Next Article in Journal
Pharmacological Advances for Treatment in Hypertension
Previous Article in Journal
An In Silico Investigation of the Molecular Interactions between Volatile Anesthetics and Actin
Previous Article in Special Issue
Study on the Intervention Mechanism of Cryptotanshinone on Human A2780 Ovarian Cancer Cell Line Using GC-MS-Based Cellular Metabolomics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Active Compounds of Panax ginseng in the Improvement of Alzheimer’s Disease and Application of Spatial Metabolomics

by
Meng Zhang
,
Huazhou Niu
,
Qingqing Li
,
Lili Jiao
,
Hui Li
* and
Wei Wu
*
Jilin Ginseng Academy, Changchun University of Chinese Medicine, Changchun 130117, China
*
Authors to whom correspondence should be addressed.
Pharmaceuticals 2024, 17(1), 38; https://doi.org/10.3390/ph17010038
Submission received: 13 November 2023 / Revised: 14 December 2023 / Accepted: 24 December 2023 / Published: 26 December 2023
(This article belongs to the Special Issue Novel Applications of Metabolomics in Drug Discovery)

Abstract

:
Panax ginseng C.A. Meyer (P. ginseng) is one of the more common traditional Chinese medicines (TCMs). It contains numerous chemical components and exhibits a range of pharmacological effects. An enormous burden is placed on people’s health and life by Alzheimer’s disease (AD), a neurodegenerative condition. Recent research has shown that P. ginseng’s chemical constituents, particularly ginsenosides, have a significant beneficial impact on the prevention and management of neurological disorders. To understand the current status of research on P. ginseng to improve AD, this paper discusses the composition of P. ginseng, its mechanism of action, and its clinical application. The pathogenesis of AD includes amyloid beta protein (Aβ) generation and aggregation, tau protein hyperphosphorylation, oxidant stress, neuroinflammation, mitochondrial damage, and neurotransmitter and gut microbiota disorders. This review presents the key molecular mechanisms and signaling pathways of the active ingredients in P. ginseng involved in improving AD from the perspective of AD pathogenesis. A P. ginseng-related signaling pathway network was constructed to provide effective targets for the treatment of AD. In addition, the application of spatial metabolomics techniques in studying P. ginseng and AD is discussed. In summary, this paper discusses research perspectives for the study of P. ginseng in the treatment of AD, including a systematic and in-depth review of the mechanisms of action of the active substances in P. ginseng, and evaluates the feasibility of applying spatial metabolomics in the study of AD pathogenesis and pharmacological treatment.

1. Introduction

Panax ginseng C.A. Meyer (P. ginseng) is a perennial herb belonging to the family Araliaceae and the genus Panax. P. ginseng has been used for thousands of years; not only does it possess pharmacological properties, but it also has few side effects. It has been used worldwide as a functional food and complementary medicine, among other purposes [1]. P. ginseng comprises many biologically active components, including ginsenosides, gintonin, polysaccharides, peptides, glycoconjugates, and other compounds. The most commonly studied bioactive components claimed to have health benefits are ginsenosides [2]. Ginsenosides have been extensively used for many years to treat various diseases, especially neurodegenerative diseases, including AD. In addition, other active ingredients of P. ginseng, including gintonin and ginseng polysaccharides, have been found to have certain therapeutic effects on nervous system diseases.
AD is a neurodegenerative disease that manifests as a decline in cognitive and memory function. Several factors, including age, sex, and environmental factors, play a role in the development of the disease. It is mostly seen in people over 65 years old. In recent years, AD has attracted extensive research attention as a multifactorial neurodegenerative disease [3]. The pathogenesis and etiology of AD, a relatively common neurodegenerative disease, have not been fully elucidated, and there is a lack of effective treatments [4]. In addition, AD is thought to begin decades before symptoms become apparent [5]. Thus, for the sake of early diagnosis, prevention, and therapy, it is crucial to understand the pathogenesis of AD.
The amyloid cascade and tau proteins are the principal subjects of research with regard to the processes of AD [6]. Senile plaques (SPs), formed by the aggregation of amyloid beta protein (Aβ) and neurofibrillary tangles (NFTs), formed by the hyperphosphorylation of tau proteins, are two hallmark pathologies of AD. In addition, it has been found that the development of AD is accompanied by additional pathological mechanisms. Therefore, multitarget and multichannel therapies should be highlighted as emerging strategies for AD treatment. Moreover, there is growing evidence that components of P. ginseng are important in improving AD. The active ingredients in P. ginseng can also improve AD by acting on different signaling pathways and pathogenic mechanisms. In addition, P. ginseng extracts can be used in combination with other drugs for better therapeutic efficacy [1].
Metabolomics is a systematic approach for comprehensively analyzing small molecules in biological samples, and its application can provide clinically favorable biomarkers [7]. The application of metabolomics is important for the diagnosis of diseases, the study of disease mechanisms, and the selection of drugs [8]. Mass spectrometry imaging (MSI) techniques combine mass spectrometry with microscopic imaging technology to not only provide large amounts of untargeted chemical information, but also identify the spatial distribution of a large number of molecules [9,10]. Common mass spectrometry imaging technologies include matrix-assisted laser desorption ionization (MALDI), secondary ion mass spectrometry (SIMS), desorption electrospray (DESI), and air-flow-assisted desorption electrospray ionization (AFADESI). Spatial metabolomics with the use of mass spectrometry imaging can identify biological pathways of disease based on abnormal distribution and abundance of metabolites [11]. All of this suggests that spatial metabolomics is expected to provide favorable technical support for studying the complex pathophysiology of AD.
Although studies on the effects of P. ginseng on AD have been summarized many times [1,12], to more systematically elucidate the mechanisms of these effects, in this review, we constructed several comprehensive signaling pathway networks for each mechanism of action. In addition to providing an overview of commonly described mechanisms, this paper also emphasizes the role of gut microbes. Finally, this review summarizes recent studies and suggests the feasibility of applying spatial metabolomics to the study of P. ginseng for the treatment of AD.

2. Active Ingredients of P. ginseng

P. ginseng contains many bioactive and potentially effective therapeutic compounds. Wang et al. used the TCMSP database to screen the active ingredients of P. ginseng and found that a total of 190 chemical components have been identified, among which 22 are active components [13]. Numerous studies have found that the main types of compounds in P. ginseng include ginsenosides, gintonin, polysaccharides, peptides, and glycoconjugates. These active ingredients have been found to have various pharmacological effects and can provide good therapeutic benefits for different diseases (Figure 1). The various components and their active roles were reviewed in order to provide a theoretical basis for better development and utilization of P. ginseng.

2.1. Ginsenosides

Ginsenosides are the main active ingredient in several parts of P. ginseng. Ginsenosides are steroidal glycosides. The reported ginsenosides are mainly classified based on the type of saponin, as well as the number, type, and site of glycosyl units. They can be classified as either dammarane or oleanolic acid based on the glycosidic backbone: a tetracyclic backbone with a sugar linkage portion or a pentacyclic backbone with oleanolic acid (OA) and C17 side chain variant (C17SCV) isoforms, respectively. According to the position of the sugar connection, the dammarane skeleton is further classified as a protopanaxadiol (PPD) or protopanaxatriol (PPT) ginsenoside, whose sugar components are connected at the third or sixth position of the triterpenoid structure, respectively [14]. The structural formulae of ginsenosides are shown in Figure 2 [15]. A total of 170 ginsenosides have been identified from different parts of P. ginseng, including 59 OA, 42 PPD, 65 PPT, and 17 C17CSV types. The composition of ginsenosides in different parts is significantly different. A total of 16 types of ginsenosides are commonly present in all tissues, including 9 PPD, 6 PPT, and 1 OA types (and no C17SCV). Among them, ginsenosides Rb2, Rb1, Rc, Rd, Re, Rf, Rg1, and Ro comprise the highest content in P. ginseng, accounting for more than 70% of total saponins. In addition, malonyl saponins containing malonyl groups have also been detected in different parts of P. ginseng. In the roots, rhizomes, leaves, stems, and fruits, protopanaxadiol-type ginsenosides predominate, whereas higher proportions of malonyl- and C17SCV-type ginsenosides are found in the flowers and buds [2].
Each type of ginsenoside has different pharmacological effects. As the main active ingredients in P. ginseng, ginsenosides exhibit significant pharmacological activities, such as anticancer [16], antitumor [17], anti-inflammatory [18], antivirus [19], antiaging [20], antifatigue, and anti-diabetes [21] activities, and can be used in the treatment of cardiovascular diseases [22,23] and neurodegenerative diseases [24,25], including AD [26].
Therefore, increasing the production of ginsenosides and developing more types of ginsenosides are important in the treatment of various diseases. Traditional extraction methods are time-consuming and have low yields; alternative biotechnologies, such as hair root culture, polyploidy, cell suspension culture, protoplast culture, conventional tissue culture, and in-vitro mutagenesis, have been proposed to address these issues and greatly increase the yield of ginsenosides [27]. In addition, the method of using microbial hosts has also been proposed for more efficient and selective production of ginsenosides [28]. Rare ginsenosides also have various biological activities, but their content is extremely low in P. ginseng. Different extraction methods can have an impact on the structure of ginsenosides [29]. An effective combination biotechnology method, which includes tissue culture, fixation, and hydrolysis, has been developed for the production of rare ginsenosides, including Rh2 [30]. In the process of ginsenoside biosynthesis, glycosyltransferases play an important role in forming diverse structures and have biological activity by transferring various sugar parts into ginsenosides. In addition, UDP glycosyltransferases (UGTs) were discovered to help produce rarer ginsenosides [31].

2.2. Gintonin

Gintonin, a glycoprotein complex isolated from P. ginseng, contains three lipid-derived G-protein-coupled receptor ligands: lysophosphatidylinositol (LPI), lysophosphatidic acid (LPA), and linoleic acid (LA). They are the main functional components of gintonin.
Gintonin plays an important role in the treatment of neurodegenerative diseases, including AD. Gintonin has effects against disorders such as AD, Parkinson’s disease (PD), and Huntington’s disease (HD). It can regulate the levels of neurotransmitters, such as acetylcholine, dopamine, norepinephrine, and serotonin, exert antioxidant and anti-inflammatory effects, and induce autophagy to reduce Aβ production. Gintonin plays an ameliorative role in degenerative diseases through the above mechanisms [32]. In addition, gintonin induces rapid and transient opening of the BBB, enhancing the delivery of molecules to the brain. This effect could be used to enhance the therapeutic effect of drugs on brain diseases [33]. Gintonin can facilitate the delivery of AD treatment drugs to the brain, according to one study [34]. This means that gintonin is essential for AD treatment.

2.3. Ginseng Polysaccharides

Ginseng polysaccharides (GPs) are a class of bioactive compounds present in P. ginseng, which has so far yielded roughly 80 different kinds of polysaccharides, both acidic and neutral. The polysaccharides in P. ginseng are mostly neutral, primarily amyloid mixes, dextrans, and arabinogalactans. A small percentage are acidic; these are pectin-containing rhamnose and homo galacturonic acid [35].
Studies have shown that ginseng polysaccharides have antitumor [36], antifatigue [37], anticancer [38], anti-inflammatory [39], antioxidant [40], immune regulation [41], and intestinal microbiota regulation activities [42], and can be used in the treatment of nervous system diseases [43].

2.4. Ginseng Peptides and Main Proteins

The primary proteins found in P. ginseng are known as ginseng main proteins (GMPs). The study of low-molecular-weight parts of proteins, such as endogenous peptides, protein fragments, and protein degradation products, is called peptidomics [44]. Peptides have a simpler structure than proteins, making them easier to examine, and they can mimic the creation, processing, and destruction of proteins.
P. ginseng proteins contain antiviral, antifungal, and antifatigue properties. The characteristics and roles of precursor proteins can be better understood using peptide fragments. Peptides are anticipated to be potential molecules for evaluating the potency of medicinal plants [45].

2.5. Ginseng Glycoconjugates

Glycoconjugates are compounds that covalently connect one or more monosaccharide or oligosaccharide units with non-carbohydrate parts (proteins, lipids, etc.).
As biological molecules, glycoconjugates participate in regulating various life activities in organisms, and they have anti-inflammatory and analgesic properties. Luo et al. used P. ginseng as the raw material, used 85% ethanol reflux for extraction, and then purified the extract with a G-15 column of silk fatty amide to prepare ginseng glycopeptides (Gg). They found that the ginseng glycopeptides had anti-inflammatory and analgesic effects [46].

3. Pharmacological Effects of Active P. ginseng Compounds on Improving AD

AD is caused by many factors. The occurrence and development of AD result from the joint interaction of various pathogenic components and multiple targets. Many studies have shown that various active substances in P. ginseng, especially ginsenosides, can alleviate AD by regulating signaling pathways related to mechanisms of the disease. Previous studies intensively investigated the mechanism of P. ginseng in improving Aβ aggregation and tau hyperphosphorylation involving more signaling pathways. Research on neuroinflammation, oxidative stress, and neurotransmitters is also increasing. It is worth noting that the contribution of mitochondrial damage and intestinal microbe disorders to AD has also received extensive attention in recent years. This review summarizes the multiple pathways involved in these pathogenic mechanisms, with the expectation of providing theoretical support for the use of P. ginseng in AD treatment.

