Next Article in Journal
Film-Forming Spray of Water-Soluble Chitosan Containing Liposome-Coated Human Epidermal Growth Factor for Wound Healing
Next Article in Special Issue
Cashew Nut Shell Liquid (CNSL) as a Source of Drugs for Alzheimer’s Disease
Previous Article in Journal
Chemo-Preventive Action of Resveratrol: Suppression of p53—A Molecular Targeting Approach
Previous Article in Special Issue
Bio-Guided Fractionation of Stem Bark Extracts from Phyllanthus muellarianus: Identification of Phytocomponents with Anti-Cholinesterase Activity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Therapeutic Potential of Natural Products in Treating Neurodegenerative Disorders and Their Future Prospects and Challenges

1
Department of Environmental Medical Biology, Wonju College of Medicine, Yonsei University, Wonju 26426, Gangwon-do, Korea
2
Department of Global Medical Science, Yonsei University Graduate School, Wonju 26426, Gangwon-do, Korea
3
Department of Laboratory Medicine, Yonsei University Wonju College of Medicine, Yonsei University, Wonju 26426, Gangwon-do, Korea
*
Author to whom correspondence should be addressed.
These authors contributed equally to the work.
Molecules 2021, 26(17), 5327; https://doi.org/10.3390/molecules26175327
Submission received: 29 July 2021 / Revised: 25 August 2021 / Accepted: 27 August 2021 / Published: 2 September 2021
(This article belongs to the Special Issue Plant Derived Natural Products and Age-Related Diseases)

Abstract

:
Natural products derived from plants, as well as their bioactive compounds, have been extensively studied in recent years for their therapeutic potential in a variety of neurodegenerative diseases (NDs), including Alzheimer’s (AD), Huntington’s (HD), and Parkinson’s (PD) disease. These diseases are characterized by progressive dysfunction and loss of neuronal structure and function. There has been little progress in designing efficient treatments, despite impressive breakthroughs in our understanding of NDs. In the prevention and therapy of NDs, the use of natural products may provide great potential opportunities; however, many clinical issues have emerged regarding their use, primarily based on the lack of scientific support or proof of their effectiveness and patient safety. Since neurodegeneration is associated with a myriad of pathological processes, targeting multi-mechanisms of action and neuroprotection approaches that include preventing cell death and restoring the function of damaged neurons should be employed. In the treatment of NDs, including AD and PD, natural products have emerged as potential neuroprotective agents. This current review will highlight the therapeutic potential of numerous natural products and their bioactive compounds thatexert neuroprotective effects on the pathologies of NDs.

1. Introduction

A variety of chronic progressive central nervous system disorders triggered by deterioration and eventual loss of neurons are implicated in neurodegenerative diseases (NDs) [1]. Recently, aging of the population has contributed to the increase in NDs [1,2,3], and age-related diseases including NDs are becoming extremely important due to their irreversibility, lack of effective treatment, and accompanied social and economic burdens [4]. Parkinson’s disease (PD), well-characterized by loss of dopaminergic nigrostriatal neurons; Huntington’s disease (HD), which causes loss of spiny, medium-sized striatal neurons; and Alzheimer’s disease (AD) induced by diffuse brain atrophy, are generally known as NDs. Some disorders were often referred to as NDs, including primary dystonia or tremor [5,6]. Patients with NDs manifest with a wide variety of symptoms that often overlap and range from memory and cognitive impairment to impairment of the person’s ability to walk, communicate, and breathe; these patients often have certain clinical characteristics, such as gradual progression over the years, even reaching decades [4]. Furthermore, oxidative stress, neuroinflammation, dysfunction in mitochondria, dysfunctional protein misfolding and agglomeration, and other biological processes have been linked to neurodegeneration [4,7,8]. These biological pathways have been implicated in the development of NDs and their pathogenesis. To date, extensive research has attempted to explain the process and potential therapeutic goals in the battle against NDs [9]. Neuroprotection strategies and relative mechanisms, therefore, function best by interaction with the pathophysiological transition process to interrupt or postpone the neurodegeneration process [4,10,11]. Natural products are known and have been used for their medicinal properties since ancient times. Natural products and their bioactive compounds have been extensively researched and analyzed in recent years, with a focus on biological processes, nutritional principles, potential health, and therapeutic benefits. In recent decades, numerous studies have confirmed the protective effects of natural products and their bioactive compounds against a variety of diseases, including cardiovascular diseases, diabetes, reproductive diseases, cancers, and NDs [12,13,14]. Natural products for the treatment of NDs have emerged as potential neuroprotective agents. This current review highlights the therapeutic potential of numerous natural products and their bioactive compounds that exert neuroprotective effects on the pathologies of NDs.

2. Potential Therapeutic Targets of Natural Products against Neurodegenerative Diseases

The mechanism of neuronal damage and death has been investigated for several years, from the organ level to the molecular level. Neurotransmitter accumulation in the brain tissue, particularly glutamate, often leads to excessive brain injury, which can overstimulate nerves and cause neuronal death [15]. According to the World Health Organization figures from 2012, more than 35.6 million people suffer from dementia worldwide, with AD accounting for 60–70% of this population [16]. So far, the pathogenesis of AD has not been fully elucidated. The late onset of sporadic AD, the most prevalent type of the disease, is responsible for genetic vulnerability and environmental factors [17]. To summarize, natural products have recently received increasing interest as alternative or integrative treatment agents against AD and other NDs [18,19]. PD’s neuropathological characteristic consists mainly of the accumulation of intracellular protein aggregates, Lewy bodies, and Lewy neuritis, consisting mainly of the mistreated and aggregated forms of alpha-synuclein protein and the gradual loss of nigrostriatal neurons [20,21]. Mutations in the gene coding for the copper/zinc superoxide dismutase-1 (SOD1) enzyme are linked to amyotrophic lateral sclerosis [22,23]. In addition, HD is a hereditary autosomal dominant neurodegenerative condition marked by adult-onset motor dysfunctions, mental disorders, and cognitive loss [24,25,26]. Moreover, HD is associated with an unstable cytosine–adenine–guanine (CAG) expansion in the huntingtin gene on chromosome 4 [27]. Different biological processes, including oxidative stress and neuroinflammatory and mitochondrial dysfunctions, have been involved in the development and pathogenesis of NDs (Figure 1) [28,29]. Oxidative stress has been emphasized in the progression of AD, PD, and other NDs. In addition, oxidative stress leading to free radical attack on neural cells plays a role in calamitous neurodegeneration [1,30]. However, oxidative stress is caused by an imbalance in the formation of reactive oxygen species (ROS) and a lack of antioxidant defense capacity, resulting in cellular damage, DNA repair system impairment, and mitochondrial dysfunction [10,31]. Oxidative stress also aggravates amyloid-beta (Aβ) generation and aggregation and promotes tau protein phosphorylation that can cause a vicious pathogenic cycle for AD [32,33]. Neuroinflammatory pathways include both the innate and the adaptive immune systems of the central nervous system in connection with neurodegeneration. Furthermore, the pathophysiology of NDs can also include neuroinflammation [34,35]. The main component of the innate immune response is the microglia in the central nervous system. Microglia cause morphological changes in response to pathological changes in the nervous system, and activated microglia secrete a variety of inflammatory mediators including cytokines, chemokines, and cytotoxic molecules. These inflammatory mediators allow astrocytes to respond to the reparation and survival of the secondary inflammatory or growth factor repair response [36,37,38]. Mitochondria are the place of oxidative phosphorylation that help to maintain low cytosol Ca2+ concentration [39]. Excessive Ca2+ absorption and ROS development lead to a decline inmitochondrial membrane functionand the opening of mitochondrial pores [40]. Several environmental toxins are identified as complex I inhibitors and cause ND-related characteristics [41,42]. The direct association between mitochondrial dysfunction and PD [43,44] was deduced from a discovery of complex I deficiency in the substantia nigra of patients who had died with PD [43,44], followed by evidence of mitochondrial defects in skeletal muscles, platelets, and lymphoblasts in a proportion of cases [45]. The mitochondrial deficiency within the brain appeared to be confined to the substantia nigra. These mitochondrial functional changes occur early prior to the death of the neuron. In the caspase-independent process, the apoptotic factor is converted into the nucleus and results in fragmented DNA or chromatin condensation [46,47]. As neurodegeneration is associated with multifactorial pathological mechanisms, multiple action mechanisms are a promising strategy in ND prevention and therapy.

3. Neuroprotective Activities of Numerous Natural Products

A number of natural products have been suggested by Srivastava et al. as traditional pharmacological agents for the treatment of NDs [48]. The use of natural products for the treatment of NDs is widely reported in the literature, as they show different neuroprotective activities. Figure 2 summarizes a wide range of possible therapeutic effects of various natural products for combating NDs.

3.1. Luteolin

Luteolin (Lu) is a crystalline yellow flavonoid, common in the plant families Bryophyta, Pteridophyta, Pinophyta, and Magnoliophyta. The food sources of Lu are carrot, onion, celery, olive oil, peppermint, thyme, and oregano [49]. Some Lu molecules have a range of pharmacological properties, including antioxidant, anti-inflammatory, anti-microbial, anti-cancer, and neuroprotective properties [50,51]. These various pharmacological and antioxidant effects are combined with its ability to scavenge oxygen and nitrogen species [52]. A study showed that Lu (20–100 μM) effectively attenuated zinc-induced tau hyperphosphorylation not only through its antioxidant activity, but also through the regulatory mechanisms of the tau phosphatase/kinase system [48]. The decrease in intracellular ROS production increased SOD activity, and the restoration of mitochondrial membrane permeabilization has inhibited caspase-based apoptosis [53]. In addition, the amyloid precursor protein (APP) expression was down-regulated and decreased the secretion of Aβ [54]. In addition, Lu enhanced the nuclear factor erythroid 2-related factor 2 (Nrf2) route and induced the activation of the neuronal cell extracellular signal-regulated kinase (ERK1/2) [55,56]. One of the study reported the concentration of Lu (10–20 μM) increased the neuronal survival, which acts with greater efficacy and equal potency than vitamin E [49].

3.2. Quercetin

Quercetin (QCT) is known as a flavonoid in a wide range of food products, such as capers, apples, tomatoes, pasta, green tea, and black and red wines [48]. QCT is a potent herbal antioxidant and is one of the most common flavonoids in edible plants [57]. One study reported the therapeutic efficacy of QCT in improving learning, memory, and cognitive functions in AD [58]. Pharmacologically, QCT has anti-cancer, anti-viral, anti-inflammatory, and anti-amyloid effects [48]. QCT has been described to induce the gradual removal of end products associated with plasma with a recorded half-life of 11–28 h, enabling the body to generate QCT daily [59]. The risk of neurotoxicity can increase through a rise in the number of QCT aglycons entering the central nervous system parenchyma in liposome preparations or by allowing higher blood–brain barrier (BBB) permeability [60]. QCT has been reported to act as a memory booster in a Zebrafish model with scopolamine-induced impairment of memory, potentially improving cholinergic neurotransmission [57]. Further toxicological studies are, therefore, necessary to investigate the risk/beneficial effects of natural products such as QCT [61]. Aβ-mediated apoptosis in hippocampal cultures was significantly reduced at lower doses of QCT (5–20 μM); however, cytotoxicity was induced at high doses (40 μM) [62]. QCT (10 μM) demonstrated anti-amyloidine effects by inhibiting the formation of Aβ fibril [57]. The ability of QCT to cross the BBB and the quantities of QCT and its metabolites in the brain tissue are crucial considerations for its possible in vivo application. QCT reaches the brain, according to in vitro experiments using BBB models [63,64]. Furthermore, QCT and alpha-tocopherol coadministration has been found to promote QCT transport across the BBB [65]. Dihydroquercetin, also known as taxifolin, is a flavonoid commonly found in onions [66]. Treatment with taxifolin prevented spatial memory defects caused by oligomeric Aβ in the wild-type mice hippocampus [67]. In taxifolin-treated cerebral amyloid angiopathy mice, higher blood Aβ levels have been detected, suggesting that Aβ clearance from the brain to bloodwas made easier [68].

3.3. Resveratrol

Resveratrol (RSV) belongs toa class of polyphenolic stilbene compounds [69]. RSV is one of the most important red wine flavonoids in grapes, nuts, and other fruits [70]. About 12.5% of participants experienced headaches in the short dose study of RSV, but showed no serious adverse effects [71]. Many studies have reported cardiovascular, anti-cancer, anti-viral, blood-glucose-decreasing, and side effects of RSV [72,73,74]. RSV (10 and 20 mg/kg) primarily works by scavenging ROS as a strong antioxidant by enhancing glutathione (GSH) [75]. The loaded lipid core RSV nanocapsules are elevated compared with free RSV in brain tissue [76]. The gastrointestinal lumen absorbs RSV well, but due to its rapid metabolism and removal, it has poor bioaccessibility [77]. In different forms, the binding of RSV (50 μM) to Aβ was greater, but it was more strongly attached to monomeric Aβ 1–40 than to its fibrillary form [78]. By induction of non-amyloidogenic APP cleavage, RSV reduced Aβ and increased the clearance of Aβ [79]. RSV (100 and 200 μM) can also inhibit C-reactive protein and ERK1/2 mitogen-activated protein kinase (MAPK)[80]. RSV (2.5–40 mg/mL) inhibited the inflammatory response to lipopolysaccharide by reducing inflammatory factors, such as nitric oxide, tumor necrosis factor-α (TNF-α), interleukin (IL)-1β, and IL-6 of astrocytes [81]. Nuclear factor-kappa B (NF-κB) elimination led to a decreased downstream TNF-α and IL-6 levels [82]. A meta-analysis showed that RSV significantly decreased Profile of Mood States (POMS) including vigor and fatigue but had no significant effect on memory and cognitive performance [83]. However, other studies have shown that the BBB plays an important role in Aβ clearance and that its breakdown can result in ineffective clearance [84]. RSV increased claudin-5 expression and decreased the receptor for advanced glycation end products in vivo [85], protecting the BBB integrity [86].

