Next Article in Journal
Fucosylated Proteome Profiling Identifies a Fucosylated, Non-Ribosomal, Stress-Responsive Species of Ribosomal Protein S3
Next Article in Special Issue
Age-Dependent Hippocampal Proteomics in the APP/PS1 Alzheimer Mouse Model: A Comparative Analysis with Classical SWATH/DIA and directDIA Approaches
Previous Article in Journal
The Orai1-AC8 Interplay: How Breast Cancer Cells Escape from Orai1 Channel Inactivation
Previous Article in Special Issue
Cross-Reactivity and Sequence Homology Between Alpha-Synuclein and Food Products: A Step Further for Parkinson’s Disease Synucleinopathy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Neuroprotective Natural Products for Alzheimer’s Disease

1
Department of Pharmaceutical Sciences, College of Pharmacy and Health Sciences, Campbell University, Buies Creek, NC 27506, USA
2
Department of Pharmaceutical and Administrative Sciences, School of Pharmacy, Presbyterian College, Clinton, SC 29325, USA
*
Author to whom correspondence should be addressed.
Cells 2021, 10(6), 1309; https://doi.org/10.3390/cells10061309
Submission received: 21 April 2021 / Revised: 17 May 2021 / Accepted: 22 May 2021 / Published: 25 May 2021
(This article belongs to the Collection Advances in Neurodegenerative Disease)

Abstract

:
Alzheimer’s disease (AD) is the number one neurovegetative disease, but its treatment options are relatively few and ineffective. In efforts to discover new strategies for AD therapy, natural products have aroused interest in the research community and in the pharmaceutical industry for their neuroprotective activity, targeting different pathological mechanisms associated with AD. A wide variety of natural products from different origins have been evaluated preclinically and clinically for their neuroprotective mechanisms in preventing and attenuating the multifactorial pathologies of AD. This review mainly focuses on the possible neuroprotective mechanisms from natural products that may be beneficial in AD treatment and the natural product mixtures or extracts from different sources that have demonstrated neuroprotective activity in preclinical and/or clinical studies. It is believed that natural product mixtures or extracts containing multiple bioactive compounds that can work additively or synergistically to exhibit multiple neuroprotective mechanisms might be an effective approach in AD drug discovery.

1. Introduction

Alzheimer’s disease (AD), discovered by Dr. Alois Alzheimer in 1906 [1], is currently the number one chronic neurodegenerative disease, affecting more than six million people in the US and about 50 million people worldwide [2]. As a neurodegenerative disease, AD slowly and irreversibly destroys memory, cognition, and eventually the ability to carry out daily activities, leading to the need for full-time care, which is most common among people over 65, although it does occur in the younger population [3]. Age is considered the biggest risk factor for AD. About 3% of people aged 65–74, 17% of people between 75 and 84, and 32% of people 85 and over suffer from this disease [4,5]. With the size of the elderly population continually increasing, it is estimated that by the year 2050, there will be one new AD case every 33 s and the total number of AD cases will rise to 16 million in the US alone [2]. Being the most common cause of dementia among older adults and listed by the Centers for Disease Control and Prevention (CDC) as the sixth leading cause of death in the US, AD may actually be ranked third, just after heart disease and cancer, in major causes of death for older people [6].
In contrast to the high prevalence of AD, we only have five drugs approved by the FDA for its treatment, namely rivastigmine, galantamine, donepezil, memantine, and memantine in combination with donepezil. Furthermore, none of these drugs can reverse, stop, or even slow down the damage and destruction of neurons that cause AD symptoms and make the disease fatal. These largely unmet medical needs, together with the physical, emotional, societal, and healthcare-related impacts of AD, have challenged the research community to better understand and come up with more effective ways to manage this formidable disease. Enormous efforts have been devoted to the identification of the underlying mechanisms of this disease and the discovery of disease-modifying therapies that can halt the progression of AD [7].
Although the exact biological cause of AD in most people is still not fully understood, the hallmarks of this disease are believed to be the abnormal deposition of insoluble β-amyloid peptide (Aβ) and the accumulation of neurofibrillary tangles (NFTs) of phosphorylated tau protein in neuronal cytoplasm, which results in atrophy and neuron death [8]. Several biological processes have been linked to these neurodegenerative changes including neuroinflammation, depletion or insufficient synthesis of neurotransmitters, oxidative stress, collapse of calcium homeostasis, and abnormal ubiquitination [9]. As a result, neurons and synapses participating in memory processes are damaged, leading to a decline in learning ability and other cognitive functions. All the currently available medications for AD, except for memantine which blocks NMDA receptors in the brain from excess stimulation that can damage nerve cells, increase the amount of neurotransmitters in the brain to temporarily improve cognitive symptoms. These drugs’ effectiveness varies a lot individually and is limited in duration. Other approaches, including targeting Aβ plaques [10] and tau NFT aggregations [11], reducing oxidative stress [12] and neuroinflammation [13], etc., have been attempted, with a number of new drug candidates being evaluated in preclinical and clinical studies.
Natural sources, including plants, animals, microbes, and the marine world, provide abundant bioactive compounds with complex structures and novel pharmacological properties [14]. As one of the major sources of drug discovery, natural products and their isolated compounds have been extensively studied in efforts to develop more effective drugs for the management of AD [15]. In fact, the cholinesterase inhibitor galantamine is a natural product itself [16] and rivastigmine is a semi-synthetic derivative of a natural product called physostigmine [17]. Mixtures or extracts of natural products might have advantages compared to individual natural compounds, as they have multiple simultaneous target approaches, which could be a novel treatment option for AD, considering the complexity of its pathophysiology [18]. Mounting evidence has suggested that herbs or herbal formulations, together with mixtures obtained from other natural sources, may provide cognitive benefits to AD patients [15]. Consequently, various natural sources and their extracts are extensively employed in animal models and AD patients [19,20].
This review focuses on the natural product extracts or mixtures that may have neuroprotective effects through various mechanisms for the prevention and treatment of AD. The systematic literature search was conducted using SciFinder and PubMed as online databases until February 2021 with key words being neuroprotective natural products and AD. Among all the literature meeting our criteria, we especially focused on natural product extracts and mixtures.

2. Neuroprotective Mechanisms of Natural Products for AD

2.1. Overview of the Mechanisms Underlying AD

As an age-related progressive neurodegenerative disease accounting for most dementia cases, the intricate pathogenic mechanisms of AD are not fully understood. Besides the genetic and environmental factors which are believed to contribute to the etiology of AD, several hypotheses have been proposed to explain this complicated disorder with the most prevalent ones being the Aβ cascade hypothesis, tau hypothesis, inflammation hypothesis, cholinergic hypothesis, and oxidative hypothesis [21]. According to the Aβ cascade hypothesis, Aβ peptides are the causative agent in AD because the extracellular deposition of Aβ peptides as senile plaques (SP) and NFTs will result in neuron loss, vascular damage, and dementia [22]. Considered as another intracellular hallmark of AD, NFTs mainly consist of tau protein, which is a microtubule-associated scaffold protein enriched in the axons of neurons. Its aggregation impairs axons, causing neurodegeneration [23]. Recently, the inflammation hypothesis has emerged as the next major pathology in AD, involving the sustained immune response in the brain. Continued activation of the brain’s immune cells such as microglia results in the production and release of numerous proinflammatory cytokines, which not only leads to neuronal loss, but also facilitates both Aβ and tau pathologies [24]. Cholinergic neuronal damage has been widely accepted as a major pathological change correlating with the cognitive impairment in AD patients. Therefore, the cholinergic hypothesis suggests that a dysfunction of cholinergic neurons in the brain contributes substantially to the cognitive decline in AD [25]. This hypothesis is supported by the use of cholinesterase inhibitors in AD treatment. Oxidative stress has also been discovered to play an essential role in AD pathogenesis. Direct supporting evidence has indicated that AD is always accompanied with high cellular oxidative stress in the brain, caused by the increased generation of free radicals, increased lipid peroxidation and decreased polyunsaturated fatty acid, increased protein and DNA oxidation, as well as the accumulation and aggregation of Aβ, which also induces oxidative stress [26].

2.2. Neuroprotective Strategies for AD

With the substantial amount of evidence indicating that the primary causative factor in the pathogenesis of AD is the accumulation of Aβ [27], decreasing Aβ has become the major strategy in developing new therapeutics for AD [28]. However, successful AD therapeutic regimens may require multiple neuroprotective agents being used concomitantly. Through careful examination of the pathophysiological processes occurring in AD, several molecular targets have been identified as mediating these processes. These targets could aid in the development of potentially high-yield neuroprotective strategies [29]. Possible neuroprotective mechanisms focus on the inhibition of deleterious intraneuronal mechanisms triggered by Aβ and other toxic stimuli through specific interaction with various neuronal targets [30]. Practical neuroprotective approaches for managing AD include the discovery of small molecules to block Aβ interactions with its extracellular and intracellular targets [31], to minimize stress kinase signaling cascades [32], to prevent caspase activation [33] and pro-apoptotic protein expression [34], to inhibit excessive tau protein phosphorylation [35], to counteract cholinergic function loss [36], to promote the trophic state and neuron plasticity [37], to hinder reactive oxygen species accumulation [38], to suppress neuroinflammation [39] and to block excitotoxicity [40]. It is worth mentioning that some of the neuroprotective agents exhibit their effects through more than one approach. This is especially true with mixtures and extracts of natural products that contain more than one bioactive compound. Therefore, the neuroprotective effects from mixtures and extracts of natural products are always multidimensional and offer an advantage for the treatment of AD compared to single compound. Furthermore, the additive or synergistic action of crude extracts or mixtures can eliminate some of the side effects associated with the predominance of a single xenobiotic compound, providing a more comprehensive spectrum of activity, and minimizing the chances of pathogens developing resistance [41]

2.3. Neuroprotective Effects from Natural Products

Natural products have been shown to play neuroprotective roles through almost all of the different molecular mechanisms mentioned above (Figure 1). When focusing on the mixtures and extracts of natural products, the observed neuroprotective effects have typically been recognized as being obtained through anti-oxidative or anti-neuroinflammatory activities, preventing the aggregation of Aβ and tau protein as well as through enhancing cholinergic signaling. It is reasonable to speculate that the onset and progression of AD could be slowed down or even prevented by natural products working on multiple pathological targets [42].

2.3.1. Anti-Oxidative Neuroprotective Activity from Natural Products for AD

As molecules with radical and non-radical oxygen species, reactive oxygen species (ROS) and reactive nitrogen species (RNS) are highly chemically reactive. Maintaining a moderate concentration of these oxidative species is believed to play a crucial role in a number of physiological processes inside the human body, such as regulating the cell cycle, activating enzymes and receptors, as well as monitoring inflammation, phagocytosis, gene expression, and signal transduction [43,44,45,46]. The human body can neutralize these oxidative species and eliminate them so that their concentrations remain within the normal range. One of the major approaches includes the nuclear factor E2-related factor 2 (Nrf2) pathway. Nrf2 is a transcription factor that activates the expression of antioxidant genes in response to oxidative stress. A common response element of all these antioxidant genes is called antioxidant response element (ARE), which is highly involved in reducing oxidative stress, inflammation, and the accumulation of toxic metabolites [47]. However, when there is an imbalance between the production/accumulation of these oxidative species and their neutralization/elimination, the over-limitation concentrations of ROS and RNS can result in oxidative stress with serious pathological damage [48]. Specifically, because of the increased metabolic activity and limited cellular regeneration, oxidative stress shows more significant effects in the brain compared with other parts of the body [49]. In AD, Nrf2 expression is found to be upregulated in the neurons because of oxidative damage. At the same time, the levels of some ARE-containing gene products are reduced, indicating the disruption of this pathway [47]. Additionally, in vivo studies have also implied that inhibiting Keap1, which is the negative regulator of Nrf2, could prevent the Aβ42-mediated neurotoxicity that initiates AD [50]. Consequently, oxidative stress is widely accepted as playing a major role in progressively damaging neuron structure and impairing neuron function, which is considered one of the major causes of the development of severe neurodegenerative disorders, including AD [44]. Therefore, tremendous research efforts have been devoted to developing antioxidant therapies as neuroprotective agents for the treatment of AD [51].
Antioxidants are compounds that can react with free radicals to convert them into harmless species. Extensive research has identified oxidative stress as a primary factor in the development and progression of AD, and antioxidants as being capable of counteracting the damaging effects of oxidation [52]. Based on their occurrence in nature, antioxidants are grouped into two categories, natural and synthetic antioxidants, with most synthetic antioxidants being derived from natural products. The most well-known natural antioxidants are β-carotenoid (vitamin A), ascorbic acid (vitamin C), α-tocopherol (vitamin E), carotenoids, and flavonoids, all of which play an important role against ROS-induced damage in an organism.
Many of the natural antioxidants are obtained from plants and structurally belong to the classes of compounds called phenolics and polyphenolics, as well as carotenoids and antioxidant vitamins, among others. Phenolics and polyphenolics, which have one or more hydroxyl groups on their aromatic ring(s), have been well established as possessing high antioxidant capacity because of their capability to scavenge free radicals, donate hydrogen atoms and electrons, and chelate with metal cations [53]. The structures of these phenolics and polyphenolics, especially the hydroxyl groups substituted on the aromatic rings give them the capacity to perform their antioxidative function [54]. Being the most common group of polyphenolic compounds, flavonoids exhibit a wide range of antioxidative actions against free radical-mediated cell signaling, inflammation, and tumors [55]. As AD may be caused by this impaired cell signaling, it is not surprising that some of the natural flavonoids, derivatives of flavonoids, and natural sources rich in flavonoids are being extensively evaluated for the treatment of AD [56]. The neuroprotective effect of flavonoids is partially attributed to their antioxidative properties that can prevent free radical formation by modulating the cell signaling pathways involved in antioxidative protein expression, glutathione synthesis, as well as cell proliferation and survival [57].
Non-flavonoids, which is the other group of phenolic compounds, have slightly more variable structures compared to flavonoids. Some of the non-flavonoids such as phenolic acids, tannins, lignans, stilbenes, quinones, coumarins, and curcuminoids also exhibit high antioxidant activities, especially phenolic acids [58]. Dietary polyphenols have been demonstrated to prevent neuron damage and apoptosis in vitro and in vivo through attenuating ROS levels as a major mechanism for reducing the oxidative stress involved in the onset and progression of AD [59]. Some phenolic acid compounds isolated from wine have shown neuroprotective effects in vitro, partially via the prevention of RNS-induced stress injury [60]. Phenolic acids isolated from the medicinal plant Rosmarinus officinalis have exhibited neuroprotective activity in vitro through decreasing ROS/RNS levels and silencing Nrf2 expression [61]. As phenolic compounds are naturally occurring substances found in integral parts of the human diet like fruit, vegetables, nuts, seeds, flowers, and some herbal beverages, their dietary intake has been shown to be able to lower the incidence of chronic degenerative diseases, including AD. Phenolic compounds and their natural sources have received considerable attention recently for their antioxidant capacities as neuroprotective agents for better management of AD [51].
Carotenoids are another group of natural products that can interact with free radicals and exhibit antioxidant properties [62]. As lipophilic pigmented compounds primarily found in plants, algae, as well as in microorganisms including yeasts, fungi, archaea, and eubacteria, carotenoids have an isoprenoid skeleton which is folded and connected in a variety of ways to form numerous different structures possessing different physicochemical and biological properties, along with yellow, orange, or red colors. Their structures comprise conjugated double bonds within a long polyenic carbon chain and a nearly bilateral symmetry around the core, which allows electrons to move freely across the molecule and provides the capability to interact with free radicals to offer antioxidative activity [63]. Hence, when consumed in adequate levels, carotenoids have been demonstrated to bring many health benefits to the human body, especially the prevention of AD symptoms, partially through the reduction of ROS/RNS.
In fact, it has been found that AD patients have significantly lower serum levels of six measured carotenoids, along with other antioxidants, such as retinol and α-tocopherol. Moreover, antioxidants, including vitamins and carotenoids, may confer a reduction in oxidative damage as a neuroprotective defense mechanism in AD patients [64]. Dietary supplementation of carotenoids has been discovered to ameliorate AD symptoms in AD animal models [65]. The beneficial effects of consuming carotenoids such as lutein and zeaxanthin on reducing AD incidence was further supported by a recent clinical study [66]. A number of epidemiological studies have also connected the consumption of a carotenoid-rich diet with a decreased risk of neurodegenerative diseases, including AD [67].

2.3.2. Anti-Neuroinflammatory Neuroprotective Activity from Natural Products for AD

Although the exact pathophysiological mechanism of AD has not been elucidated yet, various possible mechanisms have been proposed to explain this multifactorial disorder including the Aβ hypothesis, tau hypothesis, cholinergic hypothesis, and inflammation hypothesis [68]. Neuroinflammation has been indicated to be associated with the deposit of Aβ in the brain, which is a major hallmark in the pathology of AD [69]. The neuroinflammatory process of AD results from elevated level of ROS, increased microglial activation, production of cytokines, and activated nuclear factor kappa B (NF-κB) [70]. Specifically, the activation of immune cells like microglia leads to the production and secretion of proinflammatory cytokines, including IFN-γ, IL-1β, and TNF-α [71], which stimulate the astrocytes nearby to generate Aβ42 oligomers [69]. Extracellular insoluble Aβ aggregates can attract more microglial cells to form microglia clusters at the Aβ deposition site [72]. These pro-inflammatory cytokines were found at elevated levels in the brain, serum, and cerebrospinal fluid of AD patients [72]. Countless studies have directly correlated the cognitive decline in AD patients with the increased levels of cytokines at all stages of the disease [73]. Best characterized as a transcription factor ubiquitously expressed to regulate the expression of many genes, NF-κB controls the encoding of proteins participating in the inflammatory and immunity processes [74]. Additionally, it is also directly involved in brain function, especially in neurodegenerative diseases like AD. The neuroprotective mechanism resulting from the activation of the Nrf2 pathway has also been correlated to the anti-inflammatory effects involving NF-κB [75] as NF-κB is a well-known negative regulator of Nrf2 [76]. Its activation has been shown to be linked with Aβ−induced neurotoxicity [77] and can be detected in the brains of AD patients [78]. Long-term usage of anti-inflammatory drugs can suppress the onset and progression of AD, indicating NF-κB as a key mediator of brain inflammation in AD [79].
The neuroprotective potential of natural products stemming from their anti-inflammatory properties, together with their antioxidant activities, has uncovered the potential to prevent and improve neurodegeneration in AD with minimal side effects compared to synthetic drugs [80]. As inflammation can also contribute to neurodegeneration and accelerate the progression of AD, natural products with anti-inflammatory properties might serve as a potential treatment to reduce the symptoms of AD, not only in the early prevention stage, but also in the disease management stage [80]. The effects of these natural products with anti-inflammatory activity are believed to have a synergistic effect involving interaction with multiple targets and modulation of multiple signaling pathways. Natural products, with their capacity to inhibit neuroinflammation, are able to produce anti-amyloid effects as a beneficial outcome in the management of AD [81]. Therefore, it is worthwhile to evaluate natural products or mixtures of natural products for their multi-target anti-inflammatory activity as potential therapeutic candidates for the prevention or management of AD. In fact, some plant-based and animal-based natural products, such as omega-3 fatty acids, have exhibited neuroprotective efficacy through their anti-inflammatory effects [82].
Several mechanisms have been reported for the anti-neuroinflammatory activity of some natural products, including the inhibition of microglia activation, the reduction of pro-inflammatory cytokine release from activated microglia, the inhibition of NF-κB, and the activation of p38 MAPK. The activity of those natural products that can activate Nrf2 and possess antioxidant properties is also at least partially attributed to their anti-neuroinflammatory activity [75].
Natural products with anti-neuroinflammatory activity can be found in almost all major structural categories like alkaloids, polyphenols, terpenes, carotenoids, and marine natural products. Alkaloids are nitrogen-containing (usually inside a ring structure) bioactive secondary metabolites with a wide variety of pharmacological activities, including anti-inflammatory effects. One example of an alkaloid with anti-neuroinflammatory activity is cryptolepine, obtained from Cryptolepis sanguinolenta. It has been reported that cryptolepine can reduce the levels of TNFα, IL-6, IL-1β, NO, and PGE2 in LPS-stimulated rat microglia via the blockage of the activation of NF-κB, p38 MAPK in the microglia [83]. Similarly, another alkaloid isolated from Radix Stephania tetrandra named tetrandrine shows promising anti-neuroinflammatory activity through the inhibition of NF-κB activation in a rat model of AD [84]. Flavonoids and other polyphenolic compounds have been extensively described to have anti-inflammatory activity. Examples in this structural group include kaempferol [85], tiliroside [86], apigenin [87], quercetin [88], epigallocatechin-3-gallate (EGCG) [89], punicalagin [90], urolithin A [91], mangiferin [92], resveratrol [93], curcumin [94], and many more. Their anti-inflammatory activity mostly involves the reduced production of pro-inflammatory mediators and the inhibition of activation of NF-κB and p38 MAPK pathways [95]. Within this group, flavonoids are considered as the most important subgroup for inhibiting neuroinflammation in AD because of their fundamental inhibitory actions on pro-inflammatory transcription factors, in addition to their capability of activating antioxidant/anti-inflammatory transcription factors. Terpenoids, having two or more isoprene units in their structures, have also been widely reported on for their anti-neuroinflammation properties in vitro and in animal models. Parthenolide, which is a sesquiterpene lactone present in Tanacetum parthenium [96], artemisinin, that is another sesquiterpene lactone found Artemisia annua [97], thymoquinone, which is the bioactive constituent of Nigella sativa, carnosic acid and carnosol, that are natural diterpenes found in Rosmarinus officinalis [98], and Ginkgo biloba extract, that is rich in polyphenolic compounds and terpene lactones [99], have all been demonstrated to inhibit neuroinflammation by reducing levels of pro-inflammatory mediators and regulating Nrf2, NF-κB and p38 MAPK pathways. Although a spice, Crocus sativus (Saffron) is famous for its wide variety of therapeutic applications, including AD [100]. Carotenoids obtained from Crocus sativus are believed to have anti-neuroinflammatory activity on cognitively impaired mice through the mechanism related to NF-κB [101]. Astaxanthin, a xanthophyll carotenoid found in Haematococcus pluvialis, Chlorella zofingiensis, Chlorococcum, and Phaffia rhodozyma, is probably the most intensively investigated marine natural product with neuroprotective potential. In BV-2 microglia, it was able to inhibit NO/iNOS and COX-2 [102], and attenuate the production of IL-6 induced by LPS through the activation of ERK1/2-MSK1 and NF-κB [103].

