. 2022, 24, 2004–2013. [Google Scholar] [CrossRef]
  277. Zhao, H.; Zhu, W.; Zhao, X.; Li, X.; Zhou, Z.; Zheng, M.; Meng, X.; Kong, L.; Zhang, S.; He, D.; et al. Efficacy of Epigallocatechin-3-Gallate in Preventing Dermatitis in Patients With Breast Cancer Receiving Postoperative Radiotherapy: A Double-Blind, Placebo-Controlled, Phase 2 Randomized Clinical Trial. JAMA Dermatol. 2022, 158, 779–786. [Google Scholar] [CrossRef]
  278. Hu, W.; Zhang, L.; Lin, F.; Lei, J.; Zhou, M.; Xu, A. Topical epigallocatechin-3-gallate in the treatment of vitiligo. Australas. J. Dermatol. 2021, 62, e404–e407. [Google Scholar] [CrossRef]
  279. Ud-Din, S.; Wilgus, T.A.; McGeorge, D.D.; Bayat, A. Pre-Emptive Priming of Human Skin Improves Cutaneous Scarring and Is Superior to Immediate and Delayed Topical Anti-Scarring Treatment Post-Wounding: A Double-Blind Randomised Placebo-Controlled Clinical Trial. Pharmaceutics 2021, 13, 510. [Google Scholar] [CrossRef]
  280. Ud-Din, S.; Foden, P.; Mazhari, M.; Al-Habba, S.; Baguneid, M.; Bulfone-Paus, S.; McGeorge, D.; Bayat, A. A Double-Blind, Randomized Trial Shows the Role of Zonal Priming and Direct Topical Application of Epigallocatechin-3-Gallate in the Modulation of Cutaneous Scarring in Human Skin. J. Investig. Dermatol. 2019, 139, 1680–1690.e16. [Google Scholar] [CrossRef]
  281. Din, U.; Sian, T.S.; Deane, C.S.; Smith, K.; Gates, A.; Lund, J.N.; Williams, J.P.; Rueda, R.; Pereira, S.L.; Atherton, P.J.; et al. Green Tea Extract Concurrent with an Oral Nutritional Supplement Acutely Enhances Muscle Microvascular Blood Flow without Altering Leg Glucose Uptake in Healthy Older Adults. Nutrients 2021, 13, 3895. [Google Scholar] [CrossRef] [PubMed]
  282. Yoshitomi, R.; Yamamoto, M.; Kumazoe, M.; Fujimura, Y.; Yonekura, M.; Shimamoto, Y.; Nakasone, A.; Kondo, S.; Hattori, H.; Haseda, A.; et al. The combined effect of green tea and α-glucosyl hesperidin in preventing obesity: A randomized placebo-controlled clinical trial. Sci. Rep. 2021, 11, 19067. [Google Scholar] [CrossRef]
  283. Bazyar, H.; Hosseini, S.A.; Saradar, S.; Mombaini, D.; Allivand, M.; Labibzadeh, M.; Alipour, M. Effects of epigallocatechin-3-gallate of Camellia sinensis leaves on blood pressure, lipid profile, atherogenic index of plasma and some inflammatory and antioxidant markers in type 2 diabetes mellitus patients: A clinical trial. J. Complement. Integr. Med. 2020, 18, 405–411. [Google Scholar] [CrossRef] [PubMed]
  284. de Morais Junior, A.; Schincaglia, R.M.; Passarelli, M.; Pimentel, G.D.; Mota, J.F. Acute Epigallocatechin-3-Gallate Supplementation Alters Postprandial Lipids after a Fast-Food Meal in Healthy Young Women: A Randomized, Double-Blind, Placebo-Controlled Crossover Study. Nutrients 2020, 12, 2533. [Google Scholar] [CrossRef]
  285. de la Torre, R.; de Sola, S.; Farré, M.; Xicota, L.; Cuenca-Royo, A.; Rodriguez, J.; León, A.; Langohr, K.; Gomis-González, M.; Hernandez, G.; et al. A phase 1, randomized double-blind, placebo controlled trial to evaluate safety and efficacy of epigallocatechin-3-gallate and cognitive training in adults with Fragile X syndrome. Clin. Nutr. 2020, 39, 378–387. [Google Scholar] [CrossRef] [PubMed]
  286. Kohri, T.; Matsumoto, N.; Yamakawa, M.; Suzuki, M.; Nanjo, F.; Hara, Y.; Oku, N. Metabolic fate of (-)-[4-(3)H]epigallocatechin gallate in rats after oral administration. J. Agric. Food Chem. 2001, 49, 4102–4112. [Google Scholar] [CrossRef]
  287. Mena, P.; Bresciani, L.; Brindani, N.; Ludwig, I.A.; Pereira-Caro, G.; Angelino, D.; Llorach, R.; Calani, L.; Brighenti, F.; Clifford, M.N.; et al. Phenyl-gamma-valerolactones and phenylvaleric acids, the main colonic metabolites of flavan-3-ols: Synthesis, analysis, bioavailability, and bioactivity. Nat. Prod. Rep. 2019, 36, 714–752. [Google Scholar] [CrossRef]
  288. Nation, D.; Sweeney, M.D.; Montagne, A.; Sagare, A.P.; D’Orazio, L.M.; Pachicano, M.; Sepehrband, F.; Nelson, A.R.; Buennagel, D.P.; Harrington, M.G.; et al. Blood–brain barrier breakdown is an early biomarker of human cognitive dysfunction. Nat. Med. 2019, 25, 270–276. [Google Scholar] [CrossRef]
  289. Takagaki, A.; Nanjo, F. Metabolism of (-)-epigallocatechin gallate by rat intestinal flora. J. Agric. Food Chem. 2010, 58, 1313–1321. [Google Scholar] [CrossRef]
  290. van’t Slot, G.; Humpf, H.U. Degradation and metabolism of catechin, epigallocatechin-3-gallate (EGCG), and related compounds by the intestinal microbiota in the pig cecum model. J. Agric. Food Chem. 2009, 57, 8041–8048. [Google Scholar] [CrossRef]
  291. Nehlig, A. The neuroprotective effects of cocoa flavanol and its influence on cognitive performance. Br. J. Clin. Pharmacol. 2013, 75, 716–727. [Google Scholar] [CrossRef]
  292. Wang, J.; Varghese, M.; Ono, K.; Yamada, M.; Levine Tzavaras, N.; Gong, B.; Hurst, W.; Blitzer, R.; Pasinetti, G.M. Cocoa extracts reduce oligomerization of amyloid-beta: Implications for cognitive improvement in Alzheimer’s disease. J. Alzheimers. Dis. 2014, 41, 643–650. [Google Scholar] [CrossRef] [PubMed]
  293. Sokolov, A.; Pavlova, M.A.; Klosterhalfen, S.; Enck, P. Chocolate and the brain: Neurobiological impact of cocoa flavanols on cognition and behavior. Neurosci. Biobehav. Rev. 2013, 37, 2445–2453. [Google Scholar] [CrossRef] [PubMed]
  294. Aarsland, D.; Khalifa, K.; Bergland, A.K.; Soennesyn, H.; Oppedal, K.; Holteng, L.B.A.; Oesterhus, R.; Nakling, A.; Jarholm, J.A.; de Lucia, C.; et al. A Randomised Placebo-Controlled Study of Purified Anthocyanins on Cognition in Individuals at Increased Risk for Dementia. Am. J. Geriatr. Psychiatry 2023, 31, 141–151. [Google Scholar] [CrossRef] [PubMed]
  295. Khalifa, K.; Bergland, A.K.; Soennesyn, H.; Oppedal, K.; Oesterhus, R.; Dalen, I.; Larsen, A.I.; Fladby, T.; Brooker, H.; Wesnes, K.A.; et al. Effects of Purified Anthocyanins in People at Risk for Dementia: Study Protocol for a Phase II Randomized Controlled Trial. Front. Neurol. 2020, 11, 916. [Google Scholar] [CrossRef]
  296. Arisi, T.; Gorski, F.; Eibel, B.; Barbosa, E.; Boll, L.; Waclawovsky, G.; Lehnen, A.M. Dietary intake of anthocyanins improves arterial stiffness, but not endothelial function, in volunteers with excess weight: A randomized clinical trial. Phytother. Res. 2022, 37, 798–808. [Google Scholar] [CrossRef]
  297. Kent, K.; Yousefi, M.; do Rosario, V.A.; Fitzgerald, Z.; Broyd, S.; Visentin, D.; Roodenrys, S.; Walton, K.; Charlton, K.E. Anthocyanin intake is associated with improved memory in older adults with mild cognitive impairment. Nutr. Res. 2022, 104, 36–43. [Google Scholar] [CrossRef]