3.4. Apigenin

Apigenin (AP) belongs to a subgroup of flavonoids, flavones, based on a skeleton of 2-phenylchromen-4-one (2-phenyl-1-benzopyran-4-one) [87]. To date, very little evidence suggests that AP in a normal diet promotes in vivo adverse metabolic reactions. AP has anti-inflammatory, antioxidant, and anti-cancer characteristics [88]. It is also a strong inhibitor of the enzyme metabolizing several prescription drugs in the body, cytochrome P450 [89]. AP is a highly soluble and intestinally permeable flavonoid. Different transport processes in the intestine can well absorb AP; however, the duodenum is the main absorption site [90]. It also functions as a cell growth, anti-carcinogenic, and enzyme inhibitor, as well as antigenotoxic, anti-inflammatory, and free radical scavenging [91]. In addition, in a double transgenic mouse model of AD (APP/PS1), a review of the neuroprotective potential of AP suggested that apigenin could enhance AD-associated memory impairment, decrease the load of Aβ plaque, and inhibit oxidative stress [92].

3.5. Genistein

Genistein is one of the most commonly known isoflavone found in numerous soy products and has been investigated for its antioxidant, anti-inflammatory, and proapoptotic properties; estrogen receptor affinity; protein tyrosine kinase (PTK) inhibition; and other cellular and physiological functions [93]. Current evidence strongly indicates that soy isoflavones protect against a variety of chronic conditions including atherosclerosis, postmenopausal estrogen deficiency, and hormonally based breast or prostate cancer [94]. Recently, some researchers have found the neuroprotective activity of genistein. A study reported that genistein (100 μM) was found to be effective against toxicity induced by the Aβ31-35 peptide in primary neuronal cells obtained from newborn Wistar rats [95]. It has been confirmed that genistein, a phytoestrogen able to cross the BBB, has antioxidants from ultraviolet light and chemical insults. Another research reported that in cultured hippocampal neurons, genistein has a neuroprotective effect against Aβ25-35-induced apoptosis [96]. Exposure to aged Aβ25-35 for 24 h has been shown to double the DCF fluorescence strength compared with controls for 24 h [97]. Emerging evidence indicates that estrogen and estrogen-like chemicals have beneficial effects on ND, especially PD. Interestingly, genistein exhibited a preventive effect on neuronal degeneration caused by increased oxidative stress [98]. In addition, genistein can cross the BBB [99], and it has proven to be safe for a long time (over 1 year) in the clinical trial at concentrations up to 150 mg/kg/day.

3.6. Hesperidin

Flavanone-glycosides rich in citrus fruit, lemon, sweet orange, and grapes are also called hesperidin (C28H34O15) [100]. Hesperidin administration for 16 wks helped boost learning and memory function by increasing the recognition index in the transgenic mouse model of APPswe/PS1dEE [101]. It corrects mitochondrial disorders caused by Aβ by lowering levels of malondialdehyde and hydrogen peroxide and restoring GSH depletion and total antioxidant ability (T-AOC). A protein kinase that has a prominent role in mitochondria and AD functions is glycogen synthase kinase-3β (GSK-3β). It has an important impact on the protein tau hyperphosphorylation and the mitochondrial target [102]. Increasing oxidative damage triggers the activation of this protein kinase. By inhibiting the restoration of this kinase, hesperidin theoretically rescued cognitive deficits and showed mitochondrial neuroprotective effects. It was the potential mechanism by which hesperidin lowered the Aβ1-40 level [100]. Hesperidin also inhibited learning and memory impairments resulting from aluminum chloride (AlCl3)-induced AD, functioning as an acetylcholinesterase inhibitor. In the rat hippocampus and brain cortex, hesperidin attenuated APP expression through the NF-κB-dependent pathway and suppressed Aβ1-40 and β-and γ-secretase levels [49,103]. The neuroprotective role of hesperidin was reported in the signals of up-regulating B-cell lymphoma 2 (Bcl2) and down-regulating Bcl-2-associated X protein (Bax) [104,105,106]. In addition, hesperidin has been reported to have neuroprotective effects in many neurological disorders, such as cerebral ischemia, HD, and PD, at 50 and 100 mg/kg oral doses [107]. The hesperidin of citrus flavonoid has neuroprotective effects andmay pass through the BBB. Hesperidin inhibits the release of glutamate and exercises an excitotoxic neuroprotection in rat hippocampus with kainic acid [108].

3.7. Uncaria Rhynchophylla

The herb Uncaria rhynchophylla, part of the Rubiaceae family, is used in traditional Chinese medicine. Uncaria rhynchophylla extract is made up of alkaloids, rhinchophylline, hirsutine, hirsuteine, corynanthine, corynoxine, and dihydrocorynantheine [109,110]. The most widely studied and named neuroprotective compositions among the alkaloids are rhinchophylline and isorhynchophylline [111,112]. In addition, the neuroprotection effect of Uncaria rhynchophylla has been reported in an experimental PD model [113]. Shim et al. documented that Uncaria rhynchophylla reduced neuronal cell death and ROS production, restored GSH levels in PC12 cells in case of toxicity of caspase-3 and 6-hydroxydopamine (6-OHDA) cells, and reduced the neuronal loss in the substantia nigra dopaminergic rats induced by 6-OHDA [113,114]. In the 1-methyl-4-phenylpyridinium (MPP+) induced SH-SY5Y and 1-methyl-4-phenyl-1, 2, 3, 6-tetrahydropyridine (MPTP) mouse cell models, Uncaria rhynchophylla has been found to improve cell viability, attenuate dopaminergic neuronal significant nigra and striatum lowering, and inhibit heat-shock protein 90 and autophagy [115]. All these results, taken together, indicate that Uncaria rhynchophylla demonstrates neuroprotective activity through multiple mechanisms of neuronal defense against damage, which may be attributable to the beneficial combined action of the active compounds in Uncaria rhynchophylla. These antioxidant compounds inhibit the anti-inflammatory effect by inflammatory mediation and the anti-apoptotic effect, modulating the event and preventing the activation of the caspase and decreasing ROS generation and improving the antioxidant protective mechanism. Rhynchophylline is an alkaloid found in certain Uncaria species (Rubiaceae). However, recent studies revealed that isorhynchophylline can easily pass the BBB. These observations suggested that isorhynchophylline may be an anti-inflammatory substance used to treat NDs [116,117].

3.8. Marine Macroalgae

Marine macroalgae are plant-like organisms, typically referred to as seaweed, that commonly live in coastal areas. The three groups can be categorized as brown (Phaeophyceae), red (Rhodophyceae), and green algae (Chlorophyceae) [118]. Phenolic compounds, proteins, peptides, pigments, amino acids, and phenols are also found in a variety of bioactive materials [119]. Numerous studies find that algae and bioactive compounds of various algae have a health impact [118,120,121]. In addition, Pangestuti et al. showed that carotenoids have a high radical scavenging function and are present in marine algae as a major antioxidant [122]. Furthermore, another study found that marine extracts increase cell viability, decrease oxidative stress, have a healthy mitochondrial membrane potential, and decrease caspase-3 activities. This indicates the neuroprotective effects and the antioxidant properties of these algae [123]. Silva et al. suggested the possibility of mediating this neuroprotective action with antioxidant compounds in algae extracts [123]. However, the researchers’ concern about the potential use for pharmaceuticals, particularly when new drug delivery systems are being developed, recently attracted their attention to marine sulfated polysaccharides [124,125]. The biological activities of sulfated polysaccharides have been identified in various studies [126,127]. In the meantime, Undaria pinnatifida fucoidan improved cell viability, prevented apoptosis via inhibition of activation of caspase-3, and enhanced dense antioxidant systems in Aβ (25–35), SOD activity, and GSH materials in PC12 cells with neurotoxicity [128]. The fucoidans have a reduced aggregation of Aβ (1–42), decreased cytotoxicity (1–42), and PC12 hydrogen peroxide caused by Aβ, decreased Aβ-induced apoptosis (1–42), and improved the role of neuritis outgrowth [129,130]. Moreover, the possibility of developing marine algae components as neuroprotective agents has not been investigated because of the BBB.

3.9. Cyanobacteria

Cyanobacteria are prokaryotic, photosynthetic, self-producing species that are closely related to bacteria and are commonly referred to as blue-green algae. They are members of the Oscillatoriaceae family. Researchers have been very attentive to their potential pharmacological properties and advantages for various medical conditions [120,131]. Spirulina platensis is a multicellular planktonic, alkaliphilic cyanobacterium. It has been widely studied and recognized for its proper nutritional components. Subsequently, it may protect itself against dopaminergic neuronal loss triggered by MPTP in substantia nigra. Spirulina platensis has anti-inflammatory and antioxidant properties that help it defend against PD caused by 6-OHDA [132,133]. However, evidence suggests that polysaccharides derived from Spirulina platensis have an antioxidant effect on dopaminergic neurons and dopamine levels, rather than inhibition of monoamine oxidase B [134]. These findings showed that Spirulina maxima extract improved cognitive impairment by inhibiting Aβ accumulation [135]. In addition, the neuroprotective role of Spirulina maxima (Sp.) against MPTP neurotoxicity, used as a model of PD [122]. Other studies have shown that Spirulina maxima extract has protected against memory damage caused by scopolamine in mice [136,137]. These results show that, via antioxidant activity, Spirulina maxima exert its neuroprotective impact [136]. In addition, oral administration of c-phycocyanin, a component of Spirulina, has an effect in the hippocampus, because it crosses the BBB [138]. These studies have shown collectively that cytoprotective activity against neurodegeneration is demonstrated by different mechanisms of action, primarily by antioxidants.

4. Role of Other Natural Products in Neurodegenerative Diseases

NDs exhibit some common characteristics despite specific clinical and etiopathogenic differences, such as irregular protein deposition, abnormal cellular transports, mitochondrial deficits, inflammation, intracellular Ca2+ overload, unregulated ROS generation, and excitotoxicity [4,139]. In the pathogenesis of all essential NDs, reactive astroglia and/or microglia are also involved [140,141]. Several natural substances have been suggested for treating NDs to complete and/or help conventional pharmacological agents [4]. Their use on NDs is commonly identified as a consequence of several different neuroprotective activities reported in the literature [142,143,144]. The main objectives include mitochondrial dysfunction, inflammation, oxidative stress, and protein malfunction among the natural products [145,146,147]. Some animal products, such as omega-3 fatty acids, inhibit cell toxicity and have anti-inflammatory effects in the treatment of AD [148]. Plant-based products, such as lunasin, polyphenols, alkaloids, and tannins, are possible therapeutic candidates for AD [149]. Resveratrol and flavonoids appear to be dietary additives that have obvious neuroprotective and other beneficial effects on human cognitive disability [69,150]. Although natural products can be extracted from different biological sources, it is not trivial to turn them into therapies. The challenges can include concerns about their stability and neuro-availability, difficulties in properly defining and quantifying the active principle, and, lastly, difficulties in organizing large-scale clinical trials to evaluate these complex products [151]. The capacity to defend against neurodegeneration has been evaluated in several differentnatural products. Table 1 and Table 2 provide a description of natural products and their bioactive compounds with various neuroprotective functions, depending on the disease being treated. Natural products and their bioactive substances with neuroprotective function in the treatment of AD are represented in Table 1. Similarly, PD treatment currently includes medicines such as Levodopa, primarily catalytically converted into dopamine by dopa decarboxylase in the brain, resulting in its therapeutic effects [152,153]. There is evidence that correlates neuronal mitochondrial dysfunction with the pathogenesis of PD [154,155]. However, this dysfunction is associated with the abnormal accumulation of α-synuclein, which causes an alteration of normal mitochondrial function, leading to neuronal degeneration and strong oxidative stress [156,157]. In addition, the presence of neuroinflammation is another peculiar characteristic of PD, which plays a significant role in the development of the disease. However, the inflammation depends also on the impaired energy metabolism at the level of the mitochondria impairment that causes the activation of the microglia and the relative generation of a plethora of pro-inflammatory mediators, including prostaglandins, cytokines, chemokines, complement, proteinases, ROS, and RNS [158]. Moreover, most patients with PD also have non-motor symptoms, including disorders of the sleep–wake cycle regulation, cognitive impairment disorders of mood and affect, autonomic dysfunction, as well as sensory impairmentand pain. Recently, the management of age-related diseases, such as PD, has been associated with consumption of functional food or food supplements. Certainly, a healthy diet rich in foods containing antioxidants, vitamins, and minerals or the use of food supplements can help to reduce the symptoms of PD and the related pathological mechanisms [158]. Mucunapruriens belongs to the family Leguminosae and is a twiner with trifoliate leaves, purple flowers, and pods covered with hairs. Seeds from Mucuna pruriens (Atmagupta) have been described as a useful therapeutic agent in different diseases of the human nervous and reproductive system, including PD in the ancient Indian medical system, Ayurveda [158]. Mucuna pruriens exhibited twice the anti-parkinsonian activity compared with synthetic levodopa, suggesting that Mucuna pruriens may contain unidentified antiparkinsonian compounds in addition to levodopa, or that it may have adjuvants that enhance the efficacy of levodopa [159]. Another therapy involves anticholinergic drugs that can block the excitability of cholinergic nerves by striatal cholinergic receptors; it has also been shown that they can suppress dopamine reuptake to increase the activity of dopaminergic neurons [160]. Natural products and their bioactive substances with neuroprotective function in the treatment of PD are shown in Table 2. The therapeutic potential of medicinal plants has been studied and evaluated in scientific circles. Numerous medicinal plants extract used in the clinical trial and their outcomes are shown in Table 3. In conclusion, as complementary or integrative therapeutic agents against AD, PD, and other NDs, natural products have recently gained greater attention [161].