3.1. Interference of Aβ Generation and Aggregation

Aβ amyloid plaques (senile plaques, SPs) produced by aggregation are important markers of AD [47,48]. Several peptides with 36–43 amino acid residues go under the generic name Aβ. Two peptides, with 40 and 42 amino acid residues, represent the two major species (Aβ1-40 and Aβ1-42) [49]. Aβ aggregation is neurotoxic, leading to membrane rupture, abnormal cell signaling, and organelle dysfunction [50]. In addition, it induces a series of deleterious responses, such as the induction of tau hyperphosphorylation, neuroinflammation, mitochondrial apoptosis, and synaptic dysfunction, which can further contribute to AD. Among the pathogenic mechanisms of P. ginseng in AD, the Aβ protein, which contains the largest number of signaling pathways, is the most widely studied. These signaling pathways mainly play a role in reducing Aβ production and increasing Aβ clearance, as shown in Table 1 and Figure 3.

3.1.1. Interference of Aβ Production by Regulating APP through Multiple Signaling Pathways

Amyloid precursor protein (APP) can be cleaved into Aβ by β-secretory and γ-secretase enzymes, or cleaved into secretory APP α (sAPPα) by α-secretase to block Aβ production. Thus, abnormal processing of APP is a key factor in Aβ accumulation. Research has shown that the active substances in P. ginseng can increase the expression of α-secretory enzymes or reduce the expression of β-secretory and γ-secretory enzymes through different signaling pathways, or reduce the expression of APP to further reduce the amount of Aβ, contributing to improving AD.
Presenilin 1 (PS1) is the key catalytic component of γ-secretase, and Beta-site APP-cleaving enzyme 1 (BACE1) is the key component of β-secretase. Therefore, inhibition of PS1 and BACE1 expression is a key target to block Aβ generation. A variety of active compounds in P. ginseng have this effect. Karpagam et al. [54] found that the ginsenosides CK, F1, Rh1, and Rh2 all have the potential to inhibit BACE1 expression. Choi et al. also found that Rb1 and Rb2 potentially have inhibitory activity against BACE1 [60]. The ginsenoside CK was found to not only reduce the expression of APP, but also inhibit the expression of BACE1 and PS1 [51]. The compound 1-(3,4-dimethoxyphenethyl)-3-(3dehydroxyl-20(s)-protopanaxadiol-3b-yl)-urea (DDPU) is one of the derivatives of PPD. DDPU can be used to decrease BACE1 protein levels by inhibiting the activation of protein kinase RNA-like endoplasmic reticulum kinase (PERK) and decreasing the phosphorylation of eukaryotic translation initiation factor-2α (eIF2 α) [61]. Similarly, the ginsenoside Rg1 was found to reduce Aβ production by inhibiting APP production and modulating α-, β-, and γ-secretase activity [65]. Quan et al. reported that the ginsenoside Rg1 can significantly reduce the expression of cyclin-dependent kinase 5 (CDK5), inhibit the phosphorylation of peroxisome proliferator-activated receptor y (PPARγ), and downregulate the expression of BACE1, decreasing Aβ levels [52]. In contrast, another study found that Rg1 decreased Aβ levels by increasing PPARγ expression in the hippocampus [65]. In addition, the ginsenosides Re [59] and Rb1 [53] have also been found to reduce Aβ production by activating PPARγ. Interestingly, Aβ deposition was shown to increase NADPH oxidase 2 (NOX2) expression and reactive oxygen species (ROS) generation, which further upregulated the expression of APP and BACE in APP/PS1 mice, resulting in a vicious circle. However, Rg1 treatment significantly decreased NOX2 expression and ROS production, thus breaking this vicious circle [57].

3.1.2. Direct Interference of Aβ Production through Various Signaling Pathways

Aβ aggregation may also be directly reduced through different signaling pathways by the active compounds in P. ginseng.
Insulin-degrading enzyme (IDE) and Neprilysin (NEP) are important Aβ-degrading enzymes that can reduce the production of Aβ. Various active compounds in P. ginseng can reduce Aβ production by increasing the activity of IDE and NEP. The ginsenosides CK [51] and Rg1 [65] can increase IDE activity and reduce Aβ expression. Rg1 can increase the expression of IDE, by reducing the expression of CDK5, inhibiting the phosphorylation of PPARγ [52]. In addition, as a metabolite of Rg1, the ginsenoside F1 enhances the expression of IDE, as well as NEP, in vitro and in vivo [55]. Meanwhile, the ginsenoside Rg3 can also enhance NEP expression [56]. Additionally, by inhibiting the Mitogen-Activated Protein Kinase (MAPK) signaling pathway and decreasing ROS levels, the ginsenoside Rd is able to lower the amounts of Aβ [58]. It was found that the ginsenoside complex Sum I can increase the expression of heat shock proteins (HSPs) by upregulating the expression of heat shock factor 1 (HSF-1). This prevents the generation of misfolded or aggregated Aβ and accelerates its breakdown [26].

3.1.3. Interference of Aβ Accumulation by Increasing Aβ Clearance

One of the key processes for the clearance of intracellular and extracellular Aβ is macroautophagy/autophagy [66]. Reducing Aβ aggregation by increasing autophagy is also an important way that P. ginseng’s active substances resist AD.
It has been reported that DDPU suppresses the phosphorylation of Phosphoinositide 3-kinase (PI3K), v-akt murine thymoma viral oncogene homologue (AKT), mammalian target of rapamycin (mTOR), and UNC-51-like kinase 1 (ULK1) to reduce Aβ [61]. The ginsenoside Rg3 could increase the expression of scavenger receptor type A (SRA) and upregulate the uptake of Aβ42 in an in-vitro model, as recently reported by Ahn et al. [62]. Furthermore, a 4.7 kDa ginseng-derived polysaccharide (GP4) was found to induce autophagy and reduce Aβ aggregation, possibly through activation of PTEN-induced kinase 1 (PINK1)/Parkin RBR E3 ubiquitin-protein ligase (Parkin) [63].

3.1.4. Improving AD by Modulating Aβ Downstream Mechanisms

Aβ aggregation further regulates downstream signaling pathways and promotes the generation of other pathogenic mechanisms. Active substances in P. ginseng have been found to not only regulate the downstream pathways of Aβ, but also improve other pathological conditions caused by Aβ as a way to combat AD.
Brain-derived neurotrophic factor (BDNF) and cyclic adenosine monophosphate (cAMP) response element binding protein (CREB) are the key downstream mediators of Aβ toxicity. It has been suggested that Aβ, by decreasing CREB phosphorylation, decreases BDNF expression and, consequently, has a toxic effect in AD [20]. The ginsenoside F1 plays a therapeutic role in AD by enhancing the expression of a phosphorylated form of CREB in the hippocampus and increasing the expression of BDNF [64]. In addition, the ginsenoside Re can improve Aβ-aggregation-induced mitochondrial damage [67]. The ginsenoside Rg1 was found to increase the content of acetylcholine in AD model rats induced by Aβ aggregation [68]. The ginsenoside CK can prevent oxidative stress caused by Aβ aggregation, playing a neuroprotective role [51].

3.2. Inhibition of Tau Hyperphosphorylation

One of the established pathogenic components of Alzheimer’s disease is the hyperphosphorylation of tau proteins, which is linked to microtubules. The active compounds in P. ginseng inhibit the hyperphosphorylation of tau proteins through various signaling pathways, thus preventing the generation of NFTs and the destruction of microtubules.

3.2.1. Manifestations of Tau Hyperphosphorylation

Neuronal structure, functionality, and plasticity are significantly influenced by the microtubule system. Changes in the tissue and structure of microtubules and microtubule regulatory proteins (such as the microtubule-associated tau protein) are characteristic of a variety of neurodegenerative diseases [69]. The deposition of tau aggregates is a pathological hallmark of AD that is closely linked both spatially and temporally to the emergence of neurodegeneration and clinical symptoms [70]. Normally, tau controls the assembly and transport of microtubules (MTs). When tau is hyperphosphorylated in AD, it separates from the MT assembly, causing destabilized MTs and impaired axonal transport. Paired helical filaments (PHFs), which are filamentous structures made of phospho–tau aggregates, combine to create aggregates of insoluble NFTs. In addition, there is growing evidence that tau hyperphosphorylation may be caused by Aβ, which eventually causes MT integrity to be compromised [71].

3.2.2. Inhibition of Tau Hyperphosphorylation by Active Compounds of P. ginseng

The active substances of P. ginseng regulate tau-related signaling pathways, as shown in Table 2 and Figure 4.
Glycogen synthase kinase (GSK-3) and protein phosphatase 2A (PP2A) are the main enzymes regulating tau phosphorylation. GSK-3 has two isoforms: alpha and beta. The ginsenoside Rb1 can exert neuroprotective effects by reducing pGSK-3 levels and increasing PP2A levels, preventing tau hyperphosphorylation [72]. In addition, it was proven that the ginsenosides Rd [73] and Rg1 [74] exert neuroprotective effects by preventing tau hyperphosphorylation by increasing the level of PP2A. CDK5 is one of the main protein kinases of tau, and the conversion of p35 to p25 can lead to the dislocation of CDK5. The ginsenoside Rb1 can effectively maintain intracellular calcium homeostasis and block the activation of calpain, prevent calpain-mediated transformation of P35 to P25, and reduce the activity of CDK5 [75]. Furthermore, Rk3 and ginsenosides have the effect of inhibiting tau protein hyperphosphorylation, but the relevant therapeutic mechanisms have not been elucidated [57,76,77].

3.3. Regulation of Neurotransmitter Levels

It has been found that another important mechanism of AD is changes in neurotransmitter levels. Neurotransmitters carry information across neuronal synapses and neuromuscular junctions [78]. AD is often accompanied by changes in the levels of neurotransmitters, including cholinergic, glutamatergic, GABAergic, serotonin, noradrenergic, and histaminergic changes. Changes in neurotransmitter levels have different effects on cognitive and memory function through different methods of action, leading to nervous system diseases. Therefore, regulating neurotransmitter levels and ensuring their stability is an effective treatment direction. In the past, studies on P. ginseng’s modulation of neurotransmitters in AD were mainly focused on ACh, with relatively few studies on other neurotransmitters, as shown in Table 3.

3.3.1. Regulation of ACh Levels

Acetylcholine (ACh), an important neurotransmitter in the cholinergic system, is one of the most widely studied neurotransmitters as a countermeasure against AD. A decreased acetylcholine level is an important factor in the occurrence of AD. It was found that Rh2, a rare ginsenoside in P. ginseng, can have neuroprotective effects by regulating cholinergic transmission and inhibiting oxidative stress, which had a therapeutic effect on a mouse model with scopolamine-induced AD [79]. In addition, it was found that the ginsenoside Rg1 could increase the amount of ACh in the hip cells of AD mice [68]. The ginsenoside Rb1 can promote the release of acetylcholine in the central nervous system [82]. The ginsenoside Re can increase the levels of extracellular acetylcholine [83]. ACh can be degraded by two kinds of cholinesterase (CHS): acetylcholinesterase (AChE) and butyrylcholinesterase (BChE). Therefore, the regulation of AChE and BChE plays a certain role in the treatment of AD, as reported by Kamecki et al. [87]. It has been found that the ginsenosides Rb1, Rb2, Rc, Re, Rg1, and Rg3 have an inhibitory effect against AChE and BChE. Among them, Re has the most significant AChE-inhibitory activity and Rg3 has the most significant BChE-inhibitory activity [60,84]. In addition, Ginseng stem and leaf saponins (GSLSs) can improve the cognitive impairment of patients with AD. Through activity screening and analysis of GSLSs, 31 ginsenosides were found, of which 27 compounds have acetylcholine–enzyme binding activity, and 11 were identified via the enzyme method. This result suggests that the inhibition of AChE may be a potential mechanism by which GSLSs improve AD [80]. It was found experimentally that PPT can improve the cognitive and memory impairment caused by scopolamine and plays a certain role in the treatment of AD. The mechanism may be that PPT inhibits the activity of acetylcholinesterase, increasing the level of acetylcholine [81]. In addition, ChAT is a key enzyme in the synthesis of ACh. Thus, the regulation of ChAT levels is also an important mechanism for improving AD. Gintonin has been found to play an ameliorative role in AD by upregulating choline acetyltransferase (ChAT) expression, increasing ACh levels, and decreasing AChE expression [85].

