Next Article in Journal
New Tricholidic Acid Triterpenoids from the Mushroom Tricholoma ustaloides Collected in an Italian Beech Wood
Next Article in Special Issue
The Metschnikowia pulcherrima Clade as a Model for Assessing Inhibition of Candida spp. and the Toxicity of Its Metabolite, Pulcherrimin
Previous Article in Journal
Bioactive Compounds and Sensory Analysis of Freeze-Dried Prickly Pear Fruits from an Inter-Andean Valley in Peru
Previous Article in Special Issue
Do Ganoderma Species Represent Novel Sources of Phenolic Based Antimicrobial Agents?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Immunomodulatory, Anti-Inflammatory, and Anti-Cancer Properties of Ginseng: A Pharmacological Update

by
Jose Antonio Valdés-González
,
Marta Sánchez
*,
Ignacio Moratilla-Rivera
,
Irene Iglesias
and
María Pilar Gómez-Serranillos
*
Departamento de Farmacología, Farmacognosia y Botánica, Facultad de Farmacia, Universidad Complutense de Madrid, Plaza Ramón y Cajal s/n, Ciudad Universitaria, 28040 Madrid, Spain
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(9), 3863; https://doi.org/10.3390/molecules28093863
Submission received: 28 February 2023 / Revised: 25 April 2023 / Accepted: 30 April 2023 / Published: 3 May 2023
(This article belongs to the Special Issue Natural Product-Inspired Molecules: From Weed to Remedy)

Abstract

:
Ginseng, a medicinal plant of the genus Panax, boasts a rich historical record of usage that dates back to the Paleolithic period. This botanical is extensively acknowledged and consumed in Eastern countries for its therapeutic properties, and, in Western countries, it is becoming increasingly popular as a remedy for fatigue and asthenia. This review provides an update on current research pertaining to ginseng and its isolated compounds, namely, ginsenosides and polysaccharides. The primary focus is on three crucial pharmacological activities, namely, immunomodulation, anti-inflammatory, and anti-cancer effects. The review encompasses studies on both isolated compounds and various ginseng extracts obtained from the root, leaves, and berries.

1. Introduction

Ginseng, a highly valued herbal medicine, has been used for over 5000 years, predominantly in Far East countries. Emperor Shen-Nung of China was the first to identify ginseng, along with several other herbs, as having medicinal properties, and developed what is considered the first pharmacopoeia in history [1]. Ginseng is the common name for the root of various species of the genus Panax (P.), belonging to the Araliacea family. The genus name, Panax, is derived from the ancient Greek “Panákeia” (Πανάκεια), which translates to “all healing” or “cure of all diseases.” Similarly, the term “ginseng” originates from the Chinese “jen-shen”, which means “plant-man”, possibly referencing the human form of the root [2,3]. The Greek term “panacea”, which shares the same etymological roots as “panax”, also implies that the ginseng root is a genuine and authentic panacea.
Archaeological research indicates that the ginseng root has been used as a medicinal plant since the Paleolithic period, dating back approximately 60,000 years ago [4]. The use of ginseng as a medicinal plant has been predominantly concentrated in China and other Asian countries, where it grows naturally. In East Asian countries, ginseng has been used as an adaptogenic drug since its discovery in the Manchurian mountains, more than 5000 years ago [5]. The Shen-Nung Benchau Jing treatise is considered the oldest pharmacopoeia in the world, and the use of ginseng is already mentioned in it, revealing that its pharmacological potential was already known [6].
Different species of ginseng have been used for centuries, due to their anti-inflammatory properties, which are closely associated with their ability to modulate the immune system. The chemical structure of the main active compounds, ginsenosides, partly explains this activity [7]. The chemical structure of ginsenosides is very similar to the structure of hydrocortisone, the most significant endogenous anti-inflammatory molecule [7,8,9] as shown in Figure 1. Both hydrocortisone and most ginsenosides consist of three aromatic rings that form a phenanthrene molecule linked to a cyclopentane to create a skeleton of cyclopentane-perhydrophenantrene. But this structure, by itself, does not justify the pharmacological activity of ginsenosides. This activity depends on the substituents and unsaturations of each ginsenoside, as we will see later. Thus, the anti-inflammatory activity varies depending on the ginsenoside studied, according to its chemical structure. Many ginsenosides have shown immunomodulatory properties, as well as polysaccharides associated with these ginsenosides. The primary immunomodulatory activity is due to these polysaccharides. One of the mechanisms by which ginseng exerts its immunomodulatory activity is by increasing nitric oxide (NO) production and a nonspecific stimulation of lymphocyte proliferation [10]. A recent study has investigated in depth the immunomodulatory capacity of the polysaccharides of North American ginseng (P. quinquefolius), using the crude extract from the root, as well as partially purified and completely isolated polysaccharides. However, the limited information available about the structure of these polysaccharides poses a challenge in determining the relationship between structure and activity, although some mechanisms have been proposed [11]. Furthermore, ginseng compounds, such as ginsenosides, polysaccharides, and alkaloids, can exert cancer-protective activity by inhibiting cell proliferation through the regulation of cell cycles and apoptosis.
Currently, there is a growing trend in the use of herbal medicinal products for the treatment of various diseases [12]. This increase can be attributed to various factors, including societal perception of these treatments as being healthier than conventional drugs. Additionally, there is a common misconception that herbal medicines present fewer adverse effects, and therefore are perceived as more “natural” treatments [13]. Furthermore, research on natural products is of great importance at present. Many plant families, genera, and species remain unexplored, presenting a potential source of molecules that could serve as appropriate treatment for diseases such as cancer, Alzheimer’s, or Parkinson’s [4,12]. Consumption habits of ginseng vary greatly from country to country, but, globally, it is valued at over two billion dollars. Ginseng is often consumed together with other natural products, such as royal jelly, vitamins, and minerals, but it can also be consumed alone [5]. All abbreviations used are listed in the Abbreviation.

2. Phytochemical Characteristic

The genus Panax consists of eleven species, including P. ginseng, P. notoginseng, P. quinquefolius, P. pseudoginseng, P. trifolius, P. zingiberensis, P. stipuleanatus, P. japonicus, P. japonicus var. angustifolius, P. japonicus var. major, and P. japonicus var. bipinnatifidus [14]. Morphologically, they are characterized by herbaceous perennial plants with palmate leaves with serrated margins. Generally, ginseng species reach 60–80 cm in height, and produce flowers ranging from white to purple, these flowers appear in an umbel at the apex. In addition, these species have a complex root system, comprising of a short rhizome and many tuberous roots [15]. While the most used species for their pharmacological properties are P. ginseng, P. notoginseng, and P. quinquefolius, all these species contain bioactive compounds of pharmacological interest [16].
Ginsenosides, the primary bioactive compounds found in ginseng, can be classified into five groups, based on their distinct chemical structures: protopanaxadiol, protopanaxatriol, ocotillo, oleanolic acid, and C-17 side-chain type [17]. The structural diversity of ginsenosides is attributed to the presence of various radicals and sugar moieties. Furthermore, the primary structure of some ginsenosides can be modified to produce secondary ginsenosides through natural microbiota [18], or by various processes during the preparation of the phytotherapeutic product, such as heating or drying. Some of these molecules can be seen in Figure 2. Additionally, the composition of each ginseng preparation can be influenced by factors such as growth temperature, altitude, and weather conditions, in the different places of cultivation [19]. Although ginsenosides are the most extensively studied molecules among the components of ginseng, other bioactive compounds, such as polysaccharides, glycolipoproteins, and alkaloids, also exhibit important pharmacological activities.
The structure/activity relationship of ginsenosides depends on each molecule, as well as on each specific bioactivity. In any case, for most studies, the molecular mechanisms are not deeply elucidated, but focus on the consequences of these mechanisms. For example, we will see that oleanane-type ginsenosides possess greater immunomodulatory activity than dammarane-type ginsenosides, or that the anti-cancer activity of ginsenosides increases the lower the number of sugar residues in the molecule. However, the precise molecular mechanism underlying these biological consequences remains, in most cases, unelucidated.
Chromatographic methods, such as High-Performance Liquid Chromatography (HPLC), Thin-Layer Chromatography/High-Performance Liquid Chromatography (TLC/HPLC), Ultra-High-Performance Liquid Chromatography (UHPLC), and Two-Dimensional High-Performance Liquid Chromatography (HPLC 2D), are commonly employed, in combination with spectroscopic analysis, to determine the molecular composition of different species of the Panax genus. Using these techniques, researchers have discovered up to 623 types of ginsenosides in the ethanol extract of the most used species, of which 437 are potentially novel ginsenosides [20]. Additionally, 945 ginsenosides, and 662 potentially novel ginsenosides, have been identified from P. notoginseng leaves [21,22].

