Next Article in Journal
Non-pharmacological Treatment of Refractory Angina and Microvascular Angina
Next Article in Special Issue
Revisiting Traumatic Brain Injury: From Molecular Mechanisms to Therapeutic Interventions
Previous Article in Journal
Use of Oral Anticoagulation and Diabetes Do Not Inhibit the Angiogenic Potential of Hypoxia Preconditioned Blood-Derived Secretomes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Neuroprotection: Targeting Multiple Pathways by Naturally Occurring Phytochemicals

by
Andleeb Khan
1,*,
Sadaf Jahan
2,
Zuha Imtiyaz
3,
Saeed Alshahrani
1,
Hafiz Antar Makeen
4,
Bader Mohammed Alshehri
2,
Ajay Kumar
5,
Azher Arafah
6 and
Muneeb U. Rehman
6
1
Department of Pharmacology and Toxicology, College of Pharmacy, Jazan University, Jazan 45142, Saudi Arabia
2
Medical Laboratories Department, College of Applied Medical Sciences, Majmaah University, Majmaah 15341, Saudi Arabia
3
Clinical Drug Development, College of Pharmacy, Taipei Medical University, Taipei 11031, Taiwan
4
Department of Clinical Pharmacy, College of Pharmacy, Jazan University, Jazan 45142, Saudi Arabia
5
Institute of Nano Science and Technology, Habitat Centre, Phase-10, Sector-64, Mohali 160062, India
6
Department of Clinical Pharmacy, College of Pharmacy, King Saud University, Riyadh 11451, Saudi Arabia
*
Author to whom correspondence should be addressed.
Biomedicines 2020, 8(8), 284; https://doi.org/10.3390/biomedicines8080284
Submission received: 15 June 2020 / Revised: 27 July 2020 / Accepted: 5 August 2020 / Published: 12 August 2020

Abstract

:
With the increase in the expectancy of the life span of humans, neurodegenerative diseases (NDs) have imposed a considerable burden on the family, society, and nation. In defiance of the breakthroughs in the knowledge of the pathogenesis and underlying mechanisms of various NDs, very little success has been achieved in developing effective therapies. This review draws a bead on the availability of the nutraceuticals to date for various NDs (Alzheimer’s disease, Parkinson’s disease, Amyotrophic lateral sclerosis, Huntington’s disease, vascular cognitive impairment, Prion disease, Spinocerebellar ataxia, Spinal muscular atrophy, Frontotemporal dementia, and Pick’s disease) focusing on their various mechanisms of action in various in vivo and in vitro models of NDs. This review is distinctive in its compilation to critically review preclinical and clinical studies of the maximum phytochemicals in amelioration and prevention of almost all kinds of neurodegenerative diseases and address their possible mechanism of action. PubMed, Embase, and Cochrane Library searches were used for preclinical studies, while ClinicalTrials.gov and PubMed were searched for clinical updates. The results from preclinical studies demonstrate the efficacious effects of the phytochemicals in various NDs while clinical reports showing mixed results with promise for phytochemical use as an adjunct to the conventional treatment in various NDs. These studies together suggest that phytochemicals can significantly act upon different mechanisms of disease such as oxidative stress, inflammation, apoptotic pathways, and gene regulation. However, further clinical studies are needed that should include the appropriate biomarkers of NDs and the effect of phytochemicals on them as well as targeting the appropriate population.

Graphical Abstract

1. Introduction

Various conditions affecting nerve cells and the nervous system due to the loss of neurons and their connecting networks are described under the superordinate phrase “Neurodegenerative diseases”. They lead to disability due to gradual neuronal death in both the central nervous system (CNS) and the peripheral nervous system (PNS). Several diseases are genetic with few coming being caused due to the exposure to various toxins and chemicals. The main symptoms associated with these disorders are related to movements (ataxia) or mental functioning (dementia) or both causing morbidity and death. This fostered the profound social and economic implications [1]. The most common NDs include cognitive and behavioral disorders, Alzheimer’s disease (AD), Parkinson’s disease (PD), Amyotrophic lateral sclerosis (ALS), Huntington’s disease (HD), Spinocerebellar ataxia (SCA), Spinal muscular atrophy (SMA), Vascular cognitive impairment Prion disease, Frontotemporal disease, Pick’s disease, etc. The treatment available for these disorders gives only symptomatic relief to the patient by extending the lifespan to a few years [2]. Still, a lot of research is in progress to find the therapeutic markers in these diseases [3].
There has been an extensive approach to non-pharmacological therapies to combat the ill effects of NDs. Many of them include disease-modifying therapies, non-invasive brain stimulation techniques, physical exercise, adaptive physical activity, complementary and alternative medicine, and several nutraceutical compounds. In a meta-analysis, it was reported that repetitive transcranial magnetic stimulation (rTMS) therapy with Parkinson’s patients resulted in mild to moderate improvement of motor activities [4]. In another study, it was reported that both rTMS and transcranial direct current stimulation (tDCS) have shown improvement in the cognitive performance of Alzheimer’s patients [5]. Fisicaro and colleagues stated in their review that rTMS can be effectively used as a non-pharmacological tool for various motor and non-motor neurorehabilitation. They suggested that this technique along with other conventional rehabilitative modalities can be effectively used in clinics but still the regime is not clear [6]. Physical exercise has been a promising therapy for motor deficits for years. It could have therapeutic potential for the prevention of mental disorders and neurodegeneration. Extensive research to explain different molecular pathways involved in this therapy is under progress, suggesting multiple pathways may be involved together in neuroprotection [7,8]. The major cornerstones in managing ataxic patients are physiotherapy and kinesitherapy in the present scenario. Studies confirm that customized programs for various exercises of coordination, balance, cognitive skills, and posture maintenance have shown extensive improvement along with the conventional pharmacological therapeutics [9]. Complementary and alternative medicines have always attracted various disease prevention including NDs [10,11]. In addition to these, the various nutraceuticals [12] and phytochemicals multi-target approaches have proved to be a promising adjunct to the present treatment [13].
The large number of pharmacological or biological activities of the phytochemicals have made them appropriate candidates for the treatment of the NDs [14]. They exhibit antioxidant (resveratrol, zingerone), anti-inflammatory property (cineole, thymoquinone), inhibitors of gamma-aminobutyric acid (GABAA) receptors (diterpenes and cyclodepsipeptides), and Monoamine oxidase-B (MAO-B) receptors (selegiline rasagiline), although not proven clinically. Combating disease symptoms by phytochemicals and herbal nutraceutical is known in traditional medicines, and several new research for preventing NDs by phytochemicals are under progress [15]. In this review, our focal point is to discuss the recent advancements in the field of neuroprotection related to the NDs by various phytochemicals and to elicit their potential mechanism of actions. In particular, we discuss phytochemicals used in the prevention of all kinds of NDs and related signaling pathways and mechanism of recovery highlighting the role of phytochemicals. The basic structures of most the common neuroprotective phytochemicals are given in Figure 1.

2. Methodology

A Medline (PubMed), Cochrane Library and Embase based literature survey was performed by using keywords of “neurodegenerative diseases, prevention, non-pharmacological therapies, and neurodegenerative diseases, phytochemicals, Alzheimer’s disease natural products, Parkinson’s disease (PD) phytochemicals, Huntington’s disease phytoprotection, natural product amyotrophic lateral sclerosis, phytochemicals cerebral ischemia, phytochemicals prion disease, spinocerebellar ataxia prevention by natural products, natural product and spinal muscular atrophy, frontotemporal dementia natural products, pick’s disease, and natural products” till June 2020.
Two independent authors (A.K. and S.J.) screened all the titles and abstracts of the retrieved data, disagreements were resolved by the consensus of a third author (Z.I.). Duplicated entries, retracted publications, studies on other diseases or conditions different from NDs or its subtypes, studies without statistical analysis, non-English written papers, publications that are not research studies (i.e., commentaries, letters, editorials, reviews, meta-analysis), and any other article that did not fit within the scope of this review, were excluded. Articles listed in the references were also reviewed in search of more data.
A total of 873 results were retrieved and screened with the above keywords. Out of these 115 publications were selected and eventually used for qualitative analysis (Figure 2) and the same are summarized in Table 1. In more detail, 28 studies work with AD, 17 with PD, 7 with ALS, 8 with HD, 20 with VCI and stroke, 4 with prion, 7 with FTD, 2 with SMA, and 1 with SCA and others on non-pharmacological therapy and neurodegenerative diseases.

3. Multifunctional Phytochemicals as Novel Therapeutic Agents for Neurodegenerative Disorders

3.1. Alzheimer’s Disease and Related Dementias

AD is a multifactorial, pernicious disease with genetic and environmental factors leading to irreversible cognitive impairment [17]. The main neuropathological features of AD comprises of neuronal loss, amyloid-β plaques, and neurofibrillary tangles in limbic and neocortical regions [18]. Various mechanisms are associated with the progression of the disorder which includes amyloid deposition, tau hyperphosphorylation and aggregation, loss of cholinergic system, oxidative stress, inflammation, apoptosis, glutamatergic excitotoxicity, and decrease in neurogenesis and neurotrophic factors [17,19,20]. In recent years, approved drugs for AD are only modulators of cholinergic and glutamatergic systems, which can only delay the symptoms to some extent [21]. Divergent antioxidant and anti-inflammatory medications are also approved for AD patients [22]. There are recent reports that herbal products and nutraceuticals can act as a multi-target approach as suggested in the traditional system of medicines [23]. Recently extensive research has been conducted which supports their potential in in vitro, in vivo, and clinical studies. In the list of natural compounds, curcumin is considered on the top because of its miraculous qualities. In this section, the mechanistic approach of the curcumin and other important natural compounds will be discussed. Curcumin is a polyphenolic compound and a natural herb which is an integrative part of every Indian kitchen. In literature, many beneficial properties are reported like anti-inflammatory, anti-cancerous, anti-apoptotic, antioxidative, neuroprotective [24]. Apart from that curcumin is also reported as a therapeutic agent to treat rheumatoid disorders, cough, and neurological disorders. Fortunately, it has intensive effects and have multiple sites of action [25]. The relationship between curcumin consumption and a lower prevalence of AD has been observed. In one study, curcumin emulates the improvement of the spatial learning and memory in the rat model and it is reported as an inhibitor of BAC1 in in vitro study. Curcumin has a promising effect in neuroblastoma SHSY5Y cell lines and also has a therapeutic effect on the AD-induced mechanisms [26]. Acrolein exposed HT22 murine hippocampal neuronal cells were protected by curcumin by following the BDNF/TrkB signaling [27]. Orally administrated curcumin downregulates the level of GSK3β and also inhibit the hippocampus Aβ plaques with the improved cognitive impairment and improved maze test in in vivo study [28]. Curcumin effectively counteracted the p25-mediated glial activation and pro-inflammatory chemokines/cytokines production in p25Tg mice. Moreover, this curcumin-mediated suppression of neuro-inflammation reduced the progression of p25-induced tau/amyloid pathology and in turn ameliorated the p25-induced cognitive impairments [29]. Therefore, through the series of research, curcumin is found to be promising phytochemicals against Aβ accumulation and would be a determined remedy against AD. Pilot study with curcumin and Ginkgo on 34 possible or probable AD patients was carried out with curcumin and Ginkgo extract. It was found that no side effects of curcumin along with that mini-mental examination scores or plasma amyloid beta40 level between 0 and 6 months [30]. AD-related pathology in the brain is well documented, the disease has also been reported to affect the retina, a developmental outgrowth of the brain, which is more accessible for imaging. Early detection of AD, utilizing curcumin fluorescence imaging to highlight Aβ aggregates in the retina. Results show that retinal fluorescence imaging is justified with brain plaque burden from PET scan [31]. Apigenin, a known phytochemical also targets the AD by its antioxidative effects. In plants, Apigenin is found as aglycon and more often in glycoside form. Apigenin inhibits the copper-mediated β-amyloid neurotoxicity (copper-induced) via antioxidative mechanism followed by MAPK signal inactivation in an AD cell model [32]. The Apigenin follows the ERK/CREB/BDNF pathway for neuroprotection [33]. On the list of phytochemical-induced mechanistic pathways, genistin also has an important role in neuroprotection. It is a component of the soya bean meal. The effect of genistin (100 μM) was studied on the in vitro culture of rat neurons [34]. It was observed that the amyloid-beta peptide instigated the damage of mitochondrial membrane potential, nuclear DNA damage, high expression of Bcl-2 associated X protein (BAX), p53, Caspases and lower expression of B-cell lymphoma-2 (Bcl-2) [35]. The 100 µM dose of genistin was found to be efficient against the deleterious effects of AD and restores the neuronal activity by following the caspase signaling pathways [36]. It is also reported to lower calcium overload and reverse the low fluidity of the neuronal membranes in the AD model. Free radicles were also reduced with the use of genistin and thus it is suggested for the targeting of anti-ROS (reactive oxygen species) pathway and protect the cells from death. Astrocytes are also studied for the efficiency of genistin [37]. Pretreatment of 5 μM concentration of genistin was used against astrocytes and it protects by working on interleukin 1 beta (IL-1β), tumor necrosis factor-α (TNF-α), cyclooxygenase-2 (COX-2), inducible nitric oxide synthase (iNOS) cascade, and reversing the effect of NDs in the AD model [38]. In another study, 50 μM concentration of genistin was effective and initiate IL-1β /TLR4/NF-κB/ IκB-α cascade against neuroinflammation induced by Aβ in the secondary culture of C6 [39]. It can downregulate TNF-α, IL-1β, TLR4, and NF-κB and promote the upregulation of IκB-α. Other phytochemicals are also having miraculous properties against many NDs specifically AD [40]. Genistein protects Aβ-induced neurotoxicity by inducing the PKC signaling pathway, which further regulates the activities of α- and β-secretase and thereby inhibits the formation and toxicity of Aβ [41]. Genistein also protected the hippocampal neurons against injury by calcium/calmodulin dependent protein kinase IV protein levels in AD model [42]. Ginkgo biloba is also considered as a neuroprotective agent. Ginkgolide A inhibits Aβ-induced neuronal degradation via JNK phosphorylation and NMDA/AMPA-induced depolarization, thus suppress the activity of plaque [43]. EGb 761 increases the neurogenesis in the adult hippocampus region and CREB phosphorylation in the transgenic mouse model of AD [44]. In one study, G. biloba demonstrated protection against a high dose of Bisphenol A (BPA) and played a role in the improvement of cognitive deficits [45]. A flavonoid, luteolin, would be discussed against the AD model as a possible therapeutic agent. In an in vivo study (rat model), it protects against the cognitive dysfunction which was induced by chronic cerebral hypoperfusion [46]. It is also found to be protective against high-fat diet-induced cognitive abnormalities. GSK 3α isoform is involved in the mechanistic signaling. The inactivation of GSK 3α isoform phosphorylates the PS1 which is the catalytic core of the γ-secretase complex causing a decrease in PS1-APP interaction and generation of Aβ. In literature, the role of luteolin against zinc-induced hyperphosphorylation of the protein τ in SH-SY5Y secondary cell culture is reported [47]. Luteolin is suggested as a promising candidate because of its antioxidative properties as well. In the AD animal model, luteolin slows down the escape latency and traveled distance parameters in the morris water maze while increasing the time spent in the target quadrant [48]. Melatonin modulates the proteins (i.e., GSTP1 and CPLX1) that are involved in depression and anxiety. The modulation improves the neuropsychiatric behaviors in the AD model [49]. Melatonin improves anxiety and depression-like behaviors and targets the proteomic changes in the triple transgenic AD mice model [50]. A multicenter, randomized, placebo-controlled clinical trial of 2 dose formulations of oral melatonin coordinated by the National Institute of Aging-funded Alzheimer’s Disease Cooperative Study, and it was found that melatonin was effective against AD [51]. Naringenin protects Aβ25-35-caused damage via ER and PI3K/Akt-Mediated Pathways [49,52]. Naringenin also protects AlCl3/D-galactose-induced neurotoxicity in the rat model of AD via inhibition of attenuation of acetylcholinesterase levels and attenuation of oxidative stress [53]. A new study suggested the therapeutic effects of piperine against oxidative insult, neuroinflammation, and neurochemical changes induced by ICV-STZ infusion in mice [54]. In the human brain, microvascular endothelial cells from fibrillar β-amyloid are protected by quercetin [55]. The flavonoid quercetin improves Alzheimer’s disease pathology and protects cognitive and emotional function in aged triple transgenic Alzheimer’s disease model [56]. in vitro AD model, resveratrol works on oxidative damage via activating mitophagy [57]. Resveratrol is found to be active against the harmful process that occurs in AβPP/PS1 mouse hippocampus and inhibit memory loss [58]. For the AD model, resveratrol is studied in randomized, double-blind, placebo-controlled, phase II trial. Resveratrol was found to be safe, effective, and can protect the BBB integrity with CNS response [59]. Thymoquinone inhibits the Aβ-induced neurotoxicity by blocking the mitochondrial dysfunction and oxidative stress [60]. S-allyl cysteine also targets the oxidative stress-induced cognitive neuronal impairment and provide possible rescue from the diseased [61]. In mice model treated with D-galactose, S-allyl cysteine involves in the alleviation of β-Amyloid, oxidative damage [62]. Eucalyptol ameliorates inflammation induced by Aβ(25–35) in differentiated PC12 cells by following the NOS-2, COX-2, and NF-κB associated pathways [63]. Silibinin and silymarin administration could recover memory impairment and could decrease Aβ plaque in the brain of APP/PS1 mice [64]. By following the BDNF/TrkB pathway, Silibinin ameliorates anxiety/depression-like behaviors in amyloid β-treated rats [65]. From all the above discussion of the phytochemicals, it is imminent that researchers would be able to hit the specific targets for providing promising therapeutic relief against AD Figure 3. The details of several phytochemicals in AD protection are enlisted in Table 1.