5. Limitations, Future Prospects, and Challenges

The capacity of neuroprotection and the development of therapeutic products and tools, including isolated natural compounds, against various NDs have been naturally developing. Despite the promising neuroprotective activity in pre-clinical settings, the translation of promising preclinical investigations to clinical use has proven difficult because human clinical studies of neurodegenerative disorders have no favorable findings. Natural products and isolated natural compounds face several challenges and weaknesses that can compromise their therapeutic efficacy, including poor bioavailability and decreased water solubility, physical and chemical instabilities, rapid metabolism, and BBB crossing. These reviews of the literature provide more details [199,200,201]. However, numerous natural compounds, including resveratrol [202] and curcumin [203,204], have been reported to have low bioavailability and limited stability due to degradation or transformation into inactive derivatives [205,206]. As a result, their efficacy is reduced. In addition, the BBB prevents access of natural compounds to the brain, thus prevents them to reach their action site. This limits their distribution to the brain tissue and results in low bioavailability [207]. Nanotechnology and nanocarriers can help improve therapeutic responses and effectiveness in the delivery of natural products and their isolated compounds, which will help solve these problems [208,209]. Nanoparticles may be used in the delivery system to increase the bioavailability of natural products and their compounds. Polymeric nanoparticles, nanogels, rigid lipid nanoparticles, crystalline nanoparticles, micelles, and dendrimer complexes are the most commonly used nanoparticles [210,211]. Several studies have been published on the use of natural nanoparticles with thesecompounds, such as epigallocatechin 3-gallate for treating AD [212], rosemary acid for HD [213], curcumin for brain disease [214].

6. Concluding Remarks

Therapeutic potential for natural products and natural bioactive compounds to be neuroprotective has been supported by various research studies. Natural products and important bioactive compounds are needed to prevent and treat various NDs without causing harmful adverse effects. Since several functional pathways are found in neurodegenerative pathologies, ND prevention and treatment approaches have an important role to play. For natural products and bioactive substances, it is preferable to use various modes of action to display neuroprotective effects. Furthermore, the ability of natural products and their bioactive compounds to cross the BBB is essential for neuroprotective activity. It is important to develop new methods and techniques, such as nanotechnology, for the delivery of natural compounds and drugs in order to enhance the role of natural products and bioactive compounds in ND prevention and therapeutic fields, in order to promote access to the brain of neuroprotective products.

Author Contributions

Conceptualization, K.-J.L., M.H.R. and J.B.; writing—original draft preparation, M.H.R.; writing—review and editing, J.B. and A.F.; preparation of the tables and figures, R.A., S.S., S.H.G. and T.T.T.; visualization, Y.J.J., C.-S.K.; supervision, K.-J.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available within the article (tables and figures).

Acknowledgments

The authors would like to thank Editage (Cactus communication Korea Ltd.) for language editing.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

6-OHDA6-hydroxydopamine
ADAlzheimer’s disease
AlCl3Aluminum chloride
APApigenin
APPAmyloid precursor protein
Amyloid-beta
BaxBcl-2-associated X proteins
BBBBlood-brain barrier
Bcl2B-cell lymphoma 2
CAGCytosine-adenine-guanine
DHMDihydromyricetin
DNADeoxyribonucleic acid
ERKExtracellular signal-regulated kinase
GSHGlutathione
GSK-3βGlycogen synthase kinase-3β
HDHuntington’s disease
ILInterleukin
L-DOPALevodopa
LuLuteolin
MAPKMitogen-activated protein kinase
MPP+1-methyl-4-phenylpyridinium
MPTP1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine
NDsNeurodegenerative diseases
NF-κBNuclear factor-kappa B
Nrf2Nuclear factor erythroid 2-related factor 2
PDParkinson’s disease
POMSProfile of Mood States
PTKProtein tyrosine kinase
QCTQuercetin
ROSReactive oxygen species
RSVResveratrol
SOD-1Superoxide dismutase-1
T-AOCTotal antioxidant ability
TNF-αTumor necrosis factor-α

