Next Article in Journal
Effects of Harvest Timing on Phytochemical Composition in Lamiaceae Plants under an Environment-Controlled System
Next Article in Special Issue
Antioxidant Capacity of Free and Bound Phenolics from Olive Leaves: In Vitro and In Vivo Responses
Previous Article in Journal
Curcumin Electrochemistry—Antioxidant Activity Assessment, Voltammetric Behavior and Quantitative Determination, Applications as Electrode Modifier
Previous Article in Special Issue
A Comparative Study on UHPLC-HRMS Profiles and Biological Activities of Inula sarana Different Extracts and Its Beta-Cyclodextrin Complex: Effective Insights for Novel Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Natural Bio-Compounds from Ganoderma lucidum and Their Beneficial Biological Actions for Anticancer Application: A Review

by
Emin Cadar
1,†,
Ticuta Negreanu-Pirjol
1,2,*,
Carolina Pascale
3,†,
Rodica Sirbu
3,*,
Irina Prasacu
4,
Bogdan-Stefan Negreanu-Pirjol
1,†,
Cezar Laurentiu Tomescu
5,6 and
Ana-Maria Ionescu
5,7
1
Faculty of Pharmacy, “Ovidius” University of Constanta, Capitan Aviator Al. Serbanescu Street, No. 6, Campus, Building C, 900470 Constanta, Romania
2
Academy of Romanian Scientists, Ilfov Street, No. 3, 050044 Bucharest, Romania
3
Organizing Institution for Doctoral University Studies of “Carol Davila”, University of Medicine and Pharmacy of Bucharest, Dionisie Lupu Street, No. 37, Sector 2, 020021 Bucharest, Romania
4
Faculty of Pharmacy, “Carol Davila” University of Medicine and Pharmacy of Bucharest, Traian Vuia Street, No. 6, Sector 2, 020956 Bucharest, Romania
5
Faculty of Medicine, “Ovidius” University of Constanta, University Alley, No. 1, Campus, Building B, 900470 Constanta, Romania
6
“Sf. Ap. Andrei” County Clinical Emergency Hospital, Tomis Bvd., No. 145, 900591 Constanta, Romania
7
Clinical Hospital C F Constanta, 1 Mai Bvd., No. 3–5, 900123 Constanta, Romania
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Antioxidants 2023, 12(11), 1907; https://doi.org/10.3390/antiox12111907
Submission received: 26 August 2023 / Revised: 19 October 2023 / Accepted: 23 October 2023 / Published: 25 October 2023

Abstract

:
Ganoderma lucidum (G. lucidum) has been known for many centuries in Asian countries under different names, varying depending on the country. The objective of this review is to investigate the scientific research on the natural active bio-compounds in extracts obtained from G. lucidum with significant biological actions in the treatment of cancer. This review presents the classes of bio-compounds existing in G. lucidum that have been reported over time in the main databases and have shown important biological actions in the treatment of cancer. The results highlight the fact that G. lucidum possesses important bioactive compounds such as polysaccharides, triterpenoids, sterols, proteins, nucleotides, fatty acids, vitamins, and minerals, which have been demonstrated to exhibit multiple anticancer effects, namely immunomodulatory, anti-proliferative, cytotoxic, and antioxidant action. The potential health benefits of G. lucidum are systematized based on biological actions. The findings present evidence regarding the lack of certainty about the effects of G. lucidum bio-compounds in treating different forms of cancer, which may be due to the use of different types of Ganoderma formulations, differences in the study populations, or due to drug–disease interactions. In the future, larger clinical trials are needed to clarify the potential benefits of pharmaceutical preparations of G. lucidum, standardized by the known active components in the prevention and treatment of cancer.

1. Introduction

Ganoderma lucidum (G. lucidum), (Fr.) Karst is a medicinal mushroom known in traditional Asian medicine under different names depending on the country: Lingzhi (China), Reishi, and Mannentake (Japan), Linh chi (Vietnam), and Yeong Ji or Yung Gee (Republic of Korea) [1]. It is considered a source of longevity and health promotion [2,3,4]. The traditional medicinal uses of these mushrooms in Chinese and Japanese folk medicine as health remedies and herbal supplements are widely recognized [5,6,7]. The knowledge and use of medicinal mushrooms have preoccupied researchers, who have developed valuable studies on their bioactive components and their importance [8,9,10,11,12]. According to the World Health Organization, it has been estimated that around three-quarters of the world’s population relies on traditional medicines to maintain their health [13]. In traditional oriental medicine, G. lucidum has been used to treat several diseases, such as asthma, bronchitis, arthritis, hypertension, insomnia, diabetes, liver disease, nephritis, and cancer [14,15,16]. In ancient Chinese medicine, G. lucidum has been used for longevity and as an anticancer and antioxidant agent [17,18,19,20]. In the last 30 years, data have been reported on the chemical composition of G. lucidum extracts, justifying their biological activity and numerous health benefits. Among the most important bioactive compounds are polysaccharides and triterpenes [21,22,23,24,25,26]. Numerous other bioactive metabolites, such as proteins, peptides, sterols, lectins, adenosine, vitamins, and metals, have also been identified [27,28,29,30,31]. However, it should be noted that there are differences in the composition of G. lucidum products due to cultivation conditions, the extraction process of the bio-compounds, and the origin or part of the fungus used (fruiting body, mycelium, or spores) [14,28,30]. Nevertheless, many bioactive compounds in G. lucidum exhibit antioxidant, antitumor, and anti-inflammatory properties [32,33,34,35,36,37]. Such benefits have been investigated in multiple studies on their effect against prostate cancer, ovarian cancer, hepatocellular carcinoma, the induction of apoptosis in colon cancer, and the inhibition of angiogenesis [38,39,40,41,42,43]. Different immunomodulatory activities have been reported [44,45,46,47]. Other researchers have also reported studies on antioxidant and antioxidative stress actions [48,49,50]. With the development of modern research techniques to identify the chemical compositions of G. lucidum compounds, data may be accessible for use in medical research. A detailed understanding of these biological mechanisms could greatly influence and extend their benefits to human health.
The present study aims to review the data on active bio-compounds with important biological actions against cancer. Factors influencing the composition of the G. lucidum fungus, such as growing conditions (in the wild, in deciduous forests, or in special crops) and territorial areas, are analyzed. The characteristics of the G. lucidum fungus, as well as the chemo-bioactive compounds identified in G. lucidum with anticancer activities, its biological activities against cancer, and its toxicity and safety, are described in separate sections. The mechanisms of anticancer effects, including other biological activities that contribute to the fight against cancer, such as immunomodulatory, antioxidant, and cytotoxic actions, are also presented. Some negative cases in which G. lucidum bio-compounds were administered to patients undergoing treatment for several conditions are included as well. In these patients, the effects of G. lucidum treatments were not beneficial due to the occurrence of adverse effects.

2. Characteristics of Ganoderma lucidum Fungus

2.1. Description and Spread of the Fungus

The first description of Ganoderma lucidum (Curtis) P. Karst. was made by Curtis in England, and this description was officially recorded by Fries [51]. Initially, Cao et al. claimed that this fungus was already known as “Lingzhi”, a medicinal fungus identified and used in China for more than 2000 years [52]. Later, based on molecular studies, it was established that the East Asian medicinal mushroom is a different species from the G. lucidum mushroom [52]. After morphological and molecular examinations, Wang et al. confirmed that the Lingzhi species from China are related to G. lucidum from the UK and to other Ganoderma species [53]. Kwon et al. conducted phylogenetic analyses of Ganoderma species and showed that there are 62 strains of Ganoderma [54]. According to the taxonomic classification established by Nahata A., the species Ganoderma lucidum (Curt: Fr.) Karst belongs to the kingdom Fungi, phylum Basidiomycota, class Agaricomycetes, order Polyporales, family Ganodermataceae, genus Ganoderma, and species lucidum [55]. G. lucidum from the UK and other related Ganoderma species have also been morphologically and molecularly examined by different researchers who have conducted important studies, such as Kim et al., Park et al., and Liao et al. [56,57,58]. Several phylogenetically related Ganoderma species have been found in North America, Europe, and Asian countries [56,58]. Gottlieb et al. also performed molecular and morphological studies on the Ganoderma species collected from South America, followed by Eyssartier et al. from France [59,60]. In 2017, Copot et al. identified the fungus G. lucidum in hilly and mountainous areas in Romania, specifically in oak forests [61]. The description of G. lucidum collected from Romania highlighted the mushroom cap, which is kidney-shaped and can be up to 20 cm in diameter, ranging in color from red to pinkish-brown when mature [61]. Towards the edges, the coloration of the cap ranged from bright yellow to white. The spore print is brown (see Figure 1a,b) [61]. Figure 1c shows the fungus G. lucidum (Reishi or Lingzhi) adapted from studies by Parepalli et al. [62].

2.2. Data on G. lucidum Cultivation

G. lucidum can also grow in greenhouses under controlled conditions. Data on cultivation methods have been reported in the literature, such as the study of Boh et al., where different biotechnological cultivation methods are presented [14]. As G. lucidum is rare in nature, cultivation in greenhouses is practiced using two methods: cultivation of fruiting bodies on wooden logs or on sawdust bags (or large plastic bottles) [14]. The main cultivation methods to produce G. lucidum (fruiting body and mycelia) are presented in Figure 2 [14].
Various other cultivation methods have also been practiced. For example, Nithya et al. conducted research on the selection of an ideal material to grow G. lucidum fungus by testing wood retting, wheat bran, sorghum, and sorghum grains combined with chalk dust and gypsum [63]. Adongbede et al. (2021) used indigenous hardwoods supplemented with rice and wheat bran as substrates for growing G. lucidum in Nigeria, where the fungus is not naturally available [64]. Matute et al. (2002) grew G. lucidum in bags using sunflower hulls as a nutrient source [65]. Yang et al. (2003) used carbohydrate and nitrogen-rich residues from a rice bran distillery to grow G. lucidum in polypropylene bags [66]. Hsieh et al. (2004) used soybean residues as a nutrient for growing G. lucidum in polypropylene bags [67]. Chang et al. (2006) reported studies on the optimization of growth methods for G. lucidum [68]. It is worth noting that cultivation conditions and substrate composition influence the biochemical composition of G. lucidum fungi qualitatively and quantitatively, as confirmed by Baskar et al. in 2011 through their studies [69].

3. Chemical Bioactive Compounds Identified in G lucidum with Anticancer Actions

In the last 30 years, numerous studies have been conducted highlighting the nutritional potential of G. lucidum mushrooms [30]. Research on the biochemical composition of this mushroom has led to the identification of several categories of compounds with health-promoting biological activity, as reported by Ahmad et al. [7,13].

3.1. Proximate Composition for G. lucidum

The values of the bioactive compound classes differ quantitatively and qualitatively within certain limits, depending on the country and area of origin of the fungus and whether it is cultivated or naturally occurring in the forest. In the case of cultivated mushrooms, bioactive compounds depend on the nutrient richness of the substrate and environmental factors, such as humidity and temperature [14]. Most phytochemical reports show that, with regard to the chemical composition, the fruiting body of G. lucidum consists of 90% water and 10% different compounds, as presented in Table 1, according to Mau et al. [70].
Table 1 summarizes the information on the biochemical composition of the fungus, as noted by multiple researchers. The content analysis of the bioactive compound classes includes data on moisture, ash, water-soluble protein content, total lipid content, total carbohydrate content, and dietary fiber content. From Bangladesh, data were reported by Rahman et al. (2020), El Sheikha (2022), and Roy et al. (2018) [21,30,71]. The interest in the nutritional potential of G. lucidum in Taiwan was proven by studies conducted by Mau et al. (2001) [70]. The total fatty acid content was reported only by Fraile-Fabero et al. for G. lingzhi from China and G. lucidum from mushroom crops from Madrid, Spain, in 2021 [72]. Comparable data were also reported by Parapelli et al. (2021) for G. lucidum from India and by Ogbe et al. (2013) from Nigeria [73,74]. Additional studies were published by Wachtel-Galor et al. (2011), by Paterson et al. (2006), and by Garuba et al. (2000) [75,76,77]. The pharmacognostic review of the active compounds isolated from basidiocarp and mycelium of G. lucidum revealed that it contains polysaccharides, triterpenes, vitamins, minerals, sterols, proteins, proteo-polysaccharides, lectins, nucleotides, and fatty acids, as reported by Ahmad (2018) [13]. Figure 3 systematizes the categories of biochemical compounds that support the biological activities occurring in the treatment of various cancer tumors [13].

3.2. Polysaccharide Content of G. lucidum

Numerous studies have identified the existence of several types of polysaccharides in the fruiting body, spores, or mycelium of G. lucidum, as reported by Liu et al. and Lin et al. [78,79]. Polysaccharides represent a class of macromolecules with diverse structures and a wide range of physicochemical properties and biological actions, according to studies by Ferreira et al., Yu et al., and Giavasis [80,81,82]. According to Ahmad, Bhat et al., and Liu et al., these polysaccharide compounds are considered some of the most potent bioactive metabolites with antitumoral effects due to their biochemical structure [13,18,78]. Polysaccharide compounds (Gl-Ps) have a rich history and have been extensively studied in recent years by Sanodiya et al., Parepalli et al., and Chen et al. [27,62,83].

3.2.1. Extraction and Purification of Polysaccharides

The most widely used method for polysaccharide extraction from spores, fruiting bodies, and mycelium is hot-water extraction, as documented by Nie et al. [84]. Figure 4 illustrates the extraction of five polysaccharide fractions from G. lucidum [85]. The fruiting body of G. lucidum was initially peeled, shredded, and sieved in order to obtain a fine powder. The extraction of Gl-Ps from the mushroom powder was performed with distilled water at 80 °C [85]. There are researchers who have used diluted saline solution or alkaline acid solutions for polysaccharide extraction, as reported by Wang et al. (2011). Other extraction methods that were applied included microwaves, ultrasound, or enzymatic methods, as in the studies of Lin et al. (2005), Huang et al. (2010), Zhao et al. (2010), and Leong et al. (2021) [86,87,88,89,90]. After filtering, the solution is subjected to precipitation with alcohol or acetone to obtain crude polysaccharides [84]. Polysaccharide purification can be achieved by several techniques, such as ethanol fractionation or different column chromatographic techniques, as illustrated by Chen et al. (2008), Huang et al. (2011), and Jiang et al. (2012) [91,92,93]. Ion-exchange chromatography (DEAE-Sepharose Fast Flow), gel filtration, and affinity chromatography were the most useful methods, as highlighted by Choong et al. [85]. Figure 4 illustrates the extraction and fractionation steps of polysaccharides from G. lucidum. Five polysaccharide fractions were obtained [85].

3.2.2. Structure of Polysaccharides from G. lucidum

The investigation of the structures and chemical properties of polysaccharides from G. lucidum involves knowledge of the composition of existing monosaccharides, branching structures, types of glycosidic linkages, chain conformations, and molecular weights [18]. As investigation techniques have developed, it was found that G. lucidum can have polysaccharides in its composition, either as pure or linked with other proteins or peptides [94,95,96,97,98]. Bhat et al. (2021) showed that homo-glucans from G. lucidum are linear or branched biopolymers, possessing a backbone formed by α- or β-linked (1→3), (1→6)-β-glucan and (1→3)-α-glucan glucose units, and may possess side chains attached in different positions [18]. Further investigations by different authors explore bioactivities as well. Bao et al. (2002) determined that the structures of the backbone chains are linear or branched bipolymers [94]. These possess a backbone consisting of α- or β-linked glucose units [94]. In 2006, new structures with heteroglucans were outlined by Cao et al., and new structures with heteropolysaccharides were documented by Sullivan et al. [95,96]. Further types of ligands were presented by Li et al. (2007), Ye et al. (2008), Ye et al. (2009), Wang et al. (2009), Ye et al. (2010), and Liu et al. (2010) [97,98,99,100,101,102]. Pan D. et al. (2012), Ma et al. (2013), and Pan K. et al. (2013) also described the types of polysaccharide structures with different types of linkages, shown in detail in Table 2 [103,104,105]. Ooi et al. (2000) and Zhang et al. (2007) pointed out in their studies that the degree of base-chain substitution and branching chain length play important roles in determining the bioactivity of β-(1→3)-linked glucans [106,107]. Table 2 illustrates various techniques and methods utilized in the extraction, fractionation, and purification of polysaccharides and provides information about their structure (the backbone and monosaccharide compositions).
Side-chain branching occurs at C-6 of the main-chain glucosyl residues, as shown in Figure 5 and indicated by Choong et al. (2019) [85]. Jia et al. (2009) argued that among all homo-glucans, β-glucans are glucose polymers that exist as unbranched (1→3)-β-linked backbones in the form of (1→3)-β-linked backbones [108]. Figure 5 shows a Gl-Ps chain [85].
The molecular weight of Gl-Ps can range from 4 × 105 to 1 × 106 Da. Sanodiya et al. (2009) demonstrated that this has a significant impact on reducing cancer progression [27]. Studies by Moradali et al., Doco et al. (2001), and Hung et al. (2005) described polysaccharide-protein or peptide complexes using modern GC-MS methods [109,110,111]. Wang et al. (2002) and Sanodiya et al. (2009) investigated monosaccharides from the raw extract of G. lucidum [24,27]. Furthermore, Wang et al. (2009) and Ye et al. (2010) also provided data on monosaccharides identified in G. lucidum [100,101]. Similarly, Dai et al. (2010) and Yang et al. (2010) investigated G. lucidum and described the monosaccharide composition as predominantly consisting of xylose, fructose, glucose, and maltose [112,113]. Several other studies have been published reporting data on the polysaccharide structure and their medical applications, including Dong et al. (2012), Liu et al. (2012), Skalicka-Wozniak et al. (2012), and Pascale et al. (2022) [114,115,116,117].

3.3. Triterpene and Triterpenoid Content of G. lucidum

Triterpenes belong to the class of terpenes that have a molecule consisting of six isoprene units and are widespread in the plant kingdom [35]. Triterpenoids are part of the triterpene class, having heteroatoms (usually oxygen atoms) in the molecule. They are important bioactive compounds in the composition of G. lucidum, as shown by Wu et al. (2023) [118]. In 2018, Gu et al. revealed that the present triterpenes synthesize structures derived from lanosterol, which has a skeleton with a tetracyclic structure and the molecular formula C30H48 [119]. Terpenoid compounds from G. lucidum that have a C30 backbone and molecular masses between 400 and 600 kDa were studied by Baby et al. and Galappaththi et al., who generically named them Ganoderma triterpenoids (Gl-Ts) [120,121]. Further research on the structure and characterization of triterpenoids from G. lucidum that elucidates their biological activities was published in 2022 by Cör et al. and Lin et al. [122,123].

3.3.1. Physicochemical Determination and Analysis of Triterpene Compounds

Studies conducted by Ghorai et al. (2012), Taofiq et al. (2017), and Chang et al. (2012) utilizing UV spectrophotometric measurements have made significant contributions to the methods for determining total terpenoids [124,125,126]. Huie et al. (2004) employed chromatographic and electrophoretic methods for the analysis of triterpenoids, and Yang et al. (2007) utilized a combined approach involving HPLC-ESI-MS [127,128]. Chen et al. (2012) highlighted the existence of multiple methods for investigating terpenoids [83]. Triterpenoid analysis was also outlined in the studies of Zhang et al. (2008), Shi et al. (2010), and Hadda et al. (2015) [129,130,131]. Che et al., Hui et al., and Zhang et al. reported novel insights into triterpenoids obtained from G. lucidum [132,133,134]. In 2023, Wu et al. emphasized the necessity of a new research strategy intertwining the concepts of chemical component analysis and pharmacological activity [118].