3.3.2. Regulation of Other Neurotransmitter Levels

The active substances in P. ginseng can also regulate the levels of other neurotransmitters, contributing to improving AD. Amino acids, especially glutamic acid (Glu) and aspartic acid (Asp), are the main excitatory neurotransmitters in the hippocampus and cortex of the brain. Excitotoxicity caused by excitatory amino acid neurotransmitters can lead to various neurodegenerative diseases. GABA is an important inhibitory neurotransmitter in the central nervous system. Reduced GABA levels may be a potential cause of behavioral and psychological symptoms in AD. Glutamatergic and GABAergic neurotransmitters involved in Aβ-induced damage have the opposite effect. Extra cellular vesicles (EVs) are the main transmitters of intercellular information. It was found that the small EVs (sEVs) of neurons are regulated by a balance in neurotransmitter levels, and sEVs released by GABA-treated neurons reduce Aβ-induced injury, while sEVs released by neurons treated with glutamate aggravate Aβ-induced toxicity [88]. Changes in glutamatergic, GABAergic, and cholinergic levels affect the dysfunction of neural activity and the distribution of amyloid and tau, and then induce AD. Among them, the GABAergic level has the greatest impact on cognitive ability, as recently reported by Khan et al. [89]. It was found that ginsenosides can increase the levels of GABA, acetylcholine, and dopamine and reduce the levels of glutamate and aspartate in the hippocampus and cortex, and increase the levels of glycine and serotonin (5-HT) in the blood. Therefore, it can be assumed that ginsenosides play a role in improving AD by modulating a variety of neurotransmitters [76]. In addition, the ginsenoside Re can improve AD by increasing levels of extracellular dopamine (DA) [83]. Mitochondrial dysfunction can lead to abnormal amino acid metabolism. Abnormal metabolism of amino acids, including alanine (Ala), aspartate (Asp), glutamate (Glu), D-glutamine (D-Gln), D-glutamate (D-Glu), and tryptophan (Trp), has been repeatedly reported in the pathogenesis of AD. Rg3 may regulate the disordered amino acid metabolism of AD [86].

3.4. Anti-Oxidative Stress and Anti-Inflammation

It has been reported that the occurrence of neurodegenerative disease is often accompanied by oxidative stress [90] and neuroinflammation [91]. Various active compounds in P. ginseng have been found to improve AD by inhibiting oxidative stress and neuroinflammation, as shown in Table 4.

3.4.1. Anti-Oxidative Stress

Oxidative stress is an important pathogenic factor in the development of AD. In addition to the effects on multiple signaling pathways, the effects of P. ginseng on oxidative stress include the scavenging of ROS and the modulation of multiple oxidoreductases, as shown in Figure 5.
Excess ROS generation is the main cause of oxidative stress. The primary ROS source in the brain is NOX2. Aβ in combination with some metal ions, such as Cu(I/II) and Fe(II/III), causes damage to the steady state of redox and produces ROS, causing damage to neurons [50]. The ginsenoside Rg1 can reduce the production of ROS, mediated by the inhibited expression of NOX2 [57]. By scavenging ROS, it was reported that Rg3 may have protected rats from the oxidative stress of AD [86]. It was found that Rk3 can attenuate intracellular ROS production [77]. Ginseng fibrous root enzymatic hydrolysate (GFREH) was first prepared by digesting fibrous ginseng roots with alkaline protease. GFREH can reduce ROS levels in worms and has antioxidant activity [94]. The ginsenoside Rg1 can downregulate the Akt/mTOR signaling pathway by reducing the production of ROS to suppress NSC aging, thus exerting a neuroprotective effect [93]. Heme oxygenase 1 (HO-1) is the main antioxidant protein in the body; it promotes the degradation of heme and the production of biliverdin, and eliminates free radicals accumulated by an oxidation–reduction imbalance, thereby exerting an antioxidant effect. It has been found that Rk3 has antioxidant activity by activating nuclear factor erythroid 2-related factor 2 (Nrf2) and its downstream proteins, including HO-1 and NAD(P)H: quinone oxidoreductase 1 (NQO1), through increased phosphorylation of adenosine monophosphate-activated protein kinase (AMPK). The effect of Nrf2 on downstream proteins is precisely regulated by Kelch-like ECH-related protein 1 (Keap1) [77]. In addition, Rg1 can alleviate H2O2-induced oxidative stress injury in N2a cells by activating the Nrf2/HO-1 signaling pathway [92]. Furthermore, the ginsenoside CK can activate the Nrf2/Keap1 signaling pathway, which enhances the expression of the downstream HO-1 molecule to deal with oxidative stress and reduce oxidative damage [51]. The decreased superoxide dismutase (SOD) and glutathione peroxidase (GSH-Px) levels and increased malondialdehyde (MDA) and lactate dehydrogenase (LDH) levels caused by Aβ aggregation cause oxidative stress and oxidative damage. CK can increase the levels of SOD and GSH-PX, and reduce the levels of MDA [51]. In addition, Rg1 can induce oxidative stress by increasing the levels of SOD and GSH-PX and decreasing the levels of LDH and MDA [92,93]. Furthermore, Rg3 can increase the levels of SOD, catalase (CAT), and GSH-Px, and decrease the levels of MDA [86]. Rh2 can significantly reverse decreased SOD activity, increased MDA levels, and decreased GSH content induced by scopolamine, contributing to reduced antioxidant stress [79]. PPT can exert antioxidant activity by increasing SOD activity and reducing MDA levels in the hippocampus [81]. GFREH can increase SOD and CAT levels, and decrease MDA and LDH levels in worms, and has an antioxidant role [94].

3.4.2. Anti-Inflammation

Neuroinflammation is an important cause of neuronal damage and is one of the causative factors in the development of AD. The preclinical, intermediate, and late phases of AD are all accompanied by a persistent process of neuroinflammation [12]. The effect of P. ginseng on inflammation is mainly due to the effects of different signaling pathways on inflammatory factors, as shown in Figure 6.
The ginsenoside Rg1 exerts anti-inflammatory effects by inhibiting the activity of the NLRP1 inflammasome and regulating the expression of proinflammatory cytokines (IL-1β and IL-18) [95]. Rg3 can ameliorate the chronic inflammatory state by reducing proinflammatory cytokines, such as TNF-α, IL-1β, COX-2, iNOS, and IL-6 [62,96]. In addition, it is reported that ginseng oligopeptides (GOPs) can reduce the proinflammatory cytokines TNF-α and IL-β, which may be due to the inhibition of genes that cause inflammation (MAPK and NF- κ B) [97].
There are many active substances in P. ginseng that exert anti-inflammatory and antioxidant effects through different methods and signaling pathways synergistically, and have been proven to improve cognitive and memory impairment. They can be used as potential drugs for AD.

3.5. Prevention of Mitochondrial Damage

Studies have shown that mitochondrial dysfunction is a pathological change that usually occurs in the early stages of AD. Mitochondrial damage usually leads to synaptic damage, apoptosis, and neuronal degeneration [98]. The effects of signaling pathways on apoptosis and anti-apoptosis proteins are mainly responsible for P. ginseng’s ability to ameliorate mitochondrial damage, as shown in Table 5 and Figure 7.

3.5.1. Relationship of Mitochondrial Damage and Other Pathogenic Mechanisms

There are interactions between mitochondrial damage and other pathogenic factors of AD. Increased Aβ and phosphorylated tau (p-tau) causes increased ROS, as reported by Pradeepkiran et al. [101]. Oxidative stress caused by increased reactive oxygen species or decreased antioxidants can cause mitochondrial DNA damage, loss of integrity, and mitochondrial dysfunction. Mitochondrial dysfunction is often accompanied by abnormal energy metabolism, in which not enough energy is produced for the brain to maintain learning and memory function. In addition, abnormal lipid metabolism causes abnormal energy metabolism, and lipid metabolism can produce more ROS and aggravate mitochondrial damage. Mitochondrial dysfunction can also cause amino acid metabolism disorders and changes to amino acid levels, and affect the occurrence of AD. Thus, the amelioration of mitochondrial damage is a potential target for improving AD.

3.5.2. Prevention of Mitochondrial Damage by Active Compounds of P. ginseng

Located on the mitochondrial membrane, Bcl-2 and Bcl-xl are crucial anti-apoptotic substances, while the apoptotic promoter Bax, which physically resembles Bcl-2, can cause cytochrome C (Cyt C) release, caspase-3 expression, and, ultimately, apoptosis [102]. Studies have shown that Bax and Bcl-2 can regulate the permeability of the mitochondrial membrane by modulating the mitochondrial permeability transition pore (MPTP), and can also prevent the release of mitochondrial Cyt C. Bax and Bcl-2 can exert their anti-apoptotic effects by both of the above mechanisms, preventing cell death [86]. For this reason, increasing Bcl-2 and Bcl-xl expression, decreasing Bax and caspase-3 expression, and decreasing Cyt C release are important mechanisms to stop mitochondrial damage. The ginsenoside CK was found to increase Bcl-2 expression and decrease Bax and caspase-3 expression [51]. Disturbances in the electron transport chain (ETC) are an important component of mitochondrial dysfunction. The ETC is an important part of the balance between intracellular ROS and cellular redox states, including NAD+ and NADH. Overproduction of ROS leads to a decrease in ETC activity, which in turn leads to a decrease in ATP synthesis. However, ATP can inhibit Cyt C activity by binding to Cyt C. Therefore, maintaining stable ETC and ATP levels is one way to stop mitochondrial damage. The ginsenoside Rg3 was found to regulate abnormal energy metabolism and restore mitochondrial ETC disorder, thus improving mitochondrial dysfunction in AD rats. In addition, Rg3 can regulate the levels of apoptosis-related factors (Bax and Bcl-2) and prevent mitochondrial damage. Furthermore, the ginsenoside Rg3 was found to improve cognitive and memory impairment induced by mitochondrial dysfunction in AD rats by regulating abnormal purine metabolism, amino acid metabolism, and anti-apoptosis activity [86]. The ginsenoside Re can downregulate Bax expression, increase the Bcl-2/Bax ratio, and suppress Cyt C release by reducing ROS production, decreasing apoptosis signal kinase 1 (ASK-1) and c-Jun N-terminal kinase (JNK) phosphorylation [67]. Upregulation of the NF-κB/NO signaling pathway can also cause mitochondrial damage. The ginsenoside Rg1 can increase the Bcl-2/Bax ratio, reduce the release of Cyt C from mitochondria to the cytoplasm, and block mitochondria-mediated apoptosis by scavenging ROS to downregulate the NF-κB/NO signaling pathway and inhibit NO production [99]. In addition, Rg1 can change a variety of mitochondrial proteins, which can reduce AD-related mitochondrial damage and produce a certain therapeutic effect. Mitochondrial proteins containing hydroxysteroid 17-beta dehydrogenase 10 (HSD17B10), alanyl-tRNA synthetase 2 (AARS2), translocase of outer mitochondrial membrane 40 (TOMM40), Cyt C oxidase subunit 5A (COX5A), voltage-dependent anion channel 1 (VDAC1), and renamed COXFA4 (NDUFA4) are associated with AD-related mitochondrial dysfunction. It has been reported that Rg1 prevents mitochondrial damage by upregulating mitochondrial proteins HSD17B10, AARS2, and TOMM40, and downregulating NDUFA4, VDAC1, and COX5A [100].

3.6. Regulation of Gut Microbiota

In recent years, studies have found a causal relationship between the gut microbiota and AD [103]. Imbalances in the gut microbiota may promote Aβ aggregation, affect neurotransmitter levels, and trigger neuroinflammation; this may be an indirect factor in the development of AD. Active substances in P. ginseng, such as ginsenosides and polysaccharides, have been found to regulate intestinal microbiota and neurotransmitter disorders caused by them; however, the mechanism by which P. ginseng regulates gut microbe disorders to improve neuroinflammation and Aβ aggregation has not yet been elucidated, as shown in Table 6.

3.6.1. Regulation of Gut Microbiota

There are some healthy microorganisms in the gut that can reduce neuroinflammation, vascular pathology, and Aβ aggregation, and the maintenance of brain balance plays a beneficial role in the treatment of AD [108]. However, under certain pathological conditions, the intestinal microenvironment may be more conducive to the overgrowth of some bacteria. GRg1 may alter the composition and abundance of gut microbes to improve AD [105]. Dushen tang (a ginseng decoction), which is composed of a single herbal material, Panax ginseng C.A. Meyer, could improve disturbances of the gut microbiota, including Bacteroidales, Lactobacillus, and Bacteroides, to improve memory impairment [107].

3.6.2. Gut Microbes and Aβ Aggregation

Gut microbe imbalances can promote lipopolysaccharide and amyloid secretion, and accelerate Aβ aggregation and inflammation [109]. A study found that the gut microbiota may stimulate the MAPK signaling pathway in the brain and promote amyloid deposition [110]. The secretion of lipopolysaccharides and amyloid proteins can interfere with the permeability of the intestinal tract and the blood–brain barrier [111]. Increased permeability of the gut and the blood–brain barrier can also allow pathogens to reach the nervous system, and may induce AD [112].

3.6.3. Gut Microbes and Regulation of Neurotransmitters

In addition, increased intestinal and blood–brain permeability affects the production and absorption of neurotransmitters, such as serotonin, GABA, kynurenine, and catecholamine [113]. Intestinal microorganisms produce neuroactive substances that affect neuronal function, host metabolism, and immunity, and then affect the neural pathways of the intestinal tract and nervous system, which has an impact on the occurrence of AD [114]. In addition, it has been found that disorders of the gut microbiota in AD patients affect tryptophan metabolism, with an increased kynurenine-to-tryptophan ratio (Kyn/Trp) [115]. P. ginseng polysaccharides were found to increase the microbial metabolite valeric acid and reduce the proportion of kynurenic acid and Kyn/Trp. However, further research is needed to determine whether they can have therapeutic effects on AD through this pathway [106]. The ginsenoside Rb1 can significantly increase the relative abundance of specific probiotics, especially Lactobacillus helveticus. Increased levels of Lactobacillus helveticus were found to upregulate the expression of GABAA (α2, β2, and γ2) and GABAB (1b and 2) receptor subunits in the hippocampus and striatum of rats, thereby exerting a neuroprotective effect [104].

