Skip to main content

A review on the cultivation, bioactive compounds, health-promoting factors and clinical trials of medicinal mushrooms Taiwanofungus camphoratus, Inonotus obliquus and Tropicoporus linteus

Abstract

Medicinal mushrooms, such as Taiwanofungus camphoratus, Inonotus obliquus, and Tropicoporus linteus, have been used in traditional medicine for therapeutic purposes and promotion of overall health in China and many East Asian countries for centuries. Modern pharmacological studies have demonstrated the large amounts of bioactive constituents (such as polysaccharides, triterpenoids, and phenolic compounds) available in these medicinal mushrooms and their potential therapeutic properties. Due to the rising demand for the health-promoting medicinal mushrooms, various cultivation methods have been explored to combat over-harvesting of the fungi. Evidence of the robust pharmacological properties, including their anticancer, hypoglycemic, hypolipidemic, antioxidant, and antiviral activities, have been provided in various studies, where the health-benefiting properties of the medicinal fungi have been further proven through numerous clinical trials. In this review, the cultivation methods, available bioactive constituents, therapeutic properties, and potential uses of T. camphoratus, I. obliquus and T. linteus are explored.

Introduction

Natural products, consisting of an extensive range of structural and chemical diversity, have been used as therapeutic agents for various human ailments and diseases for many years [1]. Natural products and their derivatives have been approved for use in various fields, including oncology, as well as in the anti-infective area [2]. The biological activities of these natural sources can be attributed to the presence of medicinal compounds found in them, such as polysaccharides and terpenoids [3].

Medicinal mushrooms, including Taiwanofungus camphoratus, Inonotus obliquus, and Tropicoporus linteus, are rich sources of these bioactive constituents. Hispidin, ergosterol peroxide, protocatechuic acid, and betulin are all biologically-active compounds that can be found in the medicinal mushrooms. Having a long history of use in traditional and folk medicine, they have been used for the amelioration of conditions ranging from cancer [4] to heart and liver disease [5]. Modern scientific research has validated their claimed properties, where the increasingly developed field of medical mushroom research demonstrated the potent activities of bioactive compounds isolated from the fungi. They contain great potential to treat a wide range of conditions, including tumor [6, 7], diabetes [8,9,10], viral infections [11,12,13], and disease related to oxidative stress [14,15,16]. Evidence of their health-benefiting properties have been established through various studies and clinical trials [17]. Therefore, due to their significant pharmacological and health-promoting properties, various efforts have been made to increase their production sustainably [18].

The market value of medicinal mushrooms and their derived products was estimated to be USD 6.0 billion in 1999, with demand increasing between 20 and 40% annually since then [19]. With a selling price of up to US$25,000/kg, the annual output value of T. camphoratus in Taiwan is estimated to be about NT$3 billion [20] and a global market value of $100 million per year [21]. For I. obliquus, the current global market value of US$28.2 billion is expected to rise to US$87 billion by 2034 [22], while total sales of various T. linteus products in Japan and South Korea alone in 2023 have exceeded 10 billion yen (US$65 million) [23].

In view of the rising demand and expanding market potential, this review aims to explore the growth and cultivation, bioactive components, therapeutic potentials, medical evidence of the pharmacological properties of medicinal mushrooms Taiwanofungus camphoratus, Inonotus obliquus, and Tropicoporus linteus, with a goal of providing a useful reference for necessary information for the further study of the fungi.

Growth and cultivation

Taiwanofungus camphoratus

T. camphoratus thrives in the natural habitat of Taiwan's mountainous regions and grows at altitudes between 450 and 2000 m, where the Cinnamomum kanehirai tree serves as its primary nutrient source [24]. A distinguishing feature of T. camphoratus is its growth pattern, whereby the fruiting body can only develop to full maturity when the host tree has naturally aged away (Fig. 1).

Fig. 1
figure 1

Fruiting bodies of (A) Taiwanofungus camphoratus, (B) Inonotus obliquus, and (C) Tropicoporus linteus on their host trees [25, 26] (Min and Kang 2021)

In the past, unsustainable methods have been practiced due to the high demand for wild T. camphoratus and the widespread use of C. kanehirai wood for furniture production. Fuelled by the need to satisfy this demand, excessive harvesting of the mushroom and its host tree has led to disastrous consequences on the availability of the fungi and its habitat [27]. Realizing the ecological importance of C. kanehirai and T. camphoratus’ reliance on this species, the Taiwanese government has intervened by setting conservation measures in place to prevent overexploitation and maintain the ecosystem’s ideal state [28]. However, due to its limited availability and difficult cultivation, T. camphoratus remains expensive and is often referred to as the “ruby in Taiwanese forests” [29].

Although between 80 and 85% of medicinal mushroom products are obtained from fruiting bodies, the continual rise in market demand has made mycelial products the wave of the future [30]. Artificial cultivation methods for T. camphoratus were established in order to reduce the effects of overexploitation and sustainably satisfy market demand. Two main cultivation techniques have been commonly utilized to meet the increasing demand, namely submerged fermentation and solid-state fermentation [31]. Each method possesses its own advantages and limitations, which have been described in Table 1. For example, the ability to precisely control physical, chemical, and biological factors using submerged fermentation allows high reproducibility of metabolites, thus increasing the ease of scale-up for industrial production purposes [32, 33]. Furthermore, recent efforts have been made to enhance the efficiencies of cultivation methods, such as the application of low-frequency alternating magnetic field to enhance mycelium growth, expression of genes involved in amino acid metabolism and synthesis, and accumulation of active metabolites in cultivated T. camphoratus [18].

Table 1 A comparison between two widely used techniques for the cultivation of medicinal mushrooms [34,35,36,37,38]

Inonotus obliquus

The Chaga mushroom, scientifically known as I. obliquus, is a parasitic fungus belonging to the family of Hymenochaetaceae that gained attention for its unique appearance and potential health benefits [39]. The development of I. obliquus can be recognized by a canker-like look in a shade of dark brown to black, resembling burnt wood, which is a result of fungal mycelium accumulation [40]. Despite its slow growth, I. obliquus ultimately reaches maturity, developing a distinctive texture and occasionally exposing its inner yellow flesh. The fungi frequently cohabit with birch trees in moist, wetland habitats through symbiotic interactions, whereby regions such as Russia, Korea, China, Eastern and Northern Europe, northern parts of the United States, and portions of Canada are included in their geographic distribution [41].

I. obliquus must be harvested using special procedures to protect the wellbeing of both the host tree and the fungi. It can be carefully removed from a tree's bark using equipments such as hatchets or hammers without harming the inner layers [42]. The optimal time to harvest I. obliquus is in late winter when the host tree is dormant and nutrient accumulation is at its greatest. This method allows for repeated regeneration of I. obliquus on the same tree while protecting the health of the host tree. A complex relationship exists between I. obliquus and its host tree, whereby the viability of the fungi declines with the health of the tree [42]. Thus, it is essential to understand the symbiotic relationship between I. obliquus and its host to enable appropriate cultivation methods for this priceless natural resource.

Due to the slow-growing nature and hence limited supply of I. obliquus, various cultivation methods have been employed to address this issue. A 2005 study on the optimum culture condition of I. obliquus provided valuable basic information on the in vitro cultivation conditions, including the media used and additions such as amino acids and yeast extract for the improvement of fungal growth and biomass production [43]. Submerged fermentation has been commonly used for the production of I. obliquus, whereby careful optimisation of optimum broth formula for submerged culturing of I. obliquus mycelium has successfully enhanced mycelial biomass yields [44]. Other efforts have also been made to determine the optimum I. obliquus strains and cultivation substrates for producing productive strains of the medicinal fungi, where the use of Betula wood as inoculation dowels was suggested to shorten production time [45].

Tropicoporus linteus

T. linteus is a well-known medicinal fungus that has drawn the interest of researchers and traditional medicine practitioners. It is known by several names, including "sanghuang" in China, "meshimakobu" in Japan, and "sangwhang" in Korea. Due to limited availability in the wild, its development and harvest present added challenges.

T. linteus is usually found on the trunks Populus Linn., Quercus Linn., Toxicodendron vernicifluum and Morus alba Linn. in its natural habitat [46]. It stands out in the fungi world due to its unusual pileate, perennial, horseshoe-shaped basidiocarps [46]. When dried, the pore surface changes from rusty brown to brown, the pial surface from a dark brown hue to black, and the tubes attain a cinnamon-yellowish-brown tint. The months of April and May are when T. linteus is most suitable for harvesting in the wild due to favourable environmental circumstances for its growth [47].

Despite its significant therapeutic value, the availability of T. linteus for traditional medicine and industrial use has been constrained by its low abundance in its natural environment, where the demand for this rare resource simply exceeds what the wild population can supply [48]. Researchers and cultivators have concentrated on artificially producing T. linteus to address this issue, where cultivation procedures to produce fruiting bodies often require growing it on solid artificial substrate [49]. Optimal submerged culture composition and conditions for maximum mycelial biomass have also been determined, enabling a more sustainable and steady supply for medical and scientific uses, and utilization of this unique fungus to its maximum potential [50]. Furthermore, artificial cultivation of T. linteus strains on oak logs has been successful, with resulting polysaccharides retaining their valuable bioactive properties, including immunomodulatory and anticancer effects [51].

Bioactive compounds

The chemical structures of bioactive compounds isolated from the three medicinal mushrooms explored in this review are presented in (Fig. 2).

Taiwanofungus camphoratus

Polysaccharides

A water-soluble polysaccharide, composed of β-D-glucan, was isolated and purified from cultured Taiwanofungus camphoratus powder [52]. The polysaccharide was found to have a molecular weight of 17.2 kDa, where further structural elucidation revealed a segmental and repetitive structure of the molecule, composed of 1,3,6-β-d-glucose, 1,3-β-d-glucose, and 1,4-α-d-galactose at a 2.21:1.00 mol ratio of glucose to galactose. The β-d-glucan was demonstrated to exert anti-inflammatory and anti-oxidation activities in lipopolysaccharide (LPS) induced-human hepatocytes to thus provide a hepatoprotective effect. An exopolysaccharide [53] and galactoglucan primarily composed of galactose (1,6-β-d-galactose) and glucose (1,6-β-d-glucose, 1,3-α-d-glucose, and 1,4-β-d-glucose), with terminal α-d-6-deoxyglucose and α-d-glucose [54] isolated from T. camphoratus also exhibited inflammation-reducing capabilities.

Fig. 2
figure 2

Chemical structures of bioactive compounds isolated from Taiwanofungus camphoratus (1–38 terpenes and terpenoids; 39–42 quinone derivatives; 43–47 maleic and succinic acid derivatives), Inonotus obliquus (38, 48–88 terpenes and terpenoids; 89–113 phenolic compounds), and Tropicoporus linteus (114–132 terpenes and terpenoids; 93, 95, 97, 99, 111, 112, 133–142 phenolic compounds)

Three other polysaccharides from the cell wall of T. camphoratus mycelia obtained through hot water, cold alkali, and hot alkali extraction were demonstrated to also mainly consist of glucose, with molar percentages of 75.32, 78.06, and 87.28 respectively [55]. The hot water extracted polysaccharide mainly consisted of 1,4-d-glucose, while those isolated through cold and hot alkali extractions were primarily composed of 1,3-d-glucose and were reported to contain potent in vitro antioxidant properties, suggesting the potential contribution of 1,3-d-glucose towards the observed activities exerted by the alkali-soluble polysaccharides. Another polysaccharide extracted from the mycelia of T. camphoratus was shown to possess antitumor activities [56]. Contrary to the polysaccharides isolated by previous researchers [52, 53, 55], this polysaccharide only contained trace amounts of glucose, as it is mainly composed of mannose, xylose, arabinose, fucose, and rhamnose in a ratio of 31.27:1.77:1.44:1.34:1.00, with a backbone composed of repeating α-1,3-, α-1,6-, α-1,2-, and α-1,4-glycosidic linkages.

Recently, sulfated polysaccharides obtained from T. camphoratus grown in zinc sulfate enriched culture conditions were studied by Lee et al. [57]. It was found that the sulfated polysaccharides not only contain inflammation-inhibiting properties but also possess anti-lung cancer activities. In a comparison of sulfated polysaccharides to non-sulfated groups, sulfated derivatives showed more robust anti-inflammatory effects than non-sulfated polysaccharides, thus suggesting the role of sulfation in the inflammation-reducing capacities of the bioactive compound [58]. In another study, a 13.5 kDa heterogalactan was extracted from T. camphoratus and found to be composed of α-d-1,6-linked galactose backbone chain with terminal α-d-mannose and α-l-fucose every six galactose residues [59]. Similar to the results observed by Chen et al. [58], the mannofucogalactan itself did not demonstrate inhibitory effects on angiogenesis, while its sulfated derivative exerted a significant dose-dependent reduction of tube formation.

Besides sulfated derivatives, selenium-enriched polysaccharides have also been analyzed [39]. It was reported that in vitro scavenging efficiency was greatly enhanced following selenium enrichment, thus making selenium-enriched T. camphoratus polysaccharides a promising antioxidant agent for food and pharmaceutical applications.

It is noteworthy to mention that besides from the variety of bioactive components, alkali-extracted dietary fiber from the residues of T. camphoratus basswood cultured fruiting body also serves as a promising source of functional ingredients [60]. The study highlighted that the dietary fiber promoted RAW 264.7 cell proliferation, phagocytosis, nitric oxide (NO) production, and release of pro-inflammatory cytokines tumour necrosis factor alpha (TNF-α) and interleukin-1 beta (IL-1β), thus revealing its immunomodulatory properties, and possibly other bioactivities.

Terpenes and terpenoids

Evaluation of new ergostane-type triterpenoids, antcamphorols A-K, (1–11) revealed ROS scavenging abilities of antcamphorols G (7), I (9), and J (10), along with cytotoxic activities of antcamphorols C and H against U251 and MCF-7 [61]. Furthermore, through isolation from dish cultures of T. camphoratus, three new (12–14) and 10 known isolated triterpenoids (15–24) were demonstrated to possess cytotoxic activities on cancer cell lines HL60, U251, SW480, and MCF-7 [62]. Interestingly, it was found that the combination of two of the isolated triterpenoids (25, 26) could enhance the cytotoxic effect of paclitaxel, an established anti-cancer agent, suggesting triterpenoids as potential sensitizers of paclitaxel for chemotherapy. Antcamphin Z (27) also displayed potent cytotoxic activities against human cancer cells (U251, HL60, SW480, and A549) with no significant effects on normal cells (LO2) [63]. Besides their cytotoxic activities, triterpenoids antcins A (28) [64], B (29), H (30), and K (31) [65] have also been reported to exert an anti-inflammatory effect.

Using chiral-supercritical fluid chromatography (SFC), R/S epimeric pairs of ergostane triterpenoids, including antcin A (28) and antcin B (29), were separated from T. camphoratus at a higher efficiency than high-performance liquid chromatography (HPLC) [66]. This thus demonstrates chiral-SFC as a more advantageous method for separation of epimeric bioactive compounds. Mechanochemical-assisted extraction has also been used for the isolation of triterpenoids from the medicinal fungus, where extracts containing triterpenoids (32, 33) obtained through this technique led to more robust antioxidant, anti-inflammatory, and immunomodulatory activities than ethanol thermal reflux-extracted compounds [67].

Steroids (34–37) isolated from the medicinal mushroom have exhibited cytotoxic activities against murine colorectal and human leukemia cancer cell lines [68]. Furthermore, ergosterol (38) obtained from T. camphoratus was recently found to significantly reduce nuclear factor kappa B (NF-κB) phosphorylation, microglial activation–associated ionized calcium-binding adapter molecule-1 (IBA-1), and LPS-induced neuron damage, therefore making it an effective anti-neuroinflammatory agent [69].

Quinone derivatives

A novel quinone derivative, coenzyme Q0 (CoQ0, 2,3-dimethoxy-5-methyl-1,4-benzoquinone) (39), was isolated from fermented culture broth of T. camphoratus by Yang et al. [70]. This new compound was found to be capable of inducing apoptosis and autophagic cell death both in vitro in human glioblastoma cells and in vivo in glioblastoma-xenografted mice, through increased reactive oxygen species (ROS) production.

Another bioactive constituent found in the mycelium of T. camphoratus is antroquinonol (40), which is a ubiquinone derivative. Antroquinonol has been reported to inhibit the migration and invasion of breast cancer cells, through suppression of matrix metalloproteinase-9 (MMP-9) and epithelial-mesenchymal transition (EMT) gene expressions [71], while antroquinonol Y (41) also exerts significant cytotoxic effects on human cancer cells (U251, HL60, SW480, and A549) without affecting normal cells (LO2) [63]. Furthermore, bioassay-guided isolation from methanolic extracts of T. camphoratus revealed a potent anti-inflammatory compound, 4-acetylantroquinonol B (42), which significantly inhibited polyinosinic-polycytidylic acid-induced NO production by RAW264.7 macrophages [72]. Hence, due to the highly potent anti-tumor properties and demonstrated bioactivities of antroquinonol, many studies have been conducted to enhance its synthesis [73,74,75,76,77,78].

Maleic and succinic acid derivatives

Antrodins (43–47) are maleimide derivatives found in T. camphoratus mycelium and have been found to inhibit Hepatitis C virus (HCV) protease [79]. Antrodin A (43) has been reported to possess hepatoprotective properties and alleviate intestinal flora dysbiosis [80], while antrodin B (44) contains anti-hepatofibrotic activity [81]. Antrodin C (45) has been found to reduce the pathology of Alzheimer’s disease [82], protect against liver fibrosis [83], and exert cytotoxic activities against colorectal cancer cells [84].

Inonotus obliquus

Polysaccharides

Polysaccharides components of Inonotus obliquus have been studied by many researchers for their variety of biological activities. Two polysaccharides (HIOP1-S and HIOP2-S) were obtained from I. obliquus, where HIOP1-S was primarily composed of glucose (29.673%) and galactose (20.547%) and HIOP2-S mainly consisted of glucose (49.881%) [85]. Interestingly, HIOP1-S consisted of both α-and β-glycosidic bonds, while HIOP2-S only contained β-type linkages. Despite the considerable divergences between the two polysaccharides, both were capable of inhibiting α-glucosidase activity and enhancing extracellular glucose consumption in insulin-resistant cells. Another study also isolated two polysaccharides from the medicinal mushroom, IOEP1 and IOEP2, that contained both α-and β-glycosidic bonds [86]. The pyran-type polysaccharides were reported to mainly consist of galactose and mannose (IOEP1) and arabinose (IOEP2), and were both also demonstrated to increase glucose consumption of insulin-resistant cells, thus exerting a hypoglycemic effect.

A polysaccharide, composed of galactose, glucose, xylose, and mannose (2.0:3.5:1.0:1.5 molar ratio), was extracted from I. obliquus and found to exert free-radical scavenging activities, antioxidant effects, and enhancement of nitric oxide release [87]. A polysaccharide extracted by Hu et al. [88] also exhibited antioxidant properties, and consisted of mannose, rhamnose, glucose, galactose, xylose, and arabinose in a molar ratio of 9.81:3.6:29.1:20.5:21.6:5.4. The antioxidative properties of another I. obliquus polysaccharide with a molecular weight of 111.9 kDa, through the regulation of the nuclear factor erythroid 2-related factor 2 (Nrf2) pathway, were found to contribute a protective effect against Alzheimer’s disease [89].

Neutral (60–73 kDa), acidic (10–31 kDa) and alkaline (> 450 kDa) polysaccharide fractions isolated from I. obliquus extracts, consisting primarily of 1,3- and 1,6-linked-β-glucose, were reported to possess immune-promoting activities [90]. Chen et al. [91] successfully obtained a novel polysaccharide from I. obliquus (IP3a). It was found that IP3a is composed of arabinose, galactose, rhamnose, and glucose (4.6:2.6:2.5:1.0), and possesses pro-apoptotic and immune-enhancing effects. A pro-apoptotic effect was also observed in a polysaccharide containing mainly glucose (74.95%), where it was found that the apoptosis-inducing effects were mediated by activation of adenosine monophosphate-activated protein kinase (AMPK) and reduction of mitochondrial membrane potential (MMP) [92].

Recently, another polysaccharide was isolated and found to have a molecular weight of 373 kDa, with a composition of 9 monosaccharides, primarily contributed by glucose (29.85%), galactose (20.70%), and mannose (14.19%) [93]. The polysaccharide was found to be capable of increasing Firmicutes, decreasing Bacteroidetes, and repairing the intestinal barrier to contribute to the protection against Type 2 diabetes mellitus. Interestingly, contrary to the findings of Su et al. [93], a polysaccharide consisting of mannose, galactose, and glucose (7.7:23.3:32.6) reduced the abundance of Firmicutes and enhanced Bacteroidetes [94], thus suggesting the impact of polysaccharide composition on biological activities.

