Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 1519 2023-06-30 04:27:57 |
2 Alzheimer’s disease; saponin; amyloid beta; inflammation; oxidative stress; apoptosis + 1520 word(s) 3039 2023-07-04 08:49:55 | |
3 layout & references + 604 word(s) 3643 2023-07-31 09:17:12 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Zhang, R.; Zeng, M.; Zhang, X.; Zheng, Y.; Lv, N.; Wang, L.; Gan, J.; Li, Y.; Jiang, X.; Yang, L. Mechanism of Saponins in Treating Alzheimer’s Disease. Encyclopedia. Available online: https://encyclopedia.pub/entry/46249 (accessed on 08 August 2024).
Zhang R, Zeng M, Zhang X, Zheng Y, Lv N, Wang L, et al. Mechanism of Saponins in Treating Alzheimer’s Disease. Encyclopedia. Available at: https://encyclopedia.pub/entry/46249. Accessed August 08, 2024.
Zhang, Ruifeng, Miao Zeng, Xiaolu Zhang, Yujia Zheng, Nuan Lv, Luming Wang, Jiali Gan, Yawen Li, Xijuan Jiang, Lin Yang. "Mechanism of Saponins in Treating Alzheimer’s Disease" Encyclopedia, https://encyclopedia.pub/entry/46249 (accessed August 08, 2024).
Zhang, R., Zeng, M., Zhang, X., Zheng, Y., Lv, N., Wang, L., Gan, J., Li, Y., Jiang, X., & Yang, L. (2023, June 30). Mechanism of Saponins in Treating Alzheimer’s Disease. In Encyclopedia. https://encyclopedia.pub/entry/46249
Zhang, Ruifeng, et al. "Mechanism of Saponins in Treating Alzheimer’s Disease." Encyclopedia. Web. 30 June, 2023.
Mechanism of Saponins in Treating Alzheimer’s Disease
Edit

Drug development for Alzheimer’s disease, the leading cause of dementia, has been a long-standing challenge. Saponins, which are steroid or triterpenoid glycosides with various pharmacological activities, have displayed therapeutic potential in treating Alzheimer’s disease. 

Alzheimer’s disease saponin amyloid beta inflammation oxidative stress apoptosis

1. Introduction

Alzheimer’s disease (AD), the leading cause of dementia, is a progressive neurodegenerative disease [1] that is clinically characterized by memory loss, cognitive impairment, and behavioral disturbances [2]. Since its discovery in 1906, AD has emerged as one of the most costly, fatal, and burdensome diseases of this century [3]. The pathogenesis of AD involves various biological processes [4] involving the abnormal deposition of amyloid beta peptide (Aβ) [5], the accumulation of neurofibrillary tangles (NFTs) [6], neuroinflammation [7], neuronal apoptosis [8], neurotransmitter abnormities [9], and oxidative stress [10]. Despite considerable efforts, drug discovery for the treatment of AD has been slow, with only acetylcholinesterase (AChE)/butyrylcholinesterase (BChE) inhibitors [11] such as galantamine, donepezil, tacrine, and rivastigmine currently available as therapies [12]. However, these treatments only delay the onset of symptoms and cannot halt disease progression and are often associated with significant side effects [13]. Therefore, the development of new therapeutic drugs is urgently needed. Saponins, a type of natural compound, have been extensively studied for their various pharmacological properties [14]. Of particular interest is their potential to enhance learning and memory in individuals with AD [15].

Saponins are naturally occurring compounds that are widely distributed in various plants [16], and they can be divided into two major groups based on their chemical structure: triterpenoid saponins and steroidal saponins [17]. Triterpenoid saponins are further subdivided into tetracyclic triterpenes and pentacyclic triterpenes and are mainly found in plants such as Pentaaceae, Leguminosae, Poria, and Platycodonaceae [18]. The main saponin skeletons of the triterpenoid saponins include dammarane, oleanane, ursane, and lupane (Figure 1A–D) [19]. Steroidal saponins, on the other hand, are mainly found in plants such as Dioscoreaceae, Liliaceae, and Scrophulariaceae [20]. The main saponin metaskeletons of steroidal saponins include spirostane, furostane, cholestane, and cardenolide (Figure 1E–H) [21]. Saponins possess multiple bioactivities, such as reduction of amyloid beta (Aβ) deposition [22], inhibition of tau protein phosphorylation [23], antioxidation [24], antiapoptosis [25], and anti-inflammation [26]. These properties make saponins promising therapeutic candidates for AD and other neurological disorders [27]. Meanwhile, the diversity of saponins found in different plants [28] makes them a valuable source of potential drugs for the treatment of AD. 

Figure 1. Representative triterpenoid saponins include (A) Oleanane, (B) Ursane, (C) Lupane, (D) Dammarane and steroidal saponins include (E) Cholestane, (F) Furostane, (G) Spirostane, (H) Cardenolide.

2. Mechanism of Saponins in Treating Alzheimer’s Disease

2.1. Inhibition of Aβ Deposition and Neurotoxicity

Aβ is a peptide that is a major component of the senile plaques found in the brains of people with AD [29]. Its accumulation is highly neurotoxic and is considered a hallmark of AD [30], which can result in impaired cognitive function, including spatial memory [31]. Aβ is generated from the APP and is typically enclosed by microglia and dystrophic synapses that aggregate around neurons [32][33]. APP is a transmembrane protein located in the synapse of neurons, which can be cleaved by both amyloid and nonamyloid pathways [34]. In the nonamyloid pathway, APP is sequentially cleaved by α-secretase (mainly ADAM10) and γ-secretase [35], resulting in P3 peptide (P3), C83 carboxy-terminal fragment (C83), APP intracellular domain (AICD), and soluble amyloid precursor protein-α (sAPPα) with beneficial neurotrophic effects [36]. Conversely, in the amyloid pathway, APP is first cleaved by β-secretase and sAPPβ is secreted [37]. Subsequently, γ-secretase cleaves the C-terminal fragment (C99) of the residual APP and eventually leads to the release of peptides of different lengths [38]. The most prevalent of them are Aβ1-40 and Aβ1-42 [39], which are neurotoxic fragments capable of oligomerization, aggregation, and subsequent plaque formation [40]. In vivo, Aβ is degraded by a variety of proteases, most notably insulin-degrading enzymes (IDE) and neprilysin (NEP) [41]. However, the production or activity of these clearance enzymes may decrease with age, leading to a failure to clear Aβ in a timely manner [42]. This phenomenon has been linked to the reduced activity of the peroxisome proliferator-activated receptor γ (PPARγ) [43], which is a transcription factor that regulates the expression of IDE and Bace1, reduces Aβ production [44], promotes Aβ clearance [45], and exerts neuroprotective effects [46]. An imbalance between Aβ production and clearance is responsible for the accumulation of Aβ in the brain [47]. In order to develop novel treatments for AD, it may be beneficial to inhibit Aβ production, enhance its clearance, or directly combat its neurotoxicity [48].
Natural saponins have shown potential in affecting Aβ metabolism through different pathways [49]. For example, some natural saponins are able to inhibit the formation of Aβ by reducing APP production. Ginsenoside Rg1, a tetracyclic triterpenoid saponin extracted from ginseng [50], has been shown to reduce Aβ deposition in APP/presenilin 1(PS1) double transgenic AD model mice by lowering APP levels [51]. Pharmacokinetic studies have also shown that ginsenoside Rg1 can cross the blood–brain barrier (BBB) and the blood–cerebrospinal fluid barrier (BCSFB), which provides theoretical support for its ability to improve learning and memory abilities [52]. Xanthoceraside, another tetracyclic triterpenoid saponin, extracted from the bark of the Xanthoceras sorbifolia Bunge commonly used in traditional Chinese medicine (TCM) to treat rheumatism [53], has been shown to reduce APP protein levels and Aβ deposition in the cerebral cortex and hippocampus, thereby ameliorating cognitive function dysfunction in an AD mouse model induced by Aβ intracerebroventricular (ICV) injection [54]. In addition, ginsenoside Rh2, another ginseng derivative used in TCM, has been reported to regulate APP expression by reducing cholesterol and lipid raft levels. This ultimately led to an increase in sAPPα levels and a decrease in Aβ concentrations [55].
Most saponins have been shown to inhibit Aβ deposition by modulating the activity of APP-processing enzymes, such as ginsenoside Rg1 [56], RAPO-1-3 [57], onjisaponin B [57], pseudoginsenoside-F11 (PF11) [58], theasaponin E1 [59], anginsenoside (20S)-Rg3 [60], and ginsenoside C-K (CK) [61]. Ginsenoside Rg1 reduces the γ-secretase responsible for Aβ production by attenuating the Aβ-mediated inhibition of cAMP response element-binding protein (CREB) phosphorylation and protein kinase A (PKA) activity [56]. It also upregulates ADAM10 while reducing BACE1 in Wistar rat models of AD induced by ovariectomy (OVX) and D-galactose (D-gal) [62]. In sAPPα-transfected HT22 cells and neuroblastoma (SH-SY5Y) cells, ginsenoside Rg1 has been found to increase the levels of sAPPα and estrogen receptor (ER) α, elevate α-secretase activity, and decrease extracellular release of Aβ. Further studies have shown that these effects are mediated by the upregulation of phosphorylated extracellular signal-regulated kinase 1/2 (pERK1/2) and phosphorylated protein kinase B (pAkt) [63] but can be reversed by ER antagonists and potentially attenuated by inhibitors of protein kinase C (PKC), MAPK, and phosphatidyl 3-kinase (PI3K) [63]. Yuan Zhi (RAPO) is a TCM formulation that is commonly used to treat dementia due to its neuroprotective effects [64]. Its active ingredient, RAPO-1-3, and onjisaponin B, an acyl saponin with a similar constituent to RAPO-1-3, have been found to reduce Aβ production by promoting the degradation of APP protein through interference with the PS1/BACE1 interaction [57]. Another compound, PF11, which is a pentacyclic triterpene abundant in ginseng [65], has also demonstrated efficacy in inhibiting APP amyloidogenic processing. By reducing the expression of BACE1, PF11 reduced Aβ deposition and ameliorated cognitive impairment and synaptic dysfunction in SAMP8 mice [58]. Theasaponin E1, a pentacyclic triterpene extracted from green tea seeds [66], has been shown to reduce the production of Aβ by increasing the activity of NEP and ADAM10 in APP (SweAPP N2a) cells while inhibiting the expression of BACE1, APP, and PS1 [59]. Furthermore, CK, which is produced through the degradation of protopanaxadiol saponins by the gut microbiota [67], has been studied for its neuroprotective properties. Currently, it is primarily obtained by glycosyl hydrolysis of proto-ginsenoside diol-type saponins [68]. In scopolamine-induced ICR mice, CK was found to reduce the expression of BACE1 and PS1, increase IDE activity, reduce Aβ expression, and improve memory function [61]. In addition, ginsenoside (20S)-Rg3, a component of heat-processed ginseng, has been found to reduce Aβ levels by increasing phosphatidylinositol 4-kinase IIα(PI4KIIα) activity and ultimately decrease the expression of γ-secretase by decreasing the association of PS1 fragments and lipid rafts in cultured primary neurons and in the brains of an AD mouse model [60].
Aside from inhibiting Aβ production, certain saponins have been found to promote the clearance of Aβ. For example, minor ginsenoside F1, a trace ginsenoside derived from Panax ginseng [69], has been shown to effectively reduce Aβ plaques and alleviate cognitive impairment in APP/PS1 mice by enhancing the expression of pCREB [70]. Similarly, Bacopaside I (BS-I), a tetracyclic triterpene and standardized extract of Bacopa monnieri [71], has been shown to be neuroprotective and improve cognitive function [72]. The aglycones of Bacopa monnieri and its derivatives have good intestinal absorption and BBB permeability [73]. Recent studies suggest that BS-I induces sufficient innate immune stimulation and phagocytosis to promote amyloid clearance, thereby reducing amyloid plaque burden in APP/PS1 mice [74]. In addition, Aβ deposition in the brain is linked to the dysfunction of the endosomal–lysosomal system dysfunction [75], which is regulated by the transcription factor EB (TFEB) [76]. PF11 has been observed to increase Aβ clearance by promoting the mammalian target of rapamycin (mTOR)-dependent TFEB-mediated lysosome biogenesis and alleviating endosomal–lysosomal system dysfunction through the conversion of Rab5 to Rab7 [77].