3.2. Parkinson’s Disease and Parkinsonism

PD is the second most prevalent NDs of the elderly after AD, characterized by dopaminergic neuronal loss in substantia nigra pars compacta region of the brain, the striatum, locus ceruleous, raphe nuclei, nucleus basalis of meynert, and hippocampus. The main sign and symptoms of PD patients are resting tremors, rigidity, bradykinesia, muscular impairment, postural instability, difficulty in walking, soft speech, sleep problems, slow handwriting [127]. The pathological findings are the presence of Lewy Bodies (LBs, aggregation of α-synuclein) with DNA mutations in the brain of PD patients [128], with neuronal cytoplasmic inclusions in the cortex and brain stem [129,130]. The treatment of PD mainly rotates around the enhanced activities of dopaminergic neurons or inhibition of cholinergic effects [131]. Several mechanisms associated with PD accounts for oxidative stress, mitochondrial dysfunction, apoptotic pathway activation etc. [132,133] There is no cure for PD but drugs available in the market increases the life expectancy with relief in the symptoms [131]. Besides, the drug therapy demonstrates many side effects in patients [134,135]. Therefore, researchers have to treat PD at molecular and cellular levels, and use such treatments that have a multifactorial approach. Several scientists have reported plant compounds to be effective in the treatment of PD. After dissecting the oxidative stress role in the pathophysiology of PD, the phytochemicals having antioxidative properties are drawing the attention of researchers [136]. This section deals with the phytochemicals which are generally considered as antioxidative agents. In many fruits and vegetables (apples, broccoli, and onions), a polyphenolic compound is detected named quercetin. It restricts the hippocampal neuronal damage resulting in improved memory and learning (by maze test) in the rat model [137]. In one study, when quercetin was administrated to the PD model, it reversed the neuronal impairment via antioxidative pathways [138]. In PD MitoPark transgenic mouse model, quercetin protects against mitochondrial dysfunction and progressive dopaminergic neurodegeneration [66]. Quercetin also plays a specific role in the upregulation of cognitive effects in the 6-hydroxydopamine (6-OHDA) rat PD model [67]. An isoflavone, genistin is also having several health benefits including nourishment of neurons [139]. It is known as antioxidative, anti-inflammatory, and neuroprotective agent. Genistin (oral dose 10, 50, or 100 mg/kg for one week) acts as a therapeutic agent against LPS-induced cognitive impairment. These dose-dependent treatments of genistin determines to lessen memory deficits, decrease the malondialdehyde levels in the hippocampus region, increases the level of superoxide dismutase, catalase, and glutathione [140]. Moreover, genistin reformed the level of AChE activity in the hippocampus against LPS insult in in vivo model. It targets the modulation of Nrf2/NF-κB/IL-6/TNFα/COX2/iNOS/TLR4/GFAP and thus promote neuronal protection via inhibiting the neuronal inflammation and improve cognitive deficits [140]. Genistein shows the neuroprotection in the A53T mutant α-synuclein overexpressed SH-SY5Y cells [68]. Genistin protects the Transgenic Drosophila Model of Parkinson’s Disease against oxidative stress [69].
Other phytochemical Naringin was studied and was found to be involved in the neuroprotection by triggering the Nrf2/ARE signaling against 6-OHDA-induced neurotoxicity [141]. In in vitro study (neuroblastoma SH-SY5Y cells), it is reported to show neuronal protection by blocking the c-Jun N-terminal kinase (JNK) phosphorylation and regulation of BAX in the rotenone-induced cell line [142]. Additionally, it also inhibits the caspase-3 and PARP cleavage [143]. Naringin can lead to the anti-inflammatory pathway and provide neuroprotection in PD [70]. In in vitro PD model of SH-SY5Y cells, curcumin imparts a role in neuroprotection against toxic insult via controlling the HSP-90 [71]. Curcumin was studied in Pheochromocytoma Cell Line 12 (PC12) cells against 1-methyl-4-phenylpyridinium (MPP+) toxicity and fixation of mitochondrial membrane potential was found against enhancement of intracellular ROS which leads to the instigation of iNOS, and over-expression of Bcl-2 [144]. Curcumin leads to the restoration of mitochondrial membrane potential (MMP) which causes the up-gradation in Cu/Zn SOD and modulates NF-κB nuclear translocation by inhibition of interleukin-6 (IL-6) and TNF-α [145]. It was further reported that it lowers the iNOS mRNA levels and inhibits the induction of NF-κB. It is found to be effective against oxidative stress via following cytochrome c release or caspase-3 activation. In in vitro study, it is reported to act on the aggregation of synuclein inhibition and aggregation of the PD associated, A53T mutant-synuclein in SH-SY5Y cells in a dose-dependent manner [146]. By inhibiting the α-synuclein aggregation in lipopolysaccharide-induced PD model curcumin plays a role as a promising therapeutic agent [72]. Epigallocatechin-3-gallate (EGCG) is also an important bioactive component with neuroprotective potential. The mode of action of EGCG is on neuroinflammation, oxidative stress, and autophagy. Likewise curcumin, EGCG also ameliorates the mitochondrial restoration by antioxidative properties and associated signaling pathways i.e., via blocking the expression of COX-2, iNOS, and all other enzymes which causes the generation of pro-inflammatory mediators [147]. EGCG protects by antioxidative activity against 6-OHDA administration in SH-SY5Y cells (in vitro dopaminergic model) [148]. EGCG also functions as antagonistic against the NF-κB transcription factor and participates to remove the poisonous accumulation of amyloidogenic α-synuclein in many in vitro models. 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) treated mice also show improvement when EGCG is administrated orally [149]. EGCG also inhibits the oxidative stress in rotenone-induced in vitro model of SH-SY5Y cells [73]. In the MPTP-induced mouse model of PD EGCG helps in the modulation of peripheral immunity [74]. A natural polyphenolic compound, resveratrol, is involved in the mitochondrial biogenesis by initiating the antioxidative cascade mechanism. It also protects neurotoxic chemicals i.e., 6-OHDA and MPTP in in vitro models [150]. This natural compound shows the promising cure against cell toxicity induced by dopamine (DA) and reverse the neuronal impairment by following the cAMP/PKA pathway and thus inhibit the α-synuclein aggregation [151]. Further study also points out the SIRT1/AMPK/PGC1-α axis as a key neuroprotective pathway and provides a rationale for exploring the therapeutic potential of resveratrol in delaying PD progression [152]. By leading the CXCR4 signaling pathway, resveratrol helps in the restoration of the 6-OHDA-induced damage in the PC 12 cells [75]. Resveratrol attenuates the PD like condition via inhibition of programmed cell death through inducing MALAT1/miR-129/SNCA signaling pathway [76]. A total of 119 participants were randomized to placebo or pure synthetic resveratrol 500 mg orally once daily with a dose-escalation by 500-mg increments every 13 weeks until a final dose of 1000 mg twice daily was reached for the final 13 weeks. Visits occurred at screening, baseline, and every 6 weeks with a total resveratrol exposure of 52 weeks. Compliance with resveratrol and with placebo was confirmed by mass spectrometry of blood samples. Resveratrol may have many other molecular effects, including anti-inflammatory, antioxidant, and anti-Aβ aggregation. 480 PD patients, randomized, double-blind three dosage groups of green tea polyphenol and one placebo control group. Delay in motor function progression, increased cognition, mood, and quality of life was reported [153]. 20 double-blind, crossover, randomized, placebo-controlled phase 1 study BIA 6-512 (trans-resveratrol) 25 mg dose, 50 mg dose, 100 mg dose To study BIA 6-512 effect on levodopa pharmacokinetics when administered with levodopa/benserazide [154]. Rutin, another phytochemical act as an anti-inflammatory agent for neuroprotection which is associated with the involvement of microglia response, cytokines level, and iNOS activation [155]. Rutin facilitates protection in 6-OHDA-induced neurotoxicity in rat pheochromocytoma (PC-12) cells [77]. Rutin also targets the oxidative stress and protects dopaminergic neurons in the animal model of PD [78]. Hesperidin can be a possible promising guide for the designing of inhibitors of mutant SOD1 for therapeutic purposes. Antioxidative and anti-inflammatory properties make hesperidin reliable herbal compound against NDs. ERK/Nrf2 are key signaling molecules for following the neuroprotective pathway [156]. Neurotoxicity induced by 6-OHDA can be suppressed by the use of hesperidin in the mice model of PD [79]. Piperine treatment attenuated rotenone-induced motor deficits, and rescued the loss of dopaminergic neurons in the substantia nigra [157]. Piperine stimulates the anti-apoptotic and anti-inflammatory pathway and shows protection in 6-OHDA-induced PD rat model [80]. PD is known as the second rank on the list of NDs with poor availability of effective treatments still, the studied phytochemicals have given hope for promising and long-lasting therapy. Figure 4. The briefing of phytochemicals and their therapeutic effects on PD are listed in Table 1.

3.3. Amyotrophic Lateral Sclerosis

ALS also called motor neuron disease (MND) or Lou Gehrig’s disease is another NDs that is heterogeneous and both upper and lower motor neurons are degenerated causing motor symptoms. The initial symptoms vary between individuals, some exhibiting spinal-onset disease with muscle weakness of limbs, and others present with bulbar-onset disease in which they have dysarthria and dysphagia [158]. The preponderance of patients face death within 3–5 years after the first sign of disease due to respiratory failure [159]. The cause of this diseases is unknown but few cases are linked to familial history with gene mutations as a major contributor [160] and associated mechanisms of oxidative stress [161], calcium toxicity [162], inflammation [163], chronic viral infections [164], and excitotoxicity [165]. Other symptoms are spasticity, sialorrhoea, pain, muscular cramps, liver problems, deep venous thrombosis, mood alterations, respiratory insufficiency, fever, vomiting, fatigue, and nausea. In the case of ALS, the ability of the brain to convert the toxic radicals into non-toxic substances gets compromised because of mutations in SOD1 [166]. Although, this kind of mutation does not lead to the progression of the disease but the mutation makes the SOD1 more prone to the misfolding which causes the accumulation of protein and thus involves the reduction in the activity of motor neurons. The treatment of ALS is often symptomatic with Riluzole (Rilutek) and Edaravone (Radicava) as the only drugs approved by the FDA. These drugs can only increase the life expectancy by three to six months. Much of the efforts of physicians and other health workers is to make the quality of life (QOL) better for ALS patients. Several studies on natural plants and their active compounds are in progress to combat the dreadful symptoms of this disease. The molecular mechanisms and potential therapeutic role of the plant compounds have to be studied and implemented.
Few phytochemicals and their possible known mechanism of actions will be discussed in this section. Ginseng root shows the protective effect in SOD-1 (G93A) transgenic mice [81]. An isoflavone, genistin can easily cross the blood–brain barrier (BBB) and attributes its role by anti-inflammatory activity via extracellular-signal-regulated kinase (ERK) and microtubule-associated protein kinase (MAPK) signaling in the microglia which leads to the neuronal inflammation. The neuroprotective role of genistin was detected in the SOD1-G93A mice [82]. Ginseng roots inhibit the onset and increase the survival of affected cells delays the onset and prolongs the survival of affected cells [167]. The most promising phytochemical, curcumin has a protective role in ALS too. It acts for mitochondrial biogenesis leading the AMP/PGC-1α and/or Nfr2 pathway [83]. The clinical studies show the effectiveness of curcumin for slowing down the progress of the disease. Curcumin helps in the improvement of aerobic metabolism and oxidative stress. The antioxidative mechanism is responsible for leading the protective effect against ALS [83]. A studied phytochemical named withaferin A shows an anti-inflammatory effect by working as antagonistic of NF-κB and is found to be effective for the improved motor activity in transgenic mice model [84]. ALS mouse model generated by the point mutation in SOD1 was studied and it was observed that by the withaferin treatment there was a late progression of the disease and more survival time in comparison to the non-treated ALS mice. The root extract of Withania was also tested against mutated transgenic mice (TDP-43A315T mice) and manifested with removal of abruptly cytoplasmic motor neurons, reduction in inflammatory markers, and improvement in cognitive performance [168]. It is reported that ALS is following numerous pathways therefore, combined therapy could be more efficient for targeting more than one signaling pathway within a single attempt. EGCG has a power of antioxidation, hampered activation of caspase-3, NF-κB and thus reduces microglial activation resulting in decreased neuroinflammation and increased neuronal survival [169]. EGCG stimulates the iNOS/NF-κB/ Caspase-3 and shows the increased number of neuronal cells, lower microglial activation in transgenic mice model [85]. One more phytochemical named celastrol also reveals the anti-inflammatory, antioxidative property. Celastrol exhibits an antioxidative effect in SOD1G93ANSC34 cells by leading MEK/ERK and PI3K/Akt signaling pathways [86] Celastrol protects the H2O2-induced oxidative stress model by activating the antioxidative pathway in ALS model [86]. The bran oil treatment helps in the significant enhancement of antioxidative enzymes and suppress the effect of ROS and malondialdehyde (MDA) in rotenone-induced ALS model [170].
ALS study was also conducted in bone marrow-derived mesenchymal stem cells (MSCs) by activating the SIRT1/AMPK pathway and it is noted a significant decrease in the ALSMSCs compared to normal healthy control originated BM-MSCs by using resveratrol. Neuro-progenitor markers were increased in resveratrol treated ALS-MSCs [87]. All these phytochemicals along with the ones enumerated in the table could be promising therapeutic agents which can be the best alternative of commercial drugs with negligible side effects Table 1.

3.4. Huntington’s Disease

HD is an annihilative, inherited, and familial illness, with progressive loss of brain and muscle functions. This is a single-gene disorder in which there is programmed degeneration of neurons within different regions of the brain. In wild type gene, there are 6-35 CAG repeats in exon 1 of the Huntingtin (Htt) gene on chromosome number 4 whereas, in the affected persons, this repeat increases to more than 36 [171]. As a result of this, there is an accumulation of Huntingtin protein in the neuron causing its death. Almost 95% of GABAergic medium spiny neurons (MSNs) projecting to substantia nigra and globus pallidus are lost causing atrophy of cortex, thalamus, and hypothalamic nuclei. The symptoms include concentration problems, short-term memory, tumbling, lack of focus, clumsiness, depression weight loss, feeding problems, difficulty in speech, uncontrolled face movements, itching etc. [172]. There is a lot of research going on in the field of HD prevention and treatment. Many experimental models have been experimentally validated to test the efficacy of different drugs. Hitherto, there is no cure for the disease, drugs available only give symptomatic relief. There have always been natural compounds linked to various diseases and a lot of them have shown promising effects in preclinical studies.
One of the major factors for the progression of HD is the abrupt function of mitochondria [173]. The activity of complex II of the respiratory chain gets reduced in the affected brain region. Protopanaxtriol limits the overproduction of free radicals and attenuates the expression of heat shock protein 70 (Hsp-70), along with the restoration of SOD activity inhibiting the formation of free radicals. Protopanaxtriol enhances the Nrf2 entry to the nucleus and increases the expression of Heme oxygenase-1 (HO-1) and NAD(P)H dehydrogenase [quinone]1(NQO1) [88]. Melatonin also has been shown to reduce neural damage in PD, AD as well as in ischemia-reperfusion injury against d-aminolevulinic acid and a variety of other neural toxins. In the case of mitochondrial injury, induced by 3-nitropropionic acid there was a restoration by the treatment with melatonin by leading the antioxidative pathway [88,174]. Melatonin also rescues the HD animal model from the oxidative stress induced by 3-nitropropionic acid [89]. Resveratrol attenuates the expression of ATG4 and permits the lipidation of LC3 and promotes the mortification of the polyQ-Htt accumulation and saves the neuronal cells against dopamine-induced toxicity [90]. EGCG inhibits the accumulation of misfolded proteins and oligomerization of mutant Htt exon 1 protein in vitro, which indicates the early interference of EGCG in the misfolded protein accumulation [175]. In another study, EGCG reduced the cytotoxicity and polyQ-mediated htt protein aggregation in a yeast model of HD [176]. EGCG attenuates the adverse effect of misfolded huntingtin protein in early-stage in HD model [91].
One more phytochemical, trans-ε-Viniferin (viniferin), was also found to be effective against mutant Htt-induced depletion of SIRT3 and thus save the cells from a mutant protein [177]. Viniferin also plays a crucial role in reducing the level of free radicals and thus participate in the protection of MMP in the cells having mutant Htt. The presence of mutant Htt indicates the lower deacetylase activity of SIRT3 which leads to decreased cellular NAD levels with the biogenesis of mitochondrial cells [178]. Viniferin enhances the mitochondrial biogenesis by activation of AMP-activated kinase [92]. It was reported that systemic 3-NP administration significantly enhances the level of lipid peroxidation (LPO), nitrite and lactate dehydrogenase (LDH), downregulate the antioxidant enzyme (superoxide dismutase and catalase) levels, and inhibit the synthesis of ATP by blocking the activity of mitochondrial complex in the different regions (striatum and cortex) of the brain [179]. Figure 5. The root extract of Withania somnifera is used against 3-NP-induced gait abnormalities, mitochondrial dysregulation, and oxidative stress, in striatum and cortex of the rat brain, in vivo model [93]. The detailed functions of phytochemicals are well explained in Table 1.