2.3.3. Anti-Aβ Aggregation Neuroprotective Activity from Natural Sources for AD

Produced through the proteolytic process of Aβ amyloid precursor protein (APP), a transmembrane protein, by β- and γ-secretases, Aβ is hypothesized to be the main cause of AD. Its accumulation in the brain is considered to be an early toxic event in AD pathogenesis that can cause memory loss and eventually lead to personality changes and cognitive decline over time [104]. The most toxic species of Aβ aggregates is believed to be the small, soluble aggregates called oligomers, which play an essential role in cell and tissue toxicity, especially in neurodegenerative diseases like AD [105]. While the exact toxicity mechanisms of amyloid oligomers remain unclear, results from several in vitro and in vivo studies revealed that high levels of amyloid oligomers were capable of over-stimulating the glutamatergic synaptic transmission and causing synapse loss [106]. They could also interact aberrantly with the cell membrane [105] and are reported to be associated with oxidative stress [107], inflammation [108], mitochondrial dysfunction [109], and disturbed calcium homeostasis, which is ultrasensitive to any changes in membrane permeability [110]. The amyloid cascade hypothesis of AD, which incorporates genetic, biochemical, and histological evidence, posits that the deposition of Aβ in the brain, due to an imbalance between its production and clearance, initiates a sequence of events that ultimately leads to AD dementia. The amyloid cascade hypothesis of AD proposes that Aβ aggregation is the primary event, and that aggregation of tau, inflammation and other changes observed in AD brains are downstream. Prohibiting the formation of Aβ aggregates, as well as reducing or removing them should be therapeutically useful in AD treatment.
Generally, the level of Aβ is determined by the rates of its production and removal through proteolytic degradation. Hence, any defects in the formation and degradation of Aβ may also affect its level and potentially contribute to AD. As the most prevalent mechanism in AD pathogenesis, the Aβ hypothesis indicates that extracellular deposits of Aβ form SP and NFTs, leading to neuronal loss and vascular damage [111]. The major proteinaceous Aβ peptide found in plaques of AD exists in the Aβ1–40 and Aβ1–42 forms, which are found to exist as monomers but in rapid equilibrium with the corresponding soluble oligomers through self-assembly [112].
One of the major focuses of drug discovery efforts in the AD area has been the development of drug candidates to regulate abnormal Aβ aggregation. Several synthetic drugs have been evaluated in clinical trials, including LY2886721 [113], AN1792 [114] and verubecestat [115], along with monoclonal antibody drugs such as donanemab, which targets Aβ3–42 [116] and aducanumab (BIIB037), which targets a conformational epitope found on Aβ [117]. However, none of them have been successful in the treatment of AD. Recently, rigorous investigation has been focused on exploring natural products for safer and inexpensive therapeutics, which can provide a better solution for managing AD [118]. There are two different major approaches in the development of such natural product-based AD therapies: targeting the secretases to stop the formation of Aβ, and interfering directly with Aβ aggregates [119].
Aβ peptides are formed after APPs get cleaved by secretases. After being synthesized in the brain, APPs translocate to the plasma membrane where they undergo specific endoproteolytic cleavages by α-, β-, and γ- secretases via either amyloidogenic or non-amyloidogenic pathways [120]. In the non-amyloidogenic pathway, α-secretase is the first enzyme to cleave APPs and generate the soluble N-terminal fragment (sAPPα) that is released into the extracellular space, and the membrane-tethered C terminal fragment CTFα which is further cleaved by γ-secretase to produce the extracellular P3 peptide and the APP intracellular domain (AICD) that is released into the cytosol [120]. In the amyloidogenic pathway, β-secretase, or beta-site amyloid precursor protein cleaving enzyme (BACE), cleaves APPs to create soluble APPβ fragment, which goes into the extracellular space, and leaves the rest of the protein (CTFβ) on the plasma membrane, which is further cleaved by γ-secretase to yield Aβ peptides with 40–43 amino acids [121]. In the non-amyloidogenic pathway, the cleavage of APPs by α-secretase first leads to the production of intracellular AICD fragments and extracellular P3 peptide with no Aβ deposition. However, in the amyloidogenic pathway, BACE cleaving APPs eventually results in the generation of Aβ peptides. Two major types of BACE, BACE1 and BACE2, have been found with similar structures. Because α- and β-secretases need to compete for the process of APP cleavage, and the cleavage of APP by α-secretase prevents the Aβ formation, two strategies have become very promising in developing drugs targeting these two secretases to overturn the neuropathological changes linked with most AD symptoms: increasing the activity of α-secretase or decreasing the activity of β-secretase [122].
Several natural compounds have demonstrated their capacity to regulate Aβ production in AD by modulating the activity of α- or β-secretase. As triterpene saponin (glycoside) components obtained from the root of ginseng (Panax ginseng), ginsenoside Rg1 can increase α-secretase level and decrease β-secretase level in vitro [123]. It can also significantly reduce the Aβ level in the cerebra of transgenic AD mice and improve the cognitive deficit of the mice [124]. From Glycyrrhiza glabra, a phenolic compound 2,2′,4′-trihydroxychalcone (TDC) has been isolated and has shown neuroprotective effects both in vitro and in vivo through decreasing BACE1 levels without affecting α- or γ-secretase levels [125]. Another phenolic compound, hispidin, extracted from the fungus Phellinus linteus, has been discovered to inhibit BACE1 in vitro without affecting the activities of α- and γ-secretases [126]. The polyphenol compound EGCG, most abundantly derived from green tea, has been proven to increase α-secretase secretion, and prevent abnormal Aβ aggregate formation in vitro [127]. At the same time, it can reduce the activity of β-secretase and inhibit the amyloidogenic pathway [128]. Extracted from black ginger or citrus peels and widely used for allergy and viral infections, polymethoxyflavones can reduce the enzymatic function of BACE1 in the amyloidogenic pathway while not affecting the activities of α-secretase as its mechanism for inhibiting BACE1-related amyloidogenesis in vitro [129]. As an isoquinoline alkaloid extracted from Coptidis rhizome and Berberis vulgaris, berberine could reduce BACE1 activity, decrease Aβ level, and improve behavioral symptoms in AD animal models via inhibiting β/γ-secretases activity and enhancing α-secretases [130]. Ligustilide, derived from Ligusticum chuanxiong, was able to promote the non-amyloidogenic pathway of APP cleavage in vitro and in vivo through increasing the activity of α-secretase [131].
During the process of Aβ peptide aggregation, phenylalanine 19 (F19) and phenylalanine 20 (F20) residues in the hydrophobic core domain are believed to be critical for the hydrophobic interactions among β-strands, which is thought to be the primary driving force for Aβ aggregation [132]. Therefore, targeting F19 and F20 residues of Aβ might be able to alter the structure of Aβ aggregates and decrease the related toxicity. Several natural products have been reported to be able to target these two phenylalanine residues and interact with Aβ through hydrogen bonds or hydrophobic interactions to destroy the normal structure of Aβ fibrils and block its toxicity in AD. Examples of these natural products include the polyphenol stilbene compound resveratrol that can decrease the formation of Aβ plaques [133], the major compound from the Sappan wood brazilin that can prevent abnormal Aβ aggregation [134], the polyphenol compound curcumin that interacts with F19 and F20 in the hydrophobic core domain of Aβ and disrupts their β-sheet structure to alleviate their toxicity [135], a polyphenol compound tannic acid that can stack with the hydrophobic domain of Aβ to prevent their polymerization into fibrils [136], destabilize the conformation of Aβ fibrils and promote the depolymerization of Aβ [137], and a group of polycyclic polyphenols theaflavin that binds to the hydrophobic core domain and hydrophobic C-terminus of Aβ to prevent the formation of toxic Aβ aggregates and induce the remodeling of misfolded Aβ aggregates [138].

2.3.4. Neuroprotective Activity Targeting Tau Protein from Natural Products for AD

The microtubule-associated protein tau is essentially disordered because of its high flexibility and lack of a stable conformation. One major hallmark of AD is the accumulation of NFTs of abnormal tau in neuronal cytoplasm. Therefore, tau needs to be detached from microtubules and then transferred into abnormal aggregates before the onset of AD occurs. This process is believed to be caused by a series of post-translational modifications such as phosphorylation, acetylation, ubiquitination, glycosylation, nitration, methylation, and so forth, with phosphorylation being the most important modification [139]. In fact, tau protein is hyperphosphorylated in the brain in AD, with 3–4 times more phosphorylation compared to a healthy brain [140]. Tau hyperphosphorylation promotes the dissociation of tau from microtubules and induces pathological tau aggregation [141]. Other abnormal post-translational modifications of tau, including acetylation [142], ubiquitination [143], methylation [144], glycation [145], etc., have also been identified in AD. Based on our knowledge about tau pathology in AD, several potential approaches have been proposed to block tau-mediated neurotoxicity, which mostly include inhibition of tau post-translational modifications and direct inhibition of tau aggregation.
Tau dephosphorylation is mainly carried out by protein phosphatase 2A (PP2A), which has reduced activity in the AD brain and cannot be easily targeted by drugs [146]. Subsequently, protein kinase inhibitors are usually developed with the intention to target other kinases to inhibit tau hyperphosphorylation or reduce tau aggregation rather than directly target PP2A [147]. Some natural drugs have been shown to inhibit tau hyperphosphorylation in AD animal models through modulating the activity of cyclin-dependent kinase-5 (CDK5) [148], glycogen synthase kinase-3 (GSK3) [149], or PP2A [150] to improve the symptoms of AD. Tongmai Yizhi Decoction, which contains six raw materials together with huperzine A, significantly decreases CDK5 and CDK5 expression in the hippocampus of model rats [151]. Safflower yellow from Carthamus tinctorius (Asteraceae) inhibits the GSK-3 activation and GSK-5 signaling pathways to prevent tau hyperphosphorylation by Aβ1–42 and improves learning and memory functions in AD model rats [148]. Geniposide isolated from the fruit of Gardenia jasminoides (Rubiaceae) reduces the hyperactivity of GSK3β induced by STZ and improves the spatial learning of rats [152]. Ginsenoside Rd from Panax ginseng increases PP2A activity and decreases okadaic acid-induced neurotoxicity, as well as tau hyperphosphorylation in vitro and in vivo [149].
Many of the tau aggregation inhibitors are natural products with antioxidant properties. Crocin from Crocus sativus (Iridaceae) can interfere with tau protein nucleation and inhibit tau protein filament formation in vitro [150]. Four simple quinones (1,4-benzoquinone, 1,4-naphthoquinone, 9,10-anthraquinone and 9,10-phenanthraquinone) from food or medicinal plants and four anthraquinone compounds (chrysophanol, emodin, aloe-emodin, and rhein) from Rheum rhabarbarum (Polygonaceae) have been found to inhibit the formation of toxic insulin oligomers in vitro, which may be helpful in preventing protein misfolding diseases like AD [153]. In vitro, the extracts of Glycyrrhiza inflata (Fabaceae) and P. ginseng (Araliaceae) are able to improve the growth of the repeat domain and axons in mutant tau protein to prevent tau aggregation. The aqueous extract of G. inflata (Fabaceae) can further upregulate unfolded protein response-mediated chaperones to reduce tau misfolding [154]. Another in vitro study has revealed that α-cyperone from rhizomes of Cyperus rotundus (Cyperaceae) exerts a significant effect on reducing the rate of tubulin polymerization and the concentration of polymerized tubulin, which may represent a beneficial neuroprotective strategy for AD [155]. Other well-known examples of natural products capable of inhibiting tau protein aggregation are curcumin, which inhibits amyloidogenic protein aggregation including Aβ and tau [156], resveratrol, which inhibits the aggregation of the repeat domain of tau along with many other neuroprotective mechanisms [157], purpurin, which inhibits tau fibrillization and breaks down the pre-formed fibrils [158], folic acid, which inhibits tau aggregation via stabilizing its native state [159], and the root extract of red ginseng [160].

2.3.5. Neuroprotective Activity Targeting Cholinergic Neurotransmission from Natural Sources for AD

The cholinergic system, which uses acetylcholine (ACh) as a neurotransmitter, is associated with a number of cognitive functions including learning and memory. The cholinergic hypothesis of AD proposes that the loss of these neurons in the brain is associated with cognitive deficits [161]. In fact, three out of the four approved single drugs work by increasing the lifespan of ACh by inhibiting the enzyme acetylcholinesterase (AChE), which is mainly responsible for metabolizing Ach, and has been recognized as the most effective drug target for developing AD drugs [162]. Other strategies targeting cholinergic neurons and acetylcholine to protect from AD-related neurotoxicity include promoting the expression of choline acetyltransferase (ChAT), which is a key enzyme in the acetylcholine biosynthetic pathway, and protecting cholinergic neurons by stimulating the expression of nerve growth factor (NGF), brain-derived neurotrophic factor (BDNF), and their receptors [163].
In a study with aqueous extracts from 80 traditional Chinese medicinal plants, Berberis bealei Fortune (Berberidaceae), Coptis chinensis Franch (Ranunculaceae) and Phellodendron chinensis (Rutaceae), all of which contain large amounts of isoquinoline alkaloids, were shown to effectively inhibit AChE function in vitro. Synergistically enhanced inhibitory activity has been found with the alkaloid combinations of berberine, coptisine, and palmatine [164]. Extracts from Huperzia serrata (Lycopodiaceae) inhibited AChE activity and ameliorated the cognitive impairment of AD mice [165]. A number of lycopodium alkaloids have been isolated and synthesized as potent AChE inhibitors [166]. Extracts from Crocus sativus (Iridaceae) exhibited moderate inhibitory activity against AChE. Crocetin, dimethylcrocetin, and safranal have all been found to possess moderate AChE inhibitory activities with IC50 values below or around 100 µM [167]. Gastrodia elata (Orchidaceae) could substantially increase ChAT expression in the medial septum and hippocampus to improve spatial memory in AD mice [168]. Compound Danshen Tablet increased ChAT expression in the brain, induced BDNF production, and activated the PKC receptor to improve spatial recognition in an AD rat model [169]. A variety of traditional Chinese medicine extracts demonstrated excellent therapeutic effects on AD through their effects on the expression of NGF, BDNF, and their related receptors in vivo [170]. The most famous ones are Bushen-Yizhi formula that could regulate NGF signal transduction and the anti-apoptotic cholinergic pathway to improve the IBO-induced memory impairment in an AD rat model [170], a bioactive components of ginger 6-shogaol that could increase the NGF levels of NGF and improve Aβ or scopolamine-induced memory impairment in animal models of dementia [171], Xanthoceras sorbifolium (Sapindaceae) extracts that could protect dendritic spines through the BDNF signal transduction pathway and improve cognition in an AD rat model [172], tanshinone IIA that could promote depolarization-induced BDNF synthesis [172], and polygonum multiflorum Thunberg complex composition-12 (PMC-12) that could increase BDNF level and synapse number in the hippocampus of an AD mice model [173].

3. Neuroprotective Natural Products for AD

When going through all the possible neuroprotective mechanisms that natural products may have to treat AD, it is clear that most of the individual natural compounds may exert their neuroprotective activity through more than one mechanism. One excellent example is the polyphenol flavonoid resveratrol whose neuroprotective potential consists of antioxidant activity [174], anti-neuroinflammatory activity [175], promoting the clearance of Aβ peptides [175], inhibition of GSK3β, and decreased brain levels of phosphorylated tau [176], as well as increasing cholinergic neurotransmission [177] and BDNF expression [178]. On the other hand, natural product mixtures or extracts, containing numerous individual compounds, may possess better neuroprotective potential as some of these compounds can work synergistically to present more profound effects on preventing neurotoxicity [164]. Moreover, extracts or mixtures of natural products directly obtained from their origins are more affordable therapeutic options with fewer side effects. In fact, some of these natural product mixtures or extracts have shown very promising neuroprotective activities in vitro and in vivo with quite a few being evaluated in clinical trials for AD right now. Here, we summarize the information about natural product mixtures or extracts with their neuroprotective activities for AD.

3.1. Neuroprotective Natural Products from Medicinal Plants for AD

Medicinal plants have been discovered to be able to decrease AD progress and symptoms [179]. Both the extracts and the active individual compounds from medicinal plants have been intensively investigated for their effects on AD [180]. The active compounds isolated from medicinal plants, such as phenolic lignans, flavonoids, tannins, and polyphenols, as well as triterpenes, sterols, and alkaloids, have exhibited various beneficial neuroprotective functions, including antioxidant, anti-neuroinflammatory, anti-amyloidogenic, anti-tau aggregation, and anticholinesterase activities [179]. Some of these active compounds, either as single components like curcumin, melatonin, resveratrol, and vitamins C and E, or as herbal extracts such as aged garlic extract, Ginkgo biloba extract, and green tea have been evaluated in AD patients with positive results [181]. Below, the information on the herbal extracts from medicinal plants that show neuroprotective effects for AD is summarized.

3.1.1. Pistacia Genus

Plants in the genus Pistacia are probably among the most precious natural resources with neuroprotective potential based upon their traditional applications. Related literature has shown that the neuroprotective effects of the genus Pistacia mainly include antioxidant, anti-neuroinflammatory, anti-Aβ aggregation, AChE inhibitory activity, and regulation of some other cellular pathways [182]. The kernel extract of Pistacia vera could inhibit cisplatin or vincristine-induced cognitive and motor impairments in vivo, which might be related to its high flavonoid and phenolic content [183]. Its seed oil has also been reported to improve memory and cognitive impairment in vivo [184]. The hydroalcoholic extract of pistachio nuts significantly improved learning and memory in the AD rat model [185]. The aqueous methanolic extract of its hull could inhibit AChE in vitro [186]. The hexane extract of the pistachio nuts showed antioxidant activity in vitro [182]. The gum extract also exhibited antioxidant neuroprotective effects in rats [182].
Pretreatment of the essential oil from the fruit of Pistacia lentiscus was reported to attenuate lipopolysaccharide-induced memory impairment in rats, decrease AChE activity and oxidative stress markers in brain tissue with the major components being 4-(3-[(2hydroxybenzoyl)amino] aniline)-4-oxobut-2-enoic acid, β-myrcene, 3-pentadecylphenol, P-tolyl ester, aminoformic acid, and β-sitosterol [187]. The aqueous extract of its leaves and the ethanolic extract of its oleoresin also possessed AChE inhibitory activity with IC50 values around 10 μg/mL [188]. Moreover, its dichloromethane extract oleoresin has also shown AChE inhibitory activity in vitro [189]. The methanolic extract could reduce Aβ-mediated cellular toxicity in SH-SY5Y cells [182]. The alcoholic extract from its leaves has also been shown to prevent oxidative damage-induced disorders, thanks to its substantial phenolic content, as well as significantly protect neuron cells against oxidative injury by Aβ25–35 and H2O2. This could almost completely protect the cells against Aβ-induced neurotoxicity [190]. The essential oil from the leaves has been demonstrated in multiple studies to possess anti-neuroinflammatory activity in vivo through various mechanisms such as inhibiting COX2 and stimulating PPAR-α etc. [191]. Taken together, this plant exerts neuroprotective effects which might be helpful in preventing and treating AD symptoms.
The ethyl acetate and aqueous extracts from the leaves of Pistacia atlantica have been described to have effective AChE inhibitory activity, probably attributed to the high phenolic content in these substances [182]. The methanolic and ethyl acetate extracts from the leaves showed powerful antioxidant properties comparable to known synthetic antioxidants which may be due to the constituents such as total flavonoids, total phenols, anthocyanins, chlorophyll, and carotenoid contents. In addition, both extracts showed moderate inhibitory activity against AChE with the ethyl acetate extract being more potent [182]. Essential oils obtained from the leaves and flowers, consisting of large amounts of monoterpenenes and oxygenated sesquiterpenes, have also been discovered to have protection against free radicals and oxidative stress as well as inhibiting AChE. For both free radical scavenging and anticholinesterase activities, leaf essential oil is reported to be better than flower oil [182]. Therefore, various extracts from this plant could potentially be used for the prevention of AD.
Several types of gall extracts of Pistacia integerrima have shown radical scavenging and cholinesterase inhibitory activity in vitro, which indicated that ethyl acetate extract exhibited the best radical scavenging activity and most potent AChE and butyrylcholinesterase inhibitory activities. Crude extract also demonstrated both AChE inhibitory and radical scavenging activities similar to those of the ethyl acetate extract. Free radical scavenging activity has been confirmed, with the two pure compounds isolated from this plant, quercetin and pyrogallol, as possessing high antioxidant activity especially for pyrogallol [192]. The in vivo study of the ethyl acetate and methanolic extracts of P. integerrima fruit together with four terebinth coffees revealed that both extracts showed moderate inhibitory activity towards butyrylcholinesterase but not acetylcholinesterase. They also exhibited radical scavenging activity at high concentrations. It has been revealed that terebinth coffee brands, with higher phenolic and flavonoid contents, possessed higher antioxidant and neuroprotective activities. The increased phenolic and flavonoid content may come from the roasting process, which suggests the roasting process of fruit as being a potential approach to improve the antioxidant properties of the extracts [182]. Similar to the other two plants in the same family, P. integerrima also has neuroprotective potential with mostly preventive effects for AD.

3.1.2. Panax Ginseng

Ginseng has been widely used in eastern countries as one of the main representatives of traditional medicine and presents a variety of pharmacological actions. Recent studies on the efficacy of Panax ginseng extract against AD have demonstrated that it could inhibit the neurotoxicity induced by Aβ in vitro and attenuate Aβ accumulation in vivo in the brain of AD animal models [193]. In a mouse AD model, the fermented ginseng extracts reduced Aβ formation in the brain and improved memory function [194]. Through inhibiting AChE, the extracts from white, red, and black ginseng were able to protect the memory dysfunction in mice caused by hippocampal Aβ oligomer injection [195]. Ginseng extract, when given orally to the mice injected with Aβ oligomer, restored the decreased synaptophysin and ChAT activity [196]. Panax ginseng extract has been evaluated in a number of clinical trials on AD patients. Red ginseng extract, administered long-term with AD drugs, gradually improved cognitive functions with minor side effects on AD patients [197]. Red ginseng extract treatment for more than 12 weeks was approved to improve the frontal cortical activity in elderly AD patients [198]. Hundreds of reports have been published for the effectiveness of various ginseng extracts including white ginseng, red ginseng, and fermented ginseng on the alleviation of AD symptoms in animal models and patients. It is indicated that certain biologically active chemicals in the ginseng extracts may prevent cognitive dysfunction by reducing Aβ formation and aggregation. Those bioactive chemicals are mainly ginsenosides and gintonin, which have been found to protect from Aβ-induced neurotoxicity and reactive oxidative stress, stimulate the formation of sAPPα instead of Aβ, exhibit anti-inflammatory function, as well as enhance cholinergic systems, hippocampal neurogenesis, and cognitive functions [199]. It is believed that the neuroprotective effects from ginseng can be used for the prevention and treatment of AD, an idea that is being evaluated clinically.

3.1.3. Phyllanthus Genus

The methanolic extract of Phyllanthus acidus (MEPA) exhibited neuroprotective effect through the improvement of cognitive functions and reduced oxidative stress via elevating the level of brain antioxidant enzymes as well as reducing lipid peroxidation and AChE activity. Thus, this plant extract can be useful in AD treatment [200]. To test the effects of Phyllanthus amarus and Cynodon dactylon in ameliorating AD-induced oxidative stress, the methanolic extracts of both plants were evaluated in an AD rat model. The results indicated that both extracts significantly increased the levels of superoxide dismutase, catalase, and NADH dehydrogenase compared to the control group. These antioxidant properties of the two herbal medicines may provide a new approach for AD treatment [201]. The ethanolic extracts of Phyllanthus emblica ripe and unripe fruits demonstrated marked beneficial effects on an AD mice model by improving the learning, memory, and antioxidant potential as well as decreasing AChE activity [202]. All the observed neuroprotective effects from P. acidus are both preventive and therapeutic for AD.