2.2. Inhibiting Aberrant Tau Protein Phosphorylation

NFTs, a well-known pathological hallmark of AD [78], are composed of excessively phosphorylated proteins linked by helical filaments known as tubulin-associated unit (tau) protein [79]. Tau protein, mainly found in neuronal axons, is a microtubule-associated protein involved in various functions such as axonal microtubule maintenance, cytoplasmic transport, and regulation of neuronal signaling [80][81]. Its optimal function depends on the low levels of normal phosphorylation [82]. However, overphosphorylation of tau protein leads to its detachment from microtubules, resulting in microtubule depolymerization, aggregation into NFTs, and ultimately neurodegeneration and cell death [83]. The correlation between NFT density and cognitive decline suggests that tau protein plays a central role in AD [84]. Furthermore, the prevalence of tau protein pathology positively correlates with cognitive impairment in AD, where a person with AD has increased levels of tau protein, hyperphosphorylated tau protein, and aggregated tau in their brain [85]. Abnormal tau protein phosphorylation plays a greater role in AD than the total tau protein [86]. As such, it is considered a critical pathogenic step in the development of NFTs and, subsequently, AD [87].
Several saponins have been found to inhibit the hyperphosphorylation of tau proteins, protect synaptic structures, and reduce neuronal damage. An octillol-type saponin, PF11, has been found to increase the activity of phosphatase-2A (PP2A) by enhancing leucine carboxyl methyltransferase-1 (LCMT-1) and improving cognitive impairment in SAMP/8 mice by inhibiting the hyperphosphorylation of tau protein at serine 396 and tyrosine 205 in the brain [58]. Streptozotocin (STZ) is a glucosamine–nitrosourea compound commonly used in AD studies to induce tau proteins’ hyperphosphorylation and cognitive impairment [88]. Tau protein phosphorylation occurs at multiple sites, usually associated with glycogen synthase kinase (GSK) 3 and cyclin-dependent kinase 5 (CDK5). PF11 was also shown to attenuate STZ-induced tau hyperphosphorylation in the brain of male Wistar rats by correcting the dysregulation of the insulin receptor substrate 1/PI3K/Akt/GSK-3β and calpain I/CDK5 signaling pathways [65]. Additionally, PF11 ameliorated STZ-induced learning and memory deficits, synaptic damage, and neuronal death [65]. Another active compound, Ginsenoside Rb1, found in ginseng [89], exhibits neuroprotective properties against different neurotoxins [90]. Research has shown that ginsenoside Rb1 counters cognitive impairment induced by aluminum oxide in ICR mice while reducing Ser396 tau phosphorylation by increasing PP2A levels and decreasing p-GSK levels [85][91]. Ginsenoside Rd, another compound found in ginseng [92], has been shown to have potential therapeutic applications in neurodegenerative diseases [93]. It has been demonstrated to reduce tau proteins’ hyperphosphorylation and neurotoxicity induced by okadaic acid (OA) by increasing PP2A activity in both rat and cortical neuronal models of AD [94].
Theasaponin E1, a vital compound found in green tea seeds, shows remarkable potential in reducing p-tau protein levels and inhibiting AD-promoting genes while activating AD-remitting genes [95]. This potential has been observed in AD models constructed from SHY5Y and glioblastoma (HTB2) cells. Notably, theasaponin E1 successfully reduced p-tau protein by inhibiting several enzymes, including GSK-3β, CDK5, c-Jun NH2-terminal kinase (JNK), MAPK, ERK1/MARK, and calmodulin-dependent protein kinase II alpha [95]. Similarly, Xanthoceraside, a triterpene saponin extracted from the shell of Xanthoceras sorbifolia Bunge, has been shown to exert neuroprotective effects by inhibiting the phosphorylation of tau protein [96]. In a study conducted on mice with AD, the administration of Xanthoceraside significantly decreased the expression of pGSK-3β and tau protein at Ser396 and Ser404, thereby effectively inhibiting the phosphorylation of tau protein [54].

2.3. Anti-Inflammatory Effect

In recent years, neuroinflammation has become recognized as a third neuropathological hallmark of AD, in addition to Aβ and NFTs [97]. Neuroinflammation is an inflammatory response to pathological insults that occur within the central nervous system (CNS), such as infection, trauma, ischemia, and toxin accumulation [98]. This response is characterized by the proliferation and activation of microglia and astrocytes in the brain [99]. Such activation is usually accompanied by the secretion of inflammatory cytokines such as interleukin-1β (IL-1β), interleukin 6 (IL-6), and tumor necrosis factor α (TNF-α) [100]. Microglia are innate immune cells that reside within the CNS and are thought to be important regulators in the inflammatory response within the brain [101]. In the early stages of AD, activated microglia can exert protective effects by phagocytosing and degrading Aβ [102][103]. However, when Aβ accumulates excessively, it binds various receptors on the surface of microglia such as cluster of differentiation (CD) 14, CD36 [104], Toll-like receptor (TLR) 4, and TLR2, to promote the release of inflammatory factors [105]. This response leads to neuroinflammation and further impairment of cognitive function. Similarly, astrocyte activation is also associated with Aβ [106]. Activated astrocytes produce various inflammatory mediators [107] and secrete small amounts of Aβ [108], resulting in a chronic vicious cycle [107]. These reactions can exacerbate neuroinflammation and damage neurons.
Multiple pathways are involved in microglia- and astrocyte-mediated neuroinflammation [109], one of which is the nuclear factor kappa-B (NF-κB) pathway [110]. This pathway activates the transcription of TNF-α, IL-1β, and IL-6, which exacerbates inflammation [111]. Microglial activation also depends on other signaling pathways such as the PI3K/Akt/mTOR pathway [112] which can promote the expression of the inflammatory mediators NO synthase (NOS) and cyclooxygenase-2 (Cox-2) [113][114]. Furthermore, MAPK/p38 [115], Janus kinase (JAK)/signal transducer, and activator of transcription (STAT) and other pathways also contribute to neuroinflammation [116]. Therefore, modulating inflammatory signaling pathways and inhibiting the production of proinflammatory factors may offer a promising approach to the treatment of AD.

2.4. Improvement in Mitochondrial Function and Antioxidative Stress

Oxidative stress refers to an imbalance between oxidation and antioxidation in the body [117], which is mainly characterized by the increased production of reactive oxygen species and reduced ability of antioxidants to combat them [118]. The brain is susceptible to oxidative damage due to its high lipid content and lacks effective antioxidant defense mechanisms, despite high oxygen consumption [119][120]. It is now well established that the level of oxidation in the brain also increases with age [121]. Interestingly, extensive oxidative damage can be observed in the stage of mild cognitive impairment that precedes the typical clinical manifestations of AD, suggesting that oxidative stress may be a central mechanism in driving the disease [122].