Terpenes and terpenoids

Inotodiol (48) and trametenolic acid (49) are common triterpenoids found in I. obliquus, and were found to exert anti-proliferative activities on AGS, MCF-7, and PC3 cancer cells [95], and alleviate liver diseases through modification of the farnesoid X receptor (FXR)/small heterodimer partner (SHP)/ sterol-regulatory element binding protein-1C (SREBP-1C) axis [96]. Besides that, betulin (50) and betulinic acid (51) also inhibited the proliferation of human adenocarcinoma and human lung carcinoma cell lines, while inotodiol (48), 3β-hydroxy-8,24-dien-21-al (52), and betulin-3-O-caffeate (53) possessed immunomodulatory capabilities [97]. In addition, Zou et al. [98] highlighted the neuroprotective effects of triterpenoids lanosterol (54), inotodiol (48), and inonotsutriol A (55), while Park et al. [99] demonstrated the NO-inhibiting effects of inotodiol (48), inonotsutriol A (55), and trametenolic acid (49).

Another bioactive compound, known as ergosterol peroxide (56), was found to exert anti-cancer properties against colorectal cancer [100], prostatic carcinoma and breast carcinoma cells [101]. Ergosterol peroxide (56), ergosterol (38), and trametenolic acid (49) have also been demonstrated to have anti-inflammatory capabilities [101]. Moreover, a wide variety of I. obliquus-obtained triterpenoids (57–66) were observed to have inhibitory effects on α-glucosidase [102].

Besides known compounds, new triterpenes and triterpenoids have been elucidated and found to also contain biological activities. Two new triterpenes, inonotusane D (67) and inonotusol G (68), isolated by Zhao et al. [103] and Liu et al. [104] respectively, exhibited cytotoxic effects against cancer cells. Besides that, another new lanostane-type triterpene, inonotusane C (69), also exhibited cytotoxic activities against A549 and HeLa cell lines [105] thus demonstrating their chemotherapeutic potentials.

A total of 12 previously uncharacterized lanostane triterpenoids (70–81) were isolated by Zou et al. [98, 106], where it was revealed that compound 77 is capable of exerting neuroprotective activities. Another study also obtained novel lanostane-type triterpenoids, inonotusols H-N (82–88), that were found to significantly inhibit NO release by microglial cells, where inonotusols I (83) and L (86) were capable of interacting with inducible-nitric oxide synthase (iNOS) protein to hence exert a neuroprotective effect [107].

Phenolic compounds

Phenolic compounds, including phenolic acids and flavonoids, are produced by the sclerotia of I. obliquus and have been found to possess potent bioactivities. A phenolic compound isolated from I. obliquus, identified as inoscavin A (89), was demonstrated to improve cell viabilities of H2O2)-injured human neuroblastoma cells to exert a neuroprotective effect through their antioxidant activities [108]. Phenolic (91–97) and polyphenol compounds (98–107), isolated by Hwang et al. [109] and Wang et al. [110] respectively were reported to contain radical scavenging activities in a dose-dependent pattern.

Melanins are another type of phenolic compound found in I. obliquus and have been associated with immunomodulatory properties. A water-soluble melanin fraction was found to be capable of inhibiting the complement cascade and suppressing NO production by murine macrophages [97]. Melanins isolated from aqueous extracts of I. obliquus also exhibited antioxidant activities, thus creating an optimum growth condition for obligate anaerobe Bifidobacterium bifidum [111].

A recent study revealed that ultraviolet (UV) radiation led to an increased accumulation of both extracellular and intracellular polyphenols and flavonoids, which led to the UV-irradiated extracts’ enhanced capabilities to scavenge free radicals [112]. Another study was also able to increase exo- and endo-polyphenols production through application of stimulatory agents, where only 1.0 g/L of the most effective agent, linoleic acid, was found to significantly improve the synthesis of ferulic acid (108), gallic acid (104), epicatechin-3-gallate (ECG) (109), epigallocatechin-3-gallate (EGCG) (110), phelligridin G (111), inoscavin B (90), and davallialactone (112) to give rise to more robust scavenging activities of the treated extracts [113]. Tween-20 was also proven to be a potent stimulatory agent for the increase in polyphenols ferulic acid (108), naringin (113), ECG (109), EGCG (110) content, which subsequently led to enhanced antioxidant activities of treated products [114].

Tropicoporus linteus

Polysaccharides

A water-soluble heteropolysaccharide was isolated from the mycelia of Tropicoporus linteus, with a composition of arabinose, xylose, glucose, and galactose in the molar ratio of 4.0:6.7:1.3:1.0 and a molecular weight of 343,000 kDa, whereby the backbone consisted of α-1,2-, α-1,4, β-1,4-, and β-1,5-glycosydic bonds [115]. The polysaccharide was capable of regulating the mitogen-activated protein kinases (MAPK) and peroxisome proliferator-activated receptor (PPAR) signalling pathways to exert an anti-inflammatory effect. Another high molecular weight polysaccharide (123.8 kDa) that is rich in galacturonic acid (35.7% of monosaccharide composition) with small amounts of arabinose, galactose, xylose, rhamnose, and fucose showed immunostimulatory activities [116].

A lower molecular weight (1,3;1,6)-β-D-polysaccharide (20.7 kDa) consisting mainly of glucose (78.88%) mannose (8.32%) and galactose (8.06%) was also obtained from T. linteus [117]. It was reported to inhibit TNF-α and IL-6 release in RAW264.7 cells, while increasing IL-10 levels, thus suggesting a possible role in the restoration of the IL6/IL10 balance. Another polysaccharide with a molecular weight of 15.5 kDa is also mainly made up of glucose (53.2%), mannose (14.9%), and galactose (13.5%) with 5.5% composition of (1,3;1,6)-β-d-glucans, where it was found to be capable of inhibiting NO production in RAW264.7 cells [118].

Three antioxidant polysaccharides, PL-W, PL-A, and PL-N, were extracted from the mycelia of T. linteus, where PL-W was composed of glucose and mannose (molar ratio 8:1), PL-A composed of glucose, mannose, xylose, and arabinose (molar ratio 8:1:1:1), while PL-N consists of xylose, arabinose, glucose, and galactose (7.8:5.5:1.8:1.0) [119]. All three polysaccharides were soluble in water at 10 g/L and possess potent antioxidant capacities in a concentration-dependent manner. This is in agreement with the findings of Yan et al. [120], where PL-N was found to have antioxidant activities, which can be enhanced by ultrasonic treatment. Besides that, two novel heteropolysaccharides, PLP1-I and PL-A11, with molecular weights of 290,000 kDa and 13.8 kDa respectively, were obtained from T. linteus mycelia [121, 122]. PLP1-I consists of only glucose and galactose (molar ratio 8.9:1.0) with a 1,4-α-d-glucopyranose backbone, while PL-A11 is composed of arabinose, xylose, mannose, and glucose (molar ratio 1.1:1.3:1.0:6.6) with a 1,4-α-d-glucopyranosyl, 1,2-α-d-xylopyranosyl, and 1,3-α-d-arabinofuranosyl backbone. Nevertheless, both polysaccharides were found to increase antioxidant enzyme activities in a dose-dependent manner. Another two polysaccharides with antioxidative activities, PLPS and C-PLPS, was recently isolated, where both contained arabinose, xylose, mannose, glucose, and galactose in molar ratios of 1.0:1.8:3.8:40.1:1.4 and 1.0:1.5:3.4:25.2:1.1 respectively [123].

A branched-type glycan with an average molecular weight of 3172.9 kDa, pyranoid sugar ring conformation, and α- and β-linkages was found to be capable of exerting an anti-diabetic effect by reducing blood glucose levels in diabetic mice [124]. This agrees with the findings of Liu et al. [125] where two T. linteus polysaccharides with backbones consisting of repeating α-d-glucose(1,4)-α-d-glucose(1,6) units (PLPS-1) and 1,3-α-d-glucose and 1,6-α-d-glucose (PLPS-2) respectively were observed to ameliorate insulin resistance.

A thorough evaluation of PLPS-1 and PLPS-2 previously conducted by Mei et al. [126] revealed that PLPS-1 consisted of glucose, arabinose, fucose, xylose in a molar ratio of 21.964:1.336:1.182:1:1, while PLPS-2 consisted of glucose, galactose, mannose, arabinose, fucose, xylose in a molar ratio of 14.368:2.594:1.956:1.552:1.466:1, where PLPS-1 was found to possess antitumor activities against S180 sarcoma cells. Another 343,000 kDa polysaccharide was reported to exert antitumor activities against HepG2 cells [127] and a hepatoprotective effect in mice [128], where structural characterization revealed the presence of seven glycosidic residues Araf (1 → , → 5) Araf (1 → , → 4) Xylp (1 → , → 2) Xylp (1 → , → 2,4) Xylp (1 → , → 4) Glcp (1 → , and → 4,6) Galp (1 → in the molar ratio of 2.28:1.83:3.17:2.64:1.03:1.36:1.00.

A polysaccharide composed of glucose, galactose, mannose, fucose, and xylose in a molar ratio of 4.36: 2.34: 2.09: 0.88: 0.42 showed prebiotic potentials by increasing amounts of beneficial bacteria, such as Bacteroides, Prevotella, and Butyricimonas, and reducing pathogenic bacteria including Escherichia, Shigella, Morganella, and Intestinimonas [129]. Besides that, a polysaccharide with a monosaccharide composition of fucose, rhamnose, arabinose, glucuronic acid, galactose, glucose, and xylose (1.4:0.5:0.9:1.6:4.7:84.8:6.0) exerted bacterio-static activities against bacteria Staphylococcus aureus, Escherichia coli, and Bacillus subtilis, demonstrating the potential of T. linteus polysaccharides for clinical applications [130].

Terpenes and terpenoids

Phellilane L (114) is a cyclopropane-containing sesquiterpenoid isolated from T. linteus that was found to possess antimicrobial activities against P. gingivalis [131]. A γ-Ionylidene-type sesquiterpenoid obtained from T. linteus, (-)-trans-γ-monocyclofarnesol (115), also exhibited antimicrobial effects on P. gingivalis [132], while phellidene E (116), ( +)-γ-ionylideneacetic acid (117) [132] and phellilane H (118) [133] did not show any detectable activities. Besides that, phellinulins A-N (119–132) were extracted from the mycelium of T. linteus, where phellinulins A (119), H (126), I (127), K (129), and M (131) were found to have hepatoprotective properties [134].

Phenolic compounds

Phellifuropyranone A (133) and phelligridin G (111) can be found in T. linteus, and possess in vitro antiproliferative activity against mouse melanoma and human lung cancer cells [135]. Several other polyphenols, including hispidin (134), phelligridimer A (135), davallialactone (112), methyldavallialactone (136), hypholomine B (137), interfungin A (138), inoscavin A (139), protocatechuic acid (97), protocatechualdehyde (95), caffeic acid (99), and ellagic acid (140) were also isolated by Lee et al. [136], where davallialactone (112), hypholomine B (137), and ellagic acid (140) were capable of inhibiting aldose reductase to reduce diabetic complications. In addition, davallialactone (112), hispidin (134), hypholomine B (137), and caffeic acid (99) were found to have antioxidant activities [137], while hispidin (134) and hypholomine B (137) also exerts hypolipidemic and hepatoprotective effects [138].

In agreement with Min et al. [137] and Chiu et al. [138], hispidin (134) have been demonstrated to be a potent antioxidant by various studies, thus providing protection against oxidative stress [139,140,141,142]. Due to the significant bioactivities and potential applications of hispidin, methods have been developed to enhance the production of hispidin, such as through solid-state fermentation using pearl barley medium [143].

Another notable polyphenol that has been detected in T. linteus is hispolon (141) [144]. Hispolon has been reported to have various pharmacological activities, such as anti-inflammatory properties, [145,146,147,148], antioxidant activities [149], and antitumor effects [150,151,152,153].

A catechol-containing phenylpropanoid derivative found in T. linteus, 3,4-dihydroxybenzalacetone (Osmundacetone) (93), has been shown to attenuate inflammation [154], exert antioxidant activities [155], suppress the metastasis and formation of cancer [156], and protect against aging-associated myocardial alterations [157]. Osmundacetone (93) and inotilone (142), have also been isolated from T. linteus, where they have been demonstrated to have antioxidant properties [158] and inhibit neuraminidase activity to thus exert an antiviral effect [13].

Health-promoting properties

Taiwanofungus camphoratus

Anticancer

T. camphoratus possesses a wide range of health-benefiting properties, which has been summarized in Table 2, whereby its anticancer properties are probably the most extensively studied aspect (Fig. 3). The ethanolic extract of T. camphoratus have been shown to exert anticancer effects on SMMC-7721 and HepG2 cells through inhibiting the activation of signal transducer and activator of transcription 3 (STAT3) [159], and B16-F0 cells via inducing cell cycle arrest and apoptosis [160]. Additionally, T. camphoratus ethanolic extracts also induced cell cycle arrest and suppressed the growth of Hep 3B and Hep J5 cells, where caspase-3 and cell cycle inhibitors p21 and p27 were also activated in the cells [161]. Three ergostane-type triterpenes isolated from the fruiting bodies of T. camphoratus caused a significant increase of HT-29 (Colon cancer) cells in the sub-G1 phase from the control value of 0.77% to about 41%, 32% and 29% respectively (Compound 1, 2 and 3), strongly indicating the ability of these triterpenes to induce apoptosis in HT-29 cells [162].

Table 2 Health-promoting properties of Taiwanofungus camphoratus extracts and bioactive compounds

Furthermore, triple-negative breast cancer (MDA-MB-231) cells were also responsive to the anticancer effects of the fermented culture of T. camphoratus, through an inhibition of the EMT, further suggesting the potential of T. camphoratus as an anticancer agent [163]. Moreover, Yang et al. [164] reported that the fermented culture of T. camphoratus also produced inhibitory effects against SKOV-3 human ovarian carcinoma cells via suppression of the HER-2/neu signalling pathway. Besides, the fermented culture of this mushroom extract also exerted a dose- and time-dependent series of apoptotic events in MCF-7 cells, as evidenced by the accumulation of cells in the sub-G1 phase, chromatin condensation, internucleosomal DNA fragmentation, and loss of cell viability [165].

Fig. 3
figure 3

Mechanisms found to be employed by Taiwanofungus camphoratus to exert anticancer effects

The crude extract of the fruiting body of cultivated T. camphoratus induced cytotoxicity on HeLa and C-33A human cervical cancer cell lines, with higher potency recorded on C-33A in comparison to HeLa cells [166]. According to the results of the caspase activity assay, the crude extract significantly increased the activities of caspase-3, -8, and -9 in HeLa and C-33A cells and led to a dose-dependent rise in cytosolic cytochrome in the cytochrome c assay, thus indicating that apoptosis involves both intrinsic and extrinsic signalling pathways [166]. Similar effects were also exerted by the crude extract on SKOV-3 and TOV-21G ovarian cancer cells where the activity of caspases-3, -8, -9 and cytochrome c were found to be increased [167]. Bad protein from the Bcl-2 family was observed to be increased in SKOV-3 cells while the expression of Bim and Bak proteins increased and Bcl-xL decreased in TOV-21G cells. Moreover, T24 bladder cancer cells were also found to be susceptible to the crude extract of T. camphoratus with a concentration of about 50 µg/mL with about half of the cells in the control being inhibited at the 72 h time-point [168].

The ethyl acetate extract of T. camphoratus was reported to be effective in the liver cancer cell line Hep 3B via the increase of cytoplasmic calcium (Ca2+) and subsequent induction of calpain and caspase-12 activation [169], as well as HepG2 and PLC/PRF/5 cells, as evidenced by the marked increase in apoptotic promoting factors [170]. In addition, Hep3B and HepG2 cells were also susceptible to the methanolic extract of T. camphoratus, where treatment for 72 h resulted in apoptosis of around 40% of the Hep3B cells and 98% of the HepG2 [171]. Furthermore, Rao et al. [172] reported that methanol extract also inhibited Jurkat leukaemia cells with IC50 value of approximately 40 µg/mL.

The solid-state cultured broth of T. camphoratus also managed to exert synergistic antiproliferative effects in C3A and PLC/PRF/5 hepatoma cells when administered in combination with cisplatin and mitomycin via the inhibition of the multidrug resistance (MDR) gene expression and the cyclooxygenase-2 (COX-2)-dependent inhibition of phospho-protein kinase B (p-Akt), which in turn resulted hepatoma cells to undergo apoptosis [173].

Hypoglycemic and hypolipidemic

T. camphoratus is also known for its antiglycemic effects, where antcin K from the fruiting body of T. camphoratus can enhance membrane glucose transporter 4 (GLUT4) and p-Akt expressions in vitro in C2C12 myotube cells and induce an antihyperglycemic effect (40 mg/kg/day) comparable to Metformin (300 mg/kg/day) in vivo in C57BL/6J mice [174]. Another compound, eburicoic acid TRR, from T. camphoratus was able to increase AMPK phosphorylation, Akt, and membrane GLUT4 level in skeletal muscles, suggesting its capability to produce antihyperglycemic effects via an insulin-dependent pathway or possibly through activation of AMPK in skeletal muscles [175]. Furthermore, the administration of sulphurenic acid obtained from T. camphoratus resulted in the rise of insulin levels in streptozotocin-induced C57BL/6J diabetic mice, possibly due to an increase in pancreatic Langerhans islets’ size and number upon treatment and caused a notable reduction in the concentration of glycated hemoglobin (HbA1c) [176]. It was noted that sulphurenic acid stimulates the phosphoinositide 3-kinase (PI3K)/Akt pathway, which suppresses hepatic glucose production to result in a hypoglycemic effect. Kuo et al. [177] also reported that dehydroeburicoic acid from T. camphoratus increased the expression of GLUT4 in C2C12 myotube cells and also resulted in 43.6–46.5% decrease in blood glucose level in Streptozotocin-induced C57BL/6J diabetic mice.

Antcin K from T. camphoratus has also been found to exert a hypolipidemic effect by decreasing the expression of fatty acid synthase levels and increasing the hepatic expression levels of carnitine palmitoyltransferase 1A (CPT-1A) and PPAR-α [174, 176]. Eburicoic acid and sulphurenic acid derived from T. camphoratus also increased the expression of PPAR-α in the liver of high-fat diet (HFD)-induced C57BL/6J mice, thus facilitating fatty acid oxidation and reduction of lipid levels [175, 176]. In addition, Kuo et al. [178] reported that Ergostatrien-3β-ol from T. camphoratus reduced visceral adipocyte size and induced hepatic ballooning degeneration. All four above-mentioned compounds were also found to SREBP-1C to downregulate transcription of genes responsible for fatty acid synthesis, suggesting the potent hypolipidemic effects of T. camphoratus [174, 175, 178, 191].

Antioxidant

Ethanolic extract of T. camphoratus fruiting body displayed strong DPPH radical and superoxide dismutase (SOD)-like scavenging activities at levels nearly equivalent and stronger than vitamin C respectively [9]. In DPPH assay, the activity at 4 mg/mL of the extract was around 83%, almost as high as that of vitamin C at 0.5 mg/mL, which demonstrated around 88% activity. In contrast, the extract’s activity at 4 mg/mL in the SOD scavenging assay was around 71–87%, higher than that of vitamin C at 0.5 mg/mL, which demonstrated around 67% activity. The potent antioxidant activities of T. camphoratus were further highlighted in another study where T. camphoratus crude oil and polysaccharide had half maximal effective concentration (EC50) values of 47 µg/mL and 22 µg/mL respectively in DPPH assay, significantly than that of tert-Butylhydroquinone and ascorbic acid at 11.2 µg/mL and 6 µg/mL [179]. Besides that, T. camphoratus mycelium extract recorded strong and dose-dependent scavenging activity, which can be attributed to its high phenolic content, observed to be at 20 mg/g [14]. Furthermore, Li et al. [35] reported that the polysaccharides isolated from T. camphoratus which were enriched with selenium recorded almost double the DPPH and ABTS radical scavenging activity in comparison to the unenriched polysaccharides, thus suggesting the possibility of raising the antioxidant properties of T. camphoratus through modifications of isolated bioactive compounds.

Hepatoprotective

Kumar et al. [180] reported that treatment with antroquinonol and mycelia of Golden-T. camphoratus displayed effective protection of the liver against alcohol-induced damage both in vitro and in vivo, whereby both antroquinonol and mycelia pretreatment considerably and dose-dependently decreased the ethanol-associated rise in alanine aminotransferase (ALT) and aspartate transaminase (AST) after 24 h in oxidative stress-induced HepG2 cells [180]. Furthermore, Chen et al. [181] revealed the potential of antroquinonol from Golden-T. camphoratus to inhibit hepatic lipid accumulation and release of inflammatory cytokines through the upregulation of hepatic adenosine diphosphate (ADP)-activated protein kinase and concurrent downregulation of SREBP1 expression and fatty acid synthase expression.

Hepatoprotective activities have also been found in antrosterol from T. camphoratus, whereby pre-treatment of chronic alcohol liver disease-induced mice could reduce serum/hepatic lipids, increase fecal lipid/bile-acid output, amplify hepatic antioxidant activities, reduce serum alcohol level, and decrease liver inflammation, fibrosis, serum AST/ALT, and TNF-α/IL-1β [182]. These combined findings provide valuable insights into the multifaceted mechanisms underpinning the hepatoprotective role of compounds isolated from T. camphoratus.