3.5. Vascular Cognitive Impairment

Vascular cognitive impairment is distinguished by a specific cognitive profile involving preserved memory with impairments in attentional and executive functioning [180]. Vascular dementia should be broadened to recognize the important part cerebrovascular disease plays in several cognitive disorders, including the hereditary vascular dementia, multi-infarct dementia, post-stroke dementia, subcortical ischemic vascular disease and dementia, mild cognitive impairment, and degenerative dementias (including Alzheimer’s disease, frontotemporal dementia, and dementia with Lewy bodies). Alzheimer’s disease (AD) and vascular dementia (VD) share key pathologies including oxidative damage, oral supplement of phytochemical medicines, which are well-known for their antioxidant properties, can be a viable therapy for both types of dementia. In this study, the therapeutic potential of the Aster ageratoides extract (AAE), was found to be effective in experimental rat models of AD and VD. These results provided evidence that AAE supplements can exert anti-AD and -VD efficacies and suggested that AAE might be used as an edible phytotherapeutic for the two major types of dementia [94]. Another study suggested that bilobalide (BB) protected against learning and memory impairments, neuronal apoptosis, and oxidative stress in a rat model of AD induced by Aβ25–35 peptide. Additionally, the inhibition of TNF-α and Aβ1–40 expression is also involved in the action mechanism of BB in this experimental model. The results indicated that BB has protective effects on Aβ25–35-induced learning and memory impairments and neuronal oxidative damage [95]. Ginkgo biloba extract (EGb 761) is widely used to treat cerebral disorders. EGb761 increases proliferation of neural stem cells in the subventricular zone and dentate gyrus, and significantly improves learning and memory in rats with vascular dementia. Studies reported the effects on endogenous NSCs in rat models of vascular dementia and NSCs in the mouse cochlear. G. biloba seemed to slow down the cognitive deterioration in patients with VCI, but the effect was shown in only one of the four neuropsychological tests administered [96]. In a randomized, double-blind, placebo-controlled clinical trial G. biloba standardized extract showed effectiveness against vascular cognitive impairment [97]. Huperzine A is a new type of Lycopodium alkaloid monomer isolated from Huperzia serrata. It inhibits acetylcholinesterase activity and increases levels of acetylcholine in the brain thereby improving cognitive function in patients with dementia [98]. A total of 92 participants were included in the two studies, with 46 in the Hup A group and 46 in the control group. The number of patients in the individual studies ranged from 14 to 78, and the duration of trial ranged from 12 to 24 weeks [99].
Stroke is itself not a neurodegenerative disease but can lead to a secondary neurodegeneration. It has been known to be a major cause of VCI. It is the third major cause of death in brain disorders worldwide [181]. There is a sudden cessation of transient or permanent reduction of the blood flow to a particular area in the brain leading to deficiency of oxygen, glucose, and energy. The main region affected in the case of Ischemia is cornu ammonis 1 (CA1) of the hippocampus [182]. There is the death of pyramidal neurons of the CA1 region, delayed to 4 to 5 days of ischemia, hence the name is “delayed neuronal death” (DND). It is followed by a series of pathological processes working together including excitotoxicity, oxidative stress, apoptosis, inflammation, gliosis of astrocytes, and microglia in the CA1 region after ischemia/reperfusion. There is a huge burden of the depletion of the endogenous antioxidants and loss of mitochondrial function engendering neuronal death via intrinsic programmed cell death causing behavioral and histological alterations [183]. All this is followed by the breakdown of BBB causing edema. Edema activates the secretion of proinflammatory cytokines like TNF-α, IL-1β, and IL-6 by activated immune cells [184]. The oxidative stress pathway along with the inflammatory and apoptotic pathway causes further loss of neurons and brain tissues expanding the infarct area [184]. The treatment of this dreaded disease may lie in the science of phytochemicals. Many studies showing prevention against stroke have been extensively published, some of them are discussed below.
Resveratrol has been used in one of the studies where it is reported that pretreatment of resveratrol improved the neuronal functions by lowering the chances of ischemic injury in in vivo model. Additionally, it also inhibits axonal deterioration and promotes neuronal growth and synaptic formation. In in vitro study, it is found to be effective against oxygen-glucose deprivation/reoxygenation (OGD/R) injury and promoted neural stem cell (NSC) proliferation [100]. SIRT 1 is involved in the Shh signaling arbitrated the effects of resveratrol on the OGD/R injury neuronal model [100]. Its administration leads to neuronal survival, reticent the apoptosis, and synaptic formations with the upregulation of SIRT1. In rat stroke model, resveratrol slows down the ischemic injury and upgraded the neuronal functions by Sonic Hedgehog Signalling [101]. Ginkgolide B stimulated the proliferation and differentiation of neural stem cells in the in vivo models of NSCs in the rat brain following cerebral ischemia, thereby increasing the proportion of neurons to improve neurological outcomes [102]. Ginkgolide B is a potential neuro-protectant and in a study, the cellular and molecular mechanisms were dissected out by following the BDNF, EGF, and Suppressor of cytokine signaling 2 (SOCS2) [185]. Another phytochemical, curcumin, extensively studied shows beneficial effect in case of stroke too. The role of inflammation is the key factor in raising the problem of the ischemic stroke. Curcumin has shown the neuroprotective effect in the in vivo model of brain ischemia by showing anti-inflammatory properties [186]. It has also shown protection against ischemic injury in N2a cells and mouse brain with stroke [103]. Curcumin plays a significant role in neuronal survival and proliferation in the model of global brain ischemia [104]. EGCG, another phytochemical, also can remove the free radicals by increasing the antioxidative enzyme level and also downregulating the MMP-9. It inhibits the neuronal damage in transient focal cerebral ischemia instigated by middle cerebral artery occlusion (MCAO) in mice. It follows the mTOR-AMPK pathway in stressed ER and impaired the oxidative stress responses and enhance autophagy-dependent survival. It also promotes cell multiplication, differentiation, neovascularization, tube formation, and cell homing by associating the vascular endothelial growth factor (VEGF), regulation of BAX, Bcl-2, LC3B, caspase-3, mTOR and Beclin-1 [187]. It has also protected HBMVECs from Ischemia/Reperfusion injury by ameliorating apoptosis and autophagy and promoting neurovascularization [105]. EGCG has a protective effect on rat brain injury induced by MCAO, possibly by modulating the PI3K/Akt/eNOS signaling pathway [106]. Baicalein also has shown a promising effect in the subacute phase of cerebral Ischemia/reperfusion (I/R) injury in a rat model of ischemia-induced by occlusion of the middle cerebral artery (MCA). Furthermore, it is showed that baicalein inhibits the expression of NF-κB by reducing the IκBα phosphorylation and nuclear translocation of NF-κB/p65, this cascade of signaling molecules is associated with the down-regulation of the pro-inflammatory factors IL-6, IL-18, and TNF-α [107]. Additionally, baicalein inhibits the phosphorylation of ERK, p38, and JNK. These key molecules were involved with the modulation of microglia/macrophage M1/M2 polarization. Studies show the antioxidative activity by decreasing the expression of caspase-3 along with increasing the Bcl-2/Bax ratio, thus protecting the neurons. Baicalein follows the PI3K/Akt/mTOR signaling pathway via decreasing the LC3-II/ LC3-I ratio [188]. Figure 6. Ligusticum chuanxiong is used as a therapy for cerebrovascular and neuronal diseases. It works against OGD-reoxy-induced injury and can induce protective HSP-70 expression via the activation of the MAPK pathway [108]. Sinomenine, a natural phytochemical shows the neuroprotection by behaving as an anti-inflammatory and anti-apoptotic agent in a mouse model of middle cerebral artery occlusion (MCAO) and astrocytes/microglia treated with oxygen-glucose deprivation (OGD). It was observed that OGD downregulates the p-AMPK and at the same time activates the inflammasome NLRP3 in mixed glial culture (in vitro model) [109,189]. Honokiol, the main biphenyl neolignan, also shows the antioxidative effects, anti-inflammatory, anti-apoptotic activity. It helps to reduce the cerebral infarction [190]. Honokiol attenuates the inflammatory reaction during cerebral ischemia-reperfusion via inhibiting the by NF-κB activation and cytokine production of glial cells [110]. Zingerone targets the oxidative stress and provide protections in focal transient ischemic rats model [111]. Zingerone (4- (4-hydroxy-3-methoxyphenyl)-2-butanone) is an innocuous, easily available, and economically cost-effective compound with promising therapeutic effects. Having anti-inflammatory, antispasmodic, antioxidative properties, zingerone protects against oxidative stress and damage. It leads to the antioxidative pathway for protection, thus help against ischemic insult [111]. Perillyl alcohol ameliorates the condition of ischemia-reperfusion injury via inhibition of the oxidative damage by following the NF-κB, COX-2, and NOS-2 associated pathway in in vivo Rat model [112]. It is also reported that piperine vanquishes the cerebral ischemia–reperfusion-induced inflammation anti-inflammatory pathway i.e., NOS-2, COX-2, and NF-κB [113]. The list of phytochemicals and the associated mechanisms are explained in Table 1.

3.6. Prion Disease

Prion disease is also known as transmissible spongiform encephalopathies (TSE) and is the only naturally occurring infectious protein misfolding disorder. Human prion disease is caused by the unnatural conversion of normal prion protein (PrPc) into an abnormal form of a protein called prions (PrPsc), it stands for pertinacious infectious particles [191]. PrPc is a surface glycoprotein expressing in multiple tissues, however, it is highly expressed in the CNS [192]. The onset and progression of the prions disease are explained by the “protein only-hypothesis”, wherein it is understood how a protein self-replicates without a nucleic acid. It claims that PrPc itself is the infectious agent [193]. During the progression of prion disease, the infectious protein PrPsc aggregates and becomes resistant to proteinase digestion. This is followed by PrPsc-seeded conversion of PrPc into PrPsc, therefore, resulting in the high accumulation of PrPsc in an individual, which leads to neurodegeneration and the clinical manifestation of the disease [194]. This disease is classified as rare both in the progression and manifestation [195].
Several studies have been done to identify the potent agents for drug development for Prion disease. In a recent study, synthetic human prion protein (PrP 106-126) was used as an inducer to manifest prion diseases like signaling. Baicalein from Scutellaria biacalenesis attenuated ROS production and also inhibited ROS-induced mitochondrial dysfunction in PrP 106-126-induced-SH-SY5Y. JNK plays an important role in PrP 106-126-induced neuronal apoptosis in SK-N-SH cells, however, baicalein inhibited the JNK signaling in SK-N-SH cells [114]. Resveratrol, a flavonoid with several bioactivities is also known to prevent the PrP 106-126-induced neurotoxicity in SH-SY5Y and SK-N-SH cells by activating autophagy. Thus preventing the mitochondrial dysfunction, apoptosis, and neuronal cell death [115]. Autophagy is a cellular mechanism opted by several phytochemicals to protect against PrP-106-126-induced neurotoxicity, hinokitiol is a monoterpene that activates autophagy and it inhibits apoptosis through stabilization of hypoxia-inducible factor (HIF)-1α, which is important for the regulation of oxygen homeostasis [116]. Rutin has been found to prevent the PrP 106-126-induced neurotoxicity in mouse hippocampal cell line (HT22). Their results showed that rutin was able to increase the expression levels of neurotrophic factors like BDNF, glial cell-derived neurotrophic factor (GDNF), and nerve growth factor (NGF). They also observed the inhibition in apoptosis through suppression of FASL, FAS, caspase 8, and by inhibiting the activity of caspase 3 [117]. Table 1 enlist various neuroprotectant phytochemicals used in the case of Prion disease.

3.7. Frontotemporal Dementia (FTD)

FTD is a term that encompasses a broad group of heterogeneous progressive neurological syndromes. These are non-Alzheimer dementias characterized by the atrophy in the frontal and temporal lobe [196]. As frontal and temporal lobes are highly affected in this disease, the functions controlled by these areas are the outcome of FTD such as behavior and language. Some people with FTD have drastic behavioral and personality changes that make them socially inappropriate and impulsive whereas some people lose the ability to use language [197]. Some other common symptoms include lack of inhibition, apathy and judgment, repetitive compulsive behavior, decline in personal hygiene, eating the inedible objects, and loss of sensitivity towards another person’s feelings [198]. FTD can be categorized into three common variant syndromes; behavioral variant FTD (bvFTD), non-fluent/agrammatic variant primary progressive aphasia (nfvPPA), semantic variant PPA (svPPA) [199]. A subtype of FTD is Pick’s disease (PiD) which is a neurodegenerative condition that results in irreversible dementia where an excessive amount of tau inclusions known as Pick bodies is a key pathological feature [200]. Pick bodies are round argyrophilic neuronal inclusions. PiD causes cortical atrophy affecting frontal and temporal poles including limbic systems, neocortex, and dentate granular cells of the hippocampus. Such neuronal degenerations result in behavioral changes/alterations and personality disorders [201]. The early signs for PiD are apragmatism and mutism. Some studies refer to PiD as Niemann-Pick disease (NPD) based on the name of the scientist who reported the first case [200].
Not many phytochemicals have been explored for their effect on FTD. Nicotine, a plant alkaloid has been found to have a correctional effect on the behavior of Grn-/- (progranulin gene) mice. Haploinsufficiency of progranulin levels is the result of any mutation in the Grn gene, which is one of the cases of FTD [118]. Mice with insufficient amounts of progranulin secrete inflammatory cytokines at elevated levels and develop microgliosis [202,203]. This study reported that daily intervention of nicotine reversed the sociability and improved the behavioral change in the Grn-/- mice [202]. A traditional Chinese herbal medicine known as Yi-Gan San also known as Yokukansan in Japan containing a mixture of seven dried herbs Atractylodes lancea, Cnidium officinale, Uncaria rhynchophylla, Angelica acutiloba, Poria cocos, Bupleurum falcatum, and Glycyrrhiza uralensis in a ratio of 4:3:3:3:4:2:1.5 [204] was found to alleviate behavioral changes in patients with FTD [205]. Extracts of Piper nigrum (fruit) and Curcuma longa (rhizome) also inhibited the AChE significantly which is considered one of the key targets for therapeutic strategies against dementia [119]. The nanoformulation of curcumin can be a promising agent against Niemann-Pick disease type C astrocytes [120]. Curcumin has been reported to have several medicinal properties, wherein it was observed that nanoformulations of curcumin were able to increase cytosolic calcium ion levels and cell viability in astrocytes [120]. Another compound δ-tocopherol was found to decreased cholesterol levels and also lipid accumulation, it also reduced lysosomal enlargement. It was also reported to increase intracellular Ca2+ levels [121]. A study by Nekohashi et al., reported that quercetin and luteolin significantly decreased the cholesterol levels as cholesterol [122] as cholesterol is one of the key characteristics of PiD. The details of phytochemicals are given in Table 1.

3.8. Spinocerebellar Ataxia (SCA)

SCA is a term used to refer to hereditary (autosomal dominant) ataxias, these are neurodegenerative conditions affecting the cerebellum (movement controlling) part of the brain and the spinal cord [206]. It comprises a large group of neurodegeneration characterized by cerebellar ataxia with oculomotor dysfunction, dysarthria, also affecting the brain stem. The resulting ataxia is an outcome of atrophy in the cerebellum [207]. All types of SCA are associated with the loss of coordination in eye and hand movements including speech. SCA usually has an adult-onset and the progression is gradual, therefore worsening with time [208]. To date, 44 types of SCA have been identified and classified as SCA1 through SCA44. Many SCAs are an outcome of CAG nucleotide repeat expansion encoding polyglutamine, resulting in the involvement of toxic polyglutamine protein (polyQ). The most recent genetic cause of SCA44 reported in 2017 [209].
The focus on nutraceuticals against this disease is not that much, however, trehalose is a natural compound abundantly found in invertebrates, plants, and microorganisms. Studies have shown that it has a neuroprotective effect exhibited through interaction with protein folding. Analog of trehalose, melibiose has been reported the combat neurotoxicity/neurodegeneration by upregulating autophagy [123]. Melibiose significantly inhibited the aggregation of polyQ decreased ROS levels. It also displayed potential in decreasing SCA17TBP/Q79 and significant inhibition to Purkinje cell aggregation in SCA17 transgenic mice [124] Table 1.

3.9. Spinal Muscular Atrophy (SMA)

SMA is a genetic disorder characterized by the atrophy in skeletal muscle. The onset of SMA takes place due to the loss of motor neurons that regulate muscle movement. It results in weakness and makes the movement or any physical activity challenges, this condition worsens with time [210]. This disease is caused by the inadequate production of survival motor neuron (SMN) which is mainly expressed by gene SMN1. SMN2 is also responsible for the production of SMN but relatively far less. [211]. Mutations in SMN1 such as alter or loss results in observed in patients with SMA [212]. The presence of additional third or more copies of the SMN2 gene partially compensates for the damage caused due to the SMN1 gene and milder symptoms are seen in patients [213]. It mainly affects infants and restrict physical movement. SMA is categorized into five types (0-IV) based on the age of the patient and the severity of the disorder [214].
Since SMA is a genetic disorder, the plant products that can reverse or correct the damage are to be identified. Brucea javanica and its major compound as found to be bruceine D were found to have SMN2 splicing-correcting property. This study showed that they corrected the splice defect and also increased the activity of SMN2 significantly. B. javanica was able to ameliorate the muscle defects [125]. Triptolide is another phytochemical that showed a significant effect in combating SMA in SMA-like mice. The results showed that triptolide isolated from Tripterygium wilfordii elevated the production of SMN protein and transcription of SMN2 both in vivo and in vitro experimental models. Hsu et al. concluded that triptolide is a modulator of SMN expression. Triptolide enhances the expression of transcript and protein levels of survival motor neurons in human SMA fibroblasts and improves survival in SMA-like mice [126] Table 1.

4. Conclusions

This review focuses on the use of phytochemicals as a preventive approach to neurodegenerative disorders as they are cost-effective, readily applicable, and easily accessible compounds with a promise in future use. On the other hand, there have always been limitations to the clinical use of these phytochemicals. The main issue focuses on the bioavailability and digestibility of these natural compounds in the body of the patients. Blood–brain barrier is another hurdle in the use of phytochemicals for problems related to the nervous system. Moreover, it is time-consuming therapy for effectiveness against the disease.
Natural products possess variable neuroprotective properties, which can be exploited to target ND therapeutically. Phytochemicals have multi-targeted applications with numerous advantages over currently targeted drugs on the market and have limited adverse reactions. They can also help in developing food supplements and tonics from the edible plants possessing neuroprotective activity. However, given the complexities and heterogeneity of phytochemicals, a cautious approach is necessary to define herbal mixtures and compositions of different compounds for treatment purposes. Keeping this in mind, more preclinical and clinical studies are being conducted. Finally, these plants and other remaining body of studies can very well be executed in not only designing safer drugs to combat the NDs but also preventing the onset. Consequently, this review will abet the researchers for further extensive study in this field.