3.6.4. Gut Microbes and Neuroinflammation

Bidirectional communication between the gut microbiota and neuroactive substances via the gut–brain axis may be involved in inflammation. Some components of the intestinal microbiota can synthesize and release cytokines to activate inflammatory signals [116]. The intestinal microbiota may also regulate the kynurenine pathway (KP) and then neuroinflammation [117]. The metabolites produced by some intestinal microorganisms can affect the activity of the microglia, further mediate neuroinflammation, and cause neuronal necrosis, and can potentially play a role in the occurrence of AD [118].
In recent years, gut microbes have occupied an important place in disease research. Studies have also identified an important role for gut microbes in the development of AD. P. ginseng and its active ingredients have been found to regulate intestinal flora disorders, but the specific mechanism of action still needs more in-depth research.

4. Clinical Studies on P. ginseng to Improve AD

To elucidate the feasibility of applying P. ginseng clinically, we examined some clinical evidence justifying the importance of studying P. ginseng for the treatment of AD.

4.1. Signs and Symptoms of AD

Patients with AD often present with cognitive deficits, loss of independence, and difficulty maintaining a normal life. Clinically, they present with Aβ protein accumulation in the cerebrospinal fluid and hyperphosphorylation of tau. In addition, temporal lobe and hippocampal atrophy may be a marker for the development of AD [119].

4.2. Current Diagnosis of AD

Initially, the onset of AD was diagnosed primarily by the presence of dementia, with the AD Assessment Scale (ADAS) and mini-mental state examination (MMSE) used for diagnosis [120]. However, a minority of AD patients do not have prominent memory impairment. Therefore, with the continuous development of related research, the diagnosis of AD has changed from the presence of dementia symptoms to the presence of biomarkers. Biomarkers have been categorized as A (Aβ), T (P-tau), and N (neurodegeneration), in what is known as the ATN framework [119].

4.3. Clinical Studies on P.ginseng to Improve AD

In addition to animal and cell experiments, based on the ADAS and MMSE, some clinical evidence has shown that P. ginseng can improve cognitive impairment, including a report by Lee [121] et al. showing that P. ginseng was able to improve cognitive performance in AD patients. In addition, a P. ginseng product made by high-temperature heat treatment, sun ginseng-135, was found to improve cognitive deficits in patients with clinical AD [120]. Despite the evidence that P. ginseng improves cognitive deficits in AD patients, the mechanism of action on ATN biomarkers and other mechanisms have not been investigated. Therefore, more in-depth research is needed in order to truly apply P. ginseng to clinical use.

5. Application of Spatial Metabolomics in AD Research

Spatial metabolomics technology is widely used in various fields and has received extensive attention in the field of medicine. It can be used to detect differences in the distribution of active ingredients in the tissues of herbs. It has played an important role in the treatment of nervous system diseases, including AD, and can be used to detect spatial and quantitative changes in neurotransmitters and Aβ levels. The study of the protein deposition process and early changes in lipid metabolism can also elucidate the distribution of therapeutic drugs in order to understand their mechanism for treating diseases. The discovery of spatial metabolomics is of great significance in the study of AD, and has fostered significant progress in the field.

5.1. Spatial Metabolomics and P. ginseng

Mass spectrometry and imaging technology can be used to analyze the chemical components of different parts of herbs and are important in the study of the distribution of active substances in different parts of P. ginseng. Ultra-high-performance liquid chromatography quadrupole/time-of-flight mass spectrometry (UPLC-QTOF MS) and DESI-MSI have been used to analyze P. ginseng, and PPT, PPD, and other compounds in P. ginseng have been used as biomarkers for identifying the years of P. ginseng. In addition, the distribution of some malonyl ginsenosides, the neutral ginsenoside Rg1, and other components in different tissues of P. ginseng has been described [122].

5.2. Spatial Metabolomics and the Diagnosis of AD

Spatial metabolomics is widely used in the field of medicine. It can be used for the detection of pathological markers [123] and the study of disease pathogenesis and treatment mechanisms. Spatial metabolomics technology plays an important role in the study of the pathogenesis and treatment mechanisms of a variety of diseases, including cancer [124], cardiovascular diseases [125], tumors [126], diabetes nephropathy [127], and nervous system diseases. Spatial metabolomics plays an important role in the study of depression [127], AD, Parkinson’s disease [128], diabetes encephalopathy [129], and other neurological diseases. It can be used to detect changes in the content and spatial distribution of Aβ, neurotransmitters, and lipids during the initiation of AD treatment. It can also be used to visualize the mechanisms and sites of action of AD treatment drugs. AFADESI-MSI and metabolomics analyses allow the mapping of hundreds of metabolites with different polarity functions involving different metabolic pathways. These metabolites include neurotransmitters, organic acids, purines, and carbohydrates. Thus, AFADESI-MSI and metabolomics could be applied to the discovery of dysfunctional metabolites in brain microregions based on the pathological study of a scopolamine treatment model of AD, which could provide spatial information on the metabolic events of diseases [130]. Amyloid plaques are an early manifestation of AD. Mass spectrometry imaging technology can be used to visualize the formation of plaques and peptide deposition in different structures of the brain and to understand the characteristics of early plaque formation, which can promote research on the early prevention of AD [131]. Mass spectrometry imaging can also be used to visualize the distribution and metabolism of neurochemicals in the brain caused by drugs. DESI-MSI was used to study the effects of P. ginseng and American ginseng on the distribution of brain neurochemicals in rats, with neurochemical substances related to the cold and warm characteristics of the two plants screened out. They were divided into warm markers, which promote energy metabolism in the body, improve the function of the endocrine system, and enhance central nervous system excitability, as well as cool markers, which reduce central nervous system excitability, weakening the metabolism and stress response. These can assist in the treatment and research of nervous system diseases [132]. Lipid metabolism dysfunction is closely related to AD pathogenesis. MALDI-MSI can be used to intuitively reveal the spatial distribution and metabolism of mouse brain regions. It can help in understanding lipid changes in the early stages of AD [133].

5.3. Spatial Metabolomics and Drug Applications

Spatial metabolomics can also help to understand the metabolic distribution of drugs in the body [134].
Spatial metabolomics plays an important role in the study of drug metabolism in vivo. LC-MS/MS combined with nano-spray DESI mass spectrometry was used to analyze the drug distribution of the ginsenoside Rg1 in rats administered with intravenous Rg1 at different times, and the distribution and content changes in the kidney, liver, lung, spleen, heart, and brain at different times were obtained, providing a theoretical basis for drug development [135].
In summary, spatial metabolomics is expected to serve as technical support when studying the mechanisms of action and the pharmacological basis of various active ingredients in P. ginseng in the treatment of AD.

6. Conclusions

P. ginseng contains various active ingredients, including ginsenosides, polysaccharides, and gintonin. These bioactive compounds have multiple pharmacological activities and are effective in improving AD.
The pathogenesis of AD includes Aβ accumulation, tau hyperphosphorylation, increased neurotransmitter levels, oxidative stress, neuroinflammation, mitochondrial apoptosis, and disordered gut microflora. The effects of each pathogenetic factor in AD are related to different signaling pathways. The components in P. ginseng contribute to the treatment of AD by acting on different targets and pathways. In this article, we constructed a signaling pathway network related to the action of active ingredients in P. ginseng and the occurrence of AD to explain the relevant processes involved in using P. ginseng for AD treatment and to provide a theoretical basis for developing it as a therapeutic herb for this disease. However, some of these mechanisms have not been fully elucidated and further studies are still needed, especially regarding the specific mechanisms by which P. ginseng improves AD by modulating gut microbe disorders. Therefore, based on the current evidence, the pathogenesis of AD and the mechanism of action of P. ginseng should be studied more thoroughly in the future to promote P. ginseng and maximize its anti-AD effects.
Evidence from clinical studies has also shown that P. ginseng can improve AD; however, the specific mechanism of action has not been elucidated. Therefore, adequate research is still needed to realize the real application of P. ginseng.
Spatial metabolomics can be applied to the study of the early pathological components of AD, including neurotransmitters and Aβ. In addition, it can be used detect the distribution of effective ingredients in P. ginseng after administration and its regulatory effect on neuroactive substances. Therefore, it is believed that spatial metabolomics can provide technical support for the more effective development of P. ginseng as a medicinal herb for the treatment of AD.