3. Immunomodulatory Activity

The immune system plays a vital role in defending the body against external threats. In humans, it is comprised of two distinct components, namely, the innate immune system and the adaptive immune system. Each system is equipped with unique cells and molecules to perform its specific functions. Macrophages and Natural Killer Cells (NK cells) are well-known cells of the innate immune system, whereas T lymphocytes and B lymphocytes are good examples of the adaptive immune system. To achieve successful immunity against pathogens, effective communication between these two systems is essential. In this context, the present discussion focuses on how various molecules found in ginseng can influence the immune system, and a summary is provided in Table 1.
Ginseng has been extensively studied for its bioactivity in modulating the immune system, a property also observed in other medicinal plants [23]. Initially, it was suggested that ginseng polysaccharides were responsible for this immunomodulatory effect [9]. However, subsequent investigations have revealed the involvement of various ginsenosides, including RT5, Rh2, oleanolic acid β-D-glucopyranosyl ester [24], Rh1 [25], Rg3 [26], and Rb1 [27], in triggering this biological response.
P. ginseng extracts have demonstrated noteworthy pharmacological bioactivities, including the improvement of macrophages’ phagocytic activity and enhanced production of NO [28,29]. Furthermore, ginseng extracts have been shown to boost interleukin 12 (IL-12) release [30]. Scaglione et al. [31] conducted a study at two levels, in vivo and with humans, demonstrating interesting results that can be immediately applied in clinical treatments, such as improvements in chemotaxis, phagocytosis index, and phagocytosis fraction. Additionally, different ginseng extracts have been observed to regulate various cytokines and molecules involved in immunomodulation, including IL-1α, IL-1β, IL-6, tumor necrosis factor α (TNF-α), NO, inducible nitric oxide synthase (iNOS), or cyclooxygenase-2 (COX-2) [32,33]. Therefore, ginseng extracts possess bioactive compounds that can regulate the immune system, resulting in a normalization of its functioning. The observed effect may be attributed to the presence of specific compounds or a synergistic activity among them, which may produce a more potent effect.
The earliest investigations into the immunomodulatory effects of ginseng attributed its bioactivity to its polysaccharide content [34,35,36,37,38,39]. The underlying mechanisms of action have been partially elucidated, with TNF-α being the primary stimulus, and other molecules, such as NO, IL-6, and IL-1β, also being stimulated [40]. In addition to the increased production of these immune-associated molecules, ginseng extracts have been shown to induce a Th1 immune response and activate pathways, such as nuclear factor κB (NF-κB), mitogen-activated protein kinases (MAPK), and phosphatidylinositol 3-kinase (PI3K) [14].
The primary distinction between ginseng extract responses depends on the mode of extraction and presentation, as well as the molecular weight of bioactive compounds. Notably, a conflicting effect was observed in vivo by Azike et al., who reported that, while the aqueous extract of ginseng polysaccharides increased TNF-α and NO levels, it also inhibited the physiological rise of these proinflammatory mediators induced by lipopolysaccharide (LPS). TNF-α levels were determined by Enzyme-Linked InmunoSorbent Assay (ELISA), and NO production was estimated by assessing nitrite accumulation with the Griess reagent. Treated animals displayed a reduction of around 50% in NO production, compared to non-treated animals [37]. This effect can be explained by considering that ginseng exerts an immunomodulatory action, instead of a stimulatory or inhibitory effect.
Ginsan, an acidic polysaccharide, has been found to possess unique abilities in stimulating the production of inflammatory mediators, such as NO, by upregulating iNOS, which sets it apart from other ginseng extracts [41]. However, ginsan also exhibits other bioactivities related to the immune response, such as the induction of T helper type 1 (Th1) cells and the release of cytokines produced by macrophages [42]. In addition, ginsan has multiple immunomodulatory effects on dendritic cells, including enhancing the expression of cluster of differentiation 86 (CD86) on their surface and increasing the levels of IL-12 and TNF-α secreted by them [43].
Recent studies have challenged the previous notion that polysaccharides are solely responsible for the immunomodulatory bioactivity of ginseng. Instead, ginsenosides and ginsenoside-like molecules have been shown to play an important role through distinct, and sometimes overlapping, mechanisms [44,45]. In vitro studies have demonstrated that ginsenoside RT5 and ginsenoside Rh2 increase IL-2 production. Interestingly, oleanane-type triterpenoids have been found to exhibit stronger immunomodulatory effects than those of the dammarane type [24]. Additionally, ginsenoside Rh2 has been found to increase the number of T cells in mice with melanoma, linking its immunomodulatory effect with anti-cancer properties [46]. Animal studies involving ginsenoside Rb2 have also revealed a higher survival rate and reduced tumor size [46].
In vitro studies have demonstrated the immunomodulatory activity of ginsenoside Rh1. Its action is complementary to that of polysaccharides, which regulate proinflammatory mediators, such as TNF-α or IL-6. Ginsenoside Rh1 reduces the expression of various proinflammatory mediators, including TNF-α, IL-1β, IL-6, IL-17, and NO. Additionally, it suppresses enzymes such as matrix metalloproteinase 1 (MMP-1), iNOS, and COX-2 [47,48]. In obese mice, Rh1 has been shown to exhibit immunomodulatory properties, by suppressing proinflammatory cytokines such as TNF-α, IL-1β, and IL-6 [49]. In hairless female mice, oral administration of Rh1 resulted in reduced IL-6 and immunoglobulin E (IgE) levels in peripheral blood analysis [50]. These findings suggest that Rh1 has some level of bioavailability through this route of administration; however, further research is necessary to clarify this point.
In various studies, other ginsenosides isolated as pure compounds have also demonstrated immunomodulatory properties. For instance, Rg3 has been found to enhance Fc gamma receptor-mediated phagocytosis in macrophages through mechanisms related to the activation of extracellular signal-regulated kinase 1/2 (ERK 1/2) and p38 [25]. Additionally, Rg3 has been shown to regulate cytokines and transcription factors, thereby maintaining homeostasis [51]. Ginsenoside Rg3 has been shown, in a study with patients with non-small cell lung cancer, to improve the immune response against COVID-19 by modulating the immune system against this disease. The same study shows the ability of ginsenoside Rg3 to regulate cell cycle, suggesting a possible application against different types of cancer [52].
Another compound, Rb1, has been found to increase both humoral and cell-mediated immune responses through the induction of antigen-presenting cells, which secrete TNF-α, and T cells, which secrete interferon gamma (IFN-γ) and IL-10. Rb1 also induces the production of immunoglobulins such as IgA, IgG1, and IgG2, and potentiates virus-triggered IFN-γ expression [27]. In another study, Rb1 was found to ameliorate the expression of TNF-α and IL-6 [53].
Rg1 has demonstrated many mechanisms through which it can carry out its activity related to immunomodulation. Studies, both in vivo and in vitro, have shown that Rg1 can activate the nuclear factor E2-related factor 2 (Nrf2) signaling pathway, which generates protection of the liver from toxins and disease [54,55,56]. Furthermore, both ginsenoside Rg1 and ginseng extracts have been shown to improve NK cells activity [57].
One of the main problems in therapy with ginsenosides is the great difficulty to absorb them. In fact, the bioavailability of ginsenosides from the intestinal mucosa is very low, and its transport through intestinal mucosa to blood is energy-dependent and non-saturable [3].
Compound K, which is not a main ginsenoside, but is closely related to them and derived from their biotransformation, is of clinical significance, due to its better bioavailability and numerous immune system activities. It attenuates NF-κB by modulating the protein kinase B (PKB or Akt)-mediated inflammatory gene expression [58], and regulates cytokines and other immune molecules such as IL-1β, IL-6, TNF-α, COX-2, and iNOS [59,60,61]. However, Compound K has also been shown to suppress humoral immune response of Th1 cells and suppresses the expression of matrix metalloproteinases and receptor activator of NF-κB ligand (RANKL) [62]. Additionally, it can inhibit β-arrestin2, hindering the transformation of macrophages from type M1 to type M2 [63].
Table 1. Immunomodulatory bioactivity of ginseng extracts and compounds. An increase is represented by (↑), and a decrease is represented by (↓).
Table 1. Immunomodulatory bioactivity of ginseng extracts and compounds. An increase is represented by (↑), and a decrease is represented by (↓).
SpeciesMolecular GroupCompound/ExtractExperimental ModelResultRef.
Panax quinquefolius L.PolysaccharidesExtractWistar rats↑ Macrophages activity[34]
PolysaccharidesExtractHuman peripheral blood mononuclear cells↑ Pro-inflammatory cytokines[35]
PolysaccharidesExtractMouse 3T3-L1 preadipocytesCytokines regulation[36]
PolysaccharidesExtractSprague–Dawley rats
Murine RAW 264.7
macrophage cell line
Cytokines regulation[37]
Panax ginseng C.A. MeyerPolysaccharidesAcidic fractionC57BL/6 mice macrophagesCytokines regulation[38]
PolysaccharidesAcidic fractionC57BL/6 miceEnhanced phagocytic effect[41]
PolysaccharidesAcidic fractionC57BL/6 mice dendritic cells↑ CD86[43]
GinsenosidesRh1Murine RAW 264.7
macrophage cell line
Glucocorticoid receptor stimulus[47]
GinsenosidesRh1Hartley guinea pigs, SD rats, and ICR mice↓ NO ↓ PGE2[48]
GinsenosidesRh1Mouse embryo fibroblasts 3T3-L1 cells↓ TNF-α ↓ IL-1β ↓ IL-6[49]
GinsenosidesRh1Hairless mice↓ Infiltration of inflammatory cells
↓ IgE levels
[50]
GinsenosidesRb1EV71 mice model↑ Cellular immune response
↑ Humoral immune response
[27]
GinsenosidesRg1C57BL/6 mice
C57BL/6 mice hepatocytes
↑ Nrf2
↑ Detoxifying enzymes
[55]
GinsenosidesRg3BALB/c miceImprove immune system[51]
GinsenosidesRg3Patients with non-small cell lung cancerRegulate cell cycle[52]
GinsenosidesStandardized
G-115 extract
BALB/c pathogen-free mice↑TLR4[40]
GinsenosidesCompound KMurine RAW 264.7
macrophage cell line
Human Embryonic Kidney cell line (HEK293 cells)
↓ iNOS ↓ TNF-α[58]
GinsenosidesCompound KSprague–Dawley rats
Kunming mice
Cytokines regulation[61]
GinsenosidesCompound KDBA/1 OlaHsd mice↓ Th1 response
(in arthritis)
[62]
GinsenosidesCompound KDBA/1 miceAlleviates inflammatory response[63]
-ExtractClinical trial↑ Chemotaxis[31]
-ExtractMurine RAW 264.7
macrophage cell line
BALB/c mice
Cytokines regulation[28]
-ExtractBalb/C mice
C57 B1/6J mice
C57 B1/6J nu/nu mice
↑ Antibody formation
↑ NK
[29]
-ExtractMurine RAW 264.7
macrophage cell line
Cytokines regulation[32]
-ExtractMurine RAW 264.7
macrophage cell line
↑ Immunomodulators[33]
Panax ginseng
C.A. Meyer
Eleutherococcus senticosus Rupr. & Maxim
-ExtractMouse J774A.1 macrophages↑ lL-12[30]