3.1.4. Ginkgo biloba L. (Ginkgoaceae)

Its leaf extract (EGb) has been well known for the capacity to improve memory and age-related deterioration, and thus has been widely used in dietary supplements. This neuroprotective effect may be associated with the ability to scavenge free radicals, prevent mitochondrial dysfunction, activate JNK and ERK pathways, and inhibit neuronal apoptosis [203]. Co-administration of EGb and donepezil exhibited better anti-amnestic effect through more augmented pro-cholinergic and antioxidative effects of both drugs in a scopolamine-induced AD rat model without any change in their systemic/brain exposure [204]. Numerous clinical trials have been carried out on EGb for single dose studies and long-term studies, with mixed results. Analysis of both positive and negative results revealed that EGb could improve the cognitive function in AD patients with mild dementia with long-term administration and appropriate dosage [205]. The phytochemicals of EGb that provide neuroprotective activity are believed to be flavonoids, organic acids, and terpenoids [206]. It is widely accepted that G. biloba possesses neuroprotective effects which can be both preventive and therapeutic for AD.

3.1.5. Hibiscus sabdariffa L. (Malvaceae)

As one of the most famous traditionally used remedies worldwide, it presents numerous pharmacological activities including sedative, antioxidant, anti-inflammatory, antidepressant, antiproliferative, antimicrobial, and neuroprotective activities. The neuroprotective effects of anthocyanin-enriched extracts from two Hibiscus varieties, white and red calyces, were tested in vitro to reveal their antioxidant potential and acetylcholinesterase inhibition activity. In vivo studies have indicated that Hibiscus extracts prevented memory impairment through the amelioration of STZ-induced neuroinflammation and amyloidogenesis. In summary, Hibiscus represents a promising safe preventive agent for AD with antioxidant, anti-inflammatory, anti-acetylcholinesterase, and anti-amyloidogenic activities. The LC/MS/MS analysis for the two Hibiscus extracts identified anthocyanins, flavonoids, and aliphatic and phenolic acids as being the major components responsible for the neuroprotective activity [207].

3.1.6. Hedera nepalensis K. (Araliaceae)

Treatment of crude extract (HNC) on the AD rat model showed an increase in the levels of catalase (CAT) and superoxide dismutase (SOD) while reducing glutathione (GSH) levels. Significantly elevated levels of dopamine and serotonin identified in the midbrain region and decreases in cognitive and memory impairment in the treatment group supported this plant as being a potential therapeutic agent for AD and diabetes [208].

3.1.7. Salvia miltiorrhiza B. (Lamiaceae)

As a widely used traditional Chinese medicine, the neuroprotective effects of this herbal medicine for AD treatment were reviewed along with its mechanisms and bioactive components [209]. Its extract protected SH-SY5Y cells against Aβ25–35-induced neurotoxicity through the inhibition of oxidative stress and the mitochondria-dependent apoptotic pathway [210]. A standardized fraction of S. miltiorrhiza, PF2401-SF, inhibited iNOS expression and NO production and exhibited anti-inflammatory activity on LPS-activated RAW 264.7 macrophages [211]. S. miltiorrhiza has also been found to efficiently induce neuron cell differentiation from rat mesenchymal stem cells [212]. While induced pluripotent stem cells (iPSCs) have the potential to differentiate into neural lineages, S. miltiorrhiza could promote the differentiation potential in vitro, and in vivo enhanced the survival and neural differentiation of transplanted iPSCs-derived neurons [213]. Its active chemical components, such as cryptotanshinone, tanshinone I, tanshinone IIA, Sal B, Sal A, and danshensu, have shown multiple neuroprotective effects including anti-Aβ, antioxidant, anti-apoptosis, anti-inflammation, enhancing cholinergic signaling, and inducing neurogenesis, which support the use of this plant as a preventive agent for AD [209].

3.1.8. Nardostachys jatamansi D. (Caprifoliaceae)

Its ethanolic extract was examined in vitro and in vivo for its neuroprotective effect. It was found that the extract, along with major component chlorogenic acid, could inhibit Aβ-induced cell death in vitro. In an in vivo study using a Drosophila AD model, this extract was able to rescue the neurological phenotypes of Aβ42-expressing flies as well as prevent Aβ42-induced cell death in the brain. Other neuroprotective effects of this extract may come from the reduced number of glial cells, decreased level of ROS, NO, and ERK phosphorylation in Aβ42-expressing flies without changing Aβ accumulation. Thus, as a promising herbal medicine for AD treatment, N. jatamansi exerts its neuroprotective activity most likely via the combination of its antioxidant and anti-neuroinflammatory properties together with the inhibitory activity towards ERK signaling [214]. Another in vivo screening program with Drosophila AD models also identified N. jatamansi as having neuroprotective effects against Aβ42 neurotoxicity [215]. All observed neuroprotective effects from this plant are both preventive and therapeutic for AD.

3.1.9. Viscum album L. (Santalaceae)

While disrupted BDNF levels have been indicated in AD pathogenesis, chronic treatment with V. album extract was found to significantly increase the BDNF levels in serum and diminish AlCl3-induced neurotoxicity in vitro and in vivo, which implied the neuroprotective effects of this plant which could be used as a preventive agent for AD [216].

3.1.10. Bacopa monnieri L. (Plantaginaceae)

As a commonly used traditional medicine, this plant contains a significant number of bioactive components including saponins and triterpenoids as main compounds together with alkaloids, sterols, and polyphenols which are famous for their antioxidant activity [217]. It was used traditionally to improve memory and cognitive function [218]. In an AD rat model, B. monnieri extract decreased cholinergic degeneration and showed cognition-enhancing effect [219]. Another study revealed that AChE could be inhibited by B. monnieri, which resulted in increased ACh levels [220]. Its extracts were also found to protect neuronal cells from β-amyloid-induced damage by lowering ROS levels [221]. A clinical trial demonstrated that the polyherbal formulation containing B. monnieri extract could improve the cognitive functions efficiently through decreasing the inflammatory level and oxidative stress in patients [222]. In summary, this neuroprotective plant could be used as a preventive and therapeutic agent for AD.

3.1.11. Convolvulus pluricaulis C. (Convolvulaceae)

The traditional medicine C. pluricaulis has been used for nervous system-related diseases such as stress, anxiety, mental fatigue, and insomnia [223]. Its ethanolic extract and the ethyl acetate and water fractions were found to significantly enhance learning and memory in rats [224]. Oral administration of C. pluricaulis alleviated the scopolamine-induced neurotoxic effect through decreasing tau and APP expression in the brain in AD rat models [225]. The active compounds that have been isolated from this plant including triterpenoids, flavanol glycosides, anthocyanins, and steroids, which are believed to provide the nootropic and memory-enhancing activity of the traditional medicine, making this plant a potential preventive agent for AD [226].

3.1.12. Centella asiatica L. (Apiaceae)

This herbal medicine is used traditionally for rejuvenating the neuronal cells and for increasing intelligence, longevity, and memory [227]. In a mice AD model, its extracts were able to reduce the β-amyloid pathology and oxidative stress in the brain [228]. Its ethanolic extracts were also reported to protect neurons against the neurotoxicity induced by Aβ1–40, decrease ROS production, and activate the antioxidative defense system by increasing the activities of various related enzymes and enhancing levels of glutathione and glutathione disulfide [229]. All these activities indicate the great potential of this traditional medicine for AD prevention and treatment [230]. The major bioactive compounds identified from C. asiatica are asiatic acid and asiaticoside, which exert antioxidant activity in vitro to reduce cytotoxicity induced by H2O2, decrease the levels of free radicals, and prevent the cell damage caused by Aβ accumulation [230,231].

3.1.13. Uncaria rhynchophylla M. (Rubiaceae)

As a medicinal herb used in the traditional Chinese medicine, U. rhynchophylla extract showed free radical scavenging activity and inhibited lipid peroxidation in an excitotoxicity animal model [232]. It has also been reported to exert protection against neuronal damage by reducing microglial activation, nNOS, iNOS, and apoptosis [233]. Moreover, it was found to inhibit Aβ fibril formation and also dissemble preformed Aβ fibrils in an AD model induced by Aβ1–40 and Aβ1–42 [234]. Bioactive compounds existing in U. rhynchophylla extract are mainly alkaloids such as rhynchophylline, isorhynchophylline, hirsutine, hirsuteine, corynanthine, corynoxine, and dihydrocorynantheine [235,236], among which rhynchophylline and isorhynchophylline are the most intensively studied and have been widely accepted as neuroprotective compounds [237]. All these neuroprotective effects suggest that this plant could be used for AD prevention and treatment.

3.1.14. Glycyrrhiza inflata B. (Fabaceae)

As one of the Glycyrrhiza species, G. inflata is known to increase mitochondrial biogenesis and decrease oxidative stress [238]. In AD cell models, its aqueous extract displayed a significant reduction in ROS and tau misfolding with low levels of cytotoxicity [154]. Another in vitro study found that G. inflata extract and its two constituents licochalcone A and liquiritigenin demonstrated potent anti-Aβ aggregation and radical-scavenging activities. Moreover, both G. inflata extract and its constituents suppressed the production of NO, TNFα, IL-1β, PGE2, and/or Iba1 in LPS-stimulated RAW 264.7 or BV-2 cells and further protected cell death. Taken together, G. inflata extract, licochalcone A, and liquiritigenin exhibit neuroprotective functions through anti-oxidative and anti-inflammatory activities to prevent neuronal apoptosis and could be used as preventive and therapeutic agents for AD [239].

3.1.15. Alpinia Oxyphylla-Schisandra Chinensis Herb Pair (ASHP)

The ethanolic extract of ASHP, together with its bioactive components, namely schisandrin (SCH) and nootkatone (NKT), was evaluated in the AD mice model. The results showed that in the object recognition task both the extract and SCH/NKT had a higher discrimination index with decreased levels of TNF-α, IL-1β, and IL-6. Thus, the neuroinflammation response was attenuated in both cases through the inhibition of the TLR4/NF-κB/NLRP3 pathway. Additionally, both treatments significantly restored the activities of glutathione S-transferase (GST), COX-2, SOD, total antioxidant capacity, and iNOS, as well as increased the levels of NO, GSH, and malondialdehyde. Both treatments were able to noticeably improve the histopathological changes of the hippocampus. Collectively, ASHP and its bioactive components exhibited neuroprotective effects by improving cognitive disorders, inhibiting the inflammatory reaction, and preventing oxidative stress, supporting their potential in AD prevention [240].

3.1.16. Buchanania axillaris D. (Anacardiaceae), Hemidesmus indicus L. (Apocynaceae) and Rhus mysorensis G. (Anacardiaceae)

The methanolic extracts of these three medicinal plants have been found to be most active against type II diabetes (T2D) and AD. Their derived fractions were tested in vitro for their inhibitory capacities against cholinesterases and α- & β-glucosidase, as well as their antioxidant potency in scavenging radicals. The results revealed that all the methanolic extracts of the test plants could inhibit AChE, BuChE, and α- and β-glucosidase in a dose-dependent manner. Evaluation of the subsequent fractionation indicated that chloroform fractions for all three extracts were most potent in inhibiting all four enzymes. These active chloroform fractions also showed great neuroprotective effects against the cell death-induced oxidative stress while not affecting cell viability. Overall, these three medicinal plants, with their methanolic extracts and the derived chloroform fractions having strong anticholinesterase, antiglucosidase, antioxidant, and neuroprotective activities, could be a multifunctional therapeutic regime for T2D and AD [241].

3.1.17. Coriandrum sativum L. (Apiaceae), N. Jatamansi, Polygonum multiflorum T. (Polygonaceae), Rehmannia glutinosa G. (Plantaginaceae), and Sorbus commixta H. (Rosaceae)

From a rapid in vivo screening platform using Drosophila AD models, these five medicinal plants out of 23 tested plants have been identified as exerting neuroprotective effects against Aβ42 neurotoxicity. Further investigation on the ethanolic extracts from P. multiflorum and S. commixta showed strong suppression of the AD neurological phenotypes. In vitro studies revealed that both ethanolic extracts increased the viability of Aβ-treated cells [215]. All observed effects support their utility in AD prevention and treatment.

3.1.18. Bojungikgi-Tang (BJIGT; Bu Zhong Yi Qi Tang in China, Hochuekkito in Japan)

This traditional oriental herbal formula consisting of eight medicinal herbs was found to be effective in treating dementia in a clinical study in South Korea [242]. Its neuroprotective effect was assessed in vitro and in vivo. The results indicated that BJIGT inhibited Aβ aggregation and enhanced BACE activity in vivo and antioxidant activity in vitro. It also exerted neuroprotective effects against H2O2-induced damage in vitro and remarkably ameliorated cognitive impairments in vivo. Additionally, it could prevent the aggregation and expression of Aβ peptides, as well as the expression of NeuN and BDNF in the hippocampi of Aβ-injected mice. Therefore, BJIGT may have great potential as a therapeutic option for treatment of AD and dementia [242].

3.1.19. Fuzhisan (FZS)

As a Chinese herbal complex which contains Scutellaria baicalensis G. (Labiatae), Ginseng root (Araliaceae), Glycyrrhiza uralensis F. (Leguminosae), and Anemone altaica F. (Araceae) [243], FZS has been clinically used for senile dementia for more than fifteen years [244]. Studies have indicated that FZS increased cognitive function in AD animal models [243] and AD patients [245]. Its neuroprotective effects are found to be associated with anti-apoptosis and anti-Aβ accumulation activities, as well as enhancing ACh levels and neurotrophic effects [246]. A recent study has also indicated that the neuroprotective function of FZS may protect against Aβ-induced neurotoxicity. All evidence implied that FZS may be clinically beneficial for AD patients.

3.2. Neuroprotective Natural Products from Food for AD

Some neuroprotective natural products, especially the antioxidant natural products, are naturally occurring in the human diet including fruits, vegetables, nuts, seeds, flowers, and some herbal beverages. It is of interest to summarize these food-related natural products for their neuroprotective properties.

3.2.1. Momordica charantia L. (Cucurbitaceae)

The fruits of four strains were dried and ground to give four mixtures, MC2, MC3, MC5, and MC5523, which were investigated in vitro and in vivo for their pharmacological functions in the treatment of AD. Neuroprotective effects have been observed for all four fruit powders while MC5523, in combination with LiCl, was found to increase the survival rate as well as enhance neuroprotection associated with anti-gliosis in an AD mice model. Moreover, MC5523 and LiCl cotreatment prevented memory deficits via reduced gliosis, oligomeric Aβ level, tau hyperphosphorylation, and neuronal loss while increasing the expression levels of synaptic-related protein and pS9-GSK3b, which provided a potential strategy for AD treatment [247]. Another study investigated the ethanolic extract of its fruits on memory impairment in an AD mice model and found that the ethanolic extracts could improve the anti-amnesic activity in mice via the inhibition of lipid peroxidation and the decreased AChE activity in the brain [248]. Most of the neuroprotective effects associated with M. chrantia are preventive for AD.

3.2.2. Benincasa hispida L. (Cucurbitaceae)

The aqueous extract of BH pulp was administered orally to AD mice before the intracerebroventricular infusion of colchicine (bilaterally), preventing SP formation. This neuroprotective effect might be attributed to the chemical constituents containing vitamins A, C, and E, flavanols, and flavonoids, which possess antioxidant activity. Thus, it was concluded that B. hispida, with its preventive potential for AD through antioxidant scavenging actions, protected rat neurons from the damage caused by colchicine-induced oxidative stress through the prevention of dentate granule cell destruction in the hippocampus and through preventing the extracellular deposition of senile plaques in a colchicine-induced AD rat model [249].

3.2.3. Allium sativum L. (Alliaceae)

The neuroprotective effects of aged garlic extract (AGE) have been well documented including antioxidant, anti-neuroinflammatory, and regulatory effects on neurotransmitter signaling in the brain that may be associated with the pathogenesis of AD [250,251]. A recent study examining the effect of AGE on Aβ1–42-induced cognitive dysfunction and neuroinflammation in an AD rat model demonstrated that in cognitively impaired rats, AGE could remarkably improve short-term recognition memory. Moreover, it reduced the activation of microglia and IL-1β level to substantially minimize the inflammatory response [252]. Low temperature-aged garlic extract has been discovered to suppress psychological stress through the modulation of stress hormones and oxidative stress response in the brain [253]. All observed effects from AGE are preventive for AD.

3.2.4. Curcuma longa L. (Zingiberaceae)

The active components in this spice are curcuminoids such as curcumin and related compounds [254]. Curcumin has been reported to have various interesting activities including anti-inflammatory, antioxidant, antitumor, and antibacterial activities [255]. Its potential in AD treatment has been reviewed before [256]. A recent review with a total of 32 studies that focused on curcumin’s effect on in vitro and in vivo AD models approved that curcumin may be a promising approach for AD [257]. Its ethanolic extract has been found to exert neuroprotective activity through antioxidant effects [258]. Another study on the herbal extract approved its neuroprotective effect as it was able to attenuate CeCl3-induced oxidative stress, enhance the activities of antioxidant enzymes, and decrease AChE activity [259]. It is hypothesized that people using turmeric frequently may have a lower incidence of AD, indicating turmeric’s preventive effect [260]. Studies have revealed a possible link between regular intake of turmeric as part of curry and excellent cognitive performance among the elderly [261]. For AD patients, turmeric treatment could also greatly improve their behavioral symptoms, supporting the therapeutic potential of turmeric for AD treatment [262].

3.2.5. Zingiber officinale R. (Zingiberaceae)

Widely used as extracts or as ingredients of ginger tea in food supplements, Z. officinale showed AChE inhibitory activity in vitro. It possesses the capacity to inhibit lipid peroxidation and exerts a neuroprotective effect against AD. In vivo studies on AD rat models with extract from Z. officinale have reported reduced lipid peroxidation levels. The neuroprotective mechanism associated with Z. officinale extract may be attributed to its ability to reduce the overstimulated NMDA receptors and prevent the production of free radicals [263], which may suggest the preventive and therapeutic potential of this plant for AD. The principal chemical components identified from Z. officinale include gingerols, shogaols, bisabolene, zingiberene, and monoterpenes [264].

3.2.6. Punica granatum (Pomegranate)

The juice and extracts from pomegranate fruit have been reported to possess neuroprotective effects against AD pathogenesis in various animal models [265,266,267]. Its neuroprotective mechanisms may involve counteracting oxidative stress [267], reducing brain inflammation [266], and decreasing accumulation of soluble Aβ42 and amyloid deposition in the hippocampus [265]. A study on the bioactive chemical components of pomegranate implied that the relevant brain-absorbable compounds responsible for the anti-AD effects of pomegranate were urolithins which showed protective effects on Aβ42-induced neurotoxicity and paralysis [268]. The observed neuroprotective effects are both preventive and therapeutic for AD.

3.2.7. Oryza sativa (Rice Berry)

Oxidative stress is well known to be associated with AD pathogenesis. The study of the rice berry, which is rich in antioxidant components, on an AD rat model revealed that it could significantly protect against memory impairment and neurodegeneration in the hippocampus. The hippocampal AChE activity and the lipid peroxidation products were also decreased, which suggested the rice berry as a potentially effective agent for AD prevention and treatment [269].

3.2.8. Vitis vinifera L. (Grape) and Red Wine

As a major component in the Mediterranean diet, the benefit of regular intake of red wine at moderate amounts for lowering AD and dementia development has been systematically reviewed and it is suggested that low to moderate wine drinking could reduce the risk of dementia and AD by at least a third, indicating its AD preventive role [270]. Evidence has been provided that extracts from grape skin and grape seed can inhibit Aβ aggregation, presenting an AD treatment potential [271]. In an AD rat model, red grape juice intake has been found to increase learning speed and improve memory [272]. Grape seed proanthocyanidins presented the capacity to ameliorate neuronal oxidative damage and cognitive impairment in an experimental AD model [273]. In vitro and in vivo studies on grape seed proanthocyanidins proved that they may be a novel therapeutic strategy for AD treatment [274]. Grape leaves’ polyphenolic extract has shown neuroprotective function through its antioxidative, anti-neuroinflammatory, and anti-amnesic activities against AlCl3-induced cerebral damages and neurocognitive dysfunction [275]. The bioactive components in grapes and red wine are believed to be polyphenols which are well established as antioxidants, with resveratrol being the most intensively studied for its neuroprotective effects.

3.2.9. Nuts Including Almond, Hazelnut, and Walnut

All of these nuts, considered as beneficial for the brain, exert neuroprotective effects that contribute to alertness, concentration, and memory [276]. Many studies have proven that nuts contain a rich matrix of bioactive chemicals and could possess the capacity to support neuronal function in the brain. The relation between nut consumption, improved cognitive performance, and lowered incidence of AD has been confirmed by several studies, supporting their AD preventive properties [277]. Adding hazelnut kernel into rats’ diet led to enhanced memory, reduced anxiety, and ameliorated neuroinflammation and apoptosis. As a dietary supplement, hazelnut has been confirmed to support healthy aging [278]. Almond supplementation in rats’ diet for one to two weeks significantly reversed amnesia induced by scopolamine. This effect was believed to have resulted from reduced AChE activity as well as lowered cholesterol and triglyceride levels, together with slightly increased glucose level [279]. Adding almond paste for 4 weeks in rats’ diet significantly improved learning and memory with enhanced brain tryptophan monoamine levels and serotonergic turnover in the brain [280]. Significantly improved memory retention was also observed in another study with 4 weeks of almond administration to rats, which was attributed to the elevated level of Ach [281]. Almond and walnut supplementation for 4 weeks attenuated cadmium-induced memory impairment in rats, possibly through cholinergic and antioxidant activities [282]. A walnut enriched diet has been shown to improve cognitive and motor performance [283]. Its neuroprotective effects are believed to be associated with the remarkably attenuated expression of proinflammatory cytokines, decreased level of AChE, significantly restored levels of antioxidant enzymes, and reduced expression of NF-κB [284]. Walnut could also increase the number of Ach receptors and upregulate expression of ChAT [285]. In rat PC12 cells, the peptides from defatted walnut protein exhibited significant free radical scavenging and cytoprotective activities against oxidative damage [286]. Walnut extract inhibited Aβ fibril formation and reduced Aβ-mediated cell death [287]. All observed neuroprotective effects associated with nuts are preventive for AD.

3.3. Neuroprotective Natural Products from Marine Sources for AD

Marine natural products, developed under adverse conditions and containing unusual structures, have been a rich source for the treatment of numerous diseases including AD. A number of marine natural products, together with their derived synthetic analogs, showed good efficacy against AD [288]. Thus, marine natural products are actively sought after as neuroprotective agents for AD.

3.3.1. Marine Macroalgae (Seaweeds)

Marine macroalgae which are plant-like organisms normally found in coastal areas contain a wide variety of bioactive chemicals such as polyphenols, polysaccharides, pigments, amino acids, peptides, and proteins. The health benefits associated with marine macroalgae and their bioactive compounds have been summarized in several excellent review papers [289]. The methanolic extract of Eisenia bicyclis (Kjellman) Setchell, a perennial brown seaweed, showed neuroprotective activity together with its ethyl acetate and n-butanol subfractions by reducing intracellular ROS production in PC12 cells induced with Aβ25–35. The components that contributed to this activity were believed to be phlorotannins eckol, phlorofucofuroeckol A, and 7-phloroeckol [290]. Another brown seaweed, Ishige foliacea has been found to contain a phlorotannin-rich fraction that improved memory impairment in mice through the additive or synergistic effect from several mechanisms including reducing brain AChE activity, suppressing oxidative stress, and activating the ERK-BDNF-CREB signaling pathway [291]. All these studies indicated that marine macroalgae present great options for AD prevention and treatment through protecting neuron damage and improving memory impairment by multiple pathways to exhibit neuroprotective activity which may result from the effects of its bioactive components and the additive or synergistic effects of those bioactive components.