Several natural saponins may act as exogenous antioxidants to combat AD by managing oxidative stress [123][124]. Ginsenoside Rh2, a rare ginsenoside with few reports on its neuroprotective effects compared to other similar molecules [125], has shown significant neuroprotective effects in a scopolamine-induced model of memory impairment in ICR mice. Ginsenoside Rh2 effectively managed MDA levels while increasing glutathione (GSH) levels and SOD activity in the brain, effectively inhibiting oxidative stress [126]. These findings suggest that ginsenoside Rh2 has great potential for the treatment of AD by inhibiting oxidative stress in the brain. Crude saponins from BF showed potent antioxidant effects in addition to anti-inflammatory effects. In addition, saponins from BF alleviated oxidative stress, reduced ROS, and MDA levels by restoring GSH-Px, GSH, and CAT levels in the Aβ ICV-induced BALB/c AD mouse model [127]. Similarly, ginsenoside Rd was found to reduce protein carbonyls and 4-hydroxy-2-nonenal (4-HNE) levels by decreasing the glutathione disulfide (GSSG)/GSH ratio, thereby counteracting oxidative stress in the brain of Aβ-induced AD mouse models [128]. Ginsenoside Rg3, a tetracyclic triterpenoid saponin, is the major bioactive component of ginseng and is widely used for the treatment of tumors [129]. Results from in vivo studies have shown that ginsenoside Rg3 can improve mitochondrial dysfunction and enhance ROS scavenging ability by increasing SOD, CAT, and GSH-Px levels, thus ameliorating oxidative stress in the brain of D-gal-induced AD rats [130]. PF11 was found to inhibit the production of both APP and Aβ in the cortex and hippocampus, while restoring the activity of SOD and GSH-Px in the brains of AD mice and reducing MDA levels in the cortex [131].
Ginsenoside Rg1 has also been found to protect primary cortical neurons against Aβ-induced increases in ROS and H2O2 levels by inhibiting oxidative stress [132]. This results in a decrease in mitochondrial membrane potential and an increase in cytochrome C (Cyt-c) oxidase activity, ultimately leading to a decrease in Cyt-c release and indicating improved mitochondrial function [132]. Another newly discovered saponin compound, huangjingsterol B from Polygonatum cyrtonema Hua, showed significant antioxidant activity [133] by reducing oxidative stress and protecting cells from Aβ-induced damage through modulating the activity of GSH-Px and SOD [133]. Similarly, researchers have found that a saponin from the fruit of Solanum anguivi can protect the P2 region of rat brain synaptosomes from oxidative damage induced by ferrous iron and sodium nitroprusside, reduce ROS generation, restore total thiol levels, and protect mitochondrial function [123]. This suggests that saponins from Solanum anguivi fruits may have antioxidant properties that benefit mitochondrial function [134]. PGS has a similar effect with its ability to upregulate intracellular heme oxygenase-1 (HO-1), SOD, CAT, and GSH-Px to inhibit Aβ-induced ROS production and reduce oxidative damage in AD brains [135]. Despite the demonstrated neuroprotective effects of Panax ginseng, the mechanisms underlying its antioxidant properties remain poorly understood. However, a recent study has shown that the primary bioactive compound in Panax ginseng, NTR1, is able to prevent the loss of mitochondrial membrane potential and reduce the accumulation of ROS induced by Aβ [136]. This suggests that NTR1 may exert neuroprotective effects by inhibiting oxidative stress. Furthermore, most previous experiments have focused on saponins extracted from Panax ginseng roots, and the pharmacological effects of other parts, such as saponins in P. notoginseng leaves (LPNS), are largely unknown [137]. LPNS, which mainly include the ginsenosides Rb1, Rb2, Rb3, and Rc, have been shown to have potent neuroprotective properties [137]. In H2O2- or oxygen and glucose deprivation (OGD)-induced astrocytes from the cortex of SD rats and SH-SY5Y cells, LPNS increased the levels of nuclear factor erythroid2-related factor 2 (Nrf2) and its downstream HO-1 and glutathione S-transferase pi 1 (GSTP1) in response to increased ROS, as well as significantly reducing lactate dehydrogenase (LDH) expression and increasing the cellular activity of astrocytes [137].

2.5. Antiapoptotic Effect

It is widely accepted that neuronal apoptosis contributes significantly to the pathogenesis of AD [138] and is even a pathological hallmark of the disease [139]. Neuronal apoptosis has been found to be abundant in AD autopsy samples [140] and several AD models have successfully reproduced this apoptosis [141]. The biochemical signature of apoptosis is mainly the activation of cysteinase triggered by intrinsic or extrinsic apoptotic pathways [142]. Apoptosis triggered by external cellular signals is called the “death receptor“ pathway, which involves the activation of caspase-8 followed by cleavage of caspase-3, leading to apoptosis [143]. In contrast, the intrinsic apoptotic pathway is driven by the mitochondria [144]. Mitochondrial damage induces the release of Cyt-c and other apoptotic factors into the cytoplasm, which then activate caspase-9 and caspase-3, resulting in apoptosis [145]. The B-cell lymphoma-2 (Bcl-2) protein family includes proapoptotic proteins, BCL2-associated X (BAX), Bcl-2 antagonist killer 1 (BAK), Bcl-2-like protein 11 (BIM), p53 upregulated modulator of apoptosis (PUMA), and Bcl2 modifying factor (BMF), and antiapoptotic proteins, such as Bcl-2, lymphoma-extra-large (Bcl-xL), Bcl-2 like 2 (Bcl-w), and myeloid cell leukemia-1 (Mcl-1), which tightly regulates this apoptotic process by controlling the ratio of proapoptotic and antiapoptotic proteins [146][147], particularly the Bax/Bcl-2 ratio [148]. In AD patients, Bcl-2 expression is reduced in the hippocampus [149], while Bax accumulates near senile plaques and tau protein fiber tangles in an AD animal model, suggesting that mitochondria-mediated apoptosis is closely linked to the pathology of AD [150].
Numerous saponins, including astragaloside (AST), NTR2, PGS, AS-IV, NTR1, Ginsenoside Rg2, and Ginsenoside Rg1, have shown antiapoptotic effects in AD-associated models. AST, a small molecule saponin derived from Astragalus membranaceus, has been widely used in China for the treatment of chronic diseases [151]. In in vivo and in vitro experiments, AST can prevent Aβ-induced apoptosis in cortical neurons by modulating the PI3K/Akt and MAPK (or ERK) pathway [152]. This effect is reversed by the PI3K inhibitor LY294002 and mimicked by the inhibition of ERK and enhanced by its inhibitor U0126, suggesting that the neuroprotective effect of AST may also be related to the inhibition of the ERK pathways [152]. Similarly, NTR2 can reverse the Aβ-induced apoptosis of primary cortical neurons via miR-27a/SOX8/axis and ameliorate cognitive dysfunction in AD mice [153]. PGS may provide neuroprotection not only by reducing inflammation but also by inhibiting apoptosis. PGS increased the expression of Bcl-2 family proteins, resulting in a reduction in Cyt-c release and caspase-9 and -3 expression, thereby inhibiting Aβ-induced apoptosis in HT22 cells. PGS achieved this by inhibiting p38, ERK, and JNK activation, as evidenced by decreased levels of p-p38/p38, p-ERK/ERK, and p-JNK/JNK, and by downregulating MAPK signaling [135]. In vitro and in vivo experiments constructed in HT-22 cells and C57BL/6 induced by amyloid beta protein fragment 1-42 oligomers (AβO) have shown that AS-IV reduced the Aβ-induced inhibition of the brain-derived neurotrophic factor (BDNF)–tyrosine kinase receptor B (TrkB) signaling pathway by regulating PPARγ [154]. This mechanism leads to the reduction in the number and fluorescence density of terminal deoxynucleotidyl transferase dUTP nick end labeling (TUNEL)-positive cells and caspase-3 levels, ultimately alleviating hippocampal neuronal apoptosis and memory impairment [154]. In Aβ-induced PC12 cells, NTR1 has also been shown to exert an antiapoptotic effect, reducing the percentage of TUNEL-positive cells and the levels of caspase-3, Bax, and the Bax/Bcl-2 ratio, and exerted an antiapoptotic effect by suppressing the MAPK pathway [136].