Author Contributions

Conceptualization: A.K. (Andleeb Khan), M.U.R.; writing—Original draft preparation: A.K. (Andleeb Khan), S.J., and Z.I.; writing—Review and editing: S.A., H.A.M., B.M.A., A.K. (Ajay Kumar), A.A. and M.U.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Wynford-Thomas, R.; Robertson, N.P. The economic burden of chronic neurological disease. J. Neurol. 2017, 264, 2345–2347. [Google Scholar] [CrossRef] [Green Version]
  2. Chen, X.; Pan, W. The Treatment Strategies for Neurodegenerative Diseases by Integrative Medicine. Integr. Med. Int. 2015, 1, 223–225. [Google Scholar] [CrossRef]
  3. Mizuno, Y. Recent Research Progress in and Future Perspective on Treatment of Parkinson’s Disease. Integr. Med. Int. 2014, 1, 67–79. [Google Scholar] [CrossRef]
  4. Shukla, A.W.; Shuster, J.J.; Chung, J.W.; Vaillancourt, D.E.; Patten, C.; Ostrem, J.; Okun, M.S. Repetitive Transcranial Magnetic Stimulation (rTMS) Therapy in Parkinson Disease: A Meta-Analysis. PM&R 2015, 8, 356–366. [Google Scholar] [CrossRef] [Green Version]
  5. Bordet, R.; Ihl, R.; Korczyn, A.D.; Lanza, G.; Jansa, J.; Hoerr, R.; Guekht, A. Towards the concept of disease-modifier in post-stroke or vascular cognitive impairment: A consensus report. BMC Med. 2017, 15, 107. [Google Scholar] [CrossRef]
  6. Fisicaro, F.; Lanza, G.; Grasso, A.A.; Pennisi, G.; Bella, R.; Paulus, W.; Pennisi, M. Repetitive transcranial magnetic stimulation in stroke rehabilitation: Review of the current evidence and pitfalls. Ther. Adv. Neurol. Disord. 2019, 12. [Google Scholar] [CrossRef] [Green Version]
  7. Liu, Y.; Yan, T.; Chu, J.M.-T.; Chen, Y.; Dunnett, S.; Ho, Y.-S.; Wong, G.T.C.; Chang, R.C.C. The beneficial effects of physical exercise in the brain and related pathophysiological mechanisms in neurodegenerative diseases. Lab. Investig. 2019, 99, 943–957. [Google Scholar] [CrossRef]
  8. Lanza, G.; Pino, M.; Fisicaro, F.; Vagli, C.; Cantone, M.; Pennisi, M.; Bella, R.; Bellomo, M. Motor activity and Becker’s muscular dystrophy: Lights and shadows. Physician Sportsmed. 2019, 48, 151–160. [Google Scholar] [CrossRef]
  9. Lanza, G.; Casabona, J.A.; Bellomo, M.; Cantone, M.; Fisicaro, F.; Bella, R.; Pennisi, G.; Bramanti, P.; Pennisi, M.; Bramanti, A. Update on intensive motor training in spinocerebellar ataxia: Time to move a step forward? J. Int. Med Res. 2019. [Google Scholar] [CrossRef] [Green Version]
  10. Lanza, G.; Centonze, S.S.; Destro, G.; Vella, V.; Bellomo, M.; Pennisi, M.; Bella, R.; Ciavardelli, D. Shiatsu as an adjuvant therapy for depression in patients with Alzheimer’s disease: A pilot study. Complement. Ther. Med. 2018, 38, 74–78. [Google Scholar] [CrossRef]
  11. Prasansuklab, A.; Brimson, J.M.; Tencomnao, T. Potential Thai medicinal plants for neurodegenerative diseases: A review focusing on the anti-glutamate toxicity effect. J. Tradit. Complement. Med. 2020. [Google Scholar] [CrossRef] [PubMed]
  12. Pennisi, M.; Lanza, G.; Cantone, M.; D’Amico, E.; Fisicaro, F.; Puglisi, V.; Vinciguerra, L.; Bella, R.; Vicari, E.; Malaguarnera, G. Acetyl-L-Carnitine in Dementia and Other Cognitive Disorders: A Critical Update. Nutrients 2020, 12, 1389. [Google Scholar] [CrossRef] [PubMed]
  13. Harvey, A.L.; Cree, I.A. High-Throughput Screening of Natural Products for Cancer Therapy. Planta Medica 2010, 76, 1080–1086. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Harvey, A.L.; Clark, R.L.; Mackay, S.P.; Johnston, B.F. Current strategies for drug discovery through natural products. Expert Opin. Drug Discov. 2010, 5, 559–568. [Google Scholar] [CrossRef]
  15. Venkatesan, R.; Ji, E.; Kim, S.Y. Phytochemicals That Regulate Neurodegenerative Disease by Targeting Neurotrophins: A Comprehensive Review. BioMed Res. Int. 2015, 2015, 1–22. [Google Scholar] [CrossRef]
  16. Moher, D.; Liberati, A.; Tetzlaff, J.; Altman, D.G.; PRISMA Group. Preferred reporting items for systematic reviews and meta-analyses: The PRISMA statement. PLoS Med. 2009, 6, e1000097. [Google Scholar] [CrossRef] [Green Version]
  17. Farkhondeh, T.; Samarghandian, S.; Pourbagher-Shahri, A.M.; Sedaghat, M. The impact of curcumin and its modified formulations on Alzheimer’s disease. J. Cell. Physiol. 2019, 234, 16953–16965. [Google Scholar] [CrossRef]
  18. Macdonald, I.R.; Rockwood, K.; Martin, E.; Darvesh, S. Cholinesterase Inhibition in Alzheimer’s Disease: Is Specificity the Answer? J. Alzheimer’s Dis. 2014, 42, 379–384. [Google Scholar] [CrossRef]
  19. Webber, K.M.; Raina, A.K.; Marlatt, M.W.; Zhu, X.; Prat, M.I.; Morelli, L.; Casadesus, G.; Perry, G.; Smith, M.A. The cell cycle in Alzheimer disease: A unique target for neuropharmacology. Mech. Ageing Dev. 2005, 126, 1019–1025. [Google Scholar] [CrossRef]
  20. Ganguli, M.; Chandra, V.; Kamboh, M.I.; Johnston, J.M.; Dodge, H.H.; Thelma, B.K.; Juyal, R.C.; Pandav, R.; Belle, S.H.; DeKosky, S.T. Apolipoprotein E Polymorphism and Alzheimer Disease. Arch. Neurol. 2000, 57, 824–830. [Google Scholar] [CrossRef] [Green Version]
  21. Kumar, A.; Singh, A. A review on Alzheimer’s disease pathophysiology and its management: An update. Pharmacol. Rep. 2015, 67, 195–203. [Google Scholar] [CrossRef] [PubMed]
  22. Oghabian, Z.; Mehrpour, O. Treatment of Aluminium Phosphide Poisoning with a Combination of Intravenous Glucagon, Digoxin and Antioxidant Agents. Sultan Qaboos Univ. Med J. 2016, 16, e352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Khani, A.H.; Sahragard, A.; Namdari, A.; Zarshenas, M.M. Botanical Sources for Alzheimer’s: A Review on Reports from Traditional Persian Medicine. Am. J. Alzheimer’s Dis. Other Dement. 2017, 32, 429–437. [Google Scholar] [CrossRef]
  24. Balasubramanian, S.; Roselin, P.; Singh, K.K.; John, Z.; Saxena, S.N. Postharvest processing and benefits of black pepper, coriander, cinnamon, fenugreek, and turmeric spices. Crit. Rev. Food Sci. Nutr. 2016, 56, 1585–1607. [Google Scholar] [CrossRef]
  25. Lin, J.-K. The Molecular Targets and Therapeutic Uses of Curcumin in Health and Disease. In Molecular Targets of Curcumin; Aggarwal, B.B., Surh, Y.J., Shishodia, S., Eds.; Springer: Boston, MA, USA, 2007; Volume 595, pp. 227–243. [Google Scholar] [CrossRef]
  26. Liu, Z.; Li, T.; Yang, D.; Smith, W.W. Curcumin protects against rotenone-induced neurotoxicity in cell and drosophila models of Parkinson’s disease. Adv. Park. Dis. 2013, 2, 18–27. [Google Scholar] [CrossRef] [Green Version]
  27. Shi, L.-Y.; Zhang, L.; Li, H.; Liu, T.-L.; Lai, J.-C.; Wu, Z.-B.; Qin, J. Protective effects of curcumin on acrolein-induced neurotoxicity in HT22 mouse hippocampal cells. Pharmacol. Rep. 2018, 70, 1040–1046. [Google Scholar] [CrossRef]
  28. Maan, G.; Sikdar, B.; Kumar, A.; Shukla, R.; Mishra, A. Role of Flavonoids in Neurodegenerative Diseases: Limitations and Future Perspectives. Curr. Top. Med. Chem. 2020, 20, 1169–1194. [Google Scholar] [CrossRef]
  29. Sundaram, J.R.; Poore, C.P.; Bin Sulaimee, N.H.; Pareek, T.; Cheong, W.F.; Wenk, M.R.; Pant, H.C.; Frautschy, S.A.; Low, C.-M.; Kesavapany, S. Curcumin Ameliorates Neuroinflammation, Neurodegeneration, and Memory Deficits in p25 Transgenic Mouse Model that Bears Hallmarks of Alzheimer’s Disease. J. Alzheimer’s Dis. 2017, 60, 1429–1442. [Google Scholar] [CrossRef]
  30. Baum, L.; Lam, C.W.K.; Cheung, S.K.-K.; Kwok, T.; Lui, V.; Tsoh, J.; Lam, L.; Leung, V.; Hui, E.; Ng, C.; et al. Six-Month Randomized, Placebo-Controlled, Double-Blind, Pilot Clinical Trial of Curcumin in Patients with Alzheimer Disease. J. Clin. Psychopharmacol. 2008, 28, 110–113. [Google Scholar] [CrossRef] [Green Version]
  31. Frost, S.; Kanagasingam, Y.; Macaulay, L.; Koronyo-Hamaoui, M.; Koronyo, Y.; Biggs, D.; Verdooner, S.; Black, K.; Taddei, K.; Shah, T.; et al. O3-13-01: Retinal amyloid fluorescence imaging predicts cerebral amyloid burden and alzheimer’s disease. Alzheimer’s Dement. 2014, 10, 234–235. [Google Scholar] [CrossRef]
  32. Zhao, L.; Wang, J.-L.; Wang, Y.-R.; Fa, X.-Z. Apigenin attenuates copper-mediated β-amyloid neurotoxicity through antioxidation, mitochondrion protection and MAPK signal inactivation in an AD cell model. Brain Res. 2013, 1492, 33–45. [Google Scholar] [CrossRef] [PubMed]
  33. Zhao, L.; Wang, J.-L.; Liu, R.; Li, X.-X.; Li, J.-F.; Zhang, L. Neuroprotective, Anti-Amyloidogenic and Neurotrophic Effects of Apigenin in an Alzheimer’s Disease Mouse Model. Molecules 2013, 18, 9949–9965. [Google Scholar] [CrossRef] [PubMed]
  34. Wiegand, H.; Wagner, A.E.; Saadatmandi, C.B.; Kruse, H.-P.; Kulling, S.; Rimbach, G. Effect of dietary genistein on Phase II and antioxidant enzymes in rat liver. Cancer Genom. Proteom. 2009, 6, 85–92. [Google Scholar]
  35. Marchenko, N.D.; Zaika, A.; Moll, U.M. Death signal-induced localization of p53 protein to mitochondria a potential role in apoptotic signaling. J. Biol. Chem. 2000, 275, 16202–16212. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Scalbert, A.; Manach, C.; Morand, C.; Rémésy, C.; Jimenez, L. Dietary Polyphenols and the Prevention of Diseases. Crit. Rev. Food Sci. Nutr. 2005, 45, 287–306. [Google Scholar] [CrossRef] [PubMed]
  37. Das, S.; Mahapatra, P.; Kumari, P.; Kushwaha, P.P.; Singh, P.; Kumar, S. Phytochemical as Hope for the Treatment of Hepatic and Neuronal Disorders. Phytochemistry 2018, 2, 289–314. [Google Scholar] [CrossRef]
  38. Farooqui, T.; Farooqui, A.A. Role of the Mediterranean Diet in the Brain and Neurodegenerative Diseases; Academic Press: Salt Lake, UT, USA, 2017; p. 484. [Google Scholar] [CrossRef]
  39. Ong, W.-Y.; Farooqui, T.; Ho, C.F.-Y.; Ng, Y.-K.; Farooqui, A.A. Use of Phytochemicals against Neuroinflammation. Neuroprotective Effects Phytochem. Neurol. Disord. 2017, 1–41. [Google Scholar] [CrossRef]
  40. Lee, T.-H.; Jung, M.; Bang, M.-H.; Chung, D.K.; Kim, J.-Y. Inhibitory effects of a spinasterol glycoside on lipopolysaccharide-induced production of nitric oxide and proinflammatory cytokines via down-regulating MAP kinase pathways and NF-κB activation in RAW264.7 macrophage cells. Int. Immunopharmacol 2012, 13, 264–270. [Google Scholar] [CrossRef]
  41. Liao, W.; Jin, G.; Zhao, M.; Yang, H. The Effect of Genistein on the Content and Activity of α- and β-Secretase and Protein Kinase C in Aβ-Injured Hippocampal Neurons. Basic Clin. Pharmacol. Toxicol. 2012, 112, 182–185. [Google Scholar] [CrossRef]
  42. Ye, S.; Wang, T.T.; Cai, B.; Wang, Y.; Li, J.; Zhan, J.X.; Shen, G.M. Genistein protects hippocampal neurons against injury by regulating calcium/calmodulin dependent protein kinase IV protein levels in Alzheimer’s disease model rats. Neural Regen. Res. 2017, 12, 1479–1484. [Google Scholar] [CrossRef]
  43. Kuo, L.-C.; Song, Y.-Q.; Yao, C.-A.; Cheng, I.H.; Chien, C.-T.; Lee, G.-C.; Yang, W.-C.; Lin, Y. Ginkgolide A. Prevents the Amyloid-β-Induced Depolarization of Cortical Neurons. J. Agric. Food Chem. 2018, 67, 81–89. [Google Scholar] [CrossRef] [PubMed]
  44. Tchantchou, F.; Xu, Y.; Wu, Y.; Christen, Y.; Luo, Y. EGb 761 enhances adult hippocampal neurogenesis and phosphorylation of CREB in transgenic mouse model of Alzheimer’s disease. FASEB J. 2007, 21, 2400–2408. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. El Tabaa, M.M.; Sokkar, S.S.; Ramadan, E.S.; El Salam, I.Z.A.; Zaid, A.M. Neuroprotective role of Ginkgo biloba against cognitive deficits associated with Bisphenol A exposure: An animal model study. Neurochem. Int. 2017, 108, 199–212. [Google Scholar] [CrossRef] [PubMed]
  46. Xu, B.; Li, X.-X.; He, G.-R.; Hu, J.-J.; Mu, X.; Tian, S.; Du, G.-H. Luteolin promotes long-term potentiation and improves cognitive functions in chronic cerebral hypoperfused rats. Eur. J. Pharmacol. 2010, 627, 99–105. [Google Scholar] [CrossRef] [PubMed]
  47. Cavallaro, V.; Baier, C.J.; Murray, M.; Estévez-Braun, A.; Murray, A.P. Neuroprotective effects of Flaveria bidentis and Lippia salsa extracts on SH-SY5Y cells. S. Afr. J. Bot. 2018, 119, 318–324. [Google Scholar] [CrossRef]
  48. Wang, H.; Wang, H.; Cheng, H.; Che, Z. Ameliorating effect of luteolin on memory impairment in an Alzheimer’s disease model. Mol. Med. Rep. 2016, 13, 4215–4220. [Google Scholar] [CrossRef] [Green Version]
  49. Pappolla, M.A.; Sos, M.; Omar, R.A.; Bick, R.J.; Hickson-Bick, D.L.M.; Reiter, R.J.; Efthimiopoulos, S.; Robakis, N.K. Melatonin Prevents Death of Neuroblastoma Cells Exposed to the Alzheimer Amyloid Peptide. J. Neurosci. 1997, 17, 1683–1690. [Google Scholar] [CrossRef]
  50. Nie, L.; Wei, G.; Peng, S.; Qu, Z.; Yang, Y.; Yang, Q.; Huang, X.; Liu, J.; Zhuang, Z.; Yang, X. Melatonin ameliorates anxiety and depression-like behaviors and modulates proteomic changes in triple transgenic mice of Alzheimer’s disease. BioFactors 2017, 43, 593–611. [Google Scholar] [CrossRef]
  51. Singer, C.; Tractenberg, R.E.; Kaye, J.A.; Schafer, K.; Gamst, A.; Grundman, M.; Thomas, R.G.; Thal, L.J. A Multicenter, Placebo-controlled Trial of Melatonin for Sleep Disturbance in Alzheimer’s Disease. Sleep 2003, 26, 893–901. [Google Scholar] [CrossRef]
  52. Zhang, N.; Hu, Z.; Zhang, Z.; Liu, G.; Wang, Y.; Ren, Y.; Wu, X.; Geng, F. Protective Role of Naringenin Against Aβ25-35-Caused Damage via ER and PI3K/Akt-Mediated Pathways. Cell. Mol. Neurobiol. 2017, 38, 549–557. [Google Scholar] [CrossRef]
  53. Haider, S.; Liaquat, L.; Ahmad, S.; Batool, Z.; Ali Siddiqui, R.; Tabassum, S.; Shahzad, S.; Rafiq, S.; Naz, N. Naringenin protects AlCl3/D-galactose induced neurotoxicity in rat model of AD via attenuation of acetylcholinesterase levels and inhibition of oxidative stress. PLoS ONE 2020, 15, e0227631. [Google Scholar] [CrossRef] [PubMed]
  54. Wang, C.; Cai, Z.; Wang, W.; Wei, M.; Kou, D.; Li, T.; Yang, Z.; Guo, H.; Le, W.; Li, S. Piperine attenuates cognitive impairment in an experimental mouse model of sporadic Alzheimer’s disease. J. Nutr. Biochem. 2019, 70, 147–155. [Google Scholar] [CrossRef] [PubMed]
  55. Li, Y.; Zhou, S.; Li, J.; Sun, Y.; Hasimu, H.; Liu, R.; Zhang, T. Quercetin protects human brain microvascular endothelial cells from fibrillar β-amyloid1–40-induced toxicity. Acta Pharm. Sin. B 2015, 5, 47–54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Sabogal-Guáqueta, A.M.; Manco, J.I.M.; Ramírez-Pineda, J.R.; Lamprea-Rodriguez, M.; Osorio, E.; Cardona-Gómez, G.P. The flavonoid quercetin ameliorates Alzheimer’s disease pathology and protects cognitive and emotional function in aged triple transgenic Alzheimer’s disease model mice. Neuropharmacology 2015, 93, 134–145. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Wang, H.; Jiang, T.; Li, W.; Gao, N.; Zhang, T. Resveratrol attenuates oxidative damage through activating mitophagy in an in vitro model of Alzheimer’s disease. Toxicol. Lett. 2018, 282, 100–108. [Google Scholar] [CrossRef]
  58. Porquet, D.; Griñán-Ferré, C.; Ferrer, I.; Camins, A.; Sanfeliu, C.; Del Valle, J.; Pallàs, M. Neuroprotective Role of Trans-Resveratrol in a Murine Model of Familial Alzheimer’s Disease. J. Alzheimer’s Dis. 2014, 42, 1209–1220. [Google Scholar] [CrossRef] [Green Version]
  59. Sawda, C.; Moussa, C.; Turner, R.S. Resveratrol for Alzheimer’s disease. Ann. NY Acad. Sci. 2017, 1403, 142–149. [Google Scholar] [CrossRef]
  60. Khan, A.; Vaibhav, K.; Javed, H.; Khan, M.M.; Tabassum, R.; Ahmed, E.; Srivastava, P.; Khuwaja, G.; Islam, F.; Siddiqui, M.S.; et al. Attenuation of Aβ-induced neurotoxicity by thymoquinone via inhibition of mitochondrial dysfunction and oxidative stress. Mol. Cell. Biochem. 2012, 369, 55–65. [Google Scholar] [CrossRef]
  61. Javed, H.; Khan, M.M.; Khan, A.; Vaibhav, K.; Ahmad, A.; Khuwaja, G.; Ahmed, E.; Raza, S.S.; Ashafaq, M.; Tabassum, R.; et al. S-allyl cysteine attenuates oxidative stress associated cognitive impairment and neurodegeneration in mouse model of streptozotocin-induced experimental dementia of Alzheimer’s type. Brain Res. 2011, 1389, 133–142. [Google Scholar] [CrossRef]
  62. Tsai, S.-J.; Chiu, C.P.; Yang, H.T.; Yin, M.C. s-Allyl cysteine, s-ethyl cysteine, and s-propyl cysteine alleviate β-amyloid, glycative, and oxidative injury in brain of mice treated by D-galactose. J. Agric. Food Chem. 2011, 59, 6319–6326. [Google Scholar] [CrossRef]