Author Contributions

Conceptualization, M.Z. and W.W.; writing—original draft preparation, M.Z.; writing—review and editing, W.W.; preparing and arranging references, M.Z., H.N. and Q.L.; finishing the figures works, M.Z., H.L. and L.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National College Students Innovation and Entrepreneurship Training Program (202110199015) and the National Natural Science Foundation of China (Grant No: 82003969).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. 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] [PubMed]
  2. Piao, X.; Zhang, H.; Kang, J.P.; Yang, D.U.; Li, Y.; Pang, S.; Jin, Y.; Yang, D.C.; Wang, Y. Advances in Saponin Diversity of Panax Ginseng. Molecules 2020, 25, 3452. [Google Scholar] [CrossRef] [PubMed]
  3. Villain, N.; Dubois, B. Alzheimer’s Disease Including Focal Presentations. Semin. Neurol. 2019, 39, 213. [Google Scholar] [CrossRef] [PubMed]
  4. Huang, X.; Li, N.; Pu, Y.; Zhang, T.; Wang, B. Neuroprotective Effects of Ginseng Phytochemicals: Recent Perspectives. Molecules 2019, 24, 2939. [Google Scholar] [CrossRef]
  5. Zhang, X.-X.; Tian, Y.; Wang, Z.-T.; Ma, Y.-H.; Tan, L.; Yu, J.-T. The Epidemiology of Alzheimer’s Disease Modifiable Risk Factors and Prevention. J. Prev. Alzheimer’s Dis. 2021, 8, 313–321. [Google Scholar] [CrossRef]
  6. Weller, J.; Budson, A. Current Understanding of Alzheimer’s Disease Diagnosis and Treatment. F1000Research 2018, 7, F1000 Faculty Rev-1161. [Google Scholar] [CrossRef]
  7. Aldars-García, L.; Gisbert, J.P.; Chaparro, M. Metabolomics Insights into Inflammatory Bowel Disease: A Comprehensive Review. Pharmaceuticals 2021, 14, 1190. [Google Scholar] [CrossRef]
  8. Wishart, D.S. Emerging Applications of Metabolomics in Drug Discovery and Precision Medicine. Nat. Rev. Drug Discov. 2016, 15, 473–484. [Google Scholar] [CrossRef]
  9. Zou, Y.; Tang, W.; Li, B. Mass Spectrometry Imaging and Its Potential in Food Microbiology. Int. J. Food Microbiol. 2022, 371, 109675. [Google Scholar] [CrossRef]
  10. Vandereyken, K.; Sifrim, A.; Thienpont, B.; Voet, T. Methods and Applications for Single-Cell and Spatial Multi-Omics. Nat. Rev. Genet. 2023, 24, 494–515. [Google Scholar] [CrossRef]
  11. He, M.J.; Pu, W.; Wang, X.; Zhang, W.; Tang, D.; Dai, Y. Comparing DESI-MSI and MALDI-MSI Mediated Spatial Metabolomics and Their Applications in Cancer Studies. Front. Oncol. 2022, 12, 891018. [Google Scholar] [CrossRef] [PubMed]
  12. Li, J.; Huang, Q.; Chen, J.; Qi, H.; Liu, J.; Chen, Z.; Zhao, D.; Wang, Z.; Li, X. Neuroprotective Potentials of Panax Ginseng Against Alzheimer’s Disease: A Review of Preclinical and Clinical Evidences. Front. Pharmacol. 2021, 12, 688490. [Google Scholar] [CrossRef] [PubMed]
  13. Wang, Y.; Liu, X. The Effective Components, Core Targets, and Key Pathways of Ginseng against Alzheimer’s Disease. Evid Based Complement. Alternat. Med. 2023, 2023, 9935942. [Google Scholar] [CrossRef] [PubMed]
  14. Chopra, P.; Chhillar, H.; Kim, Y.-J.; Jo, I.H.; Kim, S.T.; Gupta, R. Phytochemistry of Ginsenosides: Recent Advancements and Emerging Roles. Crit. Rev. Food Sci. Nutr. 2023, 63, 613–640. [Google Scholar] [CrossRef] [PubMed]
  15. Wu, W.; Song, F.; Guo, D.; Mi, J.; Qin, Q.; Yu, Q.; Liu, S. Mass Spectrometry-Based Approach in Ginseng Research: A Promising Way to Metabolomics. Curr. Anal. Chem. 2012, 8, 43–66. [Google Scholar] [CrossRef]
  16. Yao, W.; Guan, Y. Ginsenosides in Cancer: A Focus on the Regulation of Cell Metabolism. Biomed. Pharmacother. 2022, 156, 113756. [Google Scholar] [CrossRef] [PubMed]
  17. Liu, Z.; Liu, T.; Li, W.; Li, J.; Wang, C.; Zhang, K. Insights into the Antitumor Mechanism of Ginsenosides Rg3. Mol. Biol. Rep. 2021, 48, 2639–2652. [Google Scholar] [CrossRef] [PubMed]
  18. Im, D.-S. Pro-Resolving Effect of Ginsenosides as an Anti-Inflammatory Mechanism of Panax Ginseng. Biomolecules 2020, 10, 444. [Google Scholar] [CrossRef]
  19. Yang, X.; Sun, H.; Zhang, Z.; Ou, W.; Xu, F.; Luo, L.; Liu, Y.; Chen, W.; Chen, J. Antiviral Effect of Ginsenosides Rk1 against Influenza a Virus Infection by Targeting the Hemagglutinin 1-Mediated Virus Attachment. Int. J. Mol. Sci. 2023, 24, 4967. [Google Scholar] [CrossRef]
  20. De Oliveira Zanuso, B.; de Oliveira Dos Santos, A.R.; Miola, V.F.B.; Guissoni Campos, L.M.; Spilla, C.S.G.; Barbalho, S.M. Panax Ginseng and Aging Related Disorders: A Systematic Review. Exp. Gerontol. 2022, 161, 111731. [Google Scholar] [CrossRef]
  21. Zhou, P.; Xie, W.; He, S.; Sun, Y.; Meng, X.; Sun, G.; Sun, X. Ginsenoside Rb1 as an Anti-Diabetic Agent and Its Underlying Mechanism Analysis. Cells 2019, 8, 204. [Google Scholar] [CrossRef] [PubMed]
  22. Fan, W.; Huang, Y.; Zheng, H.; Li, S.; Li, Z.; Yuan, L.; Cheng, X.; He, C.; Sun, J. Ginsenosides for the Treatment of Metabolic Syndrome and Cardiovascular Diseases: Pharmacology and Mechanisms. Biomed. Pharmacother. 2020, 132, 110915. [Google Scholar] [CrossRef] [PubMed]
  23. Zhu, G.-X.; Zuo, J.-L.; Xu, L.; Li, S.-Q. Ginsenosides in Vascular Remodeling: Cellular and Molecular Mechanisms of Their Therapeutic Action. Pharmacol. Res. 2021, 169, 105647. [Google Scholar] [CrossRef] [PubMed]
  24. Lu, J.; Wang, X.; Wu, A.; Cao, Y.; Dai, X.; Liang, Y.; Li, X. Ginsenosides in Central Nervous System Diseases: Pharmacological Actions, Mechanisms, and Therapeutics. Phytother. Res. 2022, 36, 1523–1544. [Google Scholar] [CrossRef] [PubMed]
  25. Zheng, M.; Xin, Y.; Li, Y.; Xu, F.; Xi, X.; Guo, H.; Cui, X.; Cao, H.; Zhang, X.; Han, C. Ginsenosides: A Potential Neuroprotective Agent. Biomed. Res. Int. 2018, 2018, 8174345. [Google Scholar] [CrossRef]
  26. Zhi, D.; Yang, W.; Yue, J.; Xu, S.; Ma, W.; Zhao, C.; Wang, X.; Wang, D. HSF-1 Mediated Combined Ginsenosides Ameliorating Alzheimer’s Disease like Symptoms in Caernorhabditis Elegans. Nutr. Neurosci. 2022, 25, 2136–2148. [Google Scholar] [CrossRef]
  27. Gantait, S.; Mitra, M.; Chen, J.-T. Biotechnological Interventions for Ginsenosides Production. Biomolecules 2020, 10, 538. [Google Scholar] [CrossRef]
  28. Chu, L.L.; Huy, N.Q.; Tung, N.H. Microorganisms for Ginsenosides Biosynthesis: Recent Progress, Challenges, and Perspectives. Molecules 2023, 28, 1437. [Google Scholar] [CrossRef]
  29. Li, H.; Jiang, H.; Xu, L.; Deng, Y.; Xu, J.; Zhao, Y. Effects of Different Extraction Methods in Pharmacopoeia on the Content and Structure Transformation of Ginsenosides. Molecules 2022, 27, 4347. [Google Scholar] [CrossRef]
  30. Cao, L.; Wu, H.; Zhang, H.; Zhao, Q.; Yin, X.; Zheng, D.; Li, C.; Kim, M.-J.; Kim, P.; Xue, Z.; et al. Highly Efficient Production of Diverse Rare Ginsenosides Using Combinatorial Biotechnology. Biotechnol. Bioeng. 2020, 117, 1615–1627. [Google Scholar] [CrossRef]
  31. Zhao, J.-N.; Wang, R.-F.; Zhao, S.-J.; Wang, Z.-T. Advance in Glycosyltransferases, the Important Bioparts for Production of Diversified Ginsenosides. Chin. J. Nat. Med. 2020, 18, 643–658. [Google Scholar] [CrossRef] [PubMed]
  32. Choi, S.-H.; Lee, R.; Nam, S.M.; Kim, D.-G.; Cho, I.-H.; Kim, H.-C.; Cho, Y.; Rhim, H.; Nah, S.-Y. Ginseng Gintonin, Aging Societies, and Geriatric Brain Diseases. Integr. Med. Res. 2021, 10, 100450. [Google Scholar] [CrossRef] [PubMed]
  33. Kim, D.-G.; Jang, M.; Choi, S.-H.; Kim, H.-J.; Jhun, H.; Kim, H.-C.; Rhim, H.; Cho, I.-H.; Nah, S.-Y. Gintonin, a Ginseng-Derived Exogenous Lysophosphatidic Acid Receptor Ligand, Enhances Blood-Brain Barrier Permeability and Brain Delivery. Int. J. Biol. Macromol. 2018, 114, 1325–1337. [Google Scholar] [CrossRef] [PubMed]
  34. Choi, S.-H.; Lee, N.-E.; Cho, H.-J.; Lee, R.M.; Rhim, H.; Kim, H.-C.; Han, M.; Lee, E.-H.; Park, J.; Nah, S.-Y. Gintonin Facilitates Brain Delivery of Donepezil, a Therapeutic Drug for Alzheimer Disease, through Lysophosphatidic Acid 1/3 and Vascular Endothelial Growth Factor Receptors. J. Ginseng Res. 2021, 45, 264–272. [Google Scholar] [CrossRef] [PubMed]
  35. Guo, M.; Shao, S.; Wang, D.; Zhao, D.; Wang, M. Recent Progress in Polysaccharides from Panax Ginseng C. A. Meyer. Food Funct. 2021, 12, 494–518. [Google Scholar] [CrossRef]
  36. Tao, R.; Lu, K.; Zong, G.; Xia, Y.; Han, H.; Zhao, Y.; Wei, Z.; Lu, Y. Ginseng Polysaccharides: Potential Antitumor Agents. J. Ginseng Res. 2023, 47, 9–22. [Google Scholar] [CrossRef] [PubMed]
  37. Jiao, L.; Li, J.; Liu, F.; Wang, J.; Jiang, P.; Li, B.; Li, H.; Chen, C.; Wu, W. Characterisation, Chain Conformation and Antifatigue Effect of Steamed Ginseng Polysaccharides With Different Molecular Weight. Front. Pharmacol. 2021, 12, 712836. [Google Scholar] [CrossRef]
  38. Zhai, F.-G.; Liang, Q.-C.; Wu, Y.-Y.; Liu, J.-Q.; Liu, J.-W. Red Ginseng Polysaccharide Exhibits Anticancer Activity through GPX4 Downregulation-Induced Ferroptosis. Pharm. Biol. 2022, 60, 909–914. [Google Scholar] [CrossRef]
  39. Li, S.; Huo, X.; Qi, Y.; Ren, D.; Li, Z.; Qu, D.; Sun, Y. The Protective Effects of Ginseng Polysaccharides and Their Effective Subfraction against Dextran Sodium Sulfate-Induced Colitis. Foods 2022, 11, 890. [Google Scholar] [CrossRef]
  40. Chen, F.; Huang, G. Antioxidant Activity of Polysaccharides from Different Sources of Ginseng. Int. J. Biol. Macromol. 2019, 125, 906–908. [Google Scholar] [CrossRef]
  41. Lee, D.-Y.; Park, C.W.; Lee, S.J.; Park, H.-R.; Kim, S.H.; Son, S.-U.; Park, J.; Shin, K.-S. Anti-Cancer Effects of Panax Ginseng Berry Polysaccharides via Activation of Immune-Related Cells. Front. Pharmacol. 2019, 10, 1411. [Google Scholar] [CrossRef] [PubMed]
  42. Wang, D.; Shao, S.; Zhang, Y.; Zhao, D.; Wang, M. Insight Into Polysaccharides From Panax Ginseng C. A. Meyer in Improving Intestinal Inflammation: Modulating Intestinal Microbiota and Autophagy. Front. Immunol. 2021, 12, 683911. [Google Scholar] [CrossRef] [PubMed]
  43. Wang, N.; Wang, X.; He, M.; Zheng, W.; Qi, D.; Zhang, Y.; Han, C.-C. Ginseng Polysaccharides: A Potential Neuroprotective Agent. J. Ginseng Res. 2021, 45, 211–217. [Google Scholar] [CrossRef] [PubMed]
  44. Zhao, Q.; Bai, Y.; Liu, D.; Zhao, N.; Gao, H.; Zhang, X. Quinetides: Diverse Posttranslational Modified Peptides of Ribonuclease-like Storage Protein from Panax Quinquefolius as Markers for Differentiating Ginseng Species. J. Ginseng Res. 2020, 44, 680–689. [Google Scholar] [CrossRef] [PubMed]