References

  1. Vossel, K.; Ranasinghe, K.G.; Beagle, A.J.; La, A.; Ah, P.K.; Castro, M.; Mizuiri, D.; Honma, S.M.; Venkateswaran, N.; Koestler, M.; et al. Effect of Levetiracetam on Cognition in Patients with Alzheimer Disease With and Without Epileptiform Activity: A Randomized Clinical Trial. JAMA Neurol. 2021, 78, 1345–1354.
  2. Balupuri, A.; Choi, K.E.; Kang, N.S. Aggregation Mechanism of Alzheimer’s Amyloid β-Peptide Mediated by α-Strand/α-Sheet Structure. Int. J. Mol. Sci. 2020, 21, 1094.
  3. Grimm, A.; Biliouris, E.E.; Lang, U.E.; Götz, J.; Mensah-Nyagan, A.G.; Eckert, A. Sex hormone-related neurosteroids differentially rescue bioenergetic deficits induced by amyloid-β or hyperphosphorylated tau protein. Cell. Mol. Life Sci. 2016, 73, 201–215.
  4. Zheng, H.; Fridkin, M.; Youdim, M. From single target to multitarget/network therapeutics in Alzheimer’s therapy. Pharmaceuticals 2014, 7, 113–135.
  5. Tamayev, R.; Matsuda, S.; Arancio, O.; D’Adamio, L. β- but not γ-secretase proteolysis of APP causes synaptic and memory deficits in a mouse model of dementia. EMBO Mol. Med. 2012, 4, 171–179.
  6. Chandra, S.; Roy, A.; Jana, M.; Pahan, K. Cinnamic acid activates PPARα to stimulate Lysosomal biogenesis and lower Amyloid plaque pathology in an Alzheimer’s disease mouse model. Neurobiol. Dis. 2019, 124, 379–395.
  7. Polis, B.; Srikanth, K.D.; Elliott, E.; Gil-Henn, H.; Samson, A.O. L-Norvaline Reverses Cognitive Decline and Synaptic Loss in a Murine Model of Alzheimer’s Disease. Neurotherapeutics 2018, 15, 1036–1054.
  8. Steffen, J.; Krohn, M.; Schwitlick, C.; Brüning, T.; Paarmann, K.; Pietrzik, C.U.; Biverstål, H.; Jansone, B.; Langer, O.; Pahnke, J. Expression of endogenous mouse APP modulates β-amyloid deposition in hAPP-transgenic mice. Acta Neuropathol. Commun. 2017, 20, 49.
  9. Govindpani, K.; Turner, C.; Waldvogel, H.J.; Faull, R.L.M.; Kwakowsky, A. Impaired Expression of GABA Signaling Components in the Alzheimer’s Disease Middle Temporal Gyrus. Int. J. Mol. Sci. 2020, 21, 8704.
  10. Ruthirakuhan, M.; Herrmann, N.; Vieira, D.; Gallagher, D.; Lanctot, K.L. The Roles of Apathy and Depression in Predicting Alzheimer Disease: A Longitudinal Analysis in Older Adults With Mild Cognitive Impairment. Am. J. Geriatr. Psychiatry 2019, 27, 873–882.
  11. Mphahlele, M.J.; Gildenhuys, S.; Agbo, E.N. In Vitro Evaluation and Docking Studies of 5-oxo-5H-furo chromene-6-carbaldehyde Derivatives as Potential Anti-Alzheimer’s Agents. Int. J. Mol. Sci. 2019, 20, 5451.
  12. Saleem, F.; Mehmood, R.; Mehar, S.; Khan, M.T.J.; Khan, Z.U.; Ashraf, M.; Ali, M.S.; Abdullah, I.; Froeyen, M.; Mirza, M.U.; et al. Bioassay Directed Isolation, Biological Evaluation and in Silico Studies of New Isolates from Pteris cretica L. Antioxidants 2019, 8, 231.
  13. Lee, B.D.; Yoo, J.M.; Baek, S.Y.; Li, F.Y.; Sok, D.E.; Kim, M.R. 3,3’-Diindolylmethane Promotes BDNF and Antioxidant Enzyme Formation via TrkB/Akt Pathway Activation for Neuroprotection against Oxidative Stress-Induced Apoptosis in Hippocampal Neuronal Cells. Antioxidants 2019, 9, 3.
  14. Chen, C.; Zhu, H.; Kang, J.; Warusawitharana, H.K.; Chen, S.; Wang, K.; Yu, F.; Wu, Y.; He, P.; Tu, Y.; et al. Comparative Transcriptome and Phytochemical Analysis Provides Insight into Triterpene Saponin Biosynthesis in Seeds and Flowers of the Tea Plant (Camellia sinensis). Metabolites 2022, 12, 204.
  15. Yoon, S.; Iqbal, H.; Kim, S.M.; Jin, M. Phytochemicals That Act on Synaptic Plasticity as Potential Prophylaxis against Stress-Induced Depressive Disorder. Biomol. Ther. 2023, 31, 148–160.
  16. Guclu-Ustundag, O.; Mazza, G. Saponins: Properties, applications and processing. Crit. Rev. Food Sci. Nutr. 2007, 47, 231–258.
  17. Ramos-Morales, E.; de la Fuente, G.; Duval, S.; Wehrli, C.; Bouillon, M.; Lahmann, M.; Preskett, D.; Braganca, R.; Newbold, C.J. Antiprotozoal Effect of Saponins in the Rumen Can Be Enhanced by Chemical Modifications in Their Structure. Front. Microbiol. 2017, 8, 399.
  18. Ghiulai, R.; Roşca, O.J.; Antal, D.S.; Mioc, M.; Mioc, A.; Racoviceanu, R.; Macaşoi, I.; Olariu, T.; Dehelean, C.; Creţu, O.M.; et al. Tetracyclic and Pentacyclic Triterpenes with High Therapeutic Efficiency in Wound Healing Approaches. Molecules. 2020, 25, 5557.
  19. Dashbaldan, S.; Pączkowski, C.; Szakiel, A. Variations in Triterpenoid Deposition in Cuticular Waxes during Development and Maturation of Selected Fruits of Rosaceae Family. Int. J. Mol. Sci. 2020, 21, 9762.
  20. Massana-Codina, J.; Schnee, S.; Allard, P.M.; Rutz, A.; Boccard, J.; Michellod, E.; Cleroux, M.; Schurch, S.; Gindro, K.; Wolfender, J.L. Insights on the Structural and Metabolic Resistance of Potato (Solanum tuberosum) Cultivars to Tuber Black Dot (Colletotrichum coccodes). Front. Plant Sci. 2020, 11, 1287.
  21. Sy, L.K.; Lok, C.N.; Wang, J.Y.; Liu, Y.; Cheng, L.; Wan, P.K.; Leung, C.T.; Cao, B.; Kwong, W.L.; Chang, R.C.; et al. Identification of “sarsasapogenin-aglyconed” timosaponins as novel Aβ-lowering modulators of amyloid precursor protein processing. Chem. Sci. 2016, 7, 3206–3214.
  22. Liu, Y.W.; Zhu, X.; Lu, Q.; Wang, J.Y.; Li, W.; Wei, Y.Q.; Yin, X.X. Total saponins from Rhizoma Anemarrhenae ameliorate diabetes-associated cognitive decline in rats: Involvement of amyloid-beta decrease in brain. J. Ethnopharmacol. 2012, 139, 194–200.
  23. Abduljawad, A.A.; Elawad, M.A.; Elkhalifa, M.; Ahmed, A.; Hamdoon, A.; Salim, L.; Ashraf, M.; Ayaz, M.; Hassan, S.; Bungau, S. Alzheimer’s Disease as a Major Public Health Concern: Role of Dietary Saponins in Mitigating Neurodegenerative Disorders and Their Underlying Mechanisms. Molecules 2022, 27, 6804.
  24. Rao, A.V.; Gurfinkel, D.M. The bioactivity of saponins: Triterpenoid and steroidal glycosides. Drug Metabol. Drug Interact. 2000, 17, 211–235.
  25. Esmeeta, A.; Adhikary, S.; Dharshnaa, V.; Swarnamughi, P.; Ummul Maqsummiya, Z.; Banerjee, A.; Pathak, S.; Duttaroy, A.K. Plant-derived bioactive compounds in colon cancer treatment: An updated review. Biomed. Pharmacother. 2022, 153, 113384.
  26. Dandawate, P.R.; Subramaniam, D.; Padhye, S.B.; Anant, S. Bitter melon: A panacea for inflammation and cancer. Chin. J. Nat. Med. 2016, 14, 81–100.
  27. Shalini, V.T.; Neelakanta, S.J.; Sriranjini, J.S. Neuroprotection with Bacopa monnieri—A review of experimental evidence. Mol. Biol. Rep. 2021, 48, 2653–2668.
  28. Bahrami, Y.; Zhang, W. MMFC Distribution of Saponins in the Sea Cucumber Holothuria lessoni; the Body Wall Versus the Viscera, and Their Biological Activities. Mar. Drugs 2018, 16, 423.
  29. Morroni, F.; Sita, G.; Graziosi, A.; Turrini, E.; Fimognari, C.; Tarozzi, A.; Hrelia, P. Protective Effects of 6-(Methylsulfinyl) hexyl Isothiocyanate on Abeta (1-42)-Induced Cognitive Deficit, Oxidative Stress, Inflammation, and Apoptosis in Mice. Int. J. Mol. Sci. 2018, 19, 2083.
  30. Mucke, L.; Selkoe, D.J. Neurotoxicity of amyloid β-protein: Synaptic and network dysfunction. Cold Spring Harb Perspect Med. 2012, 2, a006338.
  31. Lesne, S.; Koh, M.T.; Kotilinek, L.; Kayed, R.; Glabe, C.G.; Yang, A.; Gallagher, M.; Ashe, K.H. A specific amyloid-β protein assembly in the brain impairs memory. Nature 2006, 440, 352–357.