  63. Khan, A.; Vaibhav, K.; Javed, H.; Tabassum, R.; Ahmed, E.; Khan, M.M.; Khan, M.B.; Shrivastava, P.; Islam, F.; Siddiqui, M.S.; et al. 1,8-Cineole (Eucalyptol) Mitigates Inflammation in Amyloid Beta Toxicated PC12 Cells: Relevance to Alzheimer’s Disease. Neurochem. Res. 2013, 39, 344–352. [Google Scholar] [CrossRef] [PubMed]
  64. Shen, L.; Liu, L.; Li, X.-Y.; Ji, H.-F. Regulation of gut microbiota in Alzheimer’s disease mice by silibinin and silymarin and their pharmacological implications. Appl. Microbiol. Biotechnol. 2019, 103, 7141–7149. [Google Scholar] [CrossRef] [PubMed]
  65. Song, X.; Liu, B.; Cui, L.; Zhou, B.; Liu, W.; Xu, F.; Hayashi, T.; Hattori, S.; Ushiki-Kaku, Y.; Tashiro, S.-I.; et al. Silibinin ameliorates anxiety/depression-like behaviors in amyloid β-treated rats by upregulating BDNF/TrkB pathway and attenuating autophagy in hippocampus. Physiol. Behav. 2017, 179, 487–493. [Google Scholar] [CrossRef] [PubMed]
  66. Ay, M.; Luo, J.; Langley, M.; Jin, H.; Anantharam, V.; Kanthasamy, A.; Kanthasamy, A. Molecular mechanisms underlying protective effects of quercetin against mitochondrial dysfunction and progressive dopaminergic neurodegeneration in cell culture and MitoPark transgenic mouse models of Parkinson’s Disease. J. Neurochem. 2017, 141, 766–782. [Google Scholar] [CrossRef] [PubMed]
  67. Sriraksa, N.; Wattanathorn, J.; Muchimapura, S.; Tiamkao, S.; Brown, K.; Chaisiwamongkol, K. Cognitive-Enhancing Effect of Quercetin in a Rat Model of Parkinson’s Disease Induced by 6-Hydroxydopamine. Evid. Based Complement. Altern. Med. 2011, 2012. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Zhang, S.; Liu, Z.-L.; Wu, H.-C.; Hu, Q.-L.; Zhang, S.-J.; Wang, Y.-M.; Jin, Z.-K.; Lv, L.-F.; Wu, H.-L.; Cheng, O.-M. Neuroprotective effects of genistein on SH-SY5Y cells overexpressing A53T mutant α-synuclein. Neural Regen. Res. 2018, 13, 1375–1383. [Google Scholar] [CrossRef]
  69. Siddique, Y.H.; Naz, F.; Jyoti, S.; Ali, F. Rahul Effect of Genistein on the Transgenic Drosophila Model of Parkinson’s Disease. J. Diet. Suppl. 2018, 16, 550–563. [Google Scholar] [CrossRef]
  70. Jung, U.J.; Kim, S.R. Effects of naringin, a flavanone glycoside in grapefruits and citrus fruits, on the nigrostriatal dopaminergic projection in the adult brain. Neural Regen. Res. 2014, 9, 1514–1517. [Google Scholar] [CrossRef]
  71. Sang, Q.; Liu, X.; Wang, L.; Qi, L.; Sun, W.; Wang, W.; Sun, Y.; Zhang, H. Curcumin Protects an SH-SY5Y Cell Model of Parkinson’s Disease Against Toxic Injury by Regulating HSP90. Cell. Physiol. Biochem. 2018, 51, 681–691. [Google Scholar] [CrossRef]
  72. Sharma, N.; Nehru, B. Curcumin affords neuroprotection and inhibits α-synuclein aggregation in lipopolysaccharide-induced Parkinson’s disease model. Inflammopharmacology 2017, 26, 349–360. [Google Scholar] [CrossRef]
  73. Chung, W.-G.; Miranda, C.L.; Maier, C.S. Epigallocatechin gallate (EGCG) potentiates the cytotoxicity of rotenone in neuroblastoma SH-SY5Y cells. Brain Res. 2007, 1176, 133–142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Zhou, T.; Zhu, M.; Liang, Z. (-)-Epigallocatechin-3-gallate modulates peripheral immunity in the MPTP-induced mouse model of Parkinson’s disease. Mol. Med. Rep. 2018, 17, 4883–4888. [Google Scholar] [CrossRef] [Green Version]
  75. Zhang, J.; Fan, W.; Wang, H.; Bao, L.; Li, G.; Li, T.; Song, S.; Li, H.; Hao, J.; Sun, J. Resveratrol Protects PC12 Cell against 6-OHDA Damage via CXCR4 Signaling Pathway. Evid. Based Complement. Altern. Med. 2015, 2015, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Xia, D.; Sui, R.; Zhang, Z. Administration of resveratrol improved Parkinson’s disease-like phenotype by suppressing apoptosis of neurons via modulating the MALAT1/miR-129/SNCA signaling pathway. J. Cell. Biochem. 2019, 120, 4942–4951. [Google Scholar] [CrossRef] [PubMed]
  77. Magalingam, K.B.; Radhakrishnan, A.; Haleagrahara, N. Protective effects of quercetin glycosides, rutin, and isoquercetrin against 6-hydroxydopamine (6-OHDA)-induced neurotoxicity in rat pheochromocytoma (PC-12) cells. Int. J. Immunopathol. Pharmacol. 2015, 29, 30–39. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Khan, M.M.; Raza, S.S.; Javed, H.; Ahmad, A.; Khan, A.; Islam, F.; Safhi, M.M.; Islam, F. Rutin Protects Dopaminergic Neurons from Oxidative Stress in an Animal Model of Parkinson’s Disease. Neurotox. Res. 2011, 22, 1–15. [Google Scholar] [CrossRef] [PubMed]
  79. Antunes, M.S.; Goes, A.T.; Boeira, S.P.; Prigol, M.; Jesse, C.R. Protective effect of hesperidin in a model of Parkinson’s disease induced by 6-hydroxydopamine in aged mice. Nutrition 2014, 30, 1415–1422. [Google Scholar] [CrossRef] [Green Version]
  80. Shrivastava, P.; Vaibhav, K.; Tabassum, R.; Khan, A.; Ishrat, T.; Khan, M.M.; Ahmad, A.; Islam, F.; Safhi, M.M.; Islam, F. Anti-apoptotic and Anti-inflammatory effect of Piperine on 6-OHDA induced Parkinson’s Rat model. J. Nutr. Biochem. 2013, 24, 680–687. [Google Scholar] [CrossRef]
  81. Jiang, F.; DeSilva, S.; Turnbull, J. Beneficial effect of ginseng root in SOD-1 (G93A) transgenic mice. J. Neurol. Sci. 2000, 180, 52–54. [Google Scholar] [CrossRef]
  82. Zhao, Z.; Fu, J.; Li, S.; Li, Z. Neuroprotective Effects of Genistein in a SOD1-G93A Transgenic Mouse Model of Amyotrophic Lateral Sclerosis. J. Neuroimmune Pharm. 2019, 14, 688–696. [Google Scholar] [CrossRef]
  83. Chico, L.; Caldarazzo-Ienco, E.; Bisordi, C.; LoGerfo, A.; Petrozzi, L.; Petrucci, A.; Mancuso, M.; Siciliano, G. Amyotrophic Lateral Sclerosis and Oxidative Stress: A Double-Blind Therapeutic Trial After Curcumin Supplementation. CNS Neurol. Disord. Drug Targets 2018, 17, 767–779. [Google Scholar] [CrossRef] [PubMed]
  84. Dutta, K.; Patel, P.; Julien, J.-P. Protective effects of Withania somnifera extract in SOD1G93A mouse model of amyotrophic lateral sclerosis. Exp. Neurol. 2018, 309, 193–204. [Google Scholar] [CrossRef] [PubMed]
  85. Xu, Z.; Chen, S.; Li, X.; Luo, G.; Li, L.; Le, W. Neuroprotective Effects of (-)-Epigallocatechin-3-gallate in a Transgenic Mouse Model of Amyotrophic Lateral Sclerosis. Neurochem. Res. 2006, 31, 1263–1269. [Google Scholar] [CrossRef] [PubMed]
  86. Li, Y.-H.; Liu, S.-B.; Zhang, H.-Y.; Zhou, F.-H.; Liu, Y.-X.; Lu, Q.; Yang, L. Antioxidant effects of celastrol against hydrogen peroxide-induced oxidative stress in the cell model of amyotrophic lateral sclerosis. Sheng Li Xue Bao Acta Physiol. Sin. 2017, 69, 751–758. [Google Scholar]
  87. Yun, Y.C.; Jeong, S.-G.; Kim, S.H.; Cho, G.-W. Reduced sirtuin 1/adenosine monophosphate-activated protein kinase in amyotrophic lateral sclerosis patient-derived mesenchymal stem cells can be restored by resveratrol. J. Tissue Eng. Regen. Med. 2019, 13, 110–115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Gao, Y.; Chu, S.-F.; Li, J.-P.; Zhang, Z.; Yan, J.-Q.; Wen, Z.-L.; Xia, C.-Y.; Mou, Z.; Wang, Z.-Z.; He, W.-B.; et al. Protopanaxtriol protects against 3-nitropropionic acid-induced oxidative stress in a rat model of Huntington’s disease. Acta Pharmacol. Sin. 2015, 36, 311–322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Túnez, I.; Montilla, P.; Muñoz, M.D.C.; Feijóo, M.; Salcedo, M. Protective effect of melatonin on 3-nitropropionic acid-induced oxidative stress in synaptosomes in an animal model of Huntington’s disease. J. Pineal Res. 2004, 37, 252–256. [Google Scholar] [CrossRef]
  90. Vidoni, C.; Secomandi, E.; Castiglioni, A.; Melone, M.A.B.; Isidoro, C. Resveratrol protects neuronal-like cells expressing mutant Huntingtin from dopamine toxicity by rescuing ATG4-mediated autophagosome formation. Neurochem. Int. 2018, 117, 174–187. [Google Scholar] [CrossRef]
  91. Ehrnhoefer, D.E.; Duennwald, M.; Markovic, P.; Wacker, J.L.; Engemann, S.; Roark, M.; Legieiter, J.; Marsh, J.L.; Thompson, L.M.; Lindquist, S.; et al. Green tea (−)-epigallocatechin-gallate modulates early events in huntingtin misfolding and reduces toxicity in Huntington’s disease models. Hum. Mol. Genet. 2006, 15, 2743–2751. [Google Scholar] [CrossRef]
  92. Fu, J.; Jin, J.; Cichewicz, R.H.; Hageman, S.A.; Ellis, T.K.; Xiang, L.; Peng, Q.; Jiang, M.; Arbez, N.; Hotaling, K.; et al. Trans-(−)-ϵ-Viniferin increases mitochondrial sirtuin 3 (SIRT3), activates AMP-activated protein kinase (AMPK), and protects cells in models of Huntington Disease. J. Biol. Chem. 2012, 287, 24460–24472. [Google Scholar] [CrossRef] [Green Version]
  93. Kumar, P.; Kumar, A. Possible Neuroprotective Effect of Withania somnifera Root Extract Against 3-Nitropropionic Acid-Induced Behavioral, Biochemical, and Mitochondrial Dysfunction in an Animal Model of Huntington’s Disease. J. Med. Food 2009, 12, 591–600. [Google Scholar] [CrossRef] [PubMed]
  94. Jeong, J.H.; Lee, S.E.; Lee, J.H.; Kim, H.D.; Seo, K.-H.; Kim, D.H.; Han, S.Y. Aster ageratoides Turcz. extract attenuates Alzheimer’s disease-associated cognitive deficits and vascular dementia-associated neuronal death. Anat. Cell Boil. 2020, 53, 216–227. [Google Scholar] [CrossRef] [PubMed]
  95. Li, W.-Z.; Wu, W.-Y.; Huang, H.; Yin, Y.-Y.; Wu, Y.-Y. Protective effect of bilobalide on learning and memory impairment in rats with vascular dementia. Mol. Med. Rep. 2013, 8, 935–941. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Wang, J.; Chen, W.; Wang, Y. A ginkgo biloba extract promotes proliferation of endogenous neural stem cells in vascular dementia rats. Neural Regen. Res. 2013, 8, 1655–1662. [Google Scholar] [PubMed]
  97. Demarin, V.; Kes, V.B.; Trkanjec, Z.; Budišić, M.; Pasić, M.B.; Črnac, P.; Budinčević, H. Efficacy and safety of Ginkgo biloba standardized extract in the treatment of vascular cognitive impairment: A randomized, double-blind, placebo-controlled clinical trial. Neuropsychiatr. Dis. Treat. 2017, 13, 483–490. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Xu, Z.-Q.; Liang, X.-M.; Wu, J.; Zhang, Y.-F.; Zhu, C.-X.; Jiang, X.-J. Treatment with Huperzine A Improves Cognition in Vascular Dementia Patients. Cell Biophys. 2011, 62, 55–58. [Google Scholar] [CrossRef]
  99. Xing, S.-H.; Zhu, C.-X.; Zhang, R.; An, L. Huperzine A in the Treatment of Alzheimer’s Disease and Vascular Dementia: A Meta-Analysis. Evid. Based Complement. Altern. Med. 2014, 2014. [Google Scholar] [CrossRef]
  100. Tang, F.; Guo, S.; Liao, H.; Yu, P.; Wang, L.; Song, X.; Chen, J.; Yang, Q. Resveratrol Enhances Neurite Outgrowth and Synaptogenesis Via Sonic Hedgehog Signaling Following Oxygen-Glucose Deprivation/Reoxygenation Injury. Cell. Physiol. Biochem. 2017, 43, 852–869. [Google Scholar] [CrossRef] [Green Version]
  101. Shen, C.; Cheng, W.; Yu, P.; Wang, L.; Zhou, L.; Zeng, L.; Yang, Q. Resveratrol pretreatment attenuates injury and promotes proliferation of neural stem cells following oxygen-glucose deprivation/reoxygenation by upregulating the expression of Nrf2, HO-1 and NQO1 in vitro. Mol. Med. Rep. 2016, 14, 3646–3654. [Google Scholar] [CrossRef] [Green Version]
  102. Zheng, J.-S.; Zheng, P.-D.; Mungur, R.; Zhou, H.-J.; Hassan, M.; Jiang, S.-N. Ginkgolide B promotes the proliferation and differentiation of neural stem cells following cerebral ischemia/reperfusion injury, both in vivo and in vitro. Neural Regen. Res. 2018, 13, 1204–1211. [Google Scholar] [CrossRef]
  103. Xie, C.J.; Gu, A.P.; Cai, J.; Wu, Y.; Chen, R.C. Curcumin protects neural cells against ischemic injury in N2a cells and mouse brain with ischemic stroke. Brain Behav. 2018, 8, e00921. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. De Alcântara, G.F.T.; Simões-Neto, E.; Da Cruz, G.M.P.; Nobre, M.E.P.; Neves, K.R.T.; De Andrade, G.M.; Brito, G.; Viana, G.S.D.B. Curcumin reverses neurochemical, histological and immuno-histochemical alterations in the model of global brain ischemia. J. Tradit. Complement. Med. 2016, 7, 14–23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Fu, B.; Zeng, Q.; Zhang, Z.; Qian, M.; Chen, J.; Dong, W.-L.; Li, M. Epicatechin Gallate Protects HBMVECs from Ischemia/Reperfusion Injury through Ameliorating Apoptosis and Autophagy and Promoting Neovascularization. Oxidative Med. Cell. Longev. 2019, 2019, 7824684. [Google Scholar] [CrossRef]
  106. Park, D.J.; Kang, J.B.; Koh, P.O. Epigallocatechin gallate alleviates neuronal cell damage against focal cerebral ischemia in rats. J. Vet. Med. Sci. 2020, 82, 639–645. [Google Scholar] [CrossRef] [Green Version]
  107. Yang, S.; Wang, H.; Yang, Y.; Wang, R.; Wang, Y.; Wu, C.; Du, G.-H. Baicalein administered in the subacute phase ameliorates ischemia-reperfusion-induced brain injury by reducing neuroinflammation and neuronal damage. Biomed. Pharm. 2019, 117, 109102. [Google Scholar] [CrossRef] [PubMed]
  108. Yu, J.; Jiang, Z.; Ning, L.; Zhao, Z.; Yang, N.; Chen, L.; Ma, H.; Li, L.; Fu, Y.; Zhu, H.; et al. Protective HSP70 Induction by Z-Ligustilide against Oxygen–Glucose Deprivation Injury via Activation of the MAPK Pathway but Not of HSF1. Boil. Pharm. Bull. 2015, 38, 1564–1572. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Qiu, J.; Wang, M.; Zhang, J.; Cai, Q.; Lü, D.; Li, Y.; Dong, Y.; Zhao, T.; Chen, H. The neuroprotection of Sinomenine against ischemic stroke in mice by suppressing NLRP3 inflammasome via AMPK signaling. Int. Immunopharmacol. 2016, 40, 492–500. [Google Scholar] [CrossRef] [PubMed]
  110. Zhang, P.; Liu, X.; Zhu, Y.; Chen, S.; Zhou, D.; Wang, Y.-Y. Honokiol inhibits the inflammatory reaction during cerebral ischemia reperfusion by suppressing NF-κB activation and cytokine production of glial cells. Neurosci. Lett. 2013, 534, 123–127. [Google Scholar] [CrossRef]
  111. Vaibhav, K.; Shrivastava, P.; Tabassum, R.; Khan, A.; Javed, H.; Ahmed, E.; Islam, F.; Safhi, M.M.; Islam, F. Delayed administration of zingerone mitigates the behavioral and histological alteration via repression of oxidative stress and intrinsic programmed cell death in focal transient ischemic rats. Pharmacol. Biochem. Behav. 2013, 113, 53–62. [Google Scholar] [CrossRef]
  112. Tabassum, R.; Vaibhav, K.; Shrivastava, P.; Khan, A.; Ahmed, M.E.; Ashafaq, M.; Khan, M.B.; Islam, F.; Safhi, M.M.; Islam, F. Perillyl alcohol improves functional and histological outcomes against ischemia–reperfusion injury by attenuation of oxidative stress and repression of COX-2, NOS-2 and NF-κB in middle cerebral artery occlusion rats. Eur. J. Pharmacol. 2015, 747, 190–199. [Google Scholar] [CrossRef]
  113. Vaibhav, K.; Shrivastava, P.; Javed, H.; Khan, A.; Ahmed, E.; Tabassum, R.; Khan, M.M.; Khuwaja, G.; Islam, F.; Siddiqui, M.S.; et al. Piperine suppresses cerebral ischemia–reperfusion-induced inflammation through the repression of COX-2, NOS-2, and NF-κB in middle cerebral artery occlusion rat model. Mol. Cell. Biochem. 2012, 367, 73–84. [Google Scholar] [CrossRef] [PubMed]
  114. Moon, J.-H.; Park, S.-Y. Baicalein prevents human prion protein-induced neuronal cell death by regulating JNK activation. Int. J. Mol. Med. 2014, 35, 439–445. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Jeong, J.-K.; Moon, M.-H.; Bae, B.-C.; Lee, Y.-J.; Seol, J.-W.; Kang, H.-S.; Kim, J.-S.; Kang, S.-J.; Park, S.-Y. Autophagy induced by resveratrol prevents human prion protein-mediated neurotoxicity. Neurosci. Res. 2012, 73, 99–105. [Google Scholar] [CrossRef] [PubMed]
  116. Moon, J.-H.; Lee, J.-H.; Lee, Y.-J.; Park, S.-Y. Hinokitiol protects primary neuron cells against prion peptide-induced toxicity via autophagy flux regulated by hypoxia inducing factor-1. Oncotarget 2016, 7, 29944–29957. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Na, J.-Y.; Kim, S.; Song, K.; Kwon, J. Rutin Alleviates Prion Peptide-Induced Cell Death Through Inhibiting Apoptotic Pathway Activation in Dopaminergic Neuronal Cells. Cell. Mol. Neurobiol. 2014, 34, 1071–1079. [Google Scholar] [CrossRef]
  118. Minami, S.S.; Min, S.-W.; Krabbe, G.; Wang, C.; Zhou, Y.; Asgarov, R.; Li, Y.; Martens, L.H.; Elia, L.P.; Ward, M.E.; et al. Progranulin protects against amyloid β deposition and toxicity in Alzheimer’s disease mouse models. Nat. Med. 2014, 20, 1157–1164. [Google Scholar] [CrossRef] [Green Version]