4. Anti-Inflammatory Activity

Inflammation is an integral part of the immune response, and, therefore, the immunomodulatory actions are of considerable significance in this regard [61]. While inflammation serves as a physiological response to protect against pathogenic infections, it poses a risk factor in various human diseases such as neurodegenerative, cardiovascular, autoimmune, respiratory disorders, and even certain cancers. Therefore, controlling or moderating the immune response may prove to be a useful strategy in preventing complications arising from inflammation. The results of recent studies on the anti-inflammatory activity of ginseng are shown in Table 2.
Compound K has been reported to possess anti-inflammatory properties. The mechanisms underlying these bioactivities include the downregulation of inflammatory cytokines such as IL-1β, IL-6, and TNF-α [64,65,66], as well as the modulation of reactive oxygen species (ROS) generation, Mitogen-activated protein kinases (MAPKs), NF-κB, and activator protein 1 (AP-1) [64,65,66,67,68].
Apart from specific ginsenosides, ginseng aqueous extracts have also been evaluated for their anti-inflammatory effects. For instance, in rats intoxicated with fipronil, an insecticide that triggers oxidative stress and induces inflammatory responses, Panax ginseng aqueous extract normalized the levels of reduced glutathione (GSH) and catalase (CAT) activity [69]. Ginseng aqueous extracts can be divided into two fractions, namely, saponin and non-saponin fractions. Although ginsenosides belong to the saponin fraction, which has been shown to reduce the generation of inflammatory molecules and messenger RNA (mRNA) levels of inflammatory cytokines and enzymes in vitro [70], both fractions have been reported to exert anti-inflammatory effects [71].
In the investigation of ginseng, the root is typically examined. However, studies have shown that ginseng berry extracts possess anti-inflammatory properties in macrophages induced by LPS. Specifically, ginseng berry extract reduces the production of NO and prostaglandin E2 (PGE2), as well as the expression of enzymes such as iNOS or COX-2. Moreover, it diminishes the secretion of cytokines such as IL-1β, IL-6, and TNF-α, and inhibits the translocation of NF-κB by avoiding the phosphorylation of the inhibitor factor κB (IF κBα). Additionally, ginseng berry extract suppresses the phosphorylation of ERK 1/2, c-Jun N-terminal kinase (JNK), and p38, reduces ROS, and increases the expression of enzymes such as glutathione peroxidase (GPxs), superoxide dismutase (SOD), and CAT. These effects are believed to be mediated by the inhibition of MAPKs signaling pathway [72].
Ginsenosides exhibit definite anti-inflammatory actions [73,74], particularly through their influence on the activation of inflammasomes. Inflammasomes are protein complexes formed by pattern-recognition receptors and other inflammatory mediators. Many ginsenosides have been shown to suppress inflammasome activation through various pathways. For instance, Rg3, Rd, Rg1, and 25-OCH3-protopanaxdiol, which are derived from Panax ginseng, can suppress the activation of the nucleotide-binding domain leucine-rich repeat-containing receptor, NLRP3 inflammasome. Rg3 specifically reduces the production of NO and the expression of iNOS [75], and suppresses the expression of TNF-α and the activation of NF-κB [76]. Rg1 has demonstrated the ability to suppress the activation of the nucleotide-binding domain leucine-rich repeat-containing receptor, NLRP1 inflammasome [77], and to inhibit interferon-inducible protein (AIM2) inflammasome activation in macrophages [78]. Rg1 exhibits anti-inflammatory and anti-oxidative properties, and has been shown to repair neural damages related to brain injury by inhibiting the activation of inflammasomes NLRP1 and NLRP3, or inflammasome AIM2 [76]. In addition, Rg1 demonstrates in vivo anti-inflammatory bioactivity by inhibiting interleukin receptor-associated kinase (IRAK) activation-mediated inflammatory responses. Moreover, Rg1 can inhibit the activation of the NF-κB signaling pathway [79,80]. Ginsenoside Rg5 has been shown to reduce mucin secretion, as well as to reduce the mRNA levels of the gene-encoding mucins. In an in vitro study with human lung mucoepidermoid carcinoma (NCI-H292) cells, the levels of MUC5AC (a mucin protein) were quantified, and it was observed that Rg5 reduced the secretion of MUC5AC, as well as the levels of MUC5AC mRNA. This study demonstrates that ginsenoside Rg5 alleviates inflammatory response by reducing mucin secretion [81]. Ginsenoside Rd has been shown to inhibit the expression of iNOS and COX-2 in later stages after ischemia [82]. Additionally, it reduces the generation of NO, production of PGE2, and activity of NF-κB [83]. Another saponin, chikusetsusaponine IVa, derived from Panax japonicum, a different species of ginseng, exhibits similar inhibitory actions on NLRP3 inflammasome [84,85].
Rb1 is among the most extensively studied ginsenosides. It possesses anti-inflammatory properties by suppressing IκB degradation. IκB family proteins play a crucial role in the first inhibition of NF-κB by rendering it inactive within cells. Furthermore, Rb1 prevents NLRP3 inflammasome activation and mitochondrial damage, suggesting it is an anti-inflammatory molecule [86]. Rb1 has been shown to inhibit TNF-α production in LPS-stimulated macrophages, thereby controlling inflammation and TNF-α production [87,88,89]. However, oral administration of Rb1 results in poor bioavailability, which poses challenges for its use in clinical treatment. Nevertheless, its bioavailability can be enhanced through galenic development, such as the use of various nanoformulation strategies.
Several other ginsenosides have demonstrated anti-inflammatory properties, either with similar or distinct chemical structures. For instance, Rb2 shares a very similar chemical structure to Rb1. Rb2 can also inhibit TNF-α production in cells stimulated with LPS, similar to Rb1, through the inhibition of NF-κB [90,91]. On the other hand, ginsenoside Rh1 is a considerably smaller molecule than Rb1 or Rb2, but it retains the primary structure and exhibits anti-inflammatory activity by suppressing the expression of COX-2 and iNOX [48,92].
A bacterial metabolite produced from the metabolism of ginsenoside Rh1 is ginsenoside Rh2, which also exhibits anti-inflammatory activity in microglial and astroglial cells [68,93]. Another ginsenoside derivative is ginsenoside Rp1, which has demonstrated the ability to reduce the expression of IL-1β, COX-2, and iNOS by suppressing NF-κB activity in vitro [94,95], despite being less studied, compared to other ginsenosides.
The bioactive compounds present in ginseng are widely recognized. However, ginseng also harbors other significant bioactive molecules, including a non-saponin component, known as gintonin, which has gained increasing interest. Gintonin is a complex made up of carbohydrates, proteins, and lipids, and it has been found to be divisible into six different subtypes. The principal bioactive compounds of gintonin are a series of lyso-phosphatidic acids, which act as ligands for specific membrane receptors, notably, the G-coupled protein receptors [96]. Gintonin exerts its effects through Gq protein alpha subunit 11 (GαQ/11), phospholipase C, inositol triphosphate (IP3), and release of calcium, leading to the activation of Ca2+-activated Cl—channels. Moreover, gintonin has been shown to attenuate neuroinflammation associated with diseases, such as Alzheimer’s disease, by inhibiting reactive oxygen species formation and reducing cytokine production in the brain [97].
Table 2. Anti-inflammatory bioactivity of ginseng extracts and compounds. An increase is represented by (↑), and a decrease is represented by (↓).
Table 2. Anti-inflammatory bioactivity of ginseng extracts and compounds. An increase is represented by (↑), and a decrease is represented by (↓).
SpeciesMolecular GroupCompound/ExtractExperimental ModelResultRef.
Panax ginseng C.A. MeyerGinsenosideRg1ICR miceSuppression NLRP1 inflammasome activation[75]
GinsenosideRg1Murine RAW 264.7
macrophage cell line
↓ IL-6[78]
GinsenosideRg1C57BL/6 miceInhibition NF-κB pathway[79]
GinsenosideRg1ICR miceInhibition NF-κB pathway
↓ iNOS ↓ COX-2
[80]
Panax ginseng C.A. MeyerGinsenoside
Ginsenoside
Rg3
Rg5
BV-2 microglial cells
Neuro-2a cells
NCI-H292 cells
Suppression TNF-α
and NF-κB
↓ MUC5AC and ↓ MUC5AC mRNA
[76,81]
GinsenosideRdSprague–Dawley rats↓ iNOS ↓ COX-2[82]
GinsenosideRdMurine RAW 264.7
macrophage cell line
↓ iNOS ↓ COX-2[83]
GinsenosideRb1Sprague–Dawley rats↑ IκB[86]
GinsenosideRb1Murine RAW 264.7
macrophage cell line
Suppression TNF-α[88]
GinsenosideRb1ICR miceInhibition NF-κB pathway[87]
GinsenosideRb1 y Rb2Murine RAW 264.7
macrophage cell line
↓ TNF-α[90]
GinsenosideRb2, Rd, Re and Rg1Murine N9 microglial cell line↓ TNF-α ↓ NO[91]
GinsenosideRe and Rh1Murine RAW 264.7
macrophage cell line
↓ iNOS ↓ COX-2[48]
GinsenosideRh2BV-2 microglial cells↓ NO ↓ COX-2 ↓ TNF-α
↓ IL-1β
[93]
GinsenosideRp1Murine RAW 264.7
macrophage cell line
↓ iNOS ↓ COX-2 ↓ IL-1β[94]
Panax ginseng C.A. MeyerGinsenosideRp1Murine RAW 264.7
macrophage cell line
↓ IL-1β ↓ iNOS ↓ COX-2 ↓ TNF-α[95]
-Root water extract (saponin fraction)Murine RAW 264.7
macrophage cell line
↓ iNOS ↓ COX-2 ↓ TNF-α[71]
-Berry extractMurine RAW 264.7
macrophage cell line
↓ iNOS ↓ COX-2 ↓ IL-1β
↓ IL-6 ↓ TNF-α
[72]
Glycolipoprotein complexGintoninSH-SY5Y Human neuroblastoma cell line↓ ROS formation[97]
Panax japonicum
C.A. Meyer
SaponinesChikusetsusaponine IvaTHP-1 human monocyte-like cells↓ iNOS ↓ TNF-α ↓ IL-6
↓ IL-1β
[85]
Panax notoginseng Burk.GinsenosidesRb1Murine RAW 264.7
macrophage cell line
↓TNF-α ↓ IL-6 ↓COX-2
↓ IL-1β
[89]