3.3.2. Structure of G. lucidum Triterpenoids (Gl-Ts)

Wu et al. (2023) extensively reported on the diverse types of Gl-Ts identified from the mycelia, fruit body, and spores of G. lucidum, showcasing distinct structural variations, including alcohols, aldehydes, ketones, acids, esters, and various other substituents positioned differently [118]. Xia et al. (2014) conducted an in-depth analysis of the skeletal structure of G. lucidum, revealing that a majority of these terpenoids consist of 30 carbon atoms [135]. Using information derived from Wu et al., in Figure 6, the structures of a typical terpenoid skeleton are depicted, illustrating the numbered positions corresponding to the matched carbon atoms alongside 10 additional terpenoid structures exhibiting diverse substituents [118].
In Figure 6, illustrating the typical terpenoid skeleton structure, distinct substituents are evident at positions C-3, C-7, C-11, C-12, C-15, C-20, and C-27. Fatmawati et al. (2010) elucidated the structure of a new terpenoid, Ganoderic acid Df, featuring a β-hydroxy substituent at the C-11 position, distinguishing it from all other compounds characterized by a carbonyl group at the same position [136]. At the C-3 position, potential substituents include the β-hydroxy, carbonyl, and β-acetoxy groups. At C-20, a single carbon atom can host two substituents, which may be the methyl or hydroxyl groups or even hydrogen. Wu et al. demonstrated that at the C-25 position, various carboxyl groups, such as formyl, acetyl, or butyryl, can be found [118]. Additionally, Sharma et al. (2019) presented data on triterpenes from G. lucidum, characterized by isoprene units in their composition, featuring a C30 skeleton structure of ganoderic acids, aldehydes, esters, alcohols, lactones, glycosides, ketones, and molecular masses ranging from 400 to 600 g/mol [137]. Koo et al. (2019) identified a new compound with a lanostane triterpenoid structure named Ganosidone A, which, along with eight other derivatives, was investigated for its cancer chemopreventive potential [138]. Concerning the molecular configuration of ganoderic acids, Cör et al. (2022) documented the structures illustrated in Figure 7 [122].
Murata et al. (2019) identified and extracted a novel compound with a lanostane triterpene structure from G. lucidum using NMR and MS physicochemical analysis [139]. Studies presented by Yue et al. in 2008 and 2010, as well as those by Zhang et al. and Cheng et al., revealed the cytotoxic effects exerted by ganoderic acids from G. lucidum, which alter proteins involved in cell proliferation and cell death in carcinogenesis or oxidative stress [140,141,142,143]. In 2012, Liu et al. and Rios et al. studied the lanostanoid compounds from G. lucidum for their anticancer activities [144,145]. In 2013, Li et al. isolated a novel ganoderic acid from G. lucidum mycelia and studied its characteristics. Fatmawati et al. studied the structure–activity relationship of lanostane-type triterpenoids, and Li et al. reported cytotoxic effects in a wide range of triterpene compounds [146,147,148]. In 2017, Chen et al. presented data on compounds with triterpene structures existing in G. lucidum that exhibited activity as inhibitors of biological processes [149]. In 2019, Liang et al. illustrated their findings on the structure and mechanisms of action of ganoderic acids from G. lucidum [150]. In 2021, Chinthanom et al. documented their findings on lanostane triterpenoids isolated from mycelial cultures of Ganoderma spp. that can be modified by semisynthesis, thus obtaining synthetic compounds [151]. In 2023, Pascale et al. reported data on the specific structures and pharmacological mechanisms of triterpenoids with biological activities from G. lucidum [152]. Figure 8 illustrates other structures specific to the terpenoid compounds described by Pascale et al. [152].

3.4. Vitamins, Minerals, and Sterols Content

Vitamins, minerals, and sterols were among the 400 biologically active constituents recognized by various researchers after 2018, including Cör et al., Yang et al., and Ahmad, F. In terms of quantity, the vitamins were in the following order: B1, B2, B6, β-carotene, C, D, and E [153,154,155]. Hussein et al. (2022) also studied the existence of vitamins in G. lucidum [156]. El Sheikha et al. (2022) highlighted the highest vitamin content as consisting of niacin (B3) and ascorbic acid (C) [30]. Mineral contents have been reported by several studies, such as El Sheikha et al., Roy et al., Ogbe et al., Cör et al., and Hussein et al. [30,71,74,153,156]. Table 3 displays the values for mushroom compositions reported by various authors, expressed in milligrams per 100 g (mg/100 g), parts per million (ppm), and percentages (%) [30,71,74,153,156].
Similar data were also presented by Sharif et al. (2016) and Treviño et al. [157,158]. The bioaccumulation of copper and zinc in G. lucidum was also documented by Matute et al. in 2011 [159]. In 2008, Falandysz J. reported the existence of selenium in the composition of some antioxidant enzymes with protective actions that are involved in antitumor effects [160]. In their study, conducted in 2000, Chiu et al. identified the presence of germanium in the Ganoderma extract. Their research encompassed an assessment of genotoxicity and antigenotoxicity related to this finding [161]. Du et al. (2008) studied the positive effect of selenium on the immune regulation activity of G. lucidum, and the involvement of this element in anticancer activities was documented [162].
Sterols with the chemical formula C17H28O are a group of steroids that are cyclic secondary monohydric alcohols. Since 2002, Ma et al. have reported three new lanostanoids and two ergostane sterols, which were isolated by spectroscopic methods [163]. The role of sterols is important in the body’s metabolism in regulating some hormonal and immune system functions. Akihisa et al. (2007) studied the effects of sterols and triterpene acids from G. lucidum on anti-inflammatory and antitumor actions [164]. In 2011, the sterol content of G. lucidum was researched by Liu et al., who analyzed it qualitatively and quantitatively using HPLC methods [165]. In 2015, Baby et al. showed that sterols in G. lucidum can be classified based on their skeletons [120].

3.5. Protein, Lectin, and Amino Acid Content

3.5.1. Proteins and Peptides in G. lucidum

Proteins constitute a distinct category of biochemically active molecules in G. lucidum, as demonstrated by Cör et al. in 2018 [153]. Numerous studies have reported various biological effects of the primary fungal protein, Ling-zhi-8 (LZ-8). Structural reports were published by Huang et al. in 2009, and the immunomodulatory actions of LZ-8 were documented by Hsu et al. in 2013, Lin et al. in 2014, and Yang et al. [166,167,168,169]. In 2015, Sa-ard et al. studied crude proteins from G. lucidum (both mycelia and fruiting bodies) for their antioxidant actions [170]. In 2021, Fraile-Fabero et al. conducted studies on proteins from G. lucidum [72]. Additional research by Sun et al. in 2004 revealed the existence of compounds with a polysaccharide-peptide complex structure along with phenolic components with antioxidant potential in G. lucidum [171].
Ji et al. (2007) further documented the existence of proteo-polysaccharides in Ganoderma lucidum. They studied the immunomodulatory effects, emphasizing the anticancer properties of this fungus [172]. The immunomodulatory properties of LZ-8 protein and polysaccharides were studied by Yeh et al. in 2010 and Girjal et al. in 2012 [173,174]. Zhong et al. (2015) conducted research on the involvement of the peptide-polysaccharide complex from G. lucidum in oxidative stress [175]. Subsequent data on the presence of proteins in G. lucidum and their biological actions were published by Kumakura et al. in 2019, and further insights were provided by Yu et al. in 2021 and 2023, where they specifically documented the effects of proteoglycans derived from G. lucidum [176,177,178]. Huang et al. (2022) reported a novel pentapeptide in the mycelium of Ganoderma spp. that demonstrates antioxidant properties [179].

3.5.2. Lectins from G. lucidum

Lectins are glycoproteins found in the fruiting body of G. lucidum. Lectins are carbohydrate-binding proteins that have been classified according to their origin and structure. In 2007, Thakur et al. isolated and purified a group of lectin-structured proteins with a weight of 114 kDa from the fruiting bodies of G. lucidum [180]. In 2011, Girjal et al. isolated a new bioactive lectin from G. lucidum, which also exhibited hemagglutinating activity against both human and animal erythrocytes [181]. Nikitina et al. (2017) investigated lectins in mycelia and detected hemagglutinating activity in culture medium and in a crude mycelial extract of G. lucidum [182]. Recent research conducted by Yousra et al. has demonstrated varying binding patterns of lectins from G. lucidum with glycan moieties. These patterns play a crucial role in influencing biological activities, including antitumor, antiviral, and immunomodulatory effects [183].

3.5.3. Amino Acid Content of G. lucidum

The amino acid content has been documented by several authors. According to the research of Sanodiya et al., Yousra et al., and Deepalakshmi et al., the most abundant amino acid was glutamic acid [27,183,184]. This was followed by aspartic acid, glycine, and alanine. The amino acids in G. lucidum have also been analyzed by Zhang et al. in 2018, who reported 18 different types of amino acids [185]. It should be noted that factors such as the origin of the tested samples and the species of fungus, as well as different analysis techniques, may result in different amino acid values. Compounds with antioxidant properties from G. lucidum have been analyzed in numerous studies. Thus, Kim et al. (2008), Sheikh et al. (2014), Lin et al. (2015), and Veljović et al. (2017) presented data on the chemical compounds of G. lucidum [186,187,188,189]. The amino acid values for G. lucidum, as reported by Sanodiya et al. in 2009 and Deepalakshmi et al. in 2011, are presented in Table 4 [27,184].

3.5.4. Content of Compounds with Antioxidant Properties in G. lucidum

Dong et al. (2019), Zheng et al. (2020), Rahman et al. (2020), and Kolniak-Ostek et al. (2022) reported data regarding the total triterpenoid, polysaccharide, polyphenol (TPC), and flavonoid (TFC) content of G. lucidum [21,190,191,192]. Furthermore, Kim et al. (2008) analyzed the phenolic compounds from G. lucidum using HPLC and identified 28 phenolic compounds. All these compounds are responsible for antioxidant activity [186]. Triterpenoid compounds were documented by Lin et al. in 2015 and Kolniak-Ostek et al. in 2022 [188,192]. Polysaccharide compounds were described by Lin et al. in 2015 [188]. Polyphenolic compounds were investigated by Rahman et al. in 2020 and by Kolniak-Ostek et al. in 2022 [21,192]. Flavonoid compounds and ascorbic acid were detailed in a study by Rahman et al. in 2020 [21]. Depending on the area of origin of the mushroom (cultivated or forest) and the extraction techniques used, the concentrations of the compounds listed in Table 5 vary considerably. The terpenoid content is lower in cultivated G. lucidum, depending on the cultivation method, as Kolniak-Ostek et al. showed [192]. The results on compounds responsible for antioxidant activity are presented in Table 5.
Kolniak-Ostek et al. presented more data on several phenolic acids present in G. lucidum responsible for antioxidant activity, which are presented in Table 6 [192].
The antioxidant properties of G. lucidum were analyzed using several methods, including the reporting of DPPH and ABTS radical scavenging activities, as well as the FRAP assay. The results were expressed in units of measurement. These units vary according to the different experimental methods used. However, the importance of the results lies in the unequivocal demonstration of the antioxidant activity possessed by the constituents found in G. lucidum (see Table 7).
Furthermore, Dong et al. (2019) demonstrated a strong correlation between the antioxidant capacities measured using DPPH, ABTS, and FRAP assays and the content of polyphenolics and triterpenoids in G. lucidum [190]. Similar studies have also been published by Saltarelli et al. in 2015 [193]. The properties of compounds in G. lucidum and their effects on antioxidant activity were also documented by Tang et al. (2016), Sanchez C. (2017), Mohammadifar et al. (2020), and Mustafin et al. (2022) [194,195,196,197].

3.6. Content in Nucleosides and Fatty Acids

3.6.1. Nucleosides and Nucleobases in G. lucidum

Nucleotides are monomeric units that consist of a base (purine or pyrimidine) and one or more phosphate groups. Nucleotides are formed from nucleosides by phosphorylation under the action of kinases. Nucleosides are glycosyl amines derived from a nitrogenous base and a ribose or deoxyribose. These compounds have been shown to contain uridine and uracil, both of which are capable of reducing elevated serum aldolase levels. The water-soluble fraction of Ganoderma suppresses platelet aggregation. Cheung et al. have carried out studies on the identification and role of bases and nucleoside markers [198]. Gao et al. (2007) identified six nucleobases qualitatively in fruiting body samples of G. lucidum and G. sinense [199]. Additional studies of the distribution of nucleotides and nucleobases were also carried out by Yuan et al. in 2008 and Chen et al. in 2012, who were able to document 16 nucleotides and nucleobases in the Ganoderma species [200,201]. Phan et al. (2018) demonstrated in their work that nucleobases, nucleosides, and nucleotides found in fungi play crucial roles in regulating various physiological processes in the human body through purinergic and/or pyrimidine receptors. [202]. In 2022, Sheng et al. analyzed 10 target compounds from 23 batches of Ganoderma samples from different regions of China and demonstrated that the geographical origin of the fungi might be the exclusive factor affecting the accumulation of nucleosides and nucleobases in the Ganoderma spp. mycelium [203].

3.6.2. Fatty Acid Content

The total lipid content of G. lucidum has been reported in several studies by Rahman et al. (2020), El Sheikha (2022), Roy (2018), and Ogbe et al. (2013) [21,30,71,74]. Comparative data for the fatty acid content of carpophores from G. lingzhi (from industrial cultures in China) and G. lucidum (from cultures in Madrid, Spain) were described by Fraile-Fabero et al. in 2021 [72]. Fraile-Fabero et al. found that G. lingzhi contains the highest percentage of arachidic acid, followed by the margaric, behenic, margaroleic, lignoceric, and cis-vaccenic acids, while G. lucidum contains the highest percentage of α-linolenic acid, followed by the myristic, stearic, capric, erucic, nervonic, elaidic, octadecatrienoic, octadecadienoic, and eicosatrienoic acids [72]. Data on the presence of essential fatty acids in G. lucidum were also presented by Hossain et al. in 2007 [204]. Stojković et al. (2014) conducted comparative studies between the G. lucidum species from Serbia and China [205]. Lin et al. (2017) reported hydroxy fatty acids (HFA) present in the molecular species of acylglycerols of G. lucidum [206]. Salvatore et al. (2020) identified fatty acid methyl esters (FAMEs) in the triglycerides of G. lucidum spores [207]. Table 8 displays the composition similarities between the two species. [72].
Phytochemical reports over the past four decades indicate the presence of 279 bioactive secondary metabolites (Wasser et al.), along with over 200 polysaccharides and polysaccharide-protein complexes (Baby et al.), and more than 30 steroidal compounds (Wasser et al.) within the chemical composition of G. lucidum. These compounds play a pivotal role in the development of various biological actions. [120,208,209].

4. Biological Activities against Cancer

The biocomponents present in G. lucidum contribute to the numerous biological actions of this medicinal mushroom, making it applicable in medical contexts for combating cancer. Both polysaccharide compounds (Gl-Ps) and triterpene compounds (Gl-Ts) have been shown to exhibit anticancer activities. Over a 25-year period, numerous studies have investigated the anticancer effects of G. lucidum compounds, including those reported by Wang et al. (2002), Zhang et al. (2007), and Akihisa et al. (2007) [24,107,164]. Between 2009 and 2015, other researchers published studies on the anticancer effects of G. lucidum bio-compounds, such as Cho et al., Ferreira et al., Trajkovic et al., Kao et al., Zhao et al., and Joseph et al. [15,28,34,35,37,80,89]. Between 2018 and 2023, new studies disclosed data on anticancer actions, including those by Ahmad F. and Cör et al. in 2018, Kolniac-Ostek et al. in 2022, and Ekiz et al. in 2023 [13,153,192,210]. Our aim is to review the most important studies and systematize these data on the possible mechanisms.

4.1. Anticancer Action

The anticancer activity attributed to Gl-Ps was studied by Cao et al. in 2002 and 2003 [211,212]. Many other studies have demonstrated the effect of Gl-Ps on cytokines, including those by Chen et al. and Zhu et al. [213,214]. In 2007, Zhu et al. demonstrated the beneficial immunological effects in mice attributed to Gl-Ps extracts administered at low doses [215]. You et al. studied the effects of Gl-Ps on antioxidant enzymes in ovarian cancer in rats, while Xu et al. conducted research on mediating the immunomodulatory, cytotoxic, and anti-angiogenic effects of Gl-Ps [216,217]. Polysaccharides have been identified as a distinct class of compounds present in G. lucidum compositions, demonstrating anticancer activity by stimulating host immune function, as reported by Ahmad F., Sun et al., and Wiater et al. [213,218,219]. In 2012, Zhang et al. reported the effects of Gl-Ps on suppressing hepatocyte proliferation in rats [220]. Pan et al. reported the effects of Gl-Ps on rats with gastric cancer, inducing enhanced immunity and antioxidant activity [103]. The study by Suarez-Arroyo et al. in 2013 explored inflammatory breast cancer (IBC) and the effects of Gl-Ps in breast cancer [221]. In 2015, Habijanic et al. reported on the effects of Gl-Ps in modulating cytokine responses and lymphocyte activity [222]. Several studies have delved into the mechanisms of action of G. lucidum compounds in cancer treatment, such as those published by Ahmad F., Sohretoglu et al., and Wang et al. in 2018, as well as Fu et al. in 2019 [13,223,224,225]. The triterpene compounds (Gl-Ts) identified in G. lucidum exhibit significant anticancer effects. Both Min et al. in 2000 and Gao et al. in 2002 reported that triterpene compounds displayed cytotoxic effects against tumor cells [226,227]. In 2004, Lin et al. investigated the actions of Gl-Ps from aqueous extracts and Gl-Ts from alcoholic extracts of G. lucidum, evaluating their angiogenic effects [228]. Li et al. conducted research on ganoderic acid X, which acts by inhibiting topoisomerases and inducing apoptosis [229]. In Table 9, we have compiled the effects of essential compounds (Gl-Ps and Gl-Ts) in G. lucidum that significantly contribute to its anticarcinogenic activity.
Tang et al. (2006) reported the effects of ganoderic acid T in inhibiting lung cancer tumors (95-D), inducing apoptosis, and arresting the cell cycle at the G(1) phase [230]. Gl-Ts-type triterpenes and ganoderic acids A, F, and G were investigated in 2008 by Jiang et al. for their effects on breast cancer [231]. In 2009, Trajkovic et al. documented the action of triterpenes in B-16 melanoma, while Xu et al. studied ganoderic acid T and its effects on lung cancer [34,232]. In 2017, Barbieri et al. published data on the inhibition of IL-8, IL-6, MMP-2, and MMP-9 release in cancer cells under pro-inflammatory conditions in breast cancer and melanoma cells, and Ye et al. disclosed data on the effects of ganoderic and lucidenic acids in inhibiting the growth of lung carcinoma metastases and suppressing colon inflammation [233,234]. In 2020, the anticancer effects of triterpene fractions extracted from G. lucidum on SW620 human colorectal cancer cells were reported in vitro by Li et al. [235]. The action of Gl-Ts against various cancers has been studied by Jiao et al. (breast cancer cells), Wang et al. (prostate cancer), and Liu et al. (colorectal cancer) [236,237,238]. In 2022, Shahid et al. conducted research on the anticancer activities of Gl-Ts in skin cancer [239].

4.2. Possible Mechanisms in Anticancer Actions

The medical world has shown significant interest in establishing possible mechanisms of anticancer activity, especially as the literature provides ample evidence regarding the actions of G. lucidum biocomponents against various forms of cancer. In 2018, Ahmad F. systematized the possible mechanisms by which G. lucidum bio-compounds participate in cancer treatment [13]. The steps considered by Ahmad include the activation of the host cell immune response, induction of cell differentiation, inhibition of angiogenesis, direct cytotoxicity to tumor cells, inhibition of urokinase-type plasminogen activator and receptor expression in cancer cells, and inhibition of phase II metabolizing enzymes [13].
Kao et al. elucidated the mechanisms of the anticancer action of G. lucidum compounds by discussing the distinct effects of polysaccharide and triterpene compounds separately, as well as their combined effects (see Figure 9) [35].
They showed that polysaccharide compounds (Gl-Ps) act through three mechanisms [35]: by enhancing the host immune response, by stimulating macrophage activity, and by stimulating T-lymphocytes and natural killer (NK) cells. Triterpenes (Gl-Ts) act through five mechanisms:
-
G1-phase cell cycle arrest by inhibition of β-catenin;
-
Inhibition of protein kinase C (PCK), which generates G2-phase cell cycle inhibition;
-
Induction of apoptosis in cancer cells via the mitochondrial pathway, followed by activation of caspase cascades;
-
Preventing tumor metastasis by inhibiting MMP-9 and interleukin IL-8 and by degrading the extracellular matrix (ECM);
-
Suppressing the secretion of anti-inflammatory cytokines [240].
Together, the two major classes of bio-compounds, Gl-Ps and Gl-Ts, share two common mechanisms, namely:
I.
Antioxidant actions by reducing oxidative stress generated by free radicals and reactive oxygen species (ROS) through the actions of antioxidant enzymes [35].
II.
Suppressing angiogenesis and inhibiting nitric oxide production.