  119. Murata, K.; Matsumura, S.; Yoshioka, Y.; Ueno, Y.; Matsuda, H. Screening of β-secretase and acetylcholinesterase inhibitors from plant resources. J. Nat. Med. 2014, 69, 123–129. [Google Scholar] [CrossRef]
  120. Maguire, E.; Haslett, L.J.; Welton, J.L.; Lloyd-Evans, E.; Goike, J.; Clark, E.H.; Knifton, H.R.; Shrestha, R.; Wager, K.; Webb, R.; et al. Effects of curcumin nanoformulations on cellular function in Niemann-Pick disease type C astrocytes. bioRxiv 2017, 135830. [Google Scholar] [CrossRef] [Green Version]
  121. Xu, M.; Liu, K.; Swaroop, M.; Porter, F.D.; Sidhu, R.; Finkes, S.; Ory, D.S.; Marugan, J.J.; Xiao, J.; Southall, N.T.; et al. δ-Tocopherol Reduces Lipid Accumulation in Niemann-Pick Type C1 and Wolman Cholesterol Storage Disorders. J. Boil. Chem. 2012, 287, 39349–39360. [Google Scholar] [CrossRef] [Green Version]
  122. Nekohashi, M.; Ogawa, M.; Ogihara, T.; Nakazawa, K.; Kato, H.; Misaka, T.; Abe, K.; Kobayashi, S. Luteolin and Quercetin Affect the Cholesterol Absorption Mediated by Epithelial Cholesterol Transporter Niemann–Pick C1-Like 1 in Caco-2 Cells and Rats. PLoS ONE 2014, 9, e97901. [Google Scholar] [CrossRef] [Green Version]
  123. Chen, C.-M.; Lin, C.-H.; Wu, Y.R.; Yen, C.-Y.; Huang, Y.-T.; Lin, J.-L.; Lin, C.-Y.; Chen, W.-L.; Chao, C.-Y.; Lee-Chen, G.-J.; et al. Lactulose and Melibiose Inhibit α-Synuclein Aggregation and Up-Regulate Autophagy to Reduce Neuronal Vulnerability. Cells 2020, 9, 1230. [Google Scholar] [CrossRef] [PubMed]
  124. Lee, G.-C.; Lin, C.-H.; Tao, Y.-C.; Yang, J.-M.; Hsu, K.-C.; Huang, Y.-J.; Huang, S.-H.; Kung, P.-J.; Chen, W.-L.; Wang, C.-M.; et al. The potential of lactulose and melibiose, two novel trehalase-indigestible and autophagy-inducing disaccharides, for polyQ-mediated neurodegenerative disease treatment. NeuroToxicology 2015, 48, 120–130. [Google Scholar] [CrossRef] [PubMed]
  125. Baek, J.; Jeong, H.; Ham, Y.; Jo, Y.H.; Choi, M.; Kang, M.; Son, B.; Choi, S.; Ryu, H.W.; Kim, J.; et al. Improvement of spinal muscular atrophy via correction of the SMN2 splicing defect by Brucea javanica (L.) Merr. extract and Bruceine, D. Phytomedicine 2019, 65, 153089. [Google Scholar] [CrossRef] [PubMed]
  126. Hsu, Y.-Y.; Jong, Y.-J.; Tsai, H.-H.; Tseng, Y.-T.; An, L.-M.; Lo, Y.-C. Triptolide increases transcript and protein levels of survival motor neurons in human SMA fibroblasts and improves survival in SMA-like mice. Br. J. Pharmacol. 2012, 166, 1114–1126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Leverenz, J.B.; Quinn, J.F.; Zabetian, C.; Zhang, J.; Montine, K.S.; Montine, T.J. Cognitive impairment and dementia in patients with Parkinson disease. Curr. Top. Med. Chem. 2009, 9, 903–912. [Google Scholar]
  128. Zecca, L.; Shima, T.; Stroppolo, A.; Goj, C.; Battiston, G.; Gerbasi, R.; Sarna, T.; Swartz, H. Interaction of neuromelanin and iron in substantia nigra and other areas of human brain. Neuroscience 1996, 73, 407–415. [Google Scholar] [CrossRef]
  129. Hoang, Q.Q. Pathway for Parkinson disease. Proc. Natl. Acad. Sci. USA 2014, 111, 2402–2403. [Google Scholar] [CrossRef] [Green Version]
  130. Gibb, W.R.; Lees, A.J. The relevance of the Lewy body to the pathogenesis of idiopathic Parkinson’s disease. J. Neurol. Neurosurg. Psychiatry 1988, 51, 745–752. [Google Scholar] [CrossRef] [Green Version]
  131. Albin, R.L. Parkinson’s Disease: Background, Diagnosis, and Initial Management. Clin. Geriatr. Med. 2006, 22, 735–751. [Google Scholar] [CrossRef]
  132. Dauer, W.; Przedborski, S. Parkinson’s disease: Mechanisms and models. Neuron 2003, 39, 889–909. [Google Scholar] [CrossRef] [Green Version]
  133. Yang, Y.X.; Wood, N.W.; Latchman, D.S. Molecular basis of Parkinson’s disease. NeuroReport 2009, 20, 150–156. [Google Scholar] [CrossRef] [PubMed]
  134. Reichmann, H. Modern treatment in Parkinson’s disease, a personal approach. J. Neural Transm. 2015, 123, 73–80. [Google Scholar] [CrossRef] [PubMed]
  135. Peschanski, M.; Defer, G.; Nguyen, J.P.; Ricolfi, F.; Monfort, J.C.; Hantraye, P.; Jeny, R.; Degos, J.D.; Cesaro, P.; Rémy, P.; et al. Bilateral motor improvement and alteration of L-dopa effect in two patients with Parkinson’s disease following intrastriatal transplantation of foetal ventral mesencephalon. Brain 1994, 117, 487–499. [Google Scholar] [CrossRef] [PubMed]
  136. Renaud, J.; Nabavi, S.F.; Daglia, M.; Martinoli, M.-G. Epigallocatechin-3-Gallate, a Promising Molecule for Parkinson’s Disease? Rejuvenation Res. 2015, 18, 257–269. [Google Scholar] [CrossRef]
  137. Haleagrahara, N.; Siew, C.J.; Mitra, N.K.; Kumari, M. Neuroprotective effect of bioflavonoid quercetin in 6-hydroxydopamine-induced oxidative stress biomarkers in the rat striatum. Neurosci. Lett. 2011, 500, 139–143. [Google Scholar] [CrossRef]
  138. El-Horany, H.E.; El-Latif, R.N.A.; Elbatsh, M.M.; Emam, M.N. Ameliorative Effect of Quercetin on Neurochemical and Behavioral Deficits in Rotenone Rat Model of Parkinson’s Disease: Modulating Autophagy (Quercetin on Experimental Parkinson’s Disease). J. Biochem. Mol. Toxicol. 2016, 30, 360–369. [Google Scholar] [CrossRef]
  139. Zhu, J.T.T.; Choi, R.C.Y.; Chu, G.K.Y.; Cheung, A.W.H.; Gao, Q.T.; Li, J.; Jiang, Z.Y.; Dong, T.T.X.; Tsim, K.W.K. Flavonoids Possess Neuroprotective Effects on Cultured Pheochromocytoma PC12 Cells: A Comparison of Different Flavonoids in Activating Estrogenic Effect and in Preventing β-Amyloid-Induced Cell Death. J. Agric. Food Chem. 2007, 55, 2438–2445. [Google Scholar] [CrossRef]
  140. Mirahmadi, S.-M.-S.; Shahmohammadi, A.; Rousta, A.-M.; Azadi, M.-R.; Fahanik-Babaei, J.; Baluchnejadmojarad, T.; Roghani, M. Soy isoflavone genistein attenuates lipopolysaccharide-induced cognitive impairments in the rat via exerting anti-oxidative and anti-inflammatory effects. Cytokine 2018, 104, 151–159. [Google Scholar] [CrossRef]
  141. Lou, H.; Jing, X.; Wei, X.; Shi, H.; Ren, D.; Zhang, X. Naringenin protects against 6-OHDA-induced neurotoxicity via activation of the Nrf2/ARE signaling pathway. Neuropharmacology 2014, 79, 380–388. [Google Scholar] [CrossRef]
  142. Kim, H.-J.; Song, J.Y.; Park, H.J.; Park, H.-K.; Yun, D.H.; Chung, J.-H. Naringin Protects against Rotenone-induced Apoptosis in Human Neuroblastoma SH-SY5Y Cells. Korean J. Physiol. Pharmacol. 2009, 13, 281–285. [Google Scholar] [CrossRef] [Green Version]
  143. Shen, S.-C.; Ko, C.H.; Tseng, S.-W.; Tsai, S.-H.; Chen, Y.-C. Structurally related antitumor effects of flavanones in vitro and in vivo: Involvement of caspase 3 activation, p21 gene expression, and reactive oxygen species production. Toxicol. Appl. Pharmacol. 2004, 197, 84–95. [Google Scholar] [CrossRef]
  144. Lee, W.-H.; Loo, C.-Y.; Bebawy, M.; Luk, F.; Mason, R.S.; Rohanizadeh, R. Curcumin and its Derivatives: Their Application in Neuropharmacology and Neuroscience in the 21st Century. Curr. Neuropharmacol. 2013, 11, 338–378. [Google Scholar] [CrossRef] [Green Version]
  145. Lee, H.S.; Jung, K.K.; Cho, J.Y.; Rhee, M.H.; Hong, S.; Kwon, M.; Kim, S.H.; Kang, S.Y. Neuroprotective effect of curcumin is mainly mediated by blockade of microglial cell activation. Die Pharm. 2007, 62, 937–942. [Google Scholar]
  146. Wang, M.S.; Boddapati, S.; Emadi, S.; Sierks, M.R. Curcumin reduces α-synuclein induced cytotoxicity in Parkinson’s disease cell model. BMC Neurosci. 2010, 11, 57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Dragicevic, N.; Smith, A.; Lin, X.; Yuan, F.; Copes, N.; Delic, V.; Tan, J.; Cao, C.; Shytle, R.D.; Bradshaw, P.C. Green Tea Epigallocatechin-3-Gallate (EGCG) and Other Flavonoids Reduce Alzheimer’s Amyloid-Induced Mitochondrial Dysfunction. J. Alzheimer’s Dis. 2011, 26, 507–521. [Google Scholar] [CrossRef] [PubMed]
  148. Chao, J.; Lau, W.K.-W.; Huie, M.J.; Ho, Y.-S.; Yu, M.-S.; Lai, C.S.-W.; Wang, M.; Yuen, W.-H.; Lam, W.H.; Chan, T.H.; et al. A pro-drug of the green tea polyphenol (−)-epigallocatechin-3-gallate (EGCG) prevents differentiated SH-SY5Y cells from toxicity induced by 6-hydroxydopamine. Neurosci. Lett. 2010, 469, 360–364. [Google Scholar] [CrossRef] [Green Version]
  149. Leaver, K.R.; Allbutt, H.N.; Creber, N.J.; Kassiou, M.; Henderson, J.M. Oral pre-treatment with epigallocatechin gallate in 6-OHDA lesioned rats produces subtle symptomatic relief but not neuroprotection. Brain Res. Bull. 2009, 80, 397–402. [Google Scholar] [CrossRef]
  150. Segura-Aguilar, J.; Kostrzewa, R.M. Neurotoxin Mechanisms and Processes Relevant to Parkinson’s Disease: An Update. Neurotox. Res. 2015, 27, 328–354. [Google Scholar] [CrossRef]
  151. Hedya, S.A.; Safar, M.M.; Bahgat, A.K. Cilostazol Mediated Nurr1 and Autophagy Enhancement: Neuroprotective Activity in Rat Rotenone PD Model. Mol. Neurobiol. 2018, 55, 7579–7587. [Google Scholar] [CrossRef]
  152. Ferretta, A.; Gaballo, A.; Tanzarella, P.; Piccoli, C.; Capitanio, N.; Nico, B.; Annese, T.; Di Paola, M.; Dell’Aquila, C.; De Mari, M.; et al. Effect of resveratrol on mitochondrial function: Implications in parkin-associated familiar Parkinson’s disease. Biochim. Biophys. Acta Mol. Basis Dis. 2014, 1842, 902–915. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Quik, M.; Perez, X.A.; Bordia, T. Nicotine as a potential neuroprotective agent for Parkinson’s disease. Mov. Disord. 2012, 27, 947–957. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Hendouei, F.; Moghaddam, H.S.; Mohammadi, M.R.; Taslimi, N.; Rezaei, F.; Akhondzadeh, S. Resveratrol as adjunctive therapy in treatment of irritability in children with autism: A double-blind and placebo-controlled randomized trial. J. Clin. Pharm. Ther. 2019, 45, 324–334. [Google Scholar] [CrossRef] [PubMed]
  155. Korkmaz, A.; Kolankaya, D. Inhibiting inducible nitric oxide synthase with rutin reduces renal ischemia/reperfusion injury. Can. J. Surg. 2013, 56, 6–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Roohbakhsh, A.; Parhiz, H.; Soltani, F.; Rezaee, R.; Iranshahi, M. Neuropharmacological properties and pharmacokinetics of the citrus flavonoids hesperidin and hesperetin—A mini-review. Life Sci. 2014, 113, 1–6. [Google Scholar] [CrossRef] [PubMed]
  157. Liu, J.; Chen, M.; Wang, X.; Wang, Y.; Duan, C.; Gao, G.; Lu, L.; Wu, X.; Wang, X.; Yang, H. Piperine induces autophagy by enhancing protein phosphotase 2A activity in a rotenone-induced Parkinson’s disease model. Oncotarget 2016, 7, 60823–60843. [Google Scholar] [CrossRef] [Green Version]
  158. Hardiman, O.; Al-Chalabi, A.; Chio, A.; Corr, E.M.; Logroscino, G.; Robberecht, W.; Shaw, P.J.; Simmons, Z.; van den Berg, L.H. Erratum: Amyotrophic lateral sclerosis. Nat. Rev. Dis. Primers 2017, 3, 1–19. [Google Scholar]
  159. Boillée, S.; Yamanaka, K.; Lobsiger, C.S.; Copeland, N.G.; Jenkins, N.A.; Kassiotis, G.; Kollias, G.; Cleveland, D.W. Onset and Progression in Inherited ALS Determined by Motor Neurons and Microglia. Science 2006, 312, 1389–1392. [Google Scholar] [CrossRef] [Green Version]
  160. Al-Chalabi, A.; Berg, L.H.V.D.; Veldink, J. Gene discovery in amyotrophic lateral sclerosis: Implications for clinical management. Nat. Rev. Neurol. 2016, 13, 96–104. [Google Scholar] [CrossRef] [Green Version]
  161. Simpson, E.P.; Yen, A.A.; Appel, S.H. Oxidative Stress: A common denominator in the pathogenesis of amyotrophic lateral sclerosis. Curr. Opin. Rheumatol. 2003, 15, 730–736. [Google Scholar] [CrossRef]
  162. Grosskreutz, J.; Bosch, L.V.D.; Keller, B.U. Calcium dysregulation in amyotrophic lateral sclerosis. Cell Calcium 2010, 47, 165–174. [Google Scholar] [CrossRef]
  163. McGeer, P.L.; McGeer, E.G. Inflammatory processes in amyotrophic lateral sclerosis. Muscle Nerve 2002, 26, 459–470. [Google Scholar] [CrossRef] [PubMed]
  164. Swash, M.; Schwartz, M.S. What do we really know about amyotrophic lateral sclerosis? J. Neurol. Sci. 1992, 113, 4–16. [Google Scholar] [CrossRef]
  165. Niebroj-Dobosz, I.; Janik, P. Amino acids acting as transmitters in amyotrophic lateral sclerosis (ALS). Acta Neurol. Scand. 1999, 100, 6–11. [Google Scholar] [CrossRef] [PubMed]
  166. Semmler, S.; Gagné, M.; Garg, P.; Pickles, S.R.; Baudouin, C.; Hamon-Keromen, E.; Destroismaisons, L.; Khalfallah, Y.; Chaineau, M.; Caron, E.; et al. TNF receptor-associated factor 6 interacts with ALS-linked misfolded superoxide dismutase 1 and promotes aggregation. J. Boil. Chem. 2020, 295, 3808–3825. [Google Scholar] [CrossRef] [PubMed]
  167. McGeer, E.G.; McGeer, P.L. Pharmacologic Approaches to the Treatment of Amyotrophic Lateral Sclerosis. BioDrugs 2005, 19, 31–37. [Google Scholar] [CrossRef] [PubMed]
  168. Dutta, K.; Patel, P.; Rahimian, R.; Phaneuf, D.; Julien, J.P. Withania somnifera Reverses Transactive Response DNA Binding Protein 43 Proteinopathy in a Mouse Model of Amyotrophic Lateral Sclerosis/Frontotemporal Lobar Degeneration. Neurother. J. Am. Soc. Exp. NeuroTher. 2017, 14, 447–462. [Google Scholar] [CrossRef] [Green Version]
  169. Ivanenkov, Y.A.; Balakin, K.V.; Lavrovsky, Y. Small Molecule Inhibitors of NF-B and JAK/STAT Signal Transduction Pathways as Promising Anti-Inflammatory Therapeutics. Mini Rev. Med. Chem. 2011, 11, 55–78. [Google Scholar] [CrossRef]
  170. Zhang, C.; Liang, W.; Wang, H.; Yang, Y.; Wang, T.; Wang, S.; Wang, X.; Wang, Y.; Feng, H. γ-Oryzanol mitigates oxidative stress and prevents mutant SOD1-Related neurotoxicity in Drosophila and cell models of amyotrophic lateral sclerosis. Neuropharmacology 2019, 160, 107777. [Google Scholar] [CrossRef]
  171. Szlachcic, W.J.; Switonski, P.M.; Krzyzosiak, W.J.; Figlerowicz, M.; Figiel, M. Huntington disease iPSCs show early molecular changes in intracellular signaling, the expression of oxidative stress proteins and the p53 pathway. Dis. Model. Mech. 2015, 8, 1047–1057. [Google Scholar] [CrossRef] [Green Version]
  172. Manoharan, S.; Guillemin, G.J.; Abiramasundari, R.S.; Essa, M.M.; Akbar, M. The Role of Reactive Oxygen Species in the Pathogenesis of Alzheimer’s Disease, Parkinson’s Disease, and Huntington’s Disease: A Mini Review. Oxidative Med. Cell. Longev. 2016, 2016, 1–15. [Google Scholar] [CrossRef]
  173. Friedlander, R.M. Apoptosis and Caspases in Neurodegenerative Diseases. N. Engl. J. Med. 2003, 348, 1365–1375. [Google Scholar] [CrossRef] [PubMed]
  174. Suwanjang, W.; Govitrapong, P.; Chetsawang, B.; Phansuwan-Pujito, P. The protective effect of melatonin on methamphetamine-induced calpain-dependent death pathway in human neuroblastoma SH-SY5Y cultured cells. J. Pineal Res. 2010, 48, 94–101. [Google Scholar] [CrossRef] [PubMed]
  175. Andreasen, M.; Lorenzen, N.; Otzen, D. Interactions between misfolded protein oligomers and membranes: A central topic in neurodegenerative diseases? Biochim. Biophys. Acta Biomembr. 2015, 1848, 1897–1907. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Mason, R.P.; Giorgini, F. Modeling Huntington disease in yeast: Perspectives and future directions. Prion 2011, 5, 269–276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Moumné, L.; Betuing, S.; Caboche, J. Multiple Aspects of Gene Dysregulation in Huntington’s Disease. Front. Neurol. 2013, 4, 127. [Google Scholar] [CrossRef] [Green Version]
  178. Szegő, É.M.; Outeiro, T.F.; Kazantsev, A.G. Sirtuins in Brain and Neurodegenerative Disease. Introd. Rev. Sirtuins Biol. Aging Dis. 2018, 175–195. [Google Scholar] [CrossRef]
  179. Túnez, I.; Tasset, I.; La Cruz, V.P.-D.; Santamaría, A. 3-Nitropropionic Acid as a Tool to Study the Mechanisms Involved in Huntington’s Disease: Past, Present and Future. Molecules 2010, 15, 878–916. [Google Scholar] [CrossRef] [Green Version]
  180. O’Brien, J.; Erkinjuntti, T.; Reisberg, B.; Román, G.; Sawada, T.; Pantoni, L.; Bowler, J.V.; Ballard, C.; DeCarli, C.; Gorelick, P.B.; et al. Vascular cognitive impairment. Lancet Neurol. 2003, 2, 89–98. [Google Scholar] [CrossRef]
  181. Hicks, A.; Jolkkonen, J. Challenges and possibilities of intravascular cell therapy in stroke. Acta Neurobiol. Exp. 2009, 69, 1–11. [Google Scholar]
  182. Kirino, T. Delayed neuronal death in the gerbil hippocampus following ischemia. Brain Res. 1982, 239, 57–69. [Google Scholar] [CrossRef]