Anti-inflammatory

The ethanol extract of T. camphoratus grown on germinated brown rice was reported to inhibit NO, prostaglandin E2 (PGE2), iNOS and COX2 production while inducing p38-MAPK, extracellular signal-related kinases (ERK) and NF-κB in LPS-stimulated RAW264.7 cells [183]. The same study also reported similar observations for the in vivo analysis in addition to the reduction of TNF-α and IL-6 mRNA expression levels in treated dextran sulfate sodium (DSS)-induced mice. Similar findings were also reported by Kuo et al. [184], in which significant inhibition of MMP-1, IL-6, iNOS, and NF-κB were observed in the skin of ultraviolet B-induced mice for topical administration of 25 and 100 μM ergostatrien-3β-ol isolated from T. camphoratus.

Besides, Tsai et al. [185] reported a decrease in gene expressions of iNOS, IL-6, TNF-α and NF-κB along with significantly lowered necrosis and inflammatory cell infiltration (both sub-dermis and epidermis) in skin-flap ischemia damage-induced mice treated with the methanol extract of the wood-cultured fruiting body and solid state-cultured mycelia of T. camphoratus. In addition, acetylantroquinonol B [186] and benzoids [187] were also found to inhibit the NO production in inflammation-induced RAW 264.7 cells, further demonstrating the anti-inflammatory capabilities of T. camphoratus.

Antiviral

A limited amount of research on the antiviral effects of T. camphoratus has been conducted over the past decade. According to He et al. [11], a fraction of the crude ethanol extract of T. camphoratus exhibited notable antiviral activity against the herpes simplex virus type 1 (HSV-1) and herpes simplex virus type 2 (HSV-2), significantly inhibiting the viral replication between 1–4 h upon infection of Vero cells. However, antiviral activity reduced with the addition of the extract after 5 h of infection, suggesting low inhibition of late multiplication of virus [11]. In addition, Chen et al. [188] reported that the pre-treatment of two ethanolic extracts of T. camphoratus inhibited the replication of Dengue virus in Meg-01 cells by increasing the expression of antiviral cytokine interferon alpha (IFN-α) and decreasing the release of pro-inflammatory cytokines IL-6 and IL-8.

Polysaccharides from T. camphoratus were also found to possess anti-Hepatitis B virus (HBV) activity, where polysaccharides from B86 and 35,398 strains most significantly inhibited hepatitis B e-antigen (HBeAg) synthesis, with about 30% antiHBeAg inhibition percentage for both strains at a concentration of 50 µg/mL [189]. Apart from HBV, antrodins A to E, except antrodin B, isolated from T. camphoratus exhibited potent inhibitory effects against the HCV protease, whereby antrodin A had the highest potency with a half-maximal inhibitory concentration (IC50) value of 0.9 µg/mL [190].

Inonotus obliquus

Anticancer

I. obliquus is capable of exerting a wide range of therapeutic properties, as summarized in Table 3, which includes its anticancer activities. I. obliquus has been used in folk medicine for the treatment of cancer [4], including its use to heal lip tumours [192]. In the twentieth century, the remarkable ability of I. obliquus to reduce the number of cancer cases in the population was noticed by the Soviet health authorities [40], leading to increasing amounts of scientific research to be conducted on this medicinal mushroom (Table 3).

Table 3 Health-promoting properties of Inonotus obliquus extracts and bioactive compounds

Aqueous extracts obtained from the fermentation of I. obliquus displayed a dose-dependent inhibition in the growth of HCT-15 cells via the induction of apoptosis [194] and a 51% reduction in B16-F10 melanoma cell proliferation after a 48-h exposure to 750 µg/mL of extract [195]. Further analysis revealed that I. obliquus extract may restrict cell proliferation or stimulate differentiation in B16-F10 cells by causing G0/G1 phase arrest and subsequent apoptosis process that occurs when caspase-3 is activated [195]. Another study using aqueous extracts of I. obliquus exhibited a dose-dependent inhibition in growth of HT-29 cells via induction of apoptosis, where the highest inhibition of 56% was recorded at the concentration of 1.0 mg/mL at the 48 h time-point. [196]. The anticancer properties of the aqueous extract of I. obliquus was also effective in vivo, where a consistent intake of the aqueous extract managed to decrease 25% of the metastatic tumour nodules and reduce 60.3% (Day 16) of tumour growth in 3LL carcinoma-implanted C57BL/6 mice as compared to the control group [6]. Furthermore, Yuan et al. [197] reported that 7 mg/mL of aqueous extract of the fungi exhibited about 33% inhibition in tumour growth in DLD-1 colorectal adenocarcinoma inoculated mice for 36 days.

Apart from aqueous extracts, the ethanol extracts of I. obliquus also exhibit promising anticancer effects. According to Lee et al. [198], the extracts stimulated cell cycle arrest at G1 phase, induced higher levels of p21 and p27, resulted in the inhibition of cyclin-dependent kinase (CDK), and consequently reduced retinoblastoma tumour suppressor protein (Rb) phosphorylation. The advancement of the G1 into S cell cycle was halted as a result of these effects, followed by a suppression of HT-29 cell proliferation. Ethanol extracts obtained from wild sclerotium, cultivated sclerotium, and cultivated fruiting body of I. obliquus demonstrated anticancer effects with inhibition rates ranging from 44.2 to 74.6% [199]. Moreover, Hu et al. [200] reported that DLD- 1 cells are also susceptible to the ethanol and hot water extracts of I. obliquus via the induction of apoptosis and reducing oxidative stress. In addition, in a study conducted by Nakajima et al. [201], methanol extracts from both the fruiting body and sclerotium of I. obliquus exhibited anticancer effects on IMR-90, A549, PA-1, U937, and HL-60 cells, whereby PA-1 cells were the most susceptible to both types of methanol extracts, with the methanol extract from sclerotium displaying more potent cytotoxic effects to all tested cancer cell lines.

Other constituents isolated from I. obliquus have also demonstrated promising anticancer properties. Polysaccharides isolated from I. obliquus were also found to inhibit tumour growth in SGC-7901 tumour-bearing mice [202] and exert anticancer effects in LLC1 cells in both time and dose-dependent manners [92]. Moreover, a triterpenoid named 3β-hydroxylanosta-8,24-dien-21-al induced cytotoxic effects on A549, H1264, H1299, and Calu-6 lung cancer cells with IC50 values of 5.3 µg/mL, 90.9 µg/mL, 128.0 µg/mL and 75.1 µg/mL respectively [203]. These findings collectively highlight the potent anticancer effects of a wide variety of compounds derived from I. obliquus.

Hypoglycemic and hypolipidemic

In addition to its anticancer effects, I. obliquus is also known for its hypoglycemic and hypolipidemic activities. Wang et al. [193] showed that oral administration of I. obliquus polysaccharides (IOPS) (900 mg/kg) could significantly improve glucose tolerance, reduce fasting blood glucose levels, increase hepatic glycogen levels, and alleviate insulin resistance in streptozotocin induced type 2 diabetes mellitus mice. It was also found that the IOPS increased expressions of PI3K-p85, p-Akt (ser473), and GLUT4 to therefore promote glycolysis and inhibit gluconeogenesis [193]. Furthermore, methanol extracts of I. obliquus can significantly improve blood glucose and lipid levels, as well as exert an anti-inflammatory effect in type 2 diabetic mice through downregulation of harmful groups such as Proteobacteria, upregulating levels of beneficial bacteria such as Odoribacter, and promoting short-chain fatty acid levels by increasing abundance of acid-producing bacteria such as Alistipes [204]. Moreover, Lee and Hyun [205] reported that water-soluble melanin complex isolated from I. obliquus were found to increase the insulin-stimulated glucose uptake and exert a dose-dependent increase in Akt phosphorylation and GLUT4 translation into the cytoplasm. In addition, another study by Cha et al. [206] mentioned that a diet supplemented with 50 g/kg fermented powder from I. obliquus was observed to significantly (p < 0.05) lower the glucose level in 100 mL of serum of streptozotocin-induced type 1 diabetes mellitus mice in comparison with supplementation with non-fermented chaga powder and without any supplementation.

According to Yang et al. [207], the administration of I. obliquus polysaccharides was able to decrease the total cholesterol (TC), triacylglycerol (TG), and low-density lipoprotein (LDL) levels, and consequently increased high-density lipoprotein (HDL) levels through the activation of AMPK and induction of fatty acid oxidation. This is consistent with the results obtained from a study conducted by Ye et al. [204] in which mice treated with I. obliquus methanol extract exhibited decreased LDL cholesterol (LDL-C), TC and TG levels, and higher HDL cholesterol (HDL-C) levels (p < 0.01). The polysaccharides treatment could also induce a significant reduction in the mRNA expression of SREBP-1, acetyl-CoA carboxylase (ACC), and fatty acid synthase to further contribute to its hypolipidemic activity [207]. Hypolipidemic effects were also observed in the 250 mg/kg and 500 mg/kg I. obliquus aqueous extract treatment groups, with decreased TC, TG, and LDL-C and increased liver glycogen and HDL-C in high-fat diet-inducedC57BL/6 mice via the regulation of the PI3K/Akt and AMPK/ACC signalling pathways [208]. Similar hypolipidemic effects were also reported in oligosaccharides of I. obliquus via modification of the intestinal bacteria in high-fat fed mice in which the ratio between the population of Firmicutes to Bacteroidetes faecal flora was reduced in treated groups, reducing the production of fatty acids and thus lowering blood lipids [209].

Antioxidant

Two novel isocoumarins found in I. obliquus exhibited significant antioxidant activity in FRAP, DPPH and ABTS assays as compared to Trolox, a potent antioxidative [15]. In addition, the sclerotium and mycelium extracts of I. obliquus exhibited protection from cellular DNA damage from H2O2 in healthy human lymphocytes at a similar intensity as natural antioxidants [210]. Another study by Cui et al. [211] found strong antioxidant activities in both the polyphenolic extract and another extract containing triterpenoids and steroids, with more significant activity in the former, suggesting the crucial contribution of phenols to the mushroom’s antioxidative properties.

Five polysaccharides isolated from I. obliquus also displayed antioxidant activities, in which stronger antioxidant activities were observed in the ones with higher uronic acid and proteinous substances content [212]. Furthermore, in a study by Hu et al. [213], it was found that 5 mg/mL of IOPS had a scavenging activity of 82.3% in DPPH assay and 81.3% in hydroxyl radical scavenging assay. Additionally, it also elevated SOD levels and reduced malondialdehyde (MDA) levels in a dose-dependent manner, thus bringing levels of oxidative stress close to normal levels in chronic pancreatitis-induced mice, highlighting the potential therapeutic application of I. obliquus’ antioxidative functions.

Hepatoprotective

A study conducted by Xu et al. [214] investigated the protective effects of polysaccharides isolated from I. obliquus in liver damage induced by Toxoplasma gondii infection in mice, reported a decrease in ALT, AST, MDA and NO levels, with an increase in SOD and glutathione (GSH). Besides, Hong et al. [215] reported that the aqueous extract of I. obliquus resulted in significant protection to tert-butyl hydroperoxide liver damage-induced rat hepatocytes even at a low concentration of 10 µg/mL. Moreover, d-galactosamine intoxicated-normal Chang liver cells experienced 2–2.5 times increase in viability upon treatment with 10−5 and 10−3 g/ L of melanin isolated from aqueous extract of I. obliquus [216]. Consistent findings were also obtained in vivo where decrease in fat buildup, steatosis, necrosis, and normalization of serum cholinesterase, gammaglutamyl transpeptidase, total protein and unconjugated bilirubin levels were observed in carbon tetrachloride liver damage induced-Sprague Dawley rats (Fig. 4).

Fig. 4
figure 4

Hepatoprotective properties possessed by Inonotus obliquus

Anti-inflammatory

The extract from I. obliquus grown on germinated brown rice was found to inhibit proinflammatory mediators NO, PGE2, iNOS, COX-2, TNF-α, IL-1β, and IL-6 in LPS-stimulated RAW 264.7 macrophages [217]. Besides, Ma et al. [101] reported that the petroleum ether extract of I. obliquus resulted in approximately 84.6% and 96.9% reductions in NO production and NF-κB luciferase activity respectively in LPS-stimulated RAW 264.7 macrophages, while the ethyl acetate extract stimulated 78.2% and 96.6% reductions in NO production and NF-κB luciferase activity accordingly [101]. Moreover, more than 90% suppression in histamine-induced TNF-α was discovered in histamine inflammation-induced-C57BL/6 mice upon treatment with the methanol extract of I. obliquus, thus further highlighting the anti-inflammatory properties of I. obliquus [218].

Antiviral

Shibnev et al. [219] reported that the infective property of hepatitis C virus was significantly reduced by I. obliquus water extracts, which exhibited a high antiviral effect with low cytotoxicity on porcine embryo kidney cells (SPEV), thereby shielding all of SPEV cells post-infection from virus-induced death at a level of 0.1 TCD50/cell. I. obliquus water extracts were also able to inhibit the replication of SARS-CoV-2 (nCoV/Victoria /1/2020 strain) in Vero E6 and Vero cell cultures [220], which may be explained by its significant binding affinity with the viral S1-carboxy-terminal of the SARS-CoV-2 receptor-binding domain, producing multivalent hydrogen and non-polar interactions (7.4 to 8.6 kcal/mol) [221].

Aqueous and water-alcohol extracts of I. obliquus have also been shown to possess antiviral effects against the human immunodeficiency virus type 1 (HIV-1) at a concentration of 5.0 µg/mL against virus-infected lymphoblastoid cells culture MT-4 [12]. Furthermore, I. obliquus alcohol-water pilat extract demonstrated significant in vitro anti-HSV-1 activity at concentrations ranging from 4.5 × 10–5 to 7.5 × 100 mg/ml [222]. The observed antiviral property could also be observed in vivo, whereby 90% of HSV-2 infected albino mice survived through prior intraperitoneal administration of 0.4 to 2 mg of I. obliquus aqueous extract [223].

In addition, according to Tian et al. [224], IOPS was found to exhibit antiviral effect against a large spectrum of feline viruses, which include the feline influenza virus (FIV), feline calicivirus (FCV), feline panleukopenia virus (FPV), feline herpesvirus 1 (FHV-1) and feline infectious peritonitis virus (FIPV) by interfering with the virions and/or cell receptors, thus preventing the viral entrance into the cell. This suggests the strong antiviral properties of not only crude extracts obtained from I. obliquus, but also isolated components such as polysaccharides as demonstrated above.

Tropicoporus linteus

Anticancer

The anticancer properties of Tropicoporus linteus and their underlying mechanisms have been widely studied in recent years (Fig. 5). A comparison between ethanol and water extracts of T. linteus, P. igniarius, and P. nigricans showed that the ethanol extract of T. linteus exerts strong cytotoxic effects against cholangiocarcinoma cells, with percent cell inhibition reaching as high as 98.92 ± 0.22 [225]. Besides, the ethanol extract of this mushroom exerted a significant dose-dependent inhibition in MDA-MB-231 cells, where an IC50 value of 40 mg/mL was obtained at the 24 h time point [226]. A notable observation is that 30 mg/mL of this extract alone exerted 16.7% inhibition while 10 µg/mL of 5-fluorouracil alone exerted 12.6% inhibition in cell growth, but this value spiked significantly to 53% when both of them were administered together (Table 4).

Table 4 Health-promoting properties of Tropicoporus linteus extracts and bioactive compounds

Park et al. [227] reported that the ethyl acetate extract of T. linteus grown in germinated brown rice induced apoptosis in HT-29 cells via the induction of cell cycle arrest at the G0/G1 phase and DNA fragmentation. HT-29 cells were also susceptible to the n-hexane layer of ethyl acetate extract from T. linteus which showcased potent anticancer effects with an IC50 of 69.8 µg/mL at the 48 h time point [228]. Moreover, the methanol extract of T. linteus also displayed anticancer effects in SW-480 and HCT-116 cells where an increase in anti-migratory marker E-cadherin and a decrease in pro-migratory/pro-invasive proteins N-cadherin and Vimentin were observed with higher selectivity towards SW-480 cells [229]. Besides, extracts of T. linteus grown on germinated brown rice could increase the sensitivity of KRAS-mutated SW480 colon cancer cells to cetuximab with a reduced number of colonies in clonogenic assay and increased apoptotic rate [230]. Recent research has also revealed that T. linteus possesses potent anti-breast cancer proliferative activity by downregulating marker of proliferation Kiel 67 (MKI67), Human epidermal growth factor receptor 2 (HER2), epidermal growth factor receptor (EGFR)murine double minute 2 (MDM2)TNFα, and phosphatidylinositol-4,5-bisphosphate 3-kinase catalytic subunit alpha (PI3KCA) expression [245].

A novel polysaccharide isolated from T. linteus exhibited stronger inhibitory effects than 5-fluorouracil on murine sarcoma cancer cells (S-180) at 12 and 24 h, with inhibition rates of up to 93% at 72 h at a concentration of 200 µg/mL [126]. The polysaccharide’s potent in vitro anticancer properties were found to be mediated by a decreased expression of Bcl-2 with a rise in pro-apoptotic Bax protein expression, indicating that the observed anticancer effect was exerted via stimulation of apoptosis. Furthermore, two bioactive compounds (PL-ES and PL-I-ES) isolated from T. linteus were also found to exert apoptosis-associated anticancer effects on a total of 10 different cancer cell lines where 100 g/mL of PL-ES could suppress growth in all 10 cancer cells, while 100 and 250 g/mL of PL1 was able to cause significant growth reduction in four and seven cancer cells types respectively [231].

Besides from the above mentioned constituents, hispolon from T. linteus also exhibited anticancer effects in MCF-7, MDA-MB-231, T24, and J82 cells where it suppressed p53 mutant (MDA-MB-231, T24, and J82) and wild-type (MCF7) cells regardless of p53 status [232]. The extract induced apoptosis in the mentioned cancer cells, which was found to be mediated via chromatin condensation and an increase in poly (ADP-ribose) polymerase (PARP). Additionally, Liu et al. [233] reported that Atractylenolide I isolated from T. linteus was effective in gastric cancer cachexia management, where it was found to be significantly more effective than fish-oil-enriched nutritional supplementation at improving appetite, Karnofsky performance status, and decreasing proteolysis-inducing factor positive rates, suggesting the use of T. linteus as not only an anticancer agent, but also as a functional food during cancer treatment (Fig. 5).

Fig. 5
figure 5

Reported mechanisms through which Tropicoporus linteus exert anticancer properties

Hypoglycemic and hypolipidemic

The hypoglycemic properties of T. linteus extract from mycelia of solid-state culture was observed in diabetic rat models by Liu et al. [10], as evidenced by decreased serum glucose, glycated serum proteins, and insulin levels along with increased expressions of glucokinase and GLUT2 8 weeks after treatment with 120 and 600 mg/kg. A similar study by Kim et al. [234] also found that exo-polymers produced from the submerged cultures of T. linteus were also able to induce hypoglycemic effects on streptozotocin-induced diabetic rats, which was proposed to be through reparation of streptozotocin-exposed pancreatic β-cells, thus increasing insulin secretion leading to effective clearance of blood glucose. This finding is also supported by Sonawane et al. [246], where the crude polysaccharide of another mushroom from the Phellinus family, named Phellinus badius was also expected to exert the same effects on streptozotocin-induced diabetic rats. Moreover, polysaccharides isolated from T. linteus were found to improve insulin resistance via amplification of the population of short-chain fatty acids (SCFAs)-producing bacteria in the gut of streptozotocin induced-type 2 diabetes mellitus mice, which resulted in increased levels of SCFA that sustained the intestinal barrier function and thus caused reduction in LPS content in blood [125]. Consistent finding was also reported by Zhao et al. [124] in alloxan-induced diabetic mice, where mean percentage decrease in blood glucose levels of 22.35%, 16.35%, and 15.19% at concentrations of 100, 200, and 400 mg/kg of I. obliquus polysaccharides respectively was recorded.

Besides from its hypoglycemic properties, T. linteus extract from mycelia of solid-state culture is also capable of exerting hypolipidemic effects in HFD-induced rat models, with significant reductions in the levels of TG, TC, free fatty acids, and LDL-C in groups treated with 120 mg/kg and 600 mg/kg of the extract [10]. The extract’s hypolipidemic activities are proposed to be mediated through regulation of the expression of genes involved in lipid transport and metabolism, including an increase levels of fatty acid β-oxidation-associated genes acyl-CoA oxidase 1 (ACOX1) and CPT1A, elevation of amount of LDL receptors, and reduction of rate-limiting enzyme 3-hydroxy-3-methylglutaryl-CoA reductase (HMGCR), which together contribute to its potent hypolipidemic effects [125]. Similarly, administration of ethyl acetate fraction from T. linteus mycelia (PL-EA) for four weeks was able to improve body weight, hepatic lipid accumulation, and fasting blood glucose levels in HFD-induced mice through upregulating peroxisome proliferator-activated receptor gamma coactivator 1-alpha (PGC-1α) and adiponectin while downregulating glucokinase and SREBP-1c [138]. Furthermore, the major purified compounds of hispidin and hypholomine B in PL-EA significantly reduced the level of oleic and palmitic acids-induced lipid accumulation in HepG2 cells to contribute to the alleviation of non-alcoholic fatty liver in treated groups.