  298. Krikorian, R.; Skelton, M.R.; Summer, S.S.; Shidler, M.D.; Sullivan, P.G. Blueberry Supplementation in Midlife for Dementia Risk Reduction. Nutrients 2022, 14, 1619. [Google Scholar] [CrossRef]
  299. Cremonini, E.; Daveri, E.; Iglesias, D.E.; Kang, J.; Wang, Z.; Gray, R.; Mastaloudis, A.; Kay, C.D.; Hester, S.N.; Wood, S.M.; et al. A randomized placebo-controlled cross-over study on the effects of anthocyanins on inflammatory and metabolic responses to a high-fat meal in healthy subjects. Redox Biol. 2022, 51, 102273. [Google Scholar] [CrossRef]
  300. Emamat, H.; Zahedmehr, A.; Asadian, S.; Nasrollahzadeh, J. The effect of barberry (Berberis integerrima) on lipid profile and systemic inflammation in subjects with cardiovascular risk factors: A randomized controlled trial. BMC Complement. Med. Ther. 2022, 22, 59. [Google Scholar] [CrossRef]
  301. Tian, Z.; Li, K.; Fan, D.; Zhao, Y.; Gao, X.; Ma, X.; Xu, L.; Shi, Y.; Ya, F.; Zou, J.; et al. Dose-dependent effects of anthocyanin supplementation on platelet function in subjects with dyslipidemia: A randomized clinical trial. EBioMedicine 2021, 70, 103533. [Google Scholar] [CrossRef] [PubMed]
  302. Ahles, S.; Joris, P.J.; Plat, J. Effects of Berry Anthocyanins on Cognitive Performance, Vascular Function and Cardiometabolic Risk Markers: A Systematic Review of Randomized Placebo-Controlled Intervention Studies in Humans. Int. J. Mol. Sci. 2021, 22, 6482. [Google Scholar] [CrossRef] [PubMed]
  303. Rosli, H.; Shahar, S.; Rajab, N.F.; Che Din, N.; Haron, H. The effects of polyphenols-rich tropical fruit juice on cognitive function and metabolomics profile—A randomized controlled trial in middle-aged women. Nutr. Neurosci. 2022, 25, 1577–1593. [Google Scholar] [CrossRef] [PubMed]
  304. do Rosario, V.; Fitzgerald, Z.; Broyd, S.; Paterson, A.; Roodenrys, S.; Thomas, S.; Bliokas, V.; Potter, J.; Walton, K.; Weston-Green, K.; et al. Food anthocyanins decrease concentrations of TNF-α in older adults with mild cognitive impairment: A randomized, controlled, double blind clinical trial. Nutr. Metab. Cardiovasc. Dis. 2021, 31, 950–960. [Google Scholar] [CrossRef] [PubMed]
  305. do Rosario, V.; Chang, C.; Spencer, J.; Alahakone, T.; Roodenrys, S.; Francois, M.; Weston-Green, K.; Hölzel, N.; Nichols, D.S.; Kent, K.; et al. Anthocyanins attenuate vascular and inflammatory responses to a high fat high energy meal challenge in overweight older adults: A cross-over, randomized, double-blind clinical trial. Clin. Nutr. 2021, 40, 879–889. [Google Scholar] [CrossRef] [PubMed]
  306. Nakamura, Y.; Watanabe, H.; Tanaka, A.; Nishihira, J.; Murayama, N. Effect of quercetin glycosides on cognitive functions and cerebral blood flow: A randomized, double-blind, and placebo-controlled study. Eur. Rev. Med. Pharmacol. Sci. 2022, 26, 8700–8712. [Google Scholar] [PubMed]
  307. Gonzales, M.; Garbarino, V.R.; Marques Zilli, E.; Petersen, R.C.; Kirkland, J.L.; Tchkonia, T.; Musi, N.; Seshadri, S.; Craft, S.; Orr, M.E. Senolytic Therapy to Modulate the Progression of Alzheimer’s Disease (SToMP-AD): A Pilot Clinical Trial. J. Prev. Alzheimers Dis. 2022, 9, 22–29. [Google Scholar] [CrossRef] [PubMed]
  308. Bhattacharjee, A.; Purohit, P.; Roy, P.K. Neuroprotective Drug Discovery From Phytochemicals and Metabolites for CNS Viral Infection: A Systems Biology Approach With Clinical and Imaging Validation. Front. Neurosci. 2022, 16, 917867. [Google Scholar] [CrossRef]
  309. Otsuka, Y.; Miyamoto, N.; Nagai, A.; Izumo, T.; Nakai, M.; Fukuda, M.; Arimitsu, T.; Yamada, Y.; Hashimoto, T. Effects of Quercetin Glycoside Supplementation Combined With Low-Intensity Resistance Training on Muscle Quantity and Stiffness: A Randomized, Controlled Trial. Front. Nutr. 2022, 9, 912217. [Google Scholar] [CrossRef]
  310. Sgrò, P.; Ceci, R.; Lista, M.; Patrizio, F.; Sabatini, S.; Felici, F.; Sacchetti, M.; Bazzucchi, I.; Duranti, G.; Di Luigi, L. Quercetin Modulates IGF-I and IGF-II Levels After Eccentric Exercise-Induced Muscle-Damage: A Placebo-Controlled Study. Front. Endocrinol. 2021, 12, 745959. [Google Scholar] [CrossRef]
  311. Leyva-Soto, A.; Chavez-Santoscoy, A.R.; Porras, O.; Hidalgo-Ledesma, M.; Serrano-Medina, A.; Ramírez-Rodríguez, A.A.; Castillo-Martinez, A.N. Epicatechin and quercetin exhibit in vitro antioxidant effect, improve biochemical parameters related to metabolic syndrome, and decrease cellular genotoxicity in humans. Food Res. Int. 2021, 142, 110101. [Google Scholar] [CrossRef] [PubMed]
  312. Tamtaji, O.; Milajerdi, A.; Dadgostar, E.; Kolahdooz, F.; Chamani, M.; Amirani, E.; Mirzaei, H.; Asemi, Z. The Effects of Quercetin Supplementation on Blood Pressures and Endothelial Function Among Patients with Metabolic Syndrome and Related Disorders: A Systematic Review and Meta-analysis of Randomized Controlled Trials. Curr. Pharm. Des. 2019, 25, 1372–1384. [Google Scholar] [CrossRef] [PubMed]
  313. Tabrizi, R.; Tamtaji, O.R.; Mirhosseini, N.; Lankarani, K.B.; Akbari, M.; Heydari, S.T.; Dadgostar, E.; Asemi, Z. The effects of quercetin supplementation on lipid profiles and inflammatory markers among patients with metabolic syndrome and related disorders: A systematic review and meta-analysis of randomized controlled trials. Crit. Rev. Food Sci. Nutr. 2020, 60, 1855–1868. [Google Scholar] [CrossRef]
  314. Ou, Q.; Zheng, Z.; Zhao, Y.; Lin, W. Impact of quercetin on systemic levels of inflammation: A meta-analysis of randomised controlled human trials. Int. J. Food Sci. Nutr. 2020, 71, 152–163. [Google Scholar] [CrossRef]
  315. Vahdat-Lasemi, F.; Aghaee-Bakhtiari, S.H.; Tasbandi, A.; Jaafari, M.R.; Sahebkar, A. Targeting interleukin-β by plant-derived natural products: Implications for the treatment of atherosclerotic cardiovascular disease. Phytother. Res. 2021, 35, 5596–5622. [Google Scholar] [CrossRef] [PubMed]
  316. Duan, H.; Zhang, Q.; Liu, J.; Li, R.; Wang, D.; Peng, W.; Wu, C. Suppression of apoptosis in vascular endothelial cell, the promising way for natural medicines to treat atherosclerosis. Pharmacol. Res. 2021, 168, 105599. [Google Scholar] [CrossRef] [PubMed]
  317. Zhou, D.; Luo, M.; Shang, A.; Mao, Q.Q.; Li, B.Y.; Gan, R.Y.; Li, H.B. Antioxidant Food Components for the Prevention and Treatment of Cardiovascular Diseases: Effects, Mechanisms, and Clinical Studies. Oxid. Med. Cell. Longev. 2021, 2021, 6627355. [Google Scholar] [CrossRef]
  318. Dehghani, F.; Jandaghi, S.S.S.H.; Janani, L.; Sarebanhassanabadi, M.; Emamat, H.; Vafa, M. Effects of quercetin supplementation on inflammatory factors and quality of life in post-myocardial infarction patients: A double blind, placebo-controlled, randomized clinical trial. Phytother. Res. 2021, 35, 2085–2098. [Google Scholar] [CrossRef]