  183. Abas, F.; Alkan, T.; Goren, B.; Taskapilioglu, O.; Sarandol, E.; Tolunay, S. Neuroprotective effects of postconditioning on lipid peroxidation and apoptosis after focal cerebral ischemia/reperfusion injury in rats. Turk. Neurosurg. 2010, 20, 1–8. [Google Scholar] [PubMed]
  184. Jin, W.-N.; Shi, S.X.-Y.; Li, Z.; Li, M.; Wood, K.; Gonzales, R.J.; Liu, Q. Depletion of microglia exacerbates postischemic inflammation and brain injury. Br. J. Pharmacol. 2017, 37, 2224–2236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Xiao, G.; Lyu, M.; Wang, Y.; He, S.; Liu, X.; Ni, J.; Li, L.; Fan, G.; Han, J.; Gao, X.; et al. Ginkgo Flavonol Glycosides or Ginkgolides Tend to Differentially Protect Myocardial or Cerebral Ischemia–Reperfusion Injury via Regulation of TWEAK-Fn14 Signaling in Heart and Brain. Front. Pharmacol. 2019, 10, 735. [Google Scholar] [CrossRef] [PubMed]
  186. Jiang, J.; Wang, W.; Sun, Y.J.; Hu, M.; Li, F.; Zhu, D. Neuroprotective effect of curcumin on focal cerebral ischemic rats by preventing blood–brain barrier damage. Eur. J. Pharmacol. 2007, 561, 54–62. [Google Scholar] [CrossRef] [PubMed]
  187. Nan, W.; Zhonghang, X.; Keyan, C.; Tongtong, L.; Wanshu, G.; Xu, Z.-X. Epigallocatechin-3-Gallate Reduces Neuronal Apoptosis in Rats after Middle Cerebral Artery Occlusion Injury via PI3K/AKT/eNOS Signaling Pathway. BioMed Res. Int. 2018, 2018. [Google Scholar] [CrossRef] [Green Version]
  188. Aryal, P.; Kim, K.; Park, P.-H.; Ham, S.; Cho, J.; Song, K. Baicalein induces autophagic cell death through AMPK/ULK1 activation and downregulation of mTORC1 complex components in human cancer cells. FEBS J. 2014, 281, 4644–4658. [Google Scholar] [CrossRef]
  189. Wu, P.-F.; Zhang, Z.; Wang, F.; Chen, J. Natural compounds from traditional medicinal herbs in the treatment of cerebral ischemia/reperfusion injury. Acta Pharmacol. Sin. 2010, 31, 1523–1531. [Google Scholar] [CrossRef]
  190. Xu, H.; Tang, W.; Du, G.-H.; Kokudo, N. Targeting apoptosis pathways in cancer with magnolol and honokiol, bioactive constituents of the bark of Magnolia officinalis. Drug Discov. Ther. 2011, 5, 202–210. [Google Scholar] [CrossRef] [Green Version]
  191. Barron, R. Infectious prions and proteinopathies. Prion 2017, 11, 40–47. [Google Scholar] [CrossRef]
  192. Marín-Moreno, A.; Fernández-Borges, N.; Espinosa, J.C.; Andréoletti, O.; Torres, J.M. Transmission and Replication of Prions. Prog. Mol. Biol. Transl. Sci. 2017, 150, 181–201. [Google Scholar] [CrossRef]
  193. Ma, J.; Ma, J. Prion disease and the ‘protein-only hypothesis’. Essays Biochem. 2014, 56, 181–191. [Google Scholar] [CrossRef] [PubMed]
  194. Ma, Y.; Ma, J. Immunotherapy against Prion Disease. Pathogens 2020, 9, 216. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Baiardi, S.; Rossi, M.; Capellari, S.; Parchi, P. Recent advances in the histo-molecular pathology of human prion disease. Brain Pathol. 2019, 29, 278–300. [Google Scholar] [CrossRef] [PubMed]
  196. Bott, N.; Radke, A.; Stephens, M.L.; Kramer, J.H. Frontotemporal dementia: Diagnosis, deficits and management. Neurodegener. Dis. Manag. 2014, 4, 439–454. [Google Scholar] [CrossRef] [Green Version]
  197. Kelley, R.E.; El-Khoury, R. Frontotemporal Dementia. Neurol. Clin. 2016, 34, 171–181. [Google Scholar] [CrossRef] [Green Version]
  198. Mohandas, E.; Rajmohan, V. Frontotemporal dementia: An updated overview. Indian J. Psychiatry 2009, 51, S65–S69. [Google Scholar]
  199. Miller, B.L. Frontotemporal Dementia. Front. Dement. 2014, 35, 339–374. [Google Scholar] [CrossRef] [Green Version]
  200. Kertesz, A.; Muñoz, D. Pick’s Disease, Frontotemporal Dementia, and Pick Complex. Arch. Neurol. 1998, 55, 302–304. [Google Scholar] [CrossRef] [Green Version]
  201. Frederick, J. Pick disease: A brief overview. Arch. Pathol. Lab. Med. 2006, 130, 1063–1066. [Google Scholar]
  202. Minami, S.S.; Shen, V.; Le, D.; Krabbe, G.; Asgarov, R.; Liberty, P.-C.; Lee, C.-H.; Li, J.; Donnelly-Roberts, D.; Gan, L. Reducing inflammation and rescuing FTD-related behavioral deficits in progranulin-deficient mice with α7 nicotinic acetylcholine receptor agonists. Biochem. Pharmacol. 2015, 97, 454–462. [Google Scholar] [CrossRef] [Green Version]
  203. Yin, F.; Banerjee, R.; Thomas, B.; Zhou, P.; Qian, L.; Jia, T.; Ma, X.; Ma, Y.; Iadecola, C.; Beal, M.F.; et al. Exaggerated inflammation, impaired host defense, and neuropathology in progranulin-deficient mice. J. Exp. Med. 2009, 207, 117–128. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. De Caires, S.; Steenkamp, V. Use of Yokukansan (TJ-54) in the treatment of neurological disorders: A review. Phytother. Res. 2010, 24, 1265–1270. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Kimura, T.; Hayashida, H.; Furukawa, H.; Takamatsu, J. Pilot study of pharmacological treatment for frontotemporal dementia: Effect of Yokukansan on behavioral symptoms. Psychiatry Clin. Neurosci. 2010, 64, 207–210. [Google Scholar] [CrossRef] [PubMed]
  206. Klockgether, T.; Mariotti, C.; Paulson, H.L. Spinocerebellar Ataxia. Nat. Rev. Dis. Primers 2019, 5, 24. [Google Scholar] [CrossRef] [PubMed]
  207. Schöls, L.; Bauer, P.; Schmidt, T.; Schulte, T.; Riess, O. Autosomal dominant cerebellar ataxias: Clinical features, genetics, and pathogenesis. Lancet Neurol. 2004, 3, 291–304. [Google Scholar] [CrossRef]
  208. Groth, C.L.; Berman, B.D. Spinocerebellar Ataxia 27: A Review and Characterization of an Evolving Phenotype. Tremor Other Hyperkinetic Mov. 2018, 8, 534. [Google Scholar] [CrossRef]
  209. Sullivan, R.; Yau, W.Y.; O’Connor, E.; Houlden, H. Spinocerebellar ataxia: An update. J. Neurol. 2018, 266, 533–544. [Google Scholar] [CrossRef] [Green Version]
  210. Kolb, S.J.; Kissel, J.T. Spinal Muscular Atrophy. Arch. Neurol. 2011, 68, 979. [Google Scholar] [CrossRef] [Green Version]
  211. Farrar, M.A.; Kiernan, M.C. The Genetics of Spinal Muscular Atrophy: Progress and Challenges. Neurotherapeutics 2014, 12, 290–302. [Google Scholar] [CrossRef]
  212. Rao, V.K.; Kapp, D.; Schroth, M. Gene Therapy for Spinal Muscular Atrophy: An Emerging Treatment Option for a Devastating Disease. J. Manag. Care Spéc. Pharm. 2018, 24, S3–S16. [Google Scholar] [CrossRef]
  213. Singh, R.N.; Singh, N.N. Mechanism of Splicing Regulation of Spinal Muscular Atrophy Genes. Adv. Neurobiol. 2018, 20, 31–61. [Google Scholar] [CrossRef]
  214. Fuller, H.R.; Gillingwater, T.H.; Wishart, T.M. Commonality amid diversity: Multi-study proteomic identification of conserved disease mechanisms in spinal muscular atrophy. Neuromuscul Disord. 2016, 26, 560–569. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Chemical structures of some common phytochemicals used in the prevention of neurodegenerative diseases (NDs).
Figure 1. Chemical structures of some common phytochemicals used in the prevention of neurodegenerative diseases (NDs).
Biomedicines 08 00284 g001
Figure 2. Flow diagram showing the search strategy, the number of records identified, and the included/excluded studies [16].
Figure 2. Flow diagram showing the search strategy, the number of records identified, and the included/excluded studies [16].
Biomedicines 08 00284 g002
Figure 3. Schematic representation of the several mechanisms associated with Alzheimer’s Disease (AD) and their possible prevention by some phytochemicals. The main hallmarks of AD; amyloid plague and NFT are prevented by curcumin and luteolin via GSK 3β pathway. Curcumin, genistin, and luteolin decreases the oxidative stress in the neuron. Neuroinflammation is also inhibited by genistin and apigenin. The pathway of apoptosis, activated in AD is blocked by the use of apigenin. Apigenin also activates BDNF/ERK/CREB pathway leading to neuronal survival.
Figure 3. Schematic representation of the several mechanisms associated with Alzheimer’s Disease (AD) and their possible prevention by some phytochemicals. The main hallmarks of AD; amyloid plague and NFT are prevented by curcumin and luteolin via GSK 3β pathway. Curcumin, genistin, and luteolin decreases the oxidative stress in the neuron. Neuroinflammation is also inhibited by genistin and apigenin. The pathway of apoptosis, activated in AD is blocked by the use of apigenin. Apigenin also activates BDNF/ERK/CREB pathway leading to neuronal survival.
Biomedicines 08 00284 g003
Figure 4. Diagrammatic presentation of Parkinson’s disease, cause, and role of phytochemicals to reverse the sign of PD. Mitochondrial dysfunction, oxidative stress, and uncontrolled cell death leads to the accumulation of Lewy bodies and resultant neuronal damage. Phytochemicals function with multiple targets. Curcumin targets the inducible nitric oxide synthase (iNOS) and regulates the ratio of BCL-2/BAX which leads to the restoration of mitochondrial function. Genistin increases the activity of SOD, Catalase, and GSH and protect the neurons from damage ignited by oxidative stress. Resveratrol inhibits the formation of Lewy body (aggregation of α-syn protein-a major cause of PD). Rutin protects the neuron by leading the anti-inflammatory pathway and EGCG inhibits the transcriptional factor NFκB and restores the normal function of mitochondria. EGCG also plays a role to inhibit DNA fragmentation. All over the activities of phytochemicals escape neurons from damage.
Figure 4. Diagrammatic presentation of Parkinson’s disease, cause, and role of phytochemicals to reverse the sign of PD. Mitochondrial dysfunction, oxidative stress, and uncontrolled cell death leads to the accumulation of Lewy bodies and resultant neuronal damage. Phytochemicals function with multiple targets. Curcumin targets the inducible nitric oxide synthase (iNOS) and regulates the ratio of BCL-2/BAX which leads to the restoration of mitochondrial function. Genistin increases the activity of SOD, Catalase, and GSH and protect the neurons from damage ignited by oxidative stress. Resveratrol inhibits the formation of Lewy body (aggregation of α-syn protein-a major cause of PD). Rutin protects the neuron by leading the anti-inflammatory pathway and EGCG inhibits the transcriptional factor NFκB and restores the normal function of mitochondria. EGCG also plays a role to inhibit DNA fragmentation. All over the activities of phytochemicals escape neurons from damage.
Biomedicines 08 00284 g004
Figure 5. Diagrammatic presentation of Huntington Disease progression and protection of neurons with phytochemicals: 6-35 CAG repeats lead to the formation of Huntingtin protein which consequently accumulates in the neurons and damage motor neurons. In the early stage, damaged motor neurons result in short term memory while on a later stage, this condition changes into depression, weight loss, and many more. The studied phytochemicals generally work against free radicals and restrict the accumulation of Huntingtin protein in the cells. This inhibition further leads to neuronal survival. Resveratrol works on LC3 lipidation while protopanaxtriol and viniferin leads to the antioxidative pathway and restore the function of mitochondria to protect the neurons and stop further accumulation of Huntingtin protein in neurons.
Figure 5. Diagrammatic presentation of Huntington Disease progression and protection of neurons with phytochemicals: 6-35 CAG repeats lead to the formation of Huntingtin protein which consequently accumulates in the neurons and damage motor neurons. In the early stage, damaged motor neurons result in short term memory while on a later stage, this condition changes into depression, weight loss, and many more. The studied phytochemicals generally work against free radicals and restrict the accumulation of Huntingtin protein in the cells. This inhibition further leads to neuronal survival. Resveratrol works on LC3 lipidation while protopanaxtriol and viniferin leads to the antioxidative pathway and restore the function of mitochondria to protect the neurons and stop further accumulation of Huntingtin protein in neurons.
Biomedicines 08 00284 g005
Figure 6. Overall pathogenesis of neuronal loss in ischemic stroke and their prevention by phytochemicals. Oxidative stress, found in the neurons of the ischemic brain is prevented by resveratrol and curcumin. These phytochemicals also prevent the inflammation and apoptosis induced by oxidative stress. Baicalein prevents the formation of inflammasome reducing inflammation and edema. Resveratrol activates SMO/GLI pathway leading to the development of hippocampal neuronal circuit. It also activates the PI3K/Akt/mTOR pathway leading to cell survival. Another phytochemical ginkgolide B acts on MEK/ERK/SOCS2 pathway causing inhibition of inflammation in neurons.
Figure 6. Overall pathogenesis of neuronal loss in ischemic stroke and their prevention by phytochemicals. Oxidative stress, found in the neurons of the ischemic brain is prevented by resveratrol and curcumin. These phytochemicals also prevent the inflammation and apoptosis induced by oxidative stress. Baicalein prevents the formation of inflammasome reducing inflammation and edema. Resveratrol activates SMO/GLI pathway leading to the development of hippocampal neuronal circuit. It also activates the PI3K/Akt/mTOR pathway leading to cell survival. Another phytochemical ginkgolide B acts on MEK/ERK/SOCS2 pathway causing inhibition of inflammation in neurons.
Biomedicines 08 00284 g006
Table 1. Role of various phytochemicals in different neurodegenerative diseases with possible mechanisms of action with special reference to in vitro, in vivo and clinical studies.
Table 1. Role of various phytochemicals in different neurodegenerative diseases with possible mechanisms of action with special reference to in vitro, in vivo and clinical studies.
DiseasePlantNatural CompoundIn-Vitro/In-Vivo Models/Human TrialDose & RouteMode of ActionReferences
Alzheimer’s DiseaseCurcuma longaCurcuminAcrolein exposed HT22 murine hippocampal neuronal cells5 μg/mL/30 min↓ AD-like pathologies, MDA level, ↑ levels of GSH, SOD, ↓ metalloprotease, APP, β-secretase, ↑BDNF/TrkB signaling[27]
p25 Transgenic Mouse model4 g/kg for 12 weeks; po↓ tau, amyloid accumulation, ↑ proinflammatory cytokines [29]
Pilot study with curcumin and ginkgo on 34 possible or probable AD patients1 g/day, 4 g/day
120 mg/day ginkgo leaf extract for 6 months
No significant difference in Mini-Mental State Examination scores or plasma Aβ40 levels between 0 and 6 months.
No side effect of curcumin
[30]
Longvida40 subjects: AD, MCI, healthy20 g/day for 7 daysRetinal imaging of Aβ plagues differentiated between AD and non-AD subjects[31]
Many plantsApigenin Copper induced SHSY5Y cells 10.0 mM for 24 h↑antioxidation, mitochondrion protection, ↑MAPK signaling[32]
APP/PS1 double transgenic mouse model40 mg/kg for 5 day/week ↑behavior, ↓Aβ burden, ↑ERK/CREB/BDNF pathway, ↓oxidative stress[33]
Glycine maxGenistein Rat hippocampal neuronal cells0.4 μg/mL↑ α-secretase, ↓ β-secretase, ↑PKC signaling pathway[41]
Bilateral hippocampal Aβ25–35 injected rat Model 90 mg/kg for 42 days↓ p-tau, CALM, CAMKK1, p-CAMK4 protein, escape latency[42]
Ginkgo bilobaGinkgolides A or BMouse primary cortical neurons100 or 300 µM/1.5 h↓ AMPA- and NMDA-induced depolarization, ↓JNK phosphorylation [43]
TgAPP/PS1 mice100 mg/kg for 1 month↑ cell proliferation in the hippocampus, ↓ Aβ oligomers and ↑ pCREB levels[44]
Many plantsLuteolinICV-STZ rat model10 and 20 mg/kg↑ escape latency, ↑ thickness of CA1 pyramidal layer[48]
Many plantsMelatoninmurine N2a neuroblastoma cells10 µM↑ cell viability, ↓ cell blebbing and retraction, ↓ LPO, intracellular Ca+2[49]
3xTg-AD transgenic mice10 mg/kg/day in drinking water for 1 month↑ anxiety and depression like behaviors, ↑ GST and complexin-1[50]
157 AD patients10 mg melatonin↑ nocturnal total sleep time, ↓ wake after sleep[51]
Citrus junosNaringenin25-35 treated PC-12 cells0.4 µM for 2 h↑ cell viability, ↑ ER-mediated PI3K/Akt signaling pathway, ↓ caspase-3[52]
AlCl3+D-gal rat model50 mg/kg for 2 weeks po↑ behavioral parameters, ↑ antioxidant enzymes, ↓ LPO, AChE, ↑ ACh levels, 5-HT levels and DA levels, ↓ DOPAC levels[53]