  32. Mumford, P.; Tosh, J.; Anderle, S.; Gkanatsiou, W.E.; Lau, G.; Noy, S.; Cleverley, K.; Saito, T.; Saido, T.C.; Yu, E.; et al. Genetic Mapping of APP and Amyloid-beta Biology Modulation by Trisomy 21. J. Neurosci. 2022, 42, 6453–6468.
  33. Zhao, P.; Xu, Y.; Jiang, L.L.; Fan, X.; Ku, Z.; Li, L.; Liu, X.; Deng, M.; Arase, H.; Zhu, J.J.; et al. LILRB2-mediated TREM2 signaling inhibition suppresses microglia functions. Mol. Neurodegener. 2022, 17, 44.
  34. Lee, E.K.; Kim, H.H.; Kuwano, Y.; Abdelmohsen, K.; Srikantan, S.; Subaran, S.S.; Gleichmann, M.; Mughal, M.R.; Martindale, J.L.; Yang, X.; et al. hnRNP C promotes APP translation by competing with FMRP for APP mRNA recruitment to P bodies. Nat. Struct. Mol. Biol. 2010, 17, 732–739.
  35. Sachdev, R.; Kappes-Horn, K.; Paulsen, L.; Duernberger, Y.; Pleschka, C.; Denner, P.; Kundu, B.; Reimann, J.; Vorberg, I. Endoplasmic Reticulum Stress Induces Myostatin High Molecular Weight Aggregates and Impairs Mature Myostatin Secretion. Mol. Neurobiol. 2018, 55, 8355–8373.
  36. Mockett, B.G.; Guévremont, D.; Elder, M.K.; Parfitt, K.D.; Peppercorn, K.; Morrissey, J.; Singh, A.; Hintz, T.J.; Kochen, L.; Tom Dieck, S.; et al. Glutamate Receptor Trafficking and Protein Synthesis Mediate the Facilitation of LTP by Secreted Amyloid Precursor Protein-α. J. Neurosci. 2019, 39, 3188–3203.
  37. Cox, P.A.; Davis, D.A.; Mash, D.C.; Metcalf, J.S.; Banack, S.A. Dietary exposure to an environmental toxin triggers neurofibrillary tangles and amyloid deposits in the brain. Proc. R. Soc. B. Biol. Sci. 2016, 283, 20152317.
  38. Samaey, C.; Schreurs, A.; Stroobants, S.; Balschun, D. Early Cognitive and Behavioral Deficits in Mouse Models for Tauopathy and Alzheimer’s Disease. Front. Aging Neurosci. 2019, 11, 335.
  39. Koon, N.A.; Itokazu, Y.; Yu, R.K. Ganglioside-Dependent Neural Stem Cell Proliferation in Alzheimer’s Disease Model Mice. ASN Neuro. 2015, 7, 1759091415618916.
  40. Hongo, N.; Takamura, Y.; Nishimaru, H.; Matsumoto, J.; Tobe, K.; Saito, T.; Saido, T.C.; Nishijo, H. Astaxanthin Ameliorated Parvalbumin-Positive Neuron Deficits and Alzheimer’s Disease-Related Pathological Progression in the Hippocampus of App (NL-G-F/NL-G-F) Mice. Front. Pharmacol. 2020, 11, 307.
  41. Sasaguri, H.; Hashimoto, S.; Watamura, N.; Sato, K.; Takamura, R.; Nagata, K.; Tsubuki, S.; Ohshima, T.; Yoshiki, A.; Sato, K.; et al. Recent Advances in the Modeling of Alzheimer’s Disease. Front. Neurosci. 2022, 16, 807473.
  42. Lu, X.Y.; Huang, S.; Chen, Q.B.; Zhang, D.; Li, W.; Ao, R.; Leung, F.C.; Zhang, Z.; Huang, J.; Tang, Y.; et al. Metformin Ameliorates Aβ Pathology by Insulin-Degrading Enzyme in a Transgenic Mouse Model of Alzheimer’s Disease. Oxid. Med. Cell. Longev. 2020, 2020, 2315106.
  43. Moosecker, S.; Gomes, P.; Dioli, C.; Yu, S.; Sotiropoulos, I.; Almeida, O. Activated PPARγ Abrogates Misprocessing of Amyloid Precursor Protein, Tau Missorting and Synaptotoxicity. Front. Cell. Neurosci. 2019, 13, 239.
  44. Medrano-Jiménez, E.; Jiménez-Ferrer Carrillo, I.; Pedraza-Escalona, M.; Ramírez-Serrano, C.E.; Álvarez-Arellano, L.; Cortés-Mendoza, J.; Herrera-Ruiz, M.; Jiménez-Ferrer, E.; Zamilpa, A.; Tortoriello, J.; et al. Malva parviflora extract ameliorates the deleterious effects of a high fat diet on the cognitive deficit in a mouse model of Alzheimer’s disease by restoring microglial function via a PPAR-γ-dependent mechanism. J. Neuroinflammation 2019, 16, 143.
  45. Dinda, B.; Dinda, M.; Kulsi, G.; Chakraborty, A.; Dinda, S. Therapeutic potentials of plant iridoids in Alzheimer’s and Parkinson’s diseases: A review. Eur. J. Med. Chem. 2019, 169, 185–199.
  46. Kumar, A.P.; Prabhakaran, P.; Kumar, B.R.P.; Jeyarani, V.; Dhanabal, S.P.; Justin, A. Glitazones, PPAR-γ and Neuroprotection. Mini Rev. Med. Chem. 2021, 21, 1457–1464.
  47. Mohamed, L.A.; Qosa, H.; Kaddoumi, A. Age-Related Decline in Brain and Hepatic Clearance of Amyloid-β is Rectified by the Cholinesterase Inhibitors Donepezil and Rivastigmine in Rats. ACS Chem. Neurosci. 2015, 6, 725–736.
  48. Selkoe, D.J.; Hardy, J. The amyloid hypothesis of Alzheimer’s disease at 25 years. EMBO Mol. Med. 2016, 8, 595–608.
  49. Pagano, K.; Tomaselli, S.; Molinari, H.; Ragona, L. Natural Compounds as Inhibitors of Aβ Peptide Aggregation: Chemical Requirements and Molecular Mechanisms. Front. Neurosci. 2020, 14, 619667.
  50. Alolga, R.N.; Nuer-Allornuvor, G.F.; Kuugbee, E.D.; Yin, X.; Ma, G. Ginsenoside Rg1 and the control of inflammation implications for the therapy of type 2 diabetes: A review of scientific findings and call for further research. Pharmacol. Res. 2020, 152, 104630.
  51. 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.
  52. Xue, W.; Liu, Y.; Qi, W.Y.; Gao, Y.; Li, M.; Shi, A.X.; Li, K.X. Pharmacokinetics of ginsenoside Rg1 in rat medial prefrontal cortex, hippocampus, and lateral ventricle after subcutaneous administration. J. Asian Nat. Prod. Res. 2016, 18, 587–595.
  53. Kunisawa, K.; Shan, J.; Lu, Q.; Yang, Y.; Kosuge, A.; Kurahashi, H.; Saito, K.; Zou, L.; Nabeshima, T.; Mouri, A. Loureirin C and Xanthoceraside Attenuate Depression-Like Behaviors and Expression of Interleukin-17 in the Prefrontal Cortex Induced by Chronic Unpredictable Mild Stress in Mice. Neurochem. Res. 2022, 47, 2880–2889.
  54. Jin, G.; Wang, L.H.; Ji, X.F.; Chi, T.Y.; Qi, Y.; Jiao, Q.; Xu, Q.; Zhou, X.Y.; Zhang, R.; Zou, L.B. Xanthoceraside rescues learning and memory deficits through attenuating β-amyloid deposition and tau hyperphosphorylation in APP mice. Neurosci. Lett. 2014, 573, 58–63.
  55. Qiu, J.; Li, W.; Feng, S.H.; Wang, M.; He, Z.Y. Ginsenoside Rh2 promotes nonamyloidgenic cleavage of amyloid precursor protein via a cholesterol-dependent pathway. Genet. Mol. Res. 2014, 13, 3586–3598.
  56. 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.
  57. Li, X.; Cui, J.; Yu, Y.; Li, W.; Hou, Y.; Wang, X.; Qin, D.; Zhao, C.; Yao, X.; Zhao, J.; et al. Traditional Chinese Nootropic Medicine Radix Polygalae and Its Active Constituent Onjisaponin B Reduce β-Amyloid Production and Improve Cognitive Impairments. PLoS ONE 2016, 11, e151147.
  58. Zhang, Z.; Yang, J.; Liu, C.; Xie, J.; Qiu, S.; Yang, X.; Wu, C. Pseudoginsenoside-F11 alleviates cognitive deficits and Alzheimer’s disease-type pathologies in SAMP8 mice. Pharmacol. Res. 2019, 139, 512–523.
  59. Khan, M.I.; Shin, J.H.; Kim, M.Y.; Shin, T.S.; Kim, J.D. Green Tea Seed Isolated Theasaponin E1 Ameliorates AD Promoting Neurotoxic Pathogenesis by Attenuating Aβ Peptide Levels in SweAPP N2a Cells. Molecules 2020, 25, 2334.
  60. Kang, M.S.; Baek, S.H.; Chun, Y.S.; Moore, A.Z.; Landman, N.; Berman, D.; Yang, H.O.; Morishima-Kawashima, M.; Osawa, S.; Funamoto, S.; et al. Modulation of lipid kinase PI4KIIα activity and lipid raft association of presenilin 1 underlies γ-secretase inhibition by ginsenoside (20S)-Rg3. J. Biol. Chem. 2013, 288, 20868–20882.
  61. Yang, Q.; Lin, J.; Zhang, H.; Liu, Y.; Kan, M.; Xiu, Z.; Chen, X.; Lan, X.; Li, X.; Shi, X.; et al. Ginsenoside Compound K Regulates Amyloid β via the Nrf2/Keap1 Signaling Pathway in Mice with Scopolamine Hydrobromide-Induced Memory Impairments. J. Mol. Neurosci. 2019, 67, 62–71.