  45. Zhao, N.; Cheng, M.; Lv, W.; Wu, Y.; Liu, D.; Zhang, X. Peptides as Potential Biomarkers for Authentication of Mountain-Cultivated Ginseng and Cultivated Ginseng of Different Ages Using UPLC-HRMS. J. Agric. Food Chem. 2020, 68, 2263–2275. [Google Scholar] [CrossRef]
  46. Luo, H.; Zhu, D.; Wang, Y.; Chen, Y.; Jiang, R.; Yu, P.; Qiu, Z. Study on the Structure of Ginseng Glycopeptides with Anti-Inflammatory and Analgesic Activity. Molecules 2018, 23, 1325. [Google Scholar] [CrossRef]
  47. Reiss, A.B.; Arain, H.A.; Stecker, M.M.; Siegart, N.M.; Kasselman, L.J. Amyloid Toxicity in Alzheimer’s Disease. Rev. Neurosci. 2018, 29, 613–627. [Google Scholar] [CrossRef]
  48. Lee, C.H.; Ko, M.S.; Kim, Y.S.; Ham, J.E.; Choi, J.Y.; Hwang, K.W.; Park, S.-Y. Neuroprotective Effects of Davallia Mariesii Roots and Its Active Constituents on Scopolamine-Induced Memory Impairment in In Vivo and In Vitro Studies. Pharmaceuticals 2023, 16, 1606. [Google Scholar] [CrossRef]
  49. Razgonova, M.P.; Veselov, V.V.; Zakharenko, A.M.; Golokhvast, K.S.; Nosyrev, A.E.; Cravotto, G.; Tsatsakis, A.; Spandidos, D.A. Panax Ginseng Components and the Pathogenesis of Alzheimer’s Disease. Mol. Med. Rep. 2019, 19, 2975–2998. [Google Scholar] [CrossRef]
  50. Han, J.; Du, Z.; Lim, M.H. Mechanistic Insight into the Design of Chemical Tools to Control Multiple Pathogenic Features in Alzheimer’s Disease. Acc. Chem. Res. 2021, 54, 3930–3940. [Google Scholar] [CrossRef]
  51. 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. [Google Scholar] [CrossRef] [PubMed]
  52. Quan, Q.; Li, X.; Feng, J.; Hou, J.; Li, M.; Zhang, B. Ginsenoside Rg1 Reduces β-Amyloid Levels by Inhibiting CDK5-Induced PPARγ Phosphorylation in a Neuron Model of Alzheimer’s Disease. Mol. Med. Rep. 2020, 22, 3277–3288. [Google Scholar] [CrossRef] [PubMed]
  53. Ke, C.; Peng, Y.; Yuan, Z.; Cai, J. Ginsenoside Rb1 Protected PC12 Cells from Aβ25-35-Induced Cytotoxicity via PPARγ Activation and Cholesterol Reduction. Eur. J. Pharmacol. 2021, 893, 173835. [Google Scholar] [CrossRef] [PubMed]
  54. Karpagam, V.; Sathishkumar, N.; Sathiyamoorthy, S.; Rasappan, P.; Shila, S.; Kim, Y.-J.; Yang, D.-C. Identification of BACE1 Inhibitors from Panax Ginseng Saponins-An Insilco Approach. Comput. Biol. Med. 2013, 43, 1037–1044. [Google Scholar] [CrossRef]
  55. Yun, Y.-J.; Park, B.-H.; Hou, J.; Oh, J.-P.; Han, J.-H.; Kim, S.-C. Ginsenoside F1 Protects the Brain against Amyloid Beta-Induced Toxicity by Regulating IDE and NEP. Life 2022, 12, 58. [Google Scholar] [CrossRef] [PubMed]
  56. Yang, L.; Hao, J.; Zhang, J.; Xia, W.; Dong, X.; Hu, X.; Kong, F.; Cui, X. Ginsenoside Rg3 Promotes Beta-Amyloid Peptide Degradation by Enhancing Gene Expression of Neprilysin. J. Pharm. Pharmacol. 2009, 61, 375–380. [Google Scholar] [CrossRef] [PubMed]
  57. Zhang, H.; Su, Y.; Sun, Z.; Chen, M.; Han, Y.; Li, Y.; Dong, X.; Ding, S.; Fang, Z.; Li, W.; et al. Ginsenoside Rg1 Alleviates Aβ Deposition by Inhibiting NADPH Oxidase 2 Activation in APP/PS1 Mice. J. Ginseng Res. 2021, 45, 665–675. [Google Scholar] [CrossRef] [PubMed]
  58. Mi, L.; Fan, M.; Liu, T.; Wu, D.; Wang, Y.; Li, F.; Cai, Y.; Qiu, Z.; Liu, D.; Cao, L. Ginsenoside Rd Protects Transgenic Caenorhabditis Elegans from β-Amyloid Toxicity by Activating Oxidative Resistant. Front. Pharmacol. 2022, 13, 1074397. [Google Scholar] [CrossRef]
  59. Cao, G.; Su, P.; Zhang, S.; Guo, L.; Zhang, H.; Liang, Y.; Qin, C.; Zhang, W. Ginsenoside Re Reduces Aβ Production by Activating PPARγ to Inhibit BACE1 in N2a/APP695 Cells. Eur. J. Pharmacol. 2016, 793, 101–108. [Google Scholar] [CrossRef]
  60. Choi, R.J.; Roy, A.; Jung, H.J.; Ali, M.Y.; Min, B.-S.; Park, C.H.; Yokozawa, T.; Fan, T.-P.; Choi, J.S.; Jung, H.A. BACE1 Molecular Docking and Anti-Alzheimer’s Disease Activities of Ginsenosides. J. Ethnopharmacol. 2016, 190, 219–230. [Google Scholar] [CrossRef]
  61. Guo, X.; Lv, J.; Lu, J.; Fan, L.; Huang, X.; Hu, L.; Wang, J.; Shen, X. Protopanaxadiol Derivative DDPU Improves Behavior and Cognitive Deficit in AD Mice Involving Regulation of Both ER Stress and Autophagy. Neuropharmacology 2018, 130, 77–91. [Google Scholar] [CrossRef] [PubMed]
  62. Ahn, J.W.; Jang, S.K.; Jo, B.R.; Kim, H.S.; Park, J.Y.; Park, H.Y.; Yoo, Y.-M.; Joo, S.S. A Therapeutic Intervention for Alzheimer’s Disease Using Ginsenoside Rg3: Its Role in M2 Microglial Activation and Non-Amyloidogenesis. J. Physiol. Pharmacol. 2021, 72, 185–193. [Google Scholar] [CrossRef]
  63. Zhang, S.; Liu, F.; Li, J.; Jing, C.; Lu, J.; Chen, X.; Wang, D.; Cao, D.; Zhao, D.; Sun, L. A 4.7-kDa Polysaccharide from Panax Ginseng Suppresses Aβ Pathology via Mitophagy Activation in Cross-Species Alzheimer’s Disease Models. Biomed. Pharmacother. 2023, 167, 115442. [Google Scholar] [CrossRef] [PubMed]
  64. 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. [Google Scholar] [CrossRef] [PubMed]
  65. Wu, J.-J.; Yang, Y.; Wan, Y.; Xia, J.; Xu, J.-F.; Zhang, L.; Liu, D.; Chen, L.; Tang, F.; Ao, H.; et al. New Insights into the Role and Mechanisms of Ginsenoside Rg1 in the Management of Alzheimer’s Disease. Biomed. Pharmacother. 2022, 152, 113207. [Google Scholar] [CrossRef] [PubMed]
  66. Wani, A.; Gupta, M.; Ahmad, M.; Shah, A.M.; Ahsan, A.U.; Qazi, P.H.; Malik, F.; Singh, G.; Sharma, P.R.; Kaddoumi, A.; et al. Alborixin Clears Amyloid-β by Inducing Autophagy through PTEN-Mediated Inhibition of the AKT Pathway. Autophagy 2019, 15, 1810–1828. [Google Scholar] [CrossRef] [PubMed]
  67. Liu, M.; Bai, X.; Yu, S.; Zhao, W.; Qiao, J.; Liu, Y.; Zhao, D.; Wang, J.; Wang, S. Ginsenoside Re Inhibits ROS/ASK-1 Dependent Mitochondrial Apoptosis Pathway and Activation of Nrf2-Antioxidant Response in Beta-Amyloid-Challenged SH-SY5Y Cells. Molecules 2019, 24, 2687. [Google Scholar] [CrossRef]
  68. Liu, Y.; Gao, Y.; Li, K.-X.; Xue, W. Pharmacokinetics and Acetylcholine Releasing Effects of Ginsenoside Rg1 in Hippocampus of Beta-Amyloid Model Rats. J. Asian Nat. Prod. Res. 2019, 21, 772–781. [Google Scholar] [CrossRef]
  69. Soliman, A.; Bakota, L.; Brandt, R. Microtubule-Modulating Agents in the Fight Against Neurodegeneration: Will It Ever Work? Curr. Neuropharmacol. 2022, 20, 782–798. [Google Scholar] [CrossRef]
  70. Ossenkoppele, R.; van der Kant, R.; Hansson, O. Tau Biomarkers in Alzheimer’s Disease: Towards Implementation in Clinical Practice and Trials. Lancet Neurol. 2022, 21, 726–734. [Google Scholar] [CrossRef]
  71. Shukla, V.; Skuntz, S.; Pant, H.C. Deregulated Cdk5 Activity Is Involved in Inducing Alzheimer’s Disease. Arch. Med. Res. 2012, 43, 655–662. [Google Scholar] [CrossRef] [PubMed]
  72. Zhao, H.; Di, J.; Liu, W.; Liu, H.; Lai, H.; Lü, Y. Involvement of GSK3 and PP2A in Ginsenoside Rb1′s Attenuation of Aluminum-Induced Tau Hyperphosphorylation. Behav. Brain Res. 2013, 241, 228–234. [Google Scholar] [CrossRef] [PubMed]
  73. 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] [PubMed]
  74. Li, X.; Liu, Y.; Yuan, H.-F.; Quan, Q.-K. Effects of gensenoside Rg1 on tau protein phosphorylation induced by okadaic acid in rat brain slices. J. Chin. Integr. Med. 2010, 8, 955–960. [Google Scholar] [CrossRef] [PubMed]
  75. Chen, X.; Huang, T.; Zhang, J.; Song, J.; Chen, L.; Zhu, Y. Involvement of Calpain and P25 of CDK5 Pathway in Ginsenoside Rb1′s Attenuation of Beta-Amyloid Peptide25-35-Induced Tau Hyperphosphorylation in Cortical Neurons. Brain Res. 2008, 1200, 99–106. [Google Scholar] [CrossRef] [PubMed]
  76. Zhang, Y.; Pi, Z.; Song, F.; Liu, Z. Ginsenosides Attenuate D-Galactose- and AlCl3-Inducedspatial Memory Impairment by Restoring the Dysfunction of the Neurotransmitter Systems in the Rat Model of Alzheimer’s Disease. J. Ethnopharmacol. 2016, 194, 188–195. [Google Scholar] [CrossRef]
  77. She, L.; Xiong, L.; Li, L.; Zhang, J.; Sun, J.; Wu, H.; Ren, J.; Wang, W.; Zhao, X.; Liang, G. Ginsenoside Rk3 Ameliorates Aβ-Induced Neurotoxicity in APP/PS1 Model Mice via AMPK Signaling Pathway. Biomed. Pharmacother. 2023, 158, 114192. [Google Scholar] [CrossRef] [PubMed]
  78. Kandimalla, R.; Reddy, P.H. Therapeutics of Neurotransmitters in Alzheimer’s Disease. J. Alzheimer’s Dis. 2017, 57, 1049–1069. [Google Scholar] [CrossRef]
  79. 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. [Google Scholar] [CrossRef]
  80. Yang, Y.; Liang, X.; Jin, P.; Li, N.; Zhang, Q.; Yan, W.; Zhang, H.; Sun, J. Screening and Determination for Potential Acetylcholinesterase Inhibitory Constituents from Ginseng Stem–Leaf Saponins Using Ultrafiltration (UF)-LC-ESI-MS2. Phytochem. Anal. 2019, 30, 26–33. [Google Scholar] [CrossRef]
  81. Lu, C.; Lv, J.; Dong, L.; Jiang, N.; Wang, Y.; Wang, Q.; Li, Y.; Chen, S.; Fan, B.; Wang, F.; et al. Neuroprotective Effects of 20(S)-Protopanaxatriol (PPT) on Scopolamine-Induced Cognitive Deficits in Mice. Phytother. Res. 2018, 32, 1056–1063. [Google Scholar] [CrossRef] [PubMed]
  82. Benishin, C.G.; Lee, R.; Wang, L.C.; Liu, H.J. Effects of Ginsenoside Rb1 on Central Cholinergic Metabolism. Pharmacology 1991, 42, 223–229. [Google Scholar] [CrossRef] [PubMed]
  83. Shi, J.; Xue, W.; Zhao, W.; Li, K. Pharmacokinetics and Dopamine/Acetylcholine Releasing Effects of Ginsenoside Re in Hippocampus and mPFC of Freely Moving Rats. Acta Pharmacol. Sin. 2013, 34, 214–220. [Google Scholar] [CrossRef] [PubMed]
  84. Kim, J.; Shim, J.; Lee, S.; Cho, W.-H.; Hong, E.; Lee, J.H.; Han, J.-S.; Lee, H.J.; Lee, K.W. Rg3-Enriched Ginseng Extract Ameliorates Scopolamine-Induced Learning Deficits in Mice. BMC Complement. Altern. Med. 2016, 16, 66. [Google Scholar] [CrossRef] [PubMed]
  85. Kim, H.-J.; Shin, E.-J.; Lee, B.-H.; Choi, S.-H.; Jung, S.-W.; Cho, I.-H.; Hwang, S.-H.; Kim, J.Y.; Han, J.-S.; Chung, C.; et al. Oral Administration of Gintonin Attenuates Cholinergic Impairments by Scopolamine, Amyloid-β Protein, and Mouse Model of Alzheimer’s Disease. Mol. Cells 2015, 38, 796–805. [Google Scholar] [CrossRef] [PubMed]
  86. 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. [Google Scholar] [CrossRef] [PubMed]
  87. Kamecki, F.; Knez, D.; Carvalho, D.; Marcucci, C.; Rademacher, M.; Higgs, J.; Žakelj, S.; Marcos, A.; de Tezanos Pinto, F.; Abin-Carriquiry, J.A.; et al. Multitarget 2′-Hydroxychalcones as Potential Drugs for the Treatment of Neurodegenerative Disorders and Their Comorbidities. Neuropharmacology 2021, 201, 108837. [Google Scholar] [CrossRef]