3.3.2. Spirulina Cyanobacteria

Mostly referred to as blue-green algae, Cyanobacteria are actually prokaryotic organisms more closely related to bacteria with interesting pharmacological properties [292]. Spirulina platensis is a cyanobacterium known to have abundant nutritive elements such as carotenoids, polysaccharides, polyunsaturated fatty acids, vitamins, minerals, and protein [293]. It has been reported that S. platensis protein extract was a potent antioxidant which was able to scavenge free radicals with its chelating capacity and prevent radical-mediated cell death in vitro [294]. A follow-up study revealed that the aqueous extract of S. platensis and its active component C-phycocyanin could reduce cytotoxicity and inhibit the expression of inflammation-related genes like COX-2, TNF-α, IL-6, and iNOS in vitro, suggesting its AD preventive effects [295]. Another cyanobacterium, Spirulina maxima, contains many physiologically active chemicals including carotenoids, polysaccharides, chlorophylls, C-phycocyanin, and vitamins [296]. Its extract was found to ameliorate learning and memory impairments in an AD mice model, implicating its potential in AD treatment. This neuroprotective effect was believed to come from decreased expression levels of hippocampal Aβ1–42, APP, and BACE1, as well as decreased AChE activity, suppressed hippocampal oxidative stress, increased BDNF level, and activated BDNF/PI3K/Akt signaling pathways [297]. In another study with PC12 cells treated with Aβ1–42, the Spirulina maxima extract prevented Aβ-induced oxidative stress and cell death through the activation of BDNF signaling [298].

3.3.3. Thalassospira Profundimaris

A screening program with two hundred and twenty-five marine bacterial extracts looking at both their toxicity and neuroprotective properties identified several marine bacterial extracts as promising leads, with Thalassospira profundimaris being the most potent one. Its crude extract was able to preserve synaptic structure in vitro. An in vivo follow-up study revealed its AD preventive capacity because it could block the cell cycle-related neuron death. While none of the fractions derived from the crude extract exhibited neuroprotective activity as potent as the whole extract, it was hypothesized that the synergistic action of several components might be responsible for the overall effects [299].
All the neuroprotective natural products discussed above have been summarized below in Table 1.

4. Problems and Concerns with Natural Products for AD

While natural products and their isolated natural compounds have been well established as neuroprotective agents and valuable resources for exploring novel approaches in AD treatment, many of them still remain untested and their clinical use is not easy to monitor for a wide variety of reasons [300]. First of all, the quality of the natural source materials for natural products depends not only on genetic factors, but also on other extrinsic factors including environmental conditions, harvest time, and agricultural and collection practices for the source materials, making it difficult to perform quality control on the raw materials. The general requirements and methods for quality control of the finished mixture of natural products that contain hundreds of natural constituents are even more complex.
Although several of the above-mentioned natural product extracts or mixtures such as P. ginseng, EGb, B. monnieri, AGE, and C. longa have been evaluated in various stages of clinical trials, not all of them exhibited remarkable therapeutic effects in AD patients. However, they could still be used for the prevention of AD. Some specific limitations and challenges associated with natural products and their isolated natural compounds that might affect their clinical efficacy in AD treatment are their physicochemical instability, their limited water solubility, their rapid metabolism, their low bioavailability, and their distribution to the CNS, which have been adequately summarized in several reviews [301,302,303,304]. Several neuroprotective natural compounds, especially the polyphenols like resveratrol and curcumin, are chemically unstable and can be easily degraded or converted to inactive derivatives [303]. The presence of the blood-brain barrier requires enough lipophilicity from the natural products to be able to penetrate the CNS, which may create extra obstacles for some neuroprotective natural products, including polyphenols and polysaccharides, and limit their clinical efficacy. Carotenoids and alkaloids, on the other hand, are lipophilic enough to cross the blood-brain barrier, but their poor water solubility brings other problems which lead to low bioavailability. It has also proved very challenging to translate the exciting preclinical results of the neuroprotective natural products to clinical applications. Although several natural product mixtures like EGb and ginger extracts have been evaluated in AD patients, no conclusive positive results have been obtained.
To improve the bioavailability of the neuroprotective natural products useful in AD treatment, nanotechnology and nanocarrier-based strategies have been developed in the delivery of natural product mixtures and the isolated bioactive compounds which may enhance the therapeutic response and improve clinical efficacy [305,306]. The most commonly used nanoparticles include polymeric nanoparticles, solid lipid nanoparticle, crystal nanoparticle, nanogels, liposomes, micelles, and dendrimer complexes. Several studies have described the incorporation of nanoparticle delivery systems for natural products and their bioactive compounds. One noteworthy example is the nanolipidic EGCG particles, which have been approved to almost double the neuronal α-secretase enhancing ability in vitro and improve EGCG’s oral bioavailability in vivo by more than two-fold [127].

5. Conclusions

Mounting evidence has demonstrated the great neuroprotective potentials of natural products and natural bioactive compounds in AD treatment with few harmful side effects. Although not fully understood, the pathological process associated with AD is believed to be multifactorial. Neuroprotective strategies involving multiple mechanisms of action are important for the prevention and treatment of AD. Natural product mixtures or extracts, with multiple bioactive compounds and the ability to exert multiple neuroprotective mechanisms, are preferable in AD drug discovery. With more practical and comprehensive quality control guidelines developed to ensure the safety and efficacy of natural product therapies, as well as new approaches and strategies to help promote the CNS access of these neuroprotective agents, such as the incorporation of nanotechnology in the delivery of natural products, natural product therapy could play an essential role in the prevention and treatment of AD.

Author Contributions

X.C. did the bibliographic research and wrote the manuscript. J.D., W.B. and W.L. contributed with the paper organization, writing, and discussion. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

ADAlzheimer’s Disease
amyloid β peptide
NFTneurofibrillary tangle
ROSreactive oxygen species
RNSreactive nitrogen species
Nrf2nuclear-factor erythroid 2-related factor
AREantioxidant response element
NF-κΒnuclear factor κΒ
MAPKmitogen-activated protein kinase
IL-6interleukin 6
LPSlipopolysaccharide
EGCGepigallocatechin-3-gallate
ERKextracellular signal-regulated kinase
APPamyloid precursor protein
SPsenile plaques
AICDAPP intracellular domain
BACEbeta-site amyloid precursor protein cleaving enzyme
CTF-βcarboxy-terminal fragment beta
PP2Aprotein phosphatase 2A
CDK-5cyclin-dependent kinase 5
GSK-3glycogen synthase kinase-3
AChacetylcholine
AChEacetylcholinesterase
ChATcholine acetyltransferase
BDNFbrain-derived neurotrophic factor
PKCprotein kinase C
NGFnerve growth factor
PPAR-αperoxisome proliferator-activated receptor alpha
AChTacetylcholine transferase
sAPPasoluble amyloid precursor protein a
MEPAmethanolic extract of Phyllanthus acidus
EGbGinkgo biloba
CATcatalase
SODsuperoxide dismutase
GSHglutathione
GSTglutathione S-transferase
BuChEbutyrylcholinesterase
T2Dtype II diabetes
AGEaged garlic extract