  62. Zhang, X.; Wang, J.; Xing, Y.; Gong, L.; Li, H.; Wu, Z.; Li, Y.; Wang, J.; Wang, Y.; Dong, L.; et al. Effects of ginsenoside Rg1 or 17β-estradiol on a cognitively impaired, ovariectomized rat model of Alzheimer’s disease. Neuroscience 2012, 220, 191–200.
  63. Shi, C.; Zheng, D.D.; Fang, L.; Wu, F.; Kwong, W.H.; Xu, J. Ginsenoside Rg1 promotes nonamyloidgenic cleavage of APP via estrogen receptor signaling to MAPK/ERK and PI3K/Akt. Biochim. Biophys. Acta 2012, 1820, 453–460.
  64. Lin, Z.; Gu, J.; Xiu, J.; Mi, T.; Dong, J.; Tiwari, J.K. Traditional Chinese medicine for senile dementia. Evid. Based Complement. Altern. Med. 2012, 2012, 692621.
  65. Zhu, L.; Hou, X.J.; Che, X.H.; Zhou, T.S.; Liu, X.Q.; Wu, C.F.; Yang, J.Y. Pseudoginsenoside-F11 attenuates cognitive dysfunction and tau phosphorylation in sporadic Alzheimer’s disease rat model. Acta Pharmacol. Sin. 2021, 42, 1401–1408.
  66. Kim, J.D.; Chaudhary, N.; Seo, H.J.; Kim, M.Y.; Shin, T.S. Theasaponin E1 as an effective ingredient for anti-angiogenesis and anti-obesity effects. Biosci. Biotechnol. Biochem. 2014, 78, 279–287.
  67. Oh, J.; Kim, J.S. Compound K derived from ginseng: Neuroprotection and cognitive improvement. Food Funct. 2016, 7, 4506–4515.
  68. Lee, S.; Kwon, M.C.; Jang, J.P.; Sohng, J.K.; Jung, H.J. The ginsenoside metabolite compound K inhibits growth, migration and stemness of glioblastoma cells. Int. J. Oncol. 2017, 51, 414–424.
  69. Kim, J.K.; Cui, C.H.; Yoon, M.H.; Kim, S.C.; Im, W.T. Bioconversion of major ginsenosides Rg1 to minor ginsenoside F1 using novel recombinant ginsenoside hydrolyzing glycosidase cloned from Sanguibacter keddieii and enzyme characterization. J. Biotechnol. 2012, 161, 294–301.
  70. Han, J.; Oh, J.P.; Yoo, M.; Cui, C.H.; Jeon, B.M.; Kim, S.C.; Han, J.H. Minor ginsenoside F1 improves memory in APP/PS1 mice. Mol. Brain 2019, 12, 77.
  71. Chakravarty, A.K.; Sarkar, T.; Masuda, K.; Shiojima, K.; Nakane, T.; Kawahara, N. Bacopaside I and II: Two pseudojujubogenin glycosides from Bacopa monniera. Phytochemistry 2001, 58, 553–556.
  72. Pham, H.T.N.; Tran, H.N.; Nguyen, P.T.; Le, X.T.; Nguyen, K.M.; Phan, S.V.; Yoneyama, M.; Ogita, K.; Yamaguchi, T.; Folk, W.R.; et al. Bacopa monnieri (L.) Wettst. Extract Improves Memory Performance via Promotion of Neurogenesis in the Hippocampal Dentate Gyrus of Adolescent Mice. Int. J. Mol. Sci. 2020, 21, 3365.
  73. Ramasamy, S.; Chin, S.P.; Sukumaran, S.D.; Buckle, M.J.; Kiew, L.V.; Chung, L.Y. In Silico and In Vitro Analysis of Bacoside A Aglycones and Its Derivatives as the Constituents Responsible for the Cognitive Effects of Bacopa monnieri. PLoS ONE 2015, 10, e0126565.
  74. Li, Y.; Yuan, X.; Shen, Y.; Zhao, J.; Yue, R.; Liu, F.; He, W.; Wang, R.; Shan, L.; Zhang, W. Bacopaside I ameliorates cognitive impairment in APP/PS1 mice via immune-mediated clearance of β-amyloid. Aging 2016, 8, 521–533.
  75. Seyfried, N.T.; Dammer, E.B.; Swarup, V.; Nandakumar, D.; Duong, D.M.; Yin, L.; Deng, Q.; Nguyen, T.; Hales, C.M.; Wingo, T.; et al. A Multi-network Approach Identifies Protein-Specific Co-expression in Asymptomatic and Symptomatic Alzheimer’s Disease. Cell Syst. 2017, 4, 60–72.
  76. Brown, R.A.; Voit, A.; Srikanth, M.P.; Thayer, J.A.; Kingsbury, T.J.; Jacobson, M.A.; Lipinski, M.M.; Feldman, R.A.; Awad, O. mTOR hyperactivity mediates lysosomal dysfunction in Gaucher’s disease iPSC-neuronal cells. Dis. Model. Mech. 2019, 12, dmm038596.
  77. Yao, X.C.; Xue, X.; Zhang, H.T.; Zhu, M.M.; Yang, X.W.; Wu, C.F.; Yang, J.Y. Pseudoginsenoside-F11 alleviates oligomeric β-amyloid-induced endosome-lysosome defects in microglia. Traffic 2019, 20, 61–70.
  78. Usenovic, M.; Niroomand, S.; Drolet, R.E.; Yao, L.; Gaspar, R.C.; Hatcher, N.G.; Schachter, J.; Renger, J.J.; Parmentier-Batteur, S. Internalized Tau Oligomers Cause Neurodegeneration by Inducing Accumulation of Pathogenic Tau in Human Neurons Derived from Induced Pluripotent Stem Cells. J. Neurosci. 2015, 35, 14234–14250.
  79. Rubenstein, R.; Chang, B.; Grinkina, N.; Drummond, E.; Davies, P.; Ruditzky, M.; Sharma, D.; Wang, K.; Wisniewski, T. Tau phosphorylation induced by severe closed head traumatic brain injury is linked to the cellular prion protein. Acta Neuropathol. Commun. 2017, 5, 30.
  80. Papanikolopoulou, K.; Roussou, I.G.; Gouzi, J.Y.; Samiotaki, M.; Panayotou, G.; Turin, L.; Skoulakis, E.M.C. Drosophila Tau Negatively Regulates Translation and Olfactory Long-Term Memory, but Facilitates Footshock Habituation and Cytoskeletal Homeostasis. J. Neurosci. 2019, 39, 8315–8329.
  81. Lacovich, V.; Espindola, S.L.; Alloatti, M.; Pozo, D.V.; Cromberg, L.E.; Carna, M.E.; Forte, G.; Gallo, J.M.; Bruno, L.; Stokin, G.B.; et al. Tau Isoforms Imbalance Impairs the Axonal Transport of the Amyloid Precursor Protein in Human Neurons. J. Neurosci. 2017, 37, 58–69.
  82. Barini, E.; Antico, O.; Zhao, Y.; Asta, F.; Tucci, V.; Catelani, T.; Marotta, R.; Xu, H.; Gasparini, L. Metformin promotes tau aggregation and exacerbates abnormal behavior in a mouse model of tauopathy. Mol. Neurodegener. 2016, 11, 16.
  83. Candia, R.F.; Cohen, L.S.; Morozova, V.; Corbo, C.; Alonso, A.D. Importin-Mediated Pathological Tau Nuclear Translocation Causes Disruption of the Nuclear Lamina, TDP-43 Mislocalization and Cell Death. Front. Molec. Neurosci. 2022, 15, 888420.
  84. Noel, A.; Foveau, B.; LeBlanc, A.C. Caspase-6-cleaved Tau fails to induce Tau hyperphosphorylation and aggregation, neurodegeneration, glial inflammation, and cognitive deficits. Cell Death Dis. 2021, 12, 227.
  85. van der Kant, R.; Goldstein, L.S.B.; Ossenkoppele, R. Amyloid-β-independent regulators of tau pathology in Alzheimer disease. Nat. Rev. Neurosci. 2020, 21, 21–35.
  86. Pîrşcoveanu, D.F.V.; Pirici, I.; Tudorică, V.; Bălşeanu, T.A.; Albu, V.C.; Bondari, S.; Bumbea, A.M.; Pîrşcoveanu, M. Tau protein in neurodegenerative diseases—A review. Rom. J. Morphol. Embryol. 2017, 58, 1141–1150.
  87. Voss, K.; Gamblin, T.C. GSK-3β phosphorylation of functionally distinct tau isoforms has differential, but mild effects. Mol. Neurodegener 2009, 4, 18.
  88. Mahmoudi, N.; Kiasalari, Z.; Rahmani, T.; Sanaierad, A.; Afshin-Majd, S.; Naderi, G.; Baluchnejadmojarad, T.; Roghani, M. Diosgenin Attenuates Cognitive Impairment in Streptozotocin-Induced Diabetic Rats: Underlying Mechanisms. Neuropsychobiology 2021, 80, 25–35.
  89. Ahmed, T.; Raza, S.H.; Maryam, A.; Setzer, W.N.; Braidy, N.; Nabavi, S.F.; de Oliveira, M.R.; Nabavi, S.M. Ginsenoside Rb1 as a neuroprotective agent: A review. Brain Res. Bull. 2016, 125, 30–43.
  90. Shalaby, A.M.; Alnasser, S.M.; Ahmed Khairy, D.; Alabiad, M.A.; Alorini, M.; Jaber, F.A.; Tawfeek, S.E. The neuroprotective effect of ginsenoside Rb1 on the cerebral cortex changes induced by aluminium chloride in a mouse model of Alzheimer’s disease: A histological, immunohistochemical, and biochemical study. J. Chem. Neuroanat. 2023, 129, 102248.
  91. Zhao, H.H.; Di, J.; Liu, W.S.; Liu, H.L.; Lai, H.; Lu, Y.L. Involvement of GSK3 and PP2A in ginsenoside Rb1’s attenuation of aluminum-induced tau hyperphosphorylation. Behav. Brain Res. 2013, 241, 228–234.