Piper nigrumPiperine ICV-STZ C57BL/6 mice10 mg/kg/22 days↑ behavioral parameters, ↑ antioxidant enzymes, ↓ LPO, ↑ ARG1, CD206, NE, DA, 5-HT and GABA levels and ↓ glutamate level in hippocampus, ↓ CD86 and iNOS, [54]
Many plantsQuercetin fAβ 1-40 insulted hBMECs0.3, 3 and 30 µmol/L↑ cell viability, ↑ γ-GT and ALP, ↓oxidative stress, ↑ barrier function[55]
triple transgenic AD (3xTg-AD) mice25 mg/kg every 2 days for 3 months↑ neuronal population, ↓ β-amyloid accumulation, tau accumulation, astrogliosis and microgliosis, ↑behavior parameters[56]
Grapes and other plantsResveratrol 1-42-treated PC12 cells3 μM↑ cell viability, ↑ mitophagy, ↓ apoptosis and ROS[57]
AβPPswe/PS1dE9mouse model16 mg/kg/day↑ behavior, ↓ plague pathology, ↑ p-AMPK/LKB pathway, ↓ SIRT-1, ↑ IL1β and TNF[58]
199 patients’ mild to moderate dementia due to AD.500 mg once daily with 500 mg increments every 13 weeks leading to 1000 mg twice daily.↑ CSF and plasma Aβ 40 in placebo, preserves brain barrier integrity[59]
Nigella sativaThymoquinone25-35-treated PC12 cells4 μM for 24 h↑ cell viability, ↓ LDH levels, ↓ TBARS, ↑ GSH, antioxidant enzymes, ↓ ROS, AChE, NO levels, ↑ MMP, ↓ iNOS expression[60]
Allium sativums-allyl cysteine (SAC)ICV-STZ mice model30 mg/kg i.p. for 15 days↑behavior, ↑ GSH, antioxidant enzymes, ↓ LPO, ↑ Bcl2 and ↓ p53 levels[61]
D-galactose (DG) mice1 g/L in drinking water for 7 weeks↓ Aβ1-40 and Aβ1-42, ↓ APP and BACE1 expression, ↑ PKC activity[62]
Many plantsEucalyptol25-35-treated PC12 cells10 μM/24 h↓LDH levels, ↓ TBARS, ↑ GSH, antioxidant enzymes, ↓ ROS, AChE, NO levels, ↓ cytokines, ↓ NF-κB, COX-2, iNOS, ↑cell viability [63]
Silybum marianumsilibinin APPswe/PS1dE9 double-transgenic mice100 mg/kg/15 days↑ behavior in MWM, ↓ amyloid plaque burden, ↓ gut bacterial [64]
Aβ1-42 model of rats25, 50 and 100 mg/kg for 10 days↓ depression, anxiety behavior, ↓ neuronal damage, ↑ BDNF and TrkB expression, ↑ autophagy[65]
Parkinson’s DiseaseMany plantsQuercetinThe mouse dopaminergic MN9D cell line10 and 30 µM quercetin for 24 h↑ PKD1 pro-survival signaling, Akt and CREB phosphorylation, BDNF expression, mitochondrial biogenesis[66]
6-OHDA model of rats100, 200, 300 mg/kg for 14 days↑ motor and non-motor deficits, antioxidant enzyme activities, ↑ neuron density in hippocampus, ↓ AChE [67]
Glycine maxGenisteinSH-SY5Y cells overexpressing A53T mutant α-synuclein 20 μM for 24 h↑ cell viability, ↓ MDA content, ↑ GSH, ATP and Na+K+ATPase levels, ↓ apoptosis[68]
Transgenic Drosophila Model expressing human α-synuclein10, 20, 30, and 40 µM for 24 days↑ life expectancy, ↓ loss of climbing ability, ↓ oxidative stress, ↑ dopamine content[69]
Grapefruit and other citrus fruitsNaringinMPP+-induced Parkinson’s disease rat model80 mg/kg for 4 days↑ GDNF expression and mTORC1, ↓ TNF-α[70]
Curcuma longaCurcuminMPP+-induced SH-SY5Y cell 40 µM for 24 h↑ cell viability, DA, Bcl-xl level, ↓ caspase 3, Bax level and HSP-90 levels[71]
Lipopolysaccharide-induced PD rat model40 mg/kg i.p. for 21 days↓ GFAP, NFκB, TNF-α, IL-1β, IL-1α, iNOS, oxidative stress, α-synuclein aggregates, apoptotic markers[72]
Green teaEpigallocatechin gallate (EGCG)Rotenone treated SH-SY5Y cells25 or 50 μM for 24 h↓ caspase-3 and apoptosis, ↑ cell viability, SOD[73]
MPTP mouse model25 and 50 mg/kg for 20 days↑ behavior, TH-positive neurons,
↓ TNF-α and IL-6
[74]
480 PD patients, randomized, double blindthree dosage groups of green tea polyphenol and one placebo control groupDelay in motor function progression, ↑ cognition, mood and quality of lifeClinicalTrials.gov identifier: NCT00461942
Grapes and other plantsResveratrol6-OHDA treated PC 12 cells50 μM for 24 h↑cell viability, MMP, ↓ apoptosis, CXCR4 protein levels[75]
MPTP mouse model50 mg/kg for 3 weeks↑ TH+ cells and miR-129, ↓ MALAT1, SNCA and apoptosis [76]
20 double blind, crossover, randomized, Placebo controlled phase 1 studyBIA 6-512 (trans-resveratrol) 25 mg dose, 50 mg dose, 100 mg doseTo study BIA 6-512 effect on levodopa pharmacokinetics when administered with levodopa/benserazideClinicalTrials.gov Identifier: NCT03091543
Many plantsRutin6-OHDA in PC-12 cells10, 50, and 100 µM for 8 h↑cell viability, catalase, SOD, GPx, GSH, ↓ MDA[77]
6-OHDA-induced PD rat model25 mg/kg for 21 days↑ behavior activities, GSH and dependent enzymes, DA and its metabolites, protein carbonyl, ↓ TBARS, H2O2, NO level, TNF-α and IL-1β[78]
Citrus treesHesperidin6-OHDA-induced PD mice model50 mg/kg for 28 days↑ behavior activities, GSH and dependent enzymes, SOD, Catalase, DA and its metabolites
↓ROS
[79]
Piper nigrumPiperine6-OHDA-induced PD mice model10 mg/kg for 15 days↑ behavior activities, GSH and dependent enzymes, SOD, Catalase, ↓ LPO, Caspase-3 and Caspase-9, TNF-α, and IL-1β[80]
Amyotrophic lateral sclerosisPanax and EleutherococcusGinseng rootSOD1-G93A Transgenic Mouse40, 80 mg/kg↑ onset to clinical signs and surrogate death[81]
GenisteinSOD1-G93A Transgenic Mouse 16 mg/kg 2 times per day↑ longevity, ↓ ALS symptoms, ↑ motor neurons, ↓ TLR2, TLR4, and NF-kB, p65 levels, IL-1β, IL-6, TNF-α levels[82]
Curcuma longaBrainoil, curcumin supplement42 randomised ALS patients 600 mg/day for 3 months↓ disease progression, AOPPs, oxidative damage, ↑ aerobic metabolism [83]
Withania somniferaWithania somnifera extractSOD1-G93A Transgenic Mouse5 mg of root powder p.o. ↑ longevity, motor performance, Hsp-70, Hsp-60 and Hsp-27, motor neurons, ↓ misfolded SOD1 protein, inflammation[84]
Green teaEGCGSOD1-G93A Transgenic Mouse10 mg/kg body ↓ disease onset, ↑ survival, motor neurons, ↓ activated microglia, NF-kB and caspase-9[85]
Tripterygium wilfordii and Celastrus regeliiCelastrolSOD1G93A transfected NSC34 cells 50 nmol/L for 4 h↑ cell viability, ↓ MDA, ↑ mRNA expressions of GCLC and GST, ERK1/2 and Akt [86]
Grapes and other plantsResveratrolBM-MSCs derived from ALS patients1 μM for 12 h↑ Neuro-progenitor markers, nestin, NF-M, Tuj-1, and Map-2, AMPK/SIRT1 signaling[87]
Huntington’s DiseasePanax ginsengProtopanaxatriol 3-NP-induced HD5, 10, and 20 mg/kg, po↑ weight, locomotor activity, antioxidant enzymes, HO-1, NQO1, ↓ ROS[88]
Many plantsmelatonin3-NP-induced HD1 mg/kg for 8 days↓ LPO, protein carbonyl, ↑ SOD and succinate dehyrogenase[89]
20 HD gene carrier subjects5 mg/day, 30 min before bedtime/monthTo improve sleep quality in HD gene carriersClinicalTrials.gov Identifier: NCT04421339
Grapes and other plantsResveratrolSH-SY5Y cells hyper-expressing the mutant polyQ Huntingtin (polyQ-Htt) protein100 µM for 24 h↑ cell viability, autophagy, autophagy degradation of mutant Huntingtin, ↓ ROS[90]
Double blind, randomized controlled 102 early affected HD patients800 mg/day for 1 yearTo evaluate resveratrol effect on caudate volume in HD patientsClinicalTrials.gov Identifier: NCT02336633
Green teaEGCGCell culture, yeast model and HD transgenic fliesDifferent doses↓ aggregation of mutant htt exon 1 protein, polyQ-mediated htt protein aggregation, cytotoxicity[91]
Vitis viniferatrans-(-)-ε-ViniferinSTHdhQ7/Q7, STHdhQ111/Q111, Tet-Off PC12 cells, Neuroblastoma N2a cells and Primary cortical neurons1 nM, 10nM, 100nM, 1 µM↓ cell toxicity, oxidative stress, mitochondrial dysfunction, ↑ mitochondrial genesis, SIRT3-dependent AMPK Activation[92]
Withania SomniferaW. somnifera root extract3-NP-induced HD100 and 200 mg/kg for 14 days↑ body weight, behavioral activities, SOD, catalase, mitochondrial complex (I, II, III) levels, ↓ LPO, nitrite, LDH[93]
Vascular Cognitive impairmentAster ageratoidesAster ageratoides extract (AAE)PC 12 cells treated with 50 µM glutamate or 2VO/H surgery in Sprague Dawley rats0, 10, 25, and 50 µg/mL or 50 mg/kg b.w↓ Memory impairment in vivo, ↓ hippocampal structures, neuronal excitotoxicity in vitro[94]
Ginkgo biloba Lbilobalidemale Sprague Dawley rats (2-vessel occlusion, 2-VO)2, 4 and 8 mg/kg↓ Memory impairment, ↓ nuclear condensation, ↑SOD and GSH, ↓ NOS and MDA, TNF-α[95]
EGb761bilateral common carotid arteries repeated occlusion in rats and ip injection of sodium nitroprusside50 mg/kg↓ Memory impairment, proliferation of neural stem cells in dentate gyrus and subventricular zone[96]
Ginkgo biloba extract90 patients, randomized, double-blind, placebo-controlled trial60 and 120 mg↑ CGI scores, no effect on trans cranial Doppler ultrasound, more adverse reactions in Placebo group. [97]
Huperzia serrata Huperzine randomized, double-blinded, placebo-controlled study with 78 patients with mild to moderate VaD0.1-mg bid↑ MMSE, CDR, and ADL scores [98,99]
Stroke (causes secondary neurodegeneration)Grapes and other plantsResveratrolPrimary cortical neuron cultures with OGD/R1, 5, and 20 µmol/L for 24 h↑ cell viability, neurite outgrowth and synaptogenesis via Shh signaling pathway, Sirt1 activation[100]
(MCAO/R) Model30 mg/kg for 7 days↑ behavior deficits, ↓ infarction volume, ↑ NeuN+ cells, ↓ TUNEL+ cells [101]
Ginkgo bilobaGinkgolides NSC line20, 40 or 60 mg/L↑ cell viability, process length and cell body area, sizes of NSE, GFAP and SOCS2-positive cells [102]
mouse model of myocardial I/R injury2.5 mL/kg ↑ cardiac function, ↓ LDH and AST, TWEAK expression, infarction volume[102]
Curcuma longaCurcuminMouse N2a cells hypoxia model5, 15, 25, and 35 μmol/L for 24 h↑ cell viability, mitochondrial disruption, Bcl2 expression, ↓ Tunnel positive cells, Bax and Caspase-3 expression[103]
Rat model of global brain ischemia25 and 50 mg/kg↑ DA and its metabolites DOPAC and HVANE and 5-HT, cell viability, ↓ COX-2, TNF-α expression[104]
Green teaEGCGHBMVECs OGD/R model2 μM for 24 h↑ cell viability, SOD, migration and tube formation, mRNA expression of VEGF, Bcl2, ↓ apoptosis and autophagy, ROS, LDH, MDA, mRNA expression of Bax and Caspase-3[105]
Rat MCAO model20 mg/kg↓ infarct volume, TUNEL+ cells, NO, MDA, ↑ Behavioral parameters, SOD and GPx[106]
Scutellaria baicalensisBaicalin/baicaleinRat MCAO model200 mg/kg, 24 h after reperfusion till 7 days↑ behavior deficits, microglia/macrophage M2 markers CD206 and CD 163, ↓infarct volume, microglia/ macrophage M1 markers CD86 and CD 16, MAPK and NF-κB[107]
Angelica sinensisLigustilideRat MCAO model7.5, 15 or 30 mg/kg for 3 days↑ behavior deficits, ↓ infarct volume, ↑ HSP-70 and MAPK activation[108]
Sinomenium acutumSinomenineMice MCAO model10 or 20 mg/kg daily for 3 days ↑ behavior deficits, ↓ infarct volume, apoptosis, astrocyte and microglial activation, NLRP3 inflammasome, IL-1β, IL-6, IL-18 and TNF-α generation[109]
Magnolia officinalisHonokiolRat cerebral ischemia reperfusion model0.7–70 µg/kg, 15 min after ischemia↓ Cerebral edema, p65 level, NO, TNF- α, RANTES/CCL5 levels[110]
Zingiber officinaleZingeroneRat MCAO model50, 100 mg/kg at 5 h and 12 h after initiation of surgery↑ behavior deficits, ↓ infarct volume, LPO, Caspase-3 and -9 Apaf-1, Bax, ↑ GSH and dependent enzymes, Catalase, Bcl-2[111]
Many plantsPerillyl alcoholRat MCAO model25, 50, 100 mg/kg for 7 days↑ behavior deficits, ↓ infarct volume, LPO, TNF- α, IL-1β, IL-6, COX-2, iNOS, NF-κB, ↑ GSH and dependent enzymes, Catalase[112]
Piper nigrumPiperineRat MCAO model10 mg/kg for 15 days↓ infarct volume, ↑ behavior deficits and histopathological findings, ↓ TNF- α, IL-1β, IL-6, COX-2, iNOS, NF-κB[113]
Prion DiseaseScutellaria biacalenesisBaicaleinPrP 106-126-induced-(SH-SY5Y and PrP 106-126 and SK-N-SH) cells80 µM↓ROS production, ↓ mitochondrial dysfunction, ↓ apoptosis, ↓ JNK signalling[114]
Grapes and other plantsResveratrol PrP 106-126-induced-(SH-SY5Y and PrP 106-126 and SK-N-SH) cells 2–4 µM↑ autophagy, ↓ mitochondrial dysfunction, ↓ apoptosis[115]
Cupressaceous plantsHinokitiolPrP 106-126-induced SK-N-SH8 µM↑ HIF-1α, ↑ autophagy, ↑ p62/SQSTM1, ↓ apoptosis[116]
Many plantsRutin PrP induced-HT22 cells 10 µg/mL↓ ROS and NO, ↓ caspase 3 activity, ↓ caspase 8, FAS, FASL, ↑ BDNF [117]
Frontotemporal dementiaMany plantsNicotineGrn -/-mice0.6 mg/kg P.I daily for 14 days↓ CD68, IL-1β, CD11b, ↑ sociability[118]
Curcuma longa 250–500 µg/mL↓ AChE[119]
Piper nigrum 250–500 µg/mL↓ AChE[119]
Pick’s diseaseCurcuma longaCurcuminPrimary astrocytes from NPC +/+ and NPC -/- mice30 µM↑ cytosolic Ca2+, ↑ viability[120]
Many plantsδ-tocopherolHuman fibroblast and baby hamster kidney cells40 µM↓ cholesterol accumulation, ↓ lysosomal size, ↑ intracellular Ca2+, ↑ Ca2+ deficiency [121]
Many plantsQuercetinCoca-2 cells and male Wistar rats 100 µM and 5 mg/kg (rats)↓Cholesterol uptake[122]
Many plantsLuteolinCoca-2 cells and male Wistar rats 100 µM and 5 mg/kg (rats)↓Cholesterol uptake[122]
Spinocerebellar ataxia Melibiose293 cells and SCA17 transgenic mice 100 nm - 100 µM for 6 days and daily I.P for days (24 mg/kg) ↓ polyQ aggregration. ↓ ROS, ↑ autophagy, ↓ caspase 3[123,124]
Spinal muscular atrophyBrucea javanicaBruceine DSMA mice and Δ7 mice10 to 30 mg/kg i.p once a day for 7 daysCorrecting the splice defect in SMN2, ameliorating SMN phenotype defects[125]
Tripterygium wilfordiiTriptolideSMA mice, NSC34 and N18TG2 cells0.1 mg/kg I.P daily↑ SMN, Gemin2 and Gemin3 expression levels, ↑ transcription of SMN, ↑ survival rate and ↓ SMA related defects[126]
↑prevented/loss of/induced/enhanced/improved/increased/upregulated/elicited/promoted/restored/activated. ↓down-regulated/attenuated/decreased/declined/terminated/ blocked/ prevented/ inhibited. 3xTg-AD: Enhanced susceptibility of triple transgenic Alzheimer’s disease, 5-HT: 5-hydroxytryptamine, AlCl3+D-gal: Aluminium chloride+d-galactose, ALP:Alkaline phosphatase, AMPA: α-amino-3-hydroxy-5-methylisoxazole-4-propionic acid, APP/PS1:amyloid precursor protein/presenilin 1, APP:Amyloid precursor protein, ARG1:Arginase, Bcl-xl: B-cell lymphoma-extra-large, CALM: Clathrin-assembly lymphoid myeloid leukaemia protein, CAMK4: Calcium/calmodulin-dependent protein kinase type IV, CAMKK1: Calcium/Calmodulin Dependent Protein Kinase Kinase 1, CREB: cAMP response element-binding protein, CXCR4: C-X-C chemokine receptor type 4, DOPAC: 3,4-Dihydroxyphenylacetic acid, GCLC: Glutamate-Cysteine Ligase Catalytic Subunit, GSH: Glutathione, GST: Glutathione S-transferase, hBMECs: Human brain microvascular endothelial cells, Hsp-27: Heat shock protein-27, Hsp-60: Heat shock protein-60, HSP-90: Heat shock protein-90, ICV-STZ: Intracerebroventricular streptozotocin, MALAT1: Metastasis Associated Lung Adenocarcinoma Transcript 1, Map-2: Microtubule associated protein 2, MCAO/R: middle cerebral artery occlusion- reperfusion, MDA: Malondialdehyde, MN9D: murine mesencephalon-derived dopaminergic neuronal cell line, mTORC1: mammalian target of rapamycin complex 1, MWM: Morris water maze, NE: Norepinephrine, NF-M: Neurofilament medium polypeptide, NGF: Nerve growth factor, NMDA: N-methyl-D-aspartate, NO: Nitric oxide, p-AMPK/LKB: AMP-activated protein kinase/liver kinase B, PI3K: phosphoinositide 3-kinase, PKC: Protein kinase C, RANTES/CCL5: regulated upon activation, normal T cell expressed and secreted/ Chemokine (C-C motif) ligand 5, SIRT-1: silent mating type information regulation 2 homolog 1, SNCA: Synuclein Alpha, TBARS: Thiobarbituric acid reactive substances, TH+: Tyrosine hydroxylase, TLR2: Toll-like receptor 2, TLR4: Toll-like receptor 4, TrkB: Tropomyosin receptor kinase B, Tuj-1: Neuron-specific Class III β-tubulin, TUNEL+: terminal deoxynucleotidyl transferase-mediated dUTP nick-end labelling+, TWEAK: TNF-related weak inducer of apoptosis, γ-GT: Gamma-glutamyl transferase.

Share and Cite

MDPI and ACS Style

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. https://doi.org/10.3390/biomedicines8080284

AMA Style

Khan A, Jahan S, Imtiyaz Z, Alshahrani S, Antar Makeen H, Mohammed Alshehri B, Kumar A, Arafah A, Rehman MU. Neuroprotection: Targeting Multiple Pathways by Naturally Occurring Phytochemicals. Biomedicines. 2020; 8(8):284. https://doi.org/10.3390/biomedicines8080284

Chicago/Turabian Style

Khan, Andleeb, Sadaf Jahan, Zuha Imtiyaz, Saeed Alshahrani, Hafiz Antar Makeen, Bader Mohammed Alshehri, Ajay Kumar, Azher Arafah, and Muneeb U. Rehman. 2020. "Neuroprotection: Targeting Multiple Pathways by Naturally Occurring Phytochemicals" Biomedicines 8, no. 8: 284. https://doi.org/10.3390/biomedicines8080284

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop