Next Article in Journal
Effects of Green Kiwifruit Peel Extract on Sleep-Wake Profiles in Mice: A Polysomnographic Study Based on Electroencephalogram and Electromyogram Recordings
Next Article in Special Issue
The Preventive Effect of Exercise and Oral Branched-Chain Amino Acid Supplementation on Obesity-Induced Brain Changes in Ldlr−/−.Leiden Mice
Previous Article in Journal
What Do People Know about Food, Nutrition and Health?—General Nutrition Knowledge in the Austrian Population
Previous Article in Special Issue
A Prospective Study on the Relationship between Iron Supplement Intake, Hemoglobin Concentration, and Risk of Parkinsonism
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mechanistic Insights into the Neuroprotective Potential of Sacred Ficus Trees

Department of Bionano Technology, Gachon University, 1342 Seongnam-daero, Sujeong-Gu, Seongnam 461-701, Korea
*
Authors to whom correspondence should be addressed.
Nutrients 2022, 14(22), 4731; https://doi.org/10.3390/nu14224731
Submission received: 6 September 2022 / Revised: 16 October 2022 / Accepted: 5 November 2022 / Published: 9 November 2022

Abstract

:
Ficus religiosa (Bo tree or sacred fig) and Ficus benghalensis (Indian banyan) are of immense spiritual and therapeutic importance. Various parts of these trees have been investigated for their antioxidant, antimicrobial, anticonvulsant, antidiabetic, anti-inflammatory, analgesic, hepatoprotective, dermoprotective, and nephroprotective properties. Previous reviews of Ficus mostly discussed traditional usages, photochemistry, and pharmacological activities, though comprehensive reviews of the neuroprotective potential of these Ficus species extracts and/or their important phytocompounds are lacking. The interesting phytocompounds from these trees include many bengalenosides, carotenoids, flavonoids (leucopelargonidin-3-O-β-d-glucopyranoside, leucopelargonidin-3-O-α-l-rhamnopyranoside, lupeol, cetyl behenate, and α-amyrin acetate), flavonols (kaempferol, quercetin, myricetin), leucocyanidin, phytosterols (bergapten, bergaptol, lanosterol, β-sitosterol, stigmasterol), terpenes (α-thujene, α-pinene, β-pinene, α-terpinene, limonene, β-ocimene, β-bourbonene, β-caryophyllene, α-trans-bergamotene, α-copaene, aromadendrene, α-humulene, alloaromadendrene, germacrene, γ-cadinene, and δ-cadinene), and diverse polyphenols (tannin, wax, saponin, leucoanthocyanin), contributing significantly to their pharmacological effects, ranging from antimicrobial action to neuroprotection. This review presents extensive mechanistic insights into the neuroprotective potential, especially important phytochemicals from F. religiosa and F. benghalensis. Owing to the complex pathophysiology of neurodegenerative disorders (NDDs), the currently existing drugs merely alleviate the symptoms. Hence, bioactive compounds with potent neuroprotective effects through a multitarget approach would be of great interest in developing pharmacophores for the treatment of NDDs.

1. Introduction

Neurodegeneration or neuronal atrophy results from progressive degeneration of neuronal structures and/or functions. It predominantly affects the elderly population as a spectrum of neurodegenerative disorders (NDDs), including Alzheimer’s disease (AD), Parkinson’s disease (PD), Huntington’s disease (HD), and amyotrophic lateral sclerosis (ALS). NDDs lead to a progressive decline in mental (cognition, memory, orientation, attention) and motor movements (gait, orientation, balance). Neurodegeneration affects diverse neuronal circuitries at different locations in the brain. These alterations lead to the loss of dopaminergic neurons (PD), dorsal striatum neurons (HD), both upper and lower motor neurons (ALS), and axonal neuron degeneration (temporal and hippocampal neurons in AD) [1,2,3,4]. Aggregation of proteins and peptides into their pathological forms are the key events in NDDs [5], for example, amyloid beta protein (Aβ) and Tau in AD [6]; mutant huntingtin (mHTT) protein in HD [7]; α-synuclein (αSN) in PD and TAR DNA-binding protein 43 (TDP-43) in ALS [8]. Other factors, such as neuroinflammation, oxidative stress, age, genetics, and environmental factors, also contribute to the etiopathology of NDDs [9,10]. No doubt, tremendous efforts are poured in to find the ways for early diagnoses and cures for NDDs, but only little advances have been achieved in both avenues. Recently, groundbreaking advances have been made regarding early blood diagnoses of several NDDs by targeting genetic mutations, Aβ oligomers, and Aβ42 [11,12,13,14,15]. The existing therapies target only a specific pathway/enzyme and are suitable to alleviate only the symptoms, not stop the disease progression or disease modifications. Since many medicinal plants have been documented in various traditional texts for their therapeutic effects, several compounds for treating NDDs are of plant origin, such as levodopa (in PD treatment) and galantamine (in AD treatment) obtained from Vicia faba [16] and Galanthus [17], respectively. Hence, a search for bioactive compounds that can exert a neuroprotective effect through a multitarget approach is desirable. In addition, with the possibility of detecting various oligomers of NDDs in blood by a multimer detection system (MDS), screening bioactive compounds from plant extracts would be feasible. Currently, we are investigating neuroprotective mechanism of selected extracts in our laboratory using a combination of chemical, biochemical, molecular, and cell-based assays. The data obtained are quite promising and we will be publishing the results soon.
The genus Ficus (family Moraceae) consists of over 850 species of vines, shrubs, and trees, occupying a larger part of the tropical and subtropical forest ecosystem [18,19]. Ficus trees are among the highest oxygen producers in nature with a prime photosynthesis rate as well as a rich source of mineral deposits in the leaves [20]. Among the Ficus species, Ficus religiosa (Bo tree or sacred fig) and Ficus benghalensis (Indian banyan) have a long history of spiritual significance in the Indian subcontinent. As the name suggests, F. religiosa is considered a spiritual tree in Buddhism and Hinduism, as Gautama Buddha attained enlightenment under this tree in India, and Buddhism originated. In 288 BCE, a cutting from this tree was planted in Sri Lanka and the tree is still alive, which gives an astonishing account of the age of the tree. F. benghalensis is the national tree of India, symbolizing “wisdom and eternity” [21]. Several references to Ficus are also mentioned in the Bible where Adam and Eve used Ficus leaves to cover their bodies in heaven (Genesis 3:7) [22]. All these references signify the importance and eternity of the Ficus tree.
F. religiosa is a deciduous tree with heart-shaped leaves, while F. benghalensis is evergreen, with leathery and ovate leaves. In the Indian traditional system of medicine (Ayurveda), different parts of these trees are used to treat cough, asthma, heart diseases, nose bleeding, diabetes, toothaches, constipation, fever, jaundice, wounds, gonorrhea, and skin infections [23,24,25]. These trees present a wide range of promising bioactive compounds, such as phenols, flavonoids, carotenoids, sterols, anthocyanins, alkaloids, tannins, saponins, terpenoids, and vitamins, possessing a wide range of biological properties. [18]. F. benghalensis and F. religiosa share several common therapeutic activities, such as antiulcerogenic [26,27], anticancer [28,29], antidiabetic [30], antipyretic [31,32], hypolipidemic [33,34], anthelminthic [35], anti-inflammatory [32,36], and immunomodulatory properties [37,38]. Apart from this, F. benghalensis also displays antidiarrheal [39] and antiallergic activities [40], while F. religiosa boasts bronchodilatory, anti-asthmatic [41], and anticonvulsant properties [42].
Previous reviews of these species mostly discussed the traditional usages, photochemistry, and pharmacological activities; however, comprehensive reviews of the neuroprotective potential of these Ficus species extracts and/or their important phytocompounds are lacking. This report presents an inclusive review of the existing scientific works from various databases (PubMed, Google Scholar, and Science Direct) published before July 2022 on the neuroprotective mechanisms of F. religiosa and F. benghalensis extracts, as well as the major bioactive compounds present therein.
The keywords in English included “neuroprotection”, “Ficus religiosa”, “Ficus benghalensis”, “Alzheimer’s disease”, “neurodegenerative diseases”, “extracts”, and “bioactive metabolites”. All the literature about in vitro, in vivo, and in silico studies related to protein aggregation mechanism, oxidative stress, antioxidant parameters, proinflammatory cytokines, enzyme inhibition, metabolic pathways, gene expression, neurotransmitter levels, memory, and cognition were included. Literature in a language other than English and unpublished works were excluded.

2. Phytochemicals and Therapeutic Properties

2.1. F. religiosa

All parts of the plant (roots, bark, fruits, leaves, latex, and seeds) are of therapeutic importance [43] due to the presence of several important phytochemicals. The bark of F. religiosa contains high concentrations of polyphenols (tannin, wax, saponin, leucoanthocyanin), flavonoids (leucopelargonidin-3-O-β-d-glucopyranoside, leucopelargonidin-3-O-α-l-rhamnopyranoside, lupeol, cetyl behenate, and α-amyrin acetate), flavonols (kaempferol, quercetin, myricetin), and polysterols (bergapten, bergaptol, lanosterol, β-sitosterol, stigmasterol) [44,45]. The major compounds (lupenol, γ-sitosterol and 1,2-benzenediol) identified in the stem bark extract (petroleum ether, chloroform, and methanol) by GC–MS [46] display anti-inflammatory, antidiarrheal, hypoglycemic, and antibacterial properties [30,47].
The leaves are a rich source of polyphenols (eugenol, hexanol, phytol), sesquiterpene (eudesmol), and several other compounds, such as α-copaene, linalool, salicylaldehyde, phenylacetaldehyde, allyl caproate, n-nonanal, adipoin, methylcyclopentane, 2-dione, itaconic anhydride, benzeneacetonitrile, nonadienal, nonadienol, catechol, coumarin, cinnamyl alcohol, vinyl guaiacol, α-cadinol, pentadecanal, and palmitic acid [48]. The leaves are used to relieve nose bleeding and blood in urine and stool due to coagulative/anti-fibrinolytic properties. Additionally, the leaf juice is used for treating asthma, migraine, toothache, diarrhea, wounds, and gastric disorders [49].
The fruits have laxative properties [25] and contain abundant flavonols (quercetin, kaempferol, myricetin), terpenes/terpenoids (α-thujene, α-pinene, β-pinene, α-terpinene, limonene, β-ocimene, β-bourbonene, β-caryophyllene, α-trans-bergamotene, α-copaene, aromadendrene, α-humulene, alloaromadendrene, germacrene, γ-cadinene, and δ-cadinene), and polyphenols (stigmasterol, lupeol) [50,51,52,53]. The F. religiosa latex is used to treat neuralgia and contains serine proteases (religiosin B and C) [54]. The seeds contain high concentrations of alanine, tyrosine, and threonine amino acids [55].

2.2. F. benghalensis

In traditional medicine, the F. benghalensis root and bark are used as antidiabetic, anti-inflammatory, antidiarrheal [56], cholesterol-lowering [57], anthelmintic [35], and anti-asthmatic medications [40]. The leaves seem to boost the immune system and are used to treat leucorrhoea [58]. The seeds and latex are used in treating peptic ulcers [59] and urinary disorders [60], respectively.
The important phytocomponents of the F. benghalensis leaves are rutin, β-amyrin, leucopelargonin, bengalenoside, psoralen, β-sitosterol, and bergapten [61], while β-sitosterol-α-d-glucose and myo-inositol are reported in the aerial roots [18]. Synephrine, cyanuric acid, adonitol, azelaic acid, butedioic acid, heneicosanyl oleate, α-amyrin acetate, lupeol, lanostadienyl glucosyl cetoleate, and bengalensisteroic acid ester have been identified in the methanolic bark extract [62,63]. The leaves contain furanocoumarin (psoralen, rhein, and bergapten) and quinone [61].

3. Neuroprotective Effect of Ficus religiosa

The neuroprotective effect of F. religiosa is summarized in Table 1.

3.1. Leaves

The methanolic leaf extract of F. religiosa (5–200 µg/mL) exhibits an anti-inflammatory response in LPS-stimulated microglia (BV-2 cells, a mouse microglia cell line) by inhibiting the production of proinflammatory cytokines, such as tumor necrosis factor alpha (TNF-α), interleukin beta (IL-1β) and IL-6, inflammatory mediators and nitric oxide (NO), by downregulating several signaling pathways, such as the p38 mitogen-activated protein kinase (MAPK), extracellular signal-regulated kinase (ERK), and c-Jun N-terminal kinase (JNK) ones. Additionally, the extract also suppresses the activation of nuclear factor kappa B (NF-κB), which strongly supports the neuroprotective role of F. religiosa in several NDDs [36].
The neuroprotective effect of the methanolic leaf extract of F. religiosa (MEFR) was studied on aluminum chloride (AlCl3)-induced neurotoxicity in rats. Aluminum can cross the blood–brain barrier (BBB) through receptor-mediated transfer where it disturbs the oxidative state of the brain and causes neuronal cell death [64,65]. Histological studies have revealed maximum neurodegeneration in the hippocampal CA3 region of the AlCl3-treated group. This region has a key role in memory, susceptibility to seizures, and neurodegeneration [66]. However, the group treated with MEFR (200 and 300 mg/kg body weight) presented a significant improvement in the number and quality of neurons [67].
The neuroprotective effect of the petroleum ether (PE) extract of F. religiosa leaves (PEFR) was studied on a 3-nitropropionic acid (3-NP)-induced HD mouse model [68]. Systemic intraperitoneal (i.p.) administration of 3-NP caused striatum neuronal degeneration as seen in HD [69]. Daily oral administration of the PEFR (400 mg/kg body weight) significantly enhanced cognitive and motor activities compared to the untreated group. Biochemical investigations showed that the extract reduced the levels of oxidative stress and inhibited the acetylcholine esterase (AChE) activity. However, significant results were not observed at lower doses of the PEFR. The PEFR was safe up to the dose of 4000 mg/kg. These findings regarding the neuroprotective action suggest that F. religiosa could be used as an effective therapeutic agent in the management of NDDs. The same group of researchers also evaluated the anti-Parkinson’s activity of the PEFR in haloperidol- and 6-hydroxydopamine (6-OHDA)-induced experimental rat models [68]. A significant reduction in catalepsy induced by haloperidol was observed in the group pretreated with the PEFR indicating that the extract can protect dopaminergic neurotransmission in the striatum. A substantial increase in locomotor activity was observed at 200 and 400 mg/kg. However, the extract was able to reduce levels of malondialdehyde (MDA) and increase the levels of catalase (CAT), glutathione (GSH), and superoxide dismutase (SOD) at a high dose (400 mg/kg body weight) only, suggesting the antioxidant effect of the extract in the brain of 6-OHDA-treated animals.
The anti-amnesic and nootropic properties of the ethanolic extract of F. religiosa leaves (100 mg/kg body weight) were observed in amnesia and hypoxia induced by scopolamine and sodium nitrite in rodents. The results were comparable to the positive controls Piracetam (200 mg/kg) and Mentat (100 mg/kg) [70]. However, contradictory results were observed [71] after oral supplementation of the F. religiosa leaf extract in healthy mice where the extract considerably decreased the neuromuscular performance and object recognition ability of male mice only. The conflicting results were, perhaps, due to the extract supplementation to healthy animals as compared to disease models, signifying variable effects of the F. religiosa leaf extract in rodents with different health conditions [71].
In addition, the leaves of F. religiosa were also evaluated for the anticonvulsant effect [72], but the hydroethanolic extract failed to exhibit a protective effect in pentylenetetrazol (PTZ)- and maximal electroshock (MES)-induced mouse models [73].

3.2. Root

The saponin-rich fraction (SRF) of the hydroethanolic extract of F. religiosa roots (1, 2, and 4 mg/kg) displayed an anticonvulsant effect in mouse models of convulsions [74]. The study was further extended to study the effect of the SRF on cognitive decline and associated depression in a PTZ kindling mouse model of epilepsy [75]. The extract showed marked neuroprotection in the tail suspension test. The SRF considerably raised the monoamine levels and altered the levels of neurotransmitters (noradrenaline, serotonin, γ-aminobutyric acid, dopamine) in the brain. Oral administration of the aqueous root extract of F. religiosa (25, 50, and 100 mg/kg) exhibited a dose-dependent and anticonvulsant effect against strychnine- and PTZ-induced seizures. The researchers proposed the involvement of zinc and magnesium present in the extract in the anticonvulsant activity [76].

3.3. Fruit

A very high amount of serotonin is known to be present in the fruits of F. religiosa [77]. The role of serotonergic neurotransmission in the protection from seizures by modifying various GABAergic and glutamatergic activities is well-documented [78] and the reductions in brain serotonin levels lead to increased susceptibility to seizures [79]. Anticonvulsant studies with the methanolic fruit extract of F. religiosa (25, 50, and 100 mg/kg) showed substantial dose-dependent protection in picrotoxin- and MES-induced convulsion mouse models, with the activity similar (at 100 mg/kg) to that observed in the diazepam-treated group. However, PTZ-induced seizures were not inhibited by the extract [42].
The flavonoid-rich ethyl acetate fraction of the F. religiosa fruit extract (1, 2, and 4 mg/kg i.p.) was used along with a subeffective dose of phenytoin (15 mg/kg) in a PTZ-kindled mouse model. The extract completely suppressed the seizures and reduced the oxidative stress in the brain tissue by decreasing the levels of MDA and increasing the CAT and GSH levels. The extract also decreased the activity of AChE which is responsible for its memory-enhancing effect [75].
The methanolic fruit extract of F. religiosa (10, 50, and 100 mg/kg i.p.) displayed a dose-dependent anti-amnesic effect in a scopolamine-induced amnesia model of mice [80]. Additionally, inhibition of the anti-amnesic effect of the extract by cyproheptadine (a serotonin antagonist) demonstrated the association of serotonergic pathways for memory improvement by the methanolic extract.

3.4. Bark

Numerous plants were screened for AChE inhibitory activity in vitro. Among the screened plants, the methanolic extract of the F. religiosa stem bark extract (100–400 µg/mL) displayed the most potent AChE inhibitory activity with an IC50 value of 73.69 µg/mL [81].

4. Neuroprotective Effect of Ficus benghalensis

The neuroprotective effect of F. benghalensis extract is summarized in Table 1.

4.1. Leaves

Scarce literature is available on the neuroprotective action of F. benghalensis. The methanolic leaf extract of F. benghalensis (200 and 400 mg/kg p.o.) was evaluated for neuroprotective effects against alloxan-induced diabetic neuropathy in rats [82]. The treated animals had better motor coordination in response to stimuli as compared to the disease control group.

4.2. Bark

The methanolic F. benghalensis bark extract (100, 200, and 300 mg/kg i.p.) displayed a positive anti-amnesic, anti-anxiolytic, and antidepressant effects in a scopolamine-induced behavioral animal model [62]. The phytocompounds isolated from the extract were identified by GC–MS and might interact with glutamatergic, serotonergic, cholinergic, and GABAergic systems in the brain for memory-improving, anxiety-reducing, and depression-resolving activities observed in the study [62]. The aqueous bark extract (150 and 300 mg/kg body weight) had cognitive enhancement activity in scopolamine-induced amnesia in both old and young mice without any toxicity (up to 5 g/kg) [83].

4.3. Root

Oral administration of the aqueous root extract (200 mg/kg body weight) displayed anxiolytic, memory-enhancing, muscle-relaxant, and seizure-modifying effects without any toxicity in mice [84]. The phytochemicals present in the extract were suggested to affect muscarinic receptors in the brain.