  92. Nabavi, S.F.; Sureda, A.; Habtemariam, S.; Nabavi, S.M. Ginsenoside Rd and ischemic stroke; a short review of literatures. J. Ginseng Res. 2015, 39, 299–303.
  93. Adams, M.E.; Li, D.K.; McConkey, J.P.; Davidson, R.G.; Day, B.; Duncan, C.P.; Tron, V. Evaluation of cartilage lesions by magnetic resonance imaging at 0.15 T: Comparison with anatomy and concordance with arthroscopy. J. Rheumatol. 1991, 18, 1573–1580.
  94. 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.
  95. Khan, M.I.; Khan, M.Z.; Shin, J.H.; Shin, T.S.; Lee, Y.B.; Kim, M.Y.; Kim, J.D. Neuroprotective Effects of Green Tea Seed Isolated Saponin Due to the Amelioration of Tauopathy and Alleviation of Neuroinflammation: A Therapeutic Approach to Alzheimer’s Disease. Molecules 2022, 27, 2079.
  96. Zhou, H.; Tai, J.; Xu, H.; Lu, X.; Meng, D. Xanthoceraside Could Ameliorate Alzheimer’s Disease Symptoms of Rats by Affecting the Gut Microbiota Composition and Modulating the Endogenous Metabolite Levels. Front. Pharmacol. 2019, 10, 1035.
  97. Li, F.; Eteleeb, A.M.; Buchser, W.; Sohn, C.; Wang, G.; Xiong, C.; Payne, P.R.; McDade, E.; Karch, C.M.; Harari, O.; et al. Weakly activated core neuroinflammation pathways were identified as a central signaling mechanism contributing to the chronic neurodegeneration in Alzheimer’s disease. Front. Aging Neurosci. 2022, 14, 935279.
  98. Howe, A.M.; Burke, S.; O’Reilly, M.E.; McGillicuddy, F.C.; Costello, D.A. Palmitic Acid and Oleic Acid Differently Modulate TLR2-Mediated Inflammatory Responses in Microglia and Macrophages. Mol. Neurobiol. 2022, 59, 2348–2362.
  99. Kurtys, E.; Casteels, C.; Real, C.C.; Eisel, U.; Verkuyl, J.M.; Broersen, L.M.; Klein, H.C.; Dierckx, R.; Doorduin, J.; de Vries, E. Therapeutic effects of dietary intervention on neuroinflammation and brain metabolism in a rat model of photothrombotic stroke. CNS Neurosci. Ther. 2019, 25, 36–46.
  100. Alkhatip, A.A.A.M.M.; Kamel, M.G.; Hamza, M.K.; Farag, E.M.; Yassin, H.M.; Elayashy, M.; Naguib, A.A.; Wagih, M.; Abd-Elhay, F.A.; Algameel, H.Z.; et al. The diagnostic and prognostic role of neutrophil-to-lymphocyte ratio in COVID-19: A systematic review and meta-analysis. Expert Rev. Mol. Diagn. 2021, 21, 505–514.
  101. Kumar, P.; Mishra, J.; Kumar, N. Mechanistic Role of Jak3 in Obesity-Associated Cognitive Impairments. Nutrients. 2022, 14, 3715.
  102. Merighi, S.; Nigro, M.; Travagli, A.; Gessi, S. Microglia and Alzheimer’s Disease. Int. J. Mol. Sci. 2022, 23, 12990.
  103. Tiwari, S.; Atluri, V.; Kaushik, A.; Yndart, A.; Nair, M. Alzheimer’s disease: Pathogenesis, diagnostics, and therapeutics. Int. J. Nanomed. 2019, 14, 5541–5554.
  104. Yu, Y.; Ye, R.D. Microglial Aβ receptors in Alzheimer’s disease. Cell. Mol. Neurobiol. 2015, 35, 71–83.
  105. Gambuzza, M.E.; Sofo, V.; Salmeri, F.M.; Soraci, L.; Marino, S.; Bramanti, P. Toll-like receptors in Alzheimer’s disease: A therapeutic perspective. CNS Neurol. Disord. Drug Targets 2014, 13, 1542–1558.
  106. Abubakar, M.B.; Sanusi, K.O.; Ugusman, A.; Mohamed, W.; Kamal, H.; Ibrahim, N.H.; Khoo, C.S.; Kumar, J. Alzheimer’s Disease: An Update and Insights Into Pathophysiology. Front. Aging Neurosci. 2022, 14, 742408.
  107. Subedi, L.; Venkatesan, R.; Kim, S.Y. Neuroprotective and anti-inflammatory activities of allyl isothiocyanate through attenuation of JNK/NF-κB/TNF-α signaling. Int. J. Mol. Sci. 2017, 18, 1423.
  108. Radenovic, L.; Nenadic, M.; Ulamek-Koziol, M.; Januszewski, S.; Czuczwar, S.J.; Andjus, P.R.; Pluta, R. Heterogeneity in brain distribution of activated microglia and astrocytes in a rat ischemic model of Alzheimer’s disease after 2 years of survival. Aging 2020, 12, 12251–12267.
  109. Kaur, D.; Sharma, V.; Deshmukh, R. Activation of microglia and astrocytes: A roadway to neuroinflammation and Alzheimer’s disease. Inflammopharmacology 2019, 27, 663–677.
  110. Lawrence, T. The nuclear factor NF-κB pathway in inflammation. Cold Spring Harb. Perspect. Biol. 2009, 1, a001651.
  111. Kim, Y.; Lee, H.; Kim, S.Y.; Lim, Y. Effects of Lespedeza Bicolor Extract on Regulation of AMPK Associated Hepatic Lipid Metabolism in Type 2 Diabetic Mice. Antioxidants 2019, 8, 599.
  112. Urasaki, Y.; Beaumont, C.; Talbot, J.N.; Hill, D.K.; Le, T.T. Akt3 Regulates the Tissue-Specific Response to Copaiba Essential Oil. Int. J. Mol. Sci. 2020, 21, 2851.
  113. Tyagi, A.; Kamal, M.A.; Poddar, N.K. Integrated Pathways of COX-2 and mTOR: Roles in Cell Sensing and Alzheimer’s Disease. Front. Neurosci. 2020, 14, 693.
  114. Dhapola, R.; Hota, S.S.; Sarma, P.; Bhattacharyya, A.; Medhi, B.; Reddy, D.H. Recent advances in molecular pathways and therapeutic implications targeting neuroinflammation for Alzheimer’s disease. Inflammopharmacology 2021, 29, 1669–1681.
  115. Falcicchia, C.; Tozzi, F.; Arancio, O.; Watterson, D.M.; Origlia, N. Involvement of p38 MAPK in Synaptic Function and Dysfunction. Int. J. Mol. Sci. 2020, 21, 5624.
  116. Jain, M.; Singh, M.K.; Shyam, H.; Mishra, A.; Kumar, S.; Kumar, A.; Kushwaha, J. Role of JAK/STAT in the Neuroinflammation and its Association with Neurological Disorders. Ann. Neurosci. 2021, 28, 191–200.
  117. Ayalon, I.; de Barros Marangoni, L.F.; Benichou, J.I.C.; Avisar, D.; Levy, O. Red Sea corals under Artificial Light Pollution at Night (ALAN) undergo oxidative stress and photosynthetic impairment. Glob. Chang. Biol. 2019, 25, 4194–4207.
  118. Gargouri, B.; Yousif, N.M.; Bouchard, M.; Fetoui, H.; Fiebich, B.L. Inflammatory and cytotoxic effects of bifenthrin in primary microglia and organotypic hippocampal slice cultures. J. Neuroinflamm. 2018, 15, 159.
  119. Pousa, P.A.; Souza, R.M.; Melo, P.; Correa, B.; Mendonca, T.; Simoes-E-Silva, A.C.; Miranda, D.M. Telomere Shortening and Psychiatric Disorders: A Systematic Review. Cells 2021, 10, 1423.
  120. Muke, S.; Kaikini, A.; Peshattiwar, V.; Bagle, S.; Dighe, V.; Sathaye, S. Neuroprotective Effect of Coumarin Nasal Formulation: Kindling Model Assessment of Epilepsy. Front. Pharmacol. 2018, 9, 992.
  121. Ward, S.M.; Himmelstein, D.S.; Ren, Y.; Fu, Y.; Yu, X.W.; Roberts, K.; Binder, L.I.; Sahara, N. TOC1: A valuable tool in assessing disease progression in the rTg4510 mouse model of tauopathy. Neurobiol. Dis. 2014, 67, 37–48.
  122. Ohashi, H.; Tsuji, M.; Oguchi, T.; Momma, Y.; Nohara, T.; Ito, N.; Yamamoto, K.; Nagata, M.; Kimura, A.M.; Kiuchi, Y.; et al. Combined Treatment with Curcumin and Ferulic Acid Suppressed the Aβ-Induced Neurotoxicity More than Curcumin and Ferulic Acid Alone. Int. J. Mol. Sci. 2022, 23, 9685.
  123. Cui, Y.; Li, F.; Zhu, X.; Xu, J.; Muhammad, A.; Chen, Y.; Li, D.; Liu, B.; Wang, C.; Wang, Z.; et al. Alfalfa saponins inhibit oxidative stress-induced cell apoptosis through the MAPK signaling pathway. Redox Rep. 2022, 27, 1–8.
  124. Yoo, K.Y.; Park, S.Y. Terpenoids as potential anti-Alzheimer’s disease therapeutics. Molecules 2012, 17, 3524–3538.
  125. Hsieh, Y.H.; Deng, J.S.; Chang, Y.S.; Huang, G.J. Ginsenoside Rh2 Ameliorates Lipopolysaccharide-Induced Acute Lung Injury by Regulating the TLR4/PI3K/Akt/mTOR, Raf-1/MEK/ERK, and Keap1/Nrf2/HO-1 Signaling Pathways in Mice. Nutrients 2018, 10, 1208.