5. Anti-Cancer Activity

The increase in life expectancy and modern lifestyles has led to a rise in cancer diagnoses in Western societies. The pursuit of a definitive cure for cancer appears to be a distant prospect, as different cancer types necessitate distinct treatment approaches, rendering cancer more a group of diseases than a singular entity. Hence, the discovery of compounds that either prevent cancer or aid in its treatment is of utmost significance. The findings related to the studies of the antitumor activity of ginseng are summarized in Table 3.
Various ginseng compounds, including ginsenosides [98], polysaccharides [99], and alkaloids extracted using alcohol, have exhibited different activities against distinct stages in cancer development, such as cancer cell growth, proliferation, viability, and angiogenesis processes linked to cancer. An essential factor to consider in the anti-cancer bioactivity of ginsenosides is the number of sugar moieties present. Notably, the level of anti-cancer bioactivity is inversely proportional to the quantity of sugar moieties. For instance, ginsenosides Rb1 or Rc, which have four or more sugar moieties, exhibit minimal anti-cancer bioactivity [100].
Numerous in vitro studies have investigated ginsenoside Rh1 as an anti-cancer agent [19,101]. It has been examined against various cancer types, such as human lung carcinoma (A549) or human cervix uterine adenocarcinoma (HeLa). Rh1 has been shown to exhibit antiproliferative activity on mouse fibroblast cells. The mechanism underlying this effect is the inhibition of phospholipase C, resulting in a decrease of intracellular diacylglycerol, which is an endogenous activator of protein kinase C [102]. In addition, an anti-cancer effect has been observed on a human leukemia cell line, which is attributed to the induction of apoptosis [103].
Ginsenoside Rh2 exhibits a chemical structure similar to that of dexamethasone [104]. In in vitro studies, it has been demonstrated to suppress the growth and viability of various cancer cells, induce tumor cell cycle arrest and cellular apoptosis, trigger necrosis and autophagy in cancer cells, inhibit metastasis, and suppress angiogenesis. The mechanisms associated with Rh2 have been partially elucidated and classified based on their frequency and type of action. The type of action has been divided into mechanisms related to apoptosis and autophagy, mechanisms related to cell cycle regulation, and mechanisms related to invasion and migration of cancerous cells.
Rh2 has been shown to release mitochondrial cytochrome C, reduce mitochondrial membrane potential, and activate various pathways of kinases and caspases that lead to apoptosis or autophagy processes, thereby offering protection against cancer [105,106,107]. Additionally, Rh2 has demonstrated down-regulation of cyclins, up-regulation of protein expression, and activation or regulation of different pathways leading to control cell cycle. Rh2 has also been shown to regulate the expression of proteins to prevent invasion and migration of cancerous cells. Notably, Rh2 can be used as a contributory drug to prevent drug resistance in cancer treatments. In a recent study, ginsenoside Rh2 has shown anti-cancer bioactivity against many different types of cancer, such as nasopharyngeal carcinoma, glioma, lung cancer, breast cancer, digestive system cancer, genitourinary system cancer, melanoma, or leukemia. Rh2 also helps to reduce cancer drug resistance and alleviates the side effects of classical treatments for cancer, such as chemotherapy. Depending on the type of cancer, Rh2 has been shown to intervene at one or more points in the process (cell proliferation, cell apoptosis, cell cycle, autophagy, migration, or angiogenesis) [107]. Other molecules obtained from different species of ginseng, such as 25-OH-protopanaxdiol, share this pro-apoptotic bioactivity with ginsenoside Rh2 [108].
Several ginsenosides from Panax species, namely, Rg1, Rg2, Rg3, and Rg5, exhibit potential anti-cancer properties. Rg1 has been found to inhibit oncogenes [109], while Rg2 increases the levels of p53 and p21, which play critical roles in regulating cellular division [110]. Rg3 is one of the most pharmacologically active ginsenosides, with a unique chemical structure that contains a hydroxyl group at C20, resulting in two different epimers, RRg3 and SRg3 [111]. Both epimers exhibit anti-angiogenic and anti-metastatic activities [112], with RRg3 showing greater potency and efficacy in inhibiting the migration and invasion of breast cancer cells in vitro [113]. Additionally, Rg3 has demonstrated both anti-proliferative and apoptotic bioactivities [114,115]. Among these ginsenosides, Rg5 exhibits the strongest anti-proliferative bioactivity [116].
In addition to the anti-cancer activities demonstrated by various ginsenosides as pure compounds, studies have also investigated the potential anti-cancer effects of different ginseng species, including the non-saponin fractions of their extracts. For instance, alkaloids and polysaccharides derived from Korean Red ginseng (P. ginseng) have been shown to exhibit anti-cancer activities by inhibiting the proliferation of cancer cells [117,118] and acting as an adjuvant therapy in combination with chemotherapy [103]. Moreover, the alkaloids from Korean Red ginseng have been found to exert a protective effect against chromosomal damage, promoting the repair and regrowth of radiation-damaged cells [119].
Compound K has been shown to possess anti-cancer properties through its cytotoxic effects on tumor cells [120]. It has been found to regulate tumor growth and the tumor microenvironment via various biochemical signaling pathways, and induce tumor apoptosis. These anti-cancer activities have been demonstrated in various types of cancer, including lung, bladder, colon, myeloma, and neuroblastoma [121,122]. Additionally, in lung cancer, Compound K has been observed to induce apoptosis and autophagy [123].
In P. ginseng, polyacetylene compounds have been extensively investigated, and they are mainly composed of panaxydol and panaxynol, which constitute approximately 90% of the total polyacetylene content. Among these compounds, panaxydol has shown the highest potency against cancer cells. Additionally, both panaxynol and panaxydol exhibit significant chemopreventive effects against oncogenesis and mutations that can lead to genetic instability. Another minor polyacetylene, panaxytriol, has been found to possess anti-proliferative effects against cancer cells [124].
A group of eight major ginsenosides (Re, Rg1, Rc, Rb1, Rb2, Rb3, Rd, and Rg3) extracted from P. quinquefolius, has been investigated for their anti-cancer effects [125]. Several of these ginsenosides have demonstrated anti-cancer activity, including Rg3, as we have previously discussed, as well as Rb1 [126,127], Rb3, and Rd [128]. This extract inhibited cancer cell growth and migration, as well as suppressed the invasive properties of cancer cells in a dose-dependent manner. Furthermore, it increased the expression of tumor suppressor genes p53 and p21 while down-regulating B-cell lymphoma 2 (Bcl2) and signal transducer and activator of transcription 3 (STAT3), which are crucial genes related to cancer research. Rg3, along with other ginsenosides such as Rk1 and Rg5, not present in this extract, induced cell cycle arrest in the G1 phase, resulting in the suppression of cellular proliferation, angiogenesis, and metastasis in prostate cancer [125,129,130,131].
P. notoginseng has ginsenoside-like molecules. Among these, 25-OH-PPD and 25-OCH3-PPD have been shown to have anti-cancer activity by inducing apoptosis and inhibiting proliferation. These compounds activate caspases and down-regulate mouse double-minute 2 homolog (MDM2), while increasing p21 gene expression [110,132,133,134,135]. Additionally, notoginsenoside R1 has demonstrated anti-cancer activity by inhibiting TNF-α and inhibiting proliferation of cancerous cells [132].
Table 3. Anti-cancer bioactivity of ginseng extracts and compounds. An increase is represented by (↑), and a decrease is represented by (↓).
Table 3. Anti-cancer bioactivity of ginseng extracts and compounds. An increase is represented by (↑), and a decrease is represented by (↓).
SpeciesMolecular GroupCompound/ExtractExperimental ModelResultRef.
Panax ginseng C.A. MeyerGinsenosideRh1Mouse lymphoid neoplasma cell line (P388)Cytotoxic effect[102]
GinsenosideRh1Human leukemia (THP-1) cell line↑ apoptosis[103]
GinsenosideRh2B16 melanoma cell line↓ cell growth[104]
Ginsenoside
Ginsenoside
Rh2
Rh2
Murine melanoma (B16F10) cell line, Human breast cancer line (MDA-MB-231)cell, and Hepatocyte derived cellular carcinoma (HuH-7) cell lineAnti-proliferation
Anti-invasion
Anti-metastasis
[106]
Diverse cancer modelsCell cycle, autophagy, migration and angiogenesis
Alleviates chemotherapy effects
[107]
Panax ginseng C.A. MeyerGinsenosideRg1Osteosarcoma MG-63 cellsOncogenes inhibition[109]
GinsenosideRg3Breast cancer modelAnti-proliferation[113]
GinsenosideRg5Hepatic Adenocarcinoma SK-HEP-1 cells↑ p21Cip/WAF1
↓ cyclin E ↓ CDK2
↓ CDC25A
[116]
Panax ginseng C.A. MeyerGinsenosideCompound KMouse high-metastatic melanoma B16-BL6
Human myeloid leukemia K562
Human liver cancer HepG2
Human high-metastatic lung carcinoma 95-D
↓ tumor cells[120]
GinsenosideCompound K mouse highly metastatic melanoma (B16-BL6)
Human liver cancer (HepG2)
Human myeloid leukemia (K562) Human highly metastatic lung cancer (95-D)
↓ tumor cells[121]
GinsenosideCompound KLung cancer cells A549 and H1975↑ autophagy ↑apoptosis[122]
GinsenosideCompound KBladder cancer T24 cells↑apoptosis[123]
GinsenosidesCompound K and Rb1SKOV-3 and HEYA8 cells↓ tumor cells survival[126]
GinsenosideRb3 and RdApcMin/+ mice↓ oncogenic signaling molecules (iNOS, STAT3/pSTAT3, Src/pSrc)[128]
GinsenosideRe, Rg1, Rc, Rb1, Rb2, Rb3, Rd, Rg3, Rg5 and Rk1Human lung cancer cells
Human breast cancer cells
↑ apoptosis
↓ cell proliferation
↑ p21
[133,134,135,136]
-ExtractImmortalized human keratinocytes (HaCaT )cells↑ cell viability[118]
Panax ginseng C.A. MeyerAlkaloid-C57BL/6 mouse spleen lymphocytesReparation damaged cells[119]
Polyacetylene compoundsPanaxydolMurine RAW 264.7
macrophage cell line
↓ tumor cells[124]
Panax quinquefolius
L.
GinsenosideRg3 SW-480 (Leibovitz’s L-15), HT-29 (McCoy’s 5A), and non-small cell lung (NSCLC, DMEM) Anti-proliferation[125]
Panax notoginseng
Burk
GinsenosideNotoginsenoside R1Sprague–Dawley ratsCell protection[132]

6. Conclusions

Different species of genus Panax contain numerous compounds with important properties and diverse biological activities. It is important to differentiate among species: their active compounds can differ because ginsenosides are not the only biologically active compounds, but also polysaccharides, polyacetylene compounds, and alkaloids.
Ginseng is typically consumed in combination with other medicinal plants, or enriched with vitamins and minerals, to enhance its natural properties and create a multifunctional drug.
One of the points to improve in therapy with ginsenosides is to manage a better absorption. There is an added problem as well: as ginseng is often used for its antioxidant and neuroprotective properties, it is needed to cross the blood-brain barrier. The structure of ginsenosides make this crossing difficult. Further galenic investigations are needed to improve the pharmaceutical technology which reaches the appropriate formulation to get to the nervous system.
There are still numerous compounds in ginseng that are yet to be studied, including some ginsenosides that are not present in large quantities in ginseng roots, as well as various polysaccharides and other molecules that may have significant and distinct activities, if isolated and studied individually. This presents a potential avenue for future research.

Author Contributions

Conceptualization, M.P.G.-S.; methodology, M.P.G.-S.; validation, M.P.G.-S., J.A.V.-G., M.S. and I.M.-R.; investigation, J.A.V.-G.; writing—original draft preparation, J.A.V.-G.; writing—review and editing, M.S. and I.I.; visualization, M.P.G.-S.; supervision, M.P.G.-S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Figures were designed by Flaticon.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviation

AbbreviationMeaning
PanaxP.
Nitric OxideNO
GlucoseGlu
Arabinose in pyranose formArbp
XyloseXyl
Arabinose in furanose formArbf
RhamnoseRham
High-Performance Liquid Chromatography HPLC
Thin Layer Chromatography/High Performance Liquid ChromatographyTLC/HPLC
Ultra High-Performance Liquid ChromatographyUHPLC
Two-Dimensional High-Performance Liquid ChromatographyHPLC 2D
Natural Killer cellNK cell
InterleukinIL
Tumor necrosis factor α TNF-α
Inducible nitric oxide synthaseiNOS
Cyclooxygenase-2COX-2
Nuclear factor κBNF-κB
Mitogen-activated protein kinases MAPK
Phosphatidylinositol 3-kinasePI3K
Enzyme-Linked ImmunoSorbent AssayELISA
T helper type 1Th1
Cluster of differentiation 86CD86
Matrix metalloproteinase 1MMP-1
Extracellular signal-regulated kinase 1/2ERK 1/2
interferon gammaIFN-γ
ImmunoglobulinIg
Nuclear factor E2-related factor 2Nrf2
Protein kinase BPKB or Akt
Receptor activator of NF-κB ligandRANKL
Human Embryonic Kidney cell lineHEK293 cells
Prostaglandin E2PEG2
Toll Like receptorTLR4
Reactive oxygen speciesROS
glutathioneGSH
Messenger RNAmRNA
LipopolysaccharideLPS
c-Jun N-terminal kinaseJNK
Glutathione PeroxidaseGPxs
Superoxide dismutaseSOD
Human lung mucoepidermoid carcinoma cell line NCI-H292
Catalase CAT
nucleotide-binding domain leucine-rich repeat-containing receptor NLRP
interferon-inducible protein AIM2
interleukin receptor-associated kinase IRAK
IkappaB kinase IκB
Gq protein alpha subunit GαQ
Inositol triphosphateIP3
B-cell lymphoma 2Bcl2
signal transducer and activator of transcription 3STAT3
mouse double minute 2 homologMDM2
Hepatic AdenocarcinomaSK-HEP-1 cells
Murine melanoma cell lineB16F10
Human breast cancer cell lineMDA-MB-231
Hepatocyte derived cellular carcinoma cell lineHuH-7
Immortalized human keratinocytes cell lineHaCaT

References

  1. Mancuso, C.; Santangelo, R. Panax ginseng and Panax quinquefolius: From pharmacology to toxicology. Food Chem. Toxicol. 2017, 107, 362–372. [Google Scholar] [CrossRef]
  2. Attele, A.S.; Wu, J.A.; Yuan, C.-S. Ginseng pharmacology: Multiple constituents and multiple actions. Biochem. Pharmacol. 1999, 58, 1685–1693. [Google Scholar] [CrossRef] [PubMed]
  3. Leung, K.W.; Wong, A.S.-T. Pharmacology of ginsenosides: A literature review. Chin. Med. 2010, 5, 20. [Google Scholar] [CrossRef] [PubMed]
  4. González-Burgos, E.; Fernandez-Moriano, C.; Gómez-Serranillos, M.P. Potential Neuroprotective Activity of Ginseng in Parkinson’s Disease: A Review. J. Neuroimmune Pharmacol. 2014, 10, 14–29. [Google Scholar] [CrossRef] [PubMed]
  5. Zhang, H.; Abid, S.; Ahn, J.C.; Mathiyalagan, R.; Kim, Y.-J.; Yang, D.-C.; Wang, Y. Characteristics of Panax ginseng Cultivars in Korea and China. Molecules 2020, 25, 2635. [Google Scholar] [CrossRef]
  6. Ratan, Z.A.; Youn, S.H.; Kwak, Y.-S.; Han, C.-K.; Haidere, M.F.; Kim, J.K.; Min, H.; Jung, Y.-J.; Hosseinzadeh, H.; Hyun, S.H.; et al. Adaptogenic effects of Panax ginseng on modulation of immune functions. J. Ginseng Res. 2021, 45, 32–40. [Google Scholar] [CrossRef]
  7. Wang, Y.; Mi, X.; Du, Y.; Li, S.; Yu, L.; Gao, M.; Yang, X.; Song, Z.; Yu, H.; Yang, G. Design, Synthesis, and Anti-Inflammatory Activities of 12-Dehydropyxinol Derivatives. Molecules 2023, 28, 1307. [Google Scholar] [CrossRef]
  8. Wang, M.; Li, H.; Liu, W.; Cao, H.; Hu, X.; Gao, X.; Xu, F.; Li, Z.; Hua, H.; Li, D. Dammarane-type leads panaxadiol and protopanaxadiol for drug discovery: Biological activity and structural modification. Eur. J. Med. Chem. 2020, 189, 112087. [Google Scholar] [CrossRef]
  9. Karra, A.G.; Konstantinou, M.; Tzortziou, M.; Tsialtas, I.; Kalousi, F.D.; Garagounis, C.; Hayes, J.M.; Psarra, A.-M.G. Potential Dissociative Glucocorticoid Receptor Activity for Protopanaxadiol and Protopanaxatriol. Int. J. Mol. Sci. 2018, 20, 94. [Google Scholar] [CrossRef]
  10. Nair, R.; Sellaturay, S.; Sriprasad, S. The history of ginseng in the management of erectile dysfunction in ancient China (3500-2600 BCE). Indian, J. Urol. 2012, 28, 15–20. [Google Scholar] [CrossRef]
  11. Villar, A.M.; Naval, M.V.; Gómez-Serranillos, M.P. Ginseng: Revisión. Farm. Prof. 2003, 17, 68–73. [Google Scholar]
  12. Sánchez, M.; González-Burgos, E.; Iglesias, I.; Lozano, R.; Gómez-Serranillos, M.P. Current uses and knowledge of medicinal plants in the Autonomous Community of Madrid (Spain): A descriptive cross-sectional study. BMC Complement. Med. Ther. 2020, 20, 306. [Google Scholar] [CrossRef] [PubMed]
  13. Ekor, M. The growing use of herbal medicines: Issues relating to adverse reactions and challenges in monitoring safety. Front. Pharmacol. 2014, 4, 177. [Google Scholar] [CrossRef] [PubMed]
  14. Ghosh, R.; Bryant, D.L.; Farone, A.L. Panax quinquefolius (North American Ginseng) Polysaccharides as Immunomodulators: Current Research Status and Future Directions. Molecules 2020, 25, 5854. [Google Scholar] [CrossRef]
  15. Yang, Y.; Ju, Z.; Yang, Y.; Zhang, Y.; Yang, L.; Wang, Z. Phytochemical analysis of Panax species: A review. J. Ginseng Res. 2021, 45, 1–21. [Google Scholar] [CrossRef] [PubMed]
  16. Chinese Pharmacopoeia Commission. Chinese Pharmacopoeia; China Medical Science Press: Beijing, China, 2015; Volume 1, pp. 191–193. [Google Scholar]
  17. Yang, W.-Z.; Hu, Y.; Wu, W.-Y.; Ye, M.; Guo, D.-A. Saponins in the genus Panax L. (Araliaceae): A systematic review of their chemical diversity. Phytochemistry 2014, 106, 7–24. [Google Scholar] [CrossRef]
  18. Chen, Z.; Zhang, Z.; Liu, J.; Qi, H.; Li, J.; Chen, J.; Huang, Q.; Liu, Q.; Mi, J.; Li, X. Gut Microbiota: Therapeutic Targets of Ginseng Against Multiple Disorders and Ginsenoside Transformation. Front. Cell. Infect. Microbiol. 2022, 12, 456. [Google Scholar] [CrossRef]
  19. Dong, T.T.X.; Cui, X.M.; Song, Z.H.; Zhao, K.J.; Ji, Z.N.; Lo, C.K.; Tsim, K.W.K. Chemical Assessment of Roots of Panax notoginseng in China: Regional and Seasonal Variations in Its Active Constituents. J. Agric. Food Chem. 2003, 51, 4617–4623. [Google Scholar] [CrossRef]
  20. Yang, W.-Z.; Ye, M.; Qiao, X.; Liu, C.-F.; Miao, W.-J.; Bo, T.; Tao, H.-Y.; Guo, D.-A. A strategy for efficient discovery of new natural compounds by integrating orthogonal column chromatography and liquid chromatography/mass spectrometry analysis: Its application in Panax ginseng, Panax quinquefolium and Panax notoginseng to characterize 437 potential new ginsenosides. Anal. Chim. Acta 2012, 739, 56–66. [Google Scholar] [CrossRef]
  21. Liu, J.; Wu, Y.; Ma, W.; Zhang, H.; Meng, X.; Zhang, H.; Guo, M.; Ling, X.; Li, L. Anti-Inflammatory Activity of Panax notoginseng Flower Saponins Quantified Using LC/MS/MS. Molecules 2023, 28, 2416. [Google Scholar] [CrossRef]
  22. Yao, C.-L.; Pan, H.-Q.; Wang, H.; Yao, S.; Yang, W.-Z.; Hou, J.-J.; Jin, Q.-H.; Wu, W.-Y.; Guo, D.-A. Global profiling combined with predicted metabolites screening for discovery of natural compounds: Characterization of ginsenosides in the leaves of Panax notoginseng as a case study. J. Chromatogr. A 2018, 1538, 34–44. [Google Scholar] [CrossRef]
  23. Ferreira, S.S.; Passos, C.P.; Madureira, P.; Vilanova, M.; Coimbra, M.A. Structure–function relationships of immunostimulatory polysaccharides: A review. Carbohydr. Polym. 2015, 132, 378–396. [Google Scholar] [CrossRef]
  24. Vinh, L.B.; Park, J.U.; Duy, L.X.; Nguyet, N.T.M.; Yang, S.Y.; Kim, Y.R.; Kim, Y.H. Ginsenosides from Korean red ginseng modulate T cell function via the regulation of NF-AT-mediated IL-2 production. Food Sci. Biotechnol. 2018, 28, 237–242. [Google Scholar] [CrossRef] [PubMed]
  25. Tam, D.N.H.; Truong, D.H.; Nguyen, T.T.H.; Quynh, L.N.; Tran, L.; Nguyen, H.D.; Shamandy, B.E.; Le, T.M.H.; Tran, D.K.; Sayed, D.; et al. Ginsenoside Rh1: A Systematic Review of Its Pharmacological Properties. Planta Medica 2018, 84, 139–152. [Google Scholar] [CrossRef]
  26. Xin, C.; Kim, J.; Quan, H.; Yin, M.; Jeong, S.; Choi, J.I.; Jang, E.A.; Lee, C.H.; Kim, D.H.; Bae, H.B. Ginsenoside Rg3 promotes Fc gamma receptor-mediated phagocytosis of T bacteria by macrophages via an extracellular signal-regulated kinase 1/2 and p38 mitogen-activated protein kinase-dependent mechanism. Int. Immunopharmacol. 2019, 77, 105945. [Google Scholar] [CrossRef] [PubMed]
  27. Kang, N.; Gao, H.; He, L.; Liu, Y.; Fan, H.; Xu, Q.; Yang, S. Ginsenoside Rb1 is an immune-stimulatory agent with antiviral activity against enterovirus 71. J. Ethnopharmacol. 2021, 266, 113401. [Google Scholar] [CrossRef]
  28. Lim, T.-G.; Jang, M.; Cho, C.-W.; Hong, H.-D.; Kim, K.-T.; Lee, S.-Y.; Jung, S.K.; Rhee, Y.K. White ginseng extract induces immunomodulatory effects via the MKK4-JNK pathway. Food Sci. Biotechnol. 2016, 25, 1737–1744. [Google Scholar] [CrossRef]
  29. Jie, Y.H.; Cammisuli, S.; Baggiolini, M. Immunomodulatory effects of Panax Ginseng C.A. Meyer in the mouse. Agents Actions 1984, 15, 386–391. [Google Scholar] [CrossRef] [PubMed]
  30. Wang, H.; Actor, J.K.; Indrigo, J.; Olsen, M.; Dasgupta, A. Asian and Siberian ginseng as a potential modulator of immune function: An in vitro cytokine study using mouse macrophages. Clin. Chim. Acta 2003, 327, 123–128. [Google Scholar] [CrossRef]
  31. Scaglione, F.; Ferrara, F.; Dugnani, S.; Falchi, M.; Santoro, G.; Fraschini, F. Immunomodulatory effects of two extracts of Panax ginseng C.A. Meyer. Drugs Exp. Clin. Res. 1990, 16, 537–542. [Google Scholar] [PubMed]
  32. Jang, H.I.; Shin, H.M. Wild Panax Ginseng (Panax ginseng C.A. Meyer) Protects Against Methotrexate–Induced Cell Regression by Enhancing the Immune Response in RAW 264.7 Macrophages. Am. J. Chin. Med. 2010, 38, 949–960. [Google Scholar] [CrossRef]
  33. Um, Y.; Eo, H.J.; Kim, H.J.; Kim, K.; Jeon, K.S.; Jeong, J.B. Wild simulated ginseng activates mouse macrophage, RAW264.7 cells through TRL2/4-dependent activation of MAPK, NF kappaB and PI3K/AKT pathways. J. Ethnopharmacol. 2020, 263, 113218. [Google Scholar] [CrossRef] [PubMed]
  34. Assinewe, V.; Arnason, J.; Aubry, A.; Mullin, J.; Lemaire, I. Extractable polysaccharides of Panax quinquefolius L. (North American ginseng) root stimulate TNFa production by alveolar macrophages. Phytomedicine 2002, 9, 398–404. [Google Scholar] [CrossRef] [PubMed]
  35. Lemmon, H.R.; Sham, J.; Chau, L.A.; Madrenas, J. High molecular weight polysaccharides are key immunomodulators in North American ginseng extracts: Characterization of the ginseng genetic signature in primary human immune cells. J. Ethnopharmacol. 2012, 142, 1–13. [Google Scholar] [CrossRef]
  36. Wilson, S.A.F.; Wong, M.H.T.; Stryjecki, C.; De Boer, A.; Lui, E.M.K.; Mutch, D. Unraveling the adipocyte inflammomodulatory pathways activated by North American ginseng. Int. J. Obes. 2013, 37, 350–356. [Google Scholar] [CrossRef]
  37. Azike, C.G.; Charpentier, P.A.; Lui, E.M. Stimulation and Suppression of Innate Immune Function by American Ginseng Polysaccharides: Biological Relevance and Identification of Bioactives. Pharm. Res. 2015, 32, 876–897. [Google Scholar] [CrossRef] [PubMed]
  38. Shin, J.Y.; Song, J.Y.; Yun, Y.S.; Yang, H.O.; Rhee, D.K.; Pyo, S. Immunostimulating effects of acidic polysaccharides extract of Panax ginseng on macrophage function. Immunopharmacol. Immunotoxicol. 2002, 24, 469–482. [Google Scholar] [CrossRef]
  39. Pannacci, M.; Lucini, V.; Colleoni, F.; Martucci, C.; Grosso, S.; Sacerdote, P.; Scaglione, F. Panax ginseng C.A. Mayer G115 modulates pro-inflammatory cytokine production in mice throughout the increase of macrophage toll-like receptor 4 expression during physical stress. Brain, Behav. Immun. 2006, 20, 546–551. [Google Scholar] [CrossRef]
  40. Lim, D.; Bae, K.; Jung, I.; Kim, C.; Yun, Y.; Song, J. Anti-Septicaemic Effect of Polysaccharide from Panax ginseng by Macrophage Activation. J. Infect. 2002, 45, 32–38. [Google Scholar] [CrossRef]
  41. Choi, H.S.; Kim, K.H.; Sohn, E.; Park, J.D.; Kim, B.O.; Moon, E.Y.; Rhee, D.K.; Pyo, S. Red ginseng acidic polysaccharide (RGAP) in combination with IFN-gamma results in enhanced macrophage function through activation of the NF-kappaB pathway. Biosci. Biotechnol. Biochem. 2008, 72, 1817–1825. [Google Scholar] [CrossRef]
  42. Kim, K.-H.; Lee, Y.-S.; Jung, I.-S.; Park, S.-Y.; Chung, H.-Y.; Lee, I.-R.; Yun, Y.-S. Acidic Polysaccharide from Panax ginseng, Ginsan, Induces Th1 Cell and Macrophage Cytokines and Generates LAK Cells in Synergy with rIL-2. Planta Medica 1998, 64, 110–115. [Google Scholar] [CrossRef] [PubMed]
  43. Kim, M.-H.; Byon, Y.-Y.; Ko, E.-J.; Song, J.-Y.; Yun, Y.-S.; Shin, T.; Joo, H.-G. Immunomodulatory Activity of Ginsan, a Polysaccharide of Panax Ginseng, on Dendritic Cells. Korean J. Physiol. Pharmacol. 2009, 13, 169–173. [Google Scholar] [CrossRef] [PubMed]
  44. You, L.; Cha, S.; Kim, M.-Y.; Cho, J.Y. Ginsenosides are active ingredients in Panax ginseng with immunomodulatory properties from cellular to organismal levels. J. Ginseng Res. 2022, 46, 711–721. [Google Scholar] [CrossRef] [PubMed]
  45. Luo, H.; Chen, J.; Su, C.; Zha, L. Advances in the Bioactivities of Phytochemical Saponins in the Prevention and Treatment of Atherosclerosis. Nutrients 2022, 14, 4998. [Google Scholar] [CrossRef]
  46. Wang, M.; Yan, S.-J.; Zhang, H.-T.; Li, N.; Liu, T.; Zhang, Y.-L.; Li, X.-X.; Ma, Q.; Qiu, X.-C.; Fan, Q.-Y.; et al. Ginsenoside Rh2 enhances the antitumor immunological response of a melanoma mice model. Oncol. Lett. 2016, 13, 681–685. [Google Scholar] [CrossRef]
  47. Li, J.; Du, J.; Liu, D.; Cheng, B.; Fang, F.; Weng, L.; Wang, C.; Ling, C. Ginsenoside Rh1 potentiates dexamethasone’s anti-inflammatory effects for chronic inflammatory disease by reversing dexamethasone-induced resistance. Arthritis Res. Ther. 2014, 16, R106. [Google Scholar] [CrossRef]
  48. Park, E.-K.; Choo, M.-K.; Han, M.J.; Kim, D.-H. Ginsenoside Rh1 Possesses Antiallergic and Anti-Inflammatory Activities. Int. Arch. Allergy Immunol. 2004, 133, 113–120. [Google Scholar] [CrossRef]
  49. Gu, W.; Kim, K.-A.; Kim, D.-H. Ginsenoside Rh1 Ameliorates High Fat Diet-Induced Obesity in Mice by Inhibiting Adipocyte Differentiation. Biol. Pharm. Bull. 2013, 36, 102–107. [Google Scholar] [CrossRef]
  50. Zheng, H.; Jeong, Y.; Song, J.; Ji, G.E. Oral administration of ginsenoside Rh1 inhibits the development of atopic dermatitis-like skin lesions induced by oxazolone in hairless mice. Int. Immunopharmacol. 2011, 11, 511–518. [Google Scholar] [CrossRef]
  51. Liu, X.; Zhang, Z.; Liu, J.; Wang, Y.; Zhou, Q.; Wang, S.; Wang, X. Ginsenoside Rg3 improves cyclophosphamide-induced immunocompetence in Balb/c mice. Int. Immunopharmacol. 2019, 72, 98–111. [Google Scholar] [CrossRef]
  52. Zhuang, Z.; Chen, Q.; Zhong, X.; Chen, H.; Yu, R.; Tang, Y. Ginsenoside Rg3, a promising agent for NSCLC patients in the pandemic: A large-scale data mining and systemic biological analysis. J. Ginseng Res. 2023, 47, 291–301. [Google Scholar] [CrossRef] [PubMed]
  53. Lu, S.; Zhang, Y.; Li, H.; Zhang, J.; Ci, Y.; Han, M. Ginsenoside Rb1 can ameliorate the key in nti-in cytokines TNF-alpha and IL-6 in a cancer cachexia mouse model. BMC Complement. Med. Ther. 2020, 20, 11. [Google Scholar] [CrossRef] [PubMed]
  54. Shi, J.; Weng, J.-H.; Mitchison, T.J. Immunomodulatory drug discovery from herbal medicines: Insights from organ-specific activity and xenobiotic defenses. Elife 2021, 10, e73673. [Google Scholar] [CrossRef]
  55. Ning, C.; Gao, X.; Wang, C.; Kong, Y.; Liu, Z.; Sun, H.; Sun, P.; Huo, X.; Ma, X.; Meng, Q.; et al. Ginsenoside Rg1 protects against acetaminophen-induced liver injury via activating Nrf2 signaling pathway in vivo and in vitro. Regul. Toxicol. Pharmacol. 2018, 98, 58–68. [Google Scholar] [CrossRef] [PubMed]
  56. Gao, Y.; Chu, S.; Zhang, Z.; Chen, N. Hepataprotective effects of ginsenoside Rg1—A review. J. Ethnopharmacol. 2017, 206, 178–183. [Google Scholar] [CrossRef]
  57. Kang, S.W.; Min, H. Ginseng, the ‘Immunity Boost’: The Effects of Panax ginseng on Immune System. J. Ginseng Res. 2012, 36, 354–368. [Google Scholar] [CrossRef] [PubMed]
  58. Lee, J.O.; Choi, E.; Shin, K.K.; Hong, Y.H.; Kim, H.G.; Jeong, D.; Hossain, M.A.; Kim, H.S.; Yi, Y.S.; Kim, D.; et al. Compound K, a ginsenoside metabolite, plays an antiinflammatory role in macrophages by targeting the AKT1-mediated signaling pathway. J. Ginseng Res. 2019, 43, 154–160. [Google Scholar] [CrossRef] [PubMed]
  59. Sharma, A.; Lee, H.-J. Ginsenoside Compound K: Insights into Recent Studies on Pharmacokinetics and Health-Promoting Activities. Biomolecules 2020, 10, 1028. [Google Scholar] [CrossRef]
  60. Yang, X.-D.; Yang, Y.-Y.; Ouyang, D.-S.; Yang, G.-P. A review of biotransformation and pharmacology of ginsenoside compound K. Fitoterapia 2015, 100, 208–220. [Google Scholar] [CrossRef]
  61. Chen, J.; Si, M.; Wang, Y.; Liu, L.; Zhang, Y.; Zhou, A.; Wei, W. Ginsenoside metabolite compound K exerts anti-inflammatory and analgesic effects via downregulating COX2. Inflammopharmacology 2018, 27, 157–166. [Google Scholar] [CrossRef]
  62. Lee, Y.J.; Song, K.Y.; Lee, E.Y.; Kang, H.S.; Song, Y.W. Compound K, a Metabolite of Ginsenosides, Attenuates Collagen-induced Arthritis in Mice. J. Rheum. Dis. 2015, 22, 154–166. [Google Scholar] [CrossRef]
  63. Wang, R.; Zhang, M.; Hu, S.; Liu, K.; Tai, Y.; Tao, J.; Zhou, W.; Zhao, Z.; Wang, Q.; Wei, W. Ginsenoside metabolite compound-K regulates macrophage function through inhibition of β-arrestin2. Biomed. Pharmacother. 2019, 115, 108909. [Google Scholar] [CrossRef] [PubMed]
  64. Cuong, T.T.; Yang, C.-S.; Yuk, J.-M.; Lee, H.-M.; Ko, S.-R.; Cho, B.-G.; Jo, E.-K. Glucocorticoid receptor agonist compound K regulates dectin-1-dependent inflammatory signaling through inhibition of reactive oxygen species. Life Sci. 2009, 85, 625–633. [Google Scholar] [CrossRef] [PubMed]
  65. Yang, C.-S.; Ko, S.-R.; Cho, B.-G.; Shin, N.-M.; Yuk, J.-M.; Li, S.; Kim, J.-M.; Evans, R.M.; Jung, J.-S.; Song, N.-K.; et al. The ginsenoside metabolite compound K, a novel agonist of glucocorticoid receptor, induces tolerance to endotoxin-induced lethal shock. J. Cell. Mol. Med. 2008, 12, 1739–1753. [Google Scholar] [CrossRef]
  66. Park, J.-S.; Shin, J.A.; Jung, J.-S.; Hyun, J.-W.; Van Le, T.K.; Kim, D.-H.; Park, E.-M.; Kim, H.-S. Anti-Inflammatory Mechanism of Compound K in Activated Microglia and Its Neuroprotective Effect on Experimental Stroke in Mice. J. Pharmacol. Exp. Ther. 2012, 341, 59–67. [Google Scholar] [CrossRef] [PubMed]
  67. Li, J.; Zhong, W.; Wang, W.; Hu, S.; Yuan, J.; Zhang, B.; Hu, T.; Song, G. Ginsenoside metabolite compound K promotes recovery of dextran sulfate sodium-induced colitis and inhibits inflammatory responses by suppressing NF-κB activation. PLoS ONE 2014, 9, e87810. [Google Scholar] [CrossRef] [PubMed]
  68. Choi, K.; Kim, M.; Ryu, J.; Choi, C. Ginsenosides compound K and Rh 2 inhibit tumor necrosis factor-α-induced activation of the NF-κB and JNK pathways in human astroglial cells. Neurosci. Lett. 2007, 421, 37–41. [Google Scholar] [CrossRef]
  69. Abd Eldaim, M.A.; Abd El Latif, A.S.; Hassan, A.; El-Borai, N.B. Ginseng attenuates fipronil-induced hepatorenal toxicity via its antioxidant, anti-apoptotic, and anti-inflammatory activities in rats. Environ. Sci. Pollut. Res. 2020, 27, 45008–45017. [Google Scholar] [CrossRef]
  70. Lee, J.O.; Yang, Y.; Tao, Y.; Yi, Y.S.; Cho, J.Y. Korean Red Ginseng saponin fraction exerts anti-inflammatory effects by targeting the NF-κB and AP-1 pathways. J. Ginseng Res. 2022, 46, 489–495. [Google Scholar] [CrossRef]
  71. Baek, K.-S.; Yi, Y.-S.; Son, Y.-J.; Yoo, S.; Sung, N.Y.; Kim, Y.; Hong, S.; Aravinthan, A.; Kim, J.-H.; Cho, J.Y. In vitro and in vivo anti-inflammatory activities of Korean Red Ginseng-derived components. J. Ginseng Res. 2016, 40, 437–444. [Google Scholar] [CrossRef]
  72. Byun, J.; Kim, S.K.; Ban, J.Y. Anti-Inflammatory and Anti-Oxidant Effects of Korean Ginseng Berry Extract in LPS-Activated RAW264.7 Macrophages. Am. J. Chin. Med. 2021, 49, 719–735. [Google Scholar] [CrossRef] [PubMed]
  73. Jung, D.-H.; Nahar, J.; Mathiyalagan, R.; Rupa, E.J.; Ramadhania, Z.M.; Han, Y.; Yang, D.-C.; Kang, S.C. Focused Review on Molecular Signalling Mechanisms of Ginsenosides on Anti-lung cancer and Anti-inflammatory Activities. Anti-Cancer Agents Med. Chem. 2023, 23, 3–14. [Google Scholar] [CrossRef]
  74. Liu, X.; Wang, Z.; Qian, H.; Tao, W.; Zhang, Y.; Hu, C.; Mao, W.; Guo, Q. Natural medicines of targeted rheumatoid arthritis and its action mechanism. Front. Immunol. 2022, 13, 945129. [Google Scholar] [CrossRef] [PubMed]
  75. Gao, Y.; Li, J.; Wang, J.; Li, X.; Li, J.; Chu, S.; Li, L.; Chen, N.; Zhang, L. Ginsenoside Rg1 prevent and treat inflammatory diseases: A review. Int. Immunopharmacol. 2020, 87, 106805. [Google Scholar] [CrossRef] [PubMed]
  76. Joo, S.S.; Yoo, Y.M.; Ahn, B.W.; Nam, S.Y.; Kim, Y.-B.; Hwang, K.W.; Lee, D.I. Prevention of Inflammation-Mediated Neurotoxicity by Rg3 and Its Role in Microglial Activation. Biol. Pharm. Bull. 2008, 31, 1392–1396. [Google Scholar] [CrossRef]
  77. 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]
  78. Wang, Y.; Liu, Y.; Zhang, X.-Y.; Xu, L.-H.; Ouyang, D.-Y.; Liu, K.-P.; Pan, H.; He, J.; He, X.-H. Ginsenoside Rg1 regulates innate immune responses in macrophages through differentially modulating the NF-κB and PI3K/Akt/mTOR pathways. Int. Immunopharmacol. 2014, 23, 77–84. [Google Scholar] [CrossRef] [PubMed]
  79. Gao, Y.; Chu, S.; Li, J.; Li, J.; Zhang, Z.; Xia, C.; Heng, Y.; Zhang, M.; Hu, J.; Wei, G.; et al. Anti-inflammatory function of ginsenoside Rg1 on alcoholic hepatitis through glucocorticoid receptor related nuclear factor-kappa B pathway. J. Ethnopharmacol. 2015, 173, 231–240. [Google Scholar] [CrossRef] [PubMed]
  80. Lee, S.-Y.; Jeong, J.-J.; Eun, S.-H.; Kim, D.-H. Anti-inflammatory effects of ginsenoside Rg1 and its metabolites ginsenoside Rh1 and 20(S)-protopanaxatriol in mice with TNBS-induced colitis. Eur. J. Pharmacol. 2015, 762, 333–343. [Google Scholar] [CrossRef]
  81. Heo, H.; Kim, Y.; Cha, B.; Brito, S.; Kim, H.; Kim, H.; Fatombi, B.M.; Jung, S.Y.; Lee, S.M.; Lei, L.; et al. A systematic exploration of ginsenoside Rg5 reveals anti-inflammatory functions in airway mucosa cells. J. Ginseng Res. 2023, 47, 97–105. [Google Scholar] [CrossRef]
  82. Ye, R.; Yang, Q.; Kong, X.; Han, J.; Zhang, X.; Zhang, Y.; Li, P.; Liu, J.; Shi, M.; Xiong, L.; et al. Ginsenoside Rd attenuates early oxidative damage and sequential inflammatory response after transient focal ischemia in rats. Neurochem. Int. 2011, 58, 391–398. [Google Scholar] [CrossRef] [PubMed]
  83. Kim, D.H.; Chung, J.H.; Yoon, J.S.; Ha, Y.M.; Bae, S.; Lee, E.K.; Jung, K.J.; Kim, M.S.; Kim, Y.J.; Kim, M.K.; et al. Ginsenoside Rd inhibits the expressions of iNOS and COX-2 by suppressing NF-κB in LPS-stimulated RAW264.7 cells and mouse liver. J. Ginseng Res. 2013, 37, 54–63. [Google Scholar] [CrossRef] [PubMed]
  84. Yi, Y.-S. New mechanisms of ginseng saponin-mediated anti-in ammatory action via targeting canonical in ammasome signaling pathways. J. Ethnopharmacol. 2021, 278, 114292. [Google Scholar] [CrossRef] [PubMed]
  85. Wang, H.; Qi, J.; Li, L.; Wu, T.; Wang, Y.; Wang, X.; Ning, Q. Inhibitory effects of Chikusetsusaponin IVa on lipopolysaccharide-induced pro-inflammatory responses in THP-1 cells. Int. J. Immunopathol. Pharmacol. 2015, 28, 308–317. [Google Scholar] [CrossRef]
  86. Liu, Y.; Zhu, H.; Zhou, W.; Ye, Q. Anti-in ammatory and anti-gouty-arthritic effect of free Ginsenoside Rb1 and nano Ginsenoside Rb1 against MSU induced gouty arthritis in experimental animals. Chem. Biol. Interact. 2020, 332, 109285. [Google Scholar] [CrossRef]
  87. Joh, E.-H.; Lee, I.-A.; Jung, I.-H.; Kim, D.-H. Ginsenoside Rb1 and its metabolite compound K inhibit IRAK-1 activation—The key step of inflammation. Biochem. Pharmacol. 2011, 82, 278–286. [Google Scholar] [CrossRef]
  88. Kim, J.H.; Yi, Y.-S.; Kim, M.-Y.; Cho, J.Y. Role of ginsenosides, the main active components of Panax ginseng, in inflammatory responses and diseases. J. Ginseng Res. 2017, 41, 435–443. [Google Scholar] [CrossRef]
  89. Rhule, A.; Navarro, S.; Smith, J.R.; Shepherd, D.M. Panax notoginseng attenuates LPS-induced pro-inflammatory mediators in RAW264.7 cells. J. Ethnopharmacol. 2006, 106, 121–128. [Google Scholar] [CrossRef]
  90. Cho, J.Y.; Yoo, E.S.; Baik, K.U.; Park, M.H.; Han, B.H. In Vitro Inhibitory Effect of Protopanaxadiol Ginsenosides on Tumor Necrosis Factor (TNF)-α Production and its Modulation by Known TNF-α Antagonists. Planta Medica 2001, 67, 213–218. [Google Scholar] [CrossRef]
  91. Wu, C.F.; Bi, X.L.; Yang, J.Y.; Zhan, J.Y.; Dong, Y.X.; Wang, J.H.; Wang, J.M.; Zhang, R.; Li, X. Differential effects of ginsenosides on NO and TNF-α production by LPS-activated N9 microglia. Int. Immunopharmacol. 2007, 7, 313–320. [Google Scholar] [CrossRef]
  92. Jung, J.-S.; Kim, D.-H.; Kim, H.-S. Ginsenoside Rh1 suppresses inducible nitric oxide synthase gene expression in IFN-γ-stimulated microglia via modulation of JAK/STAT and ERK signaling pathways. Biochem. Biophys. Res. Commun. 2010, 397, 323–328. [Google Scholar] [CrossRef] [PubMed]
  93. Bae, E.A.; Kim, E.J.; Park, J.S.; Kim, H.S.; Ryu, J.H.; Kim, D.H. Ginsenosides Rg3 and Rh2 inhibit the activation of AP-1 and protein kinase A pathway in lipopolysaccharide/interferon-gamma-stimulated BV-2 microglial cells. Planta Med. 2006, 72, 627–633. [Google Scholar] [CrossRef] [PubMed]
  94. Kim, B.; Lee, Y.; Park, T.; Kim, H.; Rhee, M.; Cho, J. Ginsenoside Rp1, a ginsenoside derivative, blocks lipopolysaccharide-induced interleukin-1beta production via suppression of the NF-kappaB pathway. Planta Med. 2009, 75, 321–326. [Google Scholar] [CrossRef] [PubMed]
  95. Shen, T.; Lee, J.; Park, M.H.; Lee, Y.G.; Rho, H.S.; Kwak, Y.-S.; Rhee, M.H.; Park, Y.C.; Cho, J.Y. Ginsenoside Rp 1, a ginsenoside derivative, blocks promoter activation of iNOS and Cox-2 genes by suppression of an IKKβ-mediated NF-κB pathway in HEK293 cells. J. Ginseng Res. 2011, 35, 200–208. [Google Scholar] [CrossRef]
  96. Hwang, S.H.; Shin, T.-J.; Choi, S.-H.; Cho, H.-J.; Lee, B.-H.; Pyo, M.K.; Lee, J.-H.; Kang, J.; Kim, H.-J.; Park, C.-W.; et al. Gintonin, newly identified compounds from ginseng, is novel lysophosphatidic acids-protein complexes and activates G protein-coupled lysophosphatidic acid receptors with high affinity. Mol. Cells 2012, 33, 151–162. [Google Scholar] [CrossRef]
  97. 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]
  98. Yao, W.; Guan, Y. Ginsenosides in cancer: A focus on the regulation of cell metabolism. Biomed. Pharmacother. 2022, 156, 113756. [Google Scholar] [CrossRef]
  99. 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]
  100. Nag, S.A.; Qin, J.-J.; Wang, W.; Wang, M.-H.; Wang, H.; Zhang, R. Ginsenosides as anticancer agents: In vitro and in vivo activities, structure–activity relationships, and molecular mechanisms of action. Front. Pharmacol. 2012, 3, 25. [Google Scholar] [CrossRef]
  101. Kim, Y.-S.; Kim, J.-J.; Cho, K.-H.; Jung, W.-S.; Moon, S.-K.; Park, E.-K.; Kim, N.-H. Biotransformation of ginsenoside Rb1, crocin, amygdalin, geniposide, puerarin, ginsenoside Re, hesperidin, poncirin, glycyrrhizin, and baicalin by human fecal microflora and its relation to cytotoxicity against tumor cells. J. Microbiol. Biotechnol. 2008, 18, 1109–1114. [Google Scholar]
  102. Byun, B.H.; Shin, I.; Yoon, Y.-S.; Kim, S.I.L.; Joe, C.O. Modulation of Protein Kinase C Activity in NIH 3T3 Cells by Plant Glycosides from Panax ginseng. Planta Medica 1997, 63, 389–392. [Google Scholar] [CrossRef]
  103. Popovich, D.G.; Kitts, D.D. Structure–function relationship exists for ginsenosides in reducing cell proliferation and inducing apoptosis in the human leukemia (THP-1) cell line. Arch. Biochem. Biophys. 2002, 406, 1–8. [Google Scholar] [CrossRef] [PubMed]
  104. Odashima, S.; Ohta, T.; Kohno, H.; Matsuda, T.; Kitagawa, I.; Abe, H.; Arichi, S. Control of phenotypic expression of cultured B16 melanoma cells by plant glycosides. Cancer Res 1985, 45, 2781–2784. [Google Scholar] [PubMed]
  105. Li, X.; Chu, S.; Lin, M.; Gao, Y.; Liu, Y.; Yang, S.; Zhou, X.; Zhang, Y.; Hu, Y.; Wang, H.; et al. Anticancer property of ginsenoside Rh2 from ginseng. Eur. J. Med. Chem. 2020, 203, 112627. [Google Scholar] [CrossRef] [PubMed]
  106. Song, B.-K.; Kim, K.M.; Choi, K.-D.; Im, W.-T. Production of the Rare Ginsenoside Rh2-MIX (20(S)-Rh2, 20(R)-Rh2, Rk2, and Rh3) by Enzymatic Conversion Combined with Acid Treatment and Evaluation of Its Anti-Cancer Activity. J. Microbiol. Biotechnol. 2017, 27, 1233–1241. [Google Scholar] [CrossRef] [PubMed]
  107. Xiaodan, S.; Ying, C. Role of ginsenoside Rh2 in tumor therapy and tumor microenvironment immunomodulation. Biomed. Pharmacother. 2022, 156, 113912. [Google Scholar] [CrossRef]
  108. Wang, W.; Zhao, Y.; Rayburn, E.R.; Hill, D.L.; Wang, H.; Zhang, R. In vitro anti-cancer activity and structure–activity relationships of natural products isolated from fruits of Panax ginseng. Cancer Chemother. Pharmacol. 2007, 59, 589–601. [Google Scholar] [CrossRef] [PubMed]
  109. Li, Q.-F.; Shi, S.-L.; Liu, Q.-R.; Tang, J.; Song, J.; Liang, Y. Anticancer effects of ginsenoside Rg1, cinnamic acid, and tanshinone IIA in osteosarcoma MG-63 cells: Nuclear matrix downregulation and cytoplasmic trafficking of nucleophosmin. Int. J. Biochem. Cell Biol. 2008, 40, 1918–1929. [Google Scholar] [CrossRef] [PubMed]
  110. Wang, C.-Z.; Aung, H.; Ni, M.; Wu, J.-A.; Tong, R.; Wicks, S.; He, T.-C.; Yuan, C.-S. Red American Ginseng: Ginsenoside Constituents and Antiproliferative Activities of Heat-Processed Panax quinquefolius Roots. Planta Medica 2007, 73, 669–674. [Google Scholar] [CrossRef]
  111. Kang, D.-I.; Lee, J.-Y.; Yang, J.-Y.; Jeong, S.M.; Lee, J.; Nah, S.-Y.; Kim, Y. Evidence that the tertiary structure of 20(S)-ginsenoside Rg3 with tight hydrophobic packing near the chiral center is important for Na+ channel regulation. Biochem. Biophys. Res. Commun. 2005, 333, 1194–1201. [Google Scholar] [CrossRef]
  112. Nakhjavani, M.; Smith, E.; Townsend, A.R.; Price, T.J.; Hardingham, J.E. Anti-Angiogenic Properties of Ginsenoside Rg3. Molecules 2020, 25, 4905. [Google Scholar] [CrossRef]
  113. Nakhjavani, M.; Palethorpe, H.M.; Tomita, Y.; Smith, E.; Price, T.J.; Yool, A.J.; Pei, J.V.; Townsend, A.R.; Hardingham, J.E. Stereoselective Anti-Cancer Activities of Ginsenoside Rg3 on Triple Negative Breast Cancer Cell Models. Pharmaceuticals 2019, 12, 117. [Google Scholar] [CrossRef] [PubMed]
  114. Yue, P.Y.; Wong, D.Y.; Wu, P.; Leung, P.; Mak, N.; Yeung, H.; Liu, L.; Cai, Z.; Jiang, Z.-H.; Fan, T.; et al. The angiosuppressive effects of 20(R)- ginsenoside Rg3. Biochem. Pharmacol. 2006, 72, 437–445. [Google Scholar] [CrossRef]
  115. Xu, T.-M.; Cui, M.-H.; Xin, Y.; Gu, L.-P.; Jiang, X.; Su, M.-M.; Wang, D.-D.; Wang, W.-J. Inhibitory effect of ginsenoside Rg3 on ovarian cancer metastasis. Chin. Med. J. 2008, 121, 1394–1397. [Google Scholar] [CrossRef] [PubMed]
  116. Lee, K.Y.; Lee, Y.H.; Kim, S.I.; Park, J.H.; Lee, S.K. Ginsenoside-Rg5 suppresses cyclin E-dependent protein kinase activity via up-regulating p21Cip/WAF1 and down-regulating cyclin E in SK-HEP-1 cells. Anticancer. Res. 1997, 17, 1067–1072. [Google Scholar] [PubMed]
  117. Woo, L.K.; Nakamura, Y.; Donati, L. Effect of Korean Ginseng on the growth rate of cells. Arch. Ital. Patol. Clin. Tumori 1965, 8, 53–61. [Google Scholar]
  118. Kim, H.D.; Ha, S.E.; Kang, J.R.; Park, J.K. Effect of Korean Red Ginseng Extract on Cell Death Responses in Peroxynitrite-Treated Keratinocytes. J. Ginseng Res. 2010, 34, 205–211. [Google Scholar] [CrossRef]
  119. Cho, C.K.; Kim, T.H.; Yoo, S.Y.; Koh, K.H.; Kim, M.S.; Kim, J.H.; Kim, S.H.; Yoon, H.K.; Ji, Y.H. The Effects of Akaloid Fraction of Korean Ginseng on the Radiation-Induced DNA Strand Breaks. J. Korean Soc. Ther. Radiol. 1995, 13, 113–120. [Google Scholar]
  120. Zhou, W.; Feng, M.-Q.; Li, J.-Y.; Zhou, P. Studies on the preparation, crystal structure and bioactivity of ginsenoside compound K. J. Asian Nat. Prod. Res. 2006, 8, 519–527. [Google Scholar] [CrossRef]
  121. Liu, J.; Wang, Y.; Yu, Z.; Lv, G.; Huang, X.; Lin, H.; Ma, C.; Lin, Z.; Qu, P. Functional Mechanism of Ginsenoside Compound K on Tumor Growth and Metastasis. Integr. Cancer Ther. 2022, 21, 15347354221101203. [Google Scholar] [CrossRef]
  122. Li, C.; Dong, Y.; Wang, L.; Xu, G.; Yang, Q.; Tang, X.; Qiao, Y.; Cong, Z. Ginsenoside metabolite compound K induces apoptosis and autophagy in non-small cell lung cancer cells via AMPK-mTOR and JNK pathways. Biochem. Cell Biol. 2019, 97, 406–414. [Google Scholar] [CrossRef] [PubMed]
  123. Wang, H.; Jiang, D.; Liu, J.; Ye, S.; Xiao, S.; Wang, W.; Sun, Z.; Xie, Y.; Wang, J. Compound K Induces Apoptosis of Bladder Cancer T24 Cells Via Reactive Oxygen Species-Mediated p38 MAPK Pathway. Cancer Biotherapy Radiopharm. 2013, 28, 607–614. [Google Scholar] [CrossRef] [PubMed]
  124. Hyun, S.H.; Kim, S.W.; Seo, H.W.; Youn, S.H.; Kyung, J.S.; Lee, Y.Y.; In, G.; Park, C.-K.; Han, C.-K. Physiological and pharmacological features of the non-saponin components in Korean Red Ginseng. J. Ginseng Res. 2020, 44, 527–537. [Google Scholar] [CrossRef] [PubMed]
  125. He, S.; Lyu, F.; Lou, L.; Liu, L.; Li, S.; Jakowitsch, J.; Ma, Y. Anti-tumor activities of Panax quinquefolius saponins and potential biomarkers in prostate cancer. J. Ginseng Res. 2021, 45, 273–286. [Google Scholar] [CrossRef] [PubMed]
  126. Deng, S.; Wong, C.K.C.; Lai, H.-C.; Wong, A.S.T. Ginsenoside-Rb1 targets chemotherapy-resistant ovarian cancer stem cells via simultaneous inhibition of Wnt/β-catenin signaling and epithelial-to-mesenchymal transition. Oncotarget 2017, 8, 25897–25914. [Google Scholar] [CrossRef] [PubMed]
  127. Liu, D.; Liu, T.; Teng, Y.; Chen, W.; Zhao, L.; Li, X. Ginsenoside Rb1 inhibits hypoxia-induced epithelial-mesenchymal transition in ovarian cancer cells by regulating microRNA-25. Exp. Ther. Med. 2017, 14, 2895–2902. [Google Scholar] [CrossRef] [PubMed]
  128. Huang, G.; Khan, I.; Li, X.; Chen, L.; Leong, W.; Ho, L.T.; Hsiao, W.L.W. Ginsenosides Rb3 and Rd reduce polyps formation while reinstate the dysbiotic gut microbiota and the intestinal microenvironment in ApcMin/+ mice. Sci. Rep. 2017, 7, 12552. [Google Scholar] [CrossRef]
  129. Kim, J.-K.; Kim, J.-Y.; Kim, H.-J.; Park, K.-G.; Harris, R.A.; Cho, W.-J.; Lee, J.-T.; Lee, I.-K. Scoparone Exerts Anti-Tumor Activity against DU145 Prostate Cancer Cells via Inhibition of STAT3 Activity. PLoS ONE 2013, 8, e80391. [Google Scholar] [CrossRef]
  130. Yoo, J.-H.; Kwon, H.C.; Kim, Y.-J.; Park, J.H.; Yang, H.O. KG-135, enriched with selected ginsenosides, inhibits the proliferation of human prostate cancer cells in culture and inhibits xenograft growth in athymic mice. Cancer Lett. 2010, 289, 99–110. [Google Scholar] [CrossRef]
  131. Sivoňová, M.K.; Vilčková, M.; Kliment, J.; Mahmood, S.; Jurečeková, J.; Dušenková, S.; Waczulíková, I.; Slezák, P.; Dobrota, D. Association of p53 and p21 polymorphisms with prostate cancer. Biomed. Rep. 2015, 3, 707–714. [Google Scholar] [CrossRef]
  132. Liu, W.-J.; Tang, H.-T.; Jia, Y.-T.; Ma, B.; Fu, J.-F.; Wang, Y.; Lv, K.-Y.; Xia, Z.-F. Notoginsenoside R1 Attenuates Renal Ischemia-Reperfusion Injury in Rats. Shock 2010, 34, 314–320. [Google Scholar] [CrossRef]
  133. Wang, W.; Rayburn, E.R.; Hao, M.; Zhao, Y.; Hill, D.L.; Zhang, R.; Wang, H. Experimental therapy of prostate cancer with novel natural product anti-cancer ginsenosides. Prostate 2008, 68, 809–819. [Google Scholar] [CrossRef] [PubMed]
  134. Wang, W.; Rayburn, E.R.; Hang, J.; Zhao, Y.; Wang, H.; Zhang, R. Anti-lung cancer effects of novel ginsenoside 25-OCH3-PPD. Lung Cancer 2009, 65, 306–311. [Google Scholar] [CrossRef] [PubMed]
  135. Wang, W.; Wang, H.; Rayburn, E.R.; Zhao, Y.; Hill, D.L.; Zhang, R. 20(S)-25-methoxyl-dammarane-3β, 12β, 20-triol, a novel natural product for prostate cancer therapy: Activity in vitro and in vivo and mechanisms of action. Br. J. Cancer 2008, 98, 792–802. [Google Scholar] [CrossRef]
  136. Yuan, C.-S.; Wang, C.-Z.; Li, X.-L.; Wang, Q.-F.; Mehendale, S.R.; Fishbein, A.; Han, A.H.; Sun, S. The mitochondrial pathway is involved in American ginseng-induced apoptosis of SW-480 colon cancer cells. Oncol. Rep. 2009, 21, 577–584. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Comparison of the molecular structure of a human endogenous anti-inflammatory drug (hydrocortisone) with ginsenosides.
Figure 1. Comparison of the molecular structure of a human endogenous anti-inflammatory drug (hydrocortisone) with ginsenosides.
Molecules 28 03863 g001
Figure 2. Diversity of protopanaxadiol and protopanaxtriol-type ginsenosides, characterized by the glucidic substituents they possess, including Glu (glucose), Arbp (arabinose in pyranose form), Xyl (xylose), Arbf (arabinose in furanose form), and Rham (rhamnose). The superscript notation of the sugar indicates the specific carbon involved in the bond. Other molecules mentioned in this study are also included in the figure.
Figure 2. Diversity of protopanaxadiol and protopanaxtriol-type ginsenosides, characterized by the glucidic substituents they possess, including Glu (glucose), Arbp (arabinose in pyranose form), Xyl (xylose), Arbf (arabinose in furanose form), and Rham (rhamnose). The superscript notation of the sugar indicates the specific carbon involved in the bond. Other molecules mentioned in this study are also included in the figure.
Molecules 28 03863 g002
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

Valdés-González, J.A.; Sánchez, M.; Moratilla-Rivera, I.; Iglesias, I.; Gómez-Serranillos, M.P. Immunomodulatory, Anti-Inflammatory, and Anti-Cancer Properties of Ginseng: A Pharmacological Update. Molecules 2023, 28, 3863. https://doi.org/10.3390/molecules28093863

AMA Style

Valdés-González JA, Sánchez M, Moratilla-Rivera I, Iglesias I, Gómez-Serranillos MP. Immunomodulatory, Anti-Inflammatory, and Anti-Cancer Properties of Ginseng: A Pharmacological Update. Molecules. 2023; 28(9):3863. https://doi.org/10.3390/molecules28093863

Chicago/Turabian Style

Valdés-González, Jose Antonio, Marta Sánchez, Ignacio Moratilla-Rivera, Irene Iglesias, and María Pilar Gómez-Serranillos. 2023. "Immunomodulatory, Anti-Inflammatory, and Anti-Cancer Properties of Ginseng: A Pharmacological Update" Molecules 28, no. 9: 3863. https://doi.org/10.3390/molecules28093863

Article Metrics

Back to TopTop