  319. Vaez, S.; Parivr, K.; Amidi, F.; Rudbari, N.H.; Moini, A.; Amini, N. Quercetin and polycystic ovary syndrome; inflammation, hormonal parameters and pregnancy outcome: A randomized clinical trial. Am. J. Reprod. Immunol. 2022, 89, e13644. [Google Scholar] [CrossRef]
  320. Hipólito-Reis, M.; Neto, A.C.; Neves, D. Impact of curcumin, quercetin, or resveratrol on the pathophysiology of endometriosis: A systematic review. Phytother. Res. 2022, 36, 2416–2433. [Google Scholar] [CrossRef]
  321. Shohan, M.; Nashibi, R.; Mahmoudian-Sani, M.R.; Abolnezhadian, F.; Ghafourian, M.; Alavi, S.M.; Sharhani, A.; Khodadadi, A. The therapeutic efficacy of quercetin in combination with antiviral drugs in hospitalized COVID-19 patients: A randomized controlled trial. Eur. J. Pharmacol. 2022, 914, 174615. [Google Scholar] [CrossRef] [PubMed]
  322. Henning, S.; Wang, P.; Lee, R.P.; Trang, A.; Husari, G.; Yang, J.; Grojean, E.M.; Ly, A.; Hsu, M.; Heber, D.; et al. Prospective randomized trial evaluating blood and prostate tissue concentrations of green tea polyphenols and quercetin in men with prostate cancer. Food Funct. 2020, 11, 4114–4122. [Google Scholar] [CrossRef] [PubMed]
  323. Han, M.; Barreto, T.A.; Martinez, F.J.; Comstock, A.T.; Sajjan, U.S. Randomised clinical trial to determine the safety of quercetin supplementation in patients with chronic obstructive pulmonary disease. BMJ Open. Respir. Res. 2020, 7, e000392. [Google Scholar] [CrossRef] [PubMed]
  324. Zhang, Z.; Peng, X.; Li, S.; Zhang, N.; Wang, Y.; Wei, H. Isolation and identification of quercetin degrading bacteria from human fecal microbes. PLoS ONE 2014, 9, e90531. [Google Scholar] [CrossRef] [PubMed]
  325. Peng, X.; Zhang, Z.; Zhang, N.; Liu, L.; Li, S.; Wei, H. In vitro catabolism of quercetin by human fecal bacteria and the antioxidant capacity of its catabolites. Food Nutr. Res. 2014, 15, 58. [Google Scholar] [CrossRef]
  326. Jaganath, I.; Mullen, W.; Lean, M.E.; Edwards, C.A.; Crozier, A. In vitro catabolism of rutin by human fecal bacteria and the antioxidant capacity of its catabolites. Free. Radic. Biol. Med. 2009, 47, 1180–1189. [Google Scholar] [CrossRef] [PubMed]
  327. Zhang, Z.; Peng, X.; Zhang, N.; Liu, L.; Wang, Y.; Ou, S. Cytotoxicity comparison of quercetin and its metabolites from in vitro fermentation of several gut bacteria. Food Funct. 2014, 5, 2152–2156. [Google Scholar] [CrossRef]
  328. Blaut, M.; Schoefer, L.; Braune, A. Transformation of flavonoids by intestinal microorganisms. Int. J. Vitam. Nutr. Res. 2003, 73, 79–87. [Google Scholar] [CrossRef]
  329. Jain, J.; Hasan, W.; Biswas, P.; Yadav, R.S.; Jat, D. Neuroprotective effect of quercetin against rotenone-induced neuroinflammation and alterations in mice behavior. J. Biochem. Mol. Toxicol. 2022, 36, e23165. [Google Scholar] [CrossRef]
  330. Madiha, S.; Batool, Z.; Tabassum, S.; Liaquat, L.; Sadir, S.; Shahzad, S.; Naqvi, F.; Saleem, S.; Yousuf, S.; Nawaz, A.; et al. Quercetin exhibits potent antioxidant activity, restores motor and non-motor deficits induced by rotenone toxicity. PLoS ONE 2021, 16, e0258928. [Google Scholar] [CrossRef]
  331. Paula, P.; Angelica Maria, S.G.; Luis, C.H.; Gloria Patricia, C.G. Preventive Effect of Quercetin in a Triple Transgenic Alzheimer’s Disease Mice Model. Molecules 2019, 24, 2287. [Google Scholar] [CrossRef] [PubMed]
  332. Ossola, B.; Kääriäinen, T.M.; Männistö, P. The multiple faces of quercetin in neuroprotection. Expert Opin. Drug. Saf. 2009, 8, 397–409. [Google Scholar] [CrossRef] [PubMed]
  333. Song, J.; He, Y.; Luo, C.; Feng, B.; Ran, F.; Xu, H.; Ci, Z.; Xu, R.; Han, L.; Zhang, D. New progress in the pharmacology of protocatechuic acid: A compound ingested in daily foods and herbs frequently and heavily. Pharmacol. Res. 2020, 161, 105109. [Google Scholar] [CrossRef]
  334. Chen, Y.; Wang, J.; Zou, L.; Cao, H.; Ni, X.; Xiao, J. Dietary proanthocyanidins on gastrointestinal health and the interactions with gut microbiota. Crit. Rev. Food Sci. Nutr. 2022, 3, 1–24. [Google Scholar]
  335. Ou, K.; Sarnoski, P.; Schneider, K.R.; Song, K.; Khoo, C.; Gu, L. Microbial catabolism of procyanidins by human gut microbiota. Mol. Nutr. Food Res. 2014, 58, 2196–2205. [Google Scholar] [CrossRef]
  336. Rauf, A.; Imran, M.; Abu-Izneid, T.; Iahtisham-Ul-Haq Patel, S.; Pan, X.; Naz, S.; Sanches Silva, A.; Saeed, F.; Rasul Suleria, H.A. Proanthocyanidins: A comprehensive review. Biomed. Pharmacother. 2019, 116, 108999. [Google Scholar] [CrossRef]
  337. Cecarini, V.; Cuccioloni, M.; Zheng, Y.; Bonfili, L.; Gong, C.; Angeletti, M.; Mena, P.; Del Rio, D.; Eleuteri, A.M. Flavan-3-ol Microbial Metabolites Modulate Proteolysis in Neuronal Cells Reducing Amyloid-beta (1-42) Levels. Mol. Nutr. Food Res. 2021, 65, e2100380. [Google Scholar] [CrossRef]
  338. Ruotolo, R.; Minato, I.; La Vitola, P.; Artioli, L.; Curti, C.; Franceschi, V.; Brindani, N.; Amidani, D.; Colombo, L.; Salmona, M.; et al. 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, e1900890. [Google Scholar] [CrossRef]
  339. Heiss, C.; Istas, G.; Feliciano, R.P.; Weber, T.; Wang, B.; Favari, C.; Mena, P.; Del Rio, D.; Rodriguez-Mateos, A. Daily consumption of cranberry improves endothelial function in healthy adults: A double blind randomized controlled trial. Food Funct. 2022, 13, 3812–3838. [Google Scholar] [CrossRef] [PubMed]
  340. Monagas, M.; Urpi-Sarda, M.; Sánchez-Patán, F.; Llorach, R.; Garrido, I.; Gómez-Cordovés, C.; Andres-Lacueva, C.; Bartolomé, B. Insights into the metabolism and microbial biotransformation of dietary flavan-3-ols and the bioactivity of their metabolites. Food Funct. 2010, 1, 233–253. [Google Scholar] [CrossRef]
  341. Rice-Evans, C.; Spencer, J.P.; Schroeter, H.; Rechner, A.R. Bioavailability of flavonoids and potential bioactive forms in vivo. Drug. Metabol. Drug. Interact. 2000, 17, 291–310. [Google Scholar] [CrossRef] [PubMed]
  342. Angelino, D.; Carregosa, D.; Domenech-Coca, C.; Savi, M.; Figueira, I.; Brindani, N.; Jang, S.; Lakshman, S.; Molokin, A.; Urban, J.F., Jr.; et al. 5-(Hydroxyphenyl)-γ-Valerolactone-Sulfate, a Key Microbial Metabolite of Flavan-3-ols, Is Able to Reach the Brain: Evidence from Different in Silico, In Vitro and In Vivo Experimental Models. Nutrients 2019, 11, 2678. [Google Scholar] [CrossRef] [PubMed]
  343. Di Meo, F.; Valentino, A.; Petillo, O.; Peluso, G.; Filosa, S.; Crispi, S. Bioactive Polyphenols and Neuromodulation: Molecular Mechanisms in Neurodegeneration. Int. J. Mol. Sci. 2020, 21, 2564. [Google Scholar] [CrossRef] [PubMed]
  344. Billowria, K.; Ali, R.; Rangra, N.K.; Kumar, R.; Chawla, P.A. Bioactive Flavonoids: A Comprehensive Review on Pharmacokinetics and Analytical Aspects. Crit. Rev. Anal. Chem. 2022, 5, 1–15. [Google Scholar] [CrossRef]