Antioxidant

According to Lee et al. [235], hispidin isolated from T. linteus was found to protect MIN6N β-cells of the pancreas from oxidative stress, where treatment with hispidin between the range of 10–70 µM revealed notable ROS scavenging activity in a dose-dependent manner, with the highest concentration having an activity of about 65%. The antioxidant properties of the extract were found to be mediated by an inhibition of caspase-3 activity, thus preventing the initiation of oxidative-stress induced apoptosis [235]. In addition, three constituents obtained from the culture broth of T. linteus, namely caffeic acid, inotilone, and 4-(3,4-Dihydroxyphenyl)-3-buten-2-one, were found to possess potent antioxidant activity in DPPH, ABTS and FRAP assay, while (2E,4E)-γ-Ionylideneacetic acid, Phellilane H, and (2E,4E)-4′-Hydroxy-γ-ionylideneacetic acid showed weak reducing power [133]. It should also be noted that caffeic acid, inotilone, 4-(3,4-Dihydroxyphenyl)-3-buten-2-one recorded higher antioxidant activity than butylated hydroxyanisole and Trolox in FRAP assay.

Three polysaccharides isolated from T. linteus exerted strong DPPH radical scavenging activity of between 58 to 72% at a concentration of 2 mg/mL [119]. Another study reported that hot water polysaccharide extracts of T. linteus fruiting bodies demonstrated notable antioxidant capabilities, with DPPH scavenging activities stronger than that of ascorbic acid at 5 mg/mL and above [236]. In addition, the stronger antioxidant activity of T. linteus was demonstrated by a significantly lower EC50 value than that of Aspergillus brasiliensis and Agaricus bisporus, further highlighting the potent antioxidant effects of T. linteus [236].

Hepatoprotective

Wang et al. [237] reported that the polysaccharides of T. linteus exhibited protective effects on thioacetamide liver fibrosis-induced mice via the regulation of the heat shock, oxidative stress, amino acid and nucleic acid metabolic pathways. Extracts isolated from T. linteus grown on germinated rice were also found to be capable of reducing peroxidation products in the liver of carbon tetrachloride-induced mice and decreasing cytochrome P450 2E1 (CYP2E1) protein expression [238]. Moreover, a decrease in glutamic transaminase enzyme released was also observed in H2O2-injured rat hepatocytes treated with 100 µg/ml methanol extract of T. linteus mycelial culture [239]. Further investigations revealed that the ethyl acetate fraction from the methanol extract of mycelial culture displayed 68.9 ± 5.3% protection to H2O2-injured hepatocytes and 46.8 ± 3.9% protection to galactosamine-injured hepatocytes, demonstrating the potent hepatoprotective effects of T. linteus.

Anti-inflammatory

T. linteus polysaccharides have been shown to substantially reduce the expression of key inflammatory cytokines, IL-6, IL-1β, and TNF-α, both in vivo in murine models induced with DSS and in vitro in LPS-stimulated RAW 264.7 macrophages through modulation of the MAPK and PPAR signaling pathways [115]. Similar conclusions were made by Shin et al. [240], who demonstrated the ability of T. linteus to deactivate NF-κB signaling through the inhibition of p38 MAPK, subsequently suppressing both inflammatory factors and elements associated with cartilage matrix degradation, thereby exerting a potent chondroprotective effect through simultaneous mitigation of ROS production and inflammation [240].

Extracts from T. linteus growth on germinated brown rice could inhibit NO and PGE2 production as well as reduce iNOS, NF-κB, COX-2, and TNF-α expressions to ameliorate pathological changes of colitis [241], while T. linteus grown on Panax ginseng media could suppress phosphorylation of spleen associated tyrosine kinase, GRB2 associated binding protein 2 (GAB2), and extracellular signal-regulated kinases required for degranulation and release of inflammatory cytokines [242]. Similar findings were obtained in a study by Lin et al. [145], where T. linteus-fermented broth exhibited dose-dependent inhibition of NO, NF-κB, and TNF-α in RAW264.7 cells and murine primary peritoneal exudate macrophages. A notable observation is the presence of high amounts of hispolon in the sample that showed the strongest anti-inflammatory activities, suggesting the important role of this compound in the inflammation-reducing functions of T. linteus [145]. Another compound isolated from T. linteus, inotilone, has been demonstrated to possess potent anti-inflammatory properties, capable of inhibiting NO production, protein expression of iNOS, NF-κB, and MMP-9, and suppression of LPS-induced ERK, c-Jun N-terminal kinase (JNK), and p38 phosphorylation in RAW264.7 macrophages [243]. Taken together, the abovementioned studies highlight the potent anti-inflammatory abilities of T. linteus and its potential for therapeutic applications.

Antiviral

Inhibition studies revealed a dose-dependent inhibitory activity of T. linteus methanol extract on α-glucosidase, a crucial component for the morphogenesis of enveloped viruses [244]. Due to the similarity between syncytium formation of Newcastle disease virus (NDV) and HIV, the inhibition effect of the extract on hemagglutinin-neuraminidase protein of NDV expressed on baby hamster kidney cells were measured, leading to the finding that T. linteus methanol extract was capable of suppressing NDV-induced syncytium formation at 12.5 µg/mL, similar to that of α-glucosidase inhibition [244]. This suggests the potential use of T. linteus against viruses such as HIV, HBV, and dengue virus, which are known to be highly sensitive to glucosidase inhibition.

Besides glucosidase, two constituents of T. linteus fermentation broth, inotilone as well as 4-(3,4-dihydroxyphenyl)-3-buten-2-one demonstrated strong neuraminidase inhibitory effects [13]. Microscopic findings revealed that 10 µM of both isolated constituents inhibited the formation of visible cytopathic effects, with IC50 values lower than that of Oseltamivir (61 and 52 µM respectively compared to 64 µM), suggesting the potential of T. linteus-derived constituents as stronger neuraminidase inhibitors of Influenza A H1N1 virus than the commonly used Oseltamivir treatment [13]. In addition, recent findings by Li et al. [35, 38] reported that hispidin and hypholomine B from the mycelial ethanol extract of T. linteus reduced the expression of angiotensin-converting enzyme 2 (ACE2) gene in HepG2 cells and blocked the spike receptor-binding domain to minimize the entry of the severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) virus, further demonstrating the strong antiviral and therapeutic potential of T. linteus for viral infections.

Medical evidence through clinical trials

Taiwanofungus camphoratus

The promising health-promoting and therapeutic effects of T. camphoratus have been demonstrated through both in vitro and in vivo experiments, as discussed previously. In recent years, clinical trials have been conducted to further validate the pharmacological activities of the medicinal mushroom in the biological conditions of the human body (Table 5).

Table 5 Clinical trials carried out on Taiwanofungus camphoratus, Inonotus obliquus, and Tropicoporus linteus

Chen et al. [247] found that 8 weeks of oral T. camphoratus mycelium treatment could induce significant reductions in systolic and diastolic blood pressure, which was proposed to be through interaction with the renin-angiotensin system to lower blood pressure by inhibiting renin secretion. The historical use of T. camphoratus as a traditional Chinese medicine for treating hypertension further substantiates these contemporary findings [255]. Golden-T. camphoratus administration was also capable of reducing AST, ALT, and TG levels with no significant effects on general safety parameters, suggesting the use of Golden-T. camphoratus as a safe and effective hepatoprotective agent [248].

In addition, a study conducted by Tsai et al. [249] to explore the combined effects of T. camphoratus treatment and chemotherapy found that although significant enhancements in fatigue or quality of life were not observed, T. camphoratus did contribute to improved sleep quality. Nevertheless, the potential benefits were clouded by adverse effects, including gastrointestinal discomfort and a reduction in platelet counts within a month of treatment, where the precise relationship between these effects and T. camphoratus remains to be fully elucidated. Although prior in vitro and in vivo studies such as that conducted by Huang et al. [256] have highlighted the capability of T. camphoratus to enhance the sensitivity of colon cancer cells to chemotherapy, these promising results did not translate directly to clinical trials, as evidenced by the divergent outcomes observed by Tsai et al. [249] due to potentially different interactions with the complexities of human clinical contexts. This underlines the need for cautious interpretation and further rigorous investigation to ascertain these findings in clinical applications.

Inonotus obliquus

Numerous in vitro and in vivo studies have highlighted the potential health benefits and therapeutic effects of I. obliquus, thus prompting the initiation of several clinical trials to explore the pharmacological activities of I. obliquus in human subjects.

In a study undertaken by Dosychev and Bystrova [250], it was found that 38 patients experienced a complete disappearance of psoriatic eruptions, while 8 patients exhibited significant improvement in related symptoms. This outcome resonates with a related study that highlighted the immune-promoting effect of I. obliquus, indicating its suitability in the context of molecular mechanisms of action in immunological disease driven by TNF-α, which pertains to conditions such as psoriasis [257]. Additionally, noticeable relief or complete resolution of gastrointestinal tract symptoms was observed [250], which agrees with an in vivo investigation conducted by Hu et al. [258] that reported that I. obliquus polysaccharide contributes to shaping the gut microbiota composition toward a healthier profile at the genus level.

An investigation carried out by Fedotov and Yu [251] found that I. obliquus was capable of reducing pain related to peptic ulcers, with more substantial effects seen at higher doses. However, interestingly, the cessation of I. obliquus treatment resulted in the return of pain levels to their original state. These findings are aligned with a related study conducted on rats which demonstrated the effective antiulcer activity of I. obliquus, further validating the therapeutic effect of this medicinal mushroom [259].

Despite its widespread global utilisation and the multitude of experiments conducted, the available corpus of peer-reviewed literature encompasses solely two clinical studies, both conducted over four decades ago, as previously indicated [250, 251]. It is important to underscore that no controlled trials assessing the safety of I. obliquus were identified, and no pertinent studies were found to be registered on Clinicaltrials.gov. Analogous to many mushroom supplements, the prevailing evidence concerning safety and efficacy heavily relies on the extensive historical use of I. obliquus. In light of this, further comprehensive studies, particularly through rigorous clinical trials, are imperative to establish a more nuanced and comprehensive understanding of I. obliquus's impact on the human body [260].

Tropicoporus linteus

A recent randomized, double-blind, placebo-controlled pilot trial conducted on participants with a history of upper respiratory infections validated the immunomodulatory effect of T. linteus as evidenced by heightened natural killer cell activity and elevated IL-6, immunoglobulin G1 (IgG1), IgG2, and IgM levels [252]. Another pilot clinical study revealed the ability of T. linteus to relieve pain resulting from knee osteoarthritis and return markers of inflammation to normal values, suggesting the potential of T. linteus in alleviating knee arthritis symptoms [253]. This is in line with the observation made by Kim et al. [261] where T. linteus was able to suppress collagen-induced arthritis in vivo by reducing anti-type II collagen IgG and IgG2a antibodies as well as various cytokines including IL-12, TNF-α, and IFN-γ in mice, further highlighting the pivotal role of T. linteus in preventing and treating joint inflammation.

In a study conducted by Sung et al. [254], patients with pancreatic cancer who underwent pancreatectomy were recruited over the course of 19 years and retrospectively analysed to examine the impact of T. linteus on adherence to postoperative adjuvant treatment. Remarkably, T. linteus medication increased patient adherence to postoperative chemotherapy to contribute to overall long-term oncologic outcomes, further highlighting the potential application of T. linteus as not only an anticancer agent as evidenced by in vitro and in vivo studies, but also as an adjuvant to improve outcomes to conventional chemotherapy.

Conclusion

This review has extensively examined the diverse biological activities, therapeutic effects, and pharmacological impacts of three notable mushrooms: T. camphoratus, I. obliquus, and T. linteus. Throughout numerous in vitro and in vivo studies, these mushrooms have revealed their potential through bioactive compounds that hold promise for human well-being. The growing interest in natural products as alternative treatments for conditions such as cancer and viral infections is evident from the increased scholarly and societal attention. Nonetheless, it is imperative to underscore the need for further comprehensive clinical trials that focus on safety, toxicity, and potential risks. Such investigations would offer robust scientific foundations and practical insights for optimising the biological effectiveness of these mushrooms while minimising any adverse effects. This will also encourage the development of mushroom products in not only medicinal applications, but also its exploitation in the cosmetic industry due to their excellent antioxidant, anti-aging, and moisturizing properties, ensuring a wider accessibility to the manifold benefits these mushroom species offer to a broader population.

Availability of data and materials

Not applicable.

Abbreviations

ACC:

Acetyl-CoA carboxylase

ACE:

Angiotensin-converting enzyme

ACOX:

Acyl-CoA oxidase

ADP:

Adenosine diphosphate

ALT:

Alanine aminotransferase

AMPK:

Adenosine monophosphate-activated protein kinase

AST:

Aspartate transaminase

CDK:

Cyclin-dependent kinase

COX:

Cyclooxygenase

CPT-1A:

Carnitine palmitoyltransferase 1A

CYP2E1:

Cytochrome P450 2E1

DSS:

Dextran sulfate sodium

ECG:

Epicatechin-3-gallate

EGCG:

Epigallocatechin-3-gallate

EGFR:

Epidermal growth factor receptor

EMT:

Epithelial-mesenchymal transition

ERK:

Extracellular signal-related kinases

FCV:

Feline calicivirus

FHV:

Feline herpesvirus 1

FIPV:

Feline infectious peritonitis virus

FIV:

Feline influenza virus

FPV:

Feline panleukopenia virus

FXR:

Farnesoid X receptor

GLUT4:

Glucose transporter 4

GSH:

Glutathione

HBeAg:

Hepatitis B e-antigen

HBV:

Hepatitis B virus

HCV:

Hepatitis C virus

HDL:

High-density lipoprotein

HDL-C:

High-density lipoprotein cholesterol

HER:

Human epidermal growth factor receptor

HFD:

High-fat diet

HIV:

Human immunodeficiency virus

HPLC:

High-performance liquid chromatography

HSV:

Herpes simplex virus

IBA-1:

Ionized calcium-binding adapter molecule-1

IFN:

Interferon

IL:

Interleukin

iNOS:

Inducible-nitric oxide synthase

IOPS:

I. obliquus Polysaccharides

JNK:

C-Jun N-terminal kinase

LDL:

Low-density lipoprotein

LDL-C:

Low-density lipoprotein cholesterol

LPS:

Lipopolysaccharide

MAPK:

Mitogen-activated protein kinases

MDA:

Malondialdehyde

MDM:

Murine double minute 2

MDR:

Multidrug resistance

MMP:

Matrix metalloproteinase

NDV:

Newcastle disease virus

NF-κB:

Nuclear factor kappa B

NO:

Nitric oxide

Nrf:

Nuclear factor erythroid 2–related factor 2

p-Akt:

Phospho-protein kinase B

PARP:

Poly (ADP-ribose) polymerase

PGC-1α:

Peroxisome proliferator-activated receptor gamma coactivator 1-alpha

PGE2:

Prostaglandin E2

PI3K:

Phosphoinositide 3-kinase

PI3KCA:

Phosphatidylinositol-4,5-bisphosphate 3-kinase catalytic subunit alpha

PPAR:

Peroxisome proliferator-activated receptor

Rb:

Retinoblastoma tumour suppressor protein

ROS:

Reactive oxygen species

SCFA:

Short-chain fatty acid

SFC:

Supercritical fluid chromatography

SHP:

Small heterodimer partner

SOD:

Superoxide dismutase

SREBP-1C:

Sterol-regulatory element binding protein-1C

STAT3:

Signal transducer and activator of transcription 3

TC:

Total cholesterol

TG:

Triacylglycerol

TNF:

Tumour necrosis factor

UV:

Ultraviolet

References

  1. Koparde AA, Jad RC, Magdum CS. Natural products in drug discovery. In: Perveen S, Al-Taweel A, editors. Pharmacognosy-medicinal plants. IntechOpen. 2019. https://doi.org/10.5772/intechopen.82860.

  2. Newman DJ, Cragg GM. Natural products as sources of new drugs over the nearly four decades from 01/1981 to 09/2019. J Nat Prod. 2020;83(3):770–803. https://doi.org/10.1021/acs.jnatprod.9b01285.

    Article  CAS  PubMed  Google Scholar 

  3. Cieśla Ł, Moaddel R. Comparison of analytical techniques for the identification of bioactive compounds from natural products. Nat Prod Rep. 2016;33:1131–45. https://doi.org/10.1039/c6np00016a.

    Article  CAS  PubMed  Google Scholar 

  4. Al-Tuwaijri MM, Basal WT, Sabry NM, Moussa TA, Das B, Eid JI. Inonotus obliquus polysaccharides inhibited cellular growth of NCI-H23 and A549 lung cancer cells through G0/G1 cell cycle arrest and ROS mediated cell death. Egypt Acad J Biol Sci C Physiol Mol Biol. 2021;13(1):27–40. https://doi.org/10.21608/eajbsc.2021.144738.

    Article  Google Scholar 

  5. Szychowski KA, Rybczyńska-Tkaczyk K, Tobiasz J, Yelnytska-Stawasz V, Pomianek T, Gmiński J. Biological and anticancer properties of Inonotus obliquus extracts. Process Biochem. 2018;73:180–7. https://doi.org/10.1016/j.procbio.2018.07.015.

    Article  CAS  Google Scholar 

  6. Arata S, Watanabe J, Maeda M, Yamamoto M, Matsuhashi H, Mochizuki M, Kagami N, Honda K, Inagaki M. Continuous intake of the Chaga mushroom (Inonotus obliquus) aqueous extract suppresses cancer progression and maintains body temperature in mice. Heliyon. 2016;2(5): e00111. https://doi.org/10.1016/j.heliyon.2016.e00111.

    Article  PubMed  PubMed Central  Google Scholar 

  7. Zhao Y, Zheng W. Deciphering the antitumoral potential of the bioactive metabolites from medicinal mushroom Inonotus obliquus. J Ethnopharmacol. 2021;265: 113321. https://doi.org/10.1016/j.jep.2020.113321.

    Article  CAS  PubMed  Google Scholar 

  8. Huang HT, Wang SL, Nguyen VB, Kuo YH. Isolation and identification of potent antidiabetic compounds from Antrodia cinnamomea—an edible Taiwanese mushroom. Molecules. 2018;23(11):2864. https://doi.org/10.3390/molecules23112864.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Wang C, Chen Z, Pan Y, Gao X, Chen H. Anti-diabetic effects of Inonotus obliquus polysaccharides-chromium (III) complex in type 2 diabetic mice and its sub-acute toxicity evaluation in normal mice. Food Chem Toxicol. 2017;108:498–509. https://doi.org/10.1016/j.fct.2017.01.007.

    Article  CAS  PubMed  Google Scholar 

  10. Liu Y, Wang C, Li J, Mei Y, Liang Y. Hypoglycemic and hypolipidemic effects of phellinus linteus mycelial extract from solid-state culture in a rat model of type 2 diabetes. Nutrients. 2019;11(2):296. https://doi.org/10.3390/nu11020296.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. He YC, Lu ZH, Shi P, Hao JC, Zhao ZJ, Xie HT, Mao P, Chen SJ. Anti-herpes simplex virus activities of bioactive extracts from Antrodia camphorata mycelia. Antivir Ther. 2015;21(5):377–83. https://doi.org/10.3851/IMP2988.

    Article  PubMed  Google Scholar 

  12. Shibnev VA, Garaev TM, Finogenova MP, Kalnina LB, Nosik DN. Antiviral activity of aqueous extracts of the birch fungus Inonotus obliquus on the human immunodeficiency virus. Vopr Virusol. 2015;60(2):35–8.

    CAS  PubMed  Google Scholar 

  13. Hwang BS, Lee MS, Lee SW, Lee IK, Seo GS, Choi HJ, Yun BS. Neuraminidase inhibitors from the fermentation broth of Phellinus linteus. Mycobiology. 2014;42(2):189–92. https://doi.org/10.5941/MYCO.2014.42.2.189.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Hsieh YL, Wu SP, Fang LW, Hwang TS. Effects of Antrodia camphorata extracts on anti-oxidation, anti-mutagenesis and protection of DNA against hydroxyl radical damage. BMC Complement Altern Med. 2015;15:237. https://doi.org/10.1186/s12906-015-0768-3.

    Article  PubMed  PubMed Central  Google Scholar 

  15. Chang Y, Bai M, Xue XB, Zou CX, Huang XX, Song SJ. Isolation of chemical compositions as dietary antioxidant supplements and neuroprotectants from Chaga mushroom (Inonotus obliquus). Food Biosci. 2022;47: 101623. https://doi.org/10.1016/j.fbio.2022.101623.