5. Neuroprotection by Phytochemicals

A vast library of interesting chemicals has been identified in F. religiosa and F. benghalensis (Figure 1).

5.1. Amyrin

Higher concentration of β-amyrin is present in the B. ceiba leaf extract, which was reported to ameliorate various biochemical parameters (CAT, MDA, AChE) and cognitive functions in rats with scopolamine-induced amnesia [85]. The memory-enhancing effects of α- or β-amyrin from the mouse model of scopolamine-induced cognitive impairment involved the activation of extracellular signal-regulated kinase (ERK) and inhibition of glycogen synthase kinase (GSK-3β) in the hippocampus [86]. GSK-3β has a role in tau phosphorylation, which ultimately causes their detachment from microtubules and formation of aggregates [87]; hence, inhibition of GSK-3β is of significance in NDDs. The increased activity of ERK in response to oxidative stress and abnormal phosphorylation has been observed in AD [88]. In addition, both ERK and GSK-3β are known to play an important role in synaptic plasticity and memory processes [89,90]. Moreover, amyrin (25 and 50 mg/kg p.o.) also exhibited anticonvulsant activity by increasing the latency time to 75% and 101% at two doses, probably by inhibiting the protein kinase C (PKC) pathway and by increasing taurine (116% and 76%) and tyrosine (135% and 110%) in the basal ganglia and hippocampus, respectively, and decreasing glutamate (68%), aspartate (65%), and GABA (62%) in the basal ganglia [91]. The PCK pathway negatively regulated the expression of the GABAA receptors by affecting the ion channel function and receptor trafficking [92]. Additionally, amyrin exerted potent anxiolytic and antidepressant effects through the inhibition of monoamine oxidase (MAO) and elevating the GABA levels in the hippocampus [93]. MAO-B is known to have a key role in ROS generation, and its inhibitors (selegiline and rasagiline) are used in PD and AD treatment [94]. Hence, amyrin exerts neuroprotection mainly through enhancing the antioxidant pathway.

5.2. Azelaic Acid

Azelaic acid’s neuroprotective potential was evaluated in a rotenone-induced PD model (80 mg/kg p.o.) where a significant reversal in posture, muscular rigidity, and catalepsy was observed after the treatments. Additionally, synergistic effects of azelaic acid with levodopa and carbidopa (100 mg/kg + 25 mg/kg p.o.) [95] were revealed, indicating its promising role in treating PD. Azelaic acid has also been identified as a potential biomarker in urine for personalized healthcare in AD diagnosis [96] as its level correlates negatively with Aβ42 in cerebrospinal fluid (CSF) and positively with the CSF tau levels. It is also indicative of oxidative damage in the brain of AD patients that may account for changes in brain functions.

5.3. Bergapten (5-Methoxypsoralen)

Pieces of evidence support the neuropharmacological effects of bergapten in AD and depression [97,98]. The compound inhibits the AChE and butyrylcholinesterase (BChE) activities in vitro and in silico [99,100,101]. Moreover, bergapten (25 and 50 mg/kg body weight) also improved memory in a scopolamine-induced amnesia model and the cholinergic levels in the hippocampus and the prefrontal cortex without improving motor coordination and locomotor activity [102]. Interestingly, memory improvements were observed after acute and sub-chronic administrations of bergapten, which was independent of the AChE activity and dependent solely on the antioxidant activity of the compound [102]. Additionally, it also exhibited antidepressant effects by inhibiting MAO [103]. Bergapten (25, 50, and 100 mg/kg) reversed the paclitaxel-induced neuropathic pain by restoring the levels of oxidative stress markers (GSH, GST, iNOS, LPO) and downregulating the expression of inflammatory mediators (COX-2, TNF-α, NF-κB) [104]. Briefly, neuroprotection by bergapten is exerted mainly through its antioxidant and anti-inflammatory mechanisms. Additionally, higher bioavailability and the potential to cross the BBB make it a favorable candidate for treating neurological diseases [105,106].

5.4. Eudesmol

β-Eudesmol is a major phytocompound in Atractylodes lancea rhizome extracts, which induces neurite extensions in PC-12 cells at 100 and 150 µM concentrations by promoting transient phosphorylation of MAPKs (ERK1 and 2). P38-MAPK is primarily activated by inflammatory cytokines, playing a critical role in neuronal functions [107]. In addition, β-eudesmol also promotes inositol phosphatase accumulation and encourages the activation of phosphoinositide phospholipase C (PLC), which has a key role in nerve growth factor (NGF)-induced differentiation of neurons [108]. Besides, β-eudesmol (150 µM) induced phosphorylation of the cAMP-responsive element-binding protein (CREB), which is a critical regulator of neuronal plasticity and neuroprotection, in rat pheochromocytoma cells (PC-12) [109].
For the treatment of neurological disorders, low-molecular-weight compounds which can easily cross the BBB are preferred. Being a small molecule, β-eudesmol may prove an encouraging lead compound for studying neuronal functions. In short, β-eudesmol displays a distinctive effect on the nervous system by inducing various pathways that are critical for neuronal growth, plasticity, and protection. Studies on the efficacy of β-eudesmol to cross the BBB are in progress [110].

5.5. Eugenol

In traumatic brain injury (TBI) rats, pretreatment with eugenol (25, 50, and 100 mg/kg/day p.o., seven consecutive days) ameliorated the neurochemical and behavioral symptoms. Eugenol decreased lipid peroxidation and improved memory and motor activities in the treated group [111]. Eugenol pretreatment (50 and 100 mg/kg) also mitigated the cerebral ischemia/reperfusion (I/R) damage by inducing autophagy activities through the AMPK/mTOR/P70S6K signaling pathway [112]. AMPK and mTOR are the regulators of autophagy, which works through Unc-51-like kinase 1 (ULK1) activation [113]. Activation of AMPK and inhibition of mTOR endorses the autophagy activities. Additionally, P70S6K (ribosomal protein S6 kinase beta-1) is a downstream kinase of mTOR, whose suppression may promote autophagy activities [114].
In another study, eugenol treatment (10 mg/kg bw for 5 weeks) improved the gait in acrylamide-induced neuropathic rats and restored the levels of antioxidant enzymes and dopamine in the brain [115]. Furthermore, cotreatment with eugenol (6 mg/g) in aluminum-induced toxicity in rat brains reduced the AChE, TNF-α, and caspase-3 expression [116]. Additionally, it promoted neurogenesis in the hippocampus by increasing the expression of the metallothionein gene (MT-III) and restored the levels of brain-derived neurotrophic factor (BDNF) and serotonin in the brain [117]. Since BDNF is vital for the preservation of cortical neurons, its initial loss would lead to short-term memory decline in AD [118]. BDNF expression could be induced by NGF through phosphorylating the CREB in the ERK/AKT signaling pathway [119]. Eugenol (0.1, 1, and 10 mg/kg orally) also displayed neuroprotective properties in a hydroxydopamine-induced PD model [120]. In an in vitro ThT (thioflavin T) assay, eugenol (3 mM) also suppressed amyloid formation by delaying the conversion to the β-sheet form [121].
In another interesting study, eugenol and its analogs were reported to interact with vanilloid receptors in the olfactory bulb and displayed a positive effect on memory [122]. Such a possibility can be explored for the treatment of AD and PD.
Eugenol appears to be a wonder molecule because of its ability to cross the BBB [116] and displays the neuroprotective effect through neurogenesis, antioxidant, anti-amyloid, and antiapoptotic ability by affecting multiple pathways. This multitarget approach expands the application of eugenol to multiple neurological diseases.

5.6. Kaempferol

Kaempferol is a multipotential neuroprotective agent that affects various pathways in NDDs [123]. In neurological diseases, the antioxidant activity of kaempferol inhibits various metalloproteases (MMP-2, MMP-3, MMP-9) and protects the BDNF modulations responsible for neuronal plasticity. It is known for enhancing cognitive performance in animal models by displaying anti-AChE activity [124,125]. Kaempferol and its degradation products inhibit Aβ oligomerization and plaque formation by interacting with the Aβ protein without involving the Lys16 and Lys18 residues [126] and restores the levels of oxidative stress markers (SOD, glutathione, ROS) both in vitro and in vivo [127]. It also displays its neuroprotective potential at 20 mg/kg dose by suppressing microglial activation by inhibiting the NF-κB, MAPK, AKT, and toll-like receptor 4 (TLR4) pathways in an LPS-induced striatum injury mouse model [128].
In a rotenone-induced PD rat model, kaempferol (10 and 20 mg/kg) increased the SOD and glutathione peroxidase (GSH-PX) levels and decreased lipid peroxidation. It neutralized ROS by blocking apoptosis through the JNK/MAPK pathway [129]. It also reduced the expression of proinflammatory cytokines, COX-2, and the high mobility group box 1 (HMGB1)/TLR4 inflammatory pathway [128]. As a result, motor coordination and dopamine levels increased in the PD model [130]. In summary, kaempferol displays multitarget properties (antioxidant, anti-inflammatory, anti-amyloid, antiapoptotic, and modulating enzymes involved in neurotransmission) for neuroprotection. Moreover, due to its ability to cross the BBB [131], it can prove beneficial in the treatment of NDDs.

5.7. Lanosterol

In vitro and in silico studies have indicated that lanosterol can suppress the buildup of misfolded protein aggregations/sequestosomes [132,133] by promoting autophagy activities [134]. Lanosterol (0.5 mM) also protected dopaminergic neurons from 1-methyl-4-phenylpyridinium (MPP+)-induced cell death in a PD cellular model by inducing mild mitochondrial depolarization and promoting autophagy activities. The observed reallocation of lanosterol synthase to mitochondria suggested that lanosterol might exert its survival effect by regulating mitochondrial functions [135]. In addition, reduced levels of lanosterol were observed in the striatum and ventral midbrain regions from 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP)-treated mice (PD model), suggesting a potential role in the cholesterol metabolism of NDD pathogenesis. In short, lanosterol promotes autophagy activities by sequestering misfolded proteins.

5.8. Leucoanthocyanins

Leucoanthocyanins are the intermediates in the synthesis of anthocyanins, which, in turn, are reported to contain multitarget neuroprotective properties in animal and in vitro models of NDDs [136,137,138,139] by reducing oxidative stress and modulating anti-inflammation [140,141] through the phosphoinositide 3-kinase (PI3K)/protein kinase B (Akt)/nuclear factor erythroid 2-related factor 2 (Nrf2)/heme oxygenase-1 (HO-1) pathway [55] and cyclooxygenase-2/microsomal prostaglandin E synthase-1 (COX-2/mPGES-1) [142]. The phosphorylation of PI3K activates Akt, which, in turn, disables GSK-3β by phosphorylation (p-GSK-3β). The latter helps in the translocation of Nrf2 to the nucleus, where it enhances the expression of antioxidant genes (including HO-1) [143,144,145]. Furthermore, anthocyanins also promote autophagy by upregulating expression of autophagy-related proteins through the AMP-activated protein kinase and mammalian target of rapamycin (AMPK–mTOR) signal pathway [146]. Evidence also suggests the BBB-crossing capability of anthocyanins [147] which would be an advantage in treating NDDs. In summary, leucoanthocyanins exert their neuroprotection through antioxidant, antiapoptotic, and anti-inflammatory pathways.

5.9. Limonene

Limonene (10 µg/mL) presents a neuroprotective potential against exposed primary cortical neurons to Aβ1–42 oligomers (in vitro model of AD) by mitigating ROS generation and the potassium channel (KV3.4) hyperfunction [148]. In AD, enhanced ROS eventually cause Aβ1–42-induced upregulation of KV3.4. In addition, Ca2+-induced ROS production activates KV3.4 through the NF-κB pathway [149]. Limonene present in the essential oil mixture (MO: 1% and 3%) has been known to revert cognitive deficits in the scopolamine-induced amnesia rat model by alleviating the oxidative stress markers (MDA, SOD, GSH) and inhibiting AChE (24.9%) and BchE (69.1%; IC50, 1.096 ± 0.043 µg/mL) activities. Administration of limonene (5, 25, and 50 mg/kg for 1 week) also significantly increases GABA, a key hypothalamic neurotransmitter, in the rat brain. This increased activity inhibits the release of corticosterone from the hypothalamic–pituitary–adrenal (HPA) axis under stress conditions, thus playing a vital role as an antistress agent [150]. Molecular docking models have revealed the van der Waals interaction between limonene and active side residues (Ser198, His438, Leu286, Val288, Phe329) of BchE [151]. Different plant extracts with high limonene concentrations also exhibit neuroprotection by endorsing potential anti-AchE, anti-inflammatory, and antioxidant activities [152,153,154].

5.10. Linalool

Linalool has shown a protective effect in various neurodegenerative models and is reported to have the ability to cross the BBB [155]. In in vitro experiments, linalool (10 µg/mL) protected PC-12 cells from Aβ exposure by reducing ROS and inhibiting the activity of proapoptotic caspase-3 [156]. Linalool (50 and 100 mg/kg/day for 21 days) also significantly suppressed Aβ-induced ROS, oxidative stresses, and inflammatory responses in an AD fly model [157] without altering the amount of Aβ in the brain. Alterations in the hippocampal phospholipid profiles in ischemic animals were mitigated by linalool, which helped to maintain the phospholipid homeostasis, hence recovering brain functions [158]. The oral administration of linalool (25 mg/kg for 3 months) in a triple transgenic AD mouse model (3xTg-AD) restored memory in the treated animals via reducing β-amyloidosis, astrogliosis, and tauopathy besides reducing proinflammatory markers (MAPK, inducible oxide nitric synthase (iNOS), COX-2, IL-1β) [159]. Linalool (100 µM) is known to modulate glutamatergic neurotransmission by interacting with NMDA receptors [160]. Furthermore, it proved to be neuroprotective in a glutamate-induced oxidative stress in vitro model (HT-22 cells) by reducing ROS, calcium production, and LPO levels [161]. Linalool (162, 324, 648 μM) exerts protective effects in LPS-induced BV2 microglial cells by Nrf2 activation [162] and in mice with Aβ-induced cognitive deficits by restoring the levels of oxidative stress-related enzymes, suppressing caspase-3, and upregulating the Nrf2 and HO-1 expression [163]. Together, these findings suggest that linalool could be used for the development of NDD drugs as it can cross the BBB and provide neuroprotection through anti-inflammatory, antioxidant, and antiapoptotic properties.

5.11. Lupeol

Lupeol (triterpenoid) is reported for its antioxidative, anti-inflammatory, and neuroprotective activities in a variety of animal models [164]. Lupeol competitively inhibited β-secretase (BACE-1) (IC50, 5.12 μmol/L) with a low inhibition constant (Ki 1.43 μmol/L), indicating better affinity. Molecular docking revealed the formation of hydrogen bonds between the hydroxyl group of lupeol and the Asp32 and Ser35 of BACE-1 [165], upregulating the expression of proinflammatory cytokines and interleukins (TNF-α, IL-6, IL-1β) at 25, 50, and 100 mg/kg p.o. concentrations in acetic acid-induced writhing, the formalin test, carrageenan-induced hyperalgesia, and a postoperative pain model [166], activating the Nrf2/HO-1 pathway and improving cognitive functions (at a 50 mg/kg dose) in Aβ-induced oxidative stress in mice [167]. Additionally, lupeol (50 mg/kg dose p.o.) also inhibited apoptotic signaling molecules (caspase-3, BCL2-associated X (Bax), cytochrome c) and repressed astrocytes/microglia activation in the cortex and the hippocampus of a TBI mouse model [168]. Lupeol (0.1 µM) downregulated the anti-inflammatory responses of TNF-α, iNOS, and NOD-like receptor pyrin domain-containing protein 3 (NLRP3) and upregulated the arginase, IL-6, neurotrophin (glia-derived neurotrophic factor (GDNF)), and sonic hedgehog–Gli (SHH–GLI) signaling [169] in cerebellar cultures and induced neuroprotection. Additionally, lupeol displayed a superior ADMET (absorption, distribution, metabolism, excretion, toxicity) profile and proved nontoxic, noncarcinogenic, biodegradation-resistant, as well as low inhibition by cytochrome P450 (CYP450). Most importantly, it can cross the BBB easily [170].
Hence, lupeol seems to be a potent candidate for NDD drug discovery as it has an acceptable ADMET profile and the multitarget neuroprotective approach (inhibiting neuroinflammation, reducing oxidative stress, repressing apoptosis and microglial activations).

5.12. Myo-Inositol (Vitamin B8)

Even though no direct study has been conducted to investigate the effect of myo-Inositol on the NDD model systematically, it displays a neuroprotective role in ischemic stroke injury in animals exposed to tobacco smoke and in streptozotocin-induced mice by increasing the motor functions after the stroke using in situ brain perfusion and the acute brain slice method at 0.1 μCi/mL [171,172]. In addition, myo-inositol (30 mg/kg for 28 days) ameliorated spatial learning and memory deficits by attenuating cell loss in the hippocampus in a kainic acid-induced epilepsy rat model [173] by a multitarget approach including preserving neuronal circuits and activation of GABA signaling. Myo-inositol also presented the anticonvulsive property in thiosemicarbazide models of seizures by increasing the latent time and decreasing the severity of the seizure [174].
Interestingly, myo-inositol has been considered a noninvasive early marker for assessing various asymptomatic AD stages in comparison with magnetic resonance spectroscopy (MRS) [175]. Increased myo-inositol/creatine levels have been observed in healthy apolipoprotein E E4 genotype (APOE ε4) carriers with normal CSF Aβ42 levels in comparison to ε4 non-carriers, signifying the importance of myo-inositol levels in assessing AD before a noticeable amyloid pathology [176]. Reports suggest that myo-inositol can also be transported across the BBB by simple diffusion as well as through a stereospecific transporter [177].

5.13. Myricetin

Myricetin is a lipophilic compound with the ability to cross biological membranes, including the BBB [178]. It is reported to target AD by inhibiting multiple pathways, such as neuroinflammation, autophagy, oxidative stress, chelation, anti-AchE, and Aβ depositions [179]. Myricetin (5 μM) averted Aβ1–42 oligomer-induced neurotoxicity in human neuroblastoma cells (SH-SY5Y) by exerting antioxidant effects on the cell membranes and mitochondria. Myricetin also restored mitochondrial dysfunctions by decreasing ROS, increasing the expression of manganese superoxide dismutase (Mn-SOD) and ATP generation [180]. The JNK/stress-activated protein kinase (SAPK) pathway is activated by oxidative stress and Aβ. Enhanced BACE-1 levels lead to an increase in Aβ levels [181], which eventually activates the JNK/SAPK pathway, resulting in a vicious cycle of the NDD. Furthermore, oxidative stress could also endorse serine/threonine protein phosphatase (PP2A) inhibition, promoting tau phosphorylation [182] and damaging the mitochondria [183]. Myricetin inhibited the BACE-1 activity (IC50, 2.8 μM) to cleave the amyloid precursor protein (APP), increase α-secretase, and decrease Aβ production/oligomerization [184,185]. Additionally, myricetin also interferes with the NF-κB and AMPK/SIRT1 signaling pathway and reduces the levels of inflammatory mediators (IL, TNF-α, iNOS, COX-2) in the brain [186]. Myricetin (1 and 10 μM) also aids in the removal of abnormal Aβ and tau through autophagy activation by inhibiting phosphorylation of mammalian targets of rapamycin (mTOR) in primary neuron cultures [187]. Interestingly, myricetin (25 μM) can regulate the levels of metal ions in the brain to reduce their interactions with Aβ and disassemble the formation of the metal–Aβ complex [188]. Since it was previously reported that Fe2+ could activate microglia by increasing neuroinflammation, the complexation of myricetin with Fe2+ could reduce the inflammatory processes and inhibit the expression of transferrin receptor I (TrP1), thus lowering the iron levels [6]. Lastly, myricetin also improves learning and cognition through the anti-AchE activity. Some proinflammatory cytokines (IL-1) activate AchE causing the Ach levels to decrease in the brain, affecting memory. Hence, the anti-inflammatory activity of myricetin also improves memory [189,190].