  345. Chen, B.; Zhang, W.; Lin, C.; Zhang, L. A Comprehensive Review on Beneficial Effects of Catechins on Secondary Mitochondrial Diseases. Int. J. Mol. Sci. 2022, 23, 11569. [Google Scholar] [CrossRef]
  346. Mehmood, S.; Maqsood, M.; Mahtab, N.; Khan, M.I.; Sahar, A.; Zaib, S.; Gul, S. Epigallocatechin gallate: Phytochemistry, bioavailability, utilization challenges, and strategies. J. Food Biochem. 2022, 46, e14189. [Google Scholar] [CrossRef]
  347. Rapisarda, P.; Amenta, M.; Ballistreri, G.; Fabroni, S.; Timpanaro, N. Distribution, Antioxidant Capacity, Bioavailability and Biological Properties of Anthocyanin Pigments in Blood Oranges and Other Citrus Species. Molecules 2022, 27, 8675. [Google Scholar] [CrossRef]
  348. Shen, N.; Wang, T.; Gan, Q.; Liu, S.; Wang, L.; Jin, B. Plant flavonoids: Classification, distribution, biosynthesis, and antioxidant activity. Food Chem. 2022, 383, 132531. [Google Scholar] [CrossRef]
  349. Wdowiak, K.; Walkowiak, J.; Pietrzak, R.; Bazan-Woźniak, A.; Cielecka-Piontek, J. Bioavailability of Hesperidin and Its Aglycone Hesperetin-Compounds Found in Citrus Fruits as a Parameter Conditioning the Pro-Health Potential (Neuroprotective and Antidiabetic Activity)-Mini-Review. Nutrients 2022, 14, 2647. [Google Scholar] [CrossRef]
  350. Hamsalakshmi Alex, A.; Arehally Marappa, M.; Joghee, S.; Chidambaram, S.B. Therapeutic benefits of flavonoids against neuroinflammation: A systematic review. Inflammopharmacology 2022, 30, 111–136. [Google Scholar] [CrossRef]
  351. Spencer, J. Metabolism of tea flavonoids in the gastrointestinal tract. J. Nutr. 2003, 133, 3255S–3261S. [Google Scholar] [CrossRef] [PubMed]
  352. Thilakarathna, S.; Rupasinghe, H.P. Flavonoid bioavailability and attempts for bioavailability enhancement. Nutrients 2013, 5, 3367–3387. [Google Scholar] [CrossRef] [PubMed]
  353. Zhang, X.; Song, X.; Hu, X.; Chen, F.; Ma, C. Health benefits of proanthocyanidins linking with gastrointestinal modulation: An updated review. Food Chem. 2023, 404 Pt A, 134596. [Google Scholar] [CrossRef]
  354. Zhao, Y.; Zhong, X.; Yan, J.; Sun, C.; Zhao, X.; Wang, X. Potential roles of gut microbes in biotransformation of natural products: An overview. Front. Microbiol. 2022, 13, 956378. [Google Scholar] [CrossRef] [PubMed]
  355. Ballesteros-Álvarez, J.; Nguyen, W.; Sivapatham, R.; Rane, A.; Andersen, J.K. Urolithin A reduces amyloid-beta load and improves cognitive deficits uncorrelated with plaque burden in a mouse model of Alzheimer’s disease. Geroscience 2023, 45, 1095–1113. [Google Scholar] [CrossRef] [PubMed]
  356. Cao, X.; Wan, H.; Wan, H. Urolithin A induces protective autophagy to alleviate inflammation, oxidative stress, and endoplasmic reticulum stress in pediatric pneumonia. Allergol. Immunopathol. 2022, 50, 147–153. [Google Scholar] [CrossRef]
  357. Chen, P.; Guo, Z.; Chen, F.; Wu, Y.; Zhou, B. Recent Advances and Perspectives on the Health Benefits of Urolithin B, A Bioactive Natural Product Derived from Ellagitannins. Front. Pharmacol. 2022, 13, 917266. [Google Scholar] [CrossRef]
  358. Al Khalaf, A.; Abdulrahman, A.O.; Kaleem, M.; Nur, S.M.; Asseri, A.H.; Choudhry, H.; Khan, M.I. Comparative Analysis of the Impact of Urolithins on the Composition of the Gut Microbiota in Normal-Diet Fed Rats. Nutrients 2021, 13, 3885. [Google Scholar] [CrossRef]
  359. Singh, A.; D’Amico, D.; Andreux, P.A.; Dunngalvin, G.; Kern, T.; Blanco-Bose, W.; Auwerx, J.; Aebischer, P.; Rinsch, C. Direct supplementation with Urolithin A overcomes limitations of dietary exposure and gut microbiome variability in healthy adults to achieve consistent levels across the population. Eur. J. Clin. Nutr. 2022, 76, 297–308. [Google Scholar] [CrossRef]
  360. Hasani, A.; Ebrahimzadeh, S.; Hemmati, F.; Khabbaz, A.; Hasani, A.; Gholizadeh, P. The role of Akkermansia muciniphila in obesity, diabetes and atherosclerosis. J. Med. Microbiol. 2021, 70, 001435. [Google Scholar] [CrossRef]
  361. Zhou, K. Strategies to promote abundance of Akkermansia muciniphila, an emerging probiotics in the gut, evidence from dietary intervention studies. J. Funct. Foods 2017, 33, 194–201. [Google Scholar] [CrossRef] [PubMed]
  362. Everard, A.; Belzer, C.; Geurts, L.; Ouwerkerk, J.P.; Druart, C.; Bindels, L.B.; Guiot, Y.; Derrien, M.; Muccioli, G.G.; Delzenne, N.M.; et al. Cross-talk between Akkermansia muciniphila and intestinal epithelium controls diet-induced obesity. Proc Natl Acad Sci U S A. 2013, 110, 9066–9071. [Google Scholar] [CrossRef] [PubMed]
  363. García-Villalba, R.; Beltrán, D.; Frutos, M.D.; Selma, M.V.; Espín, J.C.; Tomás-Barberán, F.A. Metabolism of different dietary phenolic compounds by the urolithin-producing human-gut bacteria Gordonibacter urolithinfaciens and Ellagibacter isourolithinifaciens. Food Funct. 2020, 11, 7012–7022. [Google Scholar] [CrossRef] [PubMed]
  364. Keranmu, A.; Pan, L.B.; Yu, H.; Fu, J.; Liu, Y.F.; Amuti, S.; Han, P.; Ma, S.R.; Xu, H.; Zhang, Z.W.; et al. The potential biological effects of quercetin based on pharmacokinetics and multi-targeted mechanism in vivo. J. Asian Nat. Prod. Res. 2022, 24, 403–431. [Google Scholar] [CrossRef] [PubMed]
  365. Lan, H.; Hong, W.; Qian, D.; Peng, F.; Li, H.; Liang, C.; Du, M.; Gu, J.; Mai, J.; Bai, B.; et al. Quercetin modulates the gut microbiota as well as the metabolome in a rat model of osteoarthritis. Bioengineered 2021, 12, 6240–6250. [Google Scholar] [CrossRef] [PubMed]
  366. Sankaranarayanan, R.; Sekhon, P.K.; Ambat, A.; Nelson, J.; Jose, D.; Bhat, G.J.; Scaria, J. Screening of Human Gut Bacterial Culture Collection Identifies Species That Biotransform Quercetin into Metabolites with Anticancer Properties. Int. J. Mol. Sci. 2021, 22, 7045. [Google Scholar] [CrossRef]
  367. Chen, Y.; Peng, F.; Xing, Z.; Chen, J.; Peng, C.; Li, D. Beneficial effects of natural flavonoids on neuroinflammation. Front. Immunol. 2022, 13, 1006434. [Google Scholar] [CrossRef]
  368. Payne, A.; Nahashon, S.; Taka, E.; Adinew, G.M.; Soliman, K.F.A. Epigallocatechin-3-Gallate (EGCG): New Therapeutic Perspectives for Neuroprotection, Aging, and Neuroinflammation for the Modern Age. Biomolecules 2022, 12, 371. [Google Scholar] [CrossRef]
  369. Burberry, A.; Wells, M.F.; Limone, F.; Couto, A.; Smith, K.S.; Keaney, J.; Gillet, G.; van Gastel, N.; Wang, J.Y.; Pietilainen, O.; et al. C9orf72 suppresses systemic and neural inflammation induced by gut bacteria. Nature 2020, 582, 89–94. [Google Scholar] [CrossRef]