    Article  CAS  Google Scholar 

  16. Wu Y, Liu H, Li Z, Huang D, Nong L, Ning Z, Hu Z, Xu C, Yan JK. Purification of polysaccharides from Phellinus linteus by using an aqueous two-phase system and evaluation of the physicochemical and antioxidant properties of polysaccharides in vitro. Prep Biochem Biotechnol. 2022;52(1):89–98. https://doi.org/10.1080/10826068.2021.1911815.

    Article  CAS  PubMed  Google Scholar 

  17. Chen WJ, Chang FW. A pilot study to assess food safety and potential cholesterol-lowering efficacy of Antrodia cinnamomea solid-state cultivated mycelium in healthy adults. Evid-Based Complementary Altern Med. 2020;2020:5865764. https://doi.org/10.1155/2020/5865764.

    Article  Google Scholar 

  18. Liu D, Zhu L, Guo Y, Zhao Y, Betchem G, Yolandani Y, Ma H. Enhancing submerged fermentation of Antrodia camphorata by low-frequency alternating magnetic field. Innov Food Sci Emerg Technol. 2023;86: 103382. https://doi.org/10.1016/j.ifset.2023.103382.

    Article  CAS  Google Scholar 

  19. Chang S, Buswell J. Medicinal mushrooms: past, present and future. In: Berovic M, Zhong JJ, editors. Biochemical engineering and biotechnology of medicinal mushrooms. Springer: Cham; 2022. p. 1–27.

    Google Scholar 

  20. Wang, S.Y. Rubies From Taiwan’s Forest - Taiwanofungus Camphoratus, Its Bioactivity And Metabolites (2021). https://canr.nchu.edu.tw/web/internation/detail.php?id=476. Accessed 20 March 2024.

  21. Lu MY, Fan WL, Wang WF, Chen T, Tang YC, Chu FH, Chang TT, Wang SY, Li MY, Chen YH, Lin ZS. Genomic and transcriptomic analyses of the medicinal fungus Antrodia cinnamomea for its metabolite biosynthesis and sexual development. Proc Natl Acad Sci. 2014;111(44):E4743–52. https://doi.org/10.1073/pnas.1417570111.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Globenewswire. Chaga mushroom-based products consumption is predicted to reach a market value of US$ 87 billion by 2033: Fact.MR Report. https://www.globenewswire.com/news-release/2023/01/10/2586042/0/en/Chaga-Mushroom-Based-Products-Consumption-Is-Predicted-To-Reach-a-Market-Value-of-US-87-Billion-by-2033-Fact-MR-Report.html (2023). Accessed 20 March 2024.

  23. iNEWS. The ancient and young "Sanghuang" has returned to people's field of vision. 2024. https://inf.news/en/health/a72f5a59a737c0ac4a2395b1cd3f97d7.html. Accessed 20 March 2024.

  24. Zeng WW, Chen TC, Liu CH, Wang SY, Shaw JF, Chen YT. Identification and isolation of an intermediate metabolite with dual antioxidant and anti-proliferative activity present in the fungus Antrodia cinnamomea cultured on an alternative medium with Cinnamomum kanehirai leaf extract. Plants. 2021;10(4):737. https://doi.org/10.3390/plants10040737.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Geethangili M, Tzeng YM. Review of pharmacological effects of Antrodia camphorata and its bioactive compounds. Evid Based Complement Alternat Med. 2011. https://doi.org/10.1093/ecam/nep108.

    Article  PubMed  PubMed Central  Google Scholar 

  26. Bal TL, Raj JR, Richter DL. Influence of Chaga (Inonotus obliquus) treatment of wood in decay tests. Curr Res Environ Appl Mycol (J Fungal Biol). 2019;9(1):85–91. https://doi.org/10.5943/cream/9/1/9.

    Article  Google Scholar 

  27. Wu CC, Chang SH, Tung CW, Ho CK, Gogorcena Y, Chu FH. Identification of hybridization and introgression between Cinnamomum kanehirae Hayata and C. camphora (L.) Presl using genotyping-by-sequencing. Sci Rep. 2020;10(1):15995. https://doi.org/10.1038/s41598-020-72775-0.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Ao ZH, Xu ZH, Lu ZM, Xu HY, Zhang XM, Dou WF. Niuchangchih (Antrodia camphorata) and its potential in treating liver diseases. J Ethnopharmacol. 2009;121(2):194–212. https://doi.org/10.1016/j.jep.2008.10.039.

    Article  PubMed  Google Scholar 

  29. Du YC, Chang FR, Wu TY, Hsu YM, El-Shazly M, Chen CF, Sung PJ, Lin YY, Lin YH, Wu YC, Lu MC. Antileukemia component, dehydroeburicoic acid from Antrodia camphorata induces DNA damage and apoptosis in vitro and in vivo models. Phytomedicine. 2012;19(8–9):788–96. https://doi.org/10.1016/j.phymed.2012.03.014.

    Article  CAS  PubMed  Google Scholar 

  30. Chang ST. Overview of mushroom cultivation and utilization as functional foods. In: Cheung PCK, editor. Mushrooms as functional foods. New Jersey: Wiley; 2008. p. 1–33. https://doi.org/10.1002/9780470367285.ch1.

    Chapter  Google Scholar 

  31. Lu MC, El-Shazly M, Wu TY, Du YC, Chang TT, Chen CF, Hsu YM, Lai KH, Chiu CP, Chang FR, Wu YC. Recent research and development of Antrodia cinnamomea. Pharmacol Ther. 2013;139(2):124–56. https://doi.org/10.1016/j.pharmthera.2013.04.001.

    Article  CAS  PubMed  Google Scholar 

  32. Elisashvili VI. Submerged cultivation of medicinal mushrooms: bioprocesses and products. Int J Med Mushr. 2012. https://doi.org/10.1615/IntJMedMushr.v14.i3.10.

    Article  Google Scholar 

  33. He YC, He KZ, Pu Q, Li J, Zhao ZJ. Optimization of cultivating conditions for triterpenoids production from Antrodia cinnmomea. Indian J Microbiol. 2012;52:648–53. https://doi.org/10.1007/s12088-012-0290-y.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Meng L, Luo B, Yang Y, Faruque MO, Zhang J, Li X, Hu X. Addition of vegetable oil to improve triterpenoids production in liquid fermentation of medicinal fungus Antrodia cinnamomea. J Fungi. 2021;7(11):926. https://doi.org/10.3390/jof7110926.

    Article  CAS  Google Scholar 

  35. Li J, Wen S, Zhang B, Wang F. Selenium enrichment of the edible medicinal mushroom Antrodia camphorata by submerged fermentation. Molecules. 2023;28(7):3036. https://doi.org/10.3390/molecules28073036.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Letti LA, Vítola FM, de Melo Pereira GV, Karp SG, Medeiros AB, da Costa ES, Bissoqui L, Soccol CR. Solid-state fermentation for the production of mushrooms. In: Bui XT, Guo W, Chiemchaisri C, Pandey A, editor. Current developments in biotechnology and bioengineering. Elsevier; 2018. p. 285318.

  37. Bakratsas G, Polydera A, Katapodis P, Stamatis H. Recent trends in submerged cultivation of mushrooms and their application as a source of nutraceuticals and food additives. Future Foods. 2021;4: 100086. https://doi.org/10.1016/j.fufo.2021.100086.

    Article  CAS  Google Scholar 

  38. Li T, Lin T, Lu T, Tseng C, Lin C, Chu H, Li I, Chen C. Phellinus linteus mycelia extract in COVID-19 prevention and identification of its key metabolic compounds profiling using UPLC-QTOF-MS/MS spectrometry. Fitoterapia. 2023;171:105695–105695. https://doi.org/10.1016/j.fitote.2023.105695.

    Article  CAS  PubMed  Google Scholar 

  39. Kwon O, Kim Y, Paek JH, Park WY, Han S, Sin H, Jin K. Chaga mushroom-induced oxalate nephropathy that clinically manifested as nephrotic syndrome: a case report. Medicine. 2022. https://doi.org/10.1097/MD.0000000000028997.

    Article  PubMed  PubMed Central  Google Scholar 

  40. Géry A, Dubreule C, André V, Rioult JP, Bouchart V, Heutte N, Eldin de Pécoulas P, Krivomaz T, Garon D. Chaga (Inonotus obliquus), a future potential medicinal fungus in oncology? A chemical study and a comparison of the cytotoxicity against human lung adenocarcinoma cells (A549) and human bronchial epithelial cells (BEAS-2B). Integr Cancer Therapies. 2018;17(3):832–43. https://doi.org/10.1177/1534735418757912.

    Article  CAS  Google Scholar 

  41. Naveed D. Chaga mushrooms: the king of mushrooms. 2021. https://blog.suvie.com/chaga-mushrooms-the-king-of-mushrooms/. Accessed 7 Aug 2023.

  42. Robles A. Knowing when to harvest chaga: a guide to harvesting chaga mushroom at the right time | ShunCy. 2023. https://shuncy.com/article/when-harvest-chaga. Accessed: 7 Aug 2023.

  43. Cho NS, Shin YS. Optimization of in vitro cultivation of Inonotus obliquus. J Korean Wood Sci Technol. 2005;33(5):92–8.

    Google Scholar 

  44. Petre A, Ene M, Vamanu E. Submerged cultivation of Inonotus obliquus mycelium using statistical design of experiments and mathematical modeling to increase biomass yield. Appl Sci. 2021;11(9):4104. https://doi.org/10.3390/app11094104.

    Article  CAS  Google Scholar 

  45. Adamson K, Kütt K, Vester M, Jürimaa K, Silm M, Drenkhan R. Mycelial growth of Inonotus obliquus on malt extract media and on wood of different host species. Scand J For Res. 2023;38(3):131–43. https://doi.org/10.1080/02827581.2023.2210292.

    Article  Google Scholar 

  46. Huang HY, Chieh SY, Tso TK, Chien TY, Lin HT, Tsai YC. Orally administered mycelial culture of Phellinus linteus exhibits antitumor effects in hepatoma cell-bearing mice. J Ethnopharmacol. 2011;133(2):460–6. https://doi.org/10.1016/j.jep.2010.10.015.

    Article  PubMed  Google Scholar 

  47. Chen W, Tan H, Liu Q, Zheng X, Zhang H, Liu Y, Xu L. A review: the bioactivities and pharmacological applications of Phellinus linteus. Molecules. 2019;24(10):1888. https://doi.org/10.3390/molecules24101888.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Fokunang ET, Annih MG, Abongwa LE, Bih ME, Vanessa TM, Fomnboh DJ, Fokunang C. Medicinal Mushroom of Potential Pharmaceutical Toxic Importance: Contribution in Phytotherapy. In: Shiomi N, editor. Current Topics in Functional Food. Intech Open. 2022. https://doi.org/10.5772/intechopen.103845.

  49. Park HJ. Phellinus linteus grown on germinated brown rice suppress metastasis and induce apoptosis of colon cancer cells by suppressing NF-κB and Wnt/β-catenin signaling pathways. J Funct Foods. 2015;14:289–98. https://doi.org/10.1016/j.jff.2014.12.033.

    Article  Google Scholar 

  50. Lee JW, Baek SJ, Kim YS. Submerged culture of Phellinus linteus for mass production of polysaccharides. Mycobiology. 2008;36(3):178–82. https://doi.org/10.4489/MYCO.2008.36.3.178.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Min GJ, Kang HW. Artificial cultivation characteristics and bioactive effects of novel Tropicoporus linteus (Syn. Phellinus linteus) strains HN00K9 and HN6036 in Korea. Mycobiology. 2021;49(2):161–72. https://doi.org/10.1080/12298093.2021.1892568.

    Article  PubMed  PubMed Central  Google Scholar 

  52. Chen Q, Tang H, Zha Z, Yin H, Wang Y, Wang Y, Li H, Yue L. β-d-glucan from Antrodia Camphorata ameliorates LPS-induced inflammation and ROS production in human hepatocytes. Int J Biol Macromol. 2017;104:768–77. https://doi.org/10.1016/j.ijbiomac.2017.05.191.

    Article  CAS  PubMed  Google Scholar 

  53. Lu CL, Lee BH, Ren YL, Ji D, Rao SQ, Li HX, Yang ZQ. Effects of exopolysaccharides from Antrodia cinnamomea on inflammation and intestinal microbiota disturbance induced by antibiotics in mice. Food Biosci. 2022;50: 102116. https://doi.org/10.1016/j.fbio.2022.102116.

    Article  CAS  Google Scholar 

  54. Tang H, Nie W, Xiao J, Zha Z, Chen Q, Yin H. Structural characterization and anti-inflammatory effect in hepatocytes of a galactoglucan from Antrodia camphorata mycelium. RSC Adv. 2019;9:7664–72. https://doi.org/10.1039/C8RA10347J.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Wang ZQ, Zhu CX, Dai AR, Chen L, You CP, Zhang BB. Chemical characterization and antioxidant properties of cell wall polysaccharides from Antrodia cinnamomea mycelia. Food Biosci. 2021;41: 100932. https://doi.org/10.1016/j.fbio.2021.100932.

    Article  CAS  Google Scholar 

  56. Zhang Y, Wang Z, Li D, Zang W, Zhu H, Wu P, Mei Y, Liang Y. A polysaccharide from Antrodia cinnamomea mycelia exerts antitumor activity through blocking of TOP1/TDP1-mediated DNA repair pathway. Int J Biol Macromol. 2018;120:1551–60. https://doi.org/10.1016/j.ijbiomac.2018.09.162.

    Article  CAS  PubMed  Google Scholar 

  57. Lee MH, Chao CH, Hsu YC, Lu MK. Production, characterization, and functions of sulfated polysaccharides from zinc sulfate enriched cultivation of Antrodia cinnamomea. Int J Biol Macromol. 2020;159:1013–21. https://doi.org/10.1016/j.ijbiomac.2020.05.068.

    Article  CAS  PubMed  Google Scholar 

  58. Chen JJ, Chao CH, Chang PC, Lu MK. Studies on anti-inflammatory activity of sulfated polysaccharides from cultivated fungi Antrodia cinnamomea. Food Hydrocolloids. 2016;53:37–45. https://doi.org/10.1016/j.foodhyd.2014.09.035.

    Article  CAS  Google Scholar 

  59. Liu Y, Ding Y, Ye M, Zhu T, Tian D, Ding K. A novel heterogalactan from Antrodia camphorata and anti-angiogenic activity of its sulfated derivative. Polymers. 2017;9(6):228. https://doi.org/10.3390/polym9060228.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Xia Y, Meng P, Liu S, Tan Z, Yang X, Liang L, Xie F, Zhang H, Wang G, Xiong Z, Lo J, Ai L. Structural and potential functional properties of alkali-extracted dietary fiber from Antrodia camphorata. Front Microbiol. 2022;13: 921164. https://doi.org/10.3389/fmicb.2022.921164.

    Article  PubMed  PubMed Central  Google Scholar 

  61. Li B, Kuang Y, He JB, Tang R, Zu LL, Leung CH, Ma DL, Qiao X, Ye M. Antcamphorols A-K, cytotoxic and ROS scavenging triterpenoids from Antrodia camphorata. J Nat Prod. 2019;83(1):45–54. https://doi.org/10.1021/acs.jnatprod.9b00580.

    Article  CAS  PubMed  Google Scholar 

  62. Li B, Kuang T, Zhang M, He JB, Xu LL, Leung CH, Ma DL, Lo JY, Qiao X, Ye M. Cytotoxic triterpenoids from Antrodia camphorata as sensitizers of paclitaxel. Org Chem Front. 2020;7:768–79. https://doi.org/10.1039/c9qo01516g.

    Article  CAS  Google Scholar 

  63. Li B, Lu YY, Lo JY, Qiao X, Ye M. Chemical constituents from the dish-cultured Antrodia camphorata and their cytotoxic activities. J Asian Nat Prod Res. 2021;23(7):666–74. https://doi.org/10.1080/10286020.2020.1810673.

    Article  CAS  PubMed  Google Scholar 

  64. Chen YC, Liu YL, Li FY, Chang CI, Wang SY, Lee KY, Li SL, Chen YP, Jinn TR, Tzen JT. Antcin A, a steroid-like compound from Antrodia camphorata, exerts anti-inflammatory effect via mimicking glucocorticoids. Acta Pharmacol Sin. 2011;32:904–11. https://doi.org/10.1038/aps.2011.36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Wang YJ, Lee SC, Hsu CH, Kuo YH, Yang CC, Lin FJ. Antcins, triterpenoids from Antrodia cinnamomea, as new agonists for peroxisome proliferator-activated receptor α. J Food Drug Anal. 2019;27(1):295–304. https://doi.org/10.1016/j.jfda.2018.11.004.

    Article  CAS  PubMed  Google Scholar 

  66. Qiao X, An R, Huang Y, Ji S, Li L, Tzeng YM, Guo DA, Ye M. Separation of 25R/S-ergostane triterpenoids in the medicinal mushroom Antrodia camphorata using analytical supercritical-fluid chromatography. J Chromatogr. 2014;1358:252–60. https://doi.org/10.1016/j.chroma.2014.06.074.

    Article  CAS  Google Scholar 

  67. He R, Wu K, Zhang A, Xie Z, Sun P. Mechanochemical-assisted extraction and pharmacological study of triterpenoids from Antrodia camphorata. Appl Sci. 2019;9(20):4281. https://doi.org/10.3390/app9204281.

    Article  CAS  Google Scholar 

  68. Lee TH, Chen CC, Chen JJ, Liao HF, Chang SH, Sung PJ, Tseng MH, Wang SY, Ko HH, Kuo YH. New and cytotoxic components from Antrodia camphorata. Molecules. 2014;19(12):21378–85. https://doi.org/10.3390/molecules191221378.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Sun P, Li W, Guo J, Peng Q, Ye X, Hu S, Liu Y, Liu W, Chen H, Qiao J, Sun B. Ergosterol isolated from Antrodia camphorata suppresses LPS-induced neuroinflammatory responses in microglia cells and ICR mice. Molecules. 2023;28(5):2406. https://doi.org/10.3390/molecules28052406.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Yang HL, Tsai CH, Shrestha S, Lee CC, Liao JW, Hseu YC. Coenzyme Q0, a novel quinone derivative of Antrodia camphorata, induces ROS-mediated cytotoxic autophagy and apoptosis against human glioblastoma cells in vitro and in vivo. Food Chem Toxicol. 2021;155: 111238. https://doi.org/10.1016/j.fct.2021.112384.

    Article  CAS  Google Scholar 

  71. Lee WT, Lee TH, Cheng CH, Chen KC, Chen YC, Lin CW. Antroquinonol from Antrodia Camphorata suppresses breast tumor migration/invasion through inhibiting ERK-AP-1-and AKT-NF-κB-dependent MMP-9 and epithelial-mesenchymal transition expressions. Food Chem Toxicol. 2015;78:33–41. https://doi.org/10.1016/j.fct.2015.01.012.

    Article  CAS  PubMed  Google Scholar 

  72. Tu PC, Jiang WP, Lin MK, Huang GJ, Li YJ, Kuo YH. Anti-inflammatory constituents of Antrodia camphorata on RAW 2647 cells induced by polyinosinic-polycytidylic acid. Molecules. 2022;27(16):5320. https://doi.org/10.3390/molecules27165320.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Hu YD, Lu RQ, Liao XR, Zhang BB, Xu GR. Stimulating the biosynthesis of antroquinonol by addition of effectors and soybean oil in submerged fermentation of Antrodia camphorata. Biotechnol Appl Biochem. 2016;63(3):398–406. https://doi.org/10.1002/bab.1387.

    Article  CAS  PubMed  Google Scholar 

  74. Liu X, Xia Y, Zhang Y, Yi Z, Meng P, Wang G, Ai L. Enhancement of antroquinonol and antrodin C productions via in situ extractive fermentation of Antrodia camphorata S-29. Appl Microbiol Biotechnol. 2019;103:8351–61. https://doi.org/10.1007/s00253-019-10034-7.

    Article  CAS  PubMed  Google Scholar 

  75. Xia Y, Chen Y, Liu X, Zhou X, Wang Z, Wang G, Xiong Z, Ai L. Enhancement of antroquinonol production during batch fermentation using pH control coupled with an oxygen vector. J Sci Food Agric. 2018;99(1):449–56. https://doi.org/10.1002/jsfa.9206.

    Article  CAS  PubMed  Google Scholar 

  76. Liu XF, Xia YJ, Lai PF, Zhang Y, Yi ZW, Xie CL, Hong YQ, Ai LZ. An increase in cell membrane permeability in the in situ extractive fermentation improves the production of antroquinonol from Antrodia camphorata S-29. J Ind Microbiol Biotechnol. 2020;47(2):197–207. https://doi.org/10.1007/s10295-020-02258-8.

    Article  CAS  PubMed  Google Scholar 

  77. Tao Y, Chen L, Pu F, Xie Z, Tong S, Yan J. An efficient high-speed countercurrent chromatography method for preparative isolation of highly potent anti-cancer compound antroquinonol from Antrodia camphorata after experimental design optimized extraction. J Sep Sci. 2021;44(13):2655–62. https://doi.org/10.1002/jssc.202100162.