  88. Dou, Y.; Xie, J.; Tan, Y.; Zhang, M.; Zhao, Y.; Liu, X. Neurotransmitter-Stimulated Neuron-Derived sEVs Have Opposite Effects on Amyloid β-Induced Neuronal Damage. J. Nanobiotechnology 2021, 19, 324. [Google Scholar] [CrossRef]
  89. Khan, A.F.; Adewale, Q.; Baumeister, T.R.; Carbonell, F.; Zilles, K.; Palomero-Gallagher, N.; Iturria-Medina, Y. Personalized Brain Models Identify Neurotransmitter Receptor Changes in Alzheimer’s Disease. Brain 2021, 145, 1785–1804. [Google Scholar] [CrossRef]
  90. Ionescu-Tucker, A.; Cotman, C.W. Emerging Roles of Oxidative Stress in Brain Aging and Alzheimer’s Disease. Neurobiol. Aging 2021, 107, 86–95. [Google Scholar] [CrossRef]
  91. Ozben, T.; Ozben, S. Neuro-Inflammation and Anti-Inflammatory Treatment Options for Alzheimer’s Disease. Clin. Biochem. 2019, 72, 87–89. [Google Scholar] [CrossRef] [PubMed]
  92. Zhao, Y.; Chen, J.; Xiao, H.; Tian, Y.; Li, H.; Tian, J.; Yang, J. Study on the Inhibition of Oxidative Stress Injury in H2O2-induced N2a Cells by Ginsenoside Rg1 via Activation of Nrf2/HO-1 Signaling Pathway. Res. Pract. Chin. Med. 2022, 36, 26–31. [Google Scholar] [CrossRef]
  93. Chen, L.; Yao, H.; Chen, X.; Wang, Z.; Xiang, Y.; Xia, J.; Liu, Y.; Wang, Y. Ginsenoside Rg1 Decreases Oxidative Stress and Down-Regulates Akt/mTOR Signalling to Attenuate Cognitive Impairment in Mice and Senescence of Neural Stem Cells Induced by d-Galactose. Neurochem. Res. 2018, 43, 430–440. [Google Scholar] [CrossRef] [PubMed]
  94. Zhu, Y.; Wang, Z.; Yu, S.; Zhao, C.; Xu, B.; Liu, R.; Xu, L.; Guo, Y. Neuroprotective Effect of Ginseng Fibrous Root Enzymatic Hydrolysate against Oxidative Stress. Molecules 2022, 27, 7824. [Google Scholar] [CrossRef] [PubMed]
  95. Zhang, Y.; Hu, W.; Zhang, B.; Yin, Y.; Zhang, J.; Huang, D.; Huang, R.; Li, W.; Li, W. Ginsenoside Rg1 Protects against Neuronal Degeneration Induced by Chronic Dexamethasone Treatment by Inhibiting NLRP-1 Inflammasomes in Mice. Int. J. Mol. Med. 2017, 40, 1134–1142. [Google Scholar] [CrossRef] [PubMed]
  96. Lee, B.; Sur, B.; Park, J.; Kim, S.-H.; Kwon, S.; Yeom, M.; Shim, I.; Lee, H.; Hahm, D.-H. Ginsenoside Rg3 Alleviates Lipopolysaccharide-Induced Learning and Memory Impairments by Anti-Inflammatory Activity in Rats. Biomol. Ther. 2013, 21, 381–390. [Google Scholar] [CrossRef] [PubMed]
  97. Xu, M.; Chen, Q.; Fan, R.; Wang, J.; Li, Y. Anti-Inflammation Effect of Small Molecule Oligopeptides Prepared from Panax Ginseng C. A. Meyer in Rats. Molecules 2019, 24, 858. [Google Scholar] [CrossRef]
  98. Piaceri, I.; Rinnoci, V.; Bagnoli, S.; Failli, Y.; Sorbi, S. Mitochondria and Alzheimer’s Disease. J. Neurol. Sci. 2012, 322, 31–34. [Google Scholar] [CrossRef]
  99. Wu, J.; Yang, H.; Zhao, Q.; Zhang, X.; Lou, Y. Ginsenoside Rg1 Exerts a Protective Effect against Aβ25-35-Induced Toxicity in Primary Cultured Rat Cortical Neurons through the NF-κB/NO Pathway. Int. J. Mol. Med. 2016, 37, 781–788. [Google Scholar] [CrossRef]
  100. Shim, J.S.; Song, M.-Y.; Yim, S.-V.; Lee, S.-E.; Park, K.-S. Global Analysis of Ginsenoside Rg1 Protective Effects in β-Amyloid-Treated Neuronal Cells. J. Ginseng Res. 2017, 41, 566. [Google Scholar] [CrossRef]
  101. Pradeepkiran, J.A.; Reddy, P.H. Defective Mitophagy in Alzheimer’s Disease. Ageing Res. Rev. 2020, 64, 101191. [Google Scholar] [CrossRef] [PubMed]
  102. Wang, X.; Qi, L.; Cheng, Y.; Ji, X.; Chi, T.; Liu, P.; Zou, L. PINK1 Overexpression Prevents Forskolin-Induced Tau Hyperphosphorylation and Oxidative Stress in a Rat Model of Alzheimer’s Disease. Acta Pharmacol. Sin. 2022, 43, 1916–1927. [Google Scholar] [CrossRef] [PubMed]
  103. Grabrucker, S.; Marizzoni, M.; Silajdžić, E.; Lopizzo, N.; Mombelli, E.; Nicolas, S.; Dohm-Hansen, S.; Scassellati, C.; Moretti, D.V.; Rosa, M.; et al. Microbiota from Alzheimer’s Patients Induce Deficits in Cognition and Hippocampal Neurogenesis. Brain 2023, 146, 4916–4934. [Google Scholar] [CrossRef] [PubMed]
  104. Chen, H.; Shen, J.; Li, H.; Zheng, X.; Kang, D.; Xu, Y.; Chen, C.; Guo, H.; Xie, L.; Wang, G.; et al. Ginsenoside Rb1 Exerts Neuroprotective Effects through Regulation of Lactobacillus Helveticus Abundance and GABAA Receptor Expression. J. Ginseng Res. 2020, 44, 86–95. [Google Scholar] [CrossRef] [PubMed]
  105. Wang, L.; Lu, J.; Zeng, Y.; Guo, Y.; Wu, C.; Zhao, H.; Zheng, H.; Jiao, J. Improving Alzheimer’s Disease by Altering Gut Microbiota in Tree Shrews with Ginsenoside Rg1. FEMS Microbiol. Lett. 2020, 367, fnaa011. [Google Scholar] [CrossRef] [PubMed]
  106. Huang, J.; Liu, D.; Wang, Y.; Liu, L.; Li, J.; Yuan, J.; Jiang, Z.; Jiang, Z.; Hsiao, W.W.; Liu, H.; et al. Ginseng Polysaccharides Alter the Gut Microbiota and Kynurenine/Tryptophan Ratio, Potentiating the Antitumour Effect of Antiprogrammed Cell Death 1/Programmed Cell Death Ligand 1 (Anti-PD-1/PD-L1) Immunotherapy. Gut 2022, 71, 734–745. [Google Scholar] [CrossRef] [PubMed]
  107. Wang, J.; He, M.; Guo, W.; Zhang, Y.; Sui, X.; Lin, J.; Liu, X.; Li, H.; Li, J.; Yang, Q.; et al. Microbiome-Metabolomics Reveals Endogenous Alterations of Energy Metabolism by the Dushen Tang to Attenuate D-Galactose-Induced Memory Impairment in Rats. Biomed. Res. Int. 2021, 2021, 6649085. [Google Scholar] [CrossRef]
  108. Janeiro, M.H.; Ramírez, M.J.; Solas, M. Dysbiosis and Alzheimer’s Disease: Cause or Treatment Opportunity? Cell. Mol. Neurobiol. 2022, 42, 377–387. [Google Scholar] [CrossRef]
  109. Narengaowa; Kong, W.; Lan, F.; Awan, U.F.; Qing, H.; Ni, J. The Oral-Gut-Brain AXIS: The Influence of Microbes in Alzheimer’s Disease. Front. Cell. Neurosci. 2021, 15, 633735. [Google Scholar] [CrossRef]
  110. Li, Z.; Zhu, H.; Guo, Y.; Du, X.; Qin, C. Gut Microbiota Regulate Cognitive Deficits and Amyloid Deposition in a Model of Alzheimer’s Disease. J. Neurochem. 2020, 155, 448–461. [Google Scholar] [CrossRef]
  111. Shabbir, U.; Arshad, M.S.; Sameen, A.; Oh, D.-H. Crosstalk between Gut and Brain in Alzheimer’s Disease: The Role of Gut Microbiota Modulation Strategies. Nutrients 2021, 13, 690. [Google Scholar] [CrossRef] [PubMed]
  112. Panza, F.; Lozupone, M.; Solfrizzi, V.; Watling, M.; Imbimbo, B.P. Time to Test Antibacterial Therapy in Alzheimer’s Disease. Brain 2019, 142, 2905–2929. [Google Scholar] [CrossRef] [PubMed]
  113. Sochocka, M.; Donskow-Łysoniewska, K.; Diniz, B.S.; Kurpas, D.; Brzozowska, E.; Leszek, J. The Gut Microbiome Alterations and Inflammation-Driven Pathogenesis of Alzheimer’s Disease—A Critical Review. Mol. Neurobiol. 2019, 56, 1841–1851. [Google Scholar] [CrossRef] [PubMed]
  114. Benakis, C.; Martin-Gallausiaux, C.; Trezzi, J.-P.; Melton, P.; Liesz, A.; Wilmes, P. The Microbiome-Gut-Brain Axis in Acute and Chronic Brain Diseases. Curr. Opin. Neurobiol. 2020, 61, 1–9. [Google Scholar] [CrossRef]
  115. Leblhuber, F.; Steiner, K.; Schuetz, B.; Fuchs, D.; Gostner, J.M. Probiotic Supplementation in Patients with Alzheimer’s DementiaAn Explorative Intervention Study. Curr. Alzheimer Res. 2018, 15, 1106–1113. [Google Scholar] [CrossRef]
  116. De, J.R. De-Paula, V.; Forlenza, A.S.; Forlenza, O.V. Relevance of Gutmicrobiota in Cognition, Behaviour and Alzheimer’s Disease. Pharmacol. Res. 2018, 136, 29–34. [Google Scholar] [CrossRef]
  117. Garcez, M.L.; Jacobs, K.R.; Guillemin, G.J. Microbiota Alterations in Alzheimer’s Disease: Involvement of the Kynurenine Pathway and Inflammation. Neurotox. Res. 2019, 36, 424–436. [Google Scholar] [CrossRef]
  118. Kesika, P.; Suganthy, N.; Sivamaruthi, B.S.; Chaiyasut, C. Role of Gut-Brain Axis, Gut Microbial Composition, and Probiotic Intervention in Alzheimer’s Disease. Life Sci. 2021, 264, 118627. [Google Scholar] [CrossRef]
  119. Scheltens, P.; Strooper, B.D.; Kivipelto, M.; Holstege, H.; Chételat, G.; Teunissen, C.E.; Cummings, J.; van der Flier, W.M. Alzheimer’s Disease. Lancet 2021, 397, 1577–1590. [Google Scholar] [CrossRef]
  120. Heo, J.-H.; Lee, S.-T.; Chu, K.; Oh, M.J.; Park, H.-J.; Shim, J.-Y.; Kim, M. Heat-Processed Ginseng Enhances the Cognitive Function in Patients with Moderately Severe Alzheimer’s Disease. Nutr. Neurosci. 2012, 15, 278–282. [Google Scholar] [CrossRef]
  121. Lee, S.-T.; Chu, K.; Sim, J.-Y.; Heo, J.-H.; Kim, M. Panax Ginseng Enhances Cognitive Performance in Alzheimer Disease. Alzheimer Dis. Assoc. Disord. 2008, 22, 222–226. [Google Scholar] [CrossRef] [PubMed]
  122. Yang, Y.; Yang, Y.; Qiu, H.; Ju, Z.; Shi, Y.; Wang, Z.; Yang, L. Localization of Constituents for Determining the Age and Parts of Ginseng through Ultraperfomance Liquid Chromatography Quadrupole/Time of Flight-Mass Spectrometry Combined with Desorption Electrospray Ionization Mass Spectrometry Imaging. J. Pharm. Biomed. Anal. 2021, 193, 113722. [Google Scholar] [CrossRef] [PubMed]
  123. Schwamborn, K.; Kriegsmann, M.; Weichert, W. MALDI Imaging Mass Spectrometry—From Bench to Bedside. Biochim. Et Biophys. Acta (BBA) Proteins Proteom. 2017, 1865, 776–783. [Google Scholar] [CrossRef] [PubMed]
  124. Connolly, L.; Jamzad, A.; Kaufmann, M.; Farquharson, C.E.; Ren, K.; Rudan, J.F.; Fichtinger, G.; Mousavi, P. Combined Mass Spectrometry and Histopathology Imaging for Perioperative Tissue Assessment in Cancer Surgery. J. Imaging 2021, 7, 203. [Google Scholar] [CrossRef]
  125. Drake, R.R.; West, C.A.; Mehta, A.S.; Angel, P.M. MALDI Mass Spectrometry Imaging of N-Linked Glycans in Tissues. In Glycobiophysics; Yamaguchi, Y., Kato, K., Eds.; Advances in Experimental Medicine and Biology; Springer: Singapore, 2018; pp. 59–76. ISBN 9789811321580. [Google Scholar]
  126. Sun, C.; Li, T.; Song, X.; Huang, L.; Zang, Q.; Xu, J.; Bi, N.; Jiao, G.; Hao, Y.; Chen, Y.; et al. Spatially Resolved Metabolomics to Discover Tumor-Associated Metabolic Alterations. Proc. Natl. Acad. Sci. USA 2019, 116, 52–57. [Google Scholar] [CrossRef] [PubMed]
  127. Wang, Z.; Fu, W.; Huo, M.; He, B.; Liu, Y.; Tian, L.; Li, W.; Zhou, Z.; Wang, B.; Xia, J.; et al. Spatial-Resolved Metabolomics Reveals Tissue-Specific Metabolic Reprogramming in Diabetic Nephropathy by Using Mass Spectrometry Imaging. Acta Pharm. Sin. B 2021, 11, 3665–3677. [Google Scholar] [CrossRef]
  128. Shariatgorji, M.; Nilsson, A.; Fridjonsdottir, E.; Vallianatou, T.; Källback, P.; Katan, L.; Sävmarker, J.; Mantas, I.; Zhang, X.; Bezard, E.; et al. Comprehensive Mapping of Neurotransmitter Networks by MALDI–MS Imaging. Nat. Methods 2019, 16, 1021–1028. [Google Scholar] [CrossRef]