  370. Mou, Y.; Du, Y.; Zhou, L.; Yue, J.; Hu, X.; Liu, Y.; Chen, S.; Lin, X.; Zhang, G.; Xiao, H.; et al. Gut Microbiota Interact with the Brain Through Systemic Chronic Inflammation: Implications on Neuroinflammation, Neurodegeneration, and Aging. Front. Immunol. 2022, 13, 796288. [Google Scholar] [CrossRef]
  371. 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. [Google Scholar] [CrossRef] [PubMed]
  372. Caballero, V.; Estévez, M.; Tomás-Barberán, F.A.; Morcuende, D.; Martín, I.; Delgado, J. Biodegradation of Punicalagin into Ellagic Acid by Selected Probiotic Bacteria: A Study of the Underlying Mechanisms by MS-Based Proteomics. J. Agric. Food Chem. 2022, 70, 16273–16285. [Google Scholar] [CrossRef] [PubMed]
  373. Yin, Z.; Zhang, J.; Chen, L.; Guo, Q.; Yang, B.; Zhang, W.; Kang, W. Anticancer Effects and Mechanisms of Action of Plumbagin: Review of Research Advances. Biomed. Res. Int. 2020, 2020, 6940953. [Google Scholar] [CrossRef] [PubMed]
  374. Zou, W.; Deng, L.; Wu, H.; Liu, Z.; Lu, W.; He, Y. Untargeted Metabolomics Profiling Reveals Beneficial Changes in Milk of Sows Supplemented with Fermented Compound Chinese Medicine Feed Additive. Animals 2022, 12, 2879. [Google Scholar] [CrossRef]
  375. Liu, Y.; Lu, Y.; Li, X.; Zhang, Z.; Sun, L.; Wang, Y.; He, Z.; Liu, Z.; Zhu, L.; Fu, L. Kaempferol suppression of acute colitis is regulated by the efflux transporters BCRP and MRP2. Eur. J. Pharm. Sci. 2022, 179, 106303. [Google Scholar] [CrossRef] [PubMed]
  376. Gao, T.; Ye, F.; Tan, Y.; Peng, M.; Yuan, F.; Liu, Z.; Zhou, D.; Yang, K.; Liu, W.; Guo, R.; et al. Metabolomics and proteomics analyses revealed mechanistic insights on the antimicrobial activity of epigallocatechin gallate against Streptococcus suis. Front. Cell. Infect. Microbiol. 2022, 12, 973282. [Google Scholar] [CrossRef] [PubMed]
  377. Yang, R.; Li, J.; Wang, J.; Wang, Y.; Ma, F.; Zhai, R.; Li, P. Kaempferol inhibits the growth of Helicobacter pylori in a manner distinct from antibiotics. J. Food Biochem. 2022, 46, e14210. [Google Scholar] [CrossRef]
  378. Santi, M.; Ortega, M.G.; Peralta, M.A. A State-of-the-art Review and Prospective Therapeutic Applications of Prenyl Flavonoids as Chemosensitizers against Antifungal Multidrug Resistance in Candida albicans. Curr. Med. Chem. 2022, 29, 4251–4281. [Google Scholar]
  379. Waditzer, M.; Bucar, F. Flavonoids as Inhibitors of Bacterial Efflux Pumps. Molecules 2021, 26, 6904. [Google Scholar] [CrossRef]
  380. Palko-Łabuz, A.; Błaszczyk, M.; Środa-Pomianek, K.; Wesołowska, O. Isobavachalcone as an Active Membrane Perturbing Agent and Inhibitor of ABCB1 Multidrug Transporter. Molecules 2021, 26, 4637. [Google Scholar] [CrossRef]
  381. 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]
  382. Wen, L.; Jiang, Y.; Yang, J.; Zhao, Y.; Tian, M.; Yang, B. Structure, bioactivity, and synthesis of methylated flavonoids. Ann. N. Y. Acad. Sci. 2017, 1398, 120–129. [Google Scholar] [CrossRef] [PubMed]
  383. Fang, J. Bioavailability of anthocyanins. Drug. Metab. Rev. 2014, 46, 508–520. [Google Scholar] [CrossRef] [PubMed]
  384. Chen, L.; Cao, H.; Huang, Q.; Xiao, J.; Teng, H. Absorption, metabolism and bioavailability of flavonoids: A review. Crit. Rev. Food Sci. Nutr. 2022, 62, 7730–7742. [Google Scholar] [CrossRef] [PubMed]
  385. Zhang, M.; Zhu, S.; Yang, W.; Huang, Q.; Ho, C.T. The biological fate and bioefficacy of citrus flavonoids: Bioavailability, biotransformation, and delivery systems. Food Funct. 2021, 12, 3307–3323. [Google Scholar] [CrossRef] [PubMed]
  386. Khan, A.; Rashid, R.; Fatima, N.; Mahmood, S.; Mir, S.; Khan, S.; Jabeen, N.; Murtaza, G. Pharmacological activities of protocatechuic acid. Acta Pol. Pharm. 2015, 72, 643–650. [Google Scholar]
  387. Semaming, Y.; Pannengpetch, P.; Chattipakorn, S.C.; Chattipakorn, N. Pharmacological properties of protocatechuic Acid and its potential roles as complementary medicine. Evid. Based Complement. Alternat. Med. 2015, 2015, 593902. [Google Scholar] [CrossRef] [PubMed]
  388. Oboh, G.; Adewuni, T.M.; Ademiluyi, A.O.; Olasehinde, T.A.; Ademosun, A.O. Phenolic Constituents and Inhibitory Effects of Hibiscus sabdariffa L. (Sorrel) Calyx on Cholinergic, Monoaminergic, and Purinergic Enzyme Activities. J. Diet. Suppl. 2018, 15, 610–922. [Google Scholar] [CrossRef]
  389. Haskell-Ramsay, C.; Schmitt, J.; Actis-Goretta, L. The Impact of Epicatechin on Human Cognition: The Role of Cerebral Blood Flow. Nutrients 2018, 10, 986. [Google Scholar] [CrossRef]
  390. Kazem, Y.; Mahmoud, M.H.; Essa, H.A.; Azmy, O.; Kandeel, W.A.; Al-Moghazy, M.; El-Attar, I.; Hasheesh, A.; Mehanna, N.S. Role of Bifidobacterium spp. intake in improving depressive mood and well-being and its link to kynurenine blood level: An interventional study. J. Complement. Integr. Med. 2021. [Google Scholar] [CrossRef]
  391. Peirce, J.; Alviña, K. The role of inflammation and the gut microbiome in depression and anxiety. J. Neurosci. Res. 2019, 97, 1223–1241. [Google Scholar] [CrossRef] [PubMed]
  392. Guan, S.; Ge, D.; Liu, T.Q.; Ma, X.H.; Cui, Z.F. Protocatechuic acid promotes cell proliferation and reduces basal apoptosis in cultured neural stem cells. Toxicol In Vitro 2009, 23, 201–208. [Google Scholar] [CrossRef] [PubMed]
  393. Jiang, C.; Li, G.; Huang, P.; Liu, Z.; Zhao, B. The Gut Microbiota and Alzheimer’s Disease. J. Alzheimers. Dis. 2017, 58, 1–15. [Google Scholar] [CrossRef] [PubMed]
  394. Kang, J.; Zivkovic, A.M. The Potential Utility of Prebiotics to Modulate Alzheimer’s Disease: A Review of the Evidence. Microorganisms 2021, 9, 2310. [Google Scholar] [CrossRef]
  395. Li, H.; Zheng, T.; Lian, F.; Xu, T.; Yin, W.; Jiang, Y. Anthocyanin-rich blueberry extracts and anthocyanin metabolite protocatechuic acid promote autophagy-lysosomal pathway and alleviate neurons damage in in vivo and in vitro models of Alzheimer’s disease. Nutrition 2022, 93, 111473. [Google Scholar] [CrossRef]
  396. Liu, S.; Gao, J.; Zhu, M.; Liu, K.; Zhang, H.L. Gut Microbiota and Dysbiosis in Alzheimer’s Disease: Implications for Pathogenesis and Treatment. Mol. Neurobiol. 2020, 57, 5026–5043. [Google Scholar] [CrossRef]
  397. Megur, A.; Baltriukienė, D.; Bukelskienė, V.; Burokas, A. The Microbiota-Gut-Brain Axis and Alzheimer’s Disease: Neuroinflammation Is to Blame? Nutrients 2020, 13, 37. [Google Scholar] [CrossRef]
  398. 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]
  399. Sochocka, M.; Donskow-Łysoniewska, K.; Diniz, B.S.; Kurpas, D.; Brzozowska, E.; Leszek, J. The Gut Microbiome Alterations and Inflammation-Driven Pathogenesis of Alzheimer’s Disease-a Critical Review. Mol. Neurobiol. 2019, 56, 1841–1851. [Google Scholar] [CrossRef]