    Article  CAS  PubMed  Google Scholar 

  78. Xia Y, Zhou X, Liang L, Liu X, Li H, Xiong Z, Wang G, Song X, Ai L. Genetic evidence for the requirements of antroquinonol biosynthesis by Antrodia camphorata during liquid-state fermentation. J Ind Microbiol Biotechnol. 2022;49(1):kuab086. https://doi.org/10.1093/jimb/kuab086.

    Article  CAS  PubMed  Google Scholar 

  79. Phuong DT, Ma CM, Hattori M, Jin JS. Inhibitory effects of antrodins A-E from Antrodia cinnamomea and their metabolites on hepatitis C virus protease. Phytotherapy Res Int J Devoted Pharmacol Toxicol Eval Nat Prod Deriv. 2009;23(4):582–4. https://doi.org/10.1002/ptr.2657.

    Article  CAS  Google Scholar 

  80. Yi ZW, Xia YJ, Liu XF, Wang GQ, Xiong ZQ, Ai LZ. Antrodin A from mycelium of Antrodia camphorata alleviates acute alcoholic liver injury and modulates intestinal flora dysbiosis in mice. J Ethnopharmacol. 2020;254: 112681. https://doi.org/10.1016/j.jep.2020.112681.

    Article  CAS  PubMed  Google Scholar 

  81. Geng Y, Wang J, Sun Q, Xie M, Lu ZM, Xu HY, Shi JS, Xu ZH. Identification of antrodin B from Antrodia camphorata as a new anti-hepatofibrotic compound using a rapid cell screening method and biological evaluation. Hepatol Res. 2015;46(3):E15–25. https://doi.org/10.1111/hepr.12516.

    Article  CAS  PubMed  Google Scholar 

  82. Tsay HJ, Liu HK, Kuo YH, Chiu CS, Liang CC, Chung CW, Chen CC, Chen YP, Shiao YJ. EK100 and antrodin c improve brain amyloid pathology in APP/PS1 transgenic mice by promoting microglial and perivascular clearance pathways. Int J Mol Sci. 2021;22(19):10413. https://doi.org/10.3390/ijms221910413.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Xu XY, Geng Y, Xu HX, Ren Y, Liu DY, Mao Y. Antrodia camphorata-derived antrodin C inhibits liver fibrosis by blocking TGF-Beta and PDGF signaling pathways. Front Mol Biosci. 2022;9: 835508. https://doi.org/10.3389/fmolb.2022.835508.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Hsieh YY, Lee KC, Cheng KC, Lee KF, Yang YL, Chu HT, Lin TW, Chen CC, Hsieh MC, Huang CY, et al. Antrodin C isolated from Antrodia Cinnamomea induced apoptosis through ROS/AKT/ERK/P38 signaling pathway and epigenetic histone acetylation of TNFα in colorectal cancer cells. Antioxidants. 2023;12(3):764. https://doi.org/10.3390/antiox12030764.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Liu P, Xue J, Tong S, Dong W, Wu P. Structure characterization and hypoglycaemic activities of two polysaccharides from Inonotus obliquus. Molecules. 2018;23(8):1948. https://doi.org/10.3390/molecules23081948.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Xue J, Tong S, Wang Z, Liu P. Chemical characterization and hypoglycaemic activities in vitro of two polysaccharides from Inonotus obliquus by submerged culture. Molecules. 2018;23(12):3261. https://doi.org/10.3390/molecules23123261.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Liu Z, Yu D, Li L, Liu X, Zhang H, Sun W, Lin CC, Chen J, Chen Z, Wang W, Jia W. Three-phase partitioning for the extraction and purification of polysaccharides from the immunomodulatory medicinal mushroom Inonotus obliquus. Molecules. 2019;24(4):403. https://doi.org/10.3390/molecules24030403.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Hu Y, Sheng Y, Yu M, Li K, Ren G, Xu X, Qu J. Antioxidant activity of Inonotus obliquus polysaccharide and its amelioration for chronic pancreatitis in mice. Int J Biol Macromol. 2016;87:348–56. https://doi.org/10.1016/j.ijbiomac.2016.03.006.

    Article  CAS  PubMed  Google Scholar 

  89. Han Y, Nan S, Fan J, Chen Q, Zhang Y. Inonotus obliquus polysaccharides protect against Alzheimer’s disease by regulating Nrf2 signaling and exerting antioxidative and antiapoptotic effects. Int J Biol Macromol. 2019;131:769–78. https://doi.org/10.1016/j.ijbiomac.2019.03.033.

    Article  CAS  PubMed  Google Scholar 

  90. Wold CW, Kjeldsen C, Corthay A, Rise F, Christensen BE, Duus JØ, Inngjerdingen KT. Structural characterization of bioactive heteropolysaccharides from the medicinal fungus Inonotus obliquus (Chaga). Carbohyd Polym. 2018;185:27–40. https://doi.org/10.1016/j.carbpol.2017.12.041.

    Article  CAS  Google Scholar 

  91. Chen Y, Huang Y, Cui Z, Liu J. Purification, characterization and biological activity of a novel polysaccharide from Inonotus obliquus. Int J Biol Macromol. 2015;79:587–94. https://doi.org/10.1016/j.ijbiomac.2015.05.016.

    Article  CAS  PubMed  Google Scholar 

  92. Jiang S, Shi F, Lin H, Ying Y, Luo L, Huang D, Luo Z. Inonotus obliquus polysaccharides induces apoptosis of lung cancer cells and alters energy metabolism via the LKB1/AMPK axis. Int J Biol Macromol. 2020;151:1277–86. https://doi.org/10.1016/j.ijbiomac.2019.10.174.

    Article  CAS  PubMed  Google Scholar 

  93. Su L, Xin C, Yang J, Dong L, Mei H, Dai X, Wang Q. A polysaccharide from Inonotus obliquus ameliorates intestinal barrier dysfunction in mice with type 2 diabetes mellitus. Int J Biol Macromol. 2022;214:312–23. https://doi.org/10.1016/j.ijbiomac.2022.06.071.

    Article  CAS  PubMed  Google Scholar 

  94. Zhang J, Cheng S, Liang J, Qu J. Polysaccharide from fermented mycelium of Inonotus obliquus attenuates the ulcerative colitis and adjusts the gut microbiota in mice. Microb Pathog. 2023;177: 105990. https://doi.org/10.1016/j.micpath.2023.105990.

    Article  CAS  PubMed  Google Scholar 

  95. Kim J, Yang SC, Hwang AY, Cho H, Hwang KT. Composition of triterpenoids in Inonotus obliquus and their anti-proliferative activity on cancer cell lines. Molecules. 2020;25(18):4066. https://doi.org/10.3390/molecules25184066.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Peng A, Liu S, Fang L, Zhu Z, Zhou Y, Yue S, Ma Z, Liu X, Xue S, Qiu Y, Qi R. Inonotus obliquus and its bioactive compounds alleviate non-alcoholic fatty liver disease via regulating FXR/SHP/SREBP-1c axis. Eur J Pharmacol. 2022;921: 174841. https://doi.org/10.1016/j.ejphar.2022.174841.

    Article  CAS  PubMed  Google Scholar 

  97. Wold CW, Gerwick WH, Wangensteen H, Inngjerdingen KT. Bioactive triterpenoids and water-soluble melanin from Inonotus obliquus (Chaga) with immunomodulatory activity. J Funct Foods. 2020;71: 104025. https://doi.org/10.1016/j.jff.2020.104025.

    Article  CAS  Google Scholar 

  98. Zou CX, Dong SH, Hou ZL, Yao GD, Lin B, Huang XX, Song SJ. Modified lanostane-type triterpenoids with neuroprotective effects from the fungus Inonotus obliquus. Bioorg Chem. 2020;105: 104438. https://doi.org/10.1016/j.bioorg.2020.104438.

    Article  CAS  PubMed  Google Scholar 

  99. Park S, Shin H, Park D, Kim H, Byun Y, Lee KY. Structure elucidation of a new triterpene from Inonotus obliquus. Magn Reson Chem. 2021;59(4):489–94. https://doi.org/10.1002/mrc.5102.

    Article  CAS  PubMed  Google Scholar 

  100. Kang JH, Jang JE, Mishra SK, Lee HJ, Nho CW, Shin D, Jin M, Kim MK, Choi C, Oh SH. Ergosterol peroxide from Chaga mushroom (Inonotus obliquus) exhibits anti-cancer activity by down-regulation of the β-catenin pathway in colorectal cancer. J Ethnopharmacol. 2015;173:303–12. https://doi.org/10.1016/j.jep.2015.07.030.

    Article  CAS  PubMed  Google Scholar 

  101. Ma L, Chen H, Dong P, Lu X. Anti-inflammatory and anticancer activities of extracts and compounds from the mushroom Inonotus obliquus. Food Chem. 2013;139(1–4):503–8. https://doi.org/10.1016/j.foodchem.2013.01.030.

    Article  CAS  PubMed  Google Scholar 

  102. Chen SD, Yong TQ, Xiao C, Gao X, Xie YZ, Hu HP, Li XM, Chen DL, Pan HH, Wu QP. Inhibitory effect of triterpenoids from the mushroom Inonotus obliquus against α-glucosidase and their interaction: Inhibition kinetics and molecular stimulations. Bioorg Chem. 2021;115: 105276. https://doi.org/10.1016/j.bioorg.2021.105276.

    Article  CAS  PubMed  Google Scholar 

  103. Zhao F, Xia G, Chen L, Zhao J, Xie Z, Qiu F, Han G. Chemical constituents from Inonotus obliquus and their antitumor activities. J Nat Med. 2016;70:721–30. https://doi.org/10.1007/s11418-016-1002-4.

    Article  CAS  PubMed  Google Scholar 

  104. Liu C, Zhao C, Pan HH, Kang J, Yu XT, Wang HQ, Li BM, Xie YZ, Chen RY. Chemical constituents from Inonotus obliquus and their biological activities. J Nat Prod. 2014;77(1):35–41. https://doi.org/10.1021/np400552w.

    Article  CAS  PubMed  Google Scholar 

  105. Zhao F, Mai Q, Ma J, Xu M, Wang X, Cui T, Qiu F, Han G. Triterpenoids from Inonotus obliquus and their antitumor activities. Fitoterapia. 2015;101:34–40. https://doi.org/10.1016/j.fitote.2014.12.005.

    Article  CAS  PubMed  Google Scholar 

  106. Zou CX, Hou ZL, Bai M, Guo R, Lin B, Wang XB, Huang XX, Song SJ. Highly modified steroids from Inonotus obliquus. Org Biomol Chem. 2020;18(20):3908–16. https://doi.org/10.1039/d0ob00474j.

    Article  CAS  PubMed  Google Scholar 

  107. Kou RW, Han R, Gao YQ, Li D, Yin X, Gao JM. Anti-neuroinflammatory polyoxygenated lanostanoids from Chaga mushroom Inonotus obliquus. Phytochemistry. 2021;184: 112647. https://doi.org/10.1016/j.phytochem.2020.112647.

    Article  CAS  PubMed  Google Scholar 

  108. Wei YM, Yang L, Mei WL, Chen HQ, Cai CH, Li W, Dong WH, Chen ZB, Dai HF. Phenolic compounds from the sclerotia of Inonotus obliquus. Nat Prod Res. 2022;36(9):2413–7. https://doi.org/10.1080/14786419.2020.1833202.

    Article  CAS  PubMed  Google Scholar 

  109. Hwang BS, Lee IK, Yun BS. Phenolic compounds from the fungus Inonotus obliquus and their antioxidant properties. J Antibiot. 2016;69:108–10. https://doi.org/10.1038/ja.2015.83.

    Article  CAS  Google Scholar 

  110. Wang Y, Ouyang F, Teng C, Qu J. Optimization for the extraction of polyphenols from Inonotus obliquus and its antioxidation activity. Prep Biochem Biotechnol. 2021;51(9):852–9. https://doi.org/10.1080/10826068.2020.1864642.

    Article  CAS  PubMed  Google Scholar 

  111. Burmasova MA, Utebaeva AA, Sysoeva EV, Sysoeva MA. Melanins of Inonotus obliquus: bifidogenic and antioxidant properties. Biomolecules. 2019;9(6):248. https://doi.org/10.3390/biom9060248.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Zhou R, Yang H, Lu T, Zhao Y, Zheng W. Ultraviolet radiation promotes the production of hispidin polyphenols by medicinal mushroom Inonotus obliquus. Fungal Biol. 2022;126(11–12):775–85. https://doi.org/10.1016/j.funbio.2022.10.001.

    Article  CAS  PubMed  Google Scholar 

  113. Xu X, Shen M, Quan L. Stimulatory agents simultaneously improving the production and antioxidant activity of polyphenols from Inonotus obliquus by submerged fermentation. Appl Biochem Biotechnol. 2015;176:1237–50. https://doi.org/10.1007/s12010-015-1642-y.

    Article  CAS  PubMed  Google Scholar 

  114. Xu X, Zhao W, Shen M. Antioxidant activity of liquid cultured Inonotus obliquus polyphenols using tween-20 as a stimulatory agent: correlation of the activity and the phenolic profiles. J Taiwan Inst Chem Eng. 2016;69:41–7. https://doi.org/10.1016/j.jtice.2016.10.011.

    Article  CAS  Google Scholar 

  115. Hu T, Lin Q, Guo T, Yang T, Zhou W, Deng X, Yan JK, Luo Y, Ju M, Luo F. Polysaccharide isolated from Phellinus linteus mycelia exerts anti-inflammatory effects via MAPK and PPAR signaling pathways. Carbohyd Polym. 2018;200:487–97. https://doi.org/10.1016/j.carbpol.2018.08.021.

    Article  CAS  Google Scholar 

  116. Suh MG, Shin HY, Jeong EJ, Kim G, Jeong SB, Ha EJ, Choi SY, Moon SK, Shin KS, Yu KW, et al. Identification of galacturonic acid-rich polysaccharide with intestinal immune system modulating activity via Peyer’s patch from postbiotics of Phellinus linteus mycelial submerged culture. Int J Biol Macromol. 2023;234: 123685. https://doi.org/10.1016/j.ijbiomac.2023.123685.

    Article  CAS  PubMed  Google Scholar 

  117. Suabjakyong P, Nishimura K, Toida T, Van Griensven LJ. Structural characterization and immunomodulatory effects of polysaccharides from Phellinus linteus and Phellinus igniarius on the IL-6/IL-10 cytokine balance of the mouse macrophage cell lines (RAW 264.7). Food Funct. 2015;6(8):2834–44. https://doi.org/10.1039/C5FO00491H.

    Article  CAS  PubMed  Google Scholar 

  118. Su CH, Lai MN, Lin CC, Ng LT. Comparative characterization of physicochemical properties and bioactivities of polysaccharides from selected medicinal mushrooms. Appl Microbiol Biotechnol. 2016;100:4385–93. https://doi.org/10.1007/s00253-015-7260-3.

    Article  CAS  PubMed  Google Scholar 

  119. Wang ZB, Pei JJ, Ma HL, Cai PF, Yan JK. Effect of extraction media on preliminary characterizations and antioxidant activities of Phellinus linteus polysaccharides. Carbohyd Polym. 2014;109:49–55. https://doi.org/10.1016/j.carbpol.2014.03.057.

    Article  CAS  Google Scholar 

  120. Yan JK, Wang YY, Ma HL, Wang ZB. Ultrasonic effects on the degradation kinetics, preliminary characterization and antioxidant activities of polysaccharides from Phellinus linteus mycelia. Ultrason Sonochem. 2016;29:251–7. https://doi.org/10.1016/j.ultsonch.2015.10.005.

    Article  CAS  PubMed  Google Scholar 

  121. Yan JK, Wang YY, Ma HL, Wang ZB, Pei JJ. Structural characteristics and antioxidant activity in vivo of a polysaccharide isolated from Phellinus linteus mycelia. J Taiwan Inst Chem Eng. 2016;65:110–7. https://doi.org/10.1016/j.jtice.2016.05.052.

    Article  CAS  Google Scholar 

  122. Yan JK, Wang YY, Wang ZB, Ma HL, Pei JJ, Wu JY. Structure and antioxidative property of a polysaccharide from an ammonium oxalate extract of Phellinus linteus. Int J Biol Macromol. 2016;91:92–9. https://doi.org/10.1016/j.ijbiomac.2016.05.063.

    Article  CAS  PubMed  Google Scholar 

  123. Wu Y, Liu H, Li Z, Huang D, Nong L, Ning Z, Hu Z, Xu C, Yan JK. Purification of polysaccharides from Phellinus linteus by using an aqueous two-phase system and evaluation of the physicochemical and antioxidant properties of polysaccharides in vitro. Prep Biochem Biotechnol. 2021;52(1):89–98. https://doi.org/10.1080/10826068.2021.1911815.

    Article  CAS  PubMed  Google Scholar 

  124. Zhao C, Liao Z, Wu X, Liu Y, Liu X, Lin Z, Huang Y, Liu B. Isolation, purification, and structural features of a polysaccharide from Phellinus linteus and its hypoglycemic effect in alloxan-induced diabetic mice. J Food Sci. 2014;79(5):H1002–10. https://doi.org/10.1111/1750-3841.12464.

    Article  CAS  PubMed  Google Scholar 

  125. Liu Y, Wang C, Li J, Li T, Zhang Y, Liang Y, Mei Y. Phellinus linteus polysaccharide extract improves insulin resistance by regulating gut microbiota composition. FASEB J. 2019;34(1):1065–78. https://doi.org/10.1096/fj.201901943RR.

    Article  CAS  PubMed  Google Scholar 

  126. Mei Y, Zhu H, Hu Q, Liu Y, Zhao S, Peng N, Liang Y. A novel polysaccharide from mycelia of cultured Phellinus linteus displays antitumor activity through apoptosis. Carbohyd Polym. 2015;124:90–7. https://doi.org/10.1016/j.carbpol.2015.02.009.

    Article  CAS  Google Scholar 

  127. Pei JJ, Wang ZB, Ma HL, Yan JK. Structural features and antitumor activity of a novel polysaccharide from alkaline extract of Phellinus linteus mycelia. Carbohyd Polym. 2015;115:472–7. https://doi.org/10.1016/j.carbpol.2014.09.017.

    Article  CAS  Google Scholar 

  128. Chen C, Liu X, Qi S, Dias ACP, Yan J, Zhang X. Hepatoprotective effect of Phellinus linteus mycelia polysaccharide (PL-N1) against acetaminophen-induced liver injury in mouse. Int J Biol Macromol. 2020;154:1276–84. https://doi.org/10.1016/j.ijbiomac.2019.11.002.

    Article  CAS  PubMed  Google Scholar 

  129. Liu T, Zhao M, Zhang Y, Wang Z, Yuan B, Zhao C, Wang M. Integrated microbiota and metabolite profiling analysis of prebiotic characteristics of Phellinus linteus polysaccharide in vitro fermentation. Int J Biol Macromol. 2023;242: 124854. https://doi.org/10.1016/j.ijbiomac.2023.124854.

    Article  CAS  PubMed  Google Scholar 

  130. Gao N, Zhang W, Hu D, Lin G, Wang J, Xue F, Wang Q, Zhao H, Dou X, Zhang L. Study on extraction, physicochemical properties, and bacterio-static activity of polysaccharides from Phellinus linteus. Molecules. 2023;28(13):5102. https://doi.org/10.3390/molecules28135102.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Ota K, Yamazaki I, Saigoku T, Fukui M, Miyata T, Kamaike K, Shirahata T, Mizuno F, Asada Y, Hirotani M, et al. Phellilane L, sesquiterpene metabolite of Phellinus linteus: Isolation, structure elucidation, and asymmetric total synthesis. J Org Chem. 2017;82(23):12377–85. https://doi.org/10.1021/acs.joc.7b02141.

    Article  CAS  PubMed  Google Scholar 

  132. Shirahata T, Ino C, Mizuno F, Asada Y, Hirotani M, Petersson GA, Ōmura S, Yoshikawa T, Kobayashi Y. γ-Ionylidene-type sesquiterpenoids possessing antimicrobial activity against Porphyromonas gingivalis from Phellinus linteus and their absolute structure determination. J Antibiot. 2017;70:695–8. https://doi.org/10.1038/ja.2017.35.

    Article  CAS  Google Scholar 

  133. Lee MS, Hwang BS, Lee IK, Seo GS, Yun BS. Chemical constituents of the culture broth of Phellinus linteus and their antioxidant activity. Mycobiology. 2015;43(1):43–8. https://doi.org/10.5941/MYCO.2015.43.1.43.

    Article  PubMed  PubMed Central  Google Scholar 

  134. Huang SC, Wang PW, Kuo PC, Hung HY, Pan TL. Hepatoprotective principles and other chemical constituents from the mycelium of Phellinus linteus. Molecules. 2018;23(7):1705. https://doi.org/10.3390/molecules23071705.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Kojima K, Ohno T, Inoue M, Mizukami H, Nagatsu A. Phellifuropyranone A: a new furopyranone compound isolated from fruit bodies of wild Phellinus linteus. Chem Pharm Bull. 2008;56(2):173–5. https://doi.org/10.1248/cpb.56.173.

    Article  CAS  Google Scholar 

  136. Lee YS, Kang YH, Jung JY, Kang IJ, Han SN, Chung JS, Shin HK, Lim SS. Inhibitory constituents of aldose reductase in the fruiting body of Phellinus linteus. Biol Pharm Bull. 2008;31(4):765–8.

    Article  CAS  PubMed  Google Scholar 

  137. Min GJ, Jeong EU, Yun BS, Kang HW. Chemical identification and antioxidant activity of phenolic compounds extracted from the fruiting body of “Hankyong Sanghwang”, Phellinus linteus KACC 93057P. J Mushroom. 2018;16(4):311–7. https://doi.org/10.14480/JM.2018.16.4.311.

    Article  Google Scholar 

  138. Chiu CH, Chang CC, Lin JJ, Chen CC, Chyau CC, Peng RY. Styrylpyrones from Phellinus linteus mycelia alleviate non-alcoholic fatty liver by modulating lipid and glucose metabolic homeostasis in high-fat and high-fructose diet-fed mice. Antioxidants. 2022;11(5):898. https://doi.org/10.3390/antiox11050898.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Chen W, Shen Y, Su H, Zheng X. Hispidin derived from Phellinus linteus affords protection against acrylamide-induced oxidative stress in Caco-2 cells. Chem Biol Interact. 2014;219:83–9. https://doi.org/10.1016/j.cbi.2014.05.010.

    Article  CAS  PubMed  Google Scholar 

  140. Kim DE, Kim B, Shin HS, Kwon HJ, Park ES. The protective effect of hispidin against hydrogen peroxide-induced apoptosis in H9c2 cardiomyoblast cells through Akt/GSK-3β and ERK1/2 signaling pathway. Exp Cell Res. 2014;327(2):264–75. https://doi.org/10.1016/j.yexcr.2014.07.037.

    Article  CAS  PubMed  Google Scholar 

  141. Park JM, Lee JS, Song JE, Sim YC, Ha SJ, Hong EK. Cytoprotective effect of hispidin against palmitate-induced lipotoxicity in C2C12 myotubes. Molecules. 2015;20(4):5456–67. https://doi.org/10.3390/molecules20045456.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Huang SY, Chang SF, Chau SF, Chiu SC. The protective effect of hispidin against hydrogen peroxide-induced oxidative stress in ARPE-19 cells via Nrf2 signaling pathway. Biomolecules. 2019;9(8):380. https://doi.org/10.3390/biom9080380.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Liang CH, Wu CY, Ho WJ, Liang ZC. Influences of carbon and nitrogen source addition, water content, and initial pH of grain medium on hispidin production of Phellinus linteus by solid-state fermentation. J Biosci Bioeng. 2020;130(6):616–21. https://doi.org/10.1016/j.jbiosc.2020.08.002.

    Article  CAS  PubMed  Google Scholar 

  144. Toopmuang P, Khamchum C, Punsuvon V. Detection and confirmation of hispolon in the mushroom Phellinus linteus. J Sci Asia. 2014;40:141–4. https://doi.org/10.2306/scienceasia1513-1874.2014.40.141.

    Article  CAS  Google Scholar 

  145. Lin CJ, Lien HM, Chang HY, Huang CL, Liu JJ, Chang YC, Chen CC, Lai CH. Biological evaluation of Phellinus linteus-fermented broths as anti-inflammatory agents. J Biosci Bioeng. 2014;118(1):88–93. https://doi.org/10.1016/j.jbiosc.2014.01.001.

    Article  CAS  PubMed  Google Scholar 

  146. Yang LY, Shen SC, Cheng KT, Subbaraju GV, Chien CC, Chen YC. Hispolon inhibition of inflammatory apoptosis through reduction of iNOS/NO production via HO-1 induction in macrophages. J Ethnopharmacol. 2014;156:61–72. https://doi.org/10.1016/j.jep.2014.07.054.

    Article  CAS  PubMed  Google Scholar 

  147. Wu MS, Chien CC, Cheng KT, Subbaraju GV, Chen YC. Hispolon suppresses LPS-or LTA-induced iNOS/NO production and apoptosis in BV-2 microglial cells. Am J Chin Med. 2017;45(08):1649–66. https://doi.org/10.1142/S0192415X17500896.

    Article  CAS  PubMed  Google Scholar 

  148. Lee EK, Koh EM, Kim YN, Song J, Song CH, Jung KJ. Immunomodulatory effect of hispolon on LPS-induced RAW264. 7 Cells and mitogen/alloantigen-stimulated spleen lymphocytes of mice. Pharmaceutics. 2022;14(7):1423. https://doi.org/10.3390/pharmaceutics14071423.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Peng S, Hou Y, Chen Z. Hispolon alleviates oxidative damage by stimulating the Nrf2 signaling pathway in PC12 cells. Arch Biochem Biophys. 2022;727: 109303. https://doi.org/10.1016/j.abb.2022.109303.

    Article  CAS  PubMed  Google Scholar 

  150. Hsieh MJ, Chien SY, Chou YE, Chen CJ, Chen J, Chen MK. Hispolon from Phellinus linteus possesses mediate caspases activation and induces human nasopharyngeal carcinomas cells apoptosis through ERK1/2, JNK1/2 and p38 MAPK pathway. Phytomedicine. 2014;21(12):1746–52. https://doi.org/10.1016/j.phymed.2014.07.013.

    Article  CAS  PubMed  Google Scholar 

  151. Arcella A, Oliva MA, Sanchez M, Staffieri S, Esposito V, Giangaspero F, Cantore G. Effects of hispolon on glioblastoma cell growth. Environ Toxicol. 2017;32(9):2113–23. https://doi.org/10.1002/tox.22419.

    Article  CAS  PubMed  Google Scholar 

  152. Masood M, Rasul A, Sarfraz I, Jabeen F, Liu S, Liu X, Wei W, Li J, Li X. Hispolon induces apoptosis against prostate DU145 cancer cells via modulation of mitochondrial and STAT3 pathways. Pak J Pharm Sci. 2019;32(5):2237–43.

    PubMed  Google Scholar 

  153. Liao KF, Chiu TL, Chang SF, Wang MJ, Chiu SC. Hispolon induces apoptosis, suppresses migration and invasion of glioblastoma cells and inhibits GBM xenograft tumor growth in vivo. Molecules. 2021;26(15):4497. https://doi.org/10.3390/molecules26154497.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Chao W, Deng JS, Huang SS, Li PY, Liang YC, Huang GJ. 3, 4-dihydroxybenzalacetone attenuates lipopolysaccharide-induced inflammation in acute lung injury via down-regulation of MMP-2 and MMP-9 activities through suppressing ROS-mediated MAPK and PI3K/AKT signaling pathways. Int Immunopharmacol. 2017;50:77–86. https://doi.org/10.1016/j.intimp.2017.06.014.

    Article  CAS  PubMed  Google Scholar 

  155. Liu YJ, Lyu JL, Kuo YH, Chiu CY, Wen KC, Chiang HM. The anti-melanogenesis effect of 3, 4-dihydroxybenzalacetone through downregulation of melanosome maturation and transportation in B16F10 and human epidermal melanocytes. Int J Mol Sci. 2021;22(6):2823. https://doi.org/10.3390/ijms22062823.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Chao W, Deng JS, Li PY, Liang YC, Huang GJ. 3, 4-Dihydroxybenzalactone suppresses human non-small cell lung carcinoma cells metastasis via suppression of epithelial to mesenchymal transition, ROS-mediated PI3K/AKT/MAPK/MMP and NFκB signaling pathways. Molecules. 2017;22(4):537. https://doi.org/10.3390/molecules22040537.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. Giridharan VV, Karupppagounder V, Arumugam S, Nakamura Y, Guha A, Barichello T, Quevedo J, Watanabe K, Konishi T, Thandavarayan RA. 3, 4-Dihydroxybenzalacetone (DBL) prevents aging-induced myocardial changes in senescence-accelerated mouse-prone 8 (SAMP8) mice. Cells. 2020;9(3):597. https://doi.org/10.3390/cells9030597.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  158. Kim JH, Kwon HJ, Kim BW. Protective effect of 4-(3, 4-dihydroxyphenyl)-3-buten-2-one from Phellinus linteus on naproxen-induced gastric antral ulcers in rats. J Microbiol Biotechnol. 2016;26(5):823–8. https://doi.org/10.4014/jmb.1511.11024.

    Article  CAS  PubMed  Google Scholar 

  159. Zhu PL, Fu XQ, Li JK, Tse AK, Guo H, Yin CL, Chou JY, Wang YP, Liu YX, Chen YJ, et al. Antrodia camphorata mycelia exert anti-liver cancer effects and inhibit STAT3 signaling in vitro and in vivo. Front Pharmacol. 2018;9:1–11. https://doi.org/10.3389/fphar.2018.01449.

    Article  CAS  Google Scholar 

  160. Wang JJ, Wu CC, Lee CL, Hsieh SL, Chen JB, Lee CI. Antimelanogenic, antioxidant and antiproliferative effects of Antrodia camphorata fruiting bodies on b16–f0 melanoma cells. PLoS ONE. 2017;12(1):1–13. https://doi.org/10.1371/journal.pone.0170924.

    Article  CAS  Google Scholar 

  161. Lin LT, Tai CJ, Su CH, Chang FM, Choong CY, Wang CK, Tai CJ. The ethanolic extract of Taiwanofungus camphoratus (Antrodia camphorata) induces cell cycle arrest and enhances cytotoxicity of cisplatin and doxorubicin on human hepatocellular carcinoma cells. Biomed Res Int. 2015;2015:1–9. https://doi.org/10.1155/2015/415269.

    Article  CAS  Google Scholar 

  162. Yeh CT, Rao YK, Yao CJ, Yeh CF, Li CH, Chuang SE, Luong JHT, Lai GM, Tzeng YM. Cytotoxic triterpenes from Antrodia camphorata and their mode of action in HT-29 human colon cancer cells. Cancer Lett. 2009;285(1):73–9. https://doi.org/10.1016/j.canlet.2009.05.002.

    Article  CAS  PubMed  Google Scholar 

  163. Hseu YC, Chang GR, Pan JY, Rajendran P, Mathew DC, Li M, Liao JW, Tzu W, Yang HL. Antrodia camphorata inhibits epithelial-to-mesenchymal transition by targeting multiple pathways in triple-negative breast cancers. J Cell Physiol. 2018;234(4):125–4139. https://doi.org/10.1002/jcp.27222.

    Article  CAS  Google Scholar 

  164. Yang HL, Lin KY, Juan YC, Kumar KJS, Way TD, Shen PC, Chen SC, Hseu YC. The anti-cancer activity of Antrodia camphorata against human ovarian carcinoma (SKOV-3) cells via modulation of HER-2/neu signaling pathway. J Ethnopharmacol. 2013;148(1):254–65. https://doi.org/10.1016/j.jep.2013.04.023.

    Article  PubMed  Google Scholar 

  165. Yang HL, Chen CS, Chang WH, Lu FJ, Lai YC, Chen CC, Hseu TH, Kuo CT, Hseu YC. Growth inhibition and induction of apoptosis in MCF-7 breast cancer cells by Antrodia camphorata. Cancer Lett. 2006;231(2):215–27. https://doi.org/10.1016/j.canlet.2005.02.004.

    Article  CAS  PubMed  Google Scholar 

  166. Yang PY, Hu DN, Liu FS. Cytotoxic effect and induction of apoptosis in human cervical cancer cells by Antrodia camphorata. Am J Chin Med. 2013;41(05):1169–80. https://doi.org/10.1142/s0192415x13500791.

    Article  PubMed  Google Scholar 

  167. Liu FS, Yang PY, Hu DN, Huang YW, Chen MJ. Antrodia camphorata induces apoptosis and enhances the cytotoxic effect of paclitaxel in human ovarian cancer cells. Int J Gynecol Cancer. 2011;21(7):1. https://doi.org/10.1097/igc.0b013e31821f742c.

    Article  CAS  Google Scholar 

  168. Peng CC, Chen KC, Peng RY, Su CH, Hsieh-Li HM. Human urinary bladder cancer T24 cells are susceptible to the Antrodia camphorata extracts. Cancer Lett. 2006;243(1):109–19. https://doi.org/10.1016/j.canlet.2005.11.021.

    Article  CAS  PubMed  Google Scholar 

  169. Kuo PL, Hsu YL, Cho CY, Ng LT, Kuo YH, Lin CC. Apoptotic effects of Antrodia cinnamomea fruiting bodies extract are mediated through calcium and calpain-dependent pathways in Hep 3B cells. Food Chem Toxicol. 2006;44(8):1316–26. https://doi.org/10.1016/j.fct.2006.02.009.

    Article  CAS  PubMed  Google Scholar 

  170. Hsu YL, Kuo YC, Kuo PL, Ng LT, Kuo YH, Lin CC. Apoptotic effects of extract from Antrodia camphorata fruiting bodies in human hepatocellular carcinoma cell lines. Cancer Lett. 2005;221(1):77–89. https://doi.org/10.1016/j.canlet.2004.08.012.

    Article  CAS  PubMed  Google Scholar 

  171. Song TY, Hsu SL, Yen GC. Induction of apoptosis in human hepatoma cells by mycelia of Antrodia camphorata in submerged culture. J Ethnopharmacol. 2005;100(1–2):158–67. https://doi.org/10.1016/j.jep.2005.02.043.

    Article  PubMed  Google Scholar 

  172. Rao YK, Fang SH, Tzeng YM. Evaluation of the anti-inflammatory and anti-proliferation tumoral cells activities of Antrodia camphorata, Cordyceps sinensis, and Cinnamomum osmophloeum bark extracts. J Ethnopharmacol. 2007;114(1):78–85. https://doi.org/10.1016/j.jep.2007.07.028.

    Article  CAS  PubMed  Google Scholar 

  173. Chang CY, Huang ZN, Yu HH, Chang LH, Li SL, Chen YP, Lee KY, Chuu JJ. The adjuvant effects of Antrodia Camphorata extracts combined with anti-tumor agents on multidrug resistant human hepatoma cells. J Ethnopharmacol. 2008;118(3):387–95. https://doi.org/10.1016/j.jep.2008.05.001.

    Article  CAS  PubMed  Google Scholar 

  174. Kuo YH, Lin CH, Shih CC, Yang CS. Antcin K, a triterpenoid compound from Antrodia camphorata, displays antidiabetic and antihyperlipidemic effects via glucose transporter 4 and AMP-activated protein kinase phosphorylation in muscles. Evid-Based Complementary Altern Med. 2016;2016:1–13. https://doi.org/10.1155/2016/4867092.

    Article  Google Scholar 

  175. Lin CH, Kuo YH, Shih CC. Eburicoic acid, a triterpenoid compound from Antrodia camphorata, displays antidiabetic and antihyperlipidemic effects in palmitate-treated C2C12 myotubes and in high-fat diet-fed mice. Int J Mol Sci. 2017;18(11):2314. https://doi.org/10.3390/ijms18112314.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  176. Lin CH, Hsiao LW, Kuo YH, Shih CC. Antidiabetic and antihyperlipidemic effects of sulphurenic acid, a triterpenoid compound from Antrodia camphorata, in streptozotocin-induced diabetic mice. Int J Mol Sci. 2019;20(19):4897. https://doi.org/10.3390/ijms20194897.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  177. Kuo YH, Lin CH, Shih CC. Antidiabetic and antihyperlipidemic properties of a triterpenoid compound, dehydroeburicoic acid, from Antrodia camphorata in vitro and in streptozotocin-induced mice. J Agric Food Chem. 2015;63(46):10140–51. https://doi.org/10.1021/acs.jafc.5b04400.

    Article  CAS  PubMed  Google Scholar 

  178. Kuo YH, Lin CH, Shih CC. Ergostatrien-3β-ol from Antrodia camphorata inhibits diabetes and hyperlipidemia in high-fat-diet treated mice via regulation of hepatic related genes, glucose transporter 4, and AMP-activated protein kinase phosphorylation. J Agric Food Chem. 2015;63(9):2479–89. https://doi.org/10.1021/acs.jafc.5b00073.

    Article  CAS  PubMed  Google Scholar 

  179. Zhang F, Cai H, Liu X, Chen F, Yang N, Hou Y, Ma Y, Luo Y, Liu J. Simultaneous determination of 19 fatty acids in Antrodia camphorata by derivatized GC–MS and evaluation of antioxidant activity of Antrodia camphorata crude oil. Arch Pharm Res. 2015. https://doi.org/10.1007/s12272-014-0533-z.

    Article  PubMed  Google Scholar 

  180. Kumar S, Chu FH, Hsieh HW, Liao JW, Li WH, Lin J, Shaw JF, Wang SY. Antroquinonol from ethanolic extract of mycelium of Antrodia cinnamomea protects hepatic cells from ethanol-induced oxidative stress through Nrf-2 activation. J Ethnopharmacol. 2011;136(1):168–77. https://doi.org/10.1016/j.jep.2011.04.030.

    Article  CAS  PubMed  Google Scholar 

  181. Chen JR, Yeh WJ, Tan HY, Yang HY. Antroquinonol attenuated abdominal and hepatic fat accumulation in rats fed an obesogenic diet. J Food Sci. 2019;84(9):2682–7. https://doi.org/10.1111/1750-3841.14746.

    Article  CAS  PubMed  Google Scholar 

  182. Chang YY, Liu YC, Kuo YH, Lin YL, Wu YHS, Chen JW, Chen YC. Effects of antrosterol from Antrodia camphorata submerged whole broth on lipid homeostasis, antioxidation, alcohol clearance, and anti-inflammation in livers of chronic-alcohol fed mice. J Ethnopharmacol. 2017;202:200–7. https://doi.org/10.1016/j.jep.2017.03.003.

    Article  CAS  PubMed  Google Scholar 

  183. Park DK, Park HJ. Ethanol extract of Antrodia camphorate grown on germinated brown rice suppresses inflammatory responses in mice with acute DSS-induced colitis. Evid Based Complementary Altern Med. 2013;2013:1–12. https://doi.org/10.1155/2013/914524.

    Article  Google Scholar 

  184. Kuo YH, Lin TY, You YJ, Wen KC, Sung PJ, Chiang HM. Antiinflammatory and antiphotodamaging effects of ergostatrien-3β-ol, isolated from Antrodia camphorata, on hairless mouse skin. Molecules. 2016;21(9):1213. https://doi.org/10.3390/molecules21091213.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  185. Tsai TC, Tung YT, Kuo YH, Liao JW, Tsai HC, Chong KY, Chen HL, Chen CM. Anti-inflammatory effects of Antrodia camphorata, a herbal medicine, in a mouse skin ischemia model. J Ethnopharmacol. 2015;159:113–21. https://doi.org/10.1016/j.jep.2014.11.015.

    Article  PubMed  Google Scholar 

  186. Tu PC, Jiang WP, Lin MK, Huang GJ, Li YJ, Kuo YH. Anti-inflammatory constituents of Antrodia camphorata on RAW 264.7 cells induced by polyinosinic-polycytidylic acid. Molecules. 2022;27(16):5320. https://doi.org/10.3390/molecules27165320.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  187. Chen YC, Chiu HL, Chao CY, Lin WH, Chao L, Huang GJ, Kuo YH. New anti-inflammatory aromatic components from Antrodia camphorata. Int J Mol Sci. 2013;14(3):4629–39. https://doi.org/10.3390/ijms14034629.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  188. Chen YJ, Tsao YC, Ho TC, Puc I, Chen CC, Perng GC, Lien HM. Antrodia cinnamomea suppress dengue virus infection through enhancing the secretion of interferon-alpha. Plants. 2022;11(19):2631. https://doi.org/10.3390/plants11192631.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  189. Lee IH, Huang RL, Chen CT, Chen HC, Hsu WC, Lu MK. Antrodia camphoratapolysaccharides exhibit anti-hepatitis B virus effects. FEMS Microbiol Lett. 2002;209(1):63–7. https://doi.org/10.1111/j.1574-6968.2002.tb11110.x.

    Article  CAS  PubMed  Google Scholar 

  190. My T, Ma C, Hattori M, Jin JS. Inhibitory effects of antrodins A-E from Antrodia cinnamomea and their metabolites on hepatitis C virus protease. Phytother Res. 2009;23(4):582–4. https://doi.org/10.1002/ptr.2657.

    Article  CAS  Google Scholar 

  191. Linden AG, Li S, Choi HY, Fang F, Fukasawa M, Uyeda K, Hammer RE, Horton JD, Engelking LJ, Liang G. Interplay between ChREBP and SREBP-1c coordinates postprandial glycolysis and lipogenesis in livers of mice. J Lipid Res. 2018;59(3):475–87. https://doi.org/10.1194/jlr.M081836.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  192. Szychowski KA, Skóra B, Pomianek T, Gmiński J. Inonotus obliquus—from folk medicine to clinical use. J Tradit Complement Med. 2021;11(4):293–302. https://doi.org/10.1016/j.jtcme.2020.08.003.

    Article  CAS  PubMed  Google Scholar 

  193. Wang J, Wang C, Li S, Li W, Yuan G, Pan Y, Chen H. Anti-diabetic effects of Inonotus obliquus polysaccharides in streptozotocin-induced type 2 diabetic mice and potential mechanism via PI3K-Akt signal pathway. Biomed Pharmacother. 2017;95:1669–77. https://doi.org/10.1016/j.biopha.2017.09.104.

    Article  CAS  PubMed  Google Scholar 

  194. Cha JY, Park SH, Heo JS, Cho YS. Effects of water extract from fermented chaga mushroom (Inonotus obliquus) on the proliferation of human cancer cell lines. J Life Sci. 2007;17(5):671–7. https://doi.org/10.5352/jls.2007.17.5.671.

    Article  Google Scholar 

  195. Youn MJ, Kim JK, Park S, Kim Y, Park C, Kim ES, Park KI, So HS, Park R. Potential anticancer properties of the water extract of Inonotus obliquus by induction of apoptosis in melanoma B16–F10 cells. J Ethnopharmacol. 2009;121(2):221–8. https://doi.org/10.1016/j.jep.2008.10.016.

    Article  PubMed  Google Scholar 

  196. Lee SH, Hwang HS, Yun JW. Antitumor activity of water extract of a mushroom, Inonotus obliquus, against HT-29 human colon cancer cells. Phytother Res. 2009;23(12):1784–9. https://doi.org/10.1002/ptr.2836.

    Article  PubMed  Google Scholar 

  197. Yuan X, Sun H, Liu Y, Shiroshita T, Kawano S, Takeshi S, Ma J, Zhang Z. Anti-cancer activity comparisons of aqueous extracts from Inonotus obliquus, Cordyceps militaris and Uncaria tomentosa in vitro and in vivo. J Pharmacogn Phytochem. 2014;2(6):19–25.

    Google Scholar 

  198. Lee HS, Kim EJ, Kim SH. Ethanol extract of Innotus obliquus (Chaga mushroom) induces G1 cell cycle arrest in HT-29 human colon cancer cells. Nurs Res Pract. 2015;9(2):111–6. https://doi.org/10.4162/nrp.2015.9.2.111.

    Article  Google Scholar 

  199. Sun Y, Yin T, Chen XH, Zhang G, Curtis RB, Lu ZH, Jiang JH. In vitro antitumor activity and structure characterization of ethanol extracts from wild and cultivated chaga medicinal mushroom, Inonotus obliquus (Pers.:Fr.) Pilát (Aphyllophoromycetideae). Int J Med Mushrooms. 2011;13(2):121–30. https://doi.org/10.1615/intjmedmushr.v13.i2.40.

    Article  CAS  PubMed  Google Scholar 

  200. Hu H, Zhang Z, Lei Z, Yang Y, Sugiura N. Comparative study of antioxidant activity and antiproliferative effect of hot water and ethanol extracts from the mushroom Inonotus obliquus. J Biosci Bioeng. 2009;107(1):42–8. https://doi.org/10.1016/j.jbiosc.2008.09.004.

    Article  CAS  PubMed  Google Scholar 

  201. Nakajima Y, Nishida H, Matsugo S, Konishi T. Cancer cell cytotoxicity of extracts and small phenolic compounds from Chaga [Inonotus obliquus (persoon) Pilat. J Med Food. 2009;12(3):501–7. https://doi.org/10.1089/jmf.2008.1149.

    Article  CAS  PubMed  Google Scholar 

  202. Fan L, Ding S, Ai L, Deng K. Antitumor and immunomodulatory activity of water-soluble polysaccharide from Inonotus obliquus. Carbohyd Polym. 2012;90(2):870–4. https://doi.org/10.1016/j.carbpol.2012.06.013.

    Article  CAS  Google Scholar 

  203. Baek J, Roh HS, Baek KH, Lee S, Lee S, Song SS, Kim KH. Bioactivity-based analysis and chemical characterization of cytotoxic constituents from Chaga mushroom (Inonotus obliquus) that induce apoptosis in human lung adenocarcinoma cells. J Ethnopharmacol. 2018;224:63–75. https://doi.org/10.1016/j.jep.2018.05.025.

    Article  CAS  PubMed  Google Scholar 

  204. Ye X, Wu K, Xu L, Cen Y, Ni J, Chen J, Zheng W, Liu W. Methanol extract of Inonotus obliquus improves type 2 diabetes mellitus through modifying intestinal flora. Front Endocrinol. 2023;13:1103972. https://doi.org/10.3389/fendo.2022.1103972.

    Article  Google Scholar 

  205. Lee JH, Hyun CK. Insulin-sensitizing and beneficial lipid-metabolic effects of the water-soluble melanin complex extracted from Inonotus obliquus. Phytother Res. 2014;28(9):1320–8. https://doi.org/10.1002/ptr.5131.

    Article  CAS  PubMed  Google Scholar 

  206. Cha JY, Jun BL, Lee CH, Yoo KS, Moon JC, Cho YS. Hypoglycemic and antioxidative effects of fermented Chaga mushroom (Inonotus obliquus) on streptozotocin-induced diabetic rats. J Life Sci. 2005;15(5):809–18. https://doi.org/10.5352/JLS.2005.15.5.809.

    Article  Google Scholar 

  207. Yang M, Hu D, Cui Z, Li H, Man C, Jiang Y. Lipid-lowering effects of Inonotus obliquus polysaccharide in vivo and in vitro. Foods. 2021;10(12):3085. https://doi.org/10.3390/foods10123085.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  208. Zhang Z, Liang X, Tong L, Lv Y, Yi H, Gong P, Tian X, Cui Q, Liu T, Zhang L. Effect of Inonotus obliquus (Fr) Pilat extract on the regulation of glycolipid metabolism via PI3K/Akt and AMPK/ACC pathways in mice. J Ethnopharmacol. 2021;273:1–9. https://doi.org/10.1016/j.jep.2021.113963.

    Article  CAS  Google Scholar 

  209. Wu D, Zhang Y, Wang D, Liu T, Zhang S, Yang C. Research of Inonotus obliquus oligosaccharide in prevention of hyperlipidemia. J Food Qual. 2021;2021:1–11. https://doi.org/10.1155/2021/1174452.

    Article  CAS  Google Scholar 

  210. Park YK, Lee HB, Jeon EJ, Jung HS, Kang MH. Chaga mushroom extract inhibits oxidative DNA damage in human lymphocytes as assessed by comet assay. BioFactors. 2004;21(1–4):109–12. https://doi.org/10.1002/biof.552210120.

    Article  CAS  PubMed  Google Scholar 

  211. Cui Y, Kim DS, Park KC. Antioxidant effect of Inonotus obliquus. J Ethnopharmacol. 2005;96(1–2):79–85. https://doi.org/10.1016/j.jep.2004.08.037.

    Article  PubMed  Google Scholar 

  212. Huang S, Ding S, Fan L. Antioxidant activities of five polysaccharides from Inonotus obliquus. Int J Biol Macromol. 2012;50(5):1183–7. https://doi.org/10.1016/j.ijbiomac.2012.03.019.

    Article  CAS  PubMed  Google Scholar 

  213. Hu Y, Sheng Y, Yu M, Li K, Ren G, Xu X, Qu J. Antioxidant activity of Inonotus obliquus polysaccharides and its amelioration for chronic pancreatitis in mice. Int J Biol Macromol. 2016;87:348–56. https://doi.org/10.1016/j.ijbiomac.2016.03.006.

    Article  CAS  PubMed  Google Scholar 

  214. Xu L, Sang R, Yu Y, Li J, Ge B, Zhang X. The polysaccharide from Inonotus obliquus protects mice from Toxoplasma gondii-induced liver injury. Int J Biol Macromol. 2019;125:1–8. https://doi.org/10.1016/j.ijbiomac.2018.11.114.

    Article  CAS  PubMed  Google Scholar 

  215. Hong KB, Noh DO, Park Y, Suh HJ. Hepatoprotective activity of water extracts from Chaga Medicinal Mushroom, Inonotus obliquus (higher basidiomycetes) against tert-butyl hydroperoxide-induced oxidative liver injury in primary cultured rat hepatocytes. Int J Med Mushrooms. 2015;17(11):1069–76. https://doi.org/10.1615/intjmedmushrooms.v17.i11.70.

    Article  PubMed  Google Scholar 

  216. Parfenov AA, Vyshtakalyuk AB, Sysoeva MA, Sysoeva EV, Latipova AD, Gumarova LF, Zobov VV. Hepatoprotective effect of Inonotus obliquus melanins: in vitro and in vivo studies. BioNanoScience. 2019;9(2):528–538. https://doi.org/10.1007/s12668-019-0595-y

  217. Debnath T, Park S, Kim D, Jo J, Lim B. Anti-oxidant and anti-inflammatory activities of Inonotus obliquus and germinated brown rice extracts. Molecules. 2013;18(8):9293–304. https://doi.org/10.3390/molecules18089293.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  218. Javed S, Mitchell K, Sidsworth D, Sellers SL, Reutens-Hernandez J, Massicotte HB, Egger KN, Lee CH, Payne GW. Inonotus obliquus attenuates histamine-induced microvascular inflammation. PLoS ONE. 2019;14(8):1–14. https://doi.org/10.1371/journal.pone.0220776.

    Article  CAS  Google Scholar 

  219. Shibnev VA, Mishin DV, Garaev TM, Finogenova NP, Botikov AG, Deryabin PG. Antiviral activity of Inonotus obliquus fungus extract towards infection caused by hepatitis C virus in cell cultures. Bull Exp Biol Med. 2011;151(5):612–4. https://doi.org/10.1007/s10517-011-1395-8.

    Article  CAS  PubMed  Google Scholar 

  220. Teplyakova T, Pyankov O, Safatov A, Ovchinnikova A, Kosogova T, Skarnovich M, Filippova E, Poteshkina A. Water extract of the Chaga Medicinal Mushroom Inonotus obliquus (Agaricomycetes) inhibits the SARS-Cov-2 replication at Vero E6 and vero cell cultures experiments. Int J Med Mushrooms. 2022;24(2):23–30. https://doi.org/10.1615/intjmedmushrooms.2021042012.

    Article  PubMed  Google Scholar 

  221. Eid JI, Das B, Al-Tuwaijri MM, Basal WT. Targeting SARS-CoV-2 with Chaga mushroom: an in silico study toward developing a natural antiviral compound. Food Sci Nutr. 2021;9(12):6513–23. https://doi.org/10.1002/fsn3.2576.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  222. Kapp K, Püssa T, Vuorela H, Välimaa H. Antiviral effect of Inonotus obliquus (Pers.:Fr.) Pilat extract against herpes simplex virus type 1 in vitro. Plant Med Int Open. 2017;4(S01):S1–202. https://doi.org/10.1055/s-0037-1608419.

    Article  Google Scholar 

  223. Razumov IA, Kazachinskaia EI, Puchkova LI, Kosogorova TA, Gorbunova IA, Loktev VB, Tepliakova TV. Protective activity of aqueous extracts from higher mushrooms against Herpes simplex virus type-2 on albino mice mode. Antibiotiki i khimioterapiia Antibiot Chemoterapy sic. 2013;58(9–10):8–12.

    CAS  Google Scholar 

  224. Tian J, Hu X, Liu D, Wu H, Qu L. Identification of Inonotus obliquus polysaccharide with broad-spectrum antiviral activity against multi-feline viruses. Int J Biol Macromol. 2017;95:160–7. https://doi.org/10.1016/j.ijbiomac.2016.11.054.

    Article  CAS  PubMed  Google Scholar 

  225. Thammavong S, Phadungkit M, Laovachirasuwan P, Naksuwankul K, Saentaweesuk W, Silsirivanit A, Wongkham S. Antioxidant and cytotoxic activity of Phellinus mushrooms from northeast Thailand. Res J Pharmacogn. 2020;8(1):91–9. https://doi.org/10.22127/rjp.2020.251089.1632.

    Article  CAS  Google Scholar 

  226. Lee WY, Hsu KF, Chiang TA, Chen CJ. Phellinus Linteus extract induces autophagy and synergizes with 5-fluorouracil to inhibit breast cancer cell growth. Nutr Cancer. 2015;67(2):275–84. https://doi.org/10.1080/01635581.2015.989374.

    Article  CAS  PubMed  Google Scholar 

  227. Park HJ, Choi SY, Hong SM, Hwang SG, Park DK. The ethyl acetate extract of Phellinus linteus grown on germinated brown rice induces G0 /G1 cell cycle arrest and apoptosis in human colon carcinoma HT29 cells. Phytother Res. 2009;24(7):1019–26. https://doi.org/10.1002/ptr.3064.

    Article  Google Scholar 

  228. Jeon TI, Jung CH, Cho JY, Park DI, Moon JH. Identification of an anticancer compound against HT-29 cells from Phellinus linteus grown on germinated brown rice. Asian Pac J Trop Biomed. 2013;3(10):785–9. https://doi.org/10.1016/s2221-1691(13)60156-2.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  229. Šeklić DS, Jovanović MM, Virijević K, Grujić J, Živanović M, Marković SD. Effects of edible mushrooms Phellinus linteus and Lentinus edodes methanol extracts on colorectal cancer cell lines. Biol Life Sci Forum. 2022;6(1):85. https://doi.org/10.3390/foods2021-11055.

    Article  Google Scholar 

  230. Park HJ, Park JB, Lee SJ, Song M. Phellinus linteus grown on germinated brown rice increases cetuximab sensitivity of KRAS-mutated colon cancer. Int J Mol Sci. 2017;18(8):1746. https://doi.org/10.3390/ijms18081746.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  231. Konno S, Chu K, Feuer N, Phillips J, Choudhury M. Potent anticancer effects of bioactive mushroom extracts (Phellinus linteus) on a variety of human cancer cells. J Clin Med Res. 2015;7(2):76–82. https://doi.org/10.14740/jocmr1996w.

    Article  PubMed  Google Scholar 

  232. Lu TL, Huang GJ, Lu TJ, Wu JB, Wu CH, Yang TC, Iizuka A, Chen YF. Hispolon from Phellinus linteus has antiproliferative effects via MDM2-recruited ERK1/2 activity in breast and bladder cancer cells. Food Chem Toxicol. 2009;47(8):2013–21. https://doi.org/10.1016/j.fct.2009.05.023.

    Article  CAS  PubMed  Google Scholar 

  233. Liu Y, Jia Z, Dong L, Wang R, Qiu G. A randomized pilot study of atractylenolide I on gastric cancer cachexia patients. Evid Based Complementary Altern Med. 2008;5(3):337–44. https://doi.org/10.1093/ecam/nem031.

    Article  Google Scholar 

  234. Kim DH, Yang BK, Sang DJ, Nam JH, Das S, Jong WY, Choi JW, Lee YS, Song CH. A preliminary study on the hypoglycemic effect of the exo - polymers produced by five different medicinal mushrooms. J Microbiol Biotechnol. 2001;11(1):167–71.

    CAS  Google Scholar 

  235. Lee JH, Lee JS, Kim JR, Jung WC, Lee KE, Lee SY, Hong EK. Hispidin isolated from Phellinus linteus protects against hydrogen peroxide-induced oxidative stress in pancreatic MIN6N β-Cells. J Med Food. 2011;14(11):1431–8. https://doi.org/10.1089/jmf.2010.1493.

    Article  CAS  PubMed  Google Scholar 

  236. Kozarski M, Klaus A, Niksic M, Jakovljevic D, Helsper JPFG, Van Griensven LJLD. Antioxidative and immunomodulating activities of polysaccharide extracts of the medicinal mushrooms Agaricus bisporus, Agaricus brasiliensis, Ganoderma lucidum and Phellinus linteus. Food Chem. 2011;129(4):1667–75. https://doi.org/10.1016/j.foodchem.2011.06.029.

    Article  CAS  Google Scholar 

  237. Wang H, Wu G, Park HJ, Jiang PP, Sit WH, van Griensven LJ, Wan JMF. Protective effect of Phellinus linteus polysaccharide extracts against thioacetamide-induced liver fibrosis in rats: a proteomics analysis. Chin Med. 2012;7(23):2–9. https://doi.org/10.1186/1749-8546-7-23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  238. Jeon TI, Hwang SG, Lim BO, Park DK. Extracts of Phellinus linteus grown on germinated brown rice suppress liver damage induced by carbon tetrachloride in rats. Biotech Lett. 2003;25(24):2093–6. https://doi.org/10.1023/b:bile.0000007071.28105.c1.

    Article  CAS  Google Scholar 

  239. Kim SH, Lee HS, Lee S, Cho J, Ze K, Sung J, Kim YC. Mycelial culture of Phellinus linteus protects primary cultured rat hepatocytes against hepatotoxins. J Ethnopharmacol. 2004;95(2–3):367–72. https://doi.org/10.1016/j.jep.2004.08.007.

    Article  CAS  PubMed  Google Scholar 

  240. Shin MR, Lee JA, Min JK, Park HJ, Park W, Byeong SBS, Roh SS. Protective effects of Phellinus linteus mycelium on the development of osteoarthritis after monosodium Iodoacetate injection. Evid Based Complementary Altern Med. 2020;2020:7240858. https://doi.org/10.1155/2020/7240858.

    Article  Google Scholar 

  241. Song M, Park HJ. Anti-inflammatory effect of Phellinus linteus grown on germinated brown rice on dextran sodium sulfate-induced acute colitis in mice and LPS-activated macrophages. J Ethnopharmacol. 2014;154(2):311–8. https://doi.org/10.1016/j.jep.2013.12.059.

    Article  PubMed  Google Scholar 

  242. Park HJ. Anti-allergic and anti-inflammatory activity of Phellinus linteus grown on Panax ginseng. Food Sci Biotechnol. 2017;26:467–72. https://doi.org/10.1007/s10068-017-0064-8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  243. Huang GJ, Huang SS, Deng JS. Anti-inflammatory activities of inotilone from Phellinus linteus through the inhibition of MMP-9, NF-κB, and MAPK activation in vitro and in vivo. PLoS ONE. 2012;7(5): e35922. https://doi.org/10.1371/journal.pone.0035922.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  244. Lee D. Screening of Phellinus linteus, a medicinal mushroom, for anti-viral activity. J Korean Soc Appl Biol Chem. 2011;54(3):475–8. https://doi.org/10.3839/jksabc.2011.073.

    Article  Google Scholar 

  245. Chalertpet K, Sangkheereeput T, Somjit P, Bankeeree W, Yanatatsaneejit P. Effect of Smilax spp. and Phellinus linteus combination on cytotoxicity and cell proliferation of breast cancer cells. BMC Complementary Med Therapies. 2023;23(1):177. https://doi.org/10.1186/s12906-023-04003-x.

    Article  CAS  Google Scholar 

  246. Sonawane H, Arya S, Ghole V, Apte K, Shelke D, Chaskar M. Hypoglycemic and anticataract activity of crude exopolysaccharides of medicinal mushroom Phellinus badius on streptozotocin-induced diabetic rats and goat eye lenses respectively. Bioact Carbohydr Diet Fibre. 2020;24:100241. https://doi.org/10.1016/j.bcdf.2020.100241.

    Article  CAS  Google Scholar 

  247. Chen CC, Li IC, Lin TW, Chang HL, Lin WH, Shen YC. Efficacy and safety of oral Antrodia cinnamomea mycelium in mildly hypertensive adults: a randomized controlled pilot clinical study. Eur J Integr Med. 2016;8(5):654–60. https://doi.org/10.1016/j.eujim.2016.06.001.

    Article  Google Scholar 

  248. Yen YT, Park JH, Kang SW, Su T, Cheng H, Wen WC, Lin SY, Tai Y, Chen PN, Tsai SC. Clinical benefits of golden-Antrodia Camphorata containing antroquinonol in liver protection and liver fat reduction after alcoholic hepatitis. Front Pharmacol. 2022;13: 757494. https://doi.org/10.3389/fphar.2022.757494.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  249. Tsai MY, Hung YC, Chen YH, Chen YH, Huang YC, Kao CW, Su YL, Chiu HHE, Rau KM. A preliminary randomised controlled study of short-term Antrodia cinnamomea treatment combined with chemotherapy for patients with advanced cancer. BMC Complement Altern Med. 2016;16:1–10. https://doi.org/10.1186/s12906-016-1312-9.

    Article  CAS  Google Scholar 

  250. Dosychev EA, Bystrova VN. Treatment of psoriasis using “Chaga” fungus preparations. Vestnik Dermatologii i Venerologii. 1973;47(5):79–83.

    CAS  PubMed  Google Scholar 

  251. Fedotov AA, Yu IR. Effect of Befungin on the central nervous system in patients with peptic ulcers. Klinicheskaya Meditsina (Moscow). 1981;59(7):22–5.

    CAS  Google Scholar 

  252. Ku YH, Kang JH. Efficacy of Phellinus linteus extract on immunity