References

  1. Rahman, M.H.; Bajgai, J.; Fadriquela, A.; Sharma, S.; Trinh Thi, T.; Akter, R.; Goh, S.H.; Kim, C.S.; Lee, K.J. Redox effects of molecular hydrogen and its therapeutic efficacy in the treatment of neurodegenerative diseases. Processes 2021, 9, 308. [Google Scholar] [CrossRef]
  2. Salvadores, N.; Court, F.A. The necroptosis pathway and its role in age-related neurodegenerative diseases: Will it open up new therapeutic avenues in the next decade? Expert Opin. Ther. Targets 2020, 24, 679–693. [Google Scholar] [CrossRef]
  3. Yildiz-Unal, A.; Korulu, S.; Karabay, A. Neuroprotective strategies against calpain-mediated neurodegeneration. Neuropsychiatr. Dis. Treat. 2015, 11, 297–310. [Google Scholar] [CrossRef] [Green Version]
  4. Di Paolo, M.; Papi, L.; Gori, F.; Turillazzi, E. Natural products in neurodegenerative diseases: A great promise but an ethical challenge. Int. J. Mol. Sci. 2019, 20, 5170. [Google Scholar] [CrossRef] [Green Version]
  5. Thirupathi, A.; Chang, Y.Z. Brain iron metabolism and cns diseases. Adv. Exp. Med. Biol. 2019, 1173, 1–19. [Google Scholar] [CrossRef]
  6. Troncoso-Escudero, P.; Sepulveda, D.; Perez-Arancibia, R.; Parra, A.V.; Arcos, J.; Grunenwald, F.; Vidal, R.L. On the right track to treat movement disorders: Promising therapeutic approaches for parkinson’s and huntington’s disease. Front. Aging Neurosci. 2020, 12, 571185. [Google Scholar] [CrossRef]
  7. Harischandra, D.S.; Ghaisas, S.; Zenitsky, G.; Jin, H.; Kanthasamy, A.; Anantharam, V.; Kanthasamy, A.G. Manganese-induced neurotoxicity: New insights into the triad of protein misfolding, mitochondrial impairment, and neuroinflammation. Front. Neurosci. 2019, 13, 654. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Fu, H.; Hardy, J.; Duff, K.E. Selective vulnerability in neurodegenerative diseases. Nat. Neurosci. 2018, 21, 1350–1358. [Google Scholar] [CrossRef] [PubMed]
  9. Gan, L.; Cookson, M.R.; Petrucelli, L.; La Spada, A.R. Converging pathways in neurodegeneration, from genetics to mechanisms. Nat. Neurosci. 2018, 21, 1300–1309. [Google Scholar] [CrossRef]
  10. Singh, A.; Kukreti, R.; Saso, L.; Kukreti, S. Oxidative stress: A key modulator in neurodegenerative diseases. Molecules 2019, 24, 1583. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Fan, J.; Dawson, T.M.; Dawson, V.L. Cell death mechanisms of neurodegeneration. Adv. Neurobiol. 2017, 15, 403–425. [Google Scholar] [CrossRef]
  12. Cory, H.; Passarelli, S.; Szeto, J.; Tamez, M.; Mattei, J. The role of polyphenols in human health and food systems: A mini-review. Front. Nutr. 2018, 5, 87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Li, X.; Chu, S.; Liu, Y.; Chen, N. Neuroprotective effects of anthraquinones from rhubarb in central nervous system diseases. Evid. Based Complement. Alternat. Med. 2019, 2019, 3790728. [Google Scholar] [CrossRef] [Green Version]
  14. Lutz, M.; Fuentes, E.; Avila, F.; Alarcon, M.; Palomo, I. Roles of phenolic compounds in the reduction of risk factors of cardiovascular diseases. Molecules 2019, 24, 366. [Google Scholar] [CrossRef] [Green Version]
  15. Pham, L.; Wright, D.K.; O’Brien, W.T.; Bain, J.; Huang, C.; Sun, M.; Casillas-Espinosa, P.M.; Shah, A.D.; Schittenhelm, R.B.; Sobey, C.G.; et al. Behavioral, axonal, and proteomic alterations following repeated mild traumatic brain injury: Novel insights using a clinically relevant rat model. Neurobiol. Dis. 2021, 148, 105151. [Google Scholar] [CrossRef] [PubMed]
  16. Bui, T.T.; Nguyen, T.H. Natural product for the treatment of alzheimer’s disease. J. Basic Clin. Physiol. Pharmacol. 2017, 28, 413–423. [Google Scholar] [CrossRef] [PubMed]
  17. Babaei, F.; Mirzababaei, M.; Nassiri-Asl, M. Quercetin in food: Possible mechanisms of its effect on memory. J. Food Sci. 2018, 83, 2280–2287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Obeso, J.A.; Stamelou, M.; Goetz, C.G.; Poewe, W.; Lang, A.E.; Weintraub, D.; Burn, D.; Halliday, G.M.; Bezard, E.; Przedborski, S.; et al. Past, present, and future of parkinson’s disease: A special essay on the 200th anniversary of the shaking palsy. Mov. Disord. 2017, 32, 1264–1310. [Google Scholar] [CrossRef]
  19. Grieco, M.; Giorgi, A.; Gentile, M.C.; d’Erme, M.; Morano, S.; Maras, B.; Filardi, T. Glucagon-like peptide-1: A focus on neurodegenerative diseases. Front. Neurosci. 2019, 13, 1112. [Google Scholar] [CrossRef] [Green Version]
  20. Shahmoradian, S.H.; Lewis, A.J.; Genoud, C.; Hench, J.; Moors, T.E.; Navarro, P.P.; Castano-Diez, D.; Schweighauser, G.; Graff-Meyer, A.; Goldie, K.N.; et al. Lewy pathology in parkinson’s disease consists of crowded organelles and lipid membranes. Nat. Neurosci. 2019, 22, 1099–1109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Hunn, B.H.; Cragg, S.J.; Bolam, J.P.; Spillantini, M.G.; Wade-Martins, R. Impaired intracellular trafficking defines early parkinson’s disease. Trends Neurosci. 2015, 38, 178–188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Hilton, J.B.; White, A.R.; Crouch, P.J. Metal-deficient sod1 in amyotrophic lateral sclerosis. J. Mol. Med. 2015, 93, 481–487. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Sirangelo, I.; Iannuzzi, C. The role of metal binding in the amyotrophic lateral sclerosis-related aggregation of copper-zinc superoxide dismutase. Molecules 2017, 22, 1429. [Google Scholar] [CrossRef] [Green Version]
  24. Bostan, A.C.; Strick, P.L. The basal ganglia and the cerebellum: Nodes in an integrated network. Nat. Rev. Neurosci. 2018, 19, 338–350. [Google Scholar] [CrossRef] [PubMed]
  25. Gil, J.M.; Rego, A.C. Mechanisms of neurodegeneration in huntington’s disease. Eur. J. Neurosci. 2008, 27, 2803–2820. [Google Scholar] [CrossRef] [Green Version]
  26. Hsu, Y.T.; Chang, Y.G.; Chern, Y. Insights into gabaaergic system alteration in huntington’s disease. Open Biol. 2018, 8, 180165. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Gardiner, S.L.; van Belzen, M.J.; Boogaard, M.W.; van Roon-Mom, W.M.; Rozing, M.P.; van Hemert, A.M.; Smit, J.H.; Beekman, A.T.; van Grootheest, G.; Schoevers, R.A.; et al. Huntingtin gene repeat size variations affect risk of lifetime depression. Transl. Psychiatry 2017, 7, 1277. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Bhat, A.H.; Dar, K.B.; Anees, S.; Zargar, M.A.; Masood, A.; Sofi, M.A.; Ganie, S.A. Oxidative stress, mitochondrial dysfunction and neurodegenerative diseases; a mechanistic insight. Biomed. Pharmacother. 2015, 74, 101–110. [Google Scholar] [CrossRef]
  29. Golpich, M.; Amini, E.; Mohamed, Z.; Azman Ali, R.; Mohamed Ibrahim, N.; Ahmadiani, A. Mitochondrial dysfunction and biogenesis in neurodegenerative diseases: Pathogenesis and treatment. CNS Neurosci. Ther. 2017, 23, 5–22. [Google Scholar] [CrossRef] [PubMed]
  30. Ullah, R.; Khan, M.; Shah, S.A.; Saeed, K.; Kim, M.O. Natural antioxidant anthocyanins-a hidden therapeutic candidate in metabolic disorders with major focus in neurodegeneration. Nutrients 2019, 11, 1195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Nita, M.; Grzybowski, A. The role of the reactive oxygen species and oxidative stress in the pathomechanism of the age-related ocular diseases and other pathologies of the anterior and posterior eye segments in adults. Oxid. Med. Cell. Longev. 2016, 2016, 3164734. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Buendia, I.; Michalska, P.; Navarro, E.; Gameiro, I.; Egea, J.; Leon, R. Nrf2-are pathway: An emerging target against oxidative stress and neuroinflammation in neurodegenerative diseases. Pharmacol. Ther. 2016, 157, 84–104. [Google Scholar] [CrossRef]
  33. Gelders, G.; Baekelandt, V.; Van der Perren, A. Linking neuroinflammation and neurodegeneration in parkinson’s disease. J. Immunol. Res. 2018, 2018, 4784268. [Google Scholar] [CrossRef] [Green Version]
  34. Leszek, J.; Barreto, G.E.; Gasiorowski, K.; Koutsouraki, E.; Avila-Rodrigues, M.; Aliev, G. Inflammatory mechanisms and oxidative stress as key factors responsible for progression of neurodegeneration: Role of brain innate immune system. CNS Neurol. Disord. Drug. Targets 2016, 15, 329–336. [Google Scholar] [CrossRef] [PubMed]
  35. Stephenson, J.; Nutma, E.; van der Valk, P.; Amor, S. Inflammation in cns neurodegenerative diseases. Immunology 2018, 154, 204–219. [Google Scholar] [CrossRef] [Green Version]
  36. Thurgur, H.; Pinteaux, E. Microglia in the neurovascular unit: Blood-brain barrier-microglia interactions after central nervous system disorders. Neuroscience 2019, 405, 55–67. [Google Scholar] [CrossRef]
  37. Voet, S.; Prinz, M.; van Loo, G. Microglia in central nervous system inflammation and multiple sclerosis pathology. Trends Mol. Med. 2019, 25, 112–123. [Google Scholar] [CrossRef]
  38. Kabba, J.A.; Xu, Y.; Christian, H.; Ruan, W.; Chenai, K.; Xiang, Y.; Zhang, L.; Saavedra, J.M.; Pang, T. Microglia: Housekeeper of the central nervous system. Cell. Mol. Neurobiol. 2018, 38, 53–71. [Google Scholar] [CrossRef]
  39. Cardenas, C.; Lovy, A.; Silva-Pavez, E.; Urra, F.; Mizzoni, C.; Ahumada-Castro, U.; Bustos, G.; Jana, F.; Cruz, P.; Farias, P.; et al. Cancer cells with defective oxidative phosphorylation require endoplasmic reticulum-to-mitochondria Ca2+ transfer for survival. Sci. Signal. 2020, 13, eaay1212. [Google Scholar] [CrossRef]
  40. Panov, A.; Dikalov, S.; Shalbuyeva, N.; Hemendinger, R.; Greenamyre, J.T.; Rosenfeld, J. Species- and tissue-specific relationships between mitochondrial permeability transition and generation of ros in brain and liver mitochondria of rats and mice. Am. J. Physiol. Cell Physiol. 2007, 292, C708–C718. [Google Scholar] [CrossRef] [Green Version]
  41. Rao, V.K.; Carlson, E.A.; Yan, S.S. Mitochondrial permeability transition pore is a potential drug target for neurodegeneration. Biochim. Biophys. Acta Mol. Basis Dis. 2014, 1842, 1267–1272. [Google Scholar] [CrossRef] [Green Version]
  42. Gadd, M.E.; Broekemeier, K.M.; Crouser, E.D.; Kumar, J.; Graff, G.; Pfeiffer, D.R. Mitochondrial ipla2 activity modulates the release of cytochrome c from mitochondria and influences the permeability transition. J. Biol. Chem. 2006, 281, 6931–6939. [Google Scholar] [CrossRef] [Green Version]
  43. Osellame, L.D.; Rahim, A.A.; Hargreaves, I.P.; Gegg, M.E.; Richard-Londt, A.; Brandner, S.; Waddington, S.N.; Schapira, A.H.; Duchen, M.R. Mitochondria and quality control defects in a mouse model of gaucher disease-links to parkinson’s disease. Cell Metab. 2013, 17, 941–953. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Poewe, W.; Seppi, K.; Tanner, C.M.; Halliday, G.M.; Brundin, P.; Volkmann, J.; Schrag, A.E.; Lang, A.E. Parkinson disease. Nat. Rev. Dis. Primers 2017, 3, 17013. [Google Scholar] [CrossRef] [PubMed]
  45. Kent, J.A.; Patel, V.; Varela, N.A. Gender disparities in health care. Mt. Sinai J. Med. 2012, 79, 555–559. [Google Scholar] [CrossRef] [PubMed]
  46. Bano, D.; Dinsdale, D.; Cabrera-Socorro, A.; Maida, S.; Lambacher, N.; McColl, B.; Ferrando-May, E.; Hengartner, M.O.; Nicotera, P. Alteration of the nuclear pore complex in Ca2+-mediated cell death. Cell. Death. Differ. 2010, 17, 119–133. [Google Scholar] [CrossRef]
  47. Verma, D.K.; Gupta, S.; Biswas, J.; Joshi, N.; Sivarama Raju, K.; Wahajuddin, M.; Singh, S. Metabolic enhancer piracetam attenuates the translocation of mitochondrion-specific proteins of caspase-independent pathway, poly [adp-ribose] polymerase 1 up-regulation and oxidative DNA fragmentation. Neurotox. Res. 2018, 34, 198–219. [Google Scholar] [CrossRef]
  48. Srivastava, P.; Yadav, R.S. Efficacy of natural compounds in neurodegenerative disorders. In The Benefits of Natural Products for Neurodegenerative Diseases; Springer: Cham, Switzerland, 2016; pp. 107–123. [Google Scholar] [CrossRef] [Green Version]
  49. Akter, R.; Chowdhury, M.; Rahman, A.; Rahman, H. Flavonoids and polyphenolic compounds as potential talented agents for the treatment of alzheimer’s disease with their antioxidant activities. Curr. Pharm. Des. 2020, 27, 345–356. [Google Scholar] [CrossRef]
  50. Wang, B.; Lu, Y.; Wang, R.; Liu, S.; Hu, X.; Wang, H. Transport and metabolic profiling studies of amentoflavone in caco-2 cells by uhplc-esi-ms/ms and uhplc-esi-q-tof-ms/ms. J. Pharm. Biomed. Anal. 2020, 189, 113441. [Google Scholar] [CrossRef]