  400. Fitzgerald, E.; Murphy, S.; Martinson, H.A. Alpha-Synuclein Pathology and the Role of the Microbiota in Parkinson’s Disease. Front. Neurosci. 2019, 13, 369. [Google Scholar] [CrossRef]
  401. Johnson, A.; Ou, Z.A.; Gordon, R.; Saminathan, H. Environmental neurotoxicants and inflammasome activation in Parkinson’s disease—A focus on the gut-brain axis. Int J Biochem Cell Biol. 2022, 142, 106113. [Google Scholar] [CrossRef] [PubMed]
  402. Toh, T.; Chong, C.W.; Lim, S.Y.; Bowman, J.; Cirstea, M.; Lin, C.H.; Chen, C.C.; Appel-Cresswell, S.; Finlay, B.B.; Tan, A.H. Gut microbiome in Parkinson’s disease: New insights from meta-analysis. Park. Relat. Disord. 2021, 94, 1–9. [Google Scholar] [CrossRef]
  403. Yap, C.; Henders, A.K.; Alvares, G.A.; Wood, D.L.A.; Krause, L.; Tyson, G.W.; Restuadi, R.; Wallace, L.; McLaren, T.; Hansell, N.K.; et al. Autism-related dietary preferences mediate autism-gut microbiome associations. Cell. 2021, 184, 5916–5931.e17. [Google Scholar] [CrossRef] [PubMed]
  404. Tu, T.; Zhao, C. Treating autism spectrum disorder by intervening with gut microbiota. J. Med. Microbiol. 2021, 70, 001469. [Google Scholar] [CrossRef]
  405. Chernikova, M.; Flores, G.D.; Kilroy, E.; Labus, J.S.; Mayer, E.A.; Aziz-Zadeh, L. The Brain-Gut-Microbiome System: Pathways and Implications for Autism Spectrum Disorder. Nutrients 2021, 13, 4497. [Google Scholar] [CrossRef] [PubMed]
  406. Li, Q.; Liang, J.; Fu, N.; Han, Y.; Qin, J. A Ketogenic Diet and the Treatment of Autism Spectrum Disorder. Front. Pediatr. 2021, 9, 650624. [Google Scholar] [CrossRef]
  407. Evans, S.; Bassis, C.M.; Hein, R.; Assari, S.; Flowers, S.A.; Kelly, M.B.; Young, V.B.; Ellingrod, V.E.; McInnis, M.G. The gut microbiome composition associates with bipolar disorder and illness severity. J. Psychiatr. Res. 2017, 87, 23–29. [Google Scholar] [CrossRef]
  408. Flowers, S.; Ward, K.M.; Clark, C.T. The Gut Microbiome in Bipolar Disorder and Pharmacotherapy Management. Neuropsychobiology 2020, 79, 43–49. [Google Scholar] [CrossRef]
  409. Marx, W.; Moseley, G.; Berk, M.; Jacka, F. Nutritional psychiatry: The present state of the evidence. Proc. Nutr. Soc. 2017, 76, 427–436. [Google Scholar] [CrossRef]
  410. Butler, M.; Mörkl, S.; Sandhu, K.V.; Cryan, J.F.; Dinan, T.G. The Gut Microbiome and Mental Health: What Should We Tell Our Patients? Le microbiote Intestinal et la Santé Mentale: Que Devrions-Nous dire à nos Patients? Can. J. Psychiatry 2019, 64, 747–760. [Google Scholar] [CrossRef]
  411. Tian, P.; Chen, Y.; Zhu, H.; Wang, L.; Qian, X.; Zou, R.; Zhao, J.; Zhang, H.; Qian, L.; Wang, Q.; et al. Bifidobacterium breve CCFM1025 attenuates major depression disorder via regulating gut microbiome and tryptophan metabolism: A randomized clinical trial. Brain Behav. Immun. 2021, 100, 233–241. [Google Scholar] [CrossRef] [PubMed]
  412. Simpson, C.; Diaz-Arteche, C.; Eliby, D.; Schwartz, O.S.; Simmons, J.G.; Cowan, C.G.M. The gut microbiota in anxiety and depression—A systematic review. Clin Psychol Rev. 2021, 83, 101943. [Google Scholar] [CrossRef] [PubMed]
  413. Del Toro-Barbosa, M.; Hurtado-Romero, A.; Garcia-Amezquita, L.E.; García-Cayuela, T. Psychobiotics: Mechanisms of Action, Evaluation Methods and Effectiveness in Applications with Food Products. Nutrients 2020, 12, 3896. [Google Scholar] [CrossRef]
  414. Dahlin, M.; Prast-Nielsen, S. The gut microbiome and epilepsy. EBioMedicine 2019, 44, 741–746. [Google Scholar] [CrossRef] [PubMed]
  415. Fan, Y.; Wang, H.; Liu, X.; Zhang, J.; Liu, G. Crosstalk between the Ketogenic Diet and Epilepsy: From the Perspective of Gut Microbiota. Mediat. Inflamm. 2019, 2019, 9373060. [Google Scholar] [CrossRef] [PubMed]
  416. Ding, M.; Lang, Y.; Shu, H.; Shao, J.; Cui, L. Microbiota-Gut-Brain Axis and Epilepsy: A Review on Mechanisms and Potential Therapeutics. Front. Immunol. 2021, 12, 742449. [Google Scholar] [CrossRef]
  417. Huang, Q.; Xia, J. Influence of the gut microbiome on inflammatory and immune response after stroke. Neurol. Sci. 2021, 42, 4937–4951. [Google Scholar] [CrossRef]
  418. Huang, Q.; Di, L.; Yu, F.; Feng, X.; Liu, Z.; Wei, M.; Luo, Y.; Xia, J. Alterations in the gut microbiome with hemorrhagic transformation in experimental stroke. CNS Neurosci. Ther. 2022, 28, 77–91. [Google Scholar] [CrossRef]
  419. Pluta, R.; Januszewski, S.; Czuczwar, S.J. The Role of Gut Microbiota in an Ischemic Stroke. Int. J. Mol. Sci. 2021, 22, 915. [Google Scholar] [CrossRef]
  420. Terao, J. Factors modulating bioavailability of quercetin-related flavonoids and the consequences of their vascular function. Biochem. Pharmacol. 2017, 139, 15–23. [Google Scholar] [CrossRef]
  421. Yamashiro, K.; Kurita, N.; Urabe, T.; Hattori, N. Role of the Gut Microbiota in Stroke Pathogenesis and Potential Therapeutic Implications. Ann. Nutr. Metab. 2021, 77 (Suppl. S2), 36–44. [Google Scholar] [CrossRef] [PubMed]
  422. Zhang, L.; Ma, J.; Yang, F.; Li, S.; Ma, W.; Chang, X.; Yang, L. Neuroprotective Effects of Quercetin on Ischemic Stroke: A Literature Review. Front. Pharmacol. 2022, 13, 854249. [Google Scholar] [CrossRef] [PubMed]
  423. Sbrana, C.; Avio, L.; Giovannetti, M. Beneficial mycorrhizal symbionts affecting the production of health-promoting phytochemicals. Electrophoresis 2014, 35, 1535–1546. [Google Scholar] [CrossRef] [PubMed]
  424. Round, J.L.; Mazmanian, S.K. The gut microbiome shapes intestinal immune responses during health and disease. Nat. Rev. Immunol. 2009, 9, 313–323. [Google Scholar] [CrossRef]
  425. Li, D.; Sun, T.; Tong, Y.; Le, J.; Yao, Q.; Tao, J.; Liu, H.; Jiao, W.; Mei, Y.; Chen, J.; et al. Gut-microbiome-expressed 3β-hydroxysteroid dehydrogenase degrades estradiol and is linked to depression in premenopausal females. Cell Metab. 2023. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic of the gut-brain axis: (a) demonstration of bidirectional communication by the parasympathetic vagus nerve, and some of the neurodegenerative conditions treated by vagal stimulation; (b) the vagal nerve transports compounds (generated from dietary flavonoids by the gut microbiome) of therapeutic value in the treatment of neurological disorders. Neurodegenerative conditions treated successfully by vagal stimulation are also shown (c).
Figure 1. Schematic of the gut-brain axis: (a) demonstration of bidirectional communication by the parasympathetic vagus nerve, and some of the neurodegenerative conditions treated by vagal stimulation; (b) the vagal nerve transports compounds (generated from dietary flavonoids by the gut microbiome) of therapeutic value in the treatment of neurological disorders. Neurodegenerative conditions treated successfully by vagal stimulation are also shown (c).