4.3. Other Biological Actions Involved in Anticancer Activities

The antitumor activity of G. lucidum bio-compounds is based on the differential effects of other biological activities generated by G. lucidum bio-compounds, as reported by Cör et al. [122,153]. Other researchers, such as Sharma et al., Ahmad, and You et al., have also conducted studies on the various factors that can lead to cancer [137,155,216]. Additionally, researchers like Xu et al., Fu et al., and Hapuarachchi et al. have analyzed various biological actions associated with cancer [217,225,241].

4.3.1. Immunomodulatory Activities

Anticancer activity based on immunomodulation activity has been reported by several researchers, such as Bao et al., Ooi et al., Moradali et al., and Lin et al. [94,106,109,168].

Effects of Gl-Ps on T- and B-Lymphocytes

Numerous studies have been reported that highlight the activating role of Gl-Ps (F3 fraction) on T-lymphocytes by increasing interleukin production: IL-1, IL-2, IL-6, and IL-12, and increasing the expression of IFN-γ and INF-α. Additionally, they enhanced DNA synthesis in mouse spleen cells, as reported by Chen et al. [213] and Wang et al. [24]. Gl-Ps from G. lucidum can activate PKC and PKA protein kinases in murine T-lymphocytes, according to Sohretoglu et al. [223]. The effect of Gl-Ps on B-lymphocytes is also crucial for tumor immunity. For example, Gl-Ps can activate B-lymphocytes by enhancing their proliferation and differentiation, leading to the production of substantial amounts of immunoglobulins in mice. Furthermore, Gl-Ps can directly stimulate the expression of PKCα and PKCγ in B-lymphocytes, as reported by Zhang et al. [242]. Activated B-lymphocytes increase the production of immunomodulatory substances, such as IL-1β and TNF-α, and reactive nitrogen species, such as NO [242].

Effect of Gl-Ps on Dendritic Cells

Dendritic cells (DCs) are professional antigen-presenting cells involved in the initiation of the primary T-lymphocyte immune response [223]. Stimulation of dendritic cell maturation by Gl-Ps from G. lucidum was documented by Sanodiya et al. [27]. Lai et al. reported that treatment with Gl-Ps F3 fraction extract improved the mixed lymphocyte response and stimulated the production of ten cytokines and six chemokines [243]. Jan et al. outlined the immunomodulatory activity of Gl-Ps, including the activation and maturation of DCs, as evidenced by increased cytokine production (IL-12, IL-6, IL-23, and IL-10) [244]. Chan et al. reported that treatment of monocytic leukemic cell lines with Gl-Ps resulted in increased leukocyte HLA-DR antigen expression and induced leukemic cell differentiation through increased superoxide production and cell cycle arrest [245].

Effect of Gl-Ps on Macrophages

Sohretoglu et al. demonstrated that Gl-Ps activate macrophages in vitro and elevate the levels of various cytokines, including IL-1β, tumor necrosis factor (TNF-α), IFN-γ, and IL-6 in the culture medium. This activation significantly increases macrophage phagocytosis and enhances macrophage-mediated tumor cytotoxicity [223]. Zhang et al. showed that the in vivo treatment of S180 sarcoma-bearing mice with Gl-Ps activated bone marrow-derived macrophages, inducing the production of immunomodulatory compounds such as IL-1β, TNF-α, and nitric oxide (NO) [242]. Hsu et al. elucidated the role of caspases in macrophage F3 fraction-induced Gl-Ps from G. lucidum spores [246]. Guo et al. discovered a novel water-soluble polysaccharide within G. lucidum spores that acts as an inducer of TNF-α and IL-6 secretion in murine peritoneal macrophages [247]. The in vivo administration of extracts from G. lucidum spores potentiated the proliferative response of splenocytes and induced antitumor activity against lung cancer in mice [247]. Hsu et al. reported that Gl-Ps induces increased secretion of the inflammatory cytokine IL-1 and stimulates the expression of pro-IL-1 and IL-1-converting enzymes in human and murine macrophages, an association linked with its anticancer activity [248].

Effect of Gl-Ps on Natural Killer (NK) Cells

Altfeld et al. defined the role of (NK) natural killer cells in innate immunity [249]. Chien et al. demonstrated that treatment with Gl-Ps resulted in increased monocyte, macrophage, and NK cell populations in human umbilical cord blood [250]. Wang et al. studied the effects of the bio-compounds from G. lucidum that, upon oral administration to mice, improved NK cell and phagocytosis activities and increased cytokine levels [251]. In additional research by Zhu et al., it was shown that Gl-Ps accelerated the recovery of bone marrow cells, red blood cells, and white blood cells, as well as splenic NK and NKT killer cells, and enhanced T- and B-lymphocyte proliferative responses [215]. The application of Gl-Ps treatments is recommended in cancer chemotherapy only at low doses [215].

4.3.2. Anti-Proliferative, Cytotoxic, and Apoptosis-Increasing Activities

Ganoderic bio-compounds have demonstrated various anti-proliferative and cytotoxic effects in studies regarding the treatment of different types of cancer [252,253,254,255]. The anti-proliferative effects of Gl-Ts compounds in G. lucidum manifest through cell cycle arrest [252,253]. Gl-Ts compounds can arrest the cell cycle in the G1 phase by inhibiting the β-catenin pathway, as reported by Wu et al., and in the G2/M phase by suppressing protein kinase C (PKC) activity, as reported by Lin et al. [252,253]. Jedinak et al. reported that ganodermanontriol inhibited the proliferation of HCT116 and HT-29 colon cancer cells by inhibiting β-catenin [254]. Li et al. (2005) identified ganoderic acid X as a compound that can arrest the cell cycle by inhibiting topoisomerase [229]. Chen et al., in 2010, reported that ganoderic acid T (GA-T) exhibits anti-proliferative effects against cancer cells in vitro and against metastasis in vivo [255]. In 2008, Chen et al. reported in the wound vacuolization assay that ganoderic acid Me (GA-Me), administered in a dose- and time-dependent manner, inhibited tumor invasion and cell adhesion to the extracellular matrix (ECM) [256]. GA-Me suppressed master metalloproteinases at the mRNA and protein levels in 95-D cells and is considered a potent antimetastatic carcinoma inhibitor [256]. Hsu et al. studied the anti-proliferative effects of lucidenic acids in human leukemic HL-60 cells [257].
Tang et al. conducted a study on the anti-proliferative effect of ganoderic acid T against cancer cells and observed enhanced cytotoxicity in lung cancer [230]. Triterpene compounds can induce apoptosis of cancer cells via the mitochondria-dependent pathway, followed by caspase activation, as reported by Kao et al. and Liu et al. in 2011 and 2012 [35,258,259]. Zhou et al. investigated the cytotoxic effects of GA-Me in human colon carcinoma, observing a dose-dependent pattern, and determined that the anticancer bioactivity of GA-Me was mediated through induced apoptosis [260].

4.3.3. Anti-Inflammatory Activities

Several inflammatory mediators, such as TNF-α, IL-6, TGF-β, and IL-10, have been shown to play roles in cancer initiation and progression [261,262]. G. lucidum biocomplexes have demonstrated dose-dependent anti-inflammatory effects [223]. Administration of the triterpene extract suppressed inflammatory cytokine secretion in macrophages with a significant reduction in inflammation in the affected tissue, as reported by Dudhgaonkar et al. in 2009 [263]. Joseph et al. found that Gl-Ps administration resulted in a 58% inhibition of inflammation, as assessed by carrageenan-induced (acute) and formalin-induced (chronic) inflammation assays [37].

4.3.4. Anti-Angiogenic Activities

Angiogenesis, the process by which new vasculature is formed from pre-existing vasculature, plays a key role in tumor growth and metastasis as well [264]. Nitric oxide is known to be an angiogenesis-inducing agent in tumors, promoting capillary formation within the tumor and allowing it to expand. In 2004, Cao et al. reported the existence of a peptide in G. lucidum that significantly reduced microvessel formation, as detected by the chorioallantois membrane assay [265]. Stanley et al. (2005) found that the G. lucidum extract prevented capillary morphogenesis by inhibiting the secretion of angiogenic factors VEGF and (TGF)-β1 [38]. Cao et al. (2006) demonstrated that the G. lucidum extract contains a polysaccharide peptide that exhibits anti-angiogenic activity by inhibiting nitric oxide production, subsequently suppressing cell multiplication in a dose-dependent manner [95].

4.3.5. Antioxidant Activities

Excess free radicals adversely affect bases in the nucleic acid structure, amino acids in the protein structure, and double bonds in unsaturated fatty acids, leading to oxidative stress, which is responsible for the alteration of DNA, RNA, proteins, and lipids. Hsieh et al. reported that bio-compounds from G. lucidum can exert chemopreventive effects through their antioxidant properties, such as free radical scavenging, as well as the ability to affect phase II detoxification enzymes [39]. Other authors, such as Smina et al., studied the antioxidant activity of Gl-Ts and demonstrated that they can reduce free radicals in cancer cells. [266]. In another study, Smina et al. reported the effect of total triterpenes from G. lucidum on the intracellular levels of reactive oxygen species (ROS) and the activities of endogenous antioxidant enzymes in spleen lymphocytes, highlighting their role in reducing radiation-induced oxidative DNA damage in spleen cells [267]. In 2001 and 2003, Lu et al. highlighted the beneficial contribution of the polysaccharides extracted from G. lucidum mycelium in reducing ROS-induced oxidative damage [268,269]. Lee et al. reported the existence of an amino polysaccharide compound in G. lucidum that can inactivate hydroxyl and superoxide anion radicals [270]. Other authors, such as XiaoPing et al. and Zhao et al., studied Gl-Ps and showed that these bio-compounds could be beneficial for glutathione peroxidase and reduce malonaldehyde levels in rats with cervical carcinoma and mice exposed to γ-irradiation [49,271].

5. Toxicity and Safety

Although there are numerous published studies on the beneficial effects of G. lucidum, there is also relatively little information reporting the toxic effects in humans. In this regard, Ahmad F. highlighted human sensitisation to Ganoderma antigen, reported in the USA, leading to allergic reactions [13,155]. When undergoing G. lucidum treatment, special attention should be paid to potential interactions with other drugs. Diabetic patients or those being treated with anticoagulants or antiplatelet drugs require special caution when being administered G. lucidum, as the anticipated effects may be altered [272].
In cancer therapy, although G. lucidum has been utilized as an anticancer agent, caution is still required when using it in conjunction with chemotherapy due to potential toxicity. Plasma concentrations of G. lucidum should be carefully monitored to detect elevated, toxic levels [273]. In an in vitro study, G. lucidum extracts were found to have toxic effects when exposed to cells at concentrations higher than those required for stimulatory results, resulting in a significant reduction in cell viability in a number of cell lines [274]. G. lucidum exhibits antihypertensive activity and may potentiate the effects of antihypertensive drugs [275]. Gl-Ps from G. lucidum have antibacterial activity and can enhance the activity of some antibiotics (e.g., tetracycline and cefazolin) [276].

6. Conclusions

G. lucidum bio-compounds are regarded as valuable in alternative cancer treatments based on non-natural products. The present work has compiled available data from various in vitro and in vivo studies on G. lucidum bio-compounds and their beneficial effects in anticancer treatment through their biological actions, such as anti-proliferative, antioxidant, immunomodulatory, anti-inflammatory, and anti-angiogenic effects. This paper presents information on the active bio-compounds in G. lucidum to obtain conclusive data and confirm their benefits regarding the mechanisms of anticancer action. Additionally, understanding the mechanisms of anticancer action, combined with other biological anticancer actions exerted by both the main bioactive compounds Gl-Ps and Gl-Ts and other compounds described in the paper (proteins, vitamins, metals, sterols, fatty acids, and nucleotide compounds), is necessary for targeted use in anticancer treatments. It is also crucial to comprehend that the biocomponents of G. lucidum are directly influenced by several factors, including the origin and culture medium, environmental conditions, temperature, humidity, and the quality of the environment from which the fungus originates. Therefore, further experimental, epidemiological, and clinical studies are needed to characterize the interactions of the administration of G. lucidum forms with different conventional anticancer drugs. More research is needed to combine G. lucidum bio-compound treatments with chemotherapy. Extensive pharmacological studies are also necessary to establish optimal dosages and assess the efficacy and safety of administration. Moreover, it is important to extend the research to identify metabolite subtypes that support the observed bioactivities, aiming to establish anticancer therapy procedures that promote general health and longevity.

Author Contributions

Conceptualization E.C., C.L.T., I.P., B.-S.N.-P. and C.P.; methodology, E.C., A.-M.I., C.L.T., B.-S.N.-P. and C.P.; writing—original draft preparation, E.C., R.S., C.P., I.P., B.-S.N.-P. and C.L.T.; writing—review and editing, R.S., C.P., I.P., B-S.N-P., A.-M.I. and E.C.; supervision T.N.-P., C.L.T., I.P. and R.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

Data is contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

Gl-PsG. lucidum polysaccharides ROSReactive Oxygen Species
Gl-TsG. lucidum triterpenoidsGA-AGanoderic acids A
TPCTotal polyphenol contentGA-FGanoderic acids F
TFCTotal flavonoid contentGA-HGanoderic acids H
ILInterleukin GA-Tganoderic acid T
TNF-αTumor Necrosis Factor AlphaGA-MeGanoderic acid Me
INF-γInterferon GammaMMP Matrix metalloproteinase;
TGF-αTransforming Growth Factor-AlfaTCLT-lymphocyte cytotoxic
TGF-βTransforming Growth Factor-BetaNKNatural killer cells;
VEGFVascular Endothelial Growth FactorPKCProtein kinase C
NONitrogen speciesECM Extracellular matrix;
MDAMalondialdehydeDCsDendritic cells
HPLC-ESI-MSLiquid chromatography coupled with electrospray ionization mass spectrometryTEAC Trolox equivalent antioxidant capacity