5.14. Pinene

Pinene, a monoterpene, can cross the BBB and affect multiple neurotransmitter systems, such as adrenergic, cholinergic, dopaminergic, GABAergic, serotoninergic, and noradrenergic functions in the brain [191]. According to a cell-based study, pretreatment of α-pinene (10 and 25 µM) in PC-12 cells inhibited ROS by increasing the expression of antioxidant enzymes and reducing apoptosis by decreasing the caspase-3 activity [192]. α-Pinene administration (50 mg/kg i.p.) in the Aβ-induced rat model reduced neuroinflammation by overturning the TNF-α/NF-κB pathway and improved memory and learning. Moreover, α-pinene also upregulated the expression of both the nicotinic acetylcholine receptor (nAChR) α7 subunit and BDNF [193], which play a role in the survival and maintenance of neurons. Additionally, α-pinene (100 mg/kg i.p.) restored the levels of antioxidant enzymes (SOD, CAT, GPX) and reduced NO, IL-6, and MDA in the brain of focal ischemic stroke model rats [194]. Furthermore, α-pinene (100 mg/kg i.p.) downregulated Bax with a corresponding upregulation of Bcl-2 expression, resulting in suppression of apoptosis in a rat model of cerebral ischemia-reperfusion [195].
In a mouse model of memory impairment, α-pinene upregulated the expression of AchE, muscarinic receptors, and antioxidant transcription factors in the hippocampus, improving spatial recognition and memory [196]. In summary, the antioxidant, anti-inflammatory, and antiapoptotic properties of α-pinene would help treat NDDs.

5.15. Psoralen

Psoralen is a natural furanocoumarin with in vitro competitive inhibitory activity against AchE (IC50, 370 μg/mL) [197]. A molecular docking study revealed a stable AchE–psoralen complex with π–π stacking (Tyr334) and hydrogen bonding (Gly119 and Gly118) interactions [197]. Psoralen also displayed inhibitory activities towards MAO-A (IC50, 15.2 µM; noncompetitive inhibition) and MAO-B (IC50, 61.8 µM; competitive inhibition) in the rat brain [198]. Since MAO is an important enzyme in maintaining levels of monoamine neurotransmitters in the brain, its reduced expression results in decreased Aβ depositions and oxidative stress [199]. In another study, psoralen and iso-psoralen-rich extracts (0.1 and 0.3 mg/kg) of Psoraleae fructus improved amnesia in scopolamine-induced rats, apparently by AchE inhibition (IC50, 1.12 mM) and activation of cholinergic neuronal functions [200]. In summary, even though moderate enzyme inhibitory activity has been observed with psoralen, it could serve as a lead molecule to synthesize other potential analogs to treat NDDs.

5.16. Quercetin

Quercetin (20 μM) decreases Aβ production in primary neuron cultures by inhibiting BACE-1 (IC50, 5.4 μM). A molecular docking study identified the interaction of quercetin with the Asp32 of BACE-1 [201], which resulted in cognitive improvement in the animal model of NDD [202]. Additionally, quercetin (20 and 40 μM) exerted neuroprotective effects by protecting proteins and lipids from oxidation in case of Aβ1–42-induced toxicity [203]. Quercetin (10 and 30 μM) also safeguarded mitochondria by reducing the production/accumulation of ROS, NO, increasing GSH, decreasing overexpression of proinflammatory cytokines, and reducing dopaminergic degeneration of neurons in MN9D cells (mouse dopaminergic cell line) and a MitoPark PD animal model [204]. A positive effect of quercetin supplementation was also observed in a 3-NP-induced Huntington’s disease model (HD). Quercetin (25 mg/kg p.o. for 21 days) restored the levels of ATP, CAT, and SOD, relieving mitochondrial oxidative stress. Histopathological studies reported diminished striatum astrogliosis and pyknotic nuclei in the HD model [205]. Quercetin supplementation seemed to mitigate the biochemical and neurochemical changes in the rat brain by altering/inhibiting inflammatory activities from NDDs [206]. The targeted downstream pathways of quercetin [207] for neuroprotection are as follows: paraoxonase 2 (PON2) [208,209], Nrf2-ARE; phosphoinositide 3 kinase (PI3K) [210], JNK/ERK [211], TNF-α [212], peroxisome proliferator-activated receptor gamma coactivator 1-alpha (PGC-1α), and SIRT1 [213], CREB [214], MAPK [215], NF-κB [210], and AMPK [216].
In essence, quercetin presents a multitarget tactic (anti-inflammatory action, antioxidant action, and enzyme inhibition) to ameliorate the symptoms of NDDs.

5.17. Rhein

Rhein (12 mg/kg) displayed antioxidative and neuroprotective effects in controlled cortical impact (CCI) rats by reducing the MDA levels and increasing SOD, CAT, and GSH [163], thereby protecting the brain from oxidative damage. It also protected the BBB through the NADPH oxidase/ROS/ERK/MMP-9 signaling pathway [217]. Additionally, in an I/R rat model, rhein (50 and 100 mg/kg/day for 3 days) restored the levels of antioxidant markers, enhanced Bcl-2, and decreased the levels of Bax and caspase-3 [218]. Rhein (100 mg/kg i.p.) also exerted an anti-inflammatory response by downregulating the inflammatory cytokines in the cortex of animals with TBI [219]. Additionally, rhein (10 mg/kg i.v.) improved the cognitive decline in an APP/PS1 mouse model of AD by activating the SIRT1/PGC-1α pathway, improving oxidative stress by regulating mitochondrial biogenesis [220]. It also exerted an anti-inflammatory response by subsiding the levels of TNF-α and IL-1β in the hippocampus and an antioxidant property by decreasing MDA, increasing GSH, the GSH/GSSG ratio, CAT, and GSH-Px in the same model [220]. The pharmacokinetics of rhein disclose that it can be easily transported by body fluids and can pass through damaged BBB which is very useful in case of brain injury [221]. In summary, rhein could be an important compound for neuronal protection through antioxidant, anti-inflammatory, and antiapoptotic properties.

5.18. Rutin

Rutin has emerged as an important pharmacological flavonoid in various NDDs [222] as it can easily cross the BBB [223]. Rutin improves memory by reducing Aβ oligomerization, oxidative stress, neurotoxicity, and neuroinflammation in several AD animal models [224,225,226,227]. The anti-amyloidogenic property could be due to the destabilization of Aβ by direct interaction of the aromatic ring of rutin with the hydrophobic β-sheet of amyloid aggregates. Rutin (100 μL rutin suspension/per 10 g bw for 30 days) decreased the tau levels by regulating phosphorylation through increased PP2A activities. It also reduced inflammation by downregulating NF-κB besides securing neuronal morphology and improving cognition through synapse preservation in the brain of the Tau-P301S mouse model [228]. In addition to the AD models, rutin (25 mg/kg bw orally for 3 weeks) protected dopaminergic neurons by reducing oxidative stress and apoptosis in the 6-OHDA-induced rat model of PD [229,230]. It also improved HD symptoms by activating autophagy and the insulin/insulin-like growth factor I (IGF-1) pathway at 15–120 μM concentrations [231]. APP phosphorylation is suppressed by insulin/IGF-I in vitro which favors the non-amyloidogenic pathway [232]. This pathway regulates neurogenesis and is known to improve neuronal survival, learning, and memory through the PI3K/Akt and ERK pathway [233,234,235].
In brief, rutin provides neuroprotection through its antioxidant, anti-inflammatory, anti-amyloid, and autophagic activation properties.

5.19. Stigmasterol

Stigmasterol, a phytosterol, demonstrated its neuroprotective effects through multitarget approach in vitro and in vivo. Studies revealed that phytosterols can cross the BBB through the scavenger receptor class B member 1 (SR-BI)-dependent pathway and via apolipoprotein E (ApoE) [236], improving the cognition [237]. It reduced amyloid plaque formations [238], inhibited the AchE activity (IC50, 644 μM) in vitro [239], and decreased the elevated ROS [240]. Stigmasterol (1 μM) displayed its neuroprotective effects in hydrogen peroxide-induced oxidative stresses in SH-SY5Y cells by modulating the sirtuin 1-forkhead box O3a (SIRT1-FoxO3a) pathway [241]. SIRT1 activates FoxO3a, which consequently stimulates the production of antioxidant enzymes (SOD, CAT), hence protecting against oxidative stress. Stigmasterol also exhibited anti-inflammatory activity in the IL-1β-treated cells by inhibiting proinflammatory cytokines without affecting the levels of anti-inflammatory cytokines suggesting its role in the NF-κB inflammatory pathway [242]. In a recent report, stigmasterol (10 and 20 μM) inhibited NF-κB and NOD-like receptor thermal protein domain-associated protein 3 (NLRP3) signaling by activating AMPK, thereby reducing the Aβ-induced inflammatory response in BV2 cells [243]. Stigmasterol (3, 10, and 30 mg/kg) also exerted ameliorating effects on the scopolamine-induced memory loss in mice through the cholinergic neurotransmission augmentation by N-methyl-D-aspartate receptor (NMDA) activation [244]. Earlier reports had shown progressive modulatory effects of Ach by binding to muscarinic receptors on NMDA, enabling NMDA receptor-mediated synaptic plasticity and long-term potentiation (LTP), which are implicated in learning and memory [245]. Besides, stigmasterol (10 mg/kg p.o.) has been reported to increase the ERK and CREB phosphorylation in the hippocampus of the scopolamine-induced memory loss model in mice, which could also influence a positive effect on memory and learning [244,246].

5.20. Synephrine

Synephrine is a sympathomimetic alkaloid with mild CNS stimulant properties [247]. When the effects of oral supplementation of synephrine (20 mg) were evaluated for cognition and exercise performances during a pre-workout, no significant results were observed on muscular endurance during exercise, but it seemed to improve cognitive functions and mental focus [248]. Synephrine has effective inhibitory properties towards AchE (IC50, 226.01 nM) and BchE (IC50, 92.22 nM) in vitro [249].

5.21. β-Caryophyllene

β-Caryophyllene is a natural sesquiterpene that can easily cross the BBB and exert neuroprotective effects [250]. Additionally, β-caryophyllene (48 mg/kg p.o. for 7 weeks) reduced Aβ deposition in the cerebral cortex and the hippocampus of APP/PS1 mice. It also reduced the levels of COX-2, TNF-α, and IL-β in the cerebral cortex. β-Caryophyllene is a known cannabinoid receptor 2 (CB2) agonist, and the activation of CB2 receptors is beneficial in reducing neuroinflammation by triggering the peroxisome proliferator-activated receptor-γ (PPARγ) pathway [251]. β-Caryophyllene (5 μM) also inhibited the hypoxia-induced neuroinflammatory processes in BV2 cells by mitigating ROS production and proinflammatory cytokines by inhibiting p38MAPK/NF-κB [252]. β-Caryophyllene (24, 72 mg/kg i.p.) prevented neuronal necrosis, downregulated the receptor-interacting protein kinase-1 and -3 (RIPK1, RIPK3), MLKL phosphorylation, HMGB1, TLR4, and proinflammatory cytokines. Thus, β-caryophyllene exerts neuroprotection by inhibiting inflammation and neuronal death in a cerebral I/R injury mouse model [253]. Caryophyllene is generally considered safe (GRAS) by the FDA for use in the food industry [254].

5.22. β-Sitosterol

β-Sitosterol displayed robust anti-AchE (55 µg/mL) and anti-BchE (50 µg/mL) activity both in vitro and in silico. The in vivo results confirmed that β-sitosterol acts as a free radical scavenger and can reach the brain and inhibit AchE and BACE [255,256]. Its antioxidant effects (15 µM) were observed in a glucose oxidase (GOX)-induced oxidative stress and lipid peroxidation model of HT22 hippocampal cells through the estrogen receptor (ER)-mediated PI3K/GSK-3β signaling pathway [257]. β-Sitosterol seemed to help the PI3K recruitment to the lipid raft, an important region of the membrane in signal transduction [258]. GSK-3β is an important downstream target of PI3K upregulation which eventually increases the intracellular glutathione, a natural antioxidant [259]. Additionally, β-sitosterol augmented the mitochondrial membrane potential (ΔΨm) and adenosine triphosphate (ATP) by integrating into the mitochondrial membrane [257].
In NDDs, neurons are damaged due to persistent neuroinflammation. In a study, β-sitosterol (8 and 16 µM) displayed anti-inflammatory properties in lipopolysaccharide (LPS)-induced BV2 microglial cells by reducing the expression of proinflammatory factors (IL-6, iNOS, TNF-α, cyclooxygenase-2 (COX-2), an inhibitor of nuclear factor kappa B: IκB, NF-κB, ERK/p38) [260]. In animal models, β-sitosterol exhibited positive effects on learning and memory [255,261,262], and prevented plaque deposition in an amyloid protein precursor/presenilin 1 (APP/PS1) model [255,261,262]. Recently, substantial anxiolytic and antidepressant effects of γ-sitosterol were observed from the strong binding affinity with the human serotonin receptor in a molecular docking study [263]. In short, the neuroprotective effects of β-sitosterol would be due to its antioxidant, anti-inflammatory, anti-amyloid, and enzyme inhibition properties. It seems to be a potential candidate for managing memory-related disorders in NDDs. Pan-assay interference structures (PAINS) are chemical compounds which could give a significant false positive signal in the drug screening processes including redox reactivity, fluorescence of small compounds, and covalent changes of target proteins [264,265]. For instance, curcumin, known as the representative PAINS, has failed more than 120 clinical studies for various diseases due to false activity in vitro and in vivo [266]. Moreover, at least 15 studies have been retracted and dozens more have been corrected since 2009 [267]. As a result, it is imperative for accurate assessment of these compounds with structural alerts and elimination from further steps of the drug discovery process [268]. Medicinal chemistry analysis of PAINS was processed by the SwissADME server [269]. Except for myricetin, quercetin, rhein, and rutin, all the phytocompounds derived from Ficus trees were predicted to have no structural alerts or false positive signals (Table 2).
The summary of neuroprotective mechanisms of the discussed phytocompounds from sacred Ficus trees is described in Table 2 and Figure 2.

6. Conclusions and Future Directions

Even though plant extracts have been used for centuries for treating a spectrum of diseases in traditional medicine, no scientific validation for their therapeutic effects has been presented. In the modern era, the use of plants for therapeutic purposes was underestimated initially. Recently, after the discovery and use of numerous important drugs from herbal sources [270], such as quinine (an antimalarial drug from the Cinchona bark), atropine (anticholinergic medicine from Belladona), digoxin (obtained from Digitalis to treat cardiac arrhythmia), colchicine (extracted from Colchicum for treating gout), and galantamine (AChE inhibitor from Galanthus spp.), the interest in plant-based research has been expanding to determine their mechanistic actions behind the therapeutic potentials and has reached a new height. In an extract, a cocktail of various bioactive compounds may exhibit synergistic effects for better therapeutic activity, questioning whether a purified compound would present similar effects. However, an understanding of a purified bioactive compound should be performed first, even if a cocktail of compounds may be needed down the line. In addition, since the pathophysiology of NDDs would be complex, linking multiple cellular events, a multitarget tactic might be a better approach going forward.
Several bioactive compounds in the sacred Ficus species display neuroprotective properties in vitro and in vivo through multiple pathways, such as antioxidant (PI3K, AKT, AMPK, PKC, ERK, HO-1, Nrf2), anti-inflammatory (SIRT1, NF-κB), anti-amyloid (APP, BACE-1), antiapoptotic (Bcl-2, Bax, caspase), and modulating enzymes in neurotransmission (AChE, BChE). Most of the studies were carried out using rodent models of NDDs that unfortunately cannot recapitulate the complete aspects of AD as it is a uniquely human disease [271]. This is one of the reasons why many drugs were unsuccessful in clinical trials. Therefore, performing research using human tissues would give a more human-centric strategy.
Additionally, the negative results of the anti-Aβ strategy (amyloid cascade hypothesis) in clinical trials on AD patients demonstrate that it is not the only pathogenic factor involved. Pulling down cerebral Aβ can only delay cognitive decline but it cannot stop it, indicating the role of other etiological factors (oligomer cascade hypothesis) in the pathogenesis of AD [272]. Hence, scientists must follow a multitarget approach, leading to the treatment of complex diseases, such as NDDs.
In conclusion, Ficus spp. extracts and bioactive compounds present effective neuroprotective properties (in vivo and in vitro) by modulating several important pathways. The results from in vivo experiments also indicate the nontoxic nature of the extracts/phytocompounds at the doses tested. Yet, they have not been translated into clinical trials. Henceforth, the need to take the research to next level is of great significance in the treatment of NDDs.