Antioxidants 12 00663 g001
Figure 2. Schematic of neural signal transduction. Depiction of a neuron and its functional components (a) and the processes that occurs when a nerve is activated and signal transduction occurs (be). Specific features of the neuron are annotated, including the neural dendrite processes (1) where signal input occurs, the nucleus (2), which regulates neural activity in the neural cell body or soma (3). The myelinated sheath (4) covering the axon (5) ensures neural signal transmission efficiency is maintained. Neural synapses (6) communicate with other neurons in the neural network. Neural transmitters, such as dopamine (7), are stored in a smart gel matrix within the synaptic vesicle supplied by a 12 span transmembrane KS-storage and transport proteoglycan, SV-2 (8). The synaptic vesicle also has a calcium sensing glycoprotein: synaptotagmin (9). When a nerve becomes activated, the cell membrane becomes depolarized in the soma and a wave of membrane depolarization travels down the axon to the synapses. An influx of Ca2+ (10) into the nerve cytosol occurs in neuronal activation; this increase in Ca2+ is detected by synaptotagmin, which mobilises the transport of synaptic vesicles to the synaptic gap by SV-2 (11), and the synaptic vesicles fuse with the de-polarised pre-synaptic membrane (12). This fusion process is regulated by synaptotagmin and SNARE complex (SNAP Receptor) proteins and the neurotransmitters are released into the synaptic gap (14) to be taken up by neurotransmitter receptors on a communicating neuron in the network and the signal is successfully transduced. This is an extremely rapid process occurring in ~50–60 milliseconds.
Figure 2. Schematic of neural signal transduction. Depiction of a neuron and its functional components (a) and the processes that occurs when a nerve is activated and signal transduction occurs (be). Specific features of the neuron are annotated, including the neural dendrite processes (1) where signal input occurs, the nucleus (2), which regulates neural activity in the neural cell body or soma (3). The myelinated sheath (4) covering the axon (5) ensures neural signal transmission efficiency is maintained. Neural synapses (6) communicate with other neurons in the neural network. Neural transmitters, such as dopamine (7), are stored in a smart gel matrix within the synaptic vesicle supplied by a 12 span transmembrane KS-storage and transport proteoglycan, SV-2 (8). The synaptic vesicle also has a calcium sensing glycoprotein: synaptotagmin (9). When a nerve becomes activated, the cell membrane becomes depolarized in the soma and a wave of membrane depolarization travels down the axon to the synapses. An influx of Ca2+ (10) into the nerve cytosol occurs in neuronal activation; this increase in Ca2+ is detected by synaptotagmin, which mobilises the transport of synaptic vesicles to the synaptic gap by SV-2 (11), and the synaptic vesicles fuse with the de-polarised pre-synaptic membrane (12). This fusion process is regulated by synaptotagmin and SNARE complex (SNAP Receptor) proteins and the neurotransmitters are released into the synaptic gap (14) to be taken up by neurotransmitter receptors on a communicating neuron in the network and the signal is successfully transduced. This is an extremely rapid process occurring in ~50–60 milliseconds.
Antioxidants 12 00663 g002
Figure 3. Neurotransmitters and flavonoid metabolites. A comparison of the structure of neurotransmitters (af) conveyed by nerves by vesicular transport, as shown in Figure 2. A few selected flavonoid metabolites generated by the gut microbiome are also shown for comparison (gk). These flavonoid metabolites display a range of activities against neurons and cerebrovascular endothelial cells, and have beneficial properties that combat neuroinflammation, are neuroprotective, and have vasodilatory properties that promote cerebral blood flow in neurological disorders. Some of these metabolites have also been shown to promote mitochondrial biogenesis, improving neural bioenergetics and neuronal function in disorders, such as AD and PD, where a cognitive decline has been observed.
Figure 3. Neurotransmitters and flavonoid metabolites. A comparison of the structure of neurotransmitters (af) conveyed by nerves by vesicular transport, as shown in Figure 2. A few selected flavonoid metabolites generated by the gut microbiome are also shown for comparison (gk). These flavonoid metabolites display a range of activities against neurons and cerebrovascular endothelial cells, and have beneficial properties that combat neuroinflammation, are neuroprotective, and have vasodilatory properties that promote cerebral blood flow in neurological disorders. Some of these metabolites have also been shown to promote mitochondrial biogenesis, improving neural bioenergetics and neuronal function in disorders, such as AD and PD, where a cognitive decline has been observed.
Antioxidants 12 00663 g003
Figure 4. The flavonoids. Classification of the flavonoids, a major sub-category of phenolic compounds (a), showing the diverse modifications (b) that occur on the A, B and C flavone ring structures (c).
Figure 4. The flavonoids. Classification of the flavonoids, a major sub-category of phenolic compounds (a), showing the diverse modifications (b) that occur on the A, B and C flavone ring structures (c).
Antioxidants 12 00663 g004
Figure 5. The many forms of quercetin. As a representative flavonoid, quercetin occurs as a glycosylated form (rutin) in plant tissues (a), which, when ingested, is converted to the aglycone form (b), isorhamnetin-3-glucuronide (c), quercetin also undergoes sulfation (d), or glucosidation (e,f), and glucuronidation (g,h), as shown at specific locations in the flavone A, B, and C rings (i). The C-ring may undergo cleavages at the positions shown when the flavone is processed by the gut microbiome.
Figure 5. The many forms of quercetin. As a representative flavonoid, quercetin occurs as a glycosylated form (rutin) in plant tissues (a), which, when ingested, is converted to the aglycone form (b), isorhamnetin-3-glucuronide (c), quercetin also undergoes sulfation (d), or glucosidation (e,f), and glucuronidation (g,h), as shown at specific locations in the flavone A, B, and C rings (i). The C-ring may undergo cleavages at the positions shown when the flavone is processed by the gut microbiome.
Antioxidants 12 00663 g005
Figure 6. Examples of some of the pharmacologically active complex polyphenolic compounds that have been identified in LeZhe Chinese complimentary medicine herbal preparations (af) and Shuang-Huang-Lian herbal preparations (gi) used to treat neurodegenerative conditions and respiratory infections.
Figure 6. Examples of some of the pharmacologically active complex polyphenolic compounds that have been identified in LeZhe Chinese complimentary medicine herbal preparations (af) and Shuang-Huang-Lian herbal preparations (gi) used to treat neurodegenerative conditions and respiratory infections.
Antioxidants 12 00663 g006
Figure 7. Examples of the varied ring structures of some of the flavonoids covered in this review. Structures of bioactive phenolic compounds identified in Chaihu-Shugan-San, Qingfei Paidu, and Ma Xing Shi Gan traditional Chinese complimentary medical formulations. Identification of phillyrin glycoside, quercetin, glycyrrhizic acid, luteolin, baicalin, baicalein, hesperidin, hesperitin, naringin, naringenin mangiferin, and licochalcone A (al) as bioactive components of such herbal formulations.
Figure 7. Examples of the varied ring structures of some of the flavonoids covered in this review. Structures of bioactive phenolic compounds identified in Chaihu-Shugan-San, Qingfei Paidu, and Ma Xing Shi Gan traditional Chinese complimentary medical formulations. Identification of phillyrin glycoside, quercetin, glycyrrhizic acid, luteolin, baicalin, baicalein, hesperidin, hesperitin, naringin, naringenin mangiferin, and licochalcone A (al) as bioactive components of such herbal formulations.
Antioxidants 12 00663 g007
Figure 8. Generation of Urolithins: complex heterocyclic elligatannin compounds, such as chebulagic acid (a), chebulinic acid (b), and punicalagin (c) occurring in plant foods are converted to ellagic acid (d) and a number of urolithin metabolites by the gut microbiome. Urolithin A (e) and B (f) are active in neural tissues.
Figure 8. Generation of Urolithins: complex heterocyclic elligatannin compounds, such as chebulagic acid (a), chebulinic acid (b), and punicalagin (c) occurring in plant foods are converted to ellagic acid (d) and a number of urolithin metabolites by the gut microbiome. Urolithin A (e) and B (f) are active in neural tissues.
Antioxidants 12 00663 g008
Figure 9. Multicyclic structural forms of the anthocyanins and their bioactive protocatechuic acid and phlorglucinaldehyde metabolites generated by the gut microbiota. Cyanidin-3-O-glucosied (a); Delphinidin-3-O-glucosied (b); Malvidin-3-O-glucosied (c); Pelargonidin-3-O-glucosied (d); protocatechuic acid (e); Phloroglucinaldehyde (f).
Figure 9. Multicyclic structural forms of the anthocyanins and their bioactive protocatechuic acid and phlorglucinaldehyde metabolites generated by the gut microbiota. Cyanidin-3-O-glucosied (a); Delphinidin-3-O-glucosied (b); Malvidin-3-O-glucosied (c); Pelargonidin-3-O-glucosied (d); protocatechuic acid (e); Phloroglucinaldehyde (f).