References

  1. Cai, Q.; Li, Y.; Pei, G. Polysaccharides from Ganoderma lucidum attenuates microglia-mediated neuroinflammation and modulate microglial phagocytosis and behavioural response. J. Neuroinflamm. 2017, 14, 63–76. [Google Scholar] [CrossRef]
  2. Li, H.J.; He, Y.L.; Zhang, D.H.; Yue, T.H.; Jiang, L.X.; Li, N.; Xu, J.W. Enhancement of ganoderic acid production by constitutively expressing Vitreoscilla hemoglobin gene in Ganoderma lucidum. J. Biotechnol. 2016, 227, 35–40. [Google Scholar] [CrossRef]
  3. Sliva, D. Cellular and physiological effects of Ganoderma lucidum (Reishi). Mini Rev. Med. Chem. 2004, 4, 873–879. [Google Scholar] [CrossRef] [PubMed]
  4. Soccol, C.R.; Bissoqui, L.Y.; Rodrigues, C.; Rubel, R.; Sella, S.R.; Leifa, F.; De Souza Vandenberghe, L.P.; Soccol, V.T. Pharmacological Properties of Biocompounds from Spores of the Lingzhi or Reishi Medicinal Mushroom Ganoderma lucidum (Agaricomycetes): A Review. Int. J. Med. Mushrooms 2016, 18, 757–767. [Google Scholar] [CrossRef] [PubMed]
  5. Zheng, W. Drug discovery from fungal metabolites: A review of the papers in this monographical issue of Mycosystema concerned with the natural resources, problems and strategies. Mycosystema 2011, 30, 151–157. [Google Scholar]
  6. Lindequist, U.; Niedermeyer, T.H.; Jülich, W.D. The pharmacological potential of mushrooms. Evid. Based Complement. Altern. Med. 2005, 2, 285–299. [Google Scholar] [CrossRef]
  7. Ahmad, M.F.; Ahmad, F.A.; Azad, Z.R.; Azaz, A.; Alam, M.I.; Ansari, J.A.; Panda, B.P. Edible mushrooms as health promoting agent. Adv. Sci. Focus 2013, 1, 189–196. [Google Scholar] [CrossRef]
  8. El Sheikha, A.F.; Hu, D.M. How to trace the geographic origin of mushrooms? Trends Food Sci. Technol. 2018, 78, 292–303. [Google Scholar] [CrossRef]
  9. Kou, F.; Ge, Y.; Wang, W.; Mei, Y.; Kao, L.; Wei, X.; Xiao, H.; Wu, X. A review of Ganoderma lucidum polysaccharides: Health benefit, structure-activity relationship, modification, and nanoparticle encapsulation. Int. J. Biol. Macromol. 2023, 243, 125199. [Google Scholar] [CrossRef]
  10. Lee, K.H.; Morris-Natschke, S.L.; Yang, X.; Huang, R.; Zhou, T.; Wu, S.F.; Shi, Q.; Itokawa, H. Recent progress of research on medicinal mushrooms, foods, and other herbal products used in traditional Chinese medicine. J. Tradit. Complement. Med. 2012, 2, 1–12. [Google Scholar] [CrossRef]
  11. Loyd, A.L.; Richter, B.S.; Jusino, M.A.; Truong, C.; Smith, M.E.; Blanchette, R.A.; Smith, J.A. Identifying the “mushroom of immortality”: Assessing the Ganoderma species composition in commercial Reishi products. Front. Microbiol. 2018, 9, 1557–1572. [Google Scholar] [CrossRef]
  12. Wasser, S.P.; Coates, P.; Blackman, M.; Cragg, G.; Levine, M.; Moss, J.; White, J. Reishi or Ling Zhi (Ganoderma lucidum). In Encyclopedia of Dietary Supplements, 1st ed.; Coates, P.M., Blackman, M.R., Cragg, G.M., Levine, M., White, J.D., Moss, J., Eds.; Marcel Dekker: New York, NY, USA, 2005; Volume 1, pp. 680–690. [Google Scholar] [CrossRef]
  13. Ahmad, M.F. Ganoderma lucidum: Persuasive biologically active constituents and their health endorsement. Biomed. Pharmacother. 2018, 107, 507–519. [Google Scholar] [CrossRef]
  14. Boh, B.; Berovic, M.; Zhang, J.; Zhi-Bin, L. Ganoderma lucidum and its pharmaceutically active compounds. Biotechnol. Ann. Rev. 2007, 13, 265–301. [Google Scholar] [CrossRef]
  15. Cho, J.H.; Noh, H.J.; Kang, D.H.; Lee, J.Y.; Lee, M.J.; Park, H.S.; Sung, G.H.; Jhune, C.S. Antioxidant activity and Cancer cell growth inhibition of Ganoderma lucidum. J. Mush. 2012, 10, 203–207. [Google Scholar]
  16. Ryu, D.H.; Cho, J.Y.; Sadiq, N.B.; Kim, J.C.; Lee, B.; Hamayun, M.; Lee, T.S.; Kim, H.S.; Park, S.P.; Nho, C.W.; et al. Optimization of antioxidant, anti-diabetic, and anti-inflammatory activities and ganoderic acid content of differentially dried Ganoderma lucidum using response surface methodology. Food Chem. 2021, 335, 127645–127654. [Google Scholar] [CrossRef] [PubMed]
  17. Cherian, E.; Sudheesh, N.P.; Janardhanan, K.K.; Patani, G. Free-radical scavenging and mitochondrial antioxidant activities of Reishi-Ganoderma lucidum (Curt: Fr) P. Karst and Arogyapacha—Trichopus zeylanicus Gaertn extracts. J. Basic Clin. Physiol. Pharmacol. 2009, 20, 289–307. [Google Scholar] [CrossRef] [PubMed]
  18. Bhat, Z.A.; Wani, A.H.; War, J.M.; Bhat, M.Y. Major bioactive properties of Ganoderma polysaccharides: A Review. Asian J. Pharm. Clin. Res. 2021, 14, 11–24. [Google Scholar] [CrossRef]
  19. Liu, X.; Yuan, J.P.; Chung, C.K.; Chen, X.J. Antitumor activity of the sporoderm-broken germinating spores of Ganoderma lucidum. Cancer Lett. 2002, 182, 155–161. [Google Scholar] [CrossRef]
  20. Kumar, V.; Yadav, H.K. Therapeutic potential of an edible macro-fungus: Ganoderma lucidum (Curtis) P. Karst. Indian J. Trad. Know. 2019, 18, 702–713. [Google Scholar]
  21. Rahman, M.A.; Al Masud, A.; Lira, N.Y.; Shakil, S. Proximate analysis, phtochemical screening and antioxidant activity of different strains of Ganoderma lucidum (Reishi Mushroom). Open J. Biol. Sci. 2020, 5, 24–27. [Google Scholar]
  22. Sudheer, S.; Alzorqi, I.; Manickam, S.; Ali, A. Bioactive Compounds of the Wonder Medicinal Mushroom “Ganoderma lucidum”. Bioactive Molecules in Food. In Reference Series in Phytochemistry; Mérillon, J.M., Ramawat, K.G., Eds.; Springer: Cham, Switzerland, 2019; pp. 1863–1893. [Google Scholar] [CrossRef]
  23. Bulam, S.; Ustun, N.S.; Peksen, A. Health benefits of Ganoderma lucidum as a medicinal mushroom. Turk. J. Agric. Food Sci. Technol. 2019, 7, 84–93. [Google Scholar] [CrossRef]
  24. Wang, Y.Y.; Khoo, K.H.; Chen, S.T.; Lin, C.C.; Wonga, C.H.; Lin, C.H. Studies on the immunomodulating and antitumor activities of Ganoderma lucidum (reishi) polysaccharides: Functional and proteomic analyses of a fucose-containing glycoprotein fraction responsible for the activities. Bioorg. Med. Chem. 2002, 10, 1057–1062. [Google Scholar] [CrossRef] [PubMed]
  25. Zhao, R.L.; He, Y.M. Network pharmacology analysis of the anti-cancer pharmacological mechanisms of Ganoderma lucidum extract with experimental support using Hepa1-6-bearing C57 BL/6 mice. J. Ethnopharmacol. 2018, 10, 287–295. [Google Scholar] [CrossRef] [PubMed]
  26. Oke, M.A.; Afolabi, F.J.; Oyeleke, O.O.; Kilani, T.A.; Adeosun, A.R.; Olanbiwoninu, A.A.; Adebayo, E.A. Ganoderma lucidum: Unutilized natural medicine and promising future solution to emerging diseases in Africa. Front. Pharmacol. 2022, 13, 952027. [Google Scholar] [CrossRef]
  27. Sanodiya, B.S.; Thakur, G.S.; Baghel, R.K.; Prasad, G.B.; Bisen, P.S. Ganoderma lucidum: A potent pharmacological macrofungus. Curr. Pharm. Biotechnol. 2009, 10, 717–742. [Google Scholar] [CrossRef] [PubMed]
  28. Ferreira, I.C.; Vaz, J.; Vasconcelos, M.H.; Martins, A. Compounds from wild mushrooms with antitumor potential. Anticancer. Agents Med. Chem. 2010, 10, 424–436. [Google Scholar] [CrossRef]
  29. Smania, E.F.A.; Monache, F.D.; Smania, A., Jr.; Yunes, R.A.; Cuneo, R.S. Antifungal activity of sterols and triterpenes isolated from Ganoderma annulare. Fitoterapia 2003, 74, 375–377. [Google Scholar] [CrossRef]
  30. El Sheikha, A.F. Nutritional Profile and Health Benefits of Ganoderma lucidum “Lingzhi, Reishi, or Mannentake” as Functional Foods: Current Scenario and Future Perspectives. Foods 2022, 11, 1030. [Google Scholar] [CrossRef]
  31. Du, M.; Hu, X.; Zhao, G.; Wang, C. Mechanism Underlying the Antitumor Activity of Proteins from the Ling Zhi or Reishi Medicinal Mushroom, Ganoderma lucidum (W. Curt.: Fr.) P. Karst. (Aphyllophoromycetideae) In Vitro. Int. J. Med. Mush. 2010, 12, 133–139. [Google Scholar] [CrossRef]
  32. Siwulski, M.; Sobieralski, K.; Golak-Siwulska, I.; Sokol, S.; Sękara, A. Ganoderma lucidum (Curt.: Fr.) Karst.—health-promoting properties. A review. Herba Pol. 2015, 61, 105–118. [Google Scholar] [CrossRef]
  33. Yuen, J.W.; Gohel, M.D. Anticancer effects of Ganoderma lucidum: A review of scientific evidence. Nutr. Cancer 2005, 53, 11–17. [Google Scholar] [CrossRef]
  34. Trajkovic, H.L.M.; Mijatovic, S.A.; Maksimovic-Ivanic, D.D.; Stojanovic, I.D.; Mamcilovic, M.B.; Tufegdzic, S.J.; Maksimovic, V.M.; Marjanovic, Z.S.; Stosic-Grujicic, S.D. Anticancer Properties of Ganoderma lucidum Methanol Extracts In Vitro and In Vivo. Nutr. Cancer 2009, 61, 696–707. [Google Scholar] [CrossRef]
  35. Kao, C.; Jesuthasan, A.C.; Bishop, K.S.; Glucina, M.P.; Ferguson, L.R. Anti-cancer activities of Ganoderma lucidum: Active ingredients and pathways. Funct. Foods Health Dis. 2013, 3, 48–65. [Google Scholar] [CrossRef]
  36. Kumar, D.S.; Senthilkumar, P.; Surendran, L.; Sudhagar, B. Ganoderma lucidum oriental mushroom mediated synthesis of gold nanoparticles conjugated with doxorubicin and evaluation of its anticancer potential on human breast cancer MCF-7/DOX cells. Int. J. Pharm. Pharm. Sci. 2017, 9, 267–274. [Google Scholar] [CrossRef]
  37. Joseph, S.; Sabulal, B.; George, V.; Antony, K.; Janardhanan, K. Antitumor and anti-inflammatory activities of polysaccharides isolated from Ganoderma lucidum. Acta Pharm. 2011, 61, 335–342. [Google Scholar] [CrossRef] [PubMed]
  38. Stanley, G.; Harvey, K.; Slivova, V.; Jiang, J. Ganoderma lucidum suppresses angiogenesis through the inhibition of secretion of VEGF and TGF-beta1 from prostate cancer cells. Biochem. Biophys. Res. Commun. 2005, 330, 46–52. [Google Scholar] [CrossRef] [PubMed]
  39. Hsieh, T.C.; Wu, J.M. Suppression of proliferation and oxidative stress by extracts of Ganoderma lucidum in the ovarian cancer cell line OVCAR-3. Int. J. Mol. Med. 2011, 28, 1065–1069. [Google Scholar] [CrossRef]
  40. Wu, X.; Guan, H.; Li, J.; Guo, J.; Hou, B. Evaluation of antitumour action of Ganoderma lucidum extract in hepatocarcinoma mice. Afr. J. Pharm. Pharmacol. 2012, 6, 2884–2887. [Google Scholar] [CrossRef]
  41. Li, A.; Shuai, X.; Jia, Z.; Li, H.; Liang, X.; Su, D.; Guo, W. Ganoderma lucidum polysaccharide extract inhibits hepatocellular carcinoma growth by down regulating regulatory T cells accumulation and function by inducing microRNA-125b. J. Transl. Med. 2015, 13, 100. [Google Scholar] [CrossRef]
  42. Liang, Z.; Guo, Y.T.; Yi, Y.J.; Wang, R.C.; Hu, Q.L.; Xiong, X.Y. Ganoderma lucidum Polysaccharides Target a Fas/Caspase Dependent Pathway to Induce Apoptosis in Human Colon Cancer Cells. Asian Pac. J. Cancer Prev. 2014, 15, 3981–3986. [Google Scholar] [CrossRef]
  43. Hsu, S.C.; Ou, C.C.; Chuang, T.C.; Li, J.W.; Lee, Y.J.; Wang, V. Ganoderma tsugae extract inhibits expression of epidermal growth factor receptor and angiogenesis in human epidermoid carcinoma cells: In vitro and in vivo. Cancer. Lett. 2009, 281, 108–116. [Google Scholar] [CrossRef]
  44. Wasser, S.P. Medicinal mushrooms as a source of antitumor and immunomodulating polysaccharides. Appl. Microbiol. Biotechnol. 2002, 60, 258–274. [Google Scholar] [CrossRef]
  45. Gao, Y.; Zhou, S.; Jiang, W.; Huang, M.; Dai, X. Effects of Ganopoly (a Ganoderma lucidum polysaccharide extract) on the immune functions in advanced-stage cancer patients. Immunol. Investig. 2003, 32, 201–215. [Google Scholar] [CrossRef]
  46. Sun, L.X.; Lin, Z.B.; Duan, X.S.; Qi, H.H.; Yang, N.; Li, M. Suppression of the production of transforming growth factor b1, interleukin-10, and vascular endothelial growth factor in the B16F10 cells by G. lucidum polysaccharides. J. Interferon Cytokine Res. 2014, 34, 667–675. [Google Scholar] [CrossRef]
  47. Shi, M.; Zhang, Z.; Yang, Y. Antioxidant and immunoregulatory activity of Ganoderma lucidum polysaccharide. Carbohydr. Polym. 2013, 95, 200–206. [Google Scholar] [CrossRef] [PubMed]
  48. Shaher, F.; Qiu, H.; Wang, S.; Hu, Y.; Wang, W.; Zhang, Y.; Wei, Y.; Al-ward, S.; Abdulghani, M.A.M.; Alenezi, S.K.; et al. Associated Targets of the Antioxidant Cardioprotection of Ganoderma lucidum in Diabetic cardiomyopathy by Using Open Targets Platform: A Systematic Review. BioMed Res. Int. 2020, 2020, 7136075. [Google Scholar] [CrossRef]
  49. XiaoPing, C.X.; Yan, C.; Shuibing, L.; You Gou, C. Free radical scavenging of Ganoderma lucidum polysaccharides and its effect on antioxidant enzymes and immunity activities in cervical carcinoma rats. Carbohydr. Polym. 2009, 77, 389–393. [Google Scholar] [CrossRef]
  50. Jones, S.; Janardhanan, K.K. Antioxidant and antitumor activity of Ganoderma lucidum (Curt.: Fr.) P. Karst. Reishi (Aphyllophoromycetideae) from South India. Int. J. Med. Mushr. 2000, 2, 195–200. [Google Scholar] [CrossRef]
  51. Fries, E.M. Systema Mycologicum, Sistens Fungorum Ordines, Genera et Species, huc Usque Cognitas, Quas ad Normam Methodi Naturalis Determinavit; Sumtibus Ernesti Mauritti: New York, NY, USA, 1821; Volume 1, pp. 1–274. [Google Scholar]
  52. Cao, Y.; Wu, S.H.; Dai, Y.C. Species clarification of the prize medicinal Ganoderma mushroom “Lingzhi”. Fungal Divers. 2012, 56, 49–62. [Google Scholar] [CrossRef]
  53. Wang, X.C.; Xi, R.J.; Li, Y.; Wang, D.M.; Yao, Y.J. The species identity of the widely cultivated Ganoderma, ’G. lucidum’ (Ling-zhi), in China. PLoS ONE 2012, 7, 40857. [Google Scholar] [CrossRef]
  54. Kwon, O.C.; Park, Y.J.; Kim, H.I.; Kong, W.S.; Cho, J.H.; Lee, C.S. Taxonomic Position and Species Identity of the Cultivated Yeongji ’Ganoderma lucidum’ in Korea. Mycobiology 2016, 44, 1–6. [Google Scholar] [CrossRef]
  55. Nahata, A. Ganoderma lucidum: A potent medicinal mushroom with numerous health benefits. Pharm. Anal. Acta 2013, 4, 159. [Google Scholar] [CrossRef]
  56. Kim, H.K.; Seo, G.S.; Kim, H.G. Comparison of characteristics of Ganoderma lucidum according to geographical origins: Consideration of morphological characteristics (II). Mycobiology 2001, 29, 80–84. [Google Scholar] [CrossRef]
  57. Park, Y.J.; Kwon, O.C.; Son, E.S.; Yoon, D.E.; Han, W.; Nam, J.Y.; Yoo, Y.B.; Lee, C.S. Genetic diversity analysis of Ganoderma species and development of a specific marker for identification of medicinal mushroom Ganoderma lucidum. Afr. J. Microbiol. Res. 2012, 6, 5417–5425. [Google Scholar] [CrossRef]
  58. Liao, B.; Chen, X.; Han, J.; Dan, Y.; Wang, L.; Jiao, W.; Song, J.; Chen, S. Identification of commercial Ganoderma (Lingzhi) species by ITS2 sequences. Chin. Med. 2015, 10, 22. [Google Scholar] [CrossRef] [PubMed]
  59. Gottlieb, A.M.; Ferrer, E.; Wright, J.E. rDNA analyses as an aid to the taxonomy of species of Ganoderma. Mycol. Res. 2000, 104, 1033–1045. [Google Scholar] [CrossRef]
  60. Eyssartier, G.; Roux, P. Le Guide des Champignons: France et Europe, 4th ed.; Pub. Belin/Humensis: Paris, France, 2017; pp. 1–1150. Available online: https://www.amazon.co.uk/champignons-France-Europe-Guillaume-Eyssartier/dp/2410010423 (accessed on 22 October 2023).
  61. Copot, O.; Tanase, C. Maxent modelling of the potential distribution of Ganoderma lucidum in north-eastern region of Romania. J. Plant Develop. 2017, 24, 133–143. [Google Scholar]
  62. Parepalli, Y.; Chavali, M.; Sami, R.; Singh, M.; Sinha, S.; Touahra, F. Ganoderma Lucidum: Extraction and Characterization of Polysaccharides, Yields and their Bioapplications. Algerian J. Res. Technol. 2021, 5, 30–43. [Google Scholar]
  63. Nithya, M.; Ambikapathy, V. Cultivation of Ganoderma lucidum (Curt. fr.) P. Karst. Int. J. Sci. Res. 2012, 3, 39–41. [Google Scholar]
  64. Adongbede, E.M.; Atoyebi, A.F. Isolation of Ganoderma lucidum (Curtis) P. Karst. From the Wild in Lagos through Tissue Culture Techniques and Cultivation on Sawdust of Six Nigerian Hardwoods. J. Appl. Sci. Environ. Manag. 2021, 25, 1497–1503. [Google Scholar] [CrossRef]
  65. Matute, R.G.; Figlas, D.; Devalis, R.; Delmastro, S.; Curvetto, N. Sunflower seed hulls as a main nutrient source for cultivating Ganoderma lucidum. Micol. Apl. Int. 2002, 14, 19–24. [Google Scholar]
  66. Yang, F.C.; Hsieh, C.; Chen, H.M. Use of stillage grain from a rice-spirit distillery in the solid state fermentation of Ganoderma lucidum. Process. Biochem. 2003, 39, 21–26. [Google Scholar] [CrossRef]
  67. Hsieh, C.; Yang, F.C. Reusing soy residue for the solid-state fermentation of Ganoderma lucidum. Bioresour. Technol. 2004, 91, 105–109. [Google Scholar] [CrossRef] [PubMed]
  68. Chang, M.Y.; Tsai, G.J.; Houng, J.Y. Optimization of the medium composition for the submerged culture of Ganoderma lucidum by Taguchi array design and steepest ascent method. Enz. Microb. Technol. 2006, 38, 407–414. [Google Scholar] [CrossRef]
  69. Baskar, G.; Sathya, S.R.; Jinnah, R.B.; Sahadevan, R. Statistical optimization of polysaccharide production by submerged cultivation of Lingzhi or Reishi medicinal mushroom, Ganoderma lucidum (W.Curt.: Fr.) P. Karst. MTCC 1039 (Aphyllophoromycetideae). Int. J. Med. Mush. 2011, 13, 41–49. [Google Scholar] [CrossRef] [PubMed]
  70. Mau, J.L.; Lin, H.C.; Chen, C.C. Non-volatile components of several medicinal mushrooms. Food Res. Int. 2001, 34, 521–526. [Google Scholar] [CrossRef]
  71. Roy, D.N.; Azad, A.; Sultana, M.F.; Anisuzzaman, A.S.M.; Khondkar, P. Nutritional profil e and mineral composition of two edible mushroom varieties consumed and cultivated in Bangladesh. J. Phytopharmacol. 2018, 1, 100–107. [Google Scholar] [CrossRef]
  72. Fraile-Fabero, R.; Ozcariz-Fermoselle, M.V.; Oria-de-Rueda-Salgueiro, J.A.; Garcia-Recio, V.; Cordoba-Diaz, D.; Jimen-Lopez, M.P.; Girbes-Juan, T. Differences in Antioxidants, Polyphenols, Protein Digestibility and Nutritional Profile between Ganoderma lingzhi from Industrial Crops in Asia and Ganoderma lucidum from Cultivation and Iberian Origin. Foods 2021, 10, 1750. [Google Scholar] [CrossRef]
  73. Parepalli, Y.; Chavali, M.; Sami, R.; Khojah, E.; Elhakem, A.; El Askary, A.; Singh, M.; Sinha, S.; El-Chaghaby, G. Evaluation of Some Active Nutrients, Biological Compounds and Health Benefits of Reishi Mushroom (Ganoderma lucidum). Int. J. Pharm. 2021, 17, 243–250. [Google Scholar] [CrossRef]
  74. Ogbe, A.O.; Obeka, A.D. Proximate, Mineral and Anti-Nutrient Composition of Wild Ganoderma lucidum: Implication on its Utilization in Poultry Production. Ir. J. Appl. Anim. Sci. 2013, 3, 161–166. [Google Scholar]
  75. Wachtel-Galor, S.; John, J.Y.; Buswell, J.A.; Benzie, I.F.F. Ganoderma lucidum (Lingzhi or Reishi). In Herbal Medicine: Biomolecular and Clinical Aspects, 2nd ed.; Benzie, I.F.F., Wachtel-Galor, S., Eds.; CRC Press/Taylor and Francis: Boca Raton, FL, USA, 2011; pp. 1–500. [Google Scholar] [CrossRef]
  76. Paterson, R.R.M. Ganoderma—a therapeutic fungal biofactory. Phytochemistry 2006, 67, 1985–2001. [Google Scholar] [CrossRef]
  77. Garuba, T.; Olahan, G.; Lateef, A.; Alaya, R.; Awolowo, M.; Sulyman, A. Proximate Composition and Chemical Profiles of Reishi Mushroom (Ganoderma lucidum (Curt: Fr.) Karst). J. Sci. Res. 2020, 12, 103–110. [Google Scholar] [CrossRef]
  78. Liu, Y.; Long, Y.; Liu, H.; Lan, Y.; Long, T.; Kuang, R.; Wang, Y.; Zhao, J. Polysaccharide prediction in Ganoderma lucidum fruiting body by hyperspectral imaging. Food Chem. X 2021, 13, 100–199. [Google Scholar] [CrossRef] [PubMed]
  79. Lin, Y.L.; Shih, C.; Cheng, P.Y.; Chin, C.L.; Liou, A.T.; Lee, P.Y.; Chiang, B.L. A Polysaccharide Purified from Ganoderma lucidum Acts as a Potent Mucosal Adjuvant That Promotes Protective Immunity Against the Lethal Challenge with Enterovirus A71. Front. Immunol. 2020, 29, 5617–5658. [Google Scholar] [CrossRef] [PubMed]
  80. Ferreira, I.C.; Heleno, S.A.; Reis, F.S.; Stojkovic, D.; Queiroz, M.J.; Vasconcelos, M.H. Chemical features of Ganoderma polysaccharides with antioxidant, antitumor and antimicrobial activities. Phytochemistry 2015, 114, 38–55. [Google Scholar] [CrossRef]
  81. Yu, H.; Liu, Y.; Zhou, S.; Yan, M.; Xue, L.; Tang, Q.; Zhang, J. Comparison of the polysaccharides form fruiting bodies, mycelia, and spore powder of Ganoderma lingzhi. Mycosystema 2016, 35, 170–177. [Google Scholar] [CrossRef]
  82. Giavasis, I. Bioactive fungal polysaccharides as potential functional ingredients in food and nutraceuticals. Curr. Opin. Biotechnol. 2014, 26, 162–173. [Google Scholar] [CrossRef]
  83. Chen, Y.; Xie, M.; Zhang, H.; Wang, Y.; Nie, S.; Li, C. Quantification of total polysaccharides and triterpenoids in Ganoderma lucidum and Ganoderma atrum by near infrared spectroscopy and chemometrics. Food Chem. 2012, 135, 268–275. [Google Scholar] [CrossRef]
  84. Nie, S.; Zhang, H.; Li, W.; Xie, M. Current development of polysaccharides from Ganoderma: Isolation, structure and bioactivities. Bioact. Carbohydr. Diet. Fibre. 2013, 1, 10–20. [Google Scholar] [CrossRef]
  85. Choong, Y.-K.; Ellan, K.; Chen, X.D.; Mohamad, S.A. Extraction and Fractionation of Polysaccharides from a Selected Mushroom Species, Ganoderma lucidum: A Critical Review. In Fractionation; Hassan-Al, H.I., Ed.; IntechOpen: London, UK, 2019; pp. 39–60. [Google Scholar] [CrossRef]
  86. Wang, J.G.; Ma, Z.C.; Zhang, L.N.; Fang, Y.P.; Jiang, F.; Phillips, G.O. Structure and chain conformation of water-soluble heteropolysaccharides from Ganoderma lucidum. Carbohydr. Polym. 2011, 86, 844–851. [Google Scholar] [CrossRef]