  51. Williamson, G.; Kay, C.D.; Crozier, A. The bioavailability, transport, and bioactivity of dietary flavonoids: A review from a historical perspective. Compr. Rev. Food Sci. Food Saf. 2018, 17, 1054–1112. [Google Scholar] [CrossRef] [Green Version]
  52. Nabavi, S.F.; Braidy, N.; Gortzi, O.; Sobarzo-Sanchez, E.; Daglia, M.; Skalicka-Wozniak, K.; Nabavi, S.M. Luteolin as an anti-inflammatory and neuroprotective agent: A brief review. Brain Res. Bull. 2015, 119, 1–11. [Google Scholar] [CrossRef] [PubMed]
  53. Yulyana, Y.; Tovmasyan, A.; Ho, I.A.; Sia, K.C.; Newman, J.P.; Ng, W.H.; Guo, C.M.; Hui, K.M.; Batinic-Haberle, I.; Lam, P.Y. Redox-active Mn porphyrin-based potent SOD mimic, MnTnBuOE-2-PyP5+, enhances carbenoxolone-mediated TRAIL-induced apoptosis in glioblastoma multiforme. Stem. Cell Rev. Rep. 2016, 12, 140–155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Zheng, H.; Koo, E.H. The amyloid precursor protein: Beyond amyloid. Mol. Neurodegener. 2006, 1, 11–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Uddin, M.S.; Hossain, M.F.; Mamun, A.A.; Shah, M.A.; Hasana, S.; Bulbul, I.J.; Sarwar, M.S.; Mansouri, R.A.; Ashraf, G.M.; Rauf, A.; et al. Exploring the multimodal role of phytochemicals in the modulation of cellular signaling pathways to combat age-related neurodegeneration. Sci. Total Environ. 2020, 725, 138313. [Google Scholar] [CrossRef] [PubMed]
  56. 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]
  57. Khan, H.; Ullah, H.; Aschner, M.; Cheang, W.S.; Akkol, E.K. Neuroprotective effects of quercetin in alzheimer’s disease. Biomolecules 2020, 10, 59. [Google Scholar] [CrossRef] [Green Version]
  58. Batiha, G.E.; Beshbishy, A.M.; Ikram, M.; Mulla, Z.S.; El-Hack, M.E.; Taha, A.E.; Algammal, A.M.; Elewa, Y.H. The pharmacological activity, biochemical properties, and pharmacokinetics of the major natural polyphenolic flavonoid: Quercetin. Foods 2020, 9, 374. [Google Scholar] [CrossRef] [Green Version]
  59. Schultke, E. The Flavonoid Quercetin and Its Potential As Neuroprotectant in the Therapy of Acute traumatic CNS Injury: An Experimental Study. Ph.D. Thesis, University of Saskatchewan, Saskatoon, SK, Canada, 2004. [Google Scholar]
  60. Jurcau, A. The role of natural antioxidants in the prevention of dementia-where do we stand and future perspectives. Nutrients 2021, 13, 282. [Google Scholar] [CrossRef]
  61. Kumar, G.P.; Anilakumar, K.R.; Naveen, S. Phytochemicals having neuroprotective properties from dietary sources and medicinal herbs. Pharmacogn. J. 2015, 7, 1–17. [Google Scholar] [CrossRef] [Green Version]
  62. Costa, L.G.; Garrick, J.M.; Roque, P.J.; Pellacani, C. Mechanisms of neuroprotection by quercetin: Counteracting oxidative stress and more. Oxid. Med. Cell. Longev. 2016, 2016, 2986796. [Google Scholar] [CrossRef] [Green Version]
  63. Faria, A.; Pestana, D.; Teixeira, D.; Azevedo, J.; Freitas, V.; Mateus, N.; Calhau, C. Flavonoid transport across RBE4 cells: A blood-brain barrier model. Cell. Mol. Biol. Lett. 2010, 1, 234–241. [Google Scholar] [CrossRef]
  64. Ishisaka, A.; Ichikawa, S.; Sakakibara, H.; Piskula, M.K.; Nakamura, T.; Kato, Y.; Ito, M.; Miyamoto, K.I.; Tsuji, A.; Kawai, Y.; et al. Accumulation of orally administered quercetin in brain tissue and its antioxidative effects in rats. Free Radic. Biol.Med. 2011, 51, 1329–1336. [Google Scholar] [CrossRef]
  65. Ferri, P.; Angelino, D.; Gennari, L.; Benedetti, S.; Ambrogini, P.; Del Grande, P.; Ninfali, P. Enhancement of flavonoid ability to cross the blood–brain barrier of rats by co-administration with α-tocopherol. Food Funct. 2015, 6, 394–400. [Google Scholar] [CrossRef]
  66. Yang, P.; Xu, F.; Li, H.F.; Wang, Y.; Li, F.C.; Shang, M.Y.; Liu, G.X.; Wang, X.; Cai, S.Q. Detection of 191 taxifolin metabolites and their distribution in rats using HPLC-ESI-IT-TOF-MS(n). Molecules 2016, 21, 1209. [Google Scholar] [CrossRef] [Green Version]
  67. Wang, Y.; Wang, Q.; Bao, X.; Ding, Y.; Shentu, J.; Cui, W. Taxifolin prevents β-amyloid-induced impairments of synaptic formation and deficits of memory via the inhibition of cytosolic phospholipase A2/prostaglandin E2 content. Metab. Brain Dis. 2018, 33, 1069–1079. [Google Scholar] [CrossRef]
  68. Saito, S.; Yamamoto, Y.; Maki, T.; Hattori, Y.; Ito, H.; Mizuno, K. Taxifolin inhibits amyloid-β oligomer formation and fully restores vascular integrity and memory in cerebral amyloid angiopathy. Acta. Neuropathol. Commun. 2017, 5, 26. [Google Scholar] [CrossRef] [PubMed]
  69. Rahman, M.H.; Akter, R.; Bhattacharya, T.; Abdel-Daim, M.M.; Alkahtani, S.; Arafah, M.W.; Al-Johani, N.S.; Alhoshani, N.M.; Alkeraishan, N.; Alhenaky, A.; et al. Resveratrol and neuroprotection: Impact and its therapeutic potential in alzheimer’s disease. Front. Pharmacol. 2020, 11, 619024. [Google Scholar] [CrossRef]
  70. Bastianetto, S.; Menard, C.; Quirion, R. Neuroprotective action of resveratrol. Biochim. Biophys. Acta 2015, 1852, 1195–1201. [Google Scholar] [CrossRef] [Green Version]
  71. Hou, Y.; Wang, K.; Wan, W.; Cheng, Y.; Pu, X.; Ye, X. Resveratrol provides neuroprotection by regulating the JAK2/STAT3/PI3K/AKT/mTOR pathway after stroke in rats. Genes Dis. 2018, 5, 245–255. [Google Scholar] [CrossRef] [PubMed]
  72. Koushki, M.; Amiri-Dashatan, N.; Ahmadi, N.; Abbaszadeh, H.A.; Rezaei-Tavirani, M. Resveratrol: A miraculous natural compound for diseases treatment. Food Sci. Nutr. 2018, 6, 2473–2490. [Google Scholar] [CrossRef] [Green Version]
  73. Iside, C.; Scafuro, M.; Nebbioso, A.; Altucci, L. Sirt1 activation by natural phytochemicals: An overview. Front. Pharmacol. 2020, 11, 1225. [Google Scholar] [CrossRef] [PubMed]
  74. Durazzo, A.; Lucarini, M.; Souto, E.B.; Cicala, C.; Caiazzo, E.; Izzo, A.A.; Novellino, E.; Santini, A. Polyphenols: A concise overview on the chemistry, occurrence, and human health. Phytother. Res. 2019, 33, 2221–2243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Ferreira, P.E.; Beraldi, E.J.; Borges, S.C.; Natali, M.R.; Buttow, N.C. Resveratrol promotes neuroprotection and attenuates oxidative and nitrosative stress in the small intestine in diabetic rats. Biomed. Pharmacother. 2018, 105, 724–733. [Google Scholar] [CrossRef] [PubMed]
  76. Ma, T.; Tan, M.S.; Yu, J.T.; Tan, L. Resveratrol as a therapeutic agent for alzheimer’s disease. Biomed. Res. Int. 2014, 2014, 350516. [Google Scholar] [CrossRef]
  77. Shahidi, F.; Peng, H. Bioaccessibility and bioavailability of phenolic compounds. J. Food Bioact. 2018, 4, 11–68. [Google Scholar] [CrossRef] [Green Version]
  78. Davidov-Pardo, G.; McClements, D.J. Resveratrol encapsulation: Designing delivery systems to overcome solubility, stability and bioavailability issues. Trends Food Sci. Technol. 2014, 38, 88–103. [Google Scholar] [CrossRef]
  79. Sathya, M.; Moorthi, P.; Premkumar, P.; Kandasamy, M.; Jayachandran, K.S.; Anusuyadevi, M. Resveratrol intervenes cholesterol and isoprenoid-mediated amyloidogenic processing of abetapp in familial alzheimer’s disease. J. Alzheimers Dis. 2017, 60, S3–S23. [Google Scholar] [CrossRef]
  80. Lopez, M.S.; Dempsey, R.J.; Vemuganti, R. Resveratrol neuroprotection in stroke and traumatic cns injury. Neurochem. Int. 2015, 89, 75–82. [Google Scholar] [CrossRef] [Green Version]
  81. Zhang, G.; Liu, Y.; Xu, L.; Sha, C.; Zhang, H.; Xu, W. Resveratrol alleviates lipopolysaccharide-induced inflammation in PC-12 cells and in rat model. BMC Biotechnol. 2019, 19, 10. [Google Scholar] [CrossRef]
  82. Hou, Y.; Zhang, Y.; Mi, Y.; Wang, J.; Zhang, H.; Xu, J.; Yang, Y.; Liu, J.; Ding, L.; Yang, J.; et al. A novel quinolyl-substituted analogue of resveratrol inhibits LPS-induced inflammatory responses in microglial cells by blocking the NF-κB/MAPK signaling pathways. Mol. Nutr. Food Res. 2019, 63, e1801380. [Google Scholar] [CrossRef] [PubMed]
  83. Salehi, B.; Mishra, A.P.; Nigam, M.; Sener, B.; Kilic, M.; Sharifi-Rad, M.; Fokou, P.V.; Martins, N.; Sharifi-Rad, J. Resveratrol: A double-edged sword in health benefits. Biomedicines 2018, 6, 91. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Deane, R.; Wu, Z.; Sagare, A.; Davis, J.; Du Yan, S.; Hamm, K.; Xu, F.; Parisi, M.; LaRue, B.; Hu, H.W.; et al. LRP/amyloid β-peptide interaction mediates differential brain efflux of Aβ isoforms. Neuron 2004, 43, 333–344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Saha, A.; Sarkar, C.; Singh, S.P.; Zhang, Z.; Munasinghe, J.; Peng, S.; Chandra, G.; Kong, E.; Mukherjee, A.B. The blood-brain barrier is disrupted in a mouse model of infantile neuronal ceroid lipofuscinosis: Amelioration by resveratrol. Hum. Mol. Genet. 2012, 21, 2233–2244. [Google Scholar] [CrossRef] [Green Version]
  86. Moussa, C.; Hebron, M.; Huang, X.; Ahn, J.; Rissman, R.A.; Aisen, P.S.; Turner, R.S. Resveratrol regulates neuro-inflammation and induces adaptive immunity in Alzheimer’s disease. J. Neuroinflamm. 2017, 14, 1. [Google Scholar] [CrossRef] [Green Version]
  87. Nabavi, S.F.; Khan, H.; D’onofrio, G.; Samec, D.; Shirooie, S.; Dehpour, A.R.; Arguelles, S.; Habtemariam, S.; Sobarzo-Sanchez, E. Apigenin as neuroprotective agent: Of mice and men. Pharmacol. Res. 2018, 128, 359–365. [Google Scholar] [CrossRef]
  88. Yan, X.; Qi, M.; Li, P.; Zhan, Y.; Shao, H. Apigenin in cancer therapy: Anti-cancer effects and mechanisms of action. Cell Biosci. 2017, 7, 50. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Wu, J.J.; Ai, C.Z.; Liu, Y.; Zhang, Y.Y.; Jiang, M.; Fan, X.R.; Lv, A.P.; Yang, L. Interactions between phytochemicals from traditional chinese medicines and human cytochrome p450 enzymes. Curr. Drug Metab. 2012, 13, 599–614. [Google Scholar] [CrossRef]
  90. Sanchez-Marzo, N.; Perez-Sanchez, A.; Ruiz-Torres, V.; Martinez-Tebar, A.; Castillo, J.; Herranz-Lopez, M.; Barrajon-Catalan, E. Antioxidant and photoprotective activity of apigenin and its potassium salt derivative in human keratinocytes and absorption in Caco-2 cell monolayers. Int. J. Mol. Sci. 2019, 20, 2148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Siddique, Y.H.; Naz, F.; Jyoti, S.; Afzal, M. Protective effect of apigenin in transgenic drosophila melanogaster model of parkinson’s disease. Pharmacology 2011, 3, 790–795. [Google Scholar]
  92. Balez, R.; Steiner, N.; Engel, M.; Munoz, S.S.; Lum, J.S.; Wu, Y.; Wang, D.; Vallotton, P.; Sachdev, P.; O’Connor, M.; et al. Neuroprotective effects of apigenin against inflammation, neuronal excitability and apoptosis in an induced pluripotent stem cell model of alzheimer’s disease. Sci. Rep. 2016, 6, 31450. [Google Scholar] [CrossRef] [Green Version]
  93. Popović, M.; Caballero-Bleda, M.; Benavente-García, O.; Castillo, J. The flavonoid apigenin delays forgetting of passive avoidance conditioning in rats. J. Psychopharmacol. 2014, 28, 498–501. [Google Scholar] [CrossRef] [PubMed]
  94. Ganai, A.A.; Farooqi, H. Bioactivity of genistein: A review of in vitro and in vivo studies. Biomed. Pharmacother. 2015, 76, 30–38. [Google Scholar] [CrossRef]
  95. Rahman, M.H.; Akter, R.; Kamal, M.A. Prospective function of different antioxidant containing natural products in the treatment of neurodegenerative disease. CNS Neurol. Disord. Drug. Targets 2020. [Google Scholar] [CrossRef]
  96. Sheeja Malar, D.; Pandima Devi, K. Dietary polyphenols for treatment of alzheimer’s disease—Future research and development. Curr. Pharm. Biotechnol. 2014, 15, 330–342. [Google Scholar] [CrossRef]
  97. You, F.; Li, Q.; Jin, G.; Zheng, Y.; Chen, J.; Yang, H. Genistein protects against Aβ 25-35 induced apoptosis of PC12 cells through JNK signaling and modulation of BCL-2 family messengers. BMC Neurosci. 2017, 18, 12. [Google Scholar] [CrossRef] [Green Version]
  98. Bagheri, M.; Joghataei, M.T.; Mohseni, S.; Roghani, M. Genistein ameliorates learning and memory deficits in amyloid beta(1-40) rat model of alzheimer’s disease. Neurobiol. Learn. Mem. 2011, 95, 270–276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Kim, K.H.; Dodsworth, C.; Paras, A.; Burton, B.K. High dose genistein aglycone therapy is safe in patients with mucopolysaccharidoses involving the central nervous system. Mol. Genet. Metab. 2013, 109, 382–385. [Google Scholar] [CrossRef]