References

  1. Hippius, H.; Neundorfer, G. The discovery of Alzheimer′s disease. Dialogues Clin. Neurosci. 2003, 5, 101–108. [Google Scholar]
  2. 2020 Alzheimer’s disease facts and figures. Alzheimers Dement. 2020. [CrossRef]
  3. Kawas, C.H.; Corrada, M.M. Alzheimer’s and dementia in the oldest-old: A century of challenges. Curr. Alzheimer Res. 2006, 3, 411–419. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Farrer, L.A.; Cupples, L.A.; Haines, J.L.; Hyman, B.; Kukull, W.A.; Mayeux, R.; Myers, R.H.; Pericak-Vance, M.A.; Risch, N.; van Duijn, C.M. Effects of age, sex, and ethnicity on the association between apolipoprotein E genotype and Alzheimer disease. A meta-analysis. APOE and Alzheimer Disease Meta Analysis Consortium. JAMA 1997, 278, 1349–1356. [Google Scholar] [CrossRef] [PubMed]
  5. Hebert, L.E.; Weuve, J.; Scherr, P.A.; Evans, D.A. Alzheimer disease in the United States (2010–2050) estimated using the 2010 census. Neurology 2013, 80, 1778–1783. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. James, B.D.; Leurgans, S.E.; Hebert, L.E.; Scherr, P.A.; Yaffe, K.; Bennett, D.A. Contribution of Alzheimer disease to mortality in the United States. Neurology 2014, 82, 1045–1050. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Long, J.M.; Holtzman, D.M. Alzheimer Disease: An Update on Pathobiology and Treatment Strategies. Cell 2019, 179, 312–339. [Google Scholar] [CrossRef]
  8. Silva, M.V.F.; Loures, C.M.G.; Alves, L.C.V.; de Souza, L.C.; Borges, K.B.G.; Carvalho, M.D.G. Alzheimer’s disease: Risk factors and potentially protective measures. J. Biomed. Sci. 2019, 26, 33. [Google Scholar] [CrossRef] [Green Version]
  9. Shal, B.; Ding, W.; Ali, H.; Kim, Y.S.; Khan, S. Anti-neuroinflammatory Potential of Natural Products in Attenuation of Alzheimer’s Disease. Front. Pharm. 2018, 9, 548. [Google Scholar] [CrossRef]
  10. Schenk, D.; Basi, G.S.; Pangalos, M.N. Treatment strategies targeting amyloid beta-protein. Cold Spring Harb. Perspect Med. 2012, 2, a006387. [Google Scholar] [CrossRef] [Green Version]
  11. Congdon, E.E.; Sigurdsson, E.M. Tau-targeting therapies for Alzheimer disease. Nat. Rev. Neurol. 2018, 14, 399–415. [Google Scholar] [CrossRef] [PubMed]
  12. Thapa, A.; Carroll, N.J. Dietary Modulation of Oxidative Stress in Alzheimer’s Disease. Int. J. Mol. Sci. 2017, 18, 1583. [Google Scholar] [CrossRef] [Green Version]
  13. Zhang, B.; Gaiteri, C.; Bodea, L.G.; Wang, Z.; McElwee, J.; Podtelezhnikov, A.A.; Zhang, C.; Xie, T.; Tran, L.; Dobrin, R.; et al. Integrated systems approach identifies genetic nodes and networks in late-onset Alzheimer’s disease. Cell 2013, 153, 707–720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Yuan, H.; Ma, Q.; Ye, L.; Piao, G. The Traditional Medicine and Modern Medicine from Natural Products. Molecules 2016, 21, 559. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Kennedy, D.O.; Wightman, E.L. Herbal extracts and phytochemicals: Plant secondary metabolites and the enhancement of human brain function. Adv. Nutr. 2011, 2, 32–50. [Google Scholar] [CrossRef] [PubMed]
  16. Scott, L.J.; Goa, K.L. Galantamine: A review of its use in Alzheimer’s disease. Drugs 2000, 60, 1095–1122. [Google Scholar] [CrossRef] [PubMed]
  17. Kumar, V. Potential medicinal plants for CNS disorders: An overview. Phytother. Res. 2006, 20, 1023–1035. [Google Scholar] [CrossRef] [PubMed]
  18. Shao, R.; Xiao, J. Natural Products for Treatment of Alzheimer’s Disease and Related Diseases: Understanding their Mechanism of Action. Curr. Neuropharmacol. 2013, 11, 337. [Google Scholar] [CrossRef]
  19. Akram, M.; Nawaz, A. Effects of medicinal plants on Alzheimer’s disease and memory deficits. Neural Regen Res. 2017, 12, 660–670. [Google Scholar] [CrossRef]
  20. Andrade, S.; Ramalho, M.J.; Loureiro, J.A.; Pereira, M.D.C. Natural Compounds for Alzheimer’s Disease Therapy: A Systematic Review of Preclinical and Clinical Studies. Int. J. Mol. Sci. 2019, 20, 2313. [Google Scholar] [CrossRef] [Green Version]
  21. Liu, P.; Xie, Y.; Meng, X.; Kang, J. History and progress of hypotheses and clinical trials for Alzheimer’s disease. Sig. Transduct Target Ther. 2019, 4, 29. [Google Scholar] [CrossRef] [PubMed]
  22. Hardy, J.A.; Higgins, G.A. Alzheimer′s disease: The amyloid cascade hypothesis. Science 1992, 256, 184–185. [Google Scholar] [CrossRef] [PubMed]
  23. Frost, B.; Jacks, R.L.; Diamond, M.I. Propagation of tau misfolding from the outside to the inside of a cell. J. Biol. Chem. 2009, 284, 12845–12852. [Google Scholar] [CrossRef] [Green Version]
  24. Kinney, J.W.; Bemiller, S.M.; Murtishaw, A.S.; Leisgang, A.M.; Salazar, A.M.; Lamb, B.T. Inflammation as a central mechanism in Alzheimer’s disease. Alzheimers Dement. 2018, 4, 575–590. [Google Scholar] [CrossRef] [PubMed]
  25. Francis, P.T.; Palmer, A.M.; Snape, M.; Wilcock, G.K. The cholinergic hypothesis of Alzheimer’s disease: A review of progress. J. Neurol. Neurosurg Psychiatry 1999, 66, 137–147. [Google Scholar] [CrossRef] [PubMed]
  26. Markesbery, W.R. Oxidative stress hypothesis in Alzheimer’s disease. Free Radic. Biol. Med. 1997, 23, 134–147. [Google Scholar] [CrossRef]
  27. Hardy, J.; Selkoe, D.J. The amyloid hypothesis of Alzheimer’s disease: Progress and problems on the road to therapeutics. Science 2002, 297, 353–356. [Google Scholar] [CrossRef] [Green Version]
  28. Yiannopoulou, K.G.; Papageorgiou, S.G. Current and Future Treatments in Alzheimer Disease: An Update. J. Cent. Nerv. Syst. Dis. 2020, 12, 1179573520907397. [Google Scholar] [CrossRef] [Green Version]
  29. Longo, F.M.; Massa, S.M. Neuroprotective strategies in Alzheimer’s disease. NeuroRx 2004, 1, 117–127. [Google Scholar] [CrossRef]
  30. Niikura, T.; Tajima, H.; Kita, Y. Neuronal cell death in Alzheimer′s disease and a neuroprotective factor, humanin. Curr. Neuropharmacol. 2006, 4, 139–147. [Google Scholar] [CrossRef]
  31. Ding, Y.; Zhao, J.; Zhang, X.; Wang, S.; Viola, K.L.; Chow, F.E.; Zhang, Y.; Lippa, C.; Klein, W.L.; Gong, Y. Amyloid Beta Oligomers Target to Extracellular and Intracellular Neuronal Synaptic Proteins in Alzheimer’s Disease. Front. Neurol. 2019, 10, 1140. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Du, Y.; Du, Y.; Zhang, Y.; Huang, Z.; Fu, M.; Li, J.; Pang, Y.; Lei, P.; Wang, Y.T.; Song, W.; et al. MKP-1 reduces Abeta generation and alleviates cognitive impairments in Alzheimer’s disease models. Signal. Transduct Target. 2019, 4, 58. [Google Scholar] [CrossRef] [Green Version]
  33. Quiroz-Baez, R.; Ferrera, P.; Rosendo-Gutierrez, R.; Moran, J.; Bermudez-Rattoni, F.; Arias, C. Caspase-12 activation is involved in amyloid-beta protein-induced synaptic toxicity. J. Alzheimers Dis. 2011, 26, 467–476. [Google Scholar] [CrossRef] [PubMed]
  34. Zhang, H.; Zhang, Y.W.; Chen, Y.; Huang, X.; Zhou, F.; Wang, W.; Xian, B.; Zhang, X.; Masliah, E.; Chen, Q.; et al. Appoptosin is a novel pro-apoptotic protein and mediates cell death in neurodegeneration. J. Neurosci. 2012, 32, 15565–15576. [Google Scholar] [CrossRef] [PubMed]
  35. Iqbal, K.; Liu, F.; Gong, C.X.; Grundke-Iqbal, I. Tau in Alzheimer disease and related tauopathies. Curr. Alzheimer Res. 2010, 7, 656–664. [Google Scholar] [CrossRef] [Green Version]
  36. Ferreira-Vieira, T.H.; Guimaraes, I.M.; Silva, F.R.; Ribeiro, F.M. Alzheimer′s disease: Targeting the Cholinergic System. Curr. Neuropharmacol. 2016, 14, 101–115. [Google Scholar] [CrossRef] [Green Version]
  37. Black, I.B. Trophic regulation of synaptic plasticity. J. Neurobiol. 1999, 41, 108–118. [Google Scholar] [CrossRef]
  38. Tonnies, E.; Trushina, E. Oxidative Stress, Synaptic Dysfunction, and Alzheimer’s Disease. J. Alzheimers Dis. 2017, 57, 1105–1121. [Google Scholar] [CrossRef] [Green Version]
  39. Heneka, M.T.; Carson, M.J.; El Khoury, J.; Landreth, G.E.; Brosseron, F.; Feinstein, D.L.; Jacobs, A.H.; Wyss-Coray, T.; Vitorica, J.; Ransohoff, R.M.; et al. Neuroinflammation in Alzheimer’s disease. Lancet Neurol. 2015, 14, 388–405. [Google Scholar] [CrossRef] [Green Version]
  40. Wang, R.; Reddy, P.H. Role of Glutamate and NMDA Receptors in Alzheimer’s Disease. J. Alzheimers Dis. 2017, 57, 1041–1048. [Google Scholar] [CrossRef] [Green Version]
  41. Olila, D.; Olwa, O.; Opuda-Asibo, J. Antibacterial and antifungal activities of extracts of Zanthoxylum chalybeum and Warburgia ugandensis, Ugandan medicinal plants. Afr. Health Sci. 2001, 1, 66–72. [Google Scholar] [PubMed]
  42. Venkatesan, R.; Ji, E.; Kim, S.Y. Phytochemicals that regulate neurodegenerative disease by targeting neurotrophins: A comprehensive review. Biomed. Res. Int. 2015, 2015, 814068. [Google Scholar] [CrossRef]
  43. Hussain, T.; Tan, B.; Yin, Y.; Blachier, F.; Tossou, M.C.; Rahu, N. Oxidative Stress and Inflammation: What Polyphenols Can Do for Us? Oxid Med. Cell Longev. 2016, 2016, 7432797. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Islam, M.T. Oxidative stress and mitochondrial dysfunction-linked neurodegenerative disorders. Neurol. Res. 2017, 39, 73–82. [Google Scholar] [CrossRef] [PubMed]
  45. Manoharan, S.; Guillemin, G.J.; Abiramasundari, R.S.; Essa, M.M.; Akbar, M.; Akbar, M.D. The Role of Reactive Oxygen Species in the Pathogenesis of Alzheimer’s Disease, Parkinson’s Disease, and Huntington’s Disease: A Mini Review. Oxid Med. Cell Longev. 2016, 2016, 8590578. [Google Scholar] [CrossRef] [PubMed]
  46. Liu, Z.; Zhou, T.; Ziegler, A.C.; Dimitrion, P.; Zuo, L. Oxidative Stress in Neurodegenerative Diseases: From Molecular Mechanisms to Clinical Applications. Oxid. Med. Cell Longev. 2017, 2017, 2525967. [Google Scholar] [CrossRef] [PubMed]
  47. Itoh, K.; Chiba, T.; Takahashi, S.; Ishii, T.; Igarashi, K.; Katoh, Y.; Oyake, T.; Hayashi, N.; Satoh, K.; Hatayama, I.; et al. An Nrf2/small Maf heterodimer mediates the induction of phase II detoxifying enzyme genes through antioxidant response elements. Biochem. Biophys. Res. Commun. 1997, 236, 313–322. [Google Scholar] [CrossRef]
  48. Rani, V.; Deep, G.; Singh, R.K.; Palle, K.; Yadav, U.C. Oxidative stress and metabolic disorders: Pathogenesis and therapeutic strategies. Life Sci. 2016, 148, 183–193. [Google Scholar] [CrossRef]
  49. Liguori, I.; Russo, G.; Curcio, F.; Bulli, G.; Aran, L.; Della-Morte, D.; Gargiulo, G.; Testa, G.; Cacciatore, F.; Bonaduce, D.; et al. Oxidative stress, aging, and diseases. Clin. Interv. Aging 2018, 13, 757–772. [Google Scholar] [CrossRef] [Green Version]
  50. 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.; et al. Direct Keap1-Nrf2 disruption as a potential therapeutic target for Alzheimer’s disease. PLoS Genet. 2017, 13, e1006593. [Google Scholar] [CrossRef] [Green Version]
  51. Teleanu, R.I.; Chircov, C.; Grumezescu, A.M.; Volceanov, A.; Teleanu, D.M. Antioxidant Therapies for Neuroprotection-A Review. J. Clin. Med. 2019, 8, 1659. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Prasad, K.N.; Hovland, A.R.; Cole, W.C.; Prasad, K.C.; Nahreini, P.; Edwards-Prasad, J.; Andreatta, C.P. Multiple antioxidants in the prevention and treatment of Alzheimer disease: Analysis of biologic rationale. Clin. Neuropharmacol. 2000, 23, 2–13. [Google Scholar] [CrossRef] [PubMed]
  53. Chandran, R.; Sajeesh, T.; Parimelazhagan, T. Total Phenolic Content, Anti-Radical property and HPLC profiles of Caralluma diffusa (Wight) N.E. Br. J. Biol. Act. Prod. Nat. 2014, 4, 188–195. [Google Scholar] [CrossRef]
  54. Cosme, P.; Rodriguez, A.B.; Espino, J.; Garrido, M. Plant Phenolics: Bioavailability as a Key Determinant of Their Potential Health-Promoting Applications. Antioxidants 2020, 9, 1263. [Google Scholar] [CrossRef]
  55. Kumar, S.; Pandey, A.K. Chemistry and biological activities of flavonoids: An overview. Sci. World J. 2013, 2013, 162750. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. De Andrade Teles, R.B.; Diniz, T.C.; Costa Pinto, T.C.; de Oliveira Junior, R.G.; Gama, E.S.M.; de Lavor, E.M.; Fernandes, A.W.C.; de Oliveira, A.P.; de Almeida Ribeiro, F.P.R.; da Silva, A.A.M.; et al. Flavonoids as Therapeutic Agents in Alzheimer’s and Parkinson’s Diseases: A Systematic Review of Preclinical Evidences. Oxid. Med. Cell Longev. 2018, 2018, 7043213. [Google Scholar] [CrossRef] [PubMed]
  57. Frandsen, J.R.; Narayanasamy, P. Neuroprotection through flavonoid: Enhancement of the glyoxalase pathway. Redox Biol. 2018, 14, 465–473. [Google Scholar] [CrossRef]
  58. Vuolo, M.M.; Lima, V.S.; Maróstica Junior, M.R. Chapter 2—Phenolic Compounds: Structure, Classification, and Antioxidant Power. In Bioactive Compounds; Campos, M.R.S., Ed.; Woodhead Publishing: Cambridge, UK, 2019; pp. 33–50. [Google Scholar]
  59. Calabrese, V.; Butterfield, D.A.; Stella, A.M. Nutritional antioxidants and the heme oxygenase pathway of stress tolerance: Novel targets for neuroprotection in Alzheimer’s disease. Ital. J. Biochem. 2003, 52, 177–181. [Google Scholar]
  60. Esteban-Fernandez, A.; Rendeiro, C.; Spencer, J.P.; Del Coso, D.G.; de Llano, M.D.; Bartolome, B.; Moreno-Arribas, M.V. Neuroprotective Effects of Selected Microbial-Derived Phenolic Metabolites and Aroma Compounds from Wine in Human SH-SY5Y Neuroblastoma Cells and Their Putative Mechanisms of Action. Front. Nutr. 2017, 4, 3. [Google Scholar] [CrossRef] [Green Version]
  61. De Oliveira, M.R.; Ferreira, G.C.; Schuck, P.F. Protective effect of carnosic acid against paraquat-induced redox impairment and mitochondrial dysfunction in SH-SY5Y cells: Role for PI3K/Akt/Nrf2 pathway. Toxicol. Vitr. 2016, 32, 41–54. [Google Scholar] [CrossRef]
  62. Fiedor, J.; Burda, K. Potential role of carotenoids as antioxidants in human health and disease. Nutrients 2014, 6, 466–488. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Young, A.J.; Lowe, G.L. Carotenoids-Antioxidant Properties. Antioxidants 2018, 7, 28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Mullan, K.; Williams, M.A.; Cardwell, C.R.; McGuinness, B.; Passmore, P.; Silvestri, G.; Woodside, J.V.; McKay, G.J. Serum concentrations of vitamin E and carotenoids are altered in Alzheimer’s disease: A case-control study. Alzheimers Dement. Transl. Res. Clin. Interv. 2017, 3, 432–439. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Obulesu, M.; Dowlathabad, M.R.; Bramhachari, P.V. Carotenoids and Alzheimer’s Disease: An insight into therapeutic role of retinoids in animal models. Neurochem. Int. 2011, 59, 535–541. [Google Scholar] [CrossRef] [PubMed]
  66. Yuan, C.; Chen, H.; Wang, Y.; Schneider, J.A.; Willett, W.C.; Morris, M.C. Dietary carotenoids related to risk of incident Alzheimer dementia (AD) and brain AD neuropathology: A community-based cohort of older adults. Am. J. Clin. Nutr. 2020, 113, 200–208. [Google Scholar] [CrossRef] [PubMed]
  67. Cho, K.S.; Shin, M.; Kim, S.; Lee, S.B. Recent Advances in Studies on the Therapeutic Potential of Dietary Carotenoids in Neurodegenerative Diseases. Oxidative Med. Cell. Longev. 2018, 2018, 4120458. [Google Scholar] [CrossRef]
  68. Yiannopoulou, K.G.; Papageorgiou, S.G. Current and future treatments for Alzheimer’s disease. Adv. Neurol. Disord. 2013, 6, 19–33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Dal Pra, I.; Chiarini, A.; Gui, L.; Chakravarthy, B.; Pacchiana, R.; Gardenal, E.; Whitfield, J.F.; Armato, U. Do astrocytes collaborate with neurons in spreading the "infectious" abeta and Tau drivers of Alzheimer’s disease? Neuroscientist 2015, 21, 9–29. [Google Scholar] [CrossRef] [PubMed]
  70. Von Bernhardi, R.; Eugenin-von Bernhardi, L.; Eugenin, J. Microglial cell dysregulation in brain aging and neurodegeneration. Front. Aging Neurosci. 2015, 7, 124. [Google Scholar] [CrossRef] [Green Version]
  71. Tan, M.S.; Yu, J.T.; Jiang, T.; Zhu, X.C.; Tan, L. The NLRP3 inflammasome in Alzheimer’s disease. Mol. Neurobiol. 2013, 48, 875–882. [Google Scholar] [CrossRef]
  72. Streit, W.J.; Mrak, R.E.; Griffin, W.S. Microglia and neuroinflammation: A pathological perspective. J. Neuroinflamm. 2004, 1, 14. [Google Scholar] [CrossRef] [Green Version]
  73. Garcez, M.L.; Mina, F.; Bellettini-Santos, T.; Carneiro, F.G.; Luz, A.P.; Schiavo, G.L.; Andrighetti, M.S.; Scheid, M.G.; Bolfe, R.P.; Budni, J. Minocycline reduces inflammatory parameters in the brain structures and serum and reverses memory impairment caused by the administration of amyloid beta (1–42) in mice. Prog. Neuropsychopharmacol. Biol. Psychiatry 2017, 77, 23–31. [Google Scholar] [CrossRef]
  74. Li, Q.; Verma, I.M. Erratum: NF-κB regulation in the immune system. Nat. Rev. Immunol. 2002, 2, 725–734. [Google Scholar] [CrossRef] [Green Version]
  75. Sandberg, M.; Patil, J.; D’Angelo, B.; Weber, S.G.; Mallard, C. NRF2-regulation in brain health and disease: Implication of cerebral inflammation. Neuropharmacology 2014, 79, 298–306. [Google Scholar] [CrossRef] [Green Version]
  76. Yu, M.; Li, H.; Liu, Q.; Liu, F.; Tang, L.; Li, C.; Yuan, Y.; Zhan, Y.; Xu, W.; Li, W.; et al. Nuclear factor p65 interacts with Keap1 to repress the Nrf2-ARE pathway. Cell Signal. 2011, 23, 883–892. [Google Scholar] [CrossRef]
  77. Longpre, F.; Garneau, P.; Christen, Y.; Ramassamy, C. Protection by EGb 761 against beta-amyloid-induced neurotoxicity: Involvement of NF-kappaB, SIRT1, and MAPKs pathways and inhibition of amyloid fibril formation. Free Radic Biol. Med. 2006, 41, 1781–1794. [Google Scholar] [CrossRef]
  78. Boissiere, F.; Hunot, S.; Faucheux, B.; Duyckaerts, C.; Hauw, J.J.; Agid, Y.; Hirsch, E.C. Nuclear translocation of NF-kappaB in cholinergic neurons of patients with Alzheimer’s disease. Neuroreport 1997, 8, 2849–2852. [Google Scholar] [CrossRef]
  79. Ju Hwang, C.; Choi, D.Y.; Park, M.H.; Hong, J.T. NF-kappaB as a Key Mediator of Brain Inflammation in Alzheimer’s Disease. CNS Neurol. Disord. Drug Targets 2019, 18, 3–10. [Google Scholar] [CrossRef]
  80. Cooper, E.L.; Ma, M.J. Alzheimer Disease: Clues from traditional and complementary medicine. J. Tradit Complement. Med. 2017, 7, 380–385. [Google Scholar] [CrossRef]
  81. Zhao, L.; Wang, J.-L.; Liu, R.; Li, X.-X.; Li, J.-F.; Zhang, L. Neuroprotective, Anti-Amyloidogenic and Neurotrophic Effects of Apigenin in an Alzheimer’s Disease Mouse Model. Molecules 2013, 18, 9949–9965. [Google Scholar] [CrossRef]
  82. Wollen, K.A. Alzheimer’s disease: The pros and cons of pharmaceutical, nutritional, botanical, and stimulatory therapies, with a discussion of treatment strategies from the perspective of patients and practitioners. Altern. Med. Rev. 2010, 15, 223–244. [Google Scholar]
  83. Olajide, O.A.; Bhatia, H.S.; de Oliveira, A.C.; Wright, C.W.; Fiebich, B.L. Inhibition of Neuroinflammation in LPS-Activated Microglia by Cryptolepine. Evid. Based Complement. Altern. Med. 2013, 2013, 459723. [Google Scholar] [CrossRef] [Green Version]
  84. He, F.Q.; Qiu, B.Y.; Li, T.K.; Xie, Q.; Cui, D.J.; Huang, X.L.; Gan, H.T. Tetrandrine suppresses amyloid-beta-induced inflammatory cytokines by inhibiting NF-kappaB pathway in murine BV2 microglial cells. Int. Immunopharmacol. 2011, 11, 1220–1225. [Google Scholar] [CrossRef]
  85. Park, S.E.; Sapkota, K.; Kim, S.; Kim, H.; Kim, S.J. Kaempferol acts through mitogen-activated protein kinases and protein kinase B/AKT to elicit protection in a model of neuroinflammation in BV2 microglial cells. Br. J. Pharm. 2011, 164, 1008–1025. [Google Scholar] [CrossRef] [Green Version]
  86. Velagapudi, R.; Ajileye, O.O.; Okorji, U.; Jain, P.; Aderogba, M.A.; Olajide, O.A. Agathisflavone isolated from Anacardium occidentale suppresses SIRT1-mediated neuroinflammation in BV2 microglia and neurotoxicity in APPSwe-transfected SH-SY5Y cells. Phytother. Res. 2018, 32, 1957–1966. [Google Scholar] [CrossRef] [Green Version]
  87. Rezai-Zadeh, K.; Ehrhart, J.; Bai, Y.; Sanberg, P.R.; Bickford, P.; Tan, J.; Shytle, R.D. Apigenin and luteolin modulate microglial activation via inhibition of STAT1-induced CD40 expression. J. Neuroinflamm. 2008, 5, 41. [Google Scholar] [CrossRef] [Green Version]
  88. Kang, C.H.; Choi, Y.H.; Moon, S.K.; Kim, W.J.; Kim, G.Y. Quercetin inhibits lipopolysaccharide-induced nitric oxide production in BV2 microglial cells by suppressing the NF-kappaB pathway and activating the Nrf2-dependent HO-1 pathway. Int. Immunopharmacol. 2013, 17, 808–813. [Google Scholar] [CrossRef]
  89. Seong, K.J.; Lee, H.G.; Kook, M.S.; Ko, H.M.; Jung, J.Y.; Kim, W.J. Epigallocatechin-3-gallate rescues LPS-impaired adult hippocampal neurogenesis through suppressing the TLR4-NF-kappaB signaling pathway in mice. Korean J. Physiol. Pharm. 2016, 20, 41–51. [Google Scholar] [CrossRef]
  90. Olajide, O.A.; Kumar, A.; Velagapudi, R.; Okorji, U.P.; Fiebich, B.L. Punicalagin inhibits neuroinflammation in LPS-activated rat primary microglia. Mol. Nutr. Food Res. 2014, 58, 1843–1851. [Google Scholar] [CrossRef]
  91. Velagapudi, R.; Lepiarz, I.; El-Bakoush, A.; Katola, F.O.; Bhatia, H.; Fiebich, B.L.; Olajide, O.A. Induction of Autophagy and Activation of SIRT-1 Deacetylation Mechanisms Mediate Neuroprotection by the Pomegranate Metabolite Urolithin A in BV2 Microglia and Differentiated 3D Human Neural Progenitor Cells. Mol. Nutr. Food Res. 2019, 63, e1801237. [Google Scholar] [CrossRef] [Green Version]
  92. Infante-Garcia, C.; Ramos-Rodriguez, J.J.; Delgado-Olmos, I.; Gamero-Carrasco, C.; Fernandez-Ponce, M.T.; Casas, L.; Mantell, C.; Garcia-Alloza, M. Long-Term Mangiferin Extract Treatment Improves Central Pathology and Cognitive Deficits in APP/PS1 Mice. Mol. Neurobiol. 2017, 54, 4696–4704. [Google Scholar] [CrossRef] [PubMed]
  93. Sun, X.Y.; Dong, Q.X.; Zhu, J.; Sun, X.; Zhang, L.F.; Qiu, M.; Yu, X.L.; Liu, R.T. Resveratrol Rescues Tau-Induced Cognitive Deficits and Neuropathology in a Mouse Model of Tauopathy. Curr. Alzheimer Res. 2019, 16, 710–722. [Google Scholar] [CrossRef] [PubMed]