  87. Lin, Y.L.; Liang, Y.C.; Lee, S.S.; Chiang, B.L. Polysaccharide purified from Ganoderma lucidum induced activation and maturation of human monocyte-derived dendritic cells by the NF-κB and p38 mitogen activated protein kinase pathways. J. Leukoc. Biol. 2005, 78, 533–543. [Google Scholar] [CrossRef] [PubMed]
  88. Huang, S.Q.; Ning, Z.X. Extraction of polysaccharide from Ganoderma lucidum and its immune enhancement activity. Int. J. Biol. Macromol. 2010, 47, 336–341. [Google Scholar] [CrossRef] [PubMed]
  89. Zhao, L.; Dong, Y.; Chen, G.; Hu, Q. Extraction, purification, characterization, and antitumor activity of polysaccharides from Ganoderma lucidum. Carbohydr. Polym. 2010, 80, 783–789. [Google Scholar] [CrossRef]
  90. Leong, Y.K.; Yang, F.C.; Chang, J.S. Extraction of polysaccharides from edible mushrooms: Emerging technologies and recent advances. Carbohydr. Polym. 2021, 251, 117006. [Google Scholar] [CrossRef]
  91. Chen, Y.; Xie, M.Y.; Nie, S.P.; Li, C.; Wang, Y. Purification, composition analysis and antioxidant activity of a polysaccharide from the fruiting bodies of Ganoderma atrum. Food Chem. 2008, 107, 231–241. [Google Scholar] [CrossRef]
  92. Huang, S.Q.; Li, J.W.; Li, Y.Q.; Wang, Z. Purification, and structural characterization of a new water-soluble neutral polysaccharide GLP-F1-1 from Ganoderma lucidum. Int. J. Biol. Macromol. 2011, 48, 165–169. [Google Scholar] [CrossRef]
  93. Jiang, H.Y.; Sun, P.L.; He, J.Z.; Shao, P. Rapid purification of polysaccharides using novel radial flow ion-exchange by response surface methodology from Ganoderma lucidum. Food Bioprod. Process. 2012, 90, 1–8. [Google Scholar] [CrossRef]
  94. Bao, X.F.; Wang, X.S.; Dong, Q.; Fang, J.N.; Li, X.Y. Structural features of immunologically active polysaccharides from Ganoderma lucidum. Phytochemistry 2002, 59, 175–181. [Google Scholar] [CrossRef]
  95. Cao, Q.Z.; Lin, Z.B. Ganoderma lucidum polysaccharides peptide inhibits the growth of vascular endothelial cell and the induction of VEGF in human lung cancer cell. Life Sci. 2006, 78, 1457–1463. [Google Scholar] [CrossRef]
  96. Sullivan, R.; Smith, J.E.; Rowan, N.J. Medicinal mushrooms and cancer therapy. Translating a traditional practice into Western medicine. Perspect. Biol. Med. 2006, 49, 159–170. [Google Scholar] [CrossRef] [PubMed]
  97. Li, Y.Q.; Fang, L.; Zhang, K.C. Structure and bioactivities of agalactose rich extracellular polysaccharide from submergedly cultured Ganoderma lucidum. Carbohydr. Polym. 2007, 68, 323–328. [Google Scholar] [CrossRef]
  98. Ye, L.B.; Zhang, J.S.; Zhou, K.; Yang, Y.; Zhou, S.J.W. Purification, NMR study and immuno stimulating property of a fucogalactan from the fruiting bodies of Ganoderma lucidum. Planta Medica 2008, 74, 1730–1734. [Google Scholar] [CrossRef] [PubMed]
  99. Ye, L.B.; Zhang, J.S.; Yang, Y.; Zhou, S.; Liu, Y.F.; Tang, Q.J.; Du, X.; Chen, H.; Pan, Y. Structural characterisation of a heteropolysaccharide by NMR spectra. Food Chem. 2009, 112, 962–966. [Google Scholar] [CrossRef]
  100. Wang, J.; Zhang, L. Structure and chain conformation of five water-soluble derivatives of a beta-D-glucan isolated from Ganoderma lucidum. Carbohydr. Res. 2009, 344, 105–112. [Google Scholar] [CrossRef]
  101. Ye, L.; Li, J.R.; Zhang, J.S.; Pan, Y.J. NMR characterization for polysaccharide moiety of a glycopeptide. Fitoterapia 2010, 81, 93–96. [Google Scholar] [CrossRef] [PubMed]
  102. Liu, W.; Wang, H.Y.; Pang, X.B.; Yao, W.B.; Gao, X.D. Characterization, and antioxidant activity of two low- molecular-weight polysaccharides purified from the fruiting bodies of Ganoderma lucidum. Int. J. Biol. Macromol. 2010, 46, 451–457. [Google Scholar] [CrossRef]
  103. Pan, D.; Wang, L.; Chen, C.; Teng, B. Structure characterization of a novel neutral polysaccharide isolated from Ganoderma lucidum fruiting bodies. Food Chem. 2012, 135, 1097–1103. [Google Scholar] [CrossRef]
  104. Ma, C.W.; Feng, M.; Zhai, X.; Hu, M. Optimization for the extraction of polysaccharides from Ganoderma lucidum and their antioxidant and antiproliferative activities. J. Taiwan Inst. Chem. Eng. 2013, 44, 886–894. [Google Scholar] [CrossRef]
  105. Pan, K.; Jiang, Q.; Liu, G.; Miao, X.; Zhong, D. Optimization extraction of Ganoderma lucidum polysaccharides and its immunity and antioxidant activities. Intl. Biol. Macromol. 2013, 55, 301–306. [Google Scholar] [CrossRef]
  106. Ooi, V.E.; Liu, F. Immunomodulation, and anti-cancer activity of polysaccharide-protein complexes. Curr. Med. Chem. 2000, 7, 715–729. [Google Scholar] [CrossRef]
  107. Zhang, M.; Cui, S.W.; Cheung, P.C.; Wang, Q. Antitumor polysaccharides from mushrooms: A review on their isolation process, structural characteristics, and antitumor activity. Trends Food Sci. Technol. 2007, 18, 4–19. [Google Scholar] [CrossRef]
  108. Jia, J.; Zhang, X.; Hu, Y.S.; Wu, Y.; Wang, Q.Z.; Li, N.N. Evaluation of in vivo antioxidant activities of Ganoderma lucidum polysaccharides in STZ diabetic rats. Food Chem. 2009, 115, 32–36. [Google Scholar] [CrossRef]
  109. Moradali, M.F.; Mostafavi, H.; Ghods, S.; Hedjaroude, G.A. Immunomodulating and anticancer agents in the realm of macromycetes fungi (macrofungi). Int. Immunopharmacol. 2007, 7, 701–724. [Google Scholar] [CrossRef]
  110. Doco, T.; O’Neill, M.A.; Pellerin, P. Determination of the neutral and acidic glycosyl-residue compositions of plant polysaccharides by GC–EI–MS analysis of the trimethylsily l methyl glycoside derivatives. Carbohydr. Polym. 2001, 46, 249–259. [Google Scholar] [CrossRef]
  111. Hung, C.C.; Santschi, P.H.; Gillow, J.B. Isolation and characterization of extracellular polysaccharides produced by Pseudomonas fluorescens Biovar II. Carbohydr. Polym. 2005, 61, 141–147. [Google Scholar] [CrossRef]
  112. Dai, J.; Wu, Y.; Chen, S.W.; Zhu, S.; Yin, H.P.; Wang, M. Sugar compositional determination of polysaccharides from Dunaliella salina by modified RP-HPLC method of precolumn derivatization with 1-phenyl-3-methyl-5-pyrazolone. Carbohydr. Polym. 2010, 82, 629–635. [Google Scholar] [CrossRef]
  113. Yang, Q.; Wang, S.W.; Xie, Y.H.; Sun, J.Y.; Wang, J.B. HPLC analysis of Ganoderma lucidum polysaccharides and its effect on antioxidant enzymes activity and Bax, Bcl-2 expression. Int. J. Biol. Macromol. 2010, 46, 167–172. [Google Scholar] [CrossRef]
  114. Dong, Q.; Wang, Y.; Shi, L.; Yao, J.; Li, J.; Ma, F. A novel water-soluble β-D-glucan isolated from the spores of Ganoderma lucidum. Carbohydr. Res. 2012, 353, 100–105. [Google Scholar] [CrossRef]
  115. Liu, Y.J.; Shena, J.; Xia, Y.M.; Zhang, J. The polysaccharides from Ganoderma lucidum: Are they always inhibitors on human hepatocarcinoma cells? Carbohydr. Polym. 2012, 90, 1210–1215. [Google Scholar] [CrossRef] [PubMed]
  116. Skalicka-Wozniak, K.; Szypowski, J.; Łos, R.; Siwulski, M.; Sobieralski, K.; Glowniak, K. Evaluation of polysaccharides content in fruit bodies and their antimicrobial activity of four Ganoderma lucidum (W Curt.: Fr.) P. Karst strains cultivated on different wood type substrates. Acta Soc. Bot. Pol. 2012, 81, 17–21. [Google Scholar] [CrossRef]
  117. Pascale, C.; Sirbu, R.; Cadar, E. Importance of Bioactive Compounds of Ganoderma lucidum Extract in Medical Field. Eur. J. Nat. Sci. Med. 2022, 5, 39–47. [Google Scholar] [CrossRef]
  118. Wu, S.; Zhang, S.; Peng, B.; Tan, D.; Wua, M.; Jinchao Weia, J.; Wanga, Y.; Luoa, H. Ganoderma lucidum: A comprehensive review of phytochemistry, efficacy, safety and clinical study. Food Sci. Hum. Wellness 2023, 13, 568–596. [Google Scholar] [CrossRef]
  119. Gu, L.; Zheng, Y.; Lian, D. Production of triterpenoids from Ganoderma lucidum: Elicitation strategy and signal transduction. Process Biochem. 2018, 69, 22–32. [Google Scholar] [CrossRef]
  120. Baby, S.; Johnson, A.J.; Govindan, B. Secondary metabolites from Ganoderma. Phytochemistry 2015, 114, 66–101. [Google Scholar] [CrossRef]
  121. Galappaththi, M.C.A.; Patabendige, N.M.; Premarathne, B.M.; Hapuarachchi, K.K.; Tibpromma, S.; Dai, D.Q.; Suwannarach, N.; Rapior, S.; Karunarathna, S.C. A Review of Ganoderma Triterpenoids and Their Bioactivities. Biomolecules 2023, 13, 24. [Google Scholar] [CrossRef] [PubMed]
  122. Cör, A.D.; Knez, Ž.; Knez, M.M. Antioxidant, antibacterial, antitumor, antifungal, antiviral, anti-inflammatory, and nevroprotective activity of Ganoderma lucidum: An overview. Front. Pharmacol. 2022, 13, 934–982. [Google Scholar] [CrossRef]
  123. Lin, Y.X.; Sun, J.T.; Liao, Z.Z.; Sun, Y.; Tian, X.G.; Jin, L.L.; Wang, C.; Leng, A.J.; Zhou, J.; Li, D.W. Triterpenoids from the fruiting bodies of Ganoderma lucidum and their inhibitory activity against FAAH. Fitoterapia 2022, 158, 1051–1061. [Google Scholar] [CrossRef]
  124. Ghorai, N.; Chakraborty, S.; Shamik, G.; Saha, S.K.; Biswas, S. Estimation of total Terpenoids concentration in plant tissues using a monoterpene, Linalool as standard reagent. Protoc. Exch. 2012, 1, 6. [Google Scholar] [CrossRef]
  125. Taofiq, O.; Heleno, S.A.; Calhelha, R.C.; José Alves, M.; Barros, L.; González-Paramás, A.M.; Barreiro, M.F.; Ferreira, I.C.F.R. The potential of Ganoderma lucidum extracts as bioactive ingredients in topical formulations, beyond its nutritional benefits. Food Chem. Toxicol. 2017, 108, 139–147. [Google Scholar] [CrossRef]
  126. Chang, C.L.; Lin, C.S.; Lai, G.H. Phytochemical characteristics, free radical scavenging activities, and neuroprotection of five medicinal plant extracts. Evidence-based Complement. Altern. Med. 2012, 2012, 984295. [Google Scholar] [CrossRef]
  127. Huie, C.W.; Di, X. Chromatographic and electrophoretic methods for Lingzhi pharmacologically active components. J. Chromatogr. B. 2004, 812, 214–257. [Google Scholar] [CrossRef]
  128. Yang, M.; Wang, X.; Guan, S.; Xia, J.; Sun, J.; Guo, H.; Guo, D.A. Analysis of triterpenoids in Ganoderma lucidum using liquid chromatography coupled with electrospray ionization mass spectrometry. J. Am. Soc. Mass Spectr. 2007, 18, 927–939. [Google Scholar] [CrossRef]
  129. Zhang, C.R.; Yang, S.P.; Yue, J.M. Sterols and triterpenoids from the spores of Ganoderma lucidum. Nat. Prod. Res. 2008, 22, 1137–1142. [Google Scholar] [CrossRef]
  130. Shi, L.; Ren, A.; Mu, D.; Zhao, M. Current progress in the study on biosynthesis and regulation of ganoderic acids. Appl. Microbiol. Biotechnol. 2010, 88, 1243–1251. [Google Scholar] [CrossRef]
  131. Hadda, M.; Djamel, C.; Akila, O. Production and qualitative analysis of triterpenoids and steroids of Ganoderma species harvested from cork oak forest of North-Eastern Algeria. Res. J. Microbiol. 2015, 10, 366–376. [Google Scholar] [CrossRef]
  132. Che, X.Q.; Li, S.P.; Zhao, J. Ganoderma triterpenoids from aqueous extract of Ganoderma lucidum. Zhongguo Zhong Yao Za Zhi 2017, 42, 1908–1915. [Google Scholar]
  133. Hui, Z.; Jiang, H.; Yan Chen, Y.; Wang, J.; Yan, J. Quality evaluation of triterpenoids in Ganoderma and related species by the quantitative analysis of multi-components by single marker method. J. Liq. Chromatogr. Relat. Technol. 2018, 41, 860–867. [Google Scholar] [CrossRef]
  134. Zhang, C.; Fu, D.; Chen, G. Comparative and chemometric analysis of correlations between the chemical fingerprints and anti-proliferative activities of ganoderic acids from three Ganoderma species. Phytochem. Anal. 2019, 30, 474–480. [Google Scholar] [CrossRef]
  135. Xia, Q.; Zhang, H.; Sun, X.; Zhao, H.; Wu, L.; Zhu, D.; Yang, G.; Shao, Y.; Zhang, X.; Mao, X.; et al. A comprehensive review of the structure elucidation and biological activity of triterpenoids from Ganoderma spp. Molecules 2014, 19, 17478–17535. [Google Scholar] [CrossRef]
  136. Fatmawati, S.; Shimizu, K.; Kondo, R. Ganoderic acid Df, a new triterpenoid with aldose reductase inhibitory activity from the fruiting body of Ganoderma lucidum. Fitoterapia 2010, 81, 1033–1036. [Google Scholar] [CrossRef]
  137. Sharma, C.; Bhardwaj, N.; Sharma, A.; Tuli, H.S.; Katial, P.; Beniwal, V.; Gupta, G.K.; Sharma, A. Bioactive metabolites of Ganoderma lucidum: Factors, mechanism and broad-spectrum therapeutic potential. J. Herb. Med. 2019, 17, 100268. [Google Scholar] [CrossRef]
  138. Koo, M.H.; Chae, H.J.; Lee, J.H. Antiinflammatory lanostane triterpenoids from Ganoderma lucidum. Nat. Prod. Res. 2019, 35, 4295–4302. [Google Scholar] [CrossRef] [PubMed]
  139. Murata, C.T.Q.; Onda, S.; Usuki, T. Extraction and isolation of ganoderic acid Sigma from Ganoderma lucidum. Tetrahedron Lett. 2016, 57, 5368–5371. [Google Scholar] [CrossRef]
  140. Yue, Q.X.; Cao, Z.W.; Guan, S.H.; Liu, X.H.; Tao, L.; Wu, W.Y.; Li, Y.X.; Yang, P.Y.; Liu, X.; Guo, D.A. Proteomics characterization of the cytotoxicity mechanism of ganoderic acid D and computer-automated estimation of the possible drug target network. Mol. Cell. Proteom. 2008, 7, 949–961. [Google Scholar] [CrossRef]
  141. Yue, Q.X.; Song, X.Y.; Ma, C.; Feng, L.X.; Guan, S.H.; Wu, W.Y.; Yang, M.; Jiang, B.H.; Liu, X.; Cui, Y.J.; et al. Effects of triterpenes from Ganoderma lucidum on protein expression profile of HeLa cells. Phytomedicine 2010, 17, 606–613. [Google Scholar] [CrossRef]
  142. Zhang, X.Q.; Ip, F.C.; Zhang, D.M. Triterpenoids with neurotrophic activity from Ganoderma lucidum. Nat. Prod. Res. 2011, 25, 1607–1613. [Google Scholar] [CrossRef]
  143. Cheng, C.R.; Yue, Q.X.; Wu, Z.Y. Cytotoxic triterpenoids from Ganoderma lucidum. Phytochemistry 2010, 71, 1579–1585. [Google Scholar] [CrossRef]
  144. Liu, J.; Shimizu, K.; Tanaka, A. Target proteins of ganoderic acid DM provides clues to various pharmacological mechanisms. Sci. Rep. 2012, 2, 905. [Google Scholar] [CrossRef]
  145. Ríos, J.L.; Andújar, I.; Recio, M.C.; Giner, R.M. Lanostanoids from fungi: A group of potential anticancer compounds. J. Nat. Prod. 2012, 75, 2016–2044. [Google Scholar] [CrossRef]
  146. Li, Y.B.; Liu, R.M.; Zhong, J.J. A new ganoderic acid from Ganoderma lucidum mycelia and its stability. Fitoterapia 2013, 84, 115–122. [Google Scholar] [CrossRef] [PubMed]
  147. Fatmawati, S.; Kondo, R.; Shimizu, K. Structure-activity relationships of lanostane-type triterpenoids from Ganoderma lingzhi as α-glucosidase inhibitors. Bioorg. Med. Chem. Lett. 2013, 23, 5900–5903. [Google Scholar] [CrossRef]
  148. Li, P.; Deng, Y.P.; Wei, X.X.; Xu, J.H. Triterpenoids from Ganoderma lucidum and their cytotoxic activities. Nat. Prod. Res. 2013, 27, 17–22. [Google Scholar] [CrossRef] [PubMed]
  149. Chen, B.; Tian, J.; Zhang, J. Triterpenes and meroterpenes from Ganoderma lucidum with inhibitory activity against HMGs reductase, aldose reductase and α-glucosidase. Fitoterapia 2017, 120, 6–16. [Google Scholar] [CrossRef]
  150. Liang, C.; Tian, D.; Liu, Y. Review of the molecular mechanisms of Ganoderma lucidum triterpenoids: Ganoderic acids A, C2, D, F, DM, X and Y. Eur. J. Med. Chem. 2019, 174, 130–141. [Google Scholar] [CrossRef]
  151. Chinthanom, P.; Vichai, V.; Dokladda, K.; Sappan, M.; Thongpanchang, C.; Isaka, M. Semisynthetic modifications of antitubercular lanostane triterpenoids from Ganoderma. J. Antibiot. 2021, 74, 435–442. [Google Scholar] [CrossRef]
  152. Pascale, C.; Sirbu, R.; Cadar, E. Therapeutical Properties of Bioactive Compounds Extracted from Ganoderma lucidum Species on Acute and Chronic Diseases. Eur. J. Nat. Sci. Med. 2023, 6, 73–86. [Google Scholar] [CrossRef]
  153. Cör, A.D.; Knez, Z.; Knez-Hrncic, M. Antitumour, Antimicrobial, Antioxidant and Antiacetylcholinesterase Effect of Ganoderma lucidum Terpenoids and Polysaccharides: A Review. Molecules 2018, 23, 649. [Google Scholar] [CrossRef] [PubMed]
  154. Yang, Y.; Zhang, H.; Zuo, J.; Gong, X.; Yi, F.; Zhu, W.; Li, L. Advances in research on the active constituents and physiological effects of Ganoderma lucidum. Biomed. Dermatol. 2019, 3, 1–17. [Google Scholar] [CrossRef]
  155. Ahmad, M.F. Ganoderma lucidum: A rational pharmacological approach to surmount cancer. J. Ethnopharmacol. 2020, 60, 113047. [Google Scholar] [CrossRef] [PubMed]
  156. Hussien, Z.G. Chemical content study in Ganoderma lucidum commercial products. Minar. Int. J. App. Sci. Technol. 2022, 4, 158–165. [Google Scholar] [CrossRef]
  157. Sharif, S.; Mustafa, G.; Munir, H. Proximate composition and micronutrient mineral profile of wild Ganoderma lucidum and four commercial exotic mushrooms by ICP-OES and LIBS. J. Food Nutr. Res. 2016, 4, 703–708. [Google Scholar]
  158. Treviño, S.; Díaz, A.; Sanchez-Lara, E.; Sanchez-Gaytan, B.L.; Perez-Aguilar, J.M.; Gonzalez-Vergara, E. Vanadium in Biological Action: Chemical, Pharmacological Aspects, and Metabolic Implications in Diabetes Mellitus. Biol. Trace Elem. Res. 2019, 188, 68–98. [Google Scholar] [CrossRef] [PubMed]
  159. Matute, R.G.; Serra, A.; Figlas, D.; Curvetto, N. Copper and Zinc Bioaccumulation and Bioavailability of Ganoderma lucidum. J. Med. Food 2011, 14, 1273–1279. [Google Scholar] [CrossRef] [PubMed]
  160. Falandysz, J. Selenium in edible mushrooms. J. Environ. Sci. Health Part C Environ. Carcinog. Ecotoxicol. Rev. 2008, 26, 256–299. [Google Scholar] [CrossRef] [PubMed]
  161. Chiu, S.W.; Wang, Z.M.; Leung, T.M.; Moore, D. Nutritional value of ganoderma extract and assessment of its genotoxicity and antigenotoxicity using comet assays of mouse lymphocytes. Food Chem. Toxicol. 2000, 38, 173–178. [Google Scholar] [CrossRef]
  162. Du, M.; Cong, W.; Hu, X.S.; Zhao, G. Positive Effect of Selenium on the Immune Regulation Activity of Ling Zhi or Reishi Medicinal Mushroom, Ganoderma lucidum (W. Curt.: Fr.) P. Karst. (Aphyllophoromycetideae), Proteins In Vitro. Int. J. Med. Mushrooms 2008, 10, 337–344. [Google Scholar] [CrossRef]
  163. Ma, J.; Ye, Q.; Hua, Y.; Zhang, D.; Cooper, R.; Chang, M.N.; Chang, J.Y.; Sun, H.H. New Lanostanoids from the Mushroom Ganoderma lucidum. Nat. Prod. 2002, 65, 72–75. [Google Scholar] [CrossRef]
  164. Akihisa, T.; Nakamura, Y.; Tagata, M.; Tokuda, H.; Yasukawa, K.; Uchiyama, E.; Suzuki, T.; Kimura, Y. Anti-Inflammatory and Anti-Tumor-Promoting Effects of Triterpene Acids and Sterols from the Fungus Ganoderma lucidum. Chem. Biodivers. 2007, 4, 224–231. [Google Scholar] [CrossRef]
  165. Liu, J.; Huang, W.; Lv, M.; Si, J.; Guo, B.; Li, S. Determination of ergosterol in Ganoderma lucidum from different varieties and cultured tree species by HPLC. J. Chin. Med. Mater. 2011, 34, 187–190. [Google Scholar]
  166. Huang, L.; Sun, F.; Liang, C.; He, Y.X.; Bao, R.; Liu, L.; Zhou, C.Z. Crystal structure of LZ-8 from the medicinal fungus Ganoderma lucidium. Proteins 2009, 75, 524–527. [Google Scholar] [CrossRef]
  167. Hsu, H.Y.; Kuan, Y.C.; Lin, T.Y.; Tsao, S.M.; Hsu, J.; Ma, L.J.; Sheu, F. Reishi Protein LZ-8 Induces FOXP3(+) Treg Expansion via a CD45-Dependent Signaling Pathway and Alleviates Acute Intestinal Inflammation in Mice. Evid. Based Comp. Alternat. Med. 2013, 2013, 513542. [Google Scholar] [CrossRef]
  168. Lin, H.J.; Chang, Y.S.; Lin, L.H.; Haung, C.F.; Wu, C.Y.; Ou, K.L. An Immunomodulatory Protein (Ling Zhi-8) from a Ganoderma lucidum Induced Acceleration of Wound Healing in Rat Liver Tissues after Monopolar Electrosurgery. Evid. Based Complement. Altern. Med. 2014, 2014, 916531. [Google Scholar] [CrossRef]
  169. Yang, Y.; Yu, T.; Tang, H.; Ren, Z.; Li, Q.; Jia, J.; Chen, H.; Fu, J.; Ding, S.; Hao, Q.; et al. Ganoderma lucidum Immune Modulator Protein rLZ-8 Could Prevent and Reverse Bone Loss in Glucocorticoids-Induced Osteoporosis Rat Model. Front. Pharmacol. 2020, 11, 731. [Google Scholar] [CrossRef]
  170. Sa-Ard, P.; Sarnthima, R.; Khammuang, S. Antioxidant, antibacterial and DNA protective activities of protein extracts from Ganoderma lucidum. J. Food Sci. Technol. 2015, 52, 2966–2973. [Google Scholar] [CrossRef]
  171. Sun, J.; He, H.; Xie, B.J. Novel Antioxidant Peptides from Fermented Mushroom Ganoderma lucidum. J. Agric. Food Chem. 2004, 52, 6646–6652. [Google Scholar] [CrossRef]
  172. Ji, Z.; Tang, Q.; Zhang, J.; Yang, Y.; Jia, W.; Pan, Y. Immunomodulation of RAW264.7 macrophages by GLIS, a proteopolysaccharide from Ganoderma lucidum. J. Ethnopharmacol. 2007, 112, 445–450. [Google Scholar] [CrossRef]
  173. Yeh, C.H.; Chen, H.C.; Yang, J.J.; Chuang, W.I.; Sheu, F. Polysaccharides PS-G and Protein LZ-8 from Reishi (Ganoderma lucidum) Exhibit Diverse Functions in Regulating Murine Macrophages and T Lymphocytes. J. Agric. Food Chem. 2010, 58, 8535–8544. [Google Scholar] [CrossRef]
  174. Girjal, V.U.; Neelagund, S.; Krishnappa, M. Antioxidant Properties of the Peptides Isolated from Ganoderma lucidum Fruiting Body. Int. J. Pept. Res. Ther. 2012, 18, 319–325. [Google Scholar] [CrossRef]
  175. Zhong, D.; Wang, H.; Liu, M.; Li, X.; Huang, M.; Zhou, H.; Lin, S.; Lin, Z.; Yang, B. Ganoderma lucidum polysaccharide peptide prevents renal ischemia reperfusion injury via counteracting oxidative stress. Sci. Rep. 2015, 5, 16910. [Google Scholar] [CrossRef]
  176. Kumakura, K.; Hori, C.; Matsuoka, H.; Kiyohiko, I.K.; Masahiro, S.M. Protein components of water extracts from fruiting bodies of the reishi mushroom Ganoderma lucidum contribute to the production of functional molecules. J. Sci. Food. Agric. 2019, 99, 529–535. [Google Scholar] [CrossRef]
  177. Yu, F.; Wang, Y.; Teng, Y.; Yang, S.; He, Y.; Zhang, Z.; Yang, H.; Ding, C.F.; Zhou, P. Interaction and Inhibition of a Ganoderma lucidum Proteoglycan on PTP1B Activity for Anti-diabetes. ACS Omega 2021, 6, 29804–29813. [Google Scholar] [CrossRef]
  178. Yu, F.; Teng, Y.; Jiaqi, L.J.; Yang, S.; Zeng, Z.; Yanming, H.; Hongjie, Y.; Chuan-Fan, D.; Ping, Z. Effects of a Ganoderma lucidum Proteoglycan on Type 2 Diabetic Rats and the Recovery of Rat Pancreatic Islets. ACS Omega 2023, 8, 17304–17316. [Google Scholar] [CrossRef]
  179. Huang, P.; Luo, F.J.; Ma, Y.C.; Wang, S.X.; Huang, J.; Qin, D.D.; Xue, F.F.; Liu, B.Y.; Wu, Q.; Wang, X.L.; et al. Dual antioxidant activity and the related mechanisms of a novel pentapeptide GLP4 from the fermented mycelia of Ganoderma lingzhi. Food Funct. 2022, 13, 9032–9048. [Google Scholar] [CrossRef]
  180. Thakur, A.; Rana, M.; Lakhanpal, T.N.; Ahmad, A.; Khan, M.I. Purification and characterization of lectin from fruiting body of Ganoderma lucidum: Lectin from Ganoderma lucidum. Biochim. Et Biophys. Acta Gen. Subj. 2007, 1770, 1404–1412. [Google Scholar] [CrossRef]
  181. Girjal, V.U.; Neelagund, S.; Krishnappa, M. Ganoderma lucidum: A source for novel bioactive lectin. Prot. Pept. Lett. 2011, 18, 1150–1157. [Google Scholar] [CrossRef]
  182. Nikitina, V.E.; Loshchinina, E.A.; Vetchinkina, E.P. Lectins from Mycelia of Basidiomycetes. Int. J. Mol. Sci. 2017, 18, 1334. [Google Scholar] [CrossRef]
  183. Yousra, A.; El-Maradny, E.M.; El-Fakharany, M.M.S.; Mona, H.H.; Heba, S.S. Lectins purified from medicinal and edible mushrooms: Insights into their antiviral activity against pathogenic viruses. Int. J. Biol. Macromol. 2021, 179, 239–258. [Google Scholar] [CrossRef]
  184. Deepalakshmi, K.; Mirunalini, S. Therapeutic properties and current medical usage of medicinal mushroom: Ganoderma lucidum. Int. J. Pharm. Sci. Res. 2011, 2, 1922–1929. [Google Scholar] [CrossRef]
  185. Zhang, H.; Jiang, H.; Zhang, X.; Yan, J. Amino Acids from Ganoderma lucidum: Extraction Optimization, Composition Analysis, Hypoglycemic and Antioxidant Activities. Curr. Pharm. Anal. 2018, 14, 562–570. [Google Scholar] [CrossRef]
  186. Kim, M.Y.; Seguin, P.; Ahn, J.K.; Kim, J.J.; Chun, S.C.; Kim, E.H.; Seo, S.H.; Kang, E.Y.; Kim, S.L.; Park, Y.J.; et al. Phenolic Compound Concentration and Antioxidant Activities of Edible and Medicinal Mushrooms from Korea. J. Agric. Food Chem. 2008, 56, 7265–7270. [Google Scholar] [CrossRef]
  187. Sheikh, I.A.; Vyas, D.; Ganaie, M.A.; Dehariya, K.; Singh, V. HPLC determination of phenolics and free radical scavenging activity of ethanolic extracts of two polypore mushrooms. Int. J. Pharm. Pharm. Sci. 2014, 6, 679–684. [Google Scholar]
  188. Lin, M.S.; Yu, Z.R.; Wang, B.J.; Wang, C.C.; Weng, Y.M.; Koo, M. Bioactive constituent characterization and antioxidant activity of Ganoderma lucidum extract fractionated by supercritical carbon dioxide. Sains Mal. 2015, 44, 1685–1691. [Google Scholar]
  189. Veljović, S.; Veljović, M.; Nikićević, N.; Despotović, S.; Radulović, S.; Nikšić, M.; Filipović, L. Chemical composition, antiproliferative and antioxidant activity of differently processed Ganoderma lucidum ethanol extracts. J. Food Sci. Technol. 2017, 54, 1312–1320. [Google Scholar] [CrossRef]
  190. Dong, Q.; He, D.; Ni, X. Comparative study on phenolic compounds, triterpenoids, and antioxidant activity of Ganoderma lucidum affected by different drying methods. Food Meas. 2019, 13, 3198–3205. [Google Scholar] [CrossRef]
  191. Zheng, S.; Zhang, W.; Liu, S. Optimization of ultrasonic-assisted extraction of polysaccharides and triterpenoids from the medicinal mushroom Ganoderma lucidum and evaluation of their in vitro antioxidant capacities. PLoS ONE 2020, 15, e0244749. [Google Scholar] [CrossRef]
  192. Kolniak-Ostek, J.; Oszmianski, J.; Szyjka, A.; Moreira, H.; Barg, E. Anticancer and Antioxidant Activities in Ganoderma lucidum Wild Mushrooms in Poland, as Well as Their Phenolic and Triterpenoid Compounds. Int. J. Mol. Sci. 2022, 23, 9359. [Google Scholar] [CrossRef]
  193. Saltarelli, R.; Ceccaroli, P.; Buffalini, M.; Vallorani, L.; Casadei, L.; Zambonelli, A.; Iotti, M.; Badalyan, S.; Stocchi, V. Biochemical characterization and antioxidant and antiproliferative activities of different Ganoderma collections. J. Mol. Microb. Biotechnol. 2015, 25, 16–25. [Google Scholar] [CrossRef]
  194. Tang, X.; Cai, W.; Xu, B. Comparison of the Chemical Profiles and Antioxidant and Antidiabetic Activities of Extracts from Two Ganoderma Species (Agaricomycetes). Int. J. Med. Mush. 2016, 18, 609–620. [Google Scholar] [CrossRef]
  195. Sánchez, C. Reactive oxygen species and antioxidant properties from mushrooms. Synth. Syst. Biotechnol. 2017, 2, 13–22. [Google Scholar] [CrossRef]
  196. Mohammadifar, S.; Fallahi, G.S.; Asef, S.M.R.; Vaziri, A. Comparison between antioxidant activity and bioactive compounds of Ganoderma applanatum (Pers.) Pat. and Ganoderma lucidum (Curt.) P. Karst from Iran. Iran. J. Plant Physiol. 2020, 11, 3417–3424. [Google Scholar]
  197. Mustafin, K.; Bisko, N.; Blieva, R.; Al-Maali, G.; Krupodorova, T.; Narmuratova, Z.; Saduyeva, Z.; Zhakipbekova, A. Antioxidant and antimicrobial potential of Ganoderma lucidum and Trametes versicolor. Turk. J. Biochem. 2022, 47, 483–489. [Google Scholar] [CrossRef]
  198. Cheung, H.Y.; Ng, C.W.; Hood, D.J. Identification and quantification of base and nucleoside markers in extracts of Ganoderma lucidum, Ganoderma japonicum and Ganoderma capsules by micellar electrokinetic chromatography. J. Chromatogr. 2001, 911, 119–126. [Google Scholar] [CrossRef] [PubMed]
  199. Gao, J.L.; Leung, K.S.; Wang, Y.T.; Lai, C.M.; Li, S.P.; Hu, L.F. Qualitative and quantitative analyses of nucleosides and nucleobases in Ganoderma spp. by HPLC-DAD-MS. J. Pharm. Biomed. Anal. 2007, 44, 801–811. [Google Scholar] [CrossRef] [PubMed]
  200. Yuan, J.P.; Zhao, S.Y.; Wang, J.H.; Kuang, H.C.; Liu, X. Distribution of Nucleosides and Nucleobases in Edible Fungi. J. Agric. Food Chem. 2008, 56, 809–815. [Google Scholar] [CrossRef] [PubMed]
  201. Chen, Y.; Bicker, W.; Wu, J.I.; Xie, M.; Lindner, W. Simultaneous Determination of 16 Nucleosides and Nucleobases by Hydrophilic Interaction Chromatography and Its Application to the Quality Evaluation of Ganoderma. J. Agric. Food Chem. 2012, 60, 4243–4252. [Google Scholar] [CrossRef] [PubMed]
  202. Phan, C.W.; Wang, J.K.; Cheah, S.C.; Naidu, M.; David, P.; Sabaratnam, V. A review of the nucleid acid constituents in mushrooms: Nucleobases, nucleosides, and nucleotides. Crit. Rev. Biotechnol. 2018, 38, 762–777. [Google Scholar] [CrossRef]
  203. Sheng, F.; Wang, A.; Luo, X.; Xiao, J.; Hu, L.; Li, L. Simultaneous determination of ten nucleosides and bases in Ganoderma by micellar electrokinetic chromatography. Food Sci. Hum. Wellness 2022, 11, 263–268. [Google Scholar] [CrossRef]
  204. Hossain, S.A.; Nuhu, A.; Siddharth, B.M.A.; Rahman, M. Essential fatty acid contents of Pleurotus ostreatus, Ganoderma lucidum and Agaricus bisporus. Bangladesh J. Mush. 2007, 2, 1–7. [Google Scholar]
  205. Stojković, D.S.; Barros, L.; Calhelha, R.C.; Glamočlija, J.; Ćirić, A.; van Griensven, L.J.; Soković, M.; Ferreira, I.C. A detailed comparative study between chemical and bioactive properties of Ganoderma lucidum from different origins. Int. J. Food Sci. Nutr. 2014, 65, 42–47. [Google Scholar] [CrossRef]
  206. Lin, J.T.; Hou, C.; Dulay, R.; Ray, K.; Chen, G. Structures of hydroxy fatty acids as the constituents of triacylglycerols in Philippine wild edible mushroom, Ganoderma lucidum. Biocatal. Agric. Biotechnol. 2017, 12, 148–151. [Google Scholar] [CrossRef]
  207. Salvatore, M.M.; Elvetico, A.; Gallo, M.; Salvatore, F.; DellaGreca, M.; Naviglio, D.; Andolfi, A. Fatty Acids from Ganoderma lucidum Spores: Extraction, Identification and Quantification. Appl. Sci. 2020, 10, 3907. [Google Scholar] [CrossRef]
  208. Wasser, S.P. Medicinal mushroom science: History, current status, future trends, and unsolved problems. Int. J. Med. Mush. 2010, 12, 1–16. [Google Scholar] [CrossRef]
  209. Wasser, S.P. Current findings, future trends, and unsolved problems in studies of medicinal mushrooms. Appl. Microbiol. Biotechnol. 2011, 89, 1323–1332. [Google Scholar] [CrossRef]
  210. Ekiz, E.; Oz, E.; Abd El-Aty, A.M.; Proestos, C.; Brennan, C.; Zeng, M.; Tomasevic, I.; Elobeid, T.; Çadırcı, K.; Bayrak, M.; et al. Exploring the Potential Medicinal Benefits of Ganoderma lucidum: From Metabolic Disorders to Coronavirus Infections. Foods 2023, 12, 1512. [Google Scholar] [CrossRef]
  211. Cao, L.Z.; Lin, Z.B. Regulation on maturation and function of dendritic cells by Ganoderma lucidum polysaccharides. Immunol. Lett. 2002, 83, 163–169. [Google Scholar] [CrossRef]
  212. Cao, L.Z.; Lin, Z.B. Regulatory effect of Ganoderma lucidum polysaccharides on cytotoxic T-lymphocytes induced by dendritic cells in vitro. Acta. Pharmacol. Sin. 2003, 24, 321–326. [Google Scholar] [PubMed]
  213. Chen, H.S.; Tsai, Y.F.; Lin, S.; Lin, C.C.; Khoo, K.H.; Lin, C.H. Studies on the immuno-modulating and anti-tumor activities of Ganoderma lucidum (Reishi) polysaccharides. Bioorg. Med. Chem. 2004, 12, 5595–5601. [Google Scholar] [CrossRef] [PubMed]
  214. Zhu, X.L.; Lin, Z.B. Effects of Ganoderma lucidum polysaccharides on proliferation and cytotoxicity of cytokine-induced killer cells. Acta Pharmacol. Sin. 2005, 26, 1130–1137. [Google Scholar] [CrossRef]
  215. Zhu, X.L.; Chen, A.F.; Lin, Z.B. Ganoderma lucidum polysaccharides enhance the function of immunological effector cells in immunosuppressed mice. J. Ethnopharmacol. 2007, 111, 219–226. [Google Scholar] [CrossRef]
  216. You, G.C.; Zongji, S.; XiaoPing, C. Modulatory effect of Ganoderma lucidum polysaccharides on serum antioxidant enzymes activities in ovarian cancer rats. Carbohyd. Polym. 2009, 78, 258–262. [Google Scholar] [CrossRef]
  217. Xu, Z.; Chen, X.; Zhong, Z.; Chen, L.; Wang, Y. Ganoderma lucidum polysaccharides: Immunomodulation and potential anti-tumor activities. Am. J. Chin. Med. 2011, 39, 15–27. [Google Scholar] [CrossRef] [PubMed]
  218. Sun, L.X.; Lin, Z.B.; Li, X.J.; Li, M.; Lu, J.; Duan, X.S. Promoting effects of Ganoderma lucidum polysaccharides on B16F10 cells to activate lymphocytes. Basic Clin. Pharmacol. Toxicol. 2011, 108, 149–154. [Google Scholar] [CrossRef]
  219. Wiater, A.; Paduch, R.; Choma, A.; Pleszczyńska, M.; Siwulski, M.; Dominik, J. Biological study on carboxymethylated (1→3)-α-D-glucans from fruiting bodies of Ganoderma lucidum. Int. J. Biol. Macromol. 2012, 51, 1014–1023. [Google Scholar] [CrossRef]
  220. Zhang, J.; Liu, Y.J.; Park, H.S.; Xia, Y.M. Antitumor activity of sulfated extracellular polysaccharides of Ganoderma lucidum from the submerged fermentation broth. Carbohydr. Polym. 2012, 87, 1539–1544. [Google Scholar] [CrossRef]
  221. Suarez-Arroyo, I.J.; Rosario-Acevedo, R.; Aguilar-Perez, A.; Clemente, P.L.; Cubano, L.A.; Serrano, J. Anti-tumor effects of Ganoderma lucidum (reishi) in inflammatory breast cancer in in vivo and in vitro models. PLoS ONE 2013, 8, e57431. [Google Scholar] [CrossRef]
  222. Habijanic, J.; Berovic, M.; Boh, B.; Plankl, M.; Wraber, B. Submerged cultivation of Ganoderma lucidum and the effects of its polysaccharides on the production of human cytokines TNF-α, IL-12, IFN-γ, IL-2, IL-4, IL-10 and IL-17. N. Biotechnol. 2015, 32, 85–95. [Google Scholar] [CrossRef]
  223. Sohretoglu, D.; Huang, S. Ganoderma lucidum Polysaccharides as An Anti-cancer Agent. Anticancer Agents Med. Chem. 2018, 18, 667–674. [Google Scholar] [CrossRef]
  224. Wang, C.; Shi, S.; Chen, Q.; Lin, S.; Wang, R.; Wang, S.; Chen, C. Antitumor and Immunomodulatory Activities of Ganoderma lucidum Polysaccharides in Glioma-Bearing Rats. Integr. Cancer Ther. 2018, 17, 674–683. [Google Scholar] [CrossRef]
  225. Fu, Y.; Shi, L.; Ding, K. Structure elucidation and anti-tumor activity in vivo of a polysaccharide from spores of Ganoderma lucidum (Fr.) Karst. Int. J. Biol. Macromol. 2019, 141, 693–699. [Google Scholar] [CrossRef]
  226. Min, B.S.; Gao, J.J.; Nakamura, N.; Hattori, M. Triterpenes from the spores of Ganoderma lucidum and their cytotoxicity against meth-A and LLC tumor cells. Chem. Pharm. Bull. 2000, 48, 1026–1033. [Google Scholar] [CrossRef]
  227. Gao, J.J.; Min, B.S.; Ahn, E.M.; Nakamura, N.; Lee, H.K.; Hattori, M. New triterpene aldehydes, lucialdehydes A—C, from Ganoderma lucidum and their cytotoxicity against murine and human tumor cells. Chem. Pharm. Bull. 2002, 50, 837–840. [Google Scholar] [CrossRef]
  228. Lin, Z.B.; Zhang, H.N. Anti-tumor and immunoregulatory activities of Ganoderma lucidum and its possible mechanisms. Acta. Pharmacol. Sin. 2004, 25, 1387–1395. [Google Scholar] [PubMed]
  229. Li, C.H.; Chen, P.Y.; Chang, U.M.; Kan, L.S.; Fang, W.H.; Tsai, K.S. Ganoderic acid X, a lanostanoid triterpene, inhibits topoisomerases and induces apoptosisofcancercells. Life Sci. 2005, 77, 252–265. [Google Scholar] [CrossRef] [PubMed]
  230. Tang, W.; Liu, J.W.; Zhao, W.M.; Wei, D.Z.; Zhong, J.J. Ganoderic acid T from Ganoderma lucidum mycelia induces mitochondria mediated apoptosis in lung cancer cells. Life Sci. 2006, 80, 205–211. [Google Scholar] [CrossRef] [PubMed]
  231. Jiang, J.; Grieb, B.; Thyagarajan, A.; Sliva, D. Ganoderic acids suppress growth and invasive behavior of breast cancer cells by modulating AP-1 and NFkappaB signaling. Int. J. Mol. Med. 2008, 21, 577–584. [Google Scholar] [CrossRef]
  232. Xu, K.; Liang, X.; Gao, F.; Zhong, J.; Liu, J. Antimetastatic effect of ganoderic acid T in vitro through inhibition of cancer cell invasion. Process Biochem. 2010, 45, 1261–1267. [Google Scholar] [CrossRef]
  233. Barbieri, A.; Quagliariello, V.; DelVecchio, V.; Falco, M.; Luciano, A.; Amruthraj, N.J. Anticancer and anti-inflammatory properties of Ganoderma lucidum extract effects on melanoma and triple-negative breast cancer treatment. Nutrients 2017, 9, 210. [Google Scholar] [CrossRef]
  234. Ye, H.; Jenson, A.B. Ganoderma lucidum on the Inhibition of Colorectal Cancer. Hematol. Med. Oncol. 2017, 2, 1–4. [Google Scholar] [CrossRef]
  235. Li, P.; Liu, L.; Huang, S.; Zhang, Y.; Xu, J.; Zhang, Z. Anti-cancer effects of a neutral triterpene fraction from Ganoderma lucidum and its active constituents on SW620 human colorectal cancer cells. Anticancer Agents Med. Chem. 2020, 20, 237–244. [Google Scholar] [CrossRef]
  236. Jiao, C.; Chen, W.; Tan, X.; Liang, H.; Li, J.; Yun, H. Ganoderma lucidum spore oil induces apoptosis of breast cancer cells in vitro and in vivo by activating caspase-3 and caspase-9. J. Ethnopharmacol. 2020, 247, 112256. [Google Scholar] [CrossRef]
  237. Wang, X.; Wang, B.; Zhou, L.; Wang, X.; Veeraraghavan, V.P.; Mohan, S.K. Ganoderma lucidum put forth anti-tumor activity against PC-3 prostate cancer cells viainhibitionofJak-1/STAT-3activity. Saudi J. Biol. Sci. 2020, 27, 2632–2637. [Google Scholar] [CrossRef]
  238. Liu, M.M.; Liu, T.; Yeung, S.; Wang, Z.; Andresen, B.; Parsa, C.; Orlando, R.; Zhou, B.; Wu, W.; Li, X.; et al. Inhibitory activity of medicinal mushroom Ganoderma lucidum on colorectal cancer by attenuating inflammation. Prec. Clin. Med. 2021, 4, 231–245. [Google Scholar] [CrossRef] [PubMed]
  239. Shahid, A.; Huang, M.; Liu, M.; Shamim, M.A.; Parsa, C.; Orlando, R. The medicinal mushroom Ganoderma lucidum attenuates UV-induced skin carcinogenesis and immunosuppression. PLoS ONE 2022, 17, e0265615. [Google Scholar] [CrossRef] [PubMed]
  240. De Marzo, A.M.; Platz, E.A.; Sutcliffe, S.; Xu, J.; Gronberg, H.; Drake, C.G.; Nakai, Y.; Isaacs, W.B.; Nelson, W.G. Inflammațion in prostate carciogenesis. Nat. Rev. Cancer 2007, 7, 256–269. [Google Scholar] [CrossRef] [PubMed]
  241. Hapuarachchi, K.K.; Wen, T.C.; Jeewon, R.; Wu, X.; Kang, J.C.; Hyde, K.D. Mycosphere Essays 7: Ganoderma lucidum—are the beneficial anti-cancer properties substantiated? Mycosphere 2016, 7, 305–332. [Google Scholar] [CrossRef]
  242. Zhang, J.; Tang, Q.; Zhou, C.; Jia, W.; Da Silva, L.; Nguyen, L.D.; Reutter, W.; Fan, H. GLIS, a bioactive proteoglycan fraction from Ganoderma lucidum, displays anti-tumour activity by increasing both humoral and cellular immune response. Life Sci. 2010, 87, 628–637. [Google Scholar] [CrossRef]
  243. Lai, C.Y.; Hung, J.T.; Lin, H.H.; Yu, A.L.; Chen, S.H.; Tsai, Y.C.; Shao, L.E.; Yang, W.B.; Yu, J. Immunomodulatory and adjuvant activities of a polysaccharide extract of Ganoderma lucidum in vivo and in vitro. Vaccine 2010, 28, 4945–4954. [Google Scholar] [CrossRef]
  244. Jan, R.H.; Lin, T.Y.; Hsu, Y.C.; Lee, S.S.; Lo, S.Y.; Chang, M.; Chen, L.K.; Lin, Y.L. Immuno-modulatory activity of Ganoderma lucidum-derived polysacharide on human monocytoid dendritic cells pulsed with Der p 1 allergen. BMC Immunol. 2012, 12, 31–41. [Google Scholar] [CrossRef]
  245. Chan, W.K.; Cheung, C.C.; Law, H.K.; Lau, Y.L.; Chan, G.C. Ganoderma lucidum polysaccharides can induce human monocytic leukemia cells into dendritic cells with immuno-stimulatory function. J. Hematol. Oncol. 2008, 21, 9–21. [Google Scholar] [CrossRef]
  246. Hsu, J.W.; Huang, H.C.; Chen, S.T.; Wong, C.H.; Juan, H.F. Ganoderma lucidum Polysaccharides Induce Macrophage-Like Differentiation in Human Leukemia THP-1 Cells via Caspase and p53 Activation. Evid. Based Complement. Altern. Med. 2011, 2011, 358717. [Google Scholar] [CrossRef]
  247. Guo, L.; Xie, J.; Ruan, Y.; Zhou, L.; Zhu, H.; Yun, X.; Jiang, Y.; Lu, L.; Chen, K.; Min, Z.; et al. Characterization and immunostimulatory activity of a polysaccharide from the spores of Ganoderma lucidum. Int. Immunopharmacol. 2009, 9, 1175–1182. [Google Scholar] [CrossRef]
  248. Hsu, H.Y.; Hua, K.F.; Lin, C.C.; Lin, C.H.; Hsu, J.; Wong, C.H. Extract of Reishi Polysaccharides Induces Cytokine Expression via TLR4-Modulated Protein Kinase Signaling Pathways. J. Immunol. 2004, 173, 5989–5999. [Google Scholar] [CrossRef] [PubMed]
  249. Altfeld, M.; Fadda, L.; Frleta, D.; Bhardwaj, N. DCs and NK cells: Critical effectors in the immune response to HIV-1. Nat. Rev. Immunol. 2011, 11, 176–186. [Google Scholar] [CrossRef] [PubMed]
  250. Chien, C.M.; Cheng, J.L.; Chang, W.T.; Tien, M.H.; Tsao, C.M.; Chang, Y.H.; Chang, H.Y.; Hsieh, J.F.; Wong, C.H.; Chen, S.T. Polysaccharides of Ganoderma lucidum alter cell immunophenotypic expression and enhance CD56+ NK-cell cytotoxicity in cord blood. Bioorg. Med. Chem. 2004, 12, 5603–5609. [Google Scholar] [CrossRef]
  251. Wang, P.Y.; Zhu, X.L.; Lin, Z.B. Antitumor and Immunomodulatory Effects of Polysaccharides from Broken-Spore of Ganoderma lucidum. Front. Pharmacol. 2012, 13, 135. [Google Scholar] [CrossRef]
  252. Wu, G.; Qian, Z.; Guo, J.; Hu, D.; Bao, J.; Xie, J.; Xu, W.; Lu, J.; Chen, X.; Wang, Y. Ganoderma lucidum Extract Induces G1 Cell Cycle Arrest, and Apoptosis in Human Breast Cancer Cells. Am. J. Chin. Med. 2012, 40, 631–642. [Google Scholar] [CrossRef] [PubMed]
  253. Lin, S.B.; Li, C.H.; Lee, S.S.; Kan, L.S. Triterpene-enriched extracts from Ganoderma lucidum inhibit growth of hepatoma cells via suppressing protein kinase C, activating mitogen-activated protein kinases and G2-phase cell cycle arrest. Life Sci. 2003, 72, 2381–2390. [Google Scholar] [CrossRef]
  254. Jedinak, A.; Thyagarajan-Sahu, A.; Jiang, J.; Sliva, D. Ganodermanontriol, a lanostanoid triterpene from Ganoderma lucidum, suppresses growth of colon cancer cells through ss-catenin signaling. Int. J. Oncol. 2011, 38, 761–767. [Google Scholar] [CrossRef] [PubMed]
  255. Chen, N.H.; Liu, J.W.; Zhong, J.J. Ganoderic acid T inhibits tumor invasion in vitro and in vivo through inhibition of MMP expression. Pharmacol. Rep. 2010, 62, 150–163. [Google Scholar] [CrossRef]
  256. Chen, N.H.; Liu, J.W.; Zhong, J.J. Ganoderic acid Me inhibits tumor invasion through down-regulating matrix metalloproteinases 2/9 gene expression. J. Pharmacol. Sci. 2008, 108, 212–216. [Google Scholar] [CrossRef]
  257. Hsu, C.L.; Yu, Y.S.; Yen, G.C. Lucidenic acid B induces apoptosis in human leukemia cells via a mitochondria-mediated pathway. J. Agric. Food Chem. 2008, 56, 3973–3980. [Google Scholar] [CrossRef]
  258. Liu, R.M.; Zhong, J.J. Ganoderic acid Mf and S induce mitochondria mediated apoptosis in human cervical carcinoma HeLa cells. Phytomed. Int. J. Phytother. Phytopharmacol. 2011, 18, 349–355. [Google Scholar] [CrossRef]
  259. Liu, R.M.; Li, Y.B.; Zhong, J.J. Cytotoxic and pro-apoptotic effects of novel ganoderic acid derivatives on human cervical cancer cells in vitro. Eur. J. Pharmacol. 2012, 681, 23–33. [Google Scholar] [CrossRef] [PubMed]
  260. Zhou, L.; Shi, P.; Chen, N.H.; Zhong, J.J. Ganoderic acid Me induces apoptosis through mitochondria dysfunctions in human colon carcinoma cells. Proc. Biochem. 2011, 46, 219–225. [Google Scholar] [CrossRef]
  261. Lu, H.; Ouyang, W.; Huang, C. Inflammation, a Key Event in Cancer Development. Mol. Cancer Res. 2006, 4, 221–233. [Google Scholar] [CrossRef] [PubMed]
  262. Shalapour, S.; Karin, M. Immunity, inflammation, and cancer: An eternal fight between good and evil. J. Clin. Investig. 2015, 125, 3347–3355. [Google Scholar] [CrossRef]
  263. Dudhgaonkar, S.; Thyagarajan, A.; Sliva, D. Suppression of the inflammatory response by triterpenes isolated from the mushroom Ganoderma lucidum. Int. Immunopharmacol. 2009, 9, 1272–1280. [Google Scholar] [CrossRef]
  264. Miron, L.; Gafton, B.; Marinca, M. Tumor angiogenesis- implications in cancer therapy. Iasi J. Surg. 2010, 6, 2. [Google Scholar]
  265. Cao, Q.Z.; Lin, Z.B. Antitumor and anti-angiogenic activity of Ganoderma lucidum polysaccharides peptide. Acta Pharmacol. Sin. 2004, 25, 833–838. [Google Scholar]
  266. Smina, T.P.; Mathew, J.; Janardhanan, K.K.; Devasagayam, T.P. Antioxidant activity and toxicity profile of total triterpenes isolated from Ganoderma lucidum (Fr.) P. Karst occurring in South India. Environ. Toxicol. Pharmacol. 2011, 32, 438–446. [Google Scholar] [CrossRef]
  267. Smina, T.P.; De, S.; Devasagayam, T.P.; Adhikari, S.; Janardhanan, K.K. Ganoderma lucidum total triterpenes prevent radiation-induced DNA damage and apoptosis in splenic lymphocytes in vitro. Mutat. Res. 2011, 726, 188–194. [Google Scholar] [CrossRef]
  268. Lu, H.M.; Uesaka, T.; Katoh, O.; Kyo, E.; Watanabe, H. Prevention of the development of preneoplastic lesions, aberrant crypt foci, by a water-soluble extract from cultured medium of Ganoderma lucidum (Rei-shi) mycelia in male F344 rats. Oncol. Rep. 2001, 8, 1341–1345. [Google Scholar] [CrossRef] [PubMed]
  269. Lu, H.; Kyo, E.; Uesaka, T.; Katoh, O.H.W. A water-soluble extract from cultured medium of Ganoderma lucidum (Rei-shi) mycelia suppresses azoxymethane-induction of colon cancers in male F344 rats. Oncol. Rep. 2003, 10, 375–379. [Google Scholar] [CrossRef]
  270. Lee, J.M.; Kwon, H.; Jeong, H.; Lee, J.W.; Lee, S.Y.; Baek, S.J.; Surh, Y.J. Inhibition of lipid peroxidation and oxidative DNA damage by Ganoderma lucidum. Phytother. Res. 2001, 15, 245–249. [Google Scholar] [CrossRef] [PubMed]
  271. Zhao, W.; Jiang, X.; Deng, W.; Lai, Y.; Wu, M.; Zhang, Z. Antioxidant activities of Ganoderma lucidum polysaccharides and their role on DNA damage in mice induced by cobalt-60 gamma-irradiation. Food Chem. Toxicol. 2012, 50, 303–309. [Google Scholar] [CrossRef]
  272. Ulbricht, C.; Isaac, R.; Milkin, T.; Poole, E.A.; Rusie, E.; Grimes Serrano, J.M.; Weissner, W.; Windsor, R.C.; Woods, J. An evidence-based systematic review of stevia by the Natural Standard Research Collaboration. Cardiovasc. Hematol. Agents Med. Chem. 2010, 8, 113–127. [Google Scholar] [CrossRef] [PubMed]
  273. Gill, S.K.; Rieder, M.J. Toxicity of a traditional Chinese medicine, Ganoderma lucidum, in children with cancer. Can. J. Clin. Pharmacol. 2008, 15, 275–285. [Google Scholar]
  274. Bishop, K.S.; Kao, C.H.; Xu, Y.; Glucina, M.P.; Paterson, R.R.M.; Ferguson, L.R. From 2000 years of Ganoderma lucidum to recent developments in nutraceuticals. Phytochemistry 2015, 114, 56–65. [Google Scholar] [CrossRef]
  275. Tran, H.B.; Yamamoto, A.; Matsumoto, S.; Ito, H.; Igami, K.; Miyazaki, T.; Kondo, R.; Shimizu, K. Hypotensive effects and angiotensin-converting enzyme inhibitory peptides of reishi (Ganoderma lingzhi) auto-digested extract. Molecules 2014, 19, 13473–13485. [Google Scholar] [CrossRef]
  276. Karwa, A.; Gaikwad, S.; Rai, M.K. Mycosynthesis of silver nanoparticles using Lingzhi or Reishi medicinal mushroom, Ganoderma lucidum (W. Curt.: Fr.) P. Karst. and their role as antimicrobials and antibiotic activity enhancers. Int. J. Med. Mush. 2011, 13, 483–491. [Google Scholar] [CrossRef]
Figure 1. Appearance of Ganoderma lucidum in Europe (Romania) and Asia [61,62]. (a) G. lucidum from Neamt County, Romania, adapted from [61]. (b) G. lucidum from Bacau County, Romania, adapted from [61]. (c) Reishi or Lingzhi from India, adapted from [62].
Figure 1. Appearance of Ganoderma lucidum in Europe (Romania) and Asia [61,62]. (a) G. lucidum from Neamt County, Romania, adapted from [61]. (b) G. lucidum from Bacau County, Romania, adapted from [61]. (c) Reishi or Lingzhi from India, adapted from [62].
Antioxidants 12 01907 g001
Figure 2. Different cultivation methods for fruit body and mycelium of Ganoderma lucidum, adapted from [14].
Figure 2. Different cultivation methods for fruit body and mycelium of Ganoderma lucidum, adapted from [14].
Antioxidants 12 01907 g002
Figure 3. Bioactive compounds in G. lucidum and biological actions involved in anticancer activity, adapted from Ahmad [13].
Figure 3. Bioactive compounds in G. lucidum and biological actions involved in anticancer activity, adapted from Ahmad [13].
Antioxidants 12 01907 g003
Figure 4. Diagram of the extraction of polysaccharides from G. lucidum, adapted from [85].
Figure 4. Diagram of the extraction of polysaccharides from G. lucidum, adapted from [85].
Antioxidants 12 01907 g004
Figure 5. The structure of a Gl-Ps chain consisting of β-(1→3) bonds with some β-(1→6) branches, adapted from [85].
Figure 5. The structure of a Gl-Ps chain consisting of β-(1→3) bonds with some β-(1→6) branches, adapted from [85].
Antioxidants 12 01907 g005
Figure 6. Skeletal structure of a typical terpenoid and other terpenoids 1–10 [118].
Figure 6. Skeletal structure of a typical terpenoid and other terpenoids 1–10 [118].
Antioxidants 12 01907 g006
Figure 7. Structure of ganoderic acids from G. lucidum, adapted from [122].
Figure 7. Structure of ganoderic acids from G. lucidum, adapted from [122].
Antioxidants 12 01907 g007
Figure 8. Specific structures for other ganoderic compounds [152].
Figure 8. Specific structures for other ganoderic compounds [152].
Antioxidants 12 01907 g008
Figure 9. Mechanisms of action of polysaccharide and triterpene extracts isolated from G. lucidum with anticancer effect, adapted from [35].
Figure 9. Mechanisms of action of polysaccharide and triterpene extracts isolated from G. lucidum with anticancer effect, adapted from [35].
Antioxidants 12 01907 g009
Table 1. Proximate composition of the fruiting body of G. lucidum from different country origins, reported as a percentage.
Table 1. Proximate composition of the fruiting body of G. lucidum from different country origins, reported as a percentage.
Mushroom Origin/
Constitutes
From
Bangladesh
From
Bangladesh
From
Taiwan
From
China
From
Spain
From
India
From
Nigeria
Moisture %12.1947---7.52.78 ± 0.05
Ash %3.936.31.81.21 ± 0.06 2.31 ± 0.12 18.78.42 ± 0.13
Water-soluble proteins %28.619.507–87.47 ± 0.22 11.70 ± 0.35 23.616.79 ± 0.13
Total lipids %2.43.003–5--5.81.52 ± 0.09
Fatty acid---1.44 ± 0.101.27 ± 0.09 -1.22 ± 0.07
Total carbohydrates %44.915.4126–289.88 ± 1.0411.02 ± 1.1642.863.27 ± 0.20
Dietary fibers %14.672.45976.81 ± 3.4669.35 ± 3.12-7.77 ± 0.34
References[21][30,71][70][72][72][73][74]
Table 2. Structural characteristics, extraction, and fractionation techniques of polysaccharides from G. lucidum.
Table 2. Structural characteristics, extraction, and fractionation techniques of polysaccharides from G. lucidum.
Mushroom Part of G. lucidumMonosaccharide CompositionBackboneExtraction and FractionationBioactivityReferences
1Fruiting bodyGlucose, rhamnose
glucose, galactose, rhamnose
β(1→3)-Glucan; -(1→3)-(1→4)-, (1→6)Heteroglycan
α-(1→4), β-(1→6)-heteroglycan
Hot-water extraction; DEAE-cellulose and gel filtration; chromatography Antioxidant activity[94]
2Fruiting body (cultivated)Mannose, rhamnose, glucose, galactoseHeteroglucans (GLP, GLP1, GLP2, GLP3, GLP4) Main glycosidic bond Ultrasonic extraction; Sevag method; ethanol precipitation; ultrafiltration membranesAntioxidant activity in vitro by DPPH scavenging activity; reducing power; Fe2+ chelating activity; ORAC[95]
3Fruiting body (cultivated)Glucose, galactose, mannose, arabinoseHeteropolysaccharide (GL-1; GL-V) (1→4)-galactan, HeteropolysaccharideSoluble in water and in ethyl-acetate;
Sevag method; dialysis
Bioactive compounds are an important source of anticancer agents[96]
4ExtracellularGalactose, mannose, glucose, arabinose, rhamnoseα-(1→4)-GalactoseDEAE-Sephcel and Sephadex G200.Enhance T- and B-lymphocyte proliferation and antibody production[97]
5Fruiting bodyGalactose, glucose, fucoseα-(1→6)-galactose
α-(1→3)-Glucose
Hot-water extraction; DEAE-Sepharose Fast-Flow and Sephacryl S-300Immunostimulatory activity of spleen lymphocyte proliferation[98]
6Fruiting bodyGlucose, galactose, arhamnoseHeteroglycan α-(1→4), β-(1→6)Hot-water extractionImmunologically active; proliferation of B-lymphocytes with important immunologic activity[99]
7Fruiting body (cultivated)Glucose, galactose, mannose, arabinose, xylose, fucoseHeteropolysaccharides glucans (1→3)-β-D-glucan with a few short (1→4)-linked glucosyl unitsExtraction and separation of fractions with hot water, cold and hot 1 M NaOHAntitumor activity against sarcoma solid tumor[100]
8Fruiting bodyGalactose, glucose, fucoseα-(1→6)-, (1→2,6)-Galactose
β-(1→3)-, (1→4,6)-Glucose
Hot-water extraction; DEAE-Sepharose Fast-Flow and Sepharose CL-6BAn immunostimulating potential[101]
9Fruiting bodyGlucose, galactose, mannoseβ-(1→3)(1→4)(1→6)-Glucan
Heteropolysaccharides
Hot-water extraction; DEAE-cellulose-32 and Sephacryl S-200 hPronounced antioxidant activity in free radicals scavenging and Fe2+ chelating [102]
10Fruiting body (wild)Galactose, rhamnose, and glucose in mole ratio of 1.00:1.15:3.22 Water-soluble polysaccharide α-(1→6)-, (1→2,6) Galactose β-(1→3)-, (1→4,6) GlucoseHot water and ethanol precipitation; DEAE-Sepharose Fast Flow and Sephacryl S-300Neutral heteropolysaccharide, which reported antihyperglycemia effects[103]
11Mycelium
(cultivated)
Rhamnose, arabinose, mannose, glucose, galactoseHeteropolysaccharide α-D-Glc (1→6), α-D-Glc,α-D-Man (rhamnose and arabinose residues in the side chain)Hot water; ethanol precipitation; Sevag method; dialysisAntitumor activity against Human hepatocarcinoma cell line (HepG2) and tumor xenografts in ICR mice[104]
12Fruiting bodyGlucoseBranched homo-glucan (GLP0; GLP1) (1→3)-β-D-glucan with (1→6)-β-D branchesHot water followed by ethanol precipitationInduced a cascade of immunomodulatory cytokines against sarcoma 180 solid tumor[105]
Table 3. Minerals of G. lucidum fruiting body.
Table 3. Minerals of G. lucidum fruiting body.
Elementsmg/100 gmg/100 g% or ppmppm
Potassium4323.5901.11 ± 0.04 (%)-
Phosphorus2254.15030.17 ± 1.29 (ppm)-
Sulfur129---
Magnesium7.951.0300.34 ± 0.01 (%)50.76 ± 1.19
Sodium2.82375229.88 ± 0.34 (ppm)-
Calcium1.888321.99 ± 0.04%-
Copper27-7.43 ± 0.13 (ppm)5.49 ± 0.35
Manganese22-71.06 ± 1.56 (ppm)20.19 ± 0.54
Iron2.2282.6121.37 ± 1.82(ppm)130.60 ± 1.63
Zinc0.7-51.49 ± 2.16 (ppm)8.45 ± 0.38
References[30,71][153][74][156]
Table 4. Amino acids content in Ganoderma lucidum mushroom.
Table 4. Amino acids content in Ganoderma lucidum mushroom.
Amino AcidAspartic AcidThreonineSerineGlutamic AcidProlineGlycineAlanineValine
AspThrSerGluProGlyAlsVal
mg AA/g protein11766541206010810061
References[27,184][27,184][27,184][27,184][27,184][27,184][27,184][27,184]
Amino acidMethionineIsoleucineLeucinePhenylalanineTyrosineHistidineLysineArginine
MetIleLeuPhe TyrHisLysArg
mg AA/g protein 636552816122122
References[27,184][27,184][27,184][27,184][27,184][27,184][27,184][27,184]
Table 5. Compounds with antioxidant activity from G. lucidum.
Table 5. Compounds with antioxidant activity from G. lucidum.
Total TriterpenoidsTotal Polysaccharides Total Polyphenol Content (TPC)Total Flavonoid Content (TFC)Ascorbic Acid
/g d.w.mg glucose equiv./g d.w.mg/100 g d.w.mg/100 g d.w.mg/100 g d.w.
196.03–643.06769.1 112.5333.3–43.49912.3834.09–38.0830.51–32.2
[188] [192] [188] [21] [192] [21] [21]
Table 6. The phenolic acids from G. lucidum [192].
Table 6. The phenolic acids from G. lucidum [192].
Nr.Phenolic AcidsQuantity (mg/100 g DW of Extract)
1Tricaffeoyl-glucosyl-glucoside13.54 ± 0.23
2Tricaffeoyl-glucosyl23.79 ± 0.24
3Caffeoyltrihexoside38.02 ± 0.30
4Protocatechuic acid hexoside19.09 ± 0.15
51-Caffeoylquinic acid505.89 ± 3.21
6trans-5-ꓑ-coumaroylquinic acid0.46 ± 0.01
75-Caffeoylquinic acid95.01 ± 0.92
8Caffeoyl-2-hydroxyethane-1.1.2-tricarboxylic acid213.89 ± 1.52
9Yunnaneic acid F1.29 ± 0.01
10Salvianolic acid B1.39 ± 0.01
Sum912.38 ± 20.14
Means ± SD (p ≤ 0.05; n = 3).
Table 7. Antioxidant capacity in the extract of G. lucidum.
Table 7. Antioxidant capacity in the extract of G. lucidum.
DPPHFRAPABTS
(%)(µMol TE/g)(μg/100 g)(µMol TE/g)(µMol TE/g)
24.04 ± 0.3351.3 ±1.04614.83 ± 0.0549.87 ± 1.5881.26 ± 1.10
[21][192][21][192][192]
Means ± SD (p ≤ 0.05; n = 3).
Table 8. Fatty acids content of two Ganoderma species [72].
Table 8. Fatty acids content of two Ganoderma species [72].
Fatty AcidsG. lingzhi ± 15%G. lucidum ± 15%
Total monounsaturated fatty acids37.528.68
Total polyunsaturated fatty acids43.8449.93
Total saturated fatty acids18.6420.77
Total99.9899.38
Table 9. Anticancer activity of polysaccharides and triterpenes from G. lucidum.
Table 9. Anticancer activity of polysaccharides and triterpenes from G. lucidum.
G. lucidum CompoundsActions and Effects of Gl-Ps on Antitumor
Activity
Mechanism of ActionReferences
1Gl-PsEffects on dendritic cellsGl-Ps acts on the maturation and function of cultured murine bone marrow-derived dendritic cells (DCs).[211]
2Gl-PsEffect on cytotoxicityGl-Ps acts with a specific T-lymphocyte cytotoxic (CTL) mechanism, which has been pulsed with the tumor antigen P815.[212]
3Gl-PsEvaluation of immunomodulatory effect on cytokines Explain the mechanism of action on macrophages in which Gl-Ps (fractions) activate kinase to induce, in turn, activation of IL-1, IL-2, and TNF-α.[213]
4Gl-PsEffect of cytokine-induced killer cells (CIK)Gl-Ps decreases the number of lymphokine-activated cytokines (LAK) and CIK-induced cytokine-killing cells.[214]
5Gl-PsActions in immunopotentiation therapy against induced immunosuppression Gl-Ps extract at low doses leads to increased immunological effector cell activity in immunosuppressed mice.[215]
6Gl-PsEffect on antioxidant enzyme activity Gl-Ps from G. lucidum significantly reduced malondialdehyde (MDA) production and increased the activity of serum antioxidant enzymes in ovarian cancer therapy in rats.[216]
7Gl-PsGl-Ps suppresses tumorigenesis, inhibits tumor growth Gl-Ps affects immune cells, including B-lymphocytes, T-lymphocytes, dendritic cells, and natural killer cells. They are mediated by immunomodulatory, anti-angiogenic, and cytotoxic effects.[217]
8Gl-PsAntitumor effects by stimulating host immune functionGl-Ps acts directly in activating lymphocytes that have been tested by incubating Gl-Ps with an antigen-deficient tumor cell line. Also, Gl-Ps acts on B16F10 melanoma cells.[218]
9Gl-PsAntitumor effects by stimulating host immune functionGl-Ps can induce lymphocyte proliferation through action on B16F10 melanoma cells and IFN-γ production.[219]
10Gl-PsAntitumor activity manifested by a mixture of Gl-Ps and sulfates.Gl-Ps sulfate showed remarkable inhibition of rat Heps proliferation.[220]
11Gl-PsTherapeutic potential in inflammatory breast cancer (IBC).Study results provide evidence that Gl-Ps treatment suppresses protein synthesis and tumor growth by affecting survival signaling pathways in mice injected with IBC cells, suggesting a natural therapeutic potential for breast cancer.[221]
12Gl-PsThe ability of isolated Gl-Ps fractions (F3) to induce innate inflammatory cytokinesEnhanced Th1 response with high levels of IFN-γ and IL-2. Cell wall Gl-Ps were inducers of innate inflammatory cytokines, and extracellular Gl-Ps demonstrated a high capacity to modulate cytokine responses to IL-17 production.[222]
13Gl-Ps Potential anticancer activityThey discussed the mechanisms of anticancer activity attributed to Gl-Ps by highlighting immunomodulatory, anti-proliferative, pro-apoptotic, antimetastatic and anti-angiogenic effects.[223]
14Gl-PsAntitumor action and immunomodulatory effects of Gl-Ps in ratsGl-Ps increased the serum concentration of Il-2, INF-γ and tumor necrosis factor-α. It increased the cytotoxic activity of natural killer cells and T cells and led to prolonged lifespan of brain glioma-bearing rats.[224]
15Gl-Ps,
spores
Antitumor action of a novel polysaccharide with an estimated average molecular weight of 1.5 × 104 Da In vivo antitumor activity tests showed that Gl-Ps could significantly inhibit S180 tumor growth in mice. No drug-related toxic reactions were observed. [225]
16Gl-TsGanoderic acids from spores and their cytotoxicity The cytotoxicity of the compounds isolated from the Ganoderma spores was carried out in vitro against Meth-A and LLC tumor cell lines.[227]
17Gl-TsAnticancer study of lucialdehydes B, C (2,3), ganodermanonol, and ganodermanondiolCytotoxic mechanism. Lucialdehyde C exhibited the most potent cytotoxicity against CLL, T-47D, sarcoma 180, and Meth-A tumor cells.[227]
18Gl-Ps
Gl-Ts
Antitumor effect of aqueous extract;
cytotoxic activity of alcoholic extract
Manifestation of a significant antitumor effect in several tumor-bearing animals;
manifestation of an anti-angiogenic effect that may be involved in the antitumor activity.
[228]
19Gl-TsAbility of ganoderic acid X (GAX) to inhibit topoisomerases and interfere with apoptosisMechanisms of chromosomal DNA degradation, cancer cell apoptosis, mitochondrial membrane disruption, and caspase-3 activation have been elucidated upon GAX treatment of HuH-7 human hepatoma cells.[229]
20Gl-TsCytotoxicity of GA-T on different human carcinomaIt was shown in vivo to significantly inhibit proliferation of lung cancer cells by inducing apoptosis by GA-T[230]
21Gl-TsThe effect of ganoderic acids A, F, and H on breast cancer cells was evaluatedGA-A, GA-F, and GA-H suppressed cell proliferation, colony formation, and invasive behavior of MDA-MB-231 cells. They have biological effects by inhibiting transcription factors AP-1 and NF-κB.[231]
22GA-TStudies of anti-invasive and antimetastatic mechanisms of GA-T in vitro in lung cancerGA-T dose-dependently inhibited 95-D cell migration by wound healing assay, promoting cell aggregation and inhibiting cell adhesion to the extracellular matrix (ECM). GA-T prevents tumor metastasis in highly metastatic lung carcinoma.[232]
23Gl-Ts.Anticancer, anti-inflammatory, and antimetastatic activities of G. lucidum extracts Gl-Ts from G. lucidum reduces the production of IL-8, IL-6, MMP-2, and MMP-9 in breast cancer and melanoma cells. They decrease cancer cell viability in a time and dose-dependent manner.[233]
24GL-TsInvestigation of Gl-Ts with activity in inhibiting growth of pulmonary carcinoma metastates and suppressing colonic inflammation The triterpene extracts exhibit inhibitory activity against foodborne carcinogen-induced mouse colon carcinogenesis. All suppressive functions were enhanced by high doses of triterpene extract.[234]
25Gl-Ts
NTF, ATF
Evaluation of anticancer effects of NTF (neutral triterpene fraction) and ATF (acidic triterpene fraction) on human colorectal cancer The cytotoxic effects of Gl-Ts on human colon cancer cells SW480, SW620, SW116, and mouse embryonic fibroblast cells NIH3T3 were studied. Compounds isolated from NTF acted as antitumorals by inducing apoptosis.[235]
27GlSOMechanistic investigation of the anticancer-gene effect of GlSO (G. lucidum spore oil) on mammary cancer cells Growth of MDA—MB-231 cells, in vitro, were inhibited by treatment with GlSO (0.2, 0.4, and 0.6 µL/mL). In vitro, GlSO increased Bax and caspase-3 expression but did not affect caspase-8 expression.[236]
27Gl-TsAnticancer potential of G. lucidum against prostate cancer (PC-3)G. lucidum has been shown to prevent prostate cancer cell growth and stimulate apoptosis in PC-3 cells by preventing STAT-3 translocation (signal transduction and activation of transcription).[237]
28Gl-TsEffects on colorectal cancer. Involves suppression of NF-κB-regulated inflammation and carcinogenesisIn vitro administration of GLSF extract at non-toxic concentrations to mice inoculated with CT27 tumor cells significantly potentiated paclitaxel-induced growth inhibition and apoptosis in CT27 and HCT-15 cells. [238]
29Gl-Ps, Gl-TsEvaluation of the effects on skin carcinogenesis analyzed on JB6 cells in SKH-1 mice Reduced incidence and multiplicity of skin tumors. In tumor-free skin tissue of mice, Gl-Ps and Gl-Ts attenuated UV-induced epidermal thickening. Gl-SF increased CD8 and Granzyme B expression.[239]
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

Cadar, E.; Negreanu-Pirjol, T.; Pascale, C.; Sirbu, R.; Prasacu, I.; Negreanu-Pirjol, B.-S.; Tomescu, C.L.; Ionescu, A.-M. Natural Bio-Compounds from Ganoderma lucidum and Their Beneficial Biological Actions for Anticancer Application: A Review. Antioxidants 2023, 12, 1907. https://doi.org/10.3390/antiox12111907

AMA Style

Cadar E, Negreanu-Pirjol T, Pascale C, Sirbu R, Prasacu I, Negreanu-Pirjol B-S, Tomescu CL, Ionescu A-M. Natural Bio-Compounds from Ganoderma lucidum and Their Beneficial Biological Actions for Anticancer Application: A Review. Antioxidants. 2023; 12(11):1907. https://doi.org/10.3390/antiox12111907

Chicago/Turabian Style

Cadar, Emin, Ticuta Negreanu-Pirjol, Carolina Pascale, Rodica Sirbu, Irina Prasacu, Bogdan-Stefan Negreanu-Pirjol, Cezar Laurentiu Tomescu, and Ana-Maria Ionescu. 2023. "Natural Bio-Compounds from Ganoderma lucidum and Their Beneficial Biological Actions for Anticancer Application: A Review" Antioxidants 12, no. 11: 1907. https://doi.org/10.3390/antiox12111907

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