  100. Hajialyani, M.; Farzaei, M.H.; Echeverria, J.; Nabavi, S.M.; Uriarte, E.; Sobarzo-Sanchez, E. Hesperidin as a neuroprotective agent: A review of animal and clinical evidence. Molecules 2019, 24, 648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Bhadbhade, A.; Cheng, D.W. Amyloid precursor protein processing in alzheimer’s disease. Iran. J. Child Neurol. 2012, 6, 1–4. [Google Scholar] [CrossRef]
  102. Zhang, Y.; Huang, N.Q.; Yan, F.; Jin, H.; Zhou, S.Y.; Shi, J.S.; Jin, F. Diabetes mellitus and alzheimer’s disease: GSK-3β as a potential link. Behav. Brain Res. 2018, 339, 57–65. [Google Scholar] [CrossRef] [PubMed]
  103. Khan, A.; Jahan, S.; Imtiyaz, Z.; Alshahrani, S.; Antar Makeen, H.; Mohammed Alshehri, B.; Kumar, A.; Arafah, A.; Rehman, M.U. Neuroprotection: Targeting multiple pathways by naturally occurring phytochemicals. Biomedicines 2020, 8, 284. [Google Scholar] [CrossRef]
  104. Justin Thenmozhi, A.; William Raja, T.R.; Manivasagam, T.; Janakiraman, U.; Essa, M.M. Hesperidin ameliorates cognitive dysfunction, oxidative stress and apoptosis against aluminium chloride induced rat model of alzheimer’s disease. Nutr. Neurosci. 2017, 20, 360–368. [Google Scholar] [CrossRef]
  105. Kim, J.; Wie, M.B.; Ahn, M.; Tanaka, A.; Matsuda, H.; Shin, T. Benefits of hesperidin in central nervous system disorders: A review. Anat. Cell Biol. 2019, 52, 369–377. [Google Scholar] [CrossRef] [PubMed]
  106. Cao, Z.; Wang, F.; Xiu, C.; Zhang, J.; Li, Y. Hypericum perforatum extract attenuates behavioral, biochemical, and neurochemical abnormalities in aluminum chloride-induced alzheimer’s disease rats. Biomed. Pharmacother. 2017, 91, 931–937. [Google Scholar] [CrossRef]
  107. Kumar, A.; Lalitha, S.; Mishra, J. Hesperidin potentiates the neuroprotective effects of diazepam and gabapentin against pentylenetetrazole-induced convulsions in mice: Possible behavioral, biochemical and mitochondrial alterations. Indian J. Pharmacol. 2014, 46, 309–315. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Chang, C.Y.; Lin, T.Y.; Lu, C.W.; Huang, S.K.; Wang, Y.C.; Chou, S.S.P.; Wang, S.J. Hesperidin inhibits glutamate release and exerts neuroprotection against excitotoxicity induced by kainic acid in the hippocampus of rats. Neurotoxicology 2015, 50, 157–169. [Google Scholar] [CrossRef]
  109. Xian, Y.F.; Lin, Z.X.; Mao, Q.Q.; Hu, Z.; Zhao, M.; Che, C.T.; Ip, S.P. Bioassay-guided isolation of neuroprotective compounds from uncaria rhynchophylla against beta-amyloid-induced neurotoxicity. Evid. Based. Complement. Alternat. Med. 2012, 2012, 802625. [Google Scholar] [CrossRef]
  110. Ling, L.Z.; Zhang, S.D. The complete chloroplast genome of the traditional chinese herb, uncaria rhynchophylla (rubiaceae). Mitochondrial. DNA B Resour. 2020, 5, 424–425. [Google Scholar] [CrossRef]
  111. Wang, Y.L.; Dong, P.P.; Liang, J.H.; Li, N.; Sun, C.P.; Tian, X.G.; Huo, X.K.; Zhang, B.J.; Ma, X.C.; Lv, C.Z. Phytochemical constituents from uncaria rhynchophylla in human carboxylesterase 2 inhibition: Kinetics and interaction mechanism merged with docking simulations. Phytomedicine 2018, 51, 120–127. [Google Scholar] [CrossRef]
  112. Yang, W.; Ip, S.P.; Liu, L.; Xian, Y.F.; Lin, Z.X. Uncaria rhynchophylla and its major constituents on central nervous system: A review on their pharmacological actions. Curr. Vasc. Pharmacol. 2020, 18, 346–357. [Google Scholar] [CrossRef] [PubMed]
  113. Shim, J.S.; Kim, H.G.; Ju, M.S.; Choi, J.G.; Jeong, S.Y.; Oh, M.S. Effects of the hook of uncaria rhynchophylla on neurotoxicity in the 6-hydroxydopamine model of parkinson’s disease. J. Ethnopharmacol. 2009, 126, 361–365. [Google Scholar] [CrossRef] [PubMed]
  114. Pal, B.; Kumar, S.S. Evaluation of anti-parkinson’s activity of uncaria rhynchophylla in 6-hydroxy dopamine lesioned rat model. Int. J. Appl. Res. 2015, 1, 203–206. [Google Scholar]
  115. Lan, Y.L.; Zhou, J.J.; Liu, J.; Huo, X.K.; Wang, Y.L.; Liang, J.H.; Zhao, J.C.; Sun, C.P.; Yu, Z.L.; Fang, L.L.; et al. Uncaria rhynchophylla ameliorates parkinson’s disease by inhibiting HSP90 expression: Insights from quantitative proteomics. Cell. Physiol. Biochem. 2018, 47, 1453–1464. [Google Scholar] [CrossRef] [PubMed]
  116. Zhang, Y.N.; Yang, Y.F.; Xu, W.; Yang, X.W. The blood-brain barrier permeability of six indole alkaloids from UncariaeRamulus Cum Uncis in the MDCK-pHaMDR cell monolayer model. Molecules 2017, 22, 1944. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Zhang, C.; Wu, X.; Xian, Y.; Zhu, L.; Lin, G.; Lin, Z.X. Evidence on integrating pharmacokinetics to find truly therapeutic agent for Alzheimer’s disease: Comparative pharmacokinetics and disposition kinetics profiles of stereoisomers isorhynchophylline and rhynchophylline in rats. Evid. Based. Complement. Altern. Med. 2019, 2019, 4016323. [Google Scholar] [CrossRef] [Green Version]
  118. Thomas, N.V.; Kim, S.K. Beneficial effects of marine algal compounds in cosmeceuticals. Mar. Drugs 2013, 11, 146–164. [Google Scholar] [CrossRef] [Green Version]
  119. Overland, M.; Mydland, L.T.; Skrede, A. Marine macroalgae as sources of protein and bioactive compounds in feed for monogastric animals. J. Sci. Food. Agric. 2019, 99, 13–24. [Google Scholar] [CrossRef] [Green Version]
  120. Singh, R.; Parihar, P.; Singh, M.; Bajguz, A.; Kumar, J.; Singh, S.; Singh, V.P.; Prasad, S.M. Uncovering potential applications of cyanobacteria and algal metabolites in biology, agriculture and medicine: Current status and future prospects. Front. Microbiol. 2017, 8, 515. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Dawczynski, C.; Schafer, U.; Leiterer, M.; Jahreis, G. Nutritional and toxicological importance of macro, trace, and ultra-trace elements in algae food products. J. Agric. Food. Chem. 2007, 55, 10470–10475. [Google Scholar] [CrossRef]
  122. Chamorro, G.; Pérez-Albiter, M.; Serrano-García, N.; Mares-Sámano, J.J.; Rojas, P. Spirulina maxima pretreatment partially protects against 1-methyl-4-phenyl-1, 2, 3, 6-tetrahydropyridine neurotoxicity. Nutr. Neurosci. 2006, 9, 207–212. [Google Scholar] [CrossRef]
  123. Silva, J.; Alves, C.; Pinteus, S.; Mendes, S.; Pedrosa, R. Neuroprotective effects of seaweeds against 6-hydroxidopamine-induced cell death on an in vitro human neuroblastoma model. BMC Complement. Altern. Med. 2018, 18, 58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Cunha, L.; Grenha, A. Sulfated seaweed polysaccharides as multifunctional materials in drug delivery applications. Mar. Drugs 2016, 14, 42. [Google Scholar] [CrossRef] [PubMed]
  125. Barbosa, A.I.; Coutinho, A.J.; Lima, S.A.; Reis, S. Marine polysaccharides in pharmaceutical applications: Fucoidan and chitosan as key players in the drug delivery match field. Mar. Drugs 2019, 17, 654. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Zhang, L.; Hao, J.; Zheng, Y.; Su, R.; Liao, Y.; Gong, X.; Liu, L.; Wang, X. Fucoidan protects dopaminergic neurons by enhancing the mitochondrial function in a rotenone-induced rat model of parkinson’s disease. Aging Dis. 2018, 9, 590–604. [Google Scholar] [CrossRef] [Green Version]
  127. Wei, H.; Gao, Z.; Zheng, L.; Zhang, C.; Liu, Z.; Yang, Y.; Teng, H.; Hou, L.; Yin, Y.; Zou, X. Protective effects of fucoidan on Aβ25-35 and d-Gal-induced neurotoxicity in PC12 Cells and d-Gal-induced cognitive dysfunction in mice. Mar. Drugs 2017, 15, 77. [Google Scholar] [CrossRef] [Green Version]
  128. Mohd Sairazi, N.S.; Sirajudeen, K.N. Natural products and their bioactive compounds: Neuroprotective potentials against neurodegenerative diseases. Evid. Based Complement. Alternat. Med. 2020, 2020, 6565396. [Google Scholar] [CrossRef]
  129. Pal, S.; Paul, S. Conformational deviation of thrombin binding G-quadruplex aptamer (TBA) in presence of divalent cation Sr2+: A classical molecular dynamics simulation study. Int. J. Biol. Macromol. 2019, 121, 350–363. [Google Scholar] [CrossRef]
  130. Hannan, M.A.; Dash, R.; Haque, M.N.; Mohibbullah, M.; Sohag, A.A.; Rahman, M.A.; Uddin, M.J.; Alam, M.; Moon, I. Neuroprotective potentials of marine algae and their bioactive metabolites: Pharmacological insights and therapeutic advances. Mar. Drugs. 2020, 18, 347. [Google Scholar] [CrossRef] [PubMed]
  131. Kiuru, P.; D’Auria, M.V.; Muller, C.D.; Tammela, P.; Vuorela, H.; Yli-Kauhaluoma, J. Exploring marine resources for bioactive compounds. Planta. Med. 2014, 80, 1234–1246. [Google Scholar] [CrossRef]
  132. Moradi-Kor, N.; Ghanbari, A.; Rashidipour, H.; Bandegi, A.R.; Youse fi, B.; Barati, M.; Kokhaei, P.; Rashidy-Pour, A. Therapeutic effects of spirulina platensis against adolescent stress-induced oxidative stress, brain-derived neurotrophic factor alterations and morphological remodeling in the amygdala of adult female rats. J. Exp. Pharmacol. 2020, 12, 75–85. [Google Scholar] [CrossRef] [Green Version]
  133. Zhang, F.; Lu, J.; Zhang, J.G.; Xie, J.X. Protective effects of a polysaccharide from spirulina platensis on dopaminergic neurons in an MPTP-induced parkinson’s disease model in C57BL/6J mice. Neural. Regen. Res. 2015, 10, 308–313. [Google Scholar] [CrossRef]
  134. Huang, C.; Zhang, Z.; Cui, W. Marine-derived natural compounds for the treatment of parkinson’s disease. Mar. Drugs 2019, 17, 221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Koh, E.J.; Kim, K.J.; Song, J.H.; Choi, J.; Lee, H.Y.; Kang, D.H.; Heo, H.J.; Lee, B.Y. Spirulina maxima extract ameliorates learning and memory impairments via inhibiting GSK-3β phosphorylation induced by intracerebroventricular injection of amyloid-β 1–42 in mice. Int. J. Mol. Sci. 2017, 18, 2401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Buono, S.; Langellotti, A.L.; Martello, A.; Rinna, F.; Fogliano, V. Functional ingredients from microalgae. Food. Funct. 2014, 5, 1669–1685. [Google Scholar] [CrossRef]
  137. Choi, W.Y.; Kang, D.H.; Lee, H.Y. Effect of fermented spirulina maxima extract on cognitive-enhancing activities in mice with scopolamine-induced dementia. Evid. Based. Complement. Alternat. Med. 2018, 2018, 7218504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Rimbau, V.; Camins, A.; Romay, C.; González, R.; Pallàs, M. Protective effects of C-phycocyanin against kainic acid-induced neuronal damage in rat hippocampus. Neurosci. Lett. 1999, 276, 75–78. [Google Scholar] [CrossRef]
  139. Koh, E.J.; Seo, Y.J.; Choi, J.; Lee, H.Y.; Kang, D.H.; Kim, K.J.; Lee, B.Y. Spirulina maxima extract prevents neurotoxicity via promoting activation of BDNF/CREB signaling pathways in neuronal cells and mice. Molecules 2017, 22, 1363. [Google Scholar] [CrossRef] [PubMed]
  140. Angeloni, C.; Vauzour, D. Natural products and neuroprotection. Int. J. Mol. Sci. 2019, 20, 5570. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Durrenberger, P.F.; Fernando, F.S.; Kashefi, S.N.; Bonnert, T.P.; Seilhean, D.; Nait-Oumesmar, B.; Schmitt, A.; Gebicke-Haerter, P.J.; Falkai, P.; Grunblatt, E.; et al. Common mechanisms in neurodegeneration and neuroinflammation: A brainnet europe gene expression microarray study. J. Neural Transm. 2015, 122, 1055–1068. [Google Scholar] [CrossRef]
  142. Schetters, S.T.; Gomez-Nicola, D.; Garcia-Vallejo, J.J.; Van Kooyk, Y. Neuroinflammation: Microglia and T cells get ready to tango. Front. Immunol. 2017, 8, 1905. [Google Scholar] [CrossRef] [Green Version]
  143. Shan, C.S.; Zhang, H.F.; Xu, Q.Q.; Shi, Y.H.; Wang, Y.; Li, Y.; Lin, Y.; Zheng, G.Q. Herbal medicine formulas for parkinson’s disease: A systematic review and meta-analysis of randomized double-blind placebo-controlled clinical trials. Front. Aging Neurosci. 2018, 10, 349. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Panda, S.S.; Jhanji, N. Natural products as potential anti-alzheimer agents. Curr. Med. Chem. 2020, 27, 5887–5917. [Google Scholar] [CrossRef] [PubMed]
  145. Leonoudakis, D.; Rane, A.; Angeli, S.; Lithgow, G.J.; Andersen, J.K.; Chinta, S.J. Anti-inflammatory and neuroprotective role of natural product securinine in activated glial cells: Implications for parkinson’s disease. Mediat. Inflamm. 2017, 2017, 8302636. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Martinez-Huelamo, M.; Rodriguez-Morato, J.; Boronat, A.; de la Torre, R. Modulation of NRF2 by olive oil and wine polyphenols and neuroprotection. Antioxidants 2017, 6, 73. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Rehman, M.U.; Wali, A.F.; Ahmad, A.; Shakeel, S.; Rasool, S.; Ali, R.; Rashid, S.M.; Madkhali, H.; Ganaie, M.A.; Khan, R. Neuroprotective strategies for neurological disorders by natural products: An update. Curr. Neuropharmacol 2019, 17, 247–267. [Google Scholar] [CrossRef] [PubMed]