  94. Zhang, J.; Zheng, Y.; Luo, Y.; Du, Y.; Zhang, X.; Fu, J. Curcumin inhibits LPS-induced neuroinflammation by promoting microglial M2 polarization via TREM2/TLR4/NF-kappaB pathways in BV2 cells. Mol. Immunol. 2019, 116, 29–37. [Google Scholar] [CrossRef]
  95. Yu, Y.; Shen, Q.; Lai, Y.; Park, S.Y.; Ou, X.; Lin, D.; Jin, M.; Zhang, W. Anti-inflammatory Effects of Curcumin in Microglial Cells. Front. Pharm. 2018, 9, 386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Wang, J.A.; Tong, M.L.; Zhao, B.; Zhu, G.; Xi, D.H.; Yang, J.P. Parthenolide ameliorates intracerebral hemorrhage-induced brain injury in rats. Phytother. Res. 2020, 34, 153–160. [Google Scholar] [CrossRef]
  97. Qiang, W.; Cai, W.; Yang, Q.; Yang, L.; Dai, Y.; Zhao, Z.; Yin, J.; Li, Y.; Li, Q.; Wang, Y.; et al. Artemisinin B Improves Learning and Memory Impairment in AD Dementia Mice by Suppressing Neuroinflammation. Neuroscience 2018, 395, 1–12. [Google Scholar] [CrossRef]
  98. Abulfadl, Y.S.; El-Maraghy, N.N.; Ahmed, A.E.; Nofal, S.; Abdel-Mottaleb, Y.; Badary, O.A. Thymoquinone alleviates the experimentally induced Alzheimer’s disease inflammation by modulation of TLRs signaling. Hum. Exp. Toxicol. 2018, 37, 1092–1104. [Google Scholar] [CrossRef]
  99. Yang, G.; Wang, Y.; Sun, J.; Zhang, K.; Liu, J. Ginkgo Biloba for Mild Cognitive Impairment and Alzheimer’s Disease: A Systematic Review and Meta-Analysis of Randomized Controlled Trials. Curr. Top. Med. Chem. 2016, 16, 520–528. [Google Scholar] [CrossRef]
  100. Leone, S.; Recinella, L.; Chiavaroli, A.; Orlando, G.; Ferrante, C.; Leporini, L.; Brunetti, L.; Menghini, L. Phytotherapic use of the Crocus sativus L. (Saffron) and its potential applications: A brief overview. Phytother. Res. 2018, 32, 2364–2375. [Google Scholar] [CrossRef]
  101. Mazumder, A.G.; Sharma, P.; Patial, V.; Singh, D. Crocin Attenuates Kindling Development and Associated Cognitive Impairments in Mice via Inhibiting Reactive Oxygen Species-Mediated NF-kappaB Activation. Basic Clin. Pharm. Toxicol. 2017, 120, 426–433. [Google Scholar] [CrossRef]
  102. Choi, S.K.; Park, Y.S.; Choi, D.K.; Chang, H.I. Effects of astaxanthin on the production of NO and the expression of COX-2 and iNOS in LPS-stimulated BV2 microglial cells. J. Microbiol. Biotechnol. 2008, 18, 1990–1996. [Google Scholar] [PubMed]
  103. Kim, J.E.; You, D.J.; Lee, C.; Ahn, C.; Seong, J.Y.; Hwang, J.I. Suppression of NF-kappaB signaling by KEAP1 regulation of IKKbeta activity through autophagic degradation and inhibition of phosphorylation. Cell Signal. 2010, 22, 1645–1654. [Google Scholar] [CrossRef]
  104. Kent, S.A.; Spires-Jones, T.L.; Durrant, C.S. The physiological roles of tau and Ab: Implications for Alzheimer’s disease pathology and therapeutics. Acta Neurophathol. 2020, 140, 417–447. [Google Scholar] [CrossRef] [PubMed]
  105. Fusco, G.; Chen, S.W.; Williamson, P.T.F.; Cascella, R.; Perni, M.; Jarvis, J.A.; Cecchi, C.; Vendruscolo, M.; Chiti, F.; Cremades, N.; et al. Structural basis of membrane disruption and cellular toxicity by alpha-synuclein oligomers. Science 2017, 358, 1440–1443. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Li, S.; Hong, S.; Shepardson, N.E.; Walsh, D.M.; Shankar, G.M.; Selkoe, D. Soluble oligomers of amyloid Beta protein facilitate hippocampal long-term depression by disrupting neuronal glutamate uptake. Neuron 2009, 62, 788–801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Cheignon, C.; Tomas, M.; Bonnefont-Rousselot, D.; Faller, P.; Hureau, C.; Collin, F. Oxidative stress and the amyloid beta peptide in Alzheimer’s disease. Redox Biol. 2018, 14, 450–464. [Google Scholar] [CrossRef]
  108. Salminen, A.; Ojala, J.; Kauppinen, A.; Kaarniranta, K.; Suuronen, T. Inflammation in Alzheimer’s disease: Amyloid-beta oligomers trigger innate immunity defence via pattern recognition receptors. Prog. Neurobiol. 2009, 87, 181–194. [Google Scholar] [CrossRef] [PubMed]
  109. Eckert, A.; Hauptmann, S.; Scherping, I.; Meinhardt, J.; Rhein, V.; Drose, S.; Brandt, U.; Fandrich, M.; Muller, W.E.; Gotz, J. Oligomeric and fibrillar species of beta-amyloid (A beta 42) both impair mitochondrial function in P301L tau transgenic mice. J. Mol. Med. 2008, 86, 1255–1267. [Google Scholar] [CrossRef] [Green Version]
  110. Ferreira, I.L.; Bajouco, L.M.; Mota, S.I.; Auberson, Y.P.; Oliveira, C.R.; Rego, A.C. Amyloid beta peptide 1-42 disturbs intracellular calcium homeostasis through activation of GluN2B-containing N-methyl-d-aspartate receptors in cortical cultures. Cell Calcium 2012, 51, 95–106. [Google Scholar] [CrossRef] [PubMed]
  111. Briley, D.; Ghirardi, V.; Woltjer, R.; Renck, A.; Zolochevska, O.; Taglialatela, G.; Micci, M.A. Preserved neurogenesis in non-demented individuals with AD neuropathology. Sci. Rep. 2016, 6, 27812. [Google Scholar] [CrossRef] [Green Version]
  112. Vergallo, A.; Megret, L.; Lista, S.; Cavedo, E.; Zetterberg, H.; Blennow, K.; Vanmechelen, E.; De Vos, A.; Habert, M.O.; Potier, M.C.; et al. Plasma amyloid beta 40/42 ratio predicts cerebral amyloidosis in cognitively normal individuals at risk for Alzheimer’s disease. Alzheimers Dement. 2019, 15, 764–775. [Google Scholar] [CrossRef] [PubMed]
  113. Satir, T.M.; Agholme, L.; Karlsson, A.; Karlsson, M.; Karila, P.; Illes, S.; Bergstrom, P.; Zetterberg, H. Partial reduction of amyloid beta production by beta-secretase inhibitors does not decrease synaptic transmission. Alzheimers Res. 2020, 12, 63. [Google Scholar] [CrossRef]
  114. Song, G.; Yang, H.; Shen, N.; Pham, P.; Brown, B.; Lin, X.; Hong, Y.; Sinu, P.; Cai, J.; Li, X.; et al. An Immunomodulatory Therapeutic Vaccine Targeting Oligomeric Amyloid-beta. J. Alzheimers Dis. 2020, 77, 1639–1653. [Google Scholar] [CrossRef] [PubMed]
  115. Egan, M.F.; Kost, J.; Tariot, P.N.; Aisen, P.S.; Cummings, J.L.; Vellas, B.; Sur, C.; Mukai, Y.; Voss, T.; Furtek, C.; et al. Randomized Trial of Verubecestat for Mild-to-Moderate Alzheimer’s Disease. N. Engl. J. Med. 2018, 378, 1691–1703. [Google Scholar] [CrossRef] [PubMed]
  116. Mintun, M.A.; Lo, A.C.; Duggan Evans, C.; Wessels, A.M.; Ardayfio, P.A.; Andersen, S.W.; Shcherbinin, S.; Sparks, J.; Sims, J.R.; Brys, M.; et al. Donanemab in Early Alzheimer’s Disease. N. Engl. J. Med. 2021. [Google Scholar] [CrossRef]
  117. Budd Haeberlein, S.; O’Gorman, J.; Chiao, P.; Bussiere, T.; von Rosenstiel, P.; Tian, Y.; Zhu, Y.; von Hehn, C.; Gheuens, S.; Skordos, L.; et al. Clinical Development of Aducanumab, an Anti-Abeta Human Monoclonal Antibody Being Investigated for the Treatment of Early Alzheimer’s Disease. J. Prev. Alzheimers Dis. 2017, 4, 255–263. [Google Scholar] [CrossRef]
  118. Deshpande, P.; Gogia, N.; Singh, A. Exploring the efficacy of natural products in alleviating Alzheimer’s disease. Neural. Regen Res. 2019, 14, 1321–1329. [Google Scholar] [CrossRef]
  119. Lee, J.H.; Ahn, N.H.; Choi, S.B.; Kwon, Y.; Yang, S.H. Natural Products Targeting Amyloid Beta in Alzheimer’s Disease. Int. J. Mol. Sci. 2021, 22, 2341. [Google Scholar] [CrossRef]
  120. Chen, G.F.; Xu, T.H.; Yan, Y.; Zhou, Y.R.; Jiang, Y.; Melcher, K.; Xu, H.E. Amyloid beta: Structure, biology and structure-based therapeutic development. Acta Pharm. Sin. 2017, 38, 1205–1235. [Google Scholar] [CrossRef]
  121. Cole, S.L.; Vassar, R. The Alzheimer’s disease beta-secretase enzyme, BACE1. Mol. Neurodegener. 2007, 2, 22. [Google Scholar] [CrossRef] [Green Version]
  122. Tyler, S.J.; Dawbarn, D.; Wilcock, G.K.; Allen, S.J. alpha- and beta-secretase: Profound changes in Alzheimer’s disease. Biochem. Biophys. Res. Commun. 2002, 299, 373–376. [Google Scholar] [CrossRef]
  123. Li, F.; Wu, X.; Li, J.; Niu, Q. Ginsenoside Rg1 ameliorates hippocampal long-term potentiation and memory in an Alzheimer’s disease model. Mol. Med. Rep. 2016, 13, 4904–4910. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Fang, F.; Chen, X.; Huang, T.; Lue, L.F.; Luddy, J.S.; Yan, S.S. Multi-faced neuroprotective effects of Ginsenoside Rg1 in an Alzheimer mouse model. Biochim. Biophys. Acta 2012, 1822, 286–292. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Zhu, Z.; Li, C.; Wang, X.; Yang, Z.; Chen, J.; Hu, L.; Jiang, H.; Shen, X. 2,2’,4’-trihydroxychalcone from Glycyrrhiza glabra as a new specific BACE1 inhibitor efficiently ameliorates memory impairment in mice. J. Neurochem. 2010, 114, 374–385. [Google Scholar] [CrossRef] [PubMed]
  126. Park, I.H.; Jeon, S.Y.; Lee, H.J.; Kim, S.I.; Song, K.S. A beta-secretase (BACE1) inhibitor hispidin from the mycelial cultures of Phellinus linteus. Planta Med. 2004, 70, 143–146. [Google Scholar] [CrossRef] [PubMed]
  127. Smith, A.; Giunta, B.; Bickford, P.C.; Fountain, M.; Tan, J.; Shytle, R.D. Nanolipidic particles improve the bioavailability and alpha-secretase inducing ability of epigallocatechin-3-gallate (EGCG) for the treatment of Alzheimer’s disease. Int. J. Pharm. 2010, 389, 207–212. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Jeon, S.Y.; Bae, K.; Seong, Y.H.; Song, K.S. Green tea catechins as a BACE1 (beta-secretase) inhibitor. Bioorg. Med. Chem. Lett. 2003, 13, 3905–3908. [Google Scholar] [CrossRef]
  129. Youn, K.; Lee, J.; Ho, C.-T.; Jun, M. Discovery of polymethoxyflavones from black ginger (Kaempferia parviflora) as potential β-secretase (BACE1) inhibitors. J. Funct. Foods 2016, 20, 567–574. [Google Scholar] [CrossRef]
  130. Cai, Z.; Wang, C.; He, W.; Chen, Y. Berberine Alleviates Amyloid-Beta Pathology in the Brain of APP/PS1 Transgenic Mice via Inhibiting beta/gamma-Secretases Activity and Enhancing alpha-Secretases. Curr. Alzheimer Res. 2018, 15, 1045–1052. [Google Scholar] [CrossRef]
  131. Xu, Y.J.; Mei, Y.; Qu, Z.L.; Zhang, S.J.; Zhao, W.; Fang, J.S.; Wu, J.; Yang, C.; Liu, S.J.; Fang, Y.Q.; et al. Ligustilide Ameliorates Memory Deficiency in APP/PS1 Transgenic Mice via Restoring Mitochondrial Dysfunction. Biomed. Res. Int. 2018, 2018, 4606752. [Google Scholar] [CrossRef] [Green Version]
  132. Tamamis, P.; Adler-Abramovich, L.; Reches, M.; Marshall, K.; Sikorski, P.; Serpell, L.; Gazit, E.; Archontis, G. Self-assembly of phenylalanine oligopeptides: Insights from experiments and simulations. Biophys. J. 2009, 96, 5020–5029. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Al-Edresi, S.; Alsalahat, I.; Freeman, S.; Aojula, H.; Penny, J. Resveratrol-mediated cleavage of amyloid beta1-42 peptide: Potential relevance to Alzheimer’s disease. Neurobiol. Aging 2020, 94, 24–33. [Google Scholar] [CrossRef] [PubMed]
  134. Guo, J.; Sun, W.; Liu, F. Brazilin inhibits the Zn(2+)-mediated aggregation of amyloid beta-protein and alleviates cytotoxicity. J. Inorg. Biochem. 2017, 177, 183–189. [Google Scholar] [CrossRef] [PubMed]
  135. Thapa, A.; Jett, S.D.; Chi, E.Y. Curcumin Attenuates Amyloid-beta Aggregate Toxicity and Modulates Amyloid-beta Aggregation Pathway. ACS Chem. Neurosci. 2016, 7, 56–68. [Google Scholar] [CrossRef]
  136. Ono, K.; Hasegawa, K.; Naiki, H.; Yamada, M. Anti-amyloidogenic activity of tannic acid and its activity to destabilize Alzheimer’s beta-amyloid fibrils in vitro. Biochim. Biophys. Acta 2004, 1690, 193–202. [Google Scholar] [CrossRef] [Green Version]
  137. Ladiwala, A.R.; Dordick, J.S.; Tessier, P.M. Aromatic small molecules remodel toxic soluble oligomers of amyloid beta through three independent pathways. J. Biol. Chem. 2011, 286, 3209–3218. [Google Scholar] [CrossRef] [Green Version]
  138. Anandhan, A.; Tamilselvam, K.; Radhiga, T.; Rao, S.; Essa, M.M.; Manivasagam, T. Theaflavin, a black tea polyphenol, protects nigral dopaminergic neurons against chronic MPTP/probenecid induced Parkinson’s disease. Brain Res. 2012, 1433, 104–113. [Google Scholar] [CrossRef]
  139. Tapia-Rojas, C.; Cabezas-Opazo, F.; Deaton, C.A.; Vergara, E.H.; Johnson, G.V.W.; Quintanilla, R.A. It’s all about tau. Prog. Neurobiol. 2019, 175, 54–76. [Google Scholar] [CrossRef]
  140. Kopke, E.; Tung, Y.C.; Shaikh, S.; Alonso, A.C.; Iqbal, K.; Grundke-Iqbal, I. Microtubule-associated protein tau. Abnormal phosphorylation of a non-paired helical filament pool in Alzheimer disease. J. Biol. Chem. 1993, 268, 24374–24384. [Google Scholar] [CrossRef]
  141. Lauretti, E.; Pratico, D. Alzheimer’s disease: Phenotypic approaches using disease models and the targeting of tau protein. Expert Opin. Targets 2020, 24, 319–330. [Google Scholar] [CrossRef]
  142. Min, S.W.; Cho, S.H.; Zhou, Y.; Schroeder, S.; Haroutunian, V.; Seeley, W.W.; Huang, E.J.; Shen, Y.; Masliah, E.; Mukherjee, C.; et al. Acetylation of tau inhibits its degradation and contributes to tauopathy. Neuron 2010, 67, 953–966. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Thomas, S.N.; Funk, K.E.; Wan, Y.; Liao, Z.; Davies, P.; Kuret, J.; Yang, A.J. Dual modification of Alzheimer’s disease PHF-tau protein by lysine methylation and ubiquitylation: A mass spectrometry approach. Acta Neuropathol. 2012, 123, 105–117. [Google Scholar] [CrossRef] [Green Version]
  144. Huseby, C.J.; Hoffman, C.N.; Cooper, G.L.; Cocuron, J.C.; Alonso, A.P.; Thomas, S.N.; Yang, A.J.; Kuret, J. Quantification of Tau Protein Lysine Methylation in Aging and Alzheimer’s Disease. J. Alzheimers Dis. 2019, 71, 979–991. [Google Scholar] [CrossRef]
  145. Sharma, A.; Weber, D.; Raupbach, J.; Dakal, T.C.; Fliessbach, K.; Ramirez, A.; Grune, T.; Wullner, U. Advanced glycation end products and protein carbonyl levels in plasma reveal sex-specific differences in Parkinson’s and Alzheimer’s disease. Redox Biol. 2020, 34, 101546. [Google Scholar] [CrossRef] [PubMed]
  146. Liu, F.; Grundke-Iqbal, I.; Iqbal, K.; Gong, C.X. Contributions of protein phosphatases PP1, PP2A, PP2B and PP5 to the regulation of tau phosphorylation. Eur. J. Neurosci. 2005, 22, 1942–1950. [Google Scholar] [CrossRef] [PubMed]
  147. Hanger, D.P.; Anderton, B.H.; Noble, W. Tau phosphorylation: The therapeutic challenge for neurodegenerative disease. Trends Mol. Med. 2009, 15, 112–119. [Google Scholar] [CrossRef] [PubMed]
  148. Ma, Q.; Ruan, Y.Y.; Xu, H.; Shi, X.M.; Wang, Z.X.; Hu, Y.L. Safflower yellow reduces lipid peroxidation, neuropathology, tau phosphorylation and ameliorates amyloid beta-induced impairment of learning and memory in rats. Biomed. Pharm. 2015, 76, 153–164. [Google Scholar] [CrossRef]
  149. Li, L.; Liu, J.; Yan, X.; Qin, K.; Shi, M.; Lin, T.; Zhu, Y.; Kang, T.; Zhao, G. Protective effects of ginsenoside Rd against okadaic acid-induced neurotoxicity in vivo and in vitro. J. Ethnopharmacol. 2011, 138, 135–141. [Google Scholar] [CrossRef]
  150. Karakani, A.M.; Riazi, G.; Mahmood Ghaffari, S.; Ahmadian, S.; Mokhtari, F.; Jalili Firuzi, M.; Zahra Bathaie, S. Inhibitory effect of corcin on aggregation of 1N/4R human tau protein in vitro. Iran. J. Basic Med. Sci. 2015, 18, 485–492. [Google Scholar]
  151. Feng, J.H.; Cai, B.C.; Guo, W.F.; Wang, M.Y.; Ma, Y.; Lu, Q.X. Neuroprotective effects of Tongmai Yizhi Decoction () against Alzheimer’s disease through attenuating cyclin-dependent kinase-5 expression. Chin. J. Integr Med. 2017, 23, 132–137. [Google Scholar] [CrossRef] [PubMed]
  152. Gao, C.; Liu, Y.; Jiang, Y.; Ding, J.; Li, L. Geniposide ameliorates learning memory deficits, reduces tau phosphorylation and decreases apoptosis via GSK3beta pathway in streptozotocin-induced alzheimer rat model. Brain Pathol. 2014, 24, 261–269. [Google Scholar] [CrossRef] [PubMed]
  153. Gong, H.; He, Z.; Peng, A.; Zhang, X.; Cheng, B.; Sun, Y.; Zheng, L.; Huang, K. Effects of several quinones on insulin aggregation. Sci. Rep. 2014, 4, 5648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Chang, K.H.; Chen, I.C.; Lin, H.Y.; Chen, H.C.; Lin, C.H.; Lin, T.H.; Weng, Y.T.; Chao, C.Y.; Wu, Y.R.; Lin, J.Y.; et al. The aqueous extract of Glycyrrhiza inflata can upregulate unfolded protein response-mediated chaperones to reduce tau misfolding in cell models of Alzheimer’s disease. Drug Des. Devel. 2016, 10, 885–896. [Google Scholar] [CrossRef] [Green Version]
  155. Azimi, A.; Ghaffari, S.M.; Riazi, G.H.; Arab, S.S.; Tavakol, M.M.; Pooyan, S. alpha-Cyperone of Cyperus rotundus is an effective candidate for reduction of inflammation by destabilization of microtubule fibers in brain. J. Ethnopharmacol. 2016, 194, 219–227. [Google Scholar] [CrossRef] [PubMed]
  156. Bijari, N.; Balalaie, S.; Akbari, V.; Golmohammadi, F.; Moradi, S.; Adibi, H.; Khodarahmi, R. Effective suppression of the modified PHF6 peptide/1N4R Tau amyloid aggregation by intact curcumin, not its degradation products: Another evidence for the pigment as preventive/therapeutic "functional food". Int. J. Biol. Macromol 2018, 120, 1009–1022. [Google Scholar] [CrossRef] [PubMed]
  157. Yu, K.C.; Kwan, P.; Cheung, S.K.K.; Ho, A.; Baum, L. Effects of Resveratrol and Morin on Insoluble Tau in Tau Transgenic Mice. Transl. Neurosci. 2018, 9, 54–60. [Google Scholar] [CrossRef] [PubMed]
  158. Viswanathan, G.K.; Shwartz, D.; Losev, Y.; Arad, E.; Shemesh, C.; Pichinuk, E.; Engel, H.; Raveh, A.; Jelinek, R.; Cooper, I.; et al. Purpurin modulates Tau-derived VQIVYK fibrillization and ameliorates Alzheimer’s disease-like symptoms in animal model. Cell Mol. Life Sci. 2020, 77, 2795–2813. [Google Scholar] [CrossRef]
  159. Ghasemzadeh, S.; Riazi, G.H. Inhibition of Tau amyloid fibril formation by folic acid: In-vitro and theoretical studies. Int. J. Biol. Macromol. 2020, 154, 1505–1516. [Google Scholar] [CrossRef]
  160. Shin, S.J.; Park, Y.H.; Jeon, S.G.; Kim, S.; Nam, Y.; Oh, S.M.; Lee, Y.Y.; Moon, M. Red Ginseng Inhibits Tau Aggregation and Promotes Tau Dissociation In Vitro. Oxid Med. Cell Longev. 2020, 2020, 7829842. [Google Scholar] [CrossRef]
  161. Gold, P.E. Acetylcholine modulation of neural systems involved in learning and memory. Neurobiol. Learn. Mem. 2003, 80, 194–210. [Google Scholar] [CrossRef]
  162. Singh, M.; Kaur, M.; Kukreja, H.; Chugh, R.; Silakari, O.; Singh, D. Acetylcholinesterase inhibitors as Alzheimer therapy: From nerve toxins to neuroprotection. Eur. J. Med. Chem. 2013, 70, 165–188. [Google Scholar] [CrossRef] [PubMed]
  163. Eyjolfsdottir, H.; Eriksdotter, M.; Linderoth, B.; Lind, G.; Juliusson, B.; Kusk, P.; Almkvist, O.; Andreasen, N.; Blennow, K.; Ferreira, D.; et al. Targeted delivery of nerve growth factor to the cholinergic basal forebrain of Alzheimer’s disease patients: Application of a second-generation encapsulated cell biodelivery device. Alzheimers Res. 2016, 8, 30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Kaufmann, D.; Kaur Dogra, A.; Tahrani, A.; Herrmann, F.; Wink, M. Extracts from Traditional Chinese Medicinal Plants Inhibit Acetylcholinesterase, a Known Alzheimer’s Disease Target. Molecules 2016, 21, 1161. [Google Scholar] [CrossRef] [Green Version]
  165. Ohba, T.; Yoshino, Y.; Ishisaka, M.; Abe, N.; Tsuruma, K.; Shimazawa, M.; Oyama, M.; Tabira, T.; Hara, H. Japanese Huperzia serrata extract and the constituent, huperzine A, ameliorate the scopolamine-induced cognitive impairment in mice. Biosci. Biotechnol. Biochem. 2015, 79, 1838–1844. [Google Scholar] [CrossRef]
  166. Yang, Y.; Wang, Z.; Wu, J.; Chen, Y. Chemical Constituents of Plants from the Genus Phlegmariurus. Chem. Biodivers. 2016, 13, 269–274. [Google Scholar] [CrossRef]
  167. Geromichalos, G.D.; Lamari, F.N.; Papandreou, M.A.; Trafalis, D.T.; Margarity, M.; Papageorgiou, A.; Sinakos, Z. Saffron as a source of novel acetylcholinesterase inhibitors: Molecular docking and in vitro enzymatic studies. J. Agric. Food Chem. 2012, 60, 6131–6138. [Google Scholar] [CrossRef] [PubMed]
  168. Huang, G.B.; Zhao, T.; Muna, S.S.; Jin, H.M.; Park, J.I.; Jo, K.S.; Lee, B.H.; Chae, S.W.; Kim, S.Y.; Park, S.H.; et al. Therapeutic potential of Gastrodia elata Blume for the treatment of Alzheimer’s disease. Neural Regen Res. 2013, 8, 1061–1070. [Google Scholar] [CrossRef]
  169. Teng, Y.; Zhang, M.Q.; Wang, W.; Liu, L.T.; Zhou, L.M.; Miao, S.K.; Wan, L.H. Compound danshen tablet ameliorated abeta25-35-induced spatial memory impairment in mice via rescuing imbalance between cytokines and neurotrophins. BMC Complement. Altern. Med. 2014, 14, 23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Hou, X.Q.; Zhang, L.; Yang, C.; Rong, C.P.; He, W.Q.; Zhang, C.X.; Li, S.; Su, R.Y.; Chang, X.; Qin, J.H.; et al. Alleviating effects of Bushen-Yizhi formula on ibotenic acid-induced cholinergic impairments in rat. Rejuvenation Res. 2015, 18, 111–127. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Moon, M.; Kim, H.G.; Choi, J.G.; Oh, H.; Lee, P.K.; Ha, S.K.; Kim, S.Y.; Park, Y.; Huh, Y.; Oh, M.S. Corrigendum to ”6-Shogaol, an active constituent of ginger, attenuates neuroinflammation and cognitive deficits in animal models of dementia” [BBRC 449 (2014) 8-13]. Biochem. Biophys. Res. Commun. 2020, 521, 545. [Google Scholar] [CrossRef]
  172. Li, Y.; Xu, J.; Xu, P.; Song, S.; Liu, P.; Chi, T.; Ji, X.; Jin, G.; Qiu, S.; Hou, Y.; et al. Xanthoceras sorbifolia extracts ameliorate dendritic spine deficiency and cognitive decline via upregulation of BDNF expression in a rat model of Alzheimer’s disease. Neurosci. Lett. 2016, 629, 208–214. [Google Scholar] [CrossRef] [PubMed]
  173. Park, H.R.; Kim, J.Y.; Lee, Y.; Chun, H.J.; Choi, Y.W.; Shin, H.K.; Choi, B.T.; Kim, C.M.; Lee, J. PMC-12, a traditional herbal medicine, enhances learning memory and hippocampal neurogenesis in mice. Neurosci. Lett. 2016, 617, 254–263. [Google Scholar] [CrossRef] [PubMed]
  174. Ma, X.; Sun, Z.; Liu, Y.; Jia, Y.; Zhang, B.; Zhang, J. Resveratrol improves cognition and reduces oxidative stress in rats with vascular dementia. Neural Regen Res. 2013, 8, 2050–2059. [Google Scholar] [CrossRef] [PubMed]
  175. Zhao, H.F.; Li, N.; Wang, Q.; Cheng, X.J.; Li, X.M.; Liu, T.T. Resveratrol decreases the insoluble Abeta1-42 level in hippocampus and protects the integrity of the blood-brain barrier in AD rats. Neuroscience 2015, 310, 641–649. [Google Scholar] [CrossRef]
  176. He, X.; Li, Z.; Rizak, J.D.; Wu, S.; Wang, Z.; He, R.; Su, M.; Qin, D.; Wang, J.; Hu, X. Resveratrol Attenuates Formaldehyde Induced Hyperphosphorylation of Tau Protein and Cytotoxicity in N2a Cells. Front. Neurosci. 2016, 10, 598. [Google Scholar] [CrossRef] [Green Version]
  177. Schmatz, R.; Mazzanti, C.M.; Spanevello, R.; Stefanello, N.; Gutierres, J.; Correa, M.; da Rosa, M.M.; Rubin, M.A.; Chitolina Schetinger, M.R.; Morsch, V.M. Resveratrol prevents memory deficits and the increase in acetylcholinesterase activity in streptozotocin-induced diabetic rats. Eur. J. Pharm. 2009, 610, 42–48. [Google Scholar] [CrossRef]
  178. Rahvar, M.; Nikseresht, M.; Shafiee, S.M.; Naghibalhossaini, F.; Rasti, M.; Panjehshahin, M.R.; Owji, A.A. Effect of oral resveratrol on the BDNF gene expression in the hippocampus of the rat brain. Neurochem. Res. 2011, 36, 761–765. [Google Scholar] [CrossRef]
  179. Howes, M.J.; Perry, N.S.; Houghton, P.J. Plants with traditional uses and activities, relevant to the management of Alzheimer’s disease and other cognitive disorders. Phytother. Res. 2003, 17, 1–18. [Google Scholar] [CrossRef]
  180. Ansari, N.; Khodagholi, F. Natural products as promising drug candidates for the treatment of Alzheimer’s disease: Molecular mechanism aspect. Curr. Neuropharmacol. 2013, 11, 414–429. [Google Scholar] [CrossRef] [Green Version]
  181. D’Onofrio, G.; Sancarlo, D.; Ruan, Q.; Yu, Z.; Panza, F.; Daniele, A.; Greco, A.; Seripa, D. Phytochemicals in the Treatment of Alzheimer’s Disease: A Systematic Review. Curr. Drug Targets 2017, 18, 1487–1498. [Google Scholar] [CrossRef]
  182. Moeini, R.; Memariani, Z.; Asadi, F.; Bozorgi, M.; Gorji, N. Pistacia Genus as a Potential Source of Neuroprotective Natural Products. Planta Med. 2019, 85, 1326–1350. [Google Scholar] [CrossRef] [PubMed]
  183. Golchin, L.; Shabani, M.; Harandi, S.; Razavinasab, M. Pistachio supplementation attenuates motor and cognition impairments induced by cisplatin or vincristine in rats. Adv. Biomed. Res. 2015, 4, 92. [Google Scholar] [CrossRef] [PubMed]
  184. Tavakoli, J.; Hajpour Soq, K.; Yousefi, A.; Estakhr, P.; Dalvi, M.; Mousavi Khaneghah, A. Antioxidant activity of Pistacia atlantica var mutica kernel oil and it’s unsaponifiable matters. J. Food Sci. Technol. 2019, 56, 5336–5345. [Google Scholar] [CrossRef] [PubMed]
  185. Abdi Gorabi, S.; Mohammadzadeh, H.; Rostampour, M. The Effects of Ripe Pistachio Hulls Hydroalcoholic Extract and Aerobic Training on Learning and Memory in Streptozotocin-induced Diabetic Male Rats. Basic Clin. Neurosci. 2020, 11, 525–534. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Gholamhoseinian, A.; Moradi, M.N.; Sharifi-Far, F. Screening the methanol extracts of some Iranian plants for acetylcholinesterase inhibitory activity. Res. Pharm. Sci. 2009, 4, 105–112. [Google Scholar] [PubMed]
  187. Ammari, M.; Othman, H.; Hajri, A.; Sakly, M.; Abdelmelek, H. Pistacia lentiscus oil attenuates memory dysfunction and decreases levels of biomarkers of oxidative stress induced by lipopolysaccharide in rats. Brain Res. Bull. 2018, 140, 140–147. [Google Scholar] [CrossRef]
  188. Murray, A.P.; Faraoni, M.B.; Castro, M.J.; Alza, N.P.; Cavallaro, V. Natural AChE Inhibitors from Plants and their Contribution to Alzheimer’s Disease Therapy. Curr. Neuropharmacol. 2013, 11, 388–413. [Google Scholar] [CrossRef] [Green Version]
  189. Adhami, H.R.; Farsam, H.; Krenn, L. Screening of medicinal plants from Iranian traditional medicine for acetylcholinesterase inhibition. Phytother. Res. 2011, 25, 1148–1152. [Google Scholar] [CrossRef]
  190. Pacifico, S.; Piccolella, S.; Marciano, S.; Galasso, S.; Nocera, P.; Piscopo, V.; Fiorentino, A.; Monaco, P. LC-MS/MS profiling of a mastic leaf phenol enriched extract and its effects on H2O2 and Abeta(25–35) oxidative injury in SK-B-NE(C)-2 cells. J. Agric. Food Chem. 2014, 62, 11957–11966. [Google Scholar] [CrossRef]
  191. Quartu, M.; Serra, M.P.; Boi, M.; Pillolla, G.; Melis, T.; Poddighe, L.; Del Fiacco, M.; Falconieri, D.; Carta, G.; Murru, E.; et al. Effect of acute administration of Pistacia lentiscus L. essential oil on rat cerebral cortex following transient bilateral common carotid artery occlusion. Lipids Health Dis. 2012, 11, 8. [Google Scholar] [CrossRef] [Green Version]
  192. Zahoor, M.; Zafar, R.; Rahman, N.U. Isolation and identification of phenolic antioxidants from Pistacia integerrima gall and their anticholine esterase activities. Heliyon 2018, 4, e01007. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Seo, J.S.; Yun, J.H.; Baek, I.S.; Leem, Y.H.; Kang, H.W.; Cho, H.K.; Lyu, Y.S.; Son, H.J.; Han, P.L. Oriental medicine Jangwonhwan reduces Abeta(1-42) level and beta-amyloid deposition in the brain of Tg-APPswe/PS1dE9 mouse model of Alzheimer disease. J. Ethnopharmacol. 2010, 128, 206–212. [Google Scholar] [CrossRef] [PubMed]
  194. Kim, J.; Kim, S.H.; Lee, D.S.; Lee, D.J.; Kim, S.H.; Chung, S.; Yang, H.O. Effects of fermented ginseng on memory impairment and beta-amyloid reduction in Alzheimer’s disease experimental models. J. Ginseng Res. 2013, 37, 100–107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Lee, M.R.; Yun, B.S.; In, O.H.; Sung, C.K. Comparative study of korean white, red, and black ginseng extract on cholinesterase inhibitory activity and cholinergic function. J. Ginseng Res. 2011, 35, 421–428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Choi, J.G.; Kim, N.; Huh, E.; Lee, H.; Oh, M.H.; Park, J.D.; Pyo, M.K.; Oh, M.S. White Ginseng Protects Mouse Hippocampal Cells Against Amyloid-Beta Oligomer Toxicity. Phytother. Res. 2017, 31, 497–506. [Google Scholar] [CrossRef] [PubMed]
  197. Heo, J.H.; Lee, S.T.; Oh, M.J.; Park, H.J.; Shim, J.Y.; Chu, K.; Kim, M. Improvement of cognitive deficit in Alzheimer’s disease patients by long term treatment with korean red ginseng. J. Ginseng Res. 2011, 35, 457–461. [Google Scholar] [CrossRef] [Green Version]
  198. Heo, J.H.; Park, M.H.; Lee, J.H. Effect of Korean Red Ginseng on Cognitive Function and Quantitative EEG in Patients with Alzheimer’s Disease: A Preliminary Study. J. Altern Complement. Med. 2016, 22, 280–285. [Google Scholar] [CrossRef]
  199. Kim, H.J.; Jung, S.W.; Kim, S.Y.; Cho, I.H.; Kim, H.C.; Rhim, H.; Kim, M.; Nah, S.Y. Panax ginseng as an adjuvant treatment for Alzheimer’s disease. J. Ginseng Res. 2018, 42, 401–411. [Google Scholar] [CrossRef]
  200. Uddin, M.S.; Mamun, A.A.; Hossain, M.S.; Ashaduzzaman, M.; Noor, M.A.A.; Hossain, M.S.; Uddin, M.J.; Sarker, J.; Asaduzzaman, M. Neuroprotective effect of Phyllanthus acidus L. on learning and memory impairment in scopolamine-induced animal model of dementia and oxidative stress: Natural wonder for regulating the development and progression of Alzheimer’s disease. Adv. Alzheimers Dis. 2016, 5, 20. [Google Scholar] [CrossRef] [Green Version]
  201. Alagan, A.; Jantan, I.; Kumolosasi, E.; Ogawa, S.; Abdullah, M.A.; Azmi, N. Protective Effects of Phyllanthus amarus Against Lipopolysaccharide-Induced Neuroinflammation and Cognitive Impairment in Rats. Front. Pharm. 2019, 10, 632. [Google Scholar] [CrossRef] [Green Version]
  202. Uddin, M.S.; Mamun, A.A.; Hossain, M.S.; Akter, F.; Iqbal, M.A.; Asaduzzaman, M. Exploring the Effect of Phyllanthus emblica L. on Cognitive Performance, Brain Antioxidant Markers and Acetylcholinesterase Activity in Rats: Promising Natural Gift for the Mitigation of Alzheimer’s Disease. Ann. Neurosci. 2016, 23, 218–229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Singh, S.K.; Srivastav, S.; Castellani, R.J.; Plascencia-Villa, G.; Perry, G. Neuroprotective and Antioxidant Effect of Ginkgo biloba Extract Against AD and Other Neurological Disorders. Neurotherapeutics 2019, 16, 666–674. [Google Scholar] [CrossRef] [PubMed]
  204. Zhao, J.; Li, K.; Wang, Y.; Li, D.; Wang, Q.; Xie, S.; Wang, J.; Zuo, Z. Enhanced anti-amnestic effect of donepezil by Ginkgo biloba extract (EGb 761) via further improvement in pro-cholinergic and antioxidative activities. J. Ethnopharmacol. 2021, 269, 113711. [Google Scholar] [CrossRef] [PubMed]
  205. Liu, H.; Ye, M.; Guo, H. An Updated Review of Randomized Clinical Trials Testing the Improvement of Cognitive Function of Ginkgo biloba Extract in Healthy People and Alzheimer’s Patients. Front. Pharm. 2019, 10, 1688. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Ude, C.; Schubert-Zsilavecz, M.; Wurglics, M. Ginkgo biloba extracts: A review of the pharmacokinetics of the active ingredients. Clin. Pharm. 2013, 52, 727–749. [Google Scholar] [CrossRef]
  207. 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. Pharm. 2020, 128, 110303. [Google Scholar] [CrossRef] [PubMed]
  208. Hashmi, W.J.; Ismail, H.; Mehmood, F.; Mirza, B. Neuroprotective, antidiabetic and antioxidant effect of Hedera nepalensis and lupeol against STZ + AlCl3 induced rats model. Daru 2018, 26, 179–190. [Google Scholar] [CrossRef]
  209. Zhang, X.Z.; Qian, S.S.; Zhang, Y.J.; Wang, R.Q. Salvia miltiorrhiza: A source for anti-Alzheimer’s disease drugs. Pharm. Biol. 2016, 54, 18–24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Yu, H.; Yao, L.; Zhou, H.; Qu, S.; Zeng, X.; Zhou, D.; Zhou, Y.; Li, X.; Liu, Z. Neuroprotection against Abeta25-35-induced apoptosis by Salvia miltiorrhiza extract in SH-SY5Y cells. Neurochem. Int. 2014, 75, 89–95. [Google Scholar] [CrossRef] [PubMed]
  211. Jiang, W.Y.; Jeon, B.H.; Kim, Y.C.; Lee, S.H.; Sohn, D.H.; Seo, G.S. PF2401-SF, standardized fraction of Salvia miltiorrhiza shows anti-inflammatory activity in macrophages and acute arthritis in vivo. Int. Immunopharmacol. 2013, 16, 160–164. [Google Scholar] [CrossRef]
  212. Hu, L.; Yu, J.; Li, F.; Chen, B.; Li, L.; Liu, G. Effects of Salvia miltorrhiza in neural differentiation of rat mesenchymal stem cells with optimized protocol. J. Ethnopharmacol. 2011, 136, 334–340. [Google Scholar] [CrossRef] [PubMed]
  213. Shu, T.; Pang, M.; Rong, L.; Zhou, W.; Wang, J.; Liu, C.; Wang, X. Effects of Salvia miltiorrhiza on neural differentiation of induced pluripotent stem cells. J. Ethnopharmacol. 2014, 153, 233–241. [Google Scholar] [CrossRef]
  214. Liu, Q.F.; Jeon, Y.; Sung, Y.W.; Lee, J.H.; Jeong, H.; Kim, Y.M.; Yun, H.S.; Chin, Y.W.; Jeon, S.; Cho, K.S.; et al. Nardostachys jatamansi Ethanol Extract Ameliorates Abeta42 Cytotoxicity. Biol. Pharm. Bull. 2018, 41, 470–477. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Liu, Q.F.; Lee, J.H.; Kim, Y.M.; Lee, S.; Hong, Y.K.; Hwang, S.; Oh, Y.; Lee, K.; Yun, H.S.; Lee, I.S.; et al. In Vivo Screening of Traditional Medicinal Plants for Neuroprotective Activity against Abeta42 Cytotoxicity by Using Drosophila Models of Alzheimer’s Disease. Biol. Pharm. Bull. 2015, 38, 1891–1901. [Google Scholar] [CrossRef] [Green Version]
  216. Szurpnicka, A.; Kowalczuk, A.; Szterk, A. Biological activity of mistletoe: In vitro and in vivo studies and mechanisms of action. Arch. Pharmacal. Res. 2020, 43, 593–629. [Google Scholar] [CrossRef]
  217. Russo, A.; Borrelli, F. Bacopa monniera, a reputed nootropic plant: An overview. Phytomedicine 2005, 12, 305–317. [Google Scholar] [CrossRef]
  218. Dhanasekaran, M.; Tharakan, B.; Holcomb, L.A.; Hitt, A.R.; Young, K.A.; Manyam, B.V. Neuroprotective mechanisms of ayurvedic antidementia botanical Bacopa monniera. Phytother. Res. 2007, 21, 965–969. [Google Scholar] [CrossRef] [PubMed]
  219. Uabundit, N.; Wattanathorn, J.; Mucimapura, S.; Ingkaninan, K. Cognitive enhancement and neuroprotective effects of Bacopa monnieri in Alzheimer’s disease model. J. Ethnopharmacol. 2010, 127, 26–31. [Google Scholar] [CrossRef] [PubMed]
  220. Bhattacharya, S.K.; Bhattacharya, A.; Kumar, A.; Ghosal, S. Antioxidant activity of Bacopa monniera in rat frontal cortex, striatum and hippocampus. Phytother. Res. 2000, 14, 174–179. [Google Scholar] [CrossRef]
  221. Limpeanchob, N.; Jaipan, S.; Rattanakaruna, S.; Phrompittayarat, W.; Ingkaninan, K. Neuroprotective effect of Bacopa monnieri on beta-amyloid-induced cell death in primary cortical culture. J. Ethnopharmacol. 2008, 120, 112–117. [Google Scholar] [CrossRef]
  222. Sadhu, A.; Upadhyay, P.; Agrawal, A.; Ilango, K.; Karmakar, D.; Singh, G.P.; Dubey, G.P. Management of cognitive determinants in senile dementia of Alzheimer’s type: Therapeutic potential of a novel polyherbal drug product. Clin. Drug Investig. 2014, 34, 857–869. [Google Scholar] [CrossRef]
  223. Malik, J.; Karan, M.; Vasisht, K. Nootropic, anxiolytic and CNS-depressant studies on different plant sources of shankhpushpi. Pharm. Biol. 2011, 49, 1234–1242. [Google Scholar] [CrossRef] [PubMed]
  224. Nahata, A.; Patil, U.K.; Dixit, V.K. Effect of Convulvulus pluricaulis Choisy. on learning behaviour and memory enhancement activity in rodents. Nat. Prod. Res. 2008, 22, 1472–1482. [Google Scholar] [CrossRef]
  225. Bihaqi, S.W.; Singh, A.P.; Tiwari, M. Supplementation of Convolvulus pluricaulis attenuates scopolamine-induced increased tau and amyloid precursor protein (AbetaPP) expression in rat brain. Indian J. Pharm. 2012, 44, 593–598. [Google Scholar] [CrossRef] [PubMed]
  226. Sethiya, N.K.; Nahata, A.; Mishra, S.H.; Dixit, V.K. An update on Shankhpushpi, a cognition-boosting Ayurvedic medicine. Zhong Xi Yi Jie He Xue Bao 2009, 7, 1001–1022. [Google Scholar] [CrossRef] [PubMed]
  227. Shinomol, G.K.; Muralidhara; Bharath, M.M. Exploring the Role of “Brahmi” (Bacopa monnieri and Centella asiatica) in Brain Function and Therapy. Recent Pat. Endocr. Metab. Immune Drug Discov. 2011, 5, 33–49. [Google Scholar] [CrossRef] [Green Version]
  228. Veerendra Kumar, M.H.; Gupta, Y.K. Effect of Centella asiatica on cognition and oxidative stress in an intracerebroventricular streptozotocin model of Alzheimer’s disease in rats. Clin. Exp. Pharm. Physiol. 2003, 30, 336–342. [Google Scholar] [CrossRef] [PubMed]
  229. Chen, C.L.; Tsai, W.H.; Chen, C.J.; Pan, T.M. Centella asiatica extract protects against amyloid beta1-40-induced neurotoxicity in neuronal cells by activating the antioxidative defence system. J. Tradit Complement. Med. 2016, 6, 362–369. [Google Scholar] [CrossRef] [Green Version]
  230. Dhanasekaran, M.; Holcomb, L.A.; Hitt, A.R.; Tharakan, B.; Porter, J.W.; Young, K.A.; Manyam, B.V. Centella asiatica extract selectively decreases amyloid beta levels in hippocampus of Alzheimer’s disease animal model. Phytother. Res. 2009, 23, 14–19. [Google Scholar] [CrossRef]
  231. Bui, T.T.; Nguyen, T.H. Natural product for the treatment of Alzheimer’s disease. J. Basic Clin. Physiol. Pharm. 2017, 28, 413–423. [Google Scholar] [CrossRef]
  232. Hsieh, C.L.; Tang, N.Y.; Chiang, S.Y.; Hsieh, C.T.; Lin, J.G. Anticonvulsive and free radical scavenging actions of two herbs, Uncaria rhynchophylla (MIQ) Jack and Gastrodia elata Bl., in kainic acid-treated rats. Life Sci. 1999, 65, 2071–2082. [Google Scholar] [CrossRef]
  233. Tang, N.Y.; Liu, C.H.; Su, S.Y.; Jan, Y.M.; Hsieh, C.T.; Cheng, C.Y.; Shyu, W.C.; Hsieh, C.L. Uncaria rhynchophylla (miq) Jack plays a role in neuronal protection in kainic acid-treated rats. Am. J. Chin. Med. 2010, 38, 251–263. [Google Scholar] [CrossRef] [PubMed]
  234. Fujiwara, H.; Iwasaki, K.; Furukawa, K.; Seki, T.; He, M.; Maruyama, M.; Tomita, N.; Kudo, Y.; Higuchi, M.; Saido, T.C.; et al. Uncaria rhynchophylla, a Chinese medicinal herb, has potent antiaggregation effects on Alzheimer’s beta-amyloid proteins. J. Neurosci. Res. 2006, 84, 427–433. [Google Scholar] [CrossRef] [PubMed]
  235. Shi, J.S.; Yu, J.X.; Chen, X.P.; Xu, R.X. Pharmacological actions of Uncaria alkaloids, rhynchophylline and isorhynchophylline. Acta Pharm. Sin. 2003, 24, 97–101. [Google Scholar]
  236. Xian, Y.F.; Lin, Z.X.; Mao, Q.Q.; Hu, Z.; Zhao, M.; Che, C.T.; Ip, S.P. Bioassay-Guided Isolation of Neuroprotective Compounds from Uncaria rhynchophylla against Beta-Amyloid-Induced Neurotoxicity. Evid. Based Complement. Altern. Med. 2012, 2012, 802625. [Google Scholar] [CrossRef]
  237. Kang, T.H.; Murakami, Y.; Matsumoto, K.; Takayama, H.; Kitajima, M.; Aimi, N.; Watanabe, H. Rhynchophylline and isorhynchophylline inhibit NMDA receptors expressed in Xenopus oocytes. Eur. J. Pharm. 2002, 455, 27–34. [Google Scholar] [CrossRef]
  238. Chen, C.M.; Weng, Y.T.; Chen, W.L.; Lin, T.H.; Chao, C.Y.; Lin, C.H.; Chen, I.C.; Lee, L.C.; Lin, H.Y.; Wu, Y.R.; et al. Aqueous extract of Glycyrrhiza inflata inhibits aggregation by upregulating PPARGC1A and NFE2L2-ARE pathways in cell models of spinocerebellar ataxia 3. Free Radic. Biol. Med. 2014, 71, 339–350. [Google Scholar] [CrossRef]
  239. Chiu, Y.J.; Lee, C.M.; Lin, T.H.; Lin, H.Y.; Lee, S.Y.; Mesri, M.; Chang, K.H.; Lin, J.Y.; Lee-Chen, G.J.; Chen, C.M. Chinese Herbal Medicine Glycyrrhiza inflataReduces Abeta Aggregation and Exerts Neuroprotection through Anti-Oxidation and Anti-Inflammation. Am. J. Chin. Med. 2018, 1–25. [Google Scholar] [CrossRef]
  240. Qi, Y.; Cheng, X.; Jing, H.; Yan, T.; Xiao, F.; Wu, B.; Bi, K.; Jia, Y. Combination of schisandrin and nootkatone exerts neuroprotective effect in Alzheimer’s disease mice model. Metab. Brain Dis. 2019, 34, 1689–1703. [Google Scholar] [CrossRef] [PubMed]
  241. Penumala, M.; Zinka, R.B.; Shaik, J.B.; Mallepalli, S.K.R.; Vadde, R.; Amooru, D.G. Phytochemical profiling and in vitro screening for anticholinesterase, antioxidant, antiglucosidase and neuroprotective effect of three traditional medicinal plants for Alzheimer’s Disease and Diabetes Mellitus dual therapy. BMC Complement. Altern Med. 2018, 18, 77. [Google Scholar] [CrossRef] [Green Version]
  242. Lim, H.S.; Kim, Y.J.; Sohn, E.; Yoon, J.; Kim, B.Y.; Jeong, S.J. Bojungikgi-Tang, a Traditional Herbal Formula, Exerts Neuroprotective Effects and Ameliorates Memory Impairments in Alzheimer’s Disease-Like Experimental Models. Nutrients 2018, 10, 1952. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Li, X.L.; Wang, D.S.; Zhao, B.Q.; Li, Q.; Qu, H.Y.; Zhang, T.; Zhou, J.P.; Sun, M.J. Effects of Chinese herbal medicine fuzhisan on aged rats. Exp. Gerontol. 2008, 43, 853–858. [Google Scholar] [CrossRef] [PubMed]
  244. Zhao, J.; Wang, D.; Duan, S.; Wang, J.; Bai, J.; Li, W. Analysis of fuzhisan and quantitation of baicalin and ginsenoside Rb(1) by HPLC-DAD-ELSD. Arch. Pharm. Res. 2009, 32, 989–996. [Google Scholar] [CrossRef]
  245. Bi, M.; Tong, S.; Zhang, Z.; Ma, Q.; Zhang, S.; Luo, Z.; Zhang, Y.; Li, X.; Wang, D. Changes in cerebral glucose metabolism in patients with mild-to-moderate Alzheimer’s disease: A pilot study with the Chinese herbal medicine fuzhisan. Neurosci. Lett. 2011, 501, 35–40. [Google Scholar] [CrossRef] [PubMed]
  246. Guo, P.; Wang, D.; Wang, X.; Feng, H.; Tang, Y.; Sun, R.; Zheng, Y.; Dong, L.; Zhao, J.; Zhang, X.; et al. Effect and mechanism of fuzhisan and donepezil on the sirtuin 1 pathway and amyloid precursor protein metabolism in PC12 cells. Mol. Med. Rep. 2016, 13, 3539–3546. [Google Scholar] [CrossRef] [PubMed]
  247. Huang, H.J.; Chen, S.L.; Chang, Y.T.; Chyuan, J.H.; Hsieh-Li, H.M. Administration of Momordica charantia Enhances the Neuroprotection and Reduces the Side Effects of LiCl in the Treatment of Alzheimer’s Disease. Nutrients 2018, 10, 1888. [Google Scholar] [CrossRef] [Green Version]
  248. Sepehri, H.; Hojati, A.; Safari, R. Effect of Bitter Melon on Spatial Memory of Rats Receiving a High-Fat Diet. J. Exp. Pharm. 2019, 11, 115–119. [Google Scholar] [CrossRef] [Green Version]
  249. Veerendra Kumar, M.H.; Gupta, Y.K. Intracerebroventricular administration of colchicine produces cognitive impairment associated with oxidative stress in rats. Pharm. Biochem. Behav. 2002, 73, 565–571. [Google Scholar] [CrossRef]
  250. Mathew, B.; Biju, R. Neuroprotective effects of garlic a review. Libyan J. Med. 2008, 3, 23–33. [Google Scholar] [CrossRef]
  251. Sripanidkulchai, B. Benefits of aged garlic extract on Alzheimer’s disease: Possible mechanisms of action. Exp. Med. 2020, 19, 1560–1564. [Google Scholar] [CrossRef] [Green Version]
  252. Nillert, N.; Pannangrong, W.; Welbat, J.U.; Chaijaroonkhanarak, W.; Sripanidkulchai, K.; Sripanidkulchai, B. Neuroprotective Effects of Aged Garlic Extract on Cognitive Dysfunction and Neuroinflammation Induced by beta-Amyloid in Rats. Nutrients 2017, 9, 24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  253. Hwang, K.A.; Hwang, Y.J.; Hwang, I.G.; Song, J.; Jun Kim, Y. Low temperature-aged garlic extract suppresses psychological stress by modulation of stress hormones and oxidative stress response in brain. J. Chin. Med. Assoc. 2019, 82, 191–195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  254. Khazdair, M.R.; Anaeigoudari, A.; Hashemzehi, M.; Mohebbati, R. Neuroprotective potency of some spice herbs, a literature review. J. Tradit. Complement. Med. 2019, 9, 98–105. [Google Scholar] [CrossRef] [PubMed]
  255. Pulido-Moran, M.; Moreno-Fernandez, J.; Ramirez-Tortosa, C.; Ramirez-Tortosa, M. Curcumin and Health. Molecules 2016, 21, 264. [Google Scholar] [CrossRef] [PubMed]
  256. Hamaguchi, T.; Ono, K.; Yamada, M. REVIEW: Curcumin and Alzheimer’s disease. CNS Neurosci. 2010, 16, 285–297. [Google Scholar] [CrossRef]
  257. Da Costa, I.M.; Freire, M.A.M.; de Paiva Cavalcanti, J.R.L.; de Araujo, D.P.; Norrara, B.; Moreira Rosa, I.M.M.; de Azevedo, E.P.; do Rego, A.C.M.; Filho, I.A.; Guzen, F.P. Supplementation with Curcuma longa Reverses Neurotoxic and Behavioral Damage in Models of Alzheimer’s Disease: A Systematic Review. Curr. Neuropharmacol. 2019, 17, 406–421. [Google Scholar] [CrossRef]
  258. Yuliani, S.; Mustofa; Partadiredja, G. The neuroprotective effects of an ethanolic turmeric (Curcuma longa L.) extract against trimethyltin-induced oxidative stress in rats. Nutr. Neurosci. 2019, 22, 797–804. [Google Scholar] [CrossRef]
  259. Kadri, Y.; Nciri, R.; Brahmi, N.; Saidi, S.; Harrath, A.H.; Alwasel, S.; Aldahmash, W.; El Feki, A.; Allagui, M.S. Protective effects of Curcuma longa against neurobehavioral and neurochemical damage caused by cerium chloride in mice. Env. Sci. Pollut. Res. Int. 2018, 25, 19555–19565. [Google Scholar] [CrossRef]
  260. Ganguli, M.; Chandra, V.; Kamboh, M.I.; Johnston, J.M.; Dodge, H.H.; Thelma, B.K.; Juyal, R.C.; Pandav, R.; Belle, S.H.; DeKosky, S.T. Apolipoprotein E polymorphism and Alzheimer disease: The Indo-US Cross-National Dementia Study. Arch. Neurol. 2000, 57, 824–830. [Google Scholar] [CrossRef] [Green Version]
  261. Ng, T.P.; Chiam, P.C.; Lee, T.; Chua, H.C.; Lim, L.; Kua, E.H. Curry consumption and cognitive function in the elderly. Am. J. Epidemiol. 2006, 164, 898–906. [Google Scholar] [CrossRef] [Green Version]
  262. Hishikawa, N.; Takahashi, Y.; Amakusa, Y.; Tanno, Y.; Tuji, Y.; Niwa, H.; Murakami, N.; Krishna, U.K. Effects of turmeric on Alzheimer’s disease with behavioral and psychological symptoms of dementia. Ayu 2012, 33, 499–504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  263. Oboh, G.; Ademiluyi, A.O.; Akinyemi, A.J. Inhibition of acetylcholinesterase activities and some pro-oxidant induced lipid peroxidation in rat brain by two varieties of ginger (Zingiber officinale). Exp. Toxicol. Pathol. 2012, 64, 315–319. [Google Scholar] [CrossRef] [PubMed]
  264. Ali, B.H.; Blunden, G.; Tanira, M.O.; Nemmar, A. Some phytochemical, pharmacological and toxicological properties of ginger (Zingiber officinale Roscoe): A review of recent research. Food Chem. Toxicol. 2008, 46, 409–420. [Google Scholar] [CrossRef] [PubMed]
  265. Hartman, R.E.; Shah, A.; Fagan, A.M.; Schwetye, K.E.; Parsadanian, M.; Schulman, R.N.; Finn, M.B.; Holtzman, D.M. Pomegranate juice decreases amyloid load and improves behavior in a mouse model of Alzheimer’s disease. Neurobiol. Dis. 2006, 24, 506–515. [Google Scholar] [CrossRef]
  266. Rojanathammanee, L.; Puig, K.L.; Combs, C.K. Pomegranate polyphenols and extract inhibit nuclear factor of activated T-cell activity and microglial activation in vitro and in a transgenic mouse model of Alzheimer disease. J. Nutr. 2013, 143, 597–605. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  267. Subash, S.; Essa, M.M.; Al-Asmi, A.; Al-Adawi, S.; Vaishnav, R.; Braidy, N.; Manivasagam, T.; Guillemin, G.J. Pomegranate from Oman Alleviates the Brain Oxidative Damage in Transgenic Mouse Model of Alzheimer’s disease. J. Tradit Complement. Med. 2014, 4, 232–238. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  268. Yuan, T.; Ma, H.; Liu, W.; Niesen, D.B.; Shah, N.; Crews, R.; Rose, K.N.; Vattem, D.A.; Seeram, N.P. Pomegranate’s Neuroprotective Effects against Alzheimer’s Disease Are Mediated by Urolithins, Its Ellagitannin-Gut Microbial Derived Metabolites. ACS Chem. Neurosci. 2016, 7, 26–33. [Google Scholar] [CrossRef]
  269. Pannangrong, W.; Wattanathorn, J.; Muchimapura, S.; Tiamkao, S.; Tong-Un, T. Purple rice berry is neuroprotective and enhances cognition in a rat model of Alzheimer’s disease. J. Med. Food 2011, 14, 688–694. [Google Scholar] [CrossRef]
  270. Peters, R.; Peters, J.; Warner, J.; Beckett, N.; Bulpitt, C. Alcohol, dementia and cognitive decline in the elderly: A systematic review. Age Ageing 2008, 37, 505–512. [Google Scholar] [CrossRef] [Green Version]
  271. Loureiro, J.A.; Andrade, S.; Duarte, A.; Neves, A.R.; Queiroz, J.F.; Nunes, C.; Sevin, E.; Fenart, L.; Gosselet, F.; Coelho, M.A.; et al. Resveratrol and Grape Extract-loaded Solid Lipid Nanoparticles for the Treatment of Alzheimer’s Disease. Molecules 2017, 22, 277. [Google Scholar] [CrossRef]
  272. Siahmard, Z.; Alaei, H.; Reisi, P.; Pilehvarian, A.A. The effect of red grape juice on Alzheimer’s disease in rats. Adv. Biomed. Res. 2012, 1, 63. [Google Scholar] [CrossRef]
  273. Sun, Q.; Jia, N.; Li, X.; Yang, J.; Chen, G. Grape seed proanthocyanidins ameliorate neuronal oxidative damage by inhibiting GSK-3beta-dependent mitochondrial permeability transition pore opening in an experimental model of sporadic Alzheimer’s disease. Aging (Albany Ny) 2019, 11, 4107–4124. [Google Scholar] [CrossRef] [PubMed]
  274. Lian, Q.; Nie, Y.; Zhang, X.; Tan, B.; Cao, H.; Chen, W.; Gao, W.; Chen, J.; Liang, Z.; Lai, H.; et al. Effects of grape seed proanthocyanidin on Alzheimer’s disease in vitro and in vivo. Exp. Med. 2016, 12, 1681–1692. [Google Scholar] [CrossRef] [Green Version]
  275. Borai, I.H.; Ezz, M.K.; Rizk, M.Z.; Aly, H.F.; El-Sherbiny, M.; Matloub, A.A.; Fouad, G.I. Therapeutic impact of grape leaves polyphenols on certain biochemical and neurological markers in AlCl3-induced Alzheimer’s disease. Biomed. Pharm. 2017, 93, 837–851. [Google Scholar] [CrossRef] [PubMed]
  276. Gorji, N.; Moeini, R.; Memariani, Z. Almond, hazelnut and walnut, three nuts for neuroprotection in Alzheimer’s disease: A neuropharmacological review of their bioactive constituents. Pharm. Res. 2018, 129, 115–127. [Google Scholar] [CrossRef] [PubMed]
  277. Rajaram, S.; Valls-Pedret, C.; Cofan, M.; Sabate, J.; Serra-Mir, M.; Perez-Heras, A.M.; Arechiga, A.; Casaroli-Marano, R.P.; Alforja, S.; Sala-Vila, A.; et al. The Walnuts and Healthy Aging Study (WAHA): Protocol for a Nutritional Intervention Trial with Walnuts on Brain Aging. Front. Aging Neurosci. 2016, 8, 333. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  278. Bahaeddin, Z.; Yans, A.; Khodagholi, F.; Hajimehdipoor, H.; Sahranavard, S. Hazelnut and neuroprotection: Improved memory and hindered anxiety in response to intra-hippocampal Abeta injection. Nutr. Neurosci. 2017, 20, 317–326. [Google Scholar] [CrossRef] [PubMed]
  279. Kulkarni, K.S.; Kasture, S.B.; Mengi, S.A. Efficacy study of Prunus amygdalus (almond) nuts in scopolamine-induced amnesia in rats. Indian J. Pharm. 2010, 42, 168–173. [Google Scholar] [CrossRef] [Green Version]
  280. Haider, S.; Batool, Z.; Haleem, D.J. Nootropic and hypophagic effects following long term intake of almonds (Prunus amygdalus) in rats. Nutr. Hosp. 2012, 27, 2109–2115. [Google Scholar] [CrossRef]
  281. Batool, Z.; Sadir, S.; Liaquat, L.; Tabassum, S.; Madiha, S.; Rafiq, S.; Tariq, S.; Batool, T.S.; Saleem, S.; Naqvi, F.; et al. Repeated administration of almonds increases brain acetylcholine levels and enhances memory function in healthy rats while attenuates memory deficits in animal model of amnesia. Brain Res. Bull. 2016, 120, 63–74. [Google Scholar] [CrossRef]
  282. Batool, Z.; Agha, F.; Ahmad, S.; Liaquat, L.; Tabassum, S.; Khaliq, S.; Anis, L.; Sajid, I.; Emad, S.; Perveen, T.; et al. Attenuation of cadmium-induced decline in spatial, habituation and recognition memory by long-term administration of almond and walnut supplementation: Role of cholinergic function. Pak. J. Pharm. Sci. 2017, 30, 273–279. [Google Scholar] [PubMed]
  283. Pribis, P.; Bailey, R.N.; Russell, A.A.; Kilsby, M.A.; Hernandez, M.; Craig, W.J.; Grajales, T.; Shavlik, D.J.; Sabate, J. Effects of walnut consumption on cognitive performance in young adults. Br. J. Nutr. 2012, 107, 1393–1401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  284. Zou, J.; Cai, P.S.; Xiong, C.M.; Ruan, J.L. Neuroprotective effect of peptides extracted from walnut (Juglans Sigilata Dode) proteins on Abeta25-35-induced memory impairment in mice. J. Huazhong Univ. Sci. Technol. Med. Sci. 2016, 36, 21–30. [Google Scholar] [CrossRef] [PubMed]
  285. Li, W.; Zhao, T.; Zhang, J.; Xu, J.; Sun-Waterhouse, D.; Zhao, M.; Su, G. Effect of walnut protein hydrolysate on scopolamine-induced learning and memory deficits in mice. J. Food Sci. Technol. 2017, 54, 3102–3110. [Google Scholar] [CrossRef]
  286. Chauhan, A.; Chauhan, V. Beneficial Effects of Walnuts on Cognition and Brain Health. Nutrients 2020, 12, 550. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  287. Muthaiyah, B.; Essa, M.M.; Chauhan, V.; Chauhan, A. Protective effects of walnut extract against amyloid beta peptide-induced cell death and oxidative stress in PC12 cells. Neurochem. Res. 2011, 36, 2096–2103. [Google Scholar] [CrossRef] [Green Version]
  288. Martins, M.; Silva, R.; Pinto, M.M.M.; Sousa, E. Marine Natural Products, Multitarget Therapy and Repurposed Agents in Alzheimer’s Disease. Pharmaceuticals 2020, 13, 242. [Google Scholar] [CrossRef]
  289. Alghazwi, M.; Kan, Y.Q.; Zhang, W.; Gai, W.P.; Garson, M.J.; Smid, S. Neuroprotective activities of natural products from marine macroalgae during 1999–2015. J. Appl. Phycol. 2016, 28, 3599–3616. [Google Scholar] [CrossRef]
  290. Ahn, B.R.; Moon, H.E.; Kim, H.R.; Jung, H.A.; Choi, J.S. Neuroprotective effect of edible brown alga Eisenia bicyclis on amyloid beta peptide-induced toxicity in PC12 cells. Arch. Pharm Res. 2012, 35, 1989–1998. [Google Scholar] [CrossRef] [PubMed]
  291. Um, M.Y.; Lim, D.W.; Son, H.J.; Cho, S.; Lee, C. Phlorotannin-rich fraction from Ishige foliacea brown seaweed prevents the scopolamine-induced memory impairment via regulation of ERK-CREB-BDNF pathway. J. Funct. Foods 2018, 40, 110–116. [Google Scholar] [CrossRef]
  292. Singh, R.; Parihar, P.; Singh, M.; Bajguz, A.; Kumar, J.; Singh, S.; Singh, V.P.; Prasad, S.M. Uncovering Potential Applications of Cyanobacteria and Algal Metabolites in Biology, Agriculture and Medicine: Current Status and Future Prospects. Front. Microbiol. 2017, 8, 515. [Google Scholar] [CrossRef] [Green Version]
  293. Mohd Sairazi, N.S.; Sirajudeen, K.N.S. Natural Products and Their Bioactive Compounds: Neuroprotective Potentials against Neurodegenerative Diseases. Evid. Based Complement. Altern. Med. 2020, 2020, 6565396. [Google Scholar] [CrossRef] [PubMed]
  294. Bermejo-Bescos, P.; Pinero-Estrada, E.; Villar del Fresno, A.M. Neuroprotection by Spirulina platensis protean extract and phycocyanin against iron-induced toxicity in SH-SY5Y neuroblastoma cells. Toxicol. Vitr. 2008, 22, 1496–1502. [Google Scholar] [CrossRef]
  295. Chen, J.C.; Liu, K.S.; Yang, T.J.; Hwang, J.H.; Chan, Y.C.; Lee, I.T. Spirulina and C-phycocyanin reduce cytotoxicity and inflammation-related genes expression of microglial cells. Nutr. Neurosci. 2012, 15, 252–256. [Google Scholar] [CrossRef] [PubMed]
  296. Lee, J.C.; Hou, M.F.; Huang, H.W.; Chang, F.R.; Yeh, C.C.; Tang, J.Y.; Chang, H.W. Marine algal natural products with anti-oxidative, anti-inflammatory, and anti-cancer properties. Cancer Cell Int. 2013, 13, 55. [Google Scholar] [CrossRef] [Green Version]
  297. Koh, E.J.; Kim, K.J.; Song, J.H.; Choi, J.; Lee, H.Y.; Kang, D.H.; Heo, H.J.; Lee, B.Y. Spirulina maxima Extract Ameliorates Learning and Memory Impairments via Inhibiting GSK-3beta Phosphorylation Induced by Intracerebroventricular Injection of Amyloid-beta 1-42 in Mice. Int. J. Mol. Sci. 2017, 18, 2401. [Google Scholar] [CrossRef] [Green Version]
  298. Koh, E.J.; Kim, K.J.; Choi, J.; Kang, D.H.; Lee, B.Y. Spirulina maxima extract prevents cell death through BDNF activation against amyloid beta 1-42 (Abeta1-42) induced neurotoxicity in PC12 cells. Neurosci. Lett. 2018, 673, 33–38. [Google Scholar] [CrossRef] [PubMed]
  299. Zhu, B.; Li, Z.; Qian, P.Y.; Herrup, K. Marine bacterial extracts as a new rich source of drugs against Alzheimer’s disease. J. Neurochem. 2020, 152, 493–508. [Google Scholar] [CrossRef] [PubMed]
  300. Ekor, M. The growing use of herbal medicines: Issues relating to adverse reactions and challenges in monitoring safety. Front. Pharm. 2014, 4, 177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  301. Singh, M.; Arseneault, M.; Sanderson, T.; Murthy, V.; Ramassamy, C. Challenges for research on polyphenols from foods in Alzheimer’s disease: Bioavailability, metabolism, and cellular and molecular mechanisms. J. Agric. Food Chem. 2008, 56, 4855–4873. [Google Scholar] [CrossRef]
  302. Zhao, D.; Simon, J.E.; Wu, Q. A critical review on grape polyphenols for neuroprotection: Strategies to enhance bioefficacy. Crit. Rev. Food Sci. Nutr. 2020, 60, 597–625. [Google Scholar] [CrossRef] [PubMed]
  303. Renaud, J.; Martinoli, M.G. Considerations for the Use of Polyphenols as Therapies in Neurodegenerative Diseases. Int. J. Mol. Sci. 2019, 20, 1883. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  304. Rein, M.J.; Renouf, M.; Cruz-Hernandez, C.; Actis-Goretta, L.; Thakkar, S.K.; da Silva Pinto, M. Bioavailability of bioactive food compounds: A challenging journey to bioefficacy. Br. J. Clin. Pharm. 2013, 75, 588–602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  305. Ovais, M.; Zia, N.; Ahmad, I.; Khalil, A.T.; Raza, A.; Ayaz, M.; Sadiq, A.; Ullah, F.; Shinwari, Z.K. Phyto-Therapeutic and Nanomedicinal Approaches to Cure Alzheimer’s Disease: Present Status and Future Opportunities. Front. Aging Neurosci. 2018, 10, 284. [Google Scholar] [CrossRef] [Green Version]
  306. Niu, X.; Chen, J.; Gao, J. Nanocarriers as a powerful vehicle to overcome blood-brain barrier in treating neurodegenerative diseases: Focus on recent advances. Asian J. Pharm. Sci. 2019, 14, 480–496. [Google Scholar] [CrossRef]
Figure 1. Neuroprotective effects from natural products for AD.
Figure 1. Neuroprotective effects from natural products for AD.
Cells 10 01309 g001
Table 1. Status of Neuroprotective Natural Products Extracts and Mixtures.
Table 1. Status of Neuroprotective Natural Products Extracts and Mixtures.
Natural ProductsExtractNeuroprotective Effects FoundResearch StatusReferences
Pistacia veraKernelInhibited cisplatin or vincristine-induced cognitive and motor impairmentsIn vivo, rats[183]
Pistacia lentiscusEssential oilAttenuated lipopolysaccharide-induced memory impairment and decreased AChE activity and oxidative stress markers in brain tissueRats[187]
Pistacia integerrimaGall extractsRadical scavenging and cholinesterase inhibitory activityIn vitro[192]
Pistacia atlanticaEthyl acetate and aqueous extractsAChE inhibitory activityIn vitro[182]
Panax ginsengRoot extractsReduced Aβ formation, inhibited AChE, restored the decreased synaptophysin and ChAT activity, reduced Aβ formation and aggregationIn vitro, mice and clinical trials[193,194,195,196,197,198,199]
Phyllanthus acidusMethanolic extractImprovement of cognitive functions and reduced oxidative stress via elevating the level of brain antioxidant enzymes, as well as reducing lipid peroxidation and AChE activityIn vitro[200]
Phyllanthus amarus, Cynodon dactylonMethanolic ExtractIncreased the levels of superoxide dismutase, catalase, and NADH dehydrogenaseRats[201]
Phyllanthus emblicaEthanolic extractImproved learning, memory, and antioxidant potential, as well as decreased AChE activityMice[202]
Ginkgo biloba L.Leaf extract (EGb)Scavenged free radicals, prevented mitochondrial dysfunction, activated JNK and ERK pathways, and inhibited neuronal apoptosisScopolamine-induced AD rat model, clinical trials[203,204,205]
Hibiscus sabdariffa L.Anthocyanin-enriched extractsPrevented memory impairment through the amelioration of STZ-induced neuroinflammation and amyloidogenesisIn vitro, mice[207]
Hedera nepalensis K.Crude extractIncreased the levels of catalase (CAT) and superoxide dismutase (SOD), while reducing glutathione (GSH) levelsRats[208]
Salvia miltiorrhiza B.Root extractInhibited oxidative stress and the mitochondria-dependent apoptotic pathway. Inhibited iNOS expression and NO production. Induced neuron cell differentiation from rat mesenchymal stem cells. Promote the differentiation potential of iPSCs and enhanced the survival and neural differentiation of transplanted iPSCs-derived neurons.In vitro, rat[210,211,212,213]
Nardostachys jatamansi D.Ethanolic extractInhibited Aβ-induced cell deathIn vitro, Drosophila model[214,215]
Viscum album L.ExtractSignificantly increased BDNF levels in the serum and diminished AlCl3-induced neurotoxicityIn vitro, mice[216]
Bacopa monnieri L.ExtractDecreased cholinergic degeneration and showed cognition-enhancing effects, protect neuronal cells from β-amyloid-induced damages by lowering ROS levels, inhibited AChERats, clinical trials[217,218,219,220,221,222]
Convolvulus pluricaulis C.Ethanolic extractDecreased tau and AβPP expression in the brainRats[223,224,225,226]
Centella asiatica L.Ethanolic extractReduced β-amyloid pathology and oxidative stress in the brain; protected neurons against the neurotoxicity induced by Aβ1–40, decreased ROS production, and activated the antioxidative defense system by increasing the activities of various related enzymes and enhancing levels of glutathione and glutathione disulfideMice[227,228,229,230,231]
Uncaria rhynchophylla M.Root extractShowed free radical scavenging activity and inhibited lipid peroxidation; reduced microglial activation, nNOS, iNOS, and apoptosis; inhibited Aβ fibril formation and also dissemble preformed Aβ fibrilsRats[232,233,234,235,236,237]
Glycyrrhiza inflata B.Aqueous extractReduction in ROS and tau misfolding, potent anti-Aβ aggregation and radical-scavenging activities. Suppressed the production of NO, TNFα, IL-1β, PGE2, and/or Iba1In vitro[154,238,239]
Alpinia oxyphylla-Schisandra chinensis herb pair (ASHP)Ethanolic extractInhibition of the TLR4/NF-kB/NLRP3 pathway; restored the activities of GST, COX-2, SOD, total antioxidant capacity, and iNOS, increased the levels of NO, GSH, and malondialdehydeMice[240]
Buchanania axillaris D., Hemidesmus indicus L. and Rhus mysorensis G.Methanolic extractsInhibit AChE, BuChE, and α- and β-glucosidase, neuroprotective effects against the cell death-induced oxidative stressIn vitro[241]
Coriandrum sativum L., N. jatamansi, Polygonum multiflorum T., Rehmannia glutinosa G., and Sorbus commixta H.Ethanolic extractNeuroprotective effects against Aβ42 neurotoxicityDrosophila AD models[215]
Bojungikgi-tang (BJIGT)Herbal formulaInhibited Ab aggregation, enhanced BACE activity in vivo, and increased antioxidant activity; prevent the aggregation and expression of Aβ peptides, NeuN and BDNF in the hippocampusIn vitro, mice, clinical trials[242]
Fuzhisan (FZS)Herbal complexAnti-apoptosis and anti-Aβ accumulation activity, enhancing ACh levels, and neurotrophic effectsRats, Clinical tirals[243,244,245,246]
Momordica charantia L.Dried and ground fruitReduced gliosis, oligomeric Aβ level, tau hyperphosphorylation, and neuronal loss, Increasing the expression levels of synaptic-related protein and pS9-GSK3bIn vitro, rats[247,248]
Benincasa hispida L.Aqueous extractPrevented SP formation; antioxidant scavenging actions; prevention of dentate granule cell destruction in the hippocampus and by preventing the extracellular SP depositionRats[249]
Allium sativum L.Aged garlic extractReduced the activation of microglia and IL-1 level, minimized the inflammatory response; suppressed psychological stress through modulation of stress hormones and oxidative stress response in the brainRats[250,251,252,253]
Curcuma longa L.Ethanolic extractAttenuate CeCl3-induced oxidative stress, enhanced the activities of antioxidant enzymes, and decreased AChE activityIn vitro, mice, rats, clinical trials[254,255,256,257,258,259,260,261,262]
Zingiber officinale R.Root extractInhibited AChE, inhibited lipid peroxidation; reduced the overstimulation of NMDA receptors and prevented the production of free radicalsRats[263,264]
Punica granatumJuice and extractsCounteraced oxidative stress, reduced brain inflammation, decreased the accumulation of soluble Aβ42, and reduced amyloid deposition in the hippocampusMice[265,266,267,268]
Oryza sativaDietary supplementDecreased hippocampal AChE activity and lipid peroxidation productsRats[269]
Vitis vinifera L.Juice, polyphenolic extractInhibited Aβ aggregation; antioxidative, anti-neuroinflammatory, and anti-amnesic activitiesRats[270,271,272,273,274,275]
Hazelnut (Corylus avellana)KernelEnhanced memory, reduced anxiety, and ameliorated neuroinflammation and apoptosisRats[276,278,279]
Almond (Prunus dulcis)PasteReduced AChE activity, lowered cholesterol and triglyceride levels; improved learning and memory with enhanced brain tryptophan monoamine levels and serotonergic turnoverRats[279,280,281,282]
Walnut (Juglans regia)Defatted proteinAttenuated expression of proinflammatory cytokines, decreased level of AChE, significantly restored levels of antioxidant enzymes, and reduced expression of NF-κBRats[283,284,285,286,287]
Eisenia bicyclisMethanolic extractReduced intracellular ROS production in PC12 cells induced with Aβ25–35In vitro[290]
Ishige foliaceaPhlorotannin-rich fractionReducing brain AChE activity, suppressed oxidative stress, and activated the ERK-BDNF-CREB signaling pathwayMice[291]
Spirulina platensisProtein and aqueous extractsScavenged free radicals and prevented radical-mediated cell death; reduced cytotoxicity and inhibited the expression of inflammation-related genes like COX-2, TNF-α, IL-6, and iNOSIn vitro[293,294,295]
Spirulina maximaEthanolic ExtractDecreased expression levels of hippocampal Aβ1–42, APP, and BACE1 decreased AChE activity, suppressed hippocampal oxidative stress, increased BDNF level, and activated BDNF/PI3K/Akt signaling pathwaysMice[296,297,298]
Thalassospira profundimarisCrude extractPreserved synaptic structure; blocked cell cycle-related neuron deathIn vitro, mice[299]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chen, X.; Drew, J.; Berney, W.; Lei, W. Neuroprotective Natural Products for Alzheimer’s Disease. Cells 2021, 10, 1309. https://doi.org/10.3390/cells10061309

AMA Style

Chen X, Drew J, Berney W, Lei W. Neuroprotective Natural Products for Alzheimer’s Disease. Cells. 2021; 10(6):1309. https://doi.org/10.3390/cells10061309

Chicago/Turabian Style

Chen, Xin, Joshua Drew, Wren Berney, and Wei Lei. 2021. "Neuroprotective Natural Products for Alzheimer’s Disease" Cells 10, no. 6: 1309. https://doi.org/10.3390/cells10061309

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