  129. Huo, M.; Wang, Z.; Fu, W.; Tian, L.; Li, W.; Zhou, Z.; Chen, Y.; Wei, J.; Abliz, Z. Spatially Resolved Metabolomics Based on Air-Flow-Assisted Desorption Electrospray Ionization–Mass Spectrometry Imaging Reveals Region-Specific Metabolic Alterations in Diabetic Encephalopathy. J. Proteome Res. 2021, 20, 3567–3579. [Google Scholar] [CrossRef]
  130. Pang, X.; Gao, S.; Ga, M.; Zhang, J.; Luo, Z.; Chen, Y.; Zhang, R.; He, J.; Abliz, Z. Mapping Metabolic Networks in the Brain by Ambient Mass Spectrometry Imaging and Metabolomics. Anal. Chem. 2021, 93, 6746–6754. [Google Scholar] [CrossRef]
  131. Michno, W.; Stringer, K.M.; Enzlein, T.; Passarelli, M.K.; Escrig, S.; Vitanova, K.; Wood, J.; Blennow, K.; Zetterberg, H.; Meibom, A.; et al. Following Spatial Aβ Aggregation Dynamics in Evolving Alzheimer’s Disease Pathology by Imaging Stable Isotope Labeling Kinetics. Sci. Adv. 2021, 7, eabg4855. [Google Scholar] [CrossRef]
  132. Huang, X.; Wang, R.; Wang, Y.; Chen, C.; Liu, S. Investigation on Property Differences of Ginseng and American Ginseng by Spatial Metabolomics of Neurochemicals with Desorption Electrospray Ionization Mass Spectrometry Imaging. J. Ethnopharmacol. 2023, 303, 116006. [Google Scholar] [CrossRef] [PubMed]
  133. Zhang, X.; Wu, C.; Tan, W. Brain Lipid Dynamics in Amyloid Precursor Protein/Presenilin 1 Mouse Model of Early Alzheimer’s Disease by Desorption Electrospray Ionization and Matrix Assisted Laser Desorption Ionization–Mass Spectrometry Imaging Techniques. J. Proteome Res. 2021, 20, 2643–2650. [Google Scholar] [CrossRef] [PubMed]
  134. Nishidate, M.; Hayashi, M.; Aikawa, H.; Tanaka, K.; Nakada, N.; Miura, S.; Ryu, S.; Higashi, T.; Ikarashi, Y.; Fujiwara, Y.; et al. Applications of MALDI Mass Spectrometry Imaging for Pharmacokinetic Studies during Drug Development. Drug Metab. Pharmacokinet. 2019, 34, 209–216. [Google Scholar] [CrossRef] [PubMed]
  135. Wei, W.; Li, Z.; Li, H.; An, Y.; Qu, H.; Yao, C.; Zhang, J.; Li, J.; Zhang, G.; Shi, Y.; et al. Exploration of Tissue Distribution of Ginsenoside Rg1 by LC-MS/MS and Nanospray Desorption Electrospray Ionization Mass Spectrometry. J. Pharm. Biomed. Anal. 2021, 198, 113999. [Google Scholar] [CrossRef]
Figure 1. Active ingredients and pharmacological effects of P. ginseng.
Figure 1. Active ingredients and pharmacological effects of P. ginseng.
Pharmaceuticals 17 00038 g001
Figure 2. The chemical structures of ginsenosides (R1 and R2 refer to hydrogen or different sugar group substitution; different ginsenosides have different R1 and R2 in their structural formula).
Figure 2. The chemical structures of ginsenosides (R1 and R2 refer to hydrogen or different sugar group substitution; different ginsenosides have different R1 and R2 in their structural formula).
Pharmaceuticals 17 00038 g002
Figure 3. Signaling pathways related to Aβ accumulation.
Figure 3. Signaling pathways related to Aβ accumulation.
Pharmaceuticals 17 00038 g003
Figure 4. Signaling pathways related to tau hyperphosphorylation.
Figure 4. Signaling pathways related to tau hyperphosphorylation.
Pharmaceuticals 17 00038 g004
Figure 5. Signaling pathways related to oxidative stress.
Figure 5. Signaling pathways related to oxidative stress.
Pharmaceuticals 17 00038 g005
Figure 6. Signaling pathways related to neuroinflammation.
Figure 6. Signaling pathways related to neuroinflammation.
Pharmaceuticals 17 00038 g006
Figure 7. Signaling pathways related to mitochondrial apoptosis.
Figure 7. Signaling pathways related to mitochondrial apoptosis.
Pharmaceuticals 17 00038 g007
Table 1. The effects of P. ginseng on the pathogenesis of AD related to Aβ.
Table 1. The effects of P. ginseng on the pathogenesis of AD related to Aβ.
P. ginseng CompoundsExperimental ModelEffectsRef.
Ginsenoside CKICR miceSuppress BACE1 and PS1; enhance IDE[51]
Ginsenoside Rg1Sprague-Dawley ratsSuppress APP, CDK5, P-PPARγ, and BACE1; enhance IDE[52]
Ginsenoside Rb1PC12 pheochromocytoma cellsEnhance PPARγ[53]
Ginsenoside CK, F1, Rh1 and Rh2Molecular dockingSuppress BACE1[54]
Ginsenoside F1Mouse neuroblastoma neuro-2a (N2a) cells
and human neuroblastoma SH-SY5Y cells
Enhance IDE and NEP[55]
Ginsenoside Rg3SK-N-SH cellsEnhance NEP[56]
Ginsenoside Rg1Wild type (WT) and APP/PS1 AD miceSuppress NOX2, ROS, APP, and BACE[57]
Ginsenoside RdC. elegansSuppress ROS and MAPK[58]
Ginsenoside ReN2a/APP695 cellsEnhance PPARγ; suppress BACE1[59]
Ginsenoside Rb1, Rb2Molecular dockingSuppress BACE1[60]
DDPUSH-SY5Y cells, APP/PS1 transgenic mice, CHO-APP cells, HEK293-APP swe cells, primary neurons, or primary astrocytesSuppress BACE1, PERK, P-eIF2 α, PI3K, P-AKT, P-mTOR, and P-ULK1[61]
Ginsenoside Rg3Neuro-2a (N2a) murine neuroblastoma
and HMO6 human microglial cells
Enhance SRA[62]
A 4.7-kDa ginseng-derived polysaccharide (GP4)SH-SY5Y cells and C. elegansEnhance PINK1 and Parkin[63]
combined ginsenosides (SumI)AD wormsEnhance HSPs and HSF-1[26]
Ginsenoside F1APPswe/PSEN1dE9 double-transgenic AD mice
with a B6 × C3 background and B6 × C3 wild type mice
Enhance P-CREB and BDNF[64]
Table 2. The effects of P. ginseng on the pathogenesis of AD related to tau proteins.
Table 2. The effects of P. ginseng on the pathogenesis of AD related to tau proteins.
P. ginseng CompoundsExperimental ModelEffectsRef.
Ginsenoside Rb1ICR miceSuppress pGSK-3; enhance PP2A[72]
Ginsenoside RdSD ratsEnhance PP2A[73]
Ginsenoside Rg1Wistar ratsEnhance PP2A[74]
Ginsenoside Rb1SD ratsSuppress calpain, CDK5, and P25; enhance P35[75]
Table 3. The effects of P. ginseng on the pathogenesis of AD related to neurotransmitters.
Table 3. The effects of P. ginseng on the pathogenesis of AD related to neurotransmitters.
P. ginseng CompoundsExperimental ModelEffectsRef.
Ginsenoside Rh2Male ICR miceEnhance ACh[79]
Ginsenoside Rg1Adult male Sprague-Dawley ratsEnhance ACh[68]
GSLSCellSuppress AChE[80]
PPTMale Institute of Cancer Research miceEnhance ACh[81]
Ginsenoside Re, Rg3In-vitro enzyme assaysSuppress AChE and BChE[60]
Ginsenoside Rb1Male Sprague-Dawley ratsEnhance ACh[82]
Ginsenoside ReAdults male Sprague-Dawley ratsEnhance DA and ACh[83]
Ginsenoside Rg3Male C57BL/6 miceSuppress AChE[84]
GintoninMale ICR or C57BL/6 miceEnhance ACh and ChAT; suppress AChE[85]
GinsenosidesMale Wistar ratsEnhance GABA, ACh, DA, Gly, and 5-HT; suppress Glu and Asp[76]
Ginsenoside Rg3Male Wistar ratsImprove Ala, Asp, Glu, D-Gln, D-Glu, and Trpab normal metabolism[86]
Table 4. The effects of P. ginseng on the pathogenesis of AD related to oxidative stress and neuroinflammation.
Table 4. The effects of P. ginseng on the pathogenesis of AD related to oxidative stress and neuroinflammation.
P. ginseng CompoundsExperimental ModelEffectsRef.
Ginsenoside CKICR miceEnhance Nrf2, keap1, HO-1, SOD, and GSH-Px; suppress MDA[51]
Ginsenoside Rk3APP/PS1 mice, PC12 cellsSuppress ROS; enhance P-AMPK, Nrf2, HO-1, NQO1, and Keap1[77]
Ginsenoside Rg1Mouse neuroblastoma N2a cellsEnhance Nrf2, HO-1, and SOD; suppress MDA and LDH[92]
Ginsenoside Rg1Male APP/PS1 miceSuppress ROS and NOX2[57]
Ginsenoside Rg1C57BL/6 mice, mice NSCsEnhance SOD and GSH-Px; suppress MDA, ROS, Akt, and mTOR[93]
Ginsenoside Rg3Male Wistar ratsEnhance SOD, CAT, and GSH-Px; suppress MDA and ROS[86]
Ginsenoside Rh2Male ICR miceEnhance SOD and GSH; suppress MDA[79]
PPTMale Institute of Cancer Research miceEnhance SOD; suppress MDA[81]
GFREHC. elegansEnhance SOD and CAT; suppress MDA, LDH and ROS,[94]
Ginsenoside Rg1Male ICR miceSuppress NLRP1 and proinflammatory cytokines (IL-1β and IL-18)[95]
Ginsenoside Rg3Male Sprague-Dawley (SD) ratsSuppress proinflammatory cytokines (TNF-α, IL-1β, and COX-2)[96]
Ginsenoside Rg3Neuro-2a (N2a) murine neuroblastoma and HMO6 human microglial cellsSuppress proinflammatory cytokines (iNOS, IL-6, and TNF-α)[62]
Ginseng oligopeptides (GOPs)Male Sprague-Dawley (SD) ratsSuppress MAPK, NF-κ B, and proinflammatory cytokines (TNF-α IL-β)[97]
Table 5. The effects of P. ginseng on the pathogenesis of AD related to mitochondrial apoptosis.
Table 5. The effects of P. ginseng on the pathogenesis of AD related to mitochondrial apoptosis.
P. ginseng CompoundsExperimental ModelEffectsRef.
Ginsenoside CKICR miceEnhance Bcl-2; suppress Bax and caspase-3[51]
Ginsenoside Rg3Male Wistar ratsImprove ETC and ATP; enhance Bcl-2 Suppress Cyt C, Bax, and ROS[86]
Ginsenoside ReSH-SY5Y human neuroblastoma cellsSuppress Bax, ROS, ASK-1, and JNK[67]
Ginsenoside Rg1Primary cultured cortical neurons were prepared
from embryonic day (D17-18) Sprague-Dawley (SD) rat fetuses
Enhance Bcl-2; suppress NF-κB, NO, ROS, and Bax[99]
Ginsenoside Rg1SH-SY5Y cellsEnhance HSD17B10, AARS2, and TOMM40; suppress NDUFA4, VDAC1, and COX5A[100]
Table 6. The effects of P. ginseng on the pathogenesis of AD related to gut microbiota.
Table 6. The effects of P. ginseng on the pathogenesis of AD related to gut microbiota.
P. ginseng Compounds and Classic Chinese FormulationExperimental ModelEffectsRef.
Ginsenoside Rb1Sprague-Dawley ratsEnhance GABAA (α2, β2, and γ2), GABAB (1b and 2), Bifidobacterium longum, Bifidobacterium dentium, Lactobacillus brevis, Lactobacillus helveticus, and Lactobacillus rhamnosus[104]
Ginsenoside Rg1Male conventional tree shrewsAlter the composition and abundance of gut microbiota[105]
Ginseng polysaccharides (GPs)C57BL/6J miceIncrease valeric acid and reduce L-canine uric acid and Kyn/Trp[106]
Dushen TangAdult male Sprague-Dawley ratsCorrect the disturbance of the gut microbiota[107]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, M.; Niu, H.; Li, Q.; Jiao, L.; Li, H.; Wu, W. Active Compounds of Panax ginseng in the Improvement of Alzheimer’s Disease and Application of Spatial Metabolomics. Pharmaceuticals 2024, 17, 38. https://doi.org/10.3390/ph17010038

AMA Style

Zhang M, Niu H, Li Q, Jiao L, Li H, Wu W. Active Compounds of Panax ginseng in the Improvement of Alzheimer’s Disease and Application of Spatial Metabolomics. Pharmaceuticals. 2024; 17(1):38. https://doi.org/10.3390/ph17010038

Chicago/Turabian Style

Zhang, Meng, Huazhou Niu, Qingqing Li, Lili Jiao, Hui Li, and Wei Wu. 2024. "Active Compounds of Panax ginseng in the Improvement of Alzheimer’s Disease and Application of Spatial Metabolomics" Pharmaceuticals 17, no. 1: 38. https://doi.org/10.3390/ph17010038

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