Author Contributions

Conceptualization, N.S. and S.S.A.A.; writing—original draft preparation, N.S.; writing—review and editing, N.S., K.H.S. and S.S.A.A.; funding acquisition, S.S.A.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Research Foundation of Korea and by the Korean Government (2020R1A2B5B01002463, 2021R1A6A1A03038996, and 2021R1A2C1093218).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cowan, M.C.; Raymond, L.A. Selective neuronal degeneration in Huntington’s disease. Curr. Top. Dev. Biol. 2006, 75, 25–71. [Google Scholar]
  2. Erkkinen, M.G.; Kim, M.-O.; GeWschwind, M.D. Clinical neurology and epidemiology of the major neurodegenerative diseases. Cold Spring Harb. Perspect. Biol. 2018, 10, a033118. [Google Scholar] [PubMed] [Green Version]
  3. Ragagnin, A.M.; Shadfar, S.; Vidal, M.; Jamali, M.S.; Atkin, J.D. Motor neuron susceptibility in ALS/FTD. Front. Neurosci. 2019, 13, 532. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Salvadores, N.; Gerónimo-Olvera, C.; Court, F.A. Axonal degeneration in AD: The contribution of Aβ and Tau. Front. Aging Neurosci. 2020, 12, 581767. [Google Scholar] [CrossRef] [PubMed]
  5. Pedersen, J.T.; Heegaard, N.H. Analysis of protein aggregation in neurodegenerative disease. Anal. Chem. 2013, 85, 4215–4227. [Google Scholar] [CrossRef]
  6. Wang, B.; Zhong, Y.; Gao, C.; Li, J. Myricetin ameliorates scopolamine-induced memory impairment in mice via inhibiting acetylcholinesterase and down-regulating brain iron. Biochem. Biophys. Res. Commun. 2017, 490, 336–342. [Google Scholar] [CrossRef]
  7. Jarosińska, O.D.; Rüdiger, S.G. Molecular strategies to target protein aggregation in Huntington’s disease. Front. Mol. Biosci. 2021, 8, 769184. [Google Scholar] [CrossRef]
  8. Robinson, J.L.; Geser, F.; Stieber, A.; Umoh, M.; Kwong, L.K.; Van Deerlin, V.M.; Lee, V.M.-Y.; Trojanowski, J.Q. TDP-43 skeins show properties of amyloid in a subset of ALS cases. Acta Neuropathol. 2013, 125, 121–131. [Google Scholar] [CrossRef]
  9. Migliore, L.; Coppedè, F. Genetics, environmental factors and the emerging role of epigenetics in neurodegenerative diseases. Mutat. Res./Fundam. Mol. Mech. Mutagen. 2009, 667, 82–97. [Google Scholar] [CrossRef]
  10. Tan, M.A.; Sharma, N.; An, S.S.A. Phyto-Carbazole Alkaloids from the Rutaceae Family as Potential Protective Agents against Neurodegenerative Diseases. Antioxidants 2022, 11, 493. [Google Scholar] [CrossRef]
  11. Bagyinszky, E.; Kang, M.J.; Pyun, J.; Van Giau, V.; An, S.S.A.; Kim, S. Early-onset Alzheimer’s disease patient with prion (PRNP) p. Val180Ile mutation. Neuropsychiatr. Dis. Treat. 2019, 15, 2003. [Google Scholar] [CrossRef] [PubMed]
  12. Giau, V.V.; Bagyinszky, E.; Yang, Y.S.; Youn, Y.C.; An, S.S.A.; Kim, S.Y. Genetic analyses of early-onset Alzheimer’s disease using next generation sequencing. Sci. Rep. 2019, 9, 8368. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Bagyinszky, E.; Yang, Y.; Van Giau, V.; Youn, Y.C.; An, S.S.A.; Kim, S. Novel prion mutation (p. Tyr225Cys) in a Korean patient with atypical Creutzfeldt–Jakob disease. Clin. Interv. Aging 2019, 14, 1387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Wang, M.J.; Yi, S.; Han, J.-y.; Park, S.Y.; Jang, J.-W.; Chun, I.K.; Kim, S.E.; Lee, B.S.; Kim, G.J.; Yu, J.S. Oligomeric forms of amyloid-β protein in plasma as a potential blood-based biomarker for Alzheimer’s disease. Alzheimer’s Res. Ther. 2017, 9, 98. [Google Scholar] [CrossRef] [Green Version]
  15. Van Giau, V.; An, S.S.A. Emergence of exosomal miRNAs as a diagnostic biomarker for Alzheimer’s disease. J. Neurol. Sci. 2016, 360, 141–152. [Google Scholar] [CrossRef]
  16. Hornykiewicz, O. A brief history of levodopa. J. Neurol. 2010, 257, 249–252. [Google Scholar] [CrossRef]
  17. Lahlou, M. The success of natural products in drug discovery. Pharmacol. Pharm. 2013, 4, 17–31. [Google Scholar] [CrossRef] [Green Version]
  18. Nawaz, H.; Waheed, R.; Nawaz, M. Phytochemical composition, antioxidant potential, and medicinal significance of Ficus. Mod. Fruit Ind. 2020, 1, 20. [Google Scholar]
  19. Rahman, A.; Khanom, A. Taxonomic and ethno-medicinal study of species from Moraceae (Mulberry) Family in Bangladesh Flora. Res. Plant Sci. 2013, 1, 53–57. [Google Scholar]
  20. Pierantoni, M.; Tenne, R.; Rephael, B.; Brumfeld, V.; van Casteren, A.; Kupczik, K.; Oron, D.; Addadi, L.; Weiner, S. Mineral deposits in Ficus leaves: Morphologies and locations in relation to function. Plant Physiol. 2018, 176, 1751–1763. [Google Scholar] [CrossRef] [Green Version]
  21. Gopukumar, S.; Praseetha, P. Ficus benghalensis Linn–the sacred Indian medicinal tree with potent pharmacological remedies. Int. J. Pharm. Sci. Rev. Res. 2015, 32, 223–227. [Google Scholar]
  22. The Holy Bible: New International Version; Zondervan: Grand Rapids, MI, USA, 1984.
  23. Chandrasekar, S.; Bhanumathy, M.; Pawar, A.; Somasundaram, T. Phytopharmacology of Ficus religiosa. Pharmacogn. Rev. 2010, 4, 195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Kmail, A.; Rahman, R.; Nisar, S.; Jilani, M.I. Banyan tree-the sacred medicinal tree with potential health and pharmacological benefits. Int. J. Chem. Biochem. Sci. 2018, 13, 52–57. [Google Scholar]
  25. Shah, N. Herbal folk medicines in Northern India. J. Ethnopharmacol. 1982, 6, 293–301. [Google Scholar] [CrossRef]
  26. Gregory, M.; Divya, B.; Mary, R.A.; Viji, M.H.; Kalaichelvan, V.; Palanivel, V. Anti–ulcer activity of Ficus religiosa leaf ethanolic extract. Asian Pac. J. Trop. Biomed. 2013, 3, 554–556. [Google Scholar] [CrossRef]
  27. Kulshreshtha, M.; Goswami, M.; Rao, C.; Ashwlayan, V.; Yadav, S. Estimation of antioxidant potential of aqueous extract of Ficus bengalensis leaf on gastric ulcer. Int. J. Pharm. Sci. Rev. Res. 2011, 9, 122–126. [Google Scholar]
  28. Haneef, J.; Thankayyan, R.S.K.; Sithul, H.; Sreeharshan, S. Bax translocation mediated mitochondrial apoptosis and caspase dependent photosensitizing effect of Ficus religiosa on cancer cells. PLoS ONE 2012, 7, e40055. [Google Scholar] [CrossRef]
  29. Saloni, J.; Sakthivel, J. Evaluation of antioxidant and anticancer potential of flavonoids from aerial roots of Ficus benghalensis Linn. Int. J. Pharm. Res. 2019, 11, 11–20. [Google Scholar]
  30. Pandit, R.; Phadke, A.; Jagtap, A. Antidiabetic effect of Ficus religiosa extract in streptozotocin-induced diabetic rats. J. Ethnopharmacol. 2010, 128, 462–466. [Google Scholar] [CrossRef]
  31. Gulecha, V.; Sivakumar, T.; Upaganlawar, A.; Mahajan, M.; Upasani, C. Screening of Ficus religiosa leaves fractions for analgesic and anti-inflammatory activities. Indian J. Pharmacol. 2011, 43, 662. [Google Scholar]
  32. Patil, V.; Pimprikar, R.; Patil, V. Pharmacognostical studies and evaluation of anti-inflammatory activity of Ficus bengalensis Linn. J. Young Pharm. 2009, 1, 49. [Google Scholar] [CrossRef] [Green Version]
  33. Daniel, R.S.; Devi, K.; Augusti, K.; Nair, C. Mechanism of action of antiatherogenic and related effects of Ficus bengalensis Linn. flavonoids in experimental animals. Indian J. Exp. Biol. 2003, 41, 296–303. [Google Scholar] [PubMed]
  34. Rathi, P.; Nath, R.; Pant, K.; Dixit, R.; Pal, R.; Kumar, R. Evaluation of Hypolipidemic and TNF-Î ± Lowering Effect of Ficus religiosa in Dyslipidemic Wistar Rats. Curr. Res. Diabetes Obes. J. 2019, 10, 104–112. [Google Scholar]
  35. Aswar, M.; Aswar, U.; Watkar, B.; Vyas, M.; Wagh, A.; Gujar, K.N. Anthelmintic activity of Ficus benghalensis. Int. J. Green Pharm. (IJGP) 2008, 2, 170–172. [Google Scholar] [CrossRef]
  36. Hyo, W.; Hye, Y.; Chau, V.; Young, H.; Young, K. Methnol extract of Ficus leaf inhibits the production of nitric oxide and Proinflammatory cytokines in LPS stimulated microglia via the MAPK pathway. Phytother. Res. 2008, 22, 1064–1069. [Google Scholar]
  37. Gabhe, S.; Tatke, P.; Khan, T. Evaluation of the immunomodulatory activity of the methanol extract of Ficus benghalensis roots in rats. Indian J. Pharmacol. 2006, 38, 271. [Google Scholar] [CrossRef] [Green Version]
  38. Mallurwar, V.; Pathak, A. Studies on immunomodulatory activity of Ficus religiosa. Indian J. Pharm. Educ. Res. 2008, 42, 341–343. [Google Scholar]
  39. Mukherjee, P.K.; Saha, K.; Murugesan, T.; Mandal, S.; Pal, M.; Saha, B. Screening of anti-diarrhoeal profile of some plant extracts of a specific region of West Bengal, India. J. Ethnopharmacol. 1998, 60, 85–89. [Google Scholar] [CrossRef]
  40. Taur, D.; Nirmal, S.; Patil, R.; Kharya, M. Antistress and antiallergic effects of Ficus bengalensis bark in asthma. Nat. Prod. Res. 2007, 21, 1266–1270. [Google Scholar] [CrossRef]
  41. Kapoor, M.; Jasani, N.; Acharya, N.; Acharya, S.; Kumar, V. Phytopharmacological evaluation and anti–asthmatic activity of Ficus religiosa leaves. Asian Pac. J. Trop. Med. 2011, 4, 642–644. [Google Scholar] [CrossRef] [Green Version]
  42. Singh, D.; Goel, R.K. Anticonvulsant effect of Ficus religiosa: Role of serotonergic pathways. J. Ethnopharmacol. 2009, 123, 330–334. [Google Scholar] [CrossRef] [PubMed]
  43. Ghimire, K.; Bastakoti, R.R. Ethnomedicinal knowledge and healthcare practices among the Tharus of Nawalparasi district in central Nepal. For. Ecol. Manag. 2009, 257, 2066–2072. [Google Scholar] [CrossRef]
  44. Al-Snafi, A.E. Pharmacology of Ficus religiosa—A review. IOSR J. Pharm. 2017, 7, 49–60. [Google Scholar] [CrossRef]
  45. Rajiv, P.; Sivaraj, R. Screening for phytochemicals and antimicrobial activity of aqueous extract of Ficus religiosa Linn. Int. J. Pharm. Pharm. Sci. 2012, 4, 207–209. [Google Scholar]
  46. Manorenjitha, M.; Norita, A.; Adillah, A.; Asmawi, M. Chemical profile of Ficus religiosa (Linn.) stem. Int. J. Life Sci. Med. Res. 2014, 4, 32. [Google Scholar]
  47. Aqil, F.; Ahmad, I. Broad-spectrum antibacterial and antifungal properties of certain traditionally used Indian medicinal plants. World J. Microbiol. Biotechnol. 2003, 19, 653–657. [Google Scholar] [CrossRef]
  48. Poudel, A.; Satyal, P.; Setzer, W.N. Composition and bioactivities of the leaf Essential oil of Ficus religiosa Linn. Am. J. Essent. Oils Nat. Prod. 2015, 2, 16–17. [Google Scholar]
  49. Naira, N.; Rohini, R.; Syed, M.; Amit, K. Wound healing activity of the hydro alcoholic extract of Ficus religiosa leaves in rats. Internet J. Altern. Med. 2009, 6, 2–7. [Google Scholar]
  50. Grison-Pigé, L.; Hossaert-McKey, M.; Greeff, J.M. Fig volatile compounds—A first comparative study. Phytochemistry 2002, 61, 61–71. [Google Scholar] [CrossRef]
  51. Makhija, I.K.; Sharma, I.P.; Khamar, D. Phytochemistry and Pharmacological properties of Ficus religiosa: An overview. Ann. Biol. Res. 2010, 1, 171–180. [Google Scholar]
  52. Murugesu, S.; Selamat, J.; Perumal, V. Phytochemistry, pharmacological properties, and recent applications of Ficus benghalensis and Ficus religiosa. Plants 2021, 10, 2749. [Google Scholar] [CrossRef] [PubMed]
  53. Rathee, D.; Rathee, S.; Rathee, P.; Deep, A.; Anandjiwala, S.; Rathee, D. HPTLC densitometric quantification of stigmasterol and lupeol from Ficus religiosa. Arab. J. Chem. 2015, 8, 366–371. [Google Scholar] [CrossRef] [Green Version]
  54. Warrier, P.K. Indian Medicinal Plants: A Compendium of 500 Species; Orient Blackswan: Andhra Pradesh, India, 1993; Volume 5. [Google Scholar]
  55. Ali, M.; Qadry, J. Amino-acid-composition of fruits and seeds of medicinal-plants. J. Indian Chem. Soc. 1987, 64, 230–231. [Google Scholar]
  56. Patel, R.; Gautam, P. Medicinal potency of Ficus benghalensis: A review. Int. J. Med. Chem. Anal. 2014, 4, 53–58. [Google Scholar]
  57. Shukla, R.; Gupta, S.; Gambhir, J.; Prabhu, K.; Murthy, P. Antioxidant effect of aqueous extract of the bark of Ficus bengalensis in hypercholesterolaemic rabbits. J. Ethnopharmacol. 2004, 92, 47–51. [Google Scholar] [CrossRef]
  58. Joseph, B.; Raj, S.J. Phytopharmacological and phytochemical properties of three Ficus species—An overview. Int. J. Pharma Bio Sci. 2010, 1, 246–253. [Google Scholar]
  59. Gill, N.; Rashmi, A.; Anmol, K.; Manpreet, K. Free radical scavenging activity and phytochemical investigation of Ficus benjamina fruit. Int. J. Univ. Pharm. Bio. Sci. 2016, 5, 14–26. [Google Scholar]
  60. Govindan, V.; Francis, G. Qualitative and quantitative determination of secondary metabolites and antioxidant potential of Ficus benghalensis linn seed. Int. J. Pharm. Pharm. Sci. 2015, 7, 118–124. [Google Scholar]
  61. Bhaskara Rao, K.; Ojha, V.; Preeti; Kumar, G.; Karthik, L. Phytochemical composition and antioxidant activity of Ficus benghalensis (Moraceae) leaf extract. J. Biol. Act. Prod. Nat. 2014, 4, 236–248. [Google Scholar]
  62. Malik, H.; Javaid, S.; Fawad Rasool, M.; Samad, N.; Rizwan Ahamad, S.; Alqahtani, F.; Imran, I. Amelioration of scopolamine-induced amnesic, anxiolytic and antidepressant effects of Ficus benghalensis in behavioral experimental models. Medicina 2020, 56, 144. [Google Scholar] [CrossRef] [Green Version]
  63. Naquvi, K.J.; Ali, M.; Ahamad, J. Two new phytosterols from the stem bark of Ficus bengalensis L. J. Saudi Chem. Soc. 2015, 19, 650–654. [Google Scholar] [CrossRef] [Green Version]
  64. Roskams, A.J.; Connor, J.R. Aluminum access to the brain: A role for transferrin and its receptor. Proc. Natl. Acad. Sci. USA 1990, 87, 9024–9027. [Google Scholar] [CrossRef] [PubMed]
  65. Oğuz, E.O.; Enli, Y.; Şahin, B.; Gönen, C.; Turgut, G. Aluminium sulphate exposure increases oxidative stress and suppresses brain development in Ross broiler chicks. Med. Sci. Monit. Int. Med. J. Exp. Clin. Res. 2012, 18, BR103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Cherubini, E.; Miles, R.M. The CA3 region of the hippocampus: How is it? What is it for? How does it do it? Front. Cell. Neurosci. 2015, 9, 19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Massand, A.; Rai, R.; Rai, A.R.; Joy, T.; Murlimanju, B.; Marathe, A. Effect of Methanolic Leaf Extract of Ficus religiosa on Neuronal Degeneration: A Pilot Study in Male Albino Wistar Rats. Indian J. Public Health Res. Dev. 2020, 11, 685. [Google Scholar] [CrossRef]
  68. Bhangale, J.O.; Acharya, N.S.; Acharya, S.R. Protective effect of Ficus religiosa (L.) against 3-nitropropionic acid induced Huntington disease. Orient. Pharm. Exp. Med. 2016, 16, 165–174. [Google Scholar] [CrossRef]
  69. Schulz, J.; Matthews, R.; Henshaw, D.; Beal, M. Neuroprotective strategies for treatment of lesions produced by mitochondrial toxins: Implications for neurodegenerative diseases. Neuroscience 1996, 71, 1043–1048. [Google Scholar] [CrossRef]
  70. Devi, W.; Sengottuvelu, S.; Haja, S.; Lalitha, V.; Sivakumar, T. Memory enhancing activities of Ficus religiosa leaves in rodents. Int. J. Res. Ayurveda Pharm. (IJRAP) 2011, 2, 834–838. [Google Scholar]
  71. Ghafoor, A.; Tahir, M.; Lone, K.P.; Faisal, B.; Latif, W. The effect of Ficus carica L.(Anjir) leaf extract on gentamicin induced nephrotoxicity in adult male albino mice. J. Ayub. Med. Coll. Abbottabad 2015, 27, 398–401. [Google Scholar]
  72. Vyawahare, N.; Khandelwal, A.; Batra, V.; Nikam, A. Herbal anticonvulsants. J. Herb. Med. Toxicol. 2007, 1, 9–14. [Google Scholar]
  73. Singh, D.; Singh, B.; Goel, R.K. Hydroethanolic leaf extract of Ficus religiosa lacks anticonvulsant activity in acute electro and chemo convulsion mice models. J. Pharm. Negat. Results 2011, 2, 58. [Google Scholar]
  74. Singh, D.; Singh, B.; Goel, R.K. Role of saponins for the anticonvulsant effect of adventitious roots of Ficus religiosa. Pharm. Biol. 2012, 50, 816–822. [Google Scholar] [CrossRef] [PubMed]
  75. Singh, D.; Mishra, A.; Goel, R.K. Effect of saponin fraction from Ficus religiosa on memory deficit, and behavioral and biochemical impairments in pentylenetetrazol kindled mice. Epilepsy. Behav. 2013, 27, 206–211. [Google Scholar] [CrossRef] [PubMed]
  76. Patil, M.S.; Patil, C.; Patil, S.; Jadhav, R. Anticonvulsant activity of aqueous root extract of Ficus religiosa. J. Ethnopharmacol. 2011, 133, 92–96. [Google Scholar] [CrossRef]
  77. Ahuja, D.; Bijjem, K.R.V.; Kalia, A.N. Bronchospasm potentiating effect of methanolic extract of Ficus religiosa fruits in guinea pigs. J. Ethnopharmacol. 2011, 133, 324–328. [Google Scholar] [CrossRef]
  78. Bagdy, G.; Kecskemeti, V.; Riba, P.; Jakus, R. Serotonin and epilepsy. J. Neurochem. 2007, 100, 857–873. [Google Scholar] [CrossRef]
  79. Statnick, M.A.; Dailey, J.W.; Jobe, P.C.; Browning, R.A. Abnormalities in brain serotonin concentration, high-affinity uptake, and tryptophan hydroxylase activity in severe-seizure genetically epilepsy-prone rats. Epilepsia 1996, 37, 311–321. [Google Scholar] [CrossRef]
  80. Kaur, H.; Singh, D.; Singh, B.; Goel, R.K. Anti-amnesic effect of Ficus religiosa in scopolamine-induced anterograde and retrograde amnesia. Pharm. Biol. 2010, 48, 234–240. [Google Scholar] [CrossRef]
  81. Vinutha, B.; Prashanth, D.; Salma, K.; Sreeja, S.; Pratiti, D.; Padmaja, R.; Radhika, S.; Amit, A.; Venkateshwarlu, K.; Deepak, M. Screening of selected Indian medicinal plants for acetylcholinesterase inhibitory activity. J. Ethnopharmacol. 2007, 109, 359–363. [Google Scholar] [CrossRef]
  82. Stalin, C.; Gunasekaran, V.; Jayabalan, G. Evaluation of Neuroprotective Effect of Ficus benghalensis against Alloxan Induced Diabetic Neuropathy in Rats. Int. J. Pharmacol. Phytochem. Ethnomed. 2016, 4, 52–60. [Google Scholar] [CrossRef] [Green Version]
  83. Chandra, P.; Sachan, N.; Chaudhary, A.; Yadav, M.; Kishore, K.; Ghosh, A.K. Acute & Sub Chronic Toxicity Studies and Pharmacological Evaluation of Ficus bengalensis L. (Family: Moraceae) on Scopolamine-Induced Memory Impairmentin Experimental Animals. Indian J. Drugs 2013, 1, 6–16. [Google Scholar]
  84. Panday, D.R.; Rauniar, G. Effect of root-extracts of Ficus benghalensis (Banyan) in memory, anxiety, muscle co-ordination and seizure in animal models. BMC Complement. Altern. Med. 2016, 16, 429. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Mostafa, N.M. β-Amyrin rich Bombax ceiba leaf extract with potential neuroprotective activity against scopolamine-induced memory impairment in rats. Rec. Nat. Prod. 2018, 12, 480. [Google Scholar] [CrossRef]
  86. Park, S.J.; Ahn, Y.J.; Oh, S.R.; Lee, Y.; Kwon, G.; Woo, H.; Lee, H.E.; Jang, D.S.; Jung, J.W.; Ryu, J.H. Amyrin attenuates scopolamine-induced cognitive impairment in mice. Biol. Pharm. Bull. 2014, 37, 1207–1213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Yu, H.-J.; Koh, S.-H. The role of PI3K/AKT pathway and its therapeutic possibility in Alzheimer’s disease. Hanyang Med. Rev. 2017, 37, 18–24. [Google Scholar] [CrossRef] [Green Version]
  88. Zhu, X.; Castellani, R.J.; Takeda, A.; Nunomura, A.; Atwood, C.S.; Perry, G.; Smith, M.A. Differential activation of neuronal ERK, JNK/SAPK and p38 in Alzheimer disease: The ‘two hit’ hypothesis. Mech. Ageing Dev. 2001, 123, 39–46. [Google Scholar] [CrossRef]
  89. Giovannini, M.G. The role of the extracellular signal-regulated kinase pathway in memory encoding. Rev. Neurosci. 2006, 17, 619–634. [Google Scholar] [CrossRef]
  90. Peineau, S.; Taghibiglou, C.; Bradley, C.; Wong, T.P.; Liu, L.; Lu, J.; Lo, E.; Wu, D.; Saule, E.; Bouschet, T. LTP inhibits LTD in the hippocampus via regulation of GSK3β. Neuron 2007, 53, 703–717. [Google Scholar] [CrossRef] [Green Version]
  91. Aragão, G.F.; Carneiro, L.M.V.; Rota-Junior, A.P.; Bandeira, P.N.; Lemos, T.L.G.d.; Viana, G.S.d.B. Alterations in brain amino acid metabolism and inhibitory effects on PKC are possibly correlated with anticonvulsant effects of the isomeric mixture of α-and β-amyrin from Protium heptaphyllum. Pharm. Biol. 2015, 53, 407–413. [Google Scholar] [CrossRef]
  92. Wu, G.; Yu, J.; Wang, L.; Ren, S.; Zhang, Y. PKC/CREB pathway mediates the expressions of GABAA receptor subunits in cultured hippocampal neurons after low-Mg2+ solution treatment. Epilepsy Res. 2018, 140, 155–161. [Google Scholar] [CrossRef]
  93. Kun, X.; Zuhua, G. Amyrin exerts potent anxiolytic and antidepressant effects via mechanisms involving monoamine oxidase and γ-aminobutyric acid in mouse hippocampus. Trop. J. Pharm. Res. 2019, 18, 1673–1681. [Google Scholar]
  94. Riederer, P.; Laux, G. MAO-inhibitors in Parkinson’s Disease. Exp. Neurobiol. 2011, 20, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Sharmaa, N.; Khuranaa, N.; Muthuramanb, A.; Utrejac, P. Azelaic acid attenuatesrotenone-induced behavioural alterations in parkinson’s disease rat model. Plant Arch. 2021, 21, 2333–2337. [Google Scholar] [CrossRef]
  96. Castor, K.; Shenoi, S.; Edminster, S.; Tran, T.; King, K.; Chui, H.; Pogoda, J.; Fonteh, A.; Harrington, M. Urine dicarboxylic acids change in pre-symptomatic Alzheimer’s disease and reflect loss of energy capacity and hippocampal volume. PLoS ONE 2020, 15, e0231765. [Google Scholar] [CrossRef] [PubMed]
  97. Budzynska, B.; Skalicka-Wozniak, K.; Kruk-Slomka, M.; Wydrzynska-Kuzma, M.; Biala, G. In Vivo modulation of the behavioral effects of nicotine by the coumarins xanthotoxin, bergapten, and umbelliferone. Psychopharmacology 2016, 233, 2289–2300. [Google Scholar] [CrossRef] [Green Version]
  98. Liang, Y.; Xie, L.; Liu, K.; Cao, Y.; Dai, X.; Wang, X.; Lu, J.; Zhang, X.; Li, X. Bergapten: A review of its pharmacology, pharmacokinetics, and toxicity. Phytother. Res. 2021, 35, 6131–6147. [Google Scholar] [CrossRef]
  99. Dincel, D.; Hatipoğlu, S.D.; Gören, A.C.; Topcu, G. Anticholinesterase furocoumarins from Heracleum platytaenium, a species endemic to the Ida Mountains. Turk. J. Chem. 2013, 37, 675–683. [Google Scholar]
  100. Orhan, I.; Tosun, F.; Şener, B. Coumarin, anthroquinone and stilbene derivatives with anticholinesterase activity. Z. Nat. C 2008, 63, 366–370. [Google Scholar] [CrossRef]
  101. Senol, F.S.; Woźniak, K.S.; Khan, M.T.H.; Orhan, I.E.; Sener, B.; Głowniak, K. An in vitro and in silico approach to cholinesterase inhibitory and antioxidant effects of the methanol extract, furanocoumarin fraction, and major coumarins of Angelica officinalis L. fruits. Phytochem. Lett. 2011, 4, 462–467. [Google Scholar] [CrossRef]
  102. Kowalczyk, J.; Kurach, Ł.; Boguszewska-Czubara, A.; Skalicka-Woźniak, K.; Kruk-Słomka, M.; Kurzepa, J.; Wydrzynska-Kuźma, M.; Biała, G.; Skiba, A.; Budzyńska, B. Bergapten improves scopolamine-induced memory impairment in mice via cholinergic and antioxidative mechanisms. Front. Neurosci. 2020, 14, 730. [Google Scholar] [CrossRef]
  103. Huong, D.T.L.; Choi, H.C.; Rho, T.C.; Lee, H.S.; Lee, M.K.; Kim, Y.H. Inhibitory activity of monoamine oxidase by coumarins from Peucedanum japonicum. Arch. Pharmacal Res. 1999, 22, 324–326. [Google Scholar] [CrossRef] [PubMed]
  104. Khan, A.-U.; Ijaz, M.U.; Shah, F.A.; Khan, A.W.; Li, S. Neuroprotective Effects of Berbamine, Bergepten, and Carveol on Paclitaxel-Induced Peripheral Neuropathy. Res. Sq. 2022, 1–23. [Google Scholar] [CrossRef]
  105. Yang, Y.-F.; Xu, W.; Song, W.; Ye, M.; Yang, X.-W. Transport of twelve coumarins from Angelicae Pubescentis Radix across a MDCK-pHaMDR cell monolayer—An in vitro model for blood-brain barrier permeability. Molecules 2015, 20, 11719–11732. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Yang, Y.-F.; Zhang, L.; Yang, X.-W. Distribution assessments of coumarins from Angelicae Pubescentis Radix in rat cerebrospinal fluid and brain by Liquid Chromatography Tandem Mass Spectrometry analysis. Molecules 2018, 23, 225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Johnson, G.L.; Lapadat, R. Mitogen-activated protein kinase pathways mediated by ERK, JNK, and p38 protein kinases. Science 2002, 298, 1911–1912. [Google Scholar] [CrossRef] [Green Version]
  108. Cowley, S.; Paterson, H.; Kemp, P.; Marshall, C.J. Activation of MAP kinase kinase is necessary and sufficient for PC12 differentiation and for transformation of NIH 3T3 cells. Cell 1994, 77, 841–852. [Google Scholar] [CrossRef]
  109. Obara, Y.; Aoki, T.; Kusano, M.; Ohizumi, Y. β-Eudesmol induces neurite outgrowth in rat pheochromocytoma cells accompanied by an activation of mitogen-activated protein kinase. J. Pharmacol. Exp. Ther. 2002, 301, 803–811. [Google Scholar] [CrossRef] [Green Version]
  110. Kim, K.Y. Anti-inflammatory and ECM gene expression modulations of β-eudesmol via NF-κB signaling pathway in normal human dermal fibroblasts. Biomed. Dermatol. 2018, 2, 3. [Google Scholar] [CrossRef]
  111. Barot, J.; Saxena, B. Therapeutic effects of eugenol in a rat model of traumatic brain injury: A behavioral, biochemical, and histological study. J. Tradit. Complement. Med. 2021, 11, 318–327. [Google Scholar] [CrossRef]
  112. Sun, X.; Wang, D.; Zhang, T.; Lu, X.; Duan, F.; Ju, L.; Zhuang, X.; Jiang, X. Eugenol attenuates cerebral ischemia-reperfusion injury by enhancing autophagy via AMPK-mTOR-P70S6K pathway. Front. Pharmacol. 2020, 11, 84. [Google Scholar] [CrossRef] [Green Version]
  113. Kim, J.; Kundu, M.; Viollet, B.; Guan, K.-L. AMPK and mTOR regulate autophagy through direct phosphorylation of Ulk1. Nat. Cell Biol. 2011, 13, 132–141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Sun, J.; Mu, Y.; Jiang, Y.; Song, R.; Yi, J.; Zhou, J.; Sun, J.; Jiao, X.; Prinz, R.A.; Li, Y. Inhibition of p70 S6 kinase activity by A77,1726 induces autophagy and enhances the degradation of superoxide dismutase 1 (SOD1) protein aggregates. Cell Death Dis. 2018, 9, 407. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Prasad, S.N. Neuroprotective efficacy of eugenol and isoeugenol in acrylamide-induced neuropathy in rats: Behavioral and biochemical evidence. Neurochem. Res. 2013, 38, 330–345. [Google Scholar] [CrossRef] [PubMed]
  116. Irie, Y.; Itokazu, N.; Anjiki, N.; Ishige, A.; Watanabe, K.; Keung, W.M. Eugenol exhibits antidepressant-like activity in mice and induces expression of metallothionein-III in the hippocampus. Brain Res. 2004, 1011, 243–246. [Google Scholar] [CrossRef] [PubMed]
  117. Said, M.M.; Abd Rabo, M.M. Neuroprotective effects of eugenol against aluminiuminduced toxicity in the rat brain. Arh. Z. Hig. Rada I Toksikol. 2017, 68, 27–36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Nagahara, A.H.; Merrill, D.A.; Coppola, G.; Tsukada, S.; Schroeder, B.E.; Shaked, G.M.; Wang, L.; Blesch, A.; Kim, A.; Conner, J.M. Neuroprotective effects of brain-derived neurotrophic factor in rodent and primate models of Alzheimer’s disease. Nat. Med. 2009, 15, 331–337. [Google Scholar] [CrossRef] [Green Version]
  119. Pandini, G.; Satriano, C.; Pietropaolo, A.; Gianì, F.; Travaglia, A.; La Mendola, D.; Nicoletti, V.G.; Rizzarelli, E. The inorganic side of NGF: Copper (II) and Zinc (II) affect the NGF mimicking signaling of the N-terminus peptides encompassing the recognition domain of TrkA receptor. Front. Neurosci. 2016, 10, 569. [Google Scholar] [CrossRef] [Green Version]
  120. Moreira Vasconcelos, C.F.; da Cunha Ferreira, N.M.; Hardy Lima Pontes, N.; de Sousa dos Reis, T.D.; Basto Souza, R.; Aragao Catunda Junior, F.E.; Vasconcelos Aguiar, L.M.; Maranguape Silva da Cunha, R. Eugenol and its association with levodopa in 6-hydroxydopamine-induced hemiparkinsonian rats: Behavioural and neurochemical alterations. Basic Clin. Pharmacol. Toxicol. 2020, 127, 287–302. [Google Scholar] [CrossRef] [PubMed]
  121. Dubey, K.; Anand, B.G.; Shekhawat, D.S.; Kar, K. Eugenol prevents amyloid formation of proteins and inhibits amyloid-induced hemolysis. Sci. Rep. 2017, 7, 40744. [Google Scholar] [CrossRef]
  122. Nagy, I.; Jancsó, G.; Urbán, L. The role of the vanilloid (capsaicin) receptor (TRPV1) in physiology and pathology. Eur. J. Pharmacol. 2004, 500, 351–369. [Google Scholar] [CrossRef]
  123. Silva dos Santos, J.; Goncalves Cirino, J.P.; de Oliveira Carvalho, P.; Ortega, M.M. The pharmacological action of kaempferol in central nervous system diseases: A review. Front. Pharmacol. 2021, 11, 565700. [Google Scholar] [CrossRef] [PubMed]
  124. Beg, T.; Jyoti, S.; Naz, F.; Ali, F.; Ali, S.K.; Reyad, A.M.; Siddique, Y.H. Protective effect of kaempferol on the transgenic Drosophila model of Alzheimer’s disease. CNS Neurol. Disord.-Drug Targets (Former. Curr. Drug Targets-CNS Neurol. Disord.) 2018, 17, 421–429. [Google Scholar] [CrossRef] [PubMed]
  125. Zarei, M.; Mohammadi, S.; Jabbari, S.; Shahidi, S. Intracerebroventricular microinjection of kaempferol on memory retention of passive avoidance learning in rats: Involvement of cholinergic mechanism (s). Int. J. Neurosci. 2019, 129, 1203–1212. [Google Scholar] [CrossRef] [PubMed]
  126. Hanaki, M.; Murakami, K.; Akagi, K.-I.; Irie, K. Structural insights into mechanisms for inhibiting amyloid β42 aggregation by non-catechol-type flavonoids. Bioorganic Med. Chem. 2016, 24, 304–313. [Google Scholar] [CrossRef]
  127. Pan, X.; Liu, X.; Zhao, H.; Wu, B.; Liu, G. Antioxidant, anti-inflammatory and neuroprotective effect of kaempferol on rotenone-induced Parkinson’s disease model of rats and SH-S5Y5 cells by preventing loss of tyrosine hydroxylase. J. Funct. Foods 2020, 74, 104140. [Google Scholar] [CrossRef]
  128. Yang, Y.-L.; Cheng, X.; Li, W.-H.; Liu, M.; Wang, Y.-H.; Du, G.-H. Kaempferol attenuates LPS-induced striatum injury in mice involving anti-neuroinflammation, maintaining BBB integrity, and down-regulating the HMGB1/TLR4 pathway. Int. J. Mol. Sci. 2019, 20, 491. [Google Scholar] [CrossRef] [Green Version]
  129. Filomeni, G.; Graziani, I.; De Zio, D.; Dini, L.; Centonze, D.; Rotilio, G.; Ciriolo, M.R. Neuroprotection of kaempferol by autophagy in models of rotenone-mediated acute toxicity: Possible implications for Parkinson’s disease. Neurobiol. Aging 2012, 33, 767–785. [Google Scholar] [CrossRef]
  130. Li, S.; Pu, X.-P. Neuroprotective effect of kaempferol against a 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-induced mouse model of Parkinson’s disease. Biol. Pharm. Bull. 2011, 34, 1291–1296. [Google Scholar] [CrossRef] [Green Version]
  131. Yang, Y.; Bai, L.; Li, X.; Xiong, J.; Xu, P.; Guo, C.; Xue, M. Transport of active flavonoids, based on cytotoxicity and lipophilicity: An evaluation using the blood–brain barrier cell and Caco-2 cell models. Toxicol. Vitr. 2014, 28, 388–396. [Google Scholar] [CrossRef]
  132. Hu, L.-D.; Wang, J.; Chen, X.-J.; Yan, Y.-B. Lanosterol modulates proteostasis via dissolving cytosolic sequestosomes/aggresome-like induced structures. Biochim. Biophys. Acta (BBA)-Mol. Cell Res. 2020, 1867, 118617. [Google Scholar] [CrossRef]
  133. Zhou, H.; Yang, Z.; Tian, X.; Chen, L.; Lee, S.; Huynh, T.; Ge, C.; Zhou, R. Lanosterol disrupts the aggregation of amyloid-β peptides. ACS Chem. Neurosci. 2019, 10, 4051–4060. [Google Scholar] [CrossRef] [PubMed]
  134. Upadhyay, A.; Amanullah, A.; Mishra, R.; Kumar, A.; Mishra, A. Lanosterol suppresses the aggregation and cytotoxicity of misfolded proteins linked with neurodegenerative diseases. Mol. Neurobiol. 2018, 55, 1169–1182. [Google Scholar] [CrossRef] [PubMed]
  135. Lim, L.; Jackson-Lewis, V.; Wong, L.; Shui, G.; Goh, A.; Kesavapany, S.; Jenner, A.; Fivaz, M.; Przedborski, S.; Wenk, M. Lanosterol induces mitochondrial uncoupling and protects dopaminergic neurons from cell death in a model for Parkinson’s disease. Cell Death Differ. 2012, 19, 416–427. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Ma, H.; Johnson, S.L.; Liu, W.; DaSilva, N.A.; Meschwitz, S.; Dain, J.A.; Seeram, N.P. Evaluation of polyphenol anthocyanin-enriched extracts of blackberry, black raspberry, blueberry, cranberry, red raspberry, and strawberry for free radical scavenging, reactive carbonyl species trapping, anti-glycation, anti-β-amyloid aggregation, and microglial neuroprotective effects. Int. J. Mol. Sci. 2018, 19, 461. [Google Scholar]
  137. Suresh, S.; Begum, R.F.; Singh, A.; Chitra, V. Anthocyanin as a Therapeutic in Alzheimer’s Disease: A Systematic Review of Preclinical Evidences. Ageing Res. Rev. 2022, 76, 101595. [Google Scholar] [CrossRef]