  148. Braak, H.; Del Tredici, K. The preclinical phase of the pathological process underlying sporadic alzheimer’s disease. Brain 2015, 138, 2814–2833. [Google Scholar] [CrossRef]
  149. Deshpande, P.; Gogia, N.; Singh, A. Exploring the efficacy of natural products in alleviating alzheimer’s disease. Neural Regen. Res. 2019, 14, 1321–1329. [Google Scholar] [CrossRef]
  150. Castelli, V.; Grassi, D.; Bocale, R.; d’Angelo, M.; Antonosante, A.; Cimini, A.; Ferri, C.; Desideri, G. Diet and brain health: Which role for polyphenols? Curr. Pharm. Des. 2018, 24, 227–238. [Google Scholar] [CrossRef]
  151. Bagli, E.; Goussia, A.; Moschos, M.M.; Agnantis, N.; Kitsos, G. Natural compounds and neuroprotection: Mechanisms of action and novel delivery systems. In Vivo 2016, 30, 535–547. [Google Scholar]
  152. Lan, J.; Liu, Z.; Liao, C.; Merkler, D.J.; Han, Q.; Li, J. A study for therapeutic treatment against parkinson’s disease via Chou’s 5-steps rule. Curr. Top. Med. Chem. 2019, 19, 2318–2333. [Google Scholar] [CrossRef]
  153. Haddad, F.; Sawalha, M.; Khawaja, Y.; Najjar, A.; Karaman, R. Dopamine and levodopa prodrugs for the treatment of parkinson’s disease. Molecules 2018, 23, 40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Park, J.S.; Davis, R.L.; Sue, C.M. Mitochondrial Dysfunction in Parkinson’s Disease: New Mechanistic Insights and Therapeutic Perspectives. Curr. Neurol. Neurosci. Rep. 2018, 18, 21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Moon, H.E.; Paek, S.H. Mitochondrial Dysfunction in Parkinson’s Disease. Exp. Neurobiol. 2015, 24, 103. [Google Scholar] [CrossRef] [PubMed]
  156. Kaushik, S.; Cuervo, A.M. Proteostasis and aging. Nat. Med. 2015, 21, 1406–1415. [Google Scholar] [CrossRef] [PubMed]
  157. Wales, P.; Pinho, R.; Lázaro, D.F.; Outeiro, T.F. Limelight on alpha-synuclein: Pathological and mechanistic implications in neurodegeneration. J. Parkinsons. Dis. 2013, 3, 415–459. [Google Scholar] [CrossRef] [Green Version]
  158. Ciulla, M.; Marinelli, L.; Cacciatore, I.; Stefano, A.D. Role of Dietary Supplements in the Management of Parkinson’s Disease. Biomolecules 2019, 9, 271. [Google Scholar] [CrossRef] [Green Version]
  159. Manyam, B.V.; Dhanasekaran, M.; Hare, T.A. Neuroprotective effects of the antiparkinson drug Mucunapruriens. Phytother. Res. 2004, 18, 706–712. [Google Scholar] [CrossRef]
  160. Brichta, L.; Greengard, P.; Flajolet, M. Advances in the pharmacological treatment of parkinson’s disease: Targeting neurotransmitter systems. Trends Neurosci. 2013, 36, 543–554. [Google Scholar] [CrossRef]
  161. Manoharan, S.; Essa, M.M.; Vinoth, A.; Kowsalya, R.; Manimaran, A.; Selvasundaram, R. Alzheimer’s disease and medicinal plants: An overview. Adv. Neurobiol. 2016, 12, 95–105. [Google Scholar] [CrossRef]
  162. Martinez-Oliveira, P.; de Oliveira, M.F.; Alves, N.; Coelho, R.P.; Pilar, B.C.; Güllich, A.A.; Ströher, D.J.; Boligon, A.; Piccoli, J.D.; Mello-Carpes, P.B. Yacon leaf extract supplementation demonstrates neuroprotective effect against memory deficit related to β-amyloid-induced neurotoxicity. J. Funct. Foods 2018, 48, 665–675. [Google Scholar] [CrossRef]
  163. Zhang, L.; Zhou, Z.; Zhai, W.; Pang, J.; Mo, Y.; Yang, G.; Qu, Z.; Hu, Y. Safflower yellow attenuates learning and memory deficits in amyloid β-induced alzheimer’s disease rats by inhibiting neuroglia cell activation and inflammatory signaling pathways. Metab. Brain Dis. 2019, 34, 927–939. [Google Scholar] [CrossRef]
  164. Ionita, R.; Postu, P.A.; Mihasan, M.; Gorgan, D.L.; Hancianu, M.; Cioanca, O.; Hritcu, L. Ameliorative effects of Matricaria chamomilla L. hydroalcoholic extract on scopolamine-induced memory impairment in rats: A behavioral and molecular study. Phytomedicine 2018, 47, 113–120. [Google Scholar] [CrossRef]
  165. Hishikawa, N.; Takahashi, Y.; Amakusa, Y.; Tanno, Y.; Tuji, Y.; Niwa, H.; Murakami, N.; Krishna, U.K. Effects of turmeric on alzheimer’s disease with behavioral and psychological symptoms of dementia. Ayu 2012, 33, 499–504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Khongsombat, O.; Nakdook, W.; Ingkaninan, K. Inhibitory effects of tabernaemontana divaricata root extract on oxidative stress and neuronal loss induced by amyloid β25–35 peptide in mice. J. Tradit. Complement. Med. 2018, 8, 184–189. [Google Scholar] [CrossRef] [PubMed]
  167. de la Rubia Orti, J.E.; Garcia-Pardo, M.P.; Drehmer, E.; Sancho Cantus, D.; Julian Rochina, M.; Aguilar, M.A.; Hu Yang, I. Improvement of main cognitive functions in patients with alzheimer’s disease after treatment with coconut oil enriched mediterranean diet: A pilot study. J. Alzheimers Dis. 2018, 65, 577–587. [Google Scholar] [CrossRef] [PubMed]
  168. Ali, T.; Yoon, G.H.; Shah, S.A.; Lee, H.Y.; Kim, M.O. Osmotin attenuates amyloid β-induced memory impairment, tau phosphorylation and neurodegeneration in the mouse hippocampus. Sci. Rep. 2015, 5, 11708. [Google Scholar] [CrossRef]
  169. Azmi, N.H.; Ismail, M.; Ismail, N.; Imam, M.U.; Alitheen, N.B.; Abdullah, M.A. Germinated brown rice alters Aβ (1-42) aggregation and modulates alzheimer’s disease-related genes in differentiated human SH-SY5Y cells. Evid. Based Complement. Alternat. Med. 2015, 2015, 153684. [Google Scholar] [CrossRef] [Green Version]
  170. Xu, S.S.; Gao, Z.X.; Weng, Z.; Du, Z.M.; Xu, W.A.; Yang, J.S.; Zhang, M.L.; Tong, Z.H.; Fang, Y.S.; Chai, X.S.; et al. Efficacy of tablet huperzine-a on memory, cognition, and behavior in alzheimer’s disease. Zhongguo Yao Li Xue Bao 1995, 16, 391–395. [Google Scholar] [PubMed]
  171. Colovic, M.B.; Krstic, D.Z.; Lazarevic-Pasti, T.D.; Bondzic, A.M.; Vasic, V.M. Acetylcholinesterase inhibitors: Pharmacology and toxicology. Curr. Neuropharmacol. 2013, 11, 315–335. [Google Scholar] [CrossRef] [Green Version]
  172. Ibrahim, F.W.; Zainudin, U.N.; Latif, M.A.; Hamid, A. Neuroprotective effects of ethyl acetate extract of Zingiber zerumbet (L.) smith against oxidative stress on paraquat-induced parkinsonism in rats. Sains Malays. 2018, 47, 2337–2347. [Google Scholar] [CrossRef]
  173. Bisht, R.; Joshi, B.C.; Kalia, A.N.; Prakash, A. Antioxidant-rich fraction of urtica dioica mediated rescue of striatal mito-oxidative damage in MPTP-induced behavioral, cellular, and neurochemical alterations in rats. Mol. Neurobiol. 2017, 54, 5632–5645. [Google Scholar] [CrossRef]
  174. Chonpathompikunlert, P.; Boonruamkaew, P.; Sukketsiri, W.; Hutamekalin, P.; Sroyraya, M. The antioxidant and neurochemical activity of apium graveolens L. and its ameliorative effect on MPTP-induced parkinson-like symptoms in mice. BMC Complement. Altern. Med. 2018, 18, 103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Karbarz, M.; Mytych, J.; Solek, P.; Stawarczyk, K.; Tabecka-Lonczynska, A.; Koziorowski, M.; Luczaj, L. Cereal grass juice in wound healing: Hormesis and cell-survival in normal fibroblasts, in contrast to toxic events in cancer cells. J. Physiol. Pharmacol. 2019, 70, 595–604. [Google Scholar] [CrossRef]
  176. Kosaraju, J.; Chinni, S.; Roy, P.D.; Kannan, E.; Antony, A.S.; Kumar, M.N. Neuroprotective effect of tinospora cordifolia ethanol extract on 6-hydroxy dopamine induced parkinsonism. Indian J. Pharmacol. 2014, 46, 176–180. [Google Scholar] [CrossRef] [Green Version]
  177. Ren, Z.X.; Zhao, Y.F.; Cao, T.; Zhen, X.C. Dihydromyricetin protects neurons in an MPTP-induced model of parkinson’s disease by suppressing glycogen synthase kinase-3 beta activity. Acta Pharmacol. Sin. 2016, 37, 1315–1324. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Ye, Q.; Wang, W.; Hao, C.; Mao, X. Agaropentaose protects SH-SY5Y cells against 6-hydroxydopamine-induced neurotoxicity through modulating NF-κB and p38MAPK signaling pathways. J. Funct. Foods 2019, 57, 222–232. [Google Scholar] [CrossRef]
  179. Abdel-Salam, O.M.; Sleem, A.A.; Youness, E.R.; Yassen, N.N.; Shaffie, N.; El-Toumy, S.A. Capsicum protects against rotenone-induced toxicity in mice brain via reduced oxidative stress and 5-lipoxygenase activation. J. Pharm. Pharmacol. Res. 2018, 2, 60–77. [Google Scholar] [CrossRef]
  180. Ojha, S.; Javed, H.; Azimullah, S.; Haque, M.E. Beta-caryophyllene, a phytocannabinoid attenuates oxidative stress, neuroinflammation, glial activation, and salvages dopaminergic neurons in a rat model of parkinson disease. Mol. Cell. Biochem. 2016, 418, 59–70. [Google Scholar] [CrossRef] [PubMed]
  181. Jiang, C.Y.; Qin, X.Y.; Yuan, M.M.; Lu, G.J.; Cheng, Y. 2,3,5,4′-tetrahydroxystilbene-2-o-beta-d-glucoside reverses stress-induced depression via inflammatory and oxidative stress pathways. Oxid. Med. Cell. Longev. 2018, 2018, 9501427. [Google Scholar] [CrossRef] [Green Version]
  182. Roy, N.K.; Parama, D.; Banik, K.; Bordoloi, D.; Devi, A.K.; Thakur, K.K.; Padmavathi, G.; Shakibaei, M.; Fan, L.; Sethi, G.; et al. An update on pharmacological potential of boswellic acids against chronic diseases. Int. J. Mol. Sci. 2019, 20, 4101. [Google Scholar] [CrossRef] [Green Version]
  183. Bhullar, K.S.; Rupasinghe, H.P. Polyphenols: Multipotent therapeutic agents in neurodegenerative diseases. Oxid. Med. Cell. Longev. 2013, 2013, 891748. [Google Scholar] [CrossRef] [Green Version]
  184. Sarbishegi, M.; Charkhat Gorgich, E.A.; Khajavi, O.; Komeili, G.; Salimi, S. The neuroprotective effects of hydro-alcoholic extract of olive (Olea europaea L.) leaf on rotenone-induced parkinson’s disease in rat. Metab. Brain Dis. 2018, 33, 79–88. [Google Scholar] [CrossRef]
  185. Tomani, J.C.; Gainkam, L.O.; Nshutiyayesu, S.; Mukazayire, M.J.; Ribeiro, S.O.; Stevigny, C.; Frederich, M.; Muganga, R.; Souopgui, J. An ethnobotanical survey and inhibitory effects on NLRP3 inflammasomes/caspase-1 of herbal recipes’ extracts traditionally used in rwanda for asthma treatment. J. Ethnopharmacol. 2018, 227, 29–40. [Google Scholar] [CrossRef]
  186. Wang, Q.; Kuang, H.; Su, Y.; Sun, Y.; Feng, J.; Guo, R.; Chan, K. Naturally derived anti-inflammatory compounds from chinese medicinal plants. J. Ethnopharmacol. 2013, 146, 9–39. [Google Scholar] [CrossRef] [PubMed]
  187. Mori, M.A.; Delattre, A.M.; Carabelli, B.; Pudell, C.; Bortolanza, M.; Staziaki, P.V.; Visentainer, J.V.; Montanher, P.F.; Del Bel, E.A.; Ferraz, A.C. Neuroprotective effect of omega-3 polyunsaturated fatty acids in the 6-OHDA model of parkinson’s disease is mediated by a reduction of inducible nitric oxide synthase. Nutr. Neurosci. 2018, 21, 341–351. [Google Scholar] [CrossRef] [PubMed]
  188. Saleh, A.S.; Wang, P.; Wang, N.; Yang, L.; Xiao, Z. Brown rice versus white rice: Nutritional quality, potential health benefits, development of food products, and preservation technologies. Compr. Rev. Food Sci. Food Saf. 2019, 18, 1070–1096. [Google Scholar] [CrossRef] [Green Version]
  189. Akhondzadeh, S.; Noroozian, M.; Mohammadi, M.; Ohadinia, S.; Jamshidi, A.H.; Khani, M. Salvia officinalis extract in the treatment of patients with mild to moderate Alzheimer’s disease: A double blind, randomized and placebo-controlled trial. J. Clin. Pharm. Ther. 2003, 28, 53–59. [Google Scholar] [CrossRef] [PubMed]
  190. Turner, R.S.; Thomas, R.G.; Craft, S.; Van Dyck, C.H.; Mintzer, J.; Reynolds, B.A.; Brewer, J.B.; Rissman, R.A.; Raman, R.; Aisen, P.S. A randomized, double-blind, placebo-controlled trial of resveratrol for Alzheimer disease. Neurology 2015, 85, 1383–1391. [Google Scholar] [CrossRef] [PubMed]
  191. Marangoni, D.; Falsini, B.; Piccardi, M.; Ambrosio, L.; Minnella, A.M.; Savastano, M.C.; Bisti, S.; Maccarone, R.; Fadda, A.; Mello, E.; et al. Functional effect of Saffron supplementation and risk genotypes in early age-related macular degeneration: A preliminary report. J. Transl. Med. 2013, 11, 228. [Google Scholar] [CrossRef] [Green Version]
  192. Di Marco, F.; Romeo, S.; Nandasena, C.; Purushothuman, S.; Adams, C.; Bisti, S.; Stone, J. The time course of action of two neuroprotectants, dietary saffron and photobiomodulation, assessed in the rat retina. Am. J. Neurodegene. Dis. 2013, 2, 208–220. [Google Scholar]