  126. Lv, J.; Lu, C.; Jiang, N.; Wang, H.; Huang, H.; Chen, Y.; Li, Y.; Liu, X. Protective effect of ginsenoside Rh2 on scopolamine-induced memory deficits through regulation of cholinergic transmission, oxidative stress and the ERK-CREB-BDNF signaling pathway. Phytother. Res. 2021, 35, 337–345.
  127. Oyeleke, M.B.; Owoyele, B.V. Saponins and flavonoids from Bacopa floribunda plant extract exhibit antioxidant and anti-inflammatory effects on amyloid β 1-42-induced Alzheimer’s disease in BALB/c mice. J. Ethnopharmacol. 2022, 288, 114997.
  128. Liu, J.; Yan, X.; Li, L.; Zhu, Y.; Qin, K.; Zhou, L.; Sun, D.; Zhang, X.; Ye, R.; Zhao, G. Ginsennoside rd attenuates cognitive dysfunction in a rat model of Alzheimer’s disease. Neurochem. Res. 2012, 37, 2738–2747.
  129. Lee, S.J.; Lee, W.J.; Chang, S.E.; Lee, G.Y. Antimelanogenic effect of ginsenoside Rg3 through extracellular signal-regulated kinase-mediated inhibition of microphthalmia-associated transcription factor. J. Ginseng Res. 2015, 39, 238–242.
  130. Zhang, Y.; Yang, X.; Wang, S.; Song, S. Ginsenoside Rg3 Prevents Cognitive Impairment by Improving Mitochondrial Dysfunction in the Rat Model of Alzheimer’s Disease. J. Agric. Food Chem. 2019, 67, 10048–10058.
  131. Wang, C.M.; Liu, M.Y.; Wang, F.; Wei, M.J.; Wang, S.; Wu, C.F.; Yang, J.Y. Anti-amnesic effect of pseudoginsenoside-F11 in two mouse models of Alzheimer’s disease. Pharmacol. Biochem. Behav. 2013, 106, 57–67.
  132. Huang, T.; Fang, F.; Chen, L.; Zhu, Y.; Zhang, J.; Chen, X.; Yan, S.S. Ginsenoside Rg1 attenuates oligomeric Aβ (1-42)-induced mitochondrial dysfunction. Curr. Alzheimer Res. 2012, 9, 388–395.
  133. Huang, L.; Wu, Y.; Yin, H.; Yang, X.; Yuan, M.; Ning, H. Two new compounds from the stewed Polygonatum cyrtonema Hua and their protective effect against Aβ (25–35) induced cytotoxicity and oxidative stress damage. Nat. Prod. Res. 2021, 35, 4945–4952.
  134. Elekofehinti, O.O.; Kamdem, J.P.; Meinerz, D.F.; Kade, I.J.; Adanlawo, I.G.; Rocha, J.B. Saponin from the fruit of Solanum anguivi protects against oxidative damage mediated by Fe (2+) and sodium nitroprusside in rat brain synaptosome P2 fraction. Arch. Pharm. Res. 2015, 10.
  135. Ji, Y.J.; Kim, S.; Kim, J.J.; Jang, G.Y.; Moon, M.; Kim, H.D. Crude Saponin from Platycodon grandiflorum Attenuates Aβ-Induced Neurotoxicity via Antioxidant, Anti-Inflammatory and Anti-Apoptotic Signaling Pathways. Antioxidants 2021, 10, 1968.
  136. Ma, B.; Meng, X.; Wang, J.; Sun, J.; Ren, X.; Qin, M.; Sun, J.; Sun, G.; Sun, X. Notoginsenoside R1 attenuates amyloid-beta-induced damage in neurons by inhibiting reactive oxygen species and modulating MAPK activation. Int. Immunopharmacol. 2014, 22, 151–159.
  137. Zhou, N.; Tang, Y.; Keep, R.F.; Ma, X.; Xiang, J. Antioxidative effects of Panax notoginseng saponins in brain cells. Phytomedicine 2014, 21, 1189–1195.
  138. Okouchi, M.; Ekshyyan, O.; Maracine, M.; Aw, T.Y. Neuronal apoptosis in neurodegeneration. Antioxid. Redox Signal. 2007, 9, 1059–1096.
  139. Sharma, V.K.; Singh, T.G.; Singh, S.; Garg, N.; Dhiman, S. Apoptotic Pathways and Alzheimer’s Disease: Probing Therapeutic Potential. Neurochem. Res. 2021, 46, 3103–3122.
  140. Mufson, E.J.; He, B.; Nadeem, M.; Perez, S.E.; Counts, S.E.; Leurgans, S.; Fritz, J.; Lah, J.; Ginsberg, S.D.; Wuu, J.; et al. Hippocampal proNGF signaling pathways and β-amyloid levels in mild cognitive impairment and Alzheimer disease. J. Neuropathol. Exp. Neurol. 2012, 71, 1018–1029.
  141. Roth, K.A. Caspases, apoptosis, and Alzheimer disease: Causation, correlation, and confusion. J. Neuropathol. Exp. Neurol. 2001, 60, 829–838.
  142. Castro, M.A.; Dalmolin, R.J.; Moreira, J.C.; Mombach, J.C.; de Almeida, R.M. Evolutionary origins of human apoptosis and genome-stability gene networks. Nucleic Acids Res. 2008, 36, 6269–6283.
  143. Park, K.R.; Lee, H.; Cho, M.; Yun, H.M. A Phytochemical Constituent, (E)-Methyl-Cinnamate Isolated from Alpinia katsumadai Hayata Suppresses Cell Survival, Migration, and Differentiation in Pre-Osteoblasts. Int. J. Mol. Sci. 2020, 21, 3700.
  144. Karousi, P.; Artemaki, P.I.; Sotiropoulou, C.D.; Christodoulou, S.; Scorilas, A.; Kontos, C.K. Identification of Two Novel Circular RNAs Deriving from BCL2L12 and Investigation of Their Potential Value as a Molecular Signature in Colorectal Cancer. Int. J. Mol. Sci. 2020, 21, 8867.
  145. Asara, Y.; Marchal, J.A.; Carrasco, E.; Boulaiz, H.; Solinas, G.; Bandiera, P.; Garcia, M.A.; Farace, C.; Montella, A.; Madeddu, R. Cadmium modifies the cell cycle and apoptotic profiles of human breast cancer cells treated with 5-fluorouracil. Int. J. Mol. Sci. 2013, 14, 16600–16616.
  146. Lanning, M.E.; Yu, W.; Yap, J.L.; Chauhan, J.; Chen, L.; Whiting, E.; Pidugu, L.S.; Atkinson, T.; Bailey, H.; Li, W.; et al. Structure-based design of N-substituted 1-hydroxy-4-sulfamoyl-2-naphthoates as selective inhibitors of the Mcl-1 oncoprotein. Eur. J. Med. Chem. 2016, 113, 273–292.
  147. Valentin, R.; Grabow, S.; Davids, M.S. The rise of apoptosis: Targeting apoptosis in hematologic malignancies. Blood 2018, 132, 1248–1264.
  148. Öztaş, E.; Kara, M.; Boran, T.; Bişirir, E.; Karaman, E.F.; Kaptan, E.; Özhan, G. Cellular Stress Pathways Are Linked to Acetamiprid-Induced Apoptosis in SH-SY5Y Neural Cells. Biology 2021, 10, 820.
  149. Nagy, Z.S.; Esiri, M.M. Apoptosis-related protein expression in the hippocampus in Alzheimer’s disease. Neurobiol. Aging 1997, 18, 565–571.
  150. MacGibbon, G.A.; Lawlor, P.A.; Sirimanne, E.S.; Walton, M.R.; Connor, B.; Young, D.; Williams, C.; Gluckman, P.; Faull, R.L.; Hughes, P.; et al. Bax expression in mammalian neurons undergoing apoptosis, and in Alzheimer’s disease hippocampus. Brain Res. 1997, 750, 223–234.
  151. Costa, I.M.; Lima, F.O.V.; Fernandes, L.C.B.; Norrara, B.; Neta, F.I.; Alves, R.D.; Cavalcanti, J.R.L.P.; Lucena, E.E.S.; Cavalcante, J.S.; Rego, A.C.M.; et al. Astragaloside IV Supplementation Promotes A Neuroprotective Effect in Experimental Models of Neurological Disorders: A Systematic Review. Curr. Neuropharmacol. 2019, 17, 648–665.
  152. Chang, C.P.; Liu, Y.F.; Lin, H.J.; Hsu, C.C.; Cheng, B.C.; Liu, W.P.; Lin, M.T.; Hsu, S.F.; Chang, L.S.; Lin, K.C. Beneficial Effect of Astragaloside on Alzheimer’s Disease Condition Using Cultured Primary Cortical Cells Under β-amyloid Exposure. Mol. Neurobiol. 2016, 53, 7329–7340.
  153. Hu, Y.; Wu, L.; Jiang, L.; Liang, N.; Zhu, X.; He, Q.; Qin, H.; Chen, W. Notoginsenoside R2 reduces Aβ25-35-induced neuronal apoptosis and inflammation via miR-27a/SOX8/β-catenin axis. Hum. Exp. Toxicol. 2021, 40, S347–S358.
  154. Wang, X.; Xu, W.; Chen, H.; Li, W.; Li, W.; Zhu, G. Astragaloside IV prevents Aβ (1-42) oligomers-induced memory impairment and hippocampal cell apoptosis by promoting PPARγ/BDNF signaling pathway. Brain Res. 2020, 1747, 147041.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , , , , , , ,
View Times: 469
Revisions: 3 times (View History)
Update Date: 31 Jul 2023
1000/1000
Video Production Service