  138. Ullah, R.; Khan, M.; Shah, S.A.; Saeed, K.; Kim, M.O. Natural antioxidant anthocyanins—A hidden therapeutic candidate in metabolic disorders with major focus in neurodegeneration. Nutrients 2019, 11, 1195. [Google Scholar] [CrossRef] [Green Version]
  139. Winter, A.N.; Bickford, P.C. Anthocyanins and their metabolites as therapeutic agents for neurodegenerative disease. Antioxidants 2019, 8, 333. [Google Scholar] [CrossRef] [Green Version]
  140. Khan, M.S.; Ikram, M.; Park, T.J.; Kim, M.O. Pathology, risk factors, and oxidative damage related to type 2 diabetes-mediated Alzheimer’s disease and the rescuing effects of the potent antioxidant anthocyanin. Oxidative Med. Cell. Longev. 2021, 2021, 4051207. [Google Scholar] [CrossRef]
  141. Miguel, M.G. Anthocyanins: Antioxidant and/or anti-inflammatory activities. J. Appl. Pharm. Sci. 2011, 1, 7–15. [Google Scholar]
  142. El-Shiekh, R.A.; Ashour, R.M.; Abd El-Haleim, E.A.; Ahmed, K.A.; Abdel-Sattar, E. Hibiscus sabdariffa L.: A potent natural neuroprotective agent for the prevention of streptozotocin-induced Alzheimer’s disease in mice. Biomed. Pharmacother. 2020, 128, 110303. [Google Scholar] [CrossRef]
  143. Kerr, F.; Sofola-Adesakin, O.; Ivanov, D.K.; Gatliff, J.; Gomez Perez-Nievas, B.; Bertrand, H.C.; Martinez, P.; Callard, R.; Snoeren, I.; Cocheme, H.M. Direct Keap1-Nrf2 disruption as a potential therapeutic target for Alzheimer’s disease. PLoS Genet. 2017, 13, e1006593. [Google Scholar] [CrossRef] [PubMed]
  144. Reichard, J.F.; Motz, G.T.; Puga, A. Heme oxygenase-1 induction by NRF2 requires inactivation of the transcriptional repressor BACH1. Nucleic Acids Res. 2007, 35, 7074–7086. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Fang, Y.; Ou, S.; Wu, T.; Zhou, L.; Tang, H.; Jiang, M.; Xu, J.; Guo, K. Lycopene alleviates oxidative stress via the PI3K/Akt/Nrf2pathway in a cell model of Alzheimer’s disease. PeerJ 2020, 8, e9308. [Google Scholar] [CrossRef] [PubMed]
  146. Li, J.; Zhao, R.; Zhao, H.; Chen, G.; Jiang, Y.; Lyu, X.; Wu, T. Reduction of aging-induced oxidative stress and activation of autophagy by bilberry anthocyanin supplementation via the AMPK–mTOR signaling pathway in aged female rats. J. Agric. Food Chem. 2019, 67, 7832–7843. [Google Scholar] [CrossRef] [PubMed]
  147. Youdim, K.A.; Dobbie, M.S.; Kuhnle, G.; Proteggente, A.R.; Abbott, N.J.; Rice-Evans, C. Interaction between flavonoids and the blood–brain barrier: In Vitro studies. J. Neurochem. 2003, 85, 180–192. [Google Scholar] [CrossRef]
  148. Piccialli, I.; Tedeschi, V.; Caputo, L.; Amato, G.; De Martino, L.; De Feo, V.; Secondo, A.; Pannaccione, A. The Antioxidant Activity of Limonene Counteracts Neurotoxicity Triggered byAβ1-42 Oligomers in Primary Cortical Neurons. Antioxidants 2021, 10, 937. [Google Scholar] [CrossRef]
  149. Pannaccione, A.; Boscia, F.; Scorziello, A.; Adornetto, A.; Castaldo, P.; Sirabella, R.; Taglialatela, M.; Di Renzo, G.; Annunziato, L. Up-regulation and increased activity of KV3. 4 channels and their accessory subunit MinK-related peptide 2 induced by amyloid peptide are involved in apoptotic neuronal death. Mol. Pharmacol. 2007, 72, 665–673. [Google Scholar] [CrossRef] [Green Version]
  150. Zhou, W.; Yoshioka, M.; Yokogoshi, H. Sub-chronic effects of s-limonene on brain neurotransmitter levels and behavior of rats. J. Nutr. Sci. Vitaminol. 2009, 55, 367–373. [Google Scholar] [CrossRef] [Green Version]
  151. Boiangiu, R.S.; Brinza, I.; Hancianu, M.; Erdogan Orhan, I.; Eren, G.; Gündüz, E.; Ertas, H.; Hritcu, L.; Cioanca, O. Cognitive facilitation and antioxidant effects of an essential oil mix on scopolamine-induced amnesia in rats: Molecular modeling of in vitro and in vivo approaches. Molecules 2020, 25, 1519. [Google Scholar] [CrossRef] [Green Version]
  152. Abuhamdah, S.; Abuhamdah, R.; Howes, M.-J.R.; Al-Olimat, S.; Ennaceur, A.; Chazot, P.L. Pharmacological and neuroprotective profile of an essential oil derived from leaves of A loysia citrodora Palau. J. Pharm. Pharmacol. 2015, 67, 1306–1315. [Google Scholar] [CrossRef] [Green Version]
  153. Lomarat, P.; Sripha, K.; Phanthong, P.; Kitphati, W.; Thirapanmethee, K.; Bunyapraphatsara, N. In Vitro biological activities of black pepper essential oil and its major components relevant to the prevention of Alzheimer’s disease. Thai J. Pharm. Sci. (TJPS) 2015, 39, 94–101. [Google Scholar]
  154. Sadiki, F.Z.; El Idrissi, M.; Cioanca, O.; Trifan, A.; Hancianu, M.; Hritcu, L.; Postu, P.A. Tetraclinis articulata essential oil mitigates cognitive deficits and brain oxidative stress in an Alzheimer’s disease amyloidosis model. Phytomedicine 2019, 56, 57–63. [Google Scholar] [CrossRef] [PubMed]
  155. Cheng, B.-H.; Sheen, L.-Y.; Chang, S.-T. Evaluation of anxiolytic potency of essential oil and S-(+)-linalool from Cinnamomum osmophloeum ct. linalool leaves in mice. J. Tradit. Complement. Med. 2015, 5, 27–34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Caputo, L.; Piccialli, I.; Ciccone, R.; de Caprariis, P.; Massa, A.; De Feo, V.; Pannaccione, A. Lavender and coriander essential oils and their main component linalool exert a protective effect against amyloid-β neurotoxicity. Phytother. Res. 2021, 35, 486–493. [Google Scholar] [CrossRef]
  157. Yuan, C.; Shin, M.; Park, Y.; Choi, B.; Jang, S.; Lim, C.; Yun, H.S.; Lee, I.-S.; Won, S.-Y.; Cho, K.S. Linalool Alleviates Aβ42-Induced Neurodegeneration via Suppressing ROS Production and Inflammation in Fly and Rat Models of Alzheimer’s Disease. Oxidative Med. Cell. Longev. 2021, 2021, 8887716. [Google Scholar] [CrossRef]
  158. Sabogal-Guáqueta, A.M.; Posada-Duque, R.; Cortes, N.C.; Arias-Londoño, J.D.; Cardona-Gómez, G.P. Changes in the hippocampal and peripheral phospholipid profiles are associated with neurodegeneration hallmarks in a long-term global cerebral ischemia model: Attenuation by Linalool. Neuropharmacology 2018, 135, 555–571. [Google Scholar] [CrossRef]
  159. Sabogal-Guáqueta, A.M.; Osorio, E.; Cardona-Gómez, G.P. Linalool reverses neuropathological and behavioral impairments in old triple transgenic Alzheimer’s mice. Neuropharmacology 2016, 102, 111–120. [Google Scholar] [CrossRef] [Green Version]
  160. Silva Brum, L.; Emanuelli, T.; Souza, D.; Elisabetsky, E. Effects of linalool on glutamate release and uptake in mouse cortical synaptosomes. Neurochem. Res. 2001, 26, 191–194. [Google Scholar] [CrossRef]
  161. Sabogal-Guáqueta, A.M.; Hobbie, F.; Keerthi, A.; Oun, A.; Kortholt, A.; Boddeke, E.; Dolga, A. Linalool attenuates oxidative stress and mitochondrial dysfunction mediated by glutamate and NMDA toxicity. Biomed. Pharmacother. 2019, 118, 109295. [Google Scholar] [CrossRef]
  162. Li, Y.; Lv, O.; Zhou, F.; Li, Q.; Wu, Z.; Zheng, Y. Linalool inhibits LPS-induced inflammation in BV2 microglia cells by activating Nrf2. Neurochem. Res. 2015, 40, 1520–1525. [Google Scholar] [CrossRef]
  163. Xu, X.; Lv, H.; Xia, Z.; Fan, R.; Zhang, C.; Wang, Y.; Wang, D. Rhein exhibits antioxidative effects similar to Rhubarb in a rat model of traumatic brain injury. BMC Complement. Altern. Med. 2017, 17, 140. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Badshah, H.; Ali, T.; Rehman, S.-U.; Amin, F.-U.; Ullah, F.; Kim, T.H.; Kim, M.O. Protective effect of lupeol against lipopolysaccharide-induced neuroinflammation via the p38/c-Jun N-terminal kinase pathway in the adult mouse brain. J. Neuroimmune Pharmacol. 2016, 11, 48–60. [Google Scholar] [CrossRef] [PubMed]
  165. Koirala, P.; Seong, S.H.; Jung, H.A.; Choi, J.S. Comparative molecular docking studies of lupeol and lupenone isolated from Pueraria lobata that inhibits BACE1: Probable remedies for Alzheimer’s disease. Asian Pac. J. Trop. Med. 2017, 10, 1117–1122. [Google Scholar] [CrossRef] [PubMed]
  166. Kaundal, M.; Akhtar, M.; Deshmukh, R. Lupeol Isolated from Betula alnoides Ameliorates Amyloid Beta Induced Neuronal Damage via Targeting Various Pathological Events and Alteration in Neurotransmitter Levels in Rat’s Brain. J. Neurol. Neurosci. 2017, 8, 195. [Google Scholar] [CrossRef] [Green Version]
  167. Ahmad, R.; Khan, A.; Lee, H.J.; Ur Rehman, I.; Khan, I.; Alam, S.I.; Kim, M.O. Lupeol, a plant-derived triterpenoid, protects mice brains against Aβ-induced oxidative stress and neurodegeneration. Biomedicines 2020, 8, 380. [Google Scholar] [CrossRef] [PubMed]
  168. Ahmad, R.; Khan, A.; Rehman, I.U.; Lee, H.J.; Khan, I.; Kim, M.O. Lupeol Treatment Attenuates Activation of Glial Cells and Oxidative-Stress-Mediated Neuropathology in Mouse Model of Traumatic Brain Injury. Int. J. Mol. Sci. 2022, 23, 6086. [Google Scholar] [CrossRef]
  169. Oliveira-Junior, M.S.; Pereira, E.P.; de Amorim, V.C.M.; Reis, L.T.C.; do Nascimento, R.P.; da Silva, V.D.A.; Costa, S.L. Lupeol inhibits LPS-induced neuroinflammation in cerebellar cultures and induces neuroprotection associated to the modulation of astrocyte response and expression of neurotrophic and inflammatory factors. Int. Immunopharmacol. 2019, 70, 302–312. [Google Scholar] [CrossRef]
  170. Malik, A.; Jamil, U.; Butt, T.T.; Waquar, S.; Gan, S.H.; Shafique, H.; Jafar, T.H. In Silico and in vitro studies of lupeol and iso-orientin as potential antidiabetic agents in a rat model. Drug Des. Dev. Ther. 2019, 13, 1501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Villalba, H.; Shah, K.; Albekairi, T.H.; Sifat, A.E.; Vaidya, B.; Abbruscato, T.J. Potential role of myo-inositol to improve ischemic stroke outcome in diabetic mouse. Brain Res. 2018, 1699, 166–176. [Google Scholar] [CrossRef]
  172. Villalba, H.; Vaidya, B.; Cucullo, L.; Abbruscato, T.J. Myo-Inositol Improves Ischemic Stroke Outcome after both Nicotine Containing Electronic Cigarette and Tobacco Smoke Exposure. FASEB J. 2020, 34, 1. [Google Scholar] [CrossRef]
  173. Kandashvili, M.; Gamkrelidze, G.; Tsverava, L.; Lordkipanidze, T.; Lepsveridze, E.; Lagani, V.; Burjanadze, M.; Dashniani, M.; Kokaia, M.; Solomonia, R. Myo-Inositol Limits Kainic Acid-Induced Epileptogenesis in Rats. Int. J. Mol. Sci. 2022, 23, 1198. [Google Scholar] [CrossRef] [PubMed]
  174. Kalacheva, A.; Gromova, O.; Grishina, T.; Gogoleva, I.; Demidov, V.; Torshin, I. An experimental study of anticonvulsant effects of myo-inositol and folic acid. Zhurnal Nevrol. Psikhiatrii Im. SS Korsakova 2016, 116, 56–61. [Google Scholar] [CrossRef] [PubMed]
  175. Voevodskaya, O.; Sundgren, P.C.; Strandberg, O.; Zetterberg, H.; Minthon, L.; Blennow, K.; Wahlund, L.-O.; Westman, E.; Hansson, O.; Group, S.B.S. Myo-inositol changes precede amyloid pathology and relate to APOE genotype in Alzheimer disease. Neurology 2016, 86, 1754–1761. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Voevodskaya, O.; Poulakis, K.; Sundgren, P.; Van Westen, D.; Palmqvist, S.; Wahlund, L.-O.; Stomrud, E.; Hansson, O.; Westman, E.; Group, S.B.S. Brain myoinositol as a potential marker of amyloid-related pathology: A longitudinal study. Neurology 2019, 92, e395–e405. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Spector, R. Myo-inositol transport through the blood-brain barrier. Neurochem. Res. 1988, 13, 785–787. [Google Scholar] [CrossRef]
  178. Yao, Y.; Lin, G.; Xie, Y.; Ma, P.; Li, G.; Meng, Q.; Wu, T. Preformulation studies of myricetin: A natural antioxidant flavonoid. Die Pharm.-Int. J. Pharm. Sci. 2014, 69, 19–26. [Google Scholar]
  179. Liu, M.; Guo, H.; Li, Z.; Zhang, C.; Zhang, X.; Cui, Q.; Tian, J. Molecular level insight into the benefit of myricetin and dihydromyricetin uptake in patients with Alzheimer’s diseases. Front. Aging Neurosci. 2020, 12, 601603. [Google Scholar] [CrossRef]
  180. Kimura, A.M.; Tsuji, M.; Yasumoto, T.; Mori, Y.; Oguchi, T.; Tsuji, Y.; Umino, M.; Umino, A.; Nishikawa, T.; Nakamura, S. Myricetin prevents high molecular weight Aβ1-42 oligomer-induced neurotoxicity through antioxidant effects in cell membranes and mitochondria. Free Radic. Biol. Med. 2021, 171, 232–244. [Google Scholar] [CrossRef]
  181. Tamagno, E.; Parola, M.; Bardini, P.; Piccini, A.; Borghi, R.; Guglielmotto, M.; Santoro, G.; Davit, A.; Danni, O.; Smith, M. β-Site APP cleaving enzyme up-regulation induced by 4-hydroxynonenal is mediated by stress-activated protein kinases pathways. J. Neurochem. 2005, 92, 628–636. [Google Scholar] [CrossRef]
  182. Su, B.; Wang, X.; Lee, H.-g.; Tabaton, M.; Perry, G.; Smith, M.A.; Zhu, X. Chronic oxidative stress causes increased tau phosphorylation in M17 neuroblastoma cells. Neurosci. Lett. 2010, 468, 267–271. [Google Scholar] [CrossRef]
  183. Lasagna-Reeves, C.A.; Castillo-Carranza, D.L.; Sengupta, U.; Clos, A.L.; Jackson, G.R.; Kayed, R. Tau oligomers impair memory and induce synaptic and mitochondrial dysfunction in wild-type mice. Mol. Neurodegener. 2011, 6, 39. [Google Scholar] [CrossRef] [Green Version]
  184. Shimmyo, Y.; Kihara, T.; Akaike, A.; Niidome, T.; Sugimoto, H. Multifunction of myricetin on Aβ: Neuroprotection via a conformational change of Aβ and reduction of Aβ via the interference of secretases. J. Neurosci. Res. 2008, 86, 368–377. [Google Scholar] [CrossRef] [PubMed]
  185. Takahashi, R.; Ono, K.; Takamura, Y.; Mizuguchi, M.; Ikeda, T.; Nishijo, H.; Yamada, M. Phenolic compounds prevent the oligomerization of α-synuclein and reduce synaptic toxicity. J. Neurochem. 2015, 134, 943–955. [Google Scholar] [CrossRef]
  186. Jing, N.; Li, X. Dihydromyricetin attenuates inflammation through TLR4/NF-kappaB pathway. Open Med. 2019, 14, 719–725. [Google Scholar] [CrossRef] [PubMed]
  187. Cao, J.; Chen, H.; Lu, W.; Wu, Y.; Wu, X.; Xia, D.; Zhu, J. Myricetin induces protective autophagy by inhibiting the phosphorylation of mTOR in HepG2 cells. Anat. Rec. 2018, 301, 786–795. [Google Scholar] [CrossRef]
  188. DeToma, A.S.; Choi, J.S.; Braymer, J.J.; Lim, M.H. Myricetin: A naturally occurring regulator of metal-induced amyloid-β aggregation and neurotoxicity. ChemBioChem 2011, 12, 1198–1201. [Google Scholar] [CrossRef]
  189. Kou, X.; Liu, X.; Chen, X.; Li, J.; Yang, X.; Fan, J.; Yang, Y.; Chen, N. Ampelopsin attenuates brain aging of D-gal-induced rats through miR-34a-mediated SIRT1/mTOR signal pathway. Oncotarget 2016, 7, 74484. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Shadfar, S.; Hwang, C.J.; Lim, M.-S.; Choi, D.-Y.; Hong, J.T. Involvement of inflammation in Alzheimer’s disease pathogenesis and therapeutic potential of anti-inflammatory agents. Arch. Pharmacal Res. 2015, 38, 2106–2119. [Google Scholar] [CrossRef]
  191. Weston-Green, K.; Clunas, H.; Naranjo, C.J. A review of the potential use of pinene and linalool as terpene-based medicines for brain health: Discovering novel therapeutics in the flavours and fragrances of cannabis. Front. Psychiatry 2021, 12, 583211. [Google Scholar] [CrossRef]
  192. Porres-Martínez, M.; González-Burgos, E.; Carretero, M.E.; Gómez-Serranillos, M.P. In Vitro neuroprotective potential of the monoterpenes α-pinene and 1,8-cineole against H2O2-induced oxidative stress in PC12 cells. Z. Nat. C 2016, 71, 191–199. [Google Scholar] [CrossRef]
  193. Khan-Mohammadi-Khorrami, M.K.; Asle-Rousta, M.; Rahnema, M.; Amini, R. Neuroprotective effect of alpha-pinene is mediated by suppression of the TNF-α/NF-κB pathway in Alzheimer’s disease rat model. J. Biochem. Mol. Toxicol. 2022, 36, e23006. [Google Scholar] [CrossRef] [PubMed]
  194. Khoshnazar, M.; Bigdeli, M.R.; Parvardeh, S.; Pouriran, R. Attenuating effect of α-pinene on neurobehavioural deficit, oxidative damage and inflammatory response following focal ischaemic stroke in rat. J. Pharm. Pharmacol. 2019, 71, 1725–1733. [Google Scholar] [CrossRef] [PubMed]
  195. Khoshnazar, M.; Parvardeh, S.; Bigdeli, M.R. Alpha-pinene exerts neuroprotective effects via anti-inflammatory and anti-apoptotic mechanisms in a rat model of focal cerebral ischemia-reperfusion. J. Stroke Cerebrovasc. Dis. 2020, 29, 104977. [Google Scholar] [CrossRef] [PubMed]
  196. Lee, G.-Y.; Lee, C.; Park, G.H.; Jang, J.-H. Amelioration of Scopolamine-Induced Learning and Memory Impairment by?-Pinene in C57BL/6 Mice. Mult. Bioactivities Tradit. Med. Herbs Treat. Neurodegener. Dis. 2017, 2017, 4926815. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Somani, G.; Kulkarni, C.; Shinde, P.; Shelke, R.; Laddha, K.; Sathaye, S. In Vitro acetylcholinesterase inhibition by psoralen using molecular docking and enzymatic studies. J. Pharm. Bioallied Sci. 2015, 7, 32. [Google Scholar]
  198. Kong, L.D.; Tan, R.X.; Woo, A.Y.H.; Cheng, C.H.K. Inhibition of rat brain monoamine oxidase activities by psoralen and isopsoralen: Implications for the treatment of affective disorders. Pharmacol. Toxicol. 2001, 88, 75–80. [Google Scholar] [CrossRef]
  199. Behl, T.; Kaur, D.; Sehgal, A.; Singh, S.; Sharma, N.; Zengin, G.; Andronie-Cioara, F.L.; Toma, M.M.; Bungau, S.; Bumbu, A.G. Role of monoamine oxidase activity in Alzheimer’s disease: An insight into the therapeutic potential of inhibitors. Molecules 2021, 26, 3724. [Google Scholar] [CrossRef]
  200. Wu, C.-R.; Chang, C.-L.; Hsieh, P.-Y.; Lin, L.-W.; Ching, H. Psoralen and isopsoralen, two coumarins of Psoraleae Fructus, can alleviate scopolamine-induced amnesia in rats. Planta Med. 2007, 73, 275–278. [Google Scholar] [CrossRef] [Green Version]
  201. Shimmyo, Y.; Kihara, T.; Akaike, A.; Niidome, T.; Sugimoto, H. Flavonols and flavones as BACE-1 inhibitors: Structure–activity relationship in cell-free, cell-based and in silico studies reveal novel pharmacophore features. Biochim. Biophys. Acta (BBA)-Gen. Subj. 2008, 1780, 819–825. [Google Scholar] [CrossRef]
  202. 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. 2012, 2012, 823206. [Google Scholar] [CrossRef] [Green Version]
  203. Ansari, M.A.; Abdul, H.M.; Joshi, G.; Opii, W.O.; Butterfield, D.A. Protective effect of quercetin in primary neurons against Aβ (1–42): Relevance to Alzheimer’s disease. J. Nutr. Biochem. 2009, 20, 269–275. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Ay, M.; Luo, J.; Langley, M.; Jin, H.; Anantharam, V.; Kanthasamy, A.; Kanthasamy, A.G. 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] [Green Version]
  205. Sandhir, R.; Mehrotra, A. Quercetin supplementation is effective in improving mitochondrial dysfunctions induced by 3-nitropropionic acid: Implications in Huntington’s disease. Biochim. Biophys. Acta (BBA)-Mol. Basis Dis. 2013, 1832, 421–430. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Kuhad, A.; Singla, S.; Arora, V.; Chopra, K. Neuroprotective effect of sesamol and quercetin against QA induced neurotoxicity: An experimental paradigm of Huntington’s disease. J. Neurol. Sci. 2013, 333, e149–e150. [Google Scholar] [CrossRef]
  207. Grewal, A.K.; Singh, T.G.; Sharma, D.; Sharma, V.; Singh, M.; Rahman, M.H.; Najda, A.; Walasek-Janusz, M.; Kamel, M.; Albadrani, G.M. Mechanistic insights and perspectives involved in neuroprotective action of quercetin. Biomed. Pharmacother. 2021, 140, 111729. [Google Scholar] [CrossRef] [PubMed]
  208. Giordano, G.; Cole, T.B.; Furlong, C.E.; Costa, L.G. Paraoxonase 2 (PON2) in the mouse central nervous system: A neuroprotective role? Toxicol. Appl. Pharmacol. 2011, 256, 369–378. [Google Scholar] [CrossRef] [Green Version]
  209. Saw, C.L.L.; Guo, Y.; Yang, A.Y.; Paredes-Gonzalez, X.; Ramirez, C.; Pung, D.; Kong, A.-N.T. The berry constituents quercetin, kaempferol, and pterostilbene synergistically attenuate reactive oxygen species: Involvement of the Nrf2-ARE signaling pathway. Food Chem. Toxicol. 2014, 72, 303–311. [Google Scholar] [CrossRef]
  210. Chen, L.; Sun, L.; Liu, Z.; Wang, H.; Xu, C. Protection afforded by quercetin against H2O2-induced apoptosis on PC12 cells via activating PI3K/Akt signal pathway. J. Recept. Signal Transduct. 2016, 36, 98–102. [Google Scholar] [CrossRef]
  211. Park, J.-Y.; Lim, M.-S.; Kim, S.-I.; Lee, H.J.; Kim, S.-S.; Kwon, Y.-S.; Chun, W. Quercetin-3-O-β-D-glucuronide suppresses lipopolysaccharide-induced JNK and ERK phosphorylation in LPS-challenged RAW264. 7 cells. Biomol. Ther. 2016, 24, 610. [Google Scholar] [CrossRef] [Green Version]
  212. Lavoie, S.; Chen, Y.; Dalton, T.P.; Gysin, R.; Cuénod, M.; Steullet, P.; Do, K.Q. Curcumin, quercetin, and tBHQ modulate glutathione levels in astrocytes and neurons: Importance of the glutamate cysteine ligase modifier subunit. J. Neurochem. 2009, 108, 1410–1422. [Google Scholar] [CrossRef]
  213. de Boer, V.C.; de Goffau, M.C.; Arts, I.C.; Hollman, P.C.; Keijer, J. SIRT1 stimulation by polyphenols is affected by their stability and metabolism. Mech. Ageing Dev. 2006, 127, 618–627. [Google Scholar] [CrossRef] [PubMed]
  214. Spencer, J.P.; Rice-Evans, C.; Williams, R.J. Modulation of pro-survival Akt/protein kinase B and ERK1/2 signaling cascades by quercetin and its in vivo metabolites underlie their action on neuronal viability. J. Biol. Chem. 2003, 278, 34783–34793. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Sun, G.Y.; Rice-Evans, C.; Williams, R.J. Quercetin attenuates inflammatory responses in BV-2 microglial cells: Role of MAPKs on the Nrf2 pathway and induction of heme oxygenase-1. PLoS ONE 2015, 10, e0141509. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Sabogal-Guáqueta, A.M.; Munoz-Manco, J.I.; 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]
  217. Wang, Y.; Fan, X.; Tang, T.; Fan, R.; Zhang, C.; Huang, Z.; Peng, W.; Gan, P.; Xiong, X.; Huang, W. Rhein and rhubarb similarly protect the blood-brain barrier after experimental traumatic brain injury via gp91phox subunit of NADPH oxidase/ROS/ERK/MMP-9 signaling pathway. Sci. Rep. 2016, 6, 37098. [Google Scholar] [CrossRef] [Green Version]
  218. Zhao, Q.; Wang, X.; Chen, A.; Cheng, X.; Zhang, G.; Sun, J.; Zhao, Y.; Huang, Y.; Zhu, Y. Rhein protects against cerebral ischemic-/reperfusion-induced oxidative stress and apoptosis in rats. Int. J. Mol. Med. 2018, 41, 2802–2812. [Google Scholar] [CrossRef]
  219. Bi, F.; Ma, H.; Ji, C.; Chang, C.; Liu, W.; Xie, K. Rhein protects against neurological deficits after traumatic brain injury in mice via inhibiting neuronal pyroptosis. Front. Pharmacol. 2020, 11, 564367. [Google Scholar] [CrossRef]
  220. Yin, Z.; Gao, D.; Du, K.; Han, C.; Liu, Y.; Wang, Y.; Gao, X. Rhein Ameliorates Cognitive Impairment in an APP/PS1 Transgenic Mouse Model of Alzheimer’s Disease by Relieving Oxidative Stress through Activating the SIRT1/PGC-1α Pathway. Oxidative Med. Cell. Longev. 2022, 2022, 2524832. [Google Scholar] [CrossRef]
  221. Wang, Y.; Fan, R.; Luo, J.; Tang, T.; Xing, Z.; Xia, Z.; Peng, W.; Wang, W.; Lv, H.; Huang, W. An ultra high performance liquid chromatography with tandem mass spectrometry method for plasma and cerebrospinal fluid pharmacokinetics of rhein in patients with traumatic brain injury after administration of rhubarb decoction. J. Sep. Sci. 2015, 38, 1100–1108. [Google Scholar] [CrossRef]
  222. Budzynska, B.; Faggio, C.; Kruk-Slomka, M.; Samec, D.; Nabavi, S.F.; Sureda, A.; Devi, K.P.; Nabavi, S.M. Rutin as neuroprotective agent: From bench to bedside. Curr. Med. Chem. 2019, 26, 5152–5164. [Google Scholar] [CrossRef]
  223. Habtemariam, S. Rutin as a natural therapy for Alzheimer’s disease: Insights into its mechanisms of action. Curr. Med. Chem. 2016, 23, 860–873. [Google Scholar] [CrossRef] [PubMed]
  224. Pan, R.-Y.; Ma, J.; Kong, X.-X.; Wang, X.-F.; Li, S.-S.; Qi, X.-L.; Yan, Y.-H.; Cheng, J.; Liu, Q.; Jin, W. Sodium rutin ameliorates Alzheimer’s disease–like pathology by enhancing microglial amyloid-β clearance. Sci. Adv. 2019, 5, eaau6328. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Wang, S.-W.; Wang, Y.-J.; Su, Y.-J.; Zhou, W.-W.; Yang, S.-G.; Zhang, R.; Zhao, M.; Li, Y.-N.; Zhang, Z.-P.; Zhan, D.-W. Rutin inhibits β-amyloid aggregation and cytotoxicity, attenuates oxidative stress, and decreases the production of nitric oxide and proinflammatory cytokines. Neurotoxicology 2012, 33, 482–490. [Google Scholar] [CrossRef] [PubMed]
  226. Xu, P.-X.; Wang, S.-W.; Yu, X.-L.; Su, Y.-J.; Wang, T.; Zhou, W.-W.; Zhang, H.; Wang, Y.-J.; Liu, R.-T. Rutin improves spatial memory in Alzheimer’s disease transgenic mice by reducing Aβ oligomer level and attenuating oxidative stress and neuroinflammation. Behav. Brain Res. 2014, 264, 173–180. [Google Scholar] [CrossRef]
  227. Yu, X.-L.; Li, Y.-N.; Zhang, H.; Su, Y.-J.; Zhou, W.-W.; Zhang, Z.-P.; Wang, S.-W.; Xu, P.-X.; Wang, Y.-J.; Liu, R.-T. Rutin inhibits amylin-induced neurocytotoxicity and oxidative stress. Food Funct. 2015, 6, 3296–3306. [Google Scholar] [CrossRef]
  228. Sun, X.-Y.; Li, L.-J.; Dong, Q.-X.; Zhu, J.; Huang, Y.-R.; Hou, S.-J.; Yu, X.-L.; Liu, R.-T. Rutin prevents tau pathology and neuroinflammation in a mouse model of Alzheimer’s disease. J. Neuroinflamm. 2021, 18, 131. [Google Scholar] [CrossRef]
  229. Baluchnejadmojarad, T.; Jamali-Raeufy, N.; Zabihnejad, S.; Rabiee, N.; Roghani, M. Troxerutin exerts neuroprotection in 6-hydroxydopamine lesion rat model of Parkinson’s disease: Possible involvement of PI3K/ERβ signaling. Eur. J. Pharmacol. 2017, 801, 72–78. [Google Scholar] [CrossRef]
  230. Khan, 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. 2012, 22, 1–15. [Google Scholar] [CrossRef]
  231. Cordeiro, L.M.; Machado, M.L.; da Silva, A.F.; Baptista, F.B.O.; da Silveira, T.L.; Soares, F.A.A.; Arantes, L.P. Rutin protects Huntington’s disease through the insulin/IGF1 (IIS) signaling pathway and autophagy activity: Study in Caenorhabditis elegans model. Food Chem. Toxicol. 2020, 141, 111323. [Google Scholar] [CrossRef]
  232. Pandini, G.; Pace, V.; Copani, A.; Squatrito, S.; Milardi, D.; Vigneri, R. Insulin has multiple antiamyloidogenic effects on human neuronal cells. Endocrinology 2013, 154, 375–387. [Google Scholar] [CrossRef] [Green Version]
  233. Kim, B.; Feldman, E.L. Insulin resistance as a key link for the increased risk of cognitive impairment in the metabolic syndrome. Exp. Mol. Med. 2015, 47, e149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Nieto-Estévez, V.; Defterali, Ç.; Vicario-Abejón, C. IGF-I: A key growth factor that regulates neurogenesis and synaptogenesis from embryonic to adult stages of the brain. Front. Neurosci. 2016, 10, 52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Fernandez, A.M.; Torres-Alemán, I. The many faces of insulin-like peptide signalling in the brain. Nat. Rev. Neurosci. 2012, 13, 225–239. [Google Scholar] [CrossRef] [PubMed]
  236. Vanmierlo, T.; Bogie, J.F.; Mailleux, J.; Vanmol, J.; Lütjohann, D.; Mulder, M.; Hendriks, J.J. Plant sterols: Friend or foe in CNS disorders? Prog. Lipid Res. 2015, 58, 26–39. [Google Scholar] [CrossRef] [PubMed]
  237. Jie, F.; Yang, X.; Wu, L.; Wang, M.; Lu, B. Linking phytosterols and oxyphytosterols from food to brain health: Origins, effects, and underlying mechanisms. Crit. Rev. Food Sci. Nutr. 2022, 62, 3613–3630. [Google Scholar] [CrossRef]
  238. Burg, V.K.; Grimm, H.S.; Rothhaar, T.L.; Grösgen, S.; Hundsdörfer, B.; Haupenthal, V.J.; Zimmer, V.C.; Mett, J.; Weingärtner, O.; Laufs, U. Plant sterols the better cholesterol in Alzheimer’s disease? A mechanistical study. J. Neurosci. 2013, 33, 16072–16087. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  239. Sultana, N.; Khalid, A. Phytochemical and enzyme inhibitory studies on indigenous medicinal plant Rhazya stricta. Nat. Prod. Res. 2010, 24, 305–314. [Google Scholar] [CrossRef]
  240. Lee, J.; Weon, J.B.; Ma, C.J. Neuroprotective activity of phytosterols isolated from Artemisia apiacea. Korean J. Pharmacogn. 2014, 45, 214–219. [Google Scholar]
  241. Pratiwi, R.; Nantasenamat, C.; Ruankham, W.; Suwanjang, W.; Prachayasittikul, V.; Prachayasittikul, S.; Phopin, K. Mechanisms and neuroprotective activities of stigmasterol against oxidative stress-induced neuronal cell death via sirtuin family. Front. Nutr. 2021, 8, 648995. [Google Scholar] [CrossRef]
  242. Gabay, O.; Sanchez, C.; Salvat, C.; Chevy, F.; Breton, M.; Nourissat, G.; Wolf, C.; Jacques, C.; Berenbaum, F. Stigmasterol: A phytosterol with potential anti-osteoarthritic properties. Osteoarthr. Cartil. 2010, 18, 106–116. [Google Scholar] [CrossRef]
  243. Jie, F.; Yang, X.; Yang, B.; Liu, Y.; Wu, L.; Lu, B. Stigmasterol attenuates inflammatory response of microglia via NF-κB and NLRP3 signaling by AMPK activation. Biomed. Pharmacother. 2022, 153, 113317. [Google Scholar] [CrossRef] [PubMed]
  244. Park, S.J.; Kim, D.H.; Jung, J.M.; Kim, J.M.; Cai, M.; Liu, X.; Hong, J.G.; Lee, C.H.; Lee, K.R.; Ryu, J.H. The ameliorating effects of stigmasterol on scopolamine-induced memory impairments in mice. Eur. J. Pharmacol. 2012, 676, 64–70. [Google Scholar] [CrossRef] [PubMed]
  245. Nakanishi, S. Molecular diversity of glutamate receptors and implications for brain function. Science 1992, 258, 597–603. [Google Scholar] [CrossRef] [PubMed]
  246. Al Rahim, M.; Nakajima, A.; Saigusa, D.; Tetsu, N.; Maruyama, Y.; Shibuya, M.; Yamakoshi, H.; Tomioka, Y.; Iwabuchi, Y.; Ohizumi, Y. 4′-Demethylnobiletin, a bioactive metabolite of nobiletin enhancing PKA/ERK/CREB signaling, rescues learning impairment associated with NMDA receptor antagonism via stimulation of the ERK cascade. Biochemistry 2009, 48, 7713–7721. [Google Scholar] [CrossRef] [PubMed]
  247. Stohs, S.J.; Preuss, H.G.; Shara, M. A review of the human clinical studies involving Citrus aurantium (bitter orange) extract and its primary protoalkaloid p-synephrine. Int. J. Med. Sci. 2012, 9, 527. [Google Scholar] [CrossRef]
  248. Jung, Y.P.; Earnest, C.P.; Koozehchian, M.; Galvan, E.; Dalton, R.; Walker, D.; Rasmussen, C.; Murano, P.S.; Greenwood, M.; Kreider, R.B. Effects of acute ingestion of a pre-workout dietary supplement with and without p-synephrine on resting energy expenditure, cognitive function and exercise performance. J. Int. Soc. Sports Nutr. 2017, 14, 3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  249. Taslimi, P.; Akıncıoglu, H.; Gülçin, İ. Synephrine and phenylephrine act as α-amylase, α-glycosidase, acetylcholinesterase, butyrylcholinesterase, and carbonic anhydrase enzymes inhibitors. J. Biochem. Mol. Toxicol. 2017, 31, e21973. [Google Scholar] [CrossRef] [PubMed]
  250. Chang, H.-J.; Kim, H.J.; Chun, H.S. Quantitative structure—Activity relationship (QSAR) for neuroprotective activity of terpenoids. Life Sci. 2007, 80, 835–841. [Google Scholar] [CrossRef]
  251. Cheng, Y.; Dong, Z.; Liu, S. β-Caryophyllene ameliorates the Alzheimer-like phenotype in APP/PS1 Mice through CB2 receptor activation and the PPARγ pathway. Pharmacology 2014, 94, 1–12. [Google Scholar] [CrossRef]
  252. Guo, K.; Mou, X.; Huang, J.; Xiong, N.; Li, H. Trans-caryophyllene suppresses hypoxia-induced neuroinflammatory responses by inhibiting NF-κB activation in microglia. J. Mol. Neurosci. 2014, 54, 41–48. [Google Scholar] [CrossRef]
  253. Rao, J.-Y.; Wang, Q.; Wang, Y.-C.; Xiang, F.; Tian, X.-C.; Liu, D.-H.; Dong, Z. β-caryophyllene alleviates cerebral ischemia/reperfusion injury in mice by activating autophagy. Zhongguo Zhong Yao Za Zhi Zhongguo Zhongyao Zazhi China J. Chin. Mater. Med. 2020, 45, 932–936. [Google Scholar]
  254. NIH. β-Caryophyllene. Summary of Data for Chemical Selection. 1997. Available online: https://ntp.niehs.nih.gov/ntp/htdocs/chem_background/exsumpdf/betacaryophyllene_508.pdf (accessed on 1 October 2022).
  255. Ayaz, M.; Junaid, M.; Ullah, F.; Subhan, F.; Sadiq, A.; Ali, G.; Ovais, M.; Shahid, M.; Ahmad, A.; Wadood, A. Anti-Alzheimer’s studies on β-sitosterol isolated from Polygonum hydropiper L. Front. Pharmacol. 2017, 8, 697. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Bari, W.U.; Zahoor, M.; Zeb, A.; Khan, I.; Nazir, Y.; Khan, A.; Rehman, N.U.; Ullah, R.; Shahat, A.A.; Mahmood, H.M. Anticholinesterase, antioxidant potentials, and molecular docking studies of isolated bioactive compounds from Grewia optiva. Int. J. Food Prop. 2019, 22, 1386–1396. [Google Scholar] [CrossRef] [Green Version]