Antioxidants 12 00663 g009
Figure 10. Degradative pathways used by C. perfringens and B. fragilis of the human gut microbiome to process quercetin into bioactive metabolites, as proposed by Peng et al. [325]. Quercetin-4’glucuronide (a) is initially degraded by β-glucuronidase to form quercetin aglycone (b) then dihydroxyphenyl acetic acid (e) and dihydroxybenzoic acid (i). Alternatatively C ring internal cleavage of quercetin aglycone into an intermediate form (c) can also be converted to hydroxyphenyl acetic acid (f) and phenyl acetic acid (j) or to hydroxyphenyl propionic acid (d) then dihydroxyphenyl acetic acid (g) then to methoxy phenylacetic acid (k) or protocatechuic acid (h).
Figure 10. Degradative pathways used by C. perfringens and B. fragilis of the human gut microbiome to process quercetin into bioactive metabolites, as proposed by Peng et al. [325]. Quercetin-4’glucuronide (a) is initially degraded by β-glucuronidase to form quercetin aglycone (b) then dihydroxyphenyl acetic acid (e) and dihydroxybenzoic acid (i). Alternatatively C ring internal cleavage of quercetin aglycone into an intermediate form (c) can also be converted to hydroxyphenyl acetic acid (f) and phenyl acetic acid (j) or to hydroxyphenyl propionic acid (d) then dihydroxyphenyl acetic acid (g) then to methoxy phenylacetic acid (k) or protocatechuic acid (h).
Antioxidants 12 00663 g010
Figure 11. Generation of tea EGCG metabolites by the gut microbiome, as proposed by Lotito et al. 2011 [227] 5(3,4,5-trihydroxyphenyl)-g-valerolactone has therapeutic pharmacological properties useful in cancer therapy, antioxidant free radical scavenging, and cerebrovascular therapeutic applications in neurodegenerative disorders. (a) Epigallocatechin gallate (EGCG); (b) (-)-Epigallocatechin-3-O-gallate-4’-O-glucuroide; (c) 4’’O-Methyl-(-)epigallocatechin-3-O-gallate; (d) (-)-Epigallocatechin; (e) 4’,4’’-Di-O-Methyl-(-)epigallocatechin-3-O-gallate; (f) 5, (3’4’5’-trihydroxyphenyl)-γ-vcalerolactone; (g) 4’’O-Methyl-(-)epigallocatechin.
Figure 11. Generation of tea EGCG metabolites by the gut microbiome, as proposed by Lotito et al. 2011 [227] 5(3,4,5-trihydroxyphenyl)-g-valerolactone has therapeutic pharmacological properties useful in cancer therapy, antioxidant free radical scavenging, and cerebrovascular therapeutic applications in neurodegenerative disorders. (a) Epigallocatechin gallate (EGCG); (b) (-)-Epigallocatechin-3-O-gallate-4’-O-glucuroide; (c) 4’’O-Methyl-(-)epigallocatechin-3-O-gallate; (d) (-)-Epigallocatechin; (e) 4’,4’’-Di-O-Methyl-(-)epigallocatechin-3-O-gallate; (f) 5, (3’4’5’-trihydroxyphenyl)-γ-vcalerolactone; (g) 4’’O-Methyl-(-)epigallocatechin.
Antioxidants 12 00663 g011
Table 1. Flavonoids that induce Nrf2 expression.
Table 1. Flavonoids that induce Nrf2 expression.
CompoundFlavonoid ClassReference
QuercetinFlavonol[119,120,121]
Myrcetin[122]
Kaempferol[123,124]
Catechin[125]
Gallocatechin[126]
ApigeninFlavone[127]
Luteolin[128,129,130]
Wogonin[131,132]
Chrysin[133,134]
Baicalin[135]
DiadzeinIsoflavone[135]
Genistein[136,137,138]
Biochanin A[139]
HesperidinFlavonone[139,140]
Hesperitin[141,142,143]
Naringenin[144,145]
PelargonidinAnthocyanidin[146]
Cyanidin[147]
Delphinidin[148]
Petunidin[149]
CardamoninChalcone[150,151]
Xanthohumol[152,153]
Isoliquiritigenin[154]
Table 2. A summation of the gut microbiome and processing of therapeutic dietary components which generates bioactive flavonoid metabolites that promote neuronal health and counter neurological deficits.
Table 2. A summation of the gut microbiome and processing of therapeutic dietary components which generates bioactive flavonoid metabolites that promote neuronal health and counter neurological deficits.
Neurological
Disorder
How the Gut Microbiome Impacts the DisorderRef.
ADInduction of Nrf2 expression by flavonoids is neuroprotective countering neuroinflammation. Flavonoids also have intrinsic antioxidant activity against generation of ROS by COX, LOX, MPO, XO. Many microbiome generated flavonoid metabolites retain or have enhanced or new bioactive properties not evident in the native flavonoid and greater bioavailability. Protocatechuic acid is present in the circulation at higher concentrations for significantly longer than native flavonoids and easily crosses the BBB, inhibits accumulation of β-amyloid plaques, hyperphosphorylation of tau protein in neurons and excessive generation of neuroinflammatory ROS, has potent antioxidant and anti-inflammatory properties, is neuroprotective, increases neuronal proliferation and inhibits apoptosis of neural stem cells. Urolithin A potently inhibits the pro-oxidant heme peroxidases MPO and LPO reducing tissue inflammation and significantly reduces phorbol myristate acetate stimulated ROS generated by neutrophils. Urolithin B is a MAO inhibitor and improves cognitive deficits. Urolithin M5 is a neuraminidase inhibitor, urolithin M6 is an inhibitor of LDH. Urolithins are neuroprotective, inhibit Aβ25-35-induced neurotoxicity and neurodegenerative MAO activity. Urolithin A promotes mitophagy and mitochondrial biogenic neuronal function. γ-valerolactones detoxify the effects of amyloid β oligomers. Some flavonoid metabolites have vasodilatory properties that improve cerebrovascular circulation and lower blood pressure.[20,53,392,393,394,395,396,397,398,399]
PDAnti-oxidant properties of flavonoids and metabolites and ability to induce Nrf2 expression is neuroprotective. Gut microbiome generated components may potentially regulate α-synuclein folding lowering the levels of misfolded α-synuclein deposition in pathological protein aggregates leading to neurotoxicity and a decline in neural function. Induction of mitochondrial biogenesis by flavonoid species promotes neuronal bioenergetics and viability.[54,400,401,402]
AutismModulation of the gut microbiota to deliver high-fat low-carbohydrate ketogenic products has proven beneficial in countering the deficits in communication and social interaction evident in autism.[403,404,405,406]
Bipolar
disorder
Diets rich in n-3 fatty acids, folate, S-adenosylmethionine, N-acetyl cysteine and probiotic mediated effects offer promising interventions in the treatment of bipolar disorder [84,407,408,409,410]
Depression
and
Anxiety
Symptoms of depression and anxiety have been shown to be linked to alterations in the microbiota and can be treated by probiotic dietary manipulation with certain food products known as psychobiotics.[390,391,411,412,413]
EpilepsyManipulation of the gut microbiome to deliver a ketogenic high-fat low-carbohydrate diet mimics the fasting state of the body and is beneficial in treatment of drug-resistant epilepsy.[55,414,415,416]
StrokeAnti-oxidant, anti-thrombotic and vasodilatory properties of flavones and flavone metabolites may lower possibility of stroke and improve vascular repair processes. The anti-oxidant, anti-inflammatory, neuroprotective properties of quercetin may minimize the incidence of ischemic stroke. Promotion of endothelial cells by flavonoids improves vascular repair processes.[417,418,419,420,421,422]
Abbreviations used: Nrf2, nuclear factor erythroid 2–related factor 2; ROS, reactive oxygen species; COX, cyclooxygenase; LOX, lipoxygenases; MPO, myeloperoxidase; XO, xanthine oxidase; LPO, lactoperoxidase; BBB, blood brain barrier; MAO, myeloperoxidase’ LDH, lactate dehydrogenase.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Melrose, J. The Potential of Flavonoids and Flavonoid Metabolites in the Treatment of Neurodegenerative Pathology in Disorders of Cognitive Decline. Antioxidants 2023, 12, 663. https://doi.org/10.3390/antiox12030663

AMA Style

Melrose J. The Potential of Flavonoids and Flavonoid Metabolites in the Treatment of Neurodegenerative Pathology in Disorders of Cognitive Decline. Antioxidants. 2023; 12(3):663. https://doi.org/10.3390/antiox12030663

Chicago/Turabian Style

Melrose, James. 2023. "The Potential of Flavonoids and Flavonoid Metabolites in the Treatment of Neurodegenerative Pathology in Disorders of Cognitive Decline" Antioxidants 12, no. 3: 663. https://doi.org/10.3390/antiox12030663

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