  193. Chen, M.; Du, Z.Y.; Zheng, X.; Li, D.L.; Zhou, R.P.; Zhang, K. Use of curcumin in diagnosis, prevention, and treatment of Alzheimer’s disease. Neural Regen. Res. 2018, 13, 742. [Google Scholar]
  194. Akhondzadeh, S.; Fallah-Pour, H.; Afkham, K.; Jamshidi, A.H.; Khalighi-Cigaroudi, F. Comparison of Crocus sativus L. and imipramine in the treatment of mild to moderate depression: A pilot double-blind randomized trial [ISRCTN45683816]. BMC Complement. Altern. Med. 2004, 4, 12. [Google Scholar] [CrossRef] [Green Version]
  195. Boskabady, M.H.; Javan, H.; Sajady, M.; Rakhshandeh, H. The possible prophylactic effect of Nigella sativa seed extract in asthmatic patients. Fund. Clin. Pharmacol. 2007, 21, 559–566. [Google Scholar] [CrossRef] [PubMed]
  196. Wattanathorn, J.; Mator, L.; Muchimapura, S.; Tongun, T.; Pasuriwong, O.; Piyawatkul, N.; Yimtae, K.; Sripanidkulchai, B.; Singkhoraard, J. Positive modulation of cognition and mood in the healthy elderly volunteer following the administration of Centellaasiatica. J. Ethnopharmacol. 2008, 116, 325–332. [Google Scholar] [CrossRef]
  197. Nathan, P.J.; Clarke, J.; Lloyd, J.; Hutchison, C.W.; Downey, L.; Stough, C. The acute effects of an extract of Bacopa monniera (Brahmi) on cognitive function in healthy normal subjects. Hum. Psychopharmacol. Clin. Exp. 2001, 16, 345–351. [Google Scholar] [CrossRef]
  198. Choudhary, D.; Bhattacharyya, S.; Bose, S. Efficacy and Safety of Ashwagandha (Withaniasomnifera (L.) Dunal) Root Extract in Improving Memory and Cognitive Functions. J. Diet. Suppl. 2017, 14, 599–612. [Google Scholar] [CrossRef] [PubMed]
  199. Conte, R.; Calarco, A.; Napoletano, A.; Valentino, A.; Margarucci, S.; Di Cristo, F.; Di Salle, A.; Peluso, G. Polyphenols nanoencapsulation for therapeutic applications. J. Biomol. Res. Ther. 2016, 5, 1000139. [Google Scholar] [CrossRef]
  200. Rigacci, S.; Stefani, M. Nutraceuticals and amyloid neurodegenerative diseases: A focus on natural phenols. Expert Rev. Neurother. 2015, 15, 41–52. [Google Scholar] [CrossRef] [PubMed]
  201. Zhao, D.; Simon, J.E.; Wu, Q. A critical review on grape polyphenols for neuroprotection: Strategies to enhance bioefficacy. Crit. Rev. Food Sci. Nutr. 2020, 60, 597–625. [Google Scholar] [CrossRef]
  202. Renaud, J.; Martinoli, M.G. Considerations for the use of polyphenols as therapies in neurodegenerative diseases. Int. J. Mol. Sci. 2019, 20, 1883. [Google Scholar] [CrossRef] [Green Version]
  203. Hu, S.; Maiti, P.; Ma, Q.; Zuo, X.; Jones, M.R.; Cole, G.M.; Frautschy, S.A. Clinical development of curcumin in neurodegenerative disease. Expert Rev. Neurother. 2015, 15, 629–637. [Google Scholar] [CrossRef] [PubMed]
  204. Rakotoarisoa, M.; Angelova, A. Amphiphilic nanocarrier systems for curcumin delivery in neurodegenerative disorders. Medicines 2018, 5, 126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Lu, Y.; Kim, S.; Park, K. In vitro-in vivo correlation: Perspectives on model development. Int. J. Pharm. 2011, 418, 142–148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Bhattacharjee, S. Dls and zeta potential–what they are and what they are not? J. Control. Release 2016, 235, 337–351. [Google Scholar] [CrossRef] [PubMed]
  207. Ovais, M.; Zia, N.; Ahmad, I.; Khalil, A.T.; Raza, A.; Ayaz, M.; Sadiq, A.; Ullah, F.; Shinwari, Z.K. Phyto-therapeutic and nanomedicinal approaches to cure alzheimer’s disease: Present status and future opportunities. Front. Aging Neurosci. 2018, 10, 284. [Google Scholar] [CrossRef] [Green Version]
  208. Niu, X.Q.; Chen, J.J.; Gao, J.Q. Nanocarriers as a powerful vehicle to overcome blood-brain barrier in treating neurodegenerative diseases: Focus on recent advances. Asian J. Pharm. Sci. 2019, 14, 480–496. [Google Scholar] [CrossRef]
  209. Varma, L.T.; Singh, N.; Gorain, B.; Choudhury, H.; Tambuwala, M.M.; Kesharwani, P.; Shukla, R. Recent advances in self-assembled nanoparticles for drug delivery. Curr. Drug Deliv. 2020, 17, 279–291. [Google Scholar] [CrossRef]
  210. Kalepu, S.; Nekkanti, V. Improved delivery of poorly soluble compounds using nanoparticle technology: A review. Drug Deliv. Transl. Res. 2016, 6, 319–332. [Google Scholar] [CrossRef]
  211. Smith, A.; Giunta, B.; Bickford, P.C.; Fountain, M.; Tan, J.; Shytle, R.D. Nanolipidic particles improve the bioavailability and alpha-secretase inducing ability of epigallocatechin-3-gallate (EGCG) for the treatment of alzheimer’s disease. Int. J. Pharm. 2010, 389, 207–212. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Lv, L.; Yang, F.; Li, H.; Yuan, J. Brain-targeted co-delivery of β-amyloid converting enzyme 1 shRNA and epigallocatechin-3-gallate by multifunctional nanocarriers for Alzheimer’s disease treatment. IUBMB Life 2020, 72, 1819–1829. [Google Scholar] [CrossRef]
  213. Bhatt, R.; Singh, D.; Prakash, A.; Mishra, N. Development, characterization and nasal delivery of rosmarinic acid-loaded solid lipid nanoparticles for the effective management of huntington’s disease. Drug Deliv. 2015, 22, 931–939. [Google Scholar] [CrossRef] [PubMed]
  214. Del Prado-Audelo, M.L.; Caballero-Floran, I.H.; Meza-Toledo, J.A.; Mendoza-Munoz, N.; Gonzalez-Torres, M.; Floran, B.; Cortes, H.; Leyva-Gomez, G. Formulations of curcumin nanoparticles for brain diseases. Biomolecules 2019, 9, 56. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Different biological processes, including oxidative stress and neuroinflammatory and mitochondrial dysfunctions, have been involved in the development and pathogenesis of NDs.
Figure 1. Different biological processes, including oxidative stress and neuroinflammatory and mitochondrial dysfunctions, have been involved in the development and pathogenesis of NDs.
Molecules 26 05327 g001
Figure 2. Therapeutic effects of numerous natural products for combating NDs.
Figure 2. Therapeutic effects of numerous natural products for combating NDs.
Molecules 26 05327 g002
Table 1. Representative natural products and their bioactive substances with neuroprotective activity in the treatment of AD related disease model.
Table 1. Representative natural products and their bioactive substances with neuroprotective activity in the treatment of AD related disease model.
Name of Plant Part Name of Model Neuroprotective Mechanisms References
Yacon (Poepp. and endl.) (Smallanthus sonchifolius) extract of the leaf Rat Memory deficits prevented [162]
Natural safflower aqueous extract Rat Short and long-term memory improved [163]
Methanolic extract of Lactucacapensis thunb. leaves Rat Lowering the degree of lipid peroxidation and protein oxidation [164]
Turmeric powder Human Improvement in the quality of life and behavioral symptoms [165]
Tabernaemontana divaricata root extract Mouse Prevented memory loss [166]
Coconut oil enriched Mediterranean diet Human Enhanced cognitive features [167]
Osmotin, a protein derived from Nicotiana tabacumMouse Increased conduct of random alternation [168]
Germinated brown rice SH-SY5Y cells Reduced production of intracellular ROS [169]
Isolated from Huperzia serrata is Huperzine A Human Improvement in functions of memory, cognition, and actions [170]
Huperzine A isolated from Huperzia serrataRat Reduce oxidative damage [171]
Table 2. Representative natural products and their bioactive substances with neuroprotective activity in the treatment of PD related disease model.
Table 2. Representative natural products and their bioactive substances with neuroprotective activity in the treatment of PD related disease model.
Name of Plant Part Name of Model Neuroprotective Mechanisms References
Smith ethyl acetate extract Zingiber zerumbet (L.) Rat Prevention of neuronal damage [172]
Urticadioica Linn. ethyl acetate fraction. Rat Enhanced motor control and alteration in oxidative protection [173]
Apium graveolens L. Mouse Improved behavioral disorder caused by MPTP [174]
Tribulus terrestris extract Mouse Improved the proportion of viable neurons [175]
Ethanol extract of Tinospora cordifoliaRat Restored locomotive operation behavioral changes caused by 6-OHDA [176]
Dihydromyricetin (DHM) (Ampelopsis grossedentata) Mouse Mitigated the deficit in the balance of movement caused by the MPTP [177]
Agaropentaose, an agaro-oligosaccharide monomer that is isolated from red algae hydrolysates of agarose SH-SY5Y cells Inhibited potential loss of mitochondrial membrane [178]
Capsicum annuum L. extract Mouse Restored development of cholinesterase in the brain [179]
β-Caryophyllene, a cannabinoid compound originating from a plant known as phytocannabinoids Rat Lipid peroxidation inhibited [180]
Viride var. of Coeloglossum. Extract from bracteatum Mouse Prevented neuronal dopaminergic loss [181]
Boswellic acids Rat Motor functions improved [182]
Rosmarinic acid isolated from callus of Perilla frutescensRat Increased tyrosine hydroxylase numbers [183]
Olive leaf extracts (Olea europaea L.) Rat Inhibited tyrosine hydroxylase-positive neuron depletion [184]
Oxalis corniculata extract Mouse Improved preservation and retrieval of memory [185]
Curcuminoids (Curcuma longa (L.) rhizomes) Mouse In the striatum, decreased proinflammatory cytokine and complete nitrite production [186]
Supplementation of fish oil (rich in omega-3 polyunsaturated fatty acids) Rat Reduced loss of substantia nigra neurons and nerve terminals in the striatum) [187]
Germinated brown rice Rat Improved the number of dopaminergic neurons that survive [188]
Table 3. Numerous medicinal plant extracts used in clinical trials and their outcomes.
Table 3. Numerous medicinal plant extracts used in clinical trials and their outcomes.
Plant Species Type of Clinical Study Clinical Outcomes Reference
Salvia officinalisRandomized, double-blind Significantly improved cognitive function[189]
ResveratrolRandomized, placebo-controlled, double-blind, multicenter 52-wk phase 2 trialResveratrol was safe and well-tolerated. Resveratrol and its major metabolites penetrated the blood–brain barrier to have CNS effects [190]
Ginkgo biloba L. Longitudinal, 3 monthly follow-ups over a 12-month periodFocal electroretinograph↑ amplitude and sensitivity amplitude that stabilized after 3 months independent of genotype[191]
Crocus sativus L. Longitudinal, open-label study, 8 monthly follow-ups over a 29 (±5)-month periodFocal electroretinograph saffron treated age-related macular degeneration patients: Visual function remained stable [192]
Curcumin longa24 older adults with physical or cognitive impairmentImprove physical function and cognitive function[193]
Crocus sativusDepressant patientsThe effect of C. Sativus similar to imipramine in the treatment of mild to moderate depression[194]
Nigella sativaAsthmaticpatientsImprovement of all asthmatic symptoms, chest wheeze and pulmonary function test values[195]
Centella asiaticaRandomized, double-blind placebo-controlled trial Improved memory function[196]
Bacopa monnieriDouble-blind, placebo-controlled trial in 38 healthy volunteers (aged 18–60 years) Significantly improved cognitive function[197]
Withania somniferaProspective, randomized, double-blind, placebo-controlled Significantly improved executive functions in adults with mild cognitive impairment[198]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rahman, M.H.; Bajgai, J.; Fadriquela, A.; Sharma, S.; Trinh, T.T.; Akter, R.; Jeong, Y.J.; Goh, S.H.; Kim, C.-S.; Lee, K.-J. Therapeutic Potential of Natural Products in Treating Neurodegenerative Disorders and Their Future Prospects and Challenges. Molecules 2021, 26, 5327. https://doi.org/10.3390/molecules26175327

AMA Style

Rahman MH, Bajgai J, Fadriquela A, Sharma S, Trinh TT, Akter R, Jeong YJ, Goh SH, Kim C-S, Lee K-J. Therapeutic Potential of Natural Products in Treating Neurodegenerative Disorders and Their Future Prospects and Challenges. Molecules. 2021; 26(17):5327. https://doi.org/10.3390/molecules26175327

Chicago/Turabian Style

Rahman, Md. Habibur, Johny Bajgai, Ailyn Fadriquela, Subham Sharma, Thuy Thi Trinh, Rokeya Akter, Yun Ju Jeong, Seong Hoon Goh, Cheol-Su Kim, and Kyu-Jae Lee. 2021. "Therapeutic Potential of Natural Products in Treating Neurodegenerative Disorders and Their Future Prospects and Challenges" Molecules 26, no. 17: 5327. https://doi.org/10.3390/molecules26175327

Article Metrics

Back to TopTop