  257. Shi, C.; Wu, F.; Zhu, X.; Xu, J. Incorporation of β-sitosterol into the membrane increases resistance to oxidative stress and lipid peroxidation via estrogen receptor-mediated PI3K/GSK3β signaling. Biochim. Biophys. Acta (BBA)-Gen. Subj. 2013, 1830, 2538–2544. [Google Scholar] [CrossRef]
  258. Schmitz, G.; Grandl, M. Update on lipid membrane microdomains. Curr. Opin. Clin. Nutr. Metab. Care 2008, 11, 106–112. [Google Scholar] [CrossRef] [PubMed]
  259. Salazar, M.; Rojo, A.I.; Velasco, D.; de Sagarra, R.M.; Cuadrado, A. Glycogen synthase kinase-3β inhibits the xenobiotic and antioxidant cell response by direct phosphorylation and nuclear exclusion of the transcription factor Nrf2. J. Biol. Chem. 2006, 281, 14841–14851. [Google Scholar] [CrossRef]
  260. Sun, Y.; Gao, L.; Hou, W.; Wu, J. β-Sitosterol alleviates inflammatory response via inhibiting the activation of ERK/p38 and NF-κB pathways in LPS-exposed BV2 cells. BioMed Res. Int. 2020, 2020, 7532306. [Google Scholar] [CrossRef]
  261. Kim, H.-J.; Fan, X.; Gabbi, C.; Yakimchuk, K.; Parini, P.; Warner, M.; Gustafsson, J.-Å. Liver X receptor β (LXRβ): A link between β-sitosterol and amyotrophic lateral sclerosis–Parkinson’s dementia. Proc. Natl. Acad. Sci. USA 2008, 105, 2094–2099. [Google Scholar] [CrossRef] [Green Version]
  262. Ye, J.-Y.; Li, L.; Hao, Q.-M.; Qin, Y.; Ma, C.-S. β-Sitosterol treatment attenuates cognitive deficits and prevents amyloid plaque deposition in amyloid protein precursor/presenilin 1 mice. Korean J. Physiol. Pharmacol. 2020, 24, 39–46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  263. Hossen, M.A.; Ali Reza, A.; Amin, M.B.; Nasrin, M.S.; Khan, T.A.; Rajib, M.H.R.; Tareq, A.M.; Haque, M.A.; Rahman, M.A.; Haque, M.A. Bioactive metabolites of Blumea lacera attenuate anxiety and depression in rodents and computer-aided model. Food Sci. Nutr. 2021, 9, 3836–3851. [Google Scholar] [CrossRef]
  264. Baell, J.; Walters, M.A. Chemistry: Chemical con artists foil drug discovery. Nature 2014, 513, 481–483. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Gilberg, E.; Stumpfe, D.; Bajorath, J. Activity profiles of analog series containing pan assay interference compounds. RSC Adv. 2017, 7, 35638–35647. [Google Scholar] [CrossRef] [Green Version]
  266. Nelson, K.M.; Dahlin, J.L.; Bisson, J.; Graham, J.; Pauli, G.F.; Walters, M.A. The essential medicinal chemistry of curcumin: Miniperspective. J. Med. Chem. 2017, 60, 1620–1637. [Google Scholar] [CrossRef] [PubMed]
  267. Baker, M. Deceptive curcumin offers cautionary tale for chemists. Nature 2017, 541, 144–145. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  268. Daina, A.; Michielin, O.; Zoete, V. SwissADME: A free web tool to evaluate pharmacokinetics, drug-likeness and medicinal chemistry friendliness of small molecules. Sci. Rep. 2017, 7, 42717. [Google Scholar] [CrossRef] [Green Version]
  269. Cock, I. The safe usage of herbal medicines: Counter-indications, cross-reactivity and toxicity. Pharmacogn. Commun. 2015, 5, 2–50. [Google Scholar]
  270. Mathur, S.; Hoskins, C. Drug development: Lessons from nature. Biomed. Rep. 2017, 6, 612–614. [Google Scholar] [CrossRef]
  271. Drummond, E.; Wisniewski, T. Alzheimer’s disease: Experimental models and reality. Acta Neuropathol. 2017, 133, 155–175. [Google Scholar] [CrossRef] [Green Version]
  272. Ricciarelli, R.; Fedele, E. The amyloid cascade hypothesis in Alzheimer’s disease: It’s time to change our mind. Curr. Neuropharmacol. 2017, 15, 926–935. [Google Scholar] [CrossRef]
Figure 1. Some important phytocompounds from F. religiosa and F. benghalensis: 1, amyrin; 2, azelaic acid; 3, bergapten; 4, eudesmol; 5, eugenol; 6, kaempferol; 7, lanosterol; 8, leucoanthocyanins; 9, limonene; 10, linalool; 11, lupeol; 12, myo-inositol; 13, myricetin; 14, pinene; 15, psoralen; 16, quercetin; 17, rhein; 18, rutin; 19, stigmasterol; 20, synephrine; 21, β-caryophyllene; 22, β-sitosterol.
Figure 1. Some important phytocompounds from F. religiosa and F. benghalensis: 1, amyrin; 2, azelaic acid; 3, bergapten; 4, eudesmol; 5, eugenol; 6, kaempferol; 7, lanosterol; 8, leucoanthocyanins; 9, limonene; 10, linalool; 11, lupeol; 12, myo-inositol; 13, myricetin; 14, pinene; 15, psoralen; 16, quercetin; 17, rhein; 18, rutin; 19, stigmasterol; 20, synephrine; 21, β-caryophyllene; 22, β-sitosterol.
Nutrients 14 04731 g001
Figure 2. Mechanism of neuroprotection by various important phytochemicals from F. religiosa and F. benghalensis. Major phytochemicals act as inhibitors against oxidative stress and inflammation pathways. Oxidative stress is modulated by triggering antioxidant enzymes or transcription factors and suppressing ROS-mediated proteins. The cascade of inflammation is hampered by the downregulation of inflammatory transcription factors and caspase-related components. Abbreviations: CAT, catalase; GPx, glutathione peroxidase; GR, glutathione reductase; GSH, glutathione; GSSG, oxidized glutathione; GST, glutathione S-transferase; SOD, superoxide dismutase; ROS, reactive oxygen species; MDA, malondialdehyde; PKC, protein kinase C; HMGB1, high mobility group box 1; TLR4, toll-like receptor 4; Bcl-2, anti-apoptotic B cell lymphoma-2; Bax, Bcl-2-associated X; NF-κB; nuclear factor kappa B; MAPKs, mitogen-activated protein kinases; PI3K, phosphoinositol-3 phosphate; AKT, protein kinase B; MMP, matrix metalloproteinase; COX-2, cyclooxygenase 2; TNF-α, tumor necrosis factor α; sMAF, small musculoaponeurotic fibrosarcoma protein; Nrf2, nuclear E2-related factor 2; ARE, antioxidant response element; HO-1, heme oxygenase-1.
Figure 2. Mechanism of neuroprotection by various important phytochemicals from F. religiosa and F. benghalensis. Major phytochemicals act as inhibitors against oxidative stress and inflammation pathways. Oxidative stress is modulated by triggering antioxidant enzymes or transcription factors and suppressing ROS-mediated proteins. The cascade of inflammation is hampered by the downregulation of inflammatory transcription factors and caspase-related components. Abbreviations: CAT, catalase; GPx, glutathione peroxidase; GR, glutathione reductase; GSH, glutathione; GSSG, oxidized glutathione; GST, glutathione S-transferase; SOD, superoxide dismutase; ROS, reactive oxygen species; MDA, malondialdehyde; PKC, protein kinase C; HMGB1, high mobility group box 1; TLR4, toll-like receptor 4; Bcl-2, anti-apoptotic B cell lymphoma-2; Bax, Bcl-2-associated X; NF-κB; nuclear factor kappa B; MAPKs, mitogen-activated protein kinases; PI3K, phosphoinositol-3 phosphate; AKT, protein kinase B; MMP, matrix metalloproteinase; COX-2, cyclooxygenase 2; TNF-α, tumor necrosis factor α; sMAF, small musculoaponeurotic fibrosarcoma protein; Nrf2, nuclear E2-related factor 2; ARE, antioxidant response element; HO-1, heme oxygenase-1.
Nutrients 14 04731 g002
Table 1. Neuroprotective mechanism of Ficus extracts.
Table 1. Neuroprotective mechanism of Ficus extracts.
NamePlant PartExtractModelDose (mg/kg)ActionRef.
F. religiosaLeavesMethanolicBV2 cell lines Inhibits proinflammatory cytokine production; downregulates MAPK/ERK/JNK/NF-κB; improves the number and quality of neurons[36,64,65,66,67,68,69,70,71,72,73]
AlCl3-induced200 and 300
Petroleum ether3-NP-, 6-OHDA-induced200 and 400Anti-AChE; reduces oxidative stress
EthanolicScopolamine-, sodium nitrite- induced100Anti-amnesic and nootropic
RootHydroethanolicPTZ-induced1, 2, 4Anticonvulsant[74,75,76]
AqueousStrychnine-, PTZ-induced25, 50, 100Anticonvulsant
FruitMethanolicMES-, picrotoxin, scopolamine-induced25, 50, 100Antiamnesic, anticonvulsant[77,78,79,80]
10, 50, 100
Ethyl acetatePTZ-induced1, 2, 4Reduces oxidative stress, anticonvulsant, anti-AChE
BarkMethanolicIn vitro Anti-AChE[81]
F. benghalensisLeavesMethanolicAlloxan-induced200 and 400Improves motor coordination[82]
BarkMethanolicScopolamine- induced100, 200, 300Anxiolytic and antidepressant[62,83]
AqueousScopolamine- induced150 and 300Cognitive enhancement
RootAqueousPTZ-, MES-induced100 and 200Anxiolytic, memory-enhancing, muscle-relaxant, seizure-modifying effect[84]
Table 2. Neuroprotective mechanism of some important phytocompounds from sacred Ficus trees.
Table 2. Neuroprotective mechanism of some important phytocompounds from sacred Ficus trees.
NameClass and MWBBB PermeabilityModelDose/ConcentrationMOAPathways AffectedMedicinal Chemistry (PAINS)Ref.
AmyrinPhytosterol 426.72 PTZ-induced seizures25 and 50 mg/kgAntioxidantERK activation, GSK inhibition, memory enhancement; MAO inhibition; elevation of GABA; inhibits PKC; increases CAT; decreases MDA; inhibits AChE0 alerts[85,86,87,88,89,90,91,92,93,94]
Azelaic acidDicarboxylic acid 188.22 Rotenone-induced PD model80 mg/kg Improves motor functions0 alerts[95,96]
BergaptenFuranocoumarin 216.19Scopolamine-induced amnesia; paclitaxel-induced neuropathic pain 25 and 50 mg/kg; 25, 50, and 100 mg/kgEnzyme inhibitionInhibits AChE, BchE, and MAO; memory enhancement; anti-depressant0 alerts[97,98,99,100,101,102,103,104,105,106]
EudesmolSesquiterpenoids 222.37 PC-12 cells100 and 150 µMNeurite extensionInduced neurite extension; MAPK activation; phosphorylation of the CREB0 alerts[107,108,109,110]
EugenolPolyphenol 164.2TBI rats; I/R damage; acrylamide-induced neuropathic rats; aluminum-induced toxicity; hydroxydopamine-induced PD model 25, 50, and 100 mg/kg; 50 and 100 mg/kg; 10 mg/kg; 6 mg/kg; 0.1, 1, and 10 mg/kgAnti-inflammatory, autophagy, antioxidantImproves memory and motor functions; decreases AChE, TNF-α, and caspase-3; increases BDNF and serotonin; inhibits amyloid formation; increases MT-III, promotes neurogenesis0 alerts[111,112,113,114,115,116,117,118,119,120,121,122]
KaempferolFlavonoid 286.23 Anti-inflammatory, autophagy, antioxidant, anti-amyloidMMP inhibitor; BDNF modulation: antioxidant; reduces inflammatory cytokines, COX-2, HMGB1/TLR4; anti-AChE; increases dopamine; inhibits Abeta accumulation0 alerts[123,124,125,126,127,128,129,130,131]
LanosterolPhytosterol 426.71 (MPP+)-induced cell death in the PD cellular model 0.5 mMAutophagySuppresses the buildup of misfolded protein aggregations/sequestosomes; promotes autophagy; mitochondrial depolarization0 alerts[132,133,134,135]
LeucoanthocyaninsAnthocyanins 242.26Kainate-induced learning impairment in rats; LPS-treated adult mice; BV-2 cells2%; 24 mg/kg; 50 and 100 μg/mlAnti-inflammatory, autophagy, antioxidantModulates the PI3K/Akt/Nrf2/HO-1 pathway; COX-2/mPGES-1; promotes autophagy by upregulating AMPK–mTOR0 alerts[136,137,138,139,140,141,142,143,144,145,146,147]
LimoneneTerpene 136.24Aβ-induced in vitro model of AD; scopolamine-induced amnesia rat model; subchronic effects in rats10 μg/mL; MO: 1% and 3%; 5, 25, and 50 mg/kgAnti-inflammatoryImproves cognition; decreases MDA, increases SOD, GSH; anti-AChE and BChE; anti-inflammatory; increases GABA0 alerts[148,149,150,151,152,153,154]
LupeolPhytosterol 426.72Acetic acid-induced writhing, formalin test, carrageenan-induced hyperalgesia, and post-operative pain model; Aβ-induced oxidative stress in mice; TBI mouse model; cerebellar cultures25, 50, and 100 mg/kg; 50 mg/kg; 50 mg/kg; 0.1 µMAnti-inflammatory, antioxidantMAPK/JNK pathway; downregulates BACE-1, upregulates proinflammatory cytokines; downregulates TNF, iNOS, NLRP3; upregulates GDNF and SHH–GLI signaling0 alerts[164,165,166,167,168,169,170]
myo-InositolCarbocyclic sugar 180.16Kainic acid-induced epilepsy rat model; ischemic stroke injury in animals exposed to tobacco smoke; i streptozotocin-induced mice 0.1 μCi/mL; 30 mg/kg Improved memory and motor functions; anticonvulsant0 alerts[171,172,173,174,175,176,177]
MyricetinFlavonoid 318.23Aβ-induced in vitro model of AD; primary neuron cultures5 μM; 1 and 10 μMAnti-inflammatory, autophagy, antioxidant, anti-amyloidDecreases NF-κB and AMPK/SIRT1 signaling; reduces the levels of inflammatory mediators; autophagy; metal ion chelation; reduces A beta, anti-AChE; restores mitochondrial dysfunction1 alert[178,179,180,181,182,183,184,185,186,187,188,189,190]
PineneTerpene 136.24Aβ-induced rat model; PC-12 cells; focal ischemic stroke model of rats; cerebral ischemia–reperfusion in rats50 mg/kg; 10 and 25 µM; 100 mg/kg; 100 mg/kgAnti-inflammatory, autophagy, antioxidant, anti-amyloidImproves cognition; increases SOD, GSH, GPX, HO-1; suppresses the TNF-α/NF-κB pathway; increases the expression of choline acetyltransferase, Bcl-1, muscarinic receptors, nAChR, BDNF, and antioxidant transcription factors; decreases Bax, caspase-30 alerts[191,192,193,194,195,196]
PsoralenCoumarin 186.16 Ccopolamine-induced amnesia in rats; in vitro, in silico0.1 and 0.3 mg/kgEnzyme inhibitionAnti-AChE; anti-MAO0 alerts[197,198,199,200]
QuercetinFlavonoid 302.23Primary neuron cultures; MitoPark PD model; 3-NP-induced HD model20 μM; 20 and 40 μM; 25 and 175 mg/kg; 25 mg/kgAnti-inflammatory, antioxidantBACE-1 inhibitor, decreases proinflammatory cytokines, increases ATP, CAT, SOD; affects PON2, Nrf2–ARE, PI3K, JNK/ERK, TNF-α, SIRT1, CREB, MAPK, NF-κB, AMPK, PGC-1α1 alert[201,202,203,204,205,206,207,208,209,210,211,212,213,214,215,216]
RheinAnthraquinone 284.22CCI rats; I/R rats; TBI rat model; APP/PS1 mouse model of AD12 mg/kg; 50 and 100 mg/kg; 100 mg/kg; 10 mg/kgAnti-inflammatory, antioxidantIncreases SOD, GSH, CAT, GSH/GSSG, GSH-Px; enhances Bcl-2; decreases Bax, caspase-3, and ROS, proinflammatory cytokines; activation of the SIRT1/PGC-1α pathway; inhibits the NADPH oxidase/ROS/ERK/MMP-9 signaling pathway1 alert[163,217,218,219,220,221]
RutinFlavonoid-3-o-glycosides 610.51Tau-P301S mouse model; 6-OHDA-induced rat model of PD; Caenorhabditis elegans model of HD100 μL; 25 mg/kg; 15–120 μM Anti-inflammatory, anti-amyloid, antioxidantImproves memory, reduces Aβ oligomerization, oxidative stress, neurotoxicity, and neuroinflammation; reduces tau; protects dopaminergic neurons; insulin/insulin-like growth factor I pathway1 alert[228,229,230,231,232,233,234,235]
StigmasterolPhytosterol 412.69SH-SY5Y cells; BV2 cells; scopolamine-induced memory loss in mice1 μM; 10 and 20 μM; 3, 10, and 30 mg/kg; 10 mg/kgAnti-inflammatory, antioxidant, anti-amyloidAnti-AChE; reduces amyloid plaques; reduces ROS; modulates the SIRT1–FoxO3a pathway; inhibits proinflammatory cytokines; represses NF-κB and NLRP3 signaling by AMPK activation; NMDA activation; ERK/CREB activation; improves memory and LTP0 alerts[239,240,241,242,243,244,245,246]
SynephrineBiogenic amine 167.21 Pre-workout; in vitro20 mgEnzyme inhibitionAnti-BChE and anti-AChE activity; improves the cognitive function0 alerts[247,248,249]
β-CaryophylleneSesquiterpene 204.36APP/PS1 mice; BV2 cells; I/R injury mouse model48 mg/kg; 5 μM; 24 and 72 mg/kgAnti-inflammatory, autophagy, antioxidantAnti-BACE and anti-AChE activity; increases the expression of Bcl-2, beclin-1, CB2R; decreases p62; decreases ROS and proinflammatory cytokines0 alerts[250,251,252,253,254]
β-SitosterolPhytosterol 414.71In vitro; HT22 cells and primarily cultured hippocampal cells; LPS- induced BV2 cells15 µM; 8 and 16 µMAnti-inflammatory, antioxidantAntioxidant; anti-AChE and BChE; prevents plaque deposition; modulates the PI3K/GSK-3β pathway; increases ΔΨm and ATP; decreases the expression of IL-6, iNOS, TNF-α, COX-2, IκB, NF-κB, ERK/p380 alerts[255,256,257,258,259,260,261,262]
indicate that the compound can cross BBB.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shim, K.H.; Sharma, N.; An, S.S.A. Mechanistic Insights into the Neuroprotective Potential of Sacred Ficus Trees. Nutrients 2022, 14, 4731. https://doi.org/10.3390/nu14224731

AMA Style

Shim KH, Sharma N, An SSA. Mechanistic Insights into the Neuroprotective Potential of Sacred Ficus Trees. Nutrients. 2022; 14(22):4731. https://doi.org/10.3390/nu14224731

Chicago/Turabian Style

Shim, Kyu Hwan, Niti Sharma, and Seong Soo A. An. 2022. "Mechanistic Insights into the Neuroprotective Potential of Sacred Ficus Trees" Nutrients 14, no. 22: 4731. https://doi.org/10.3390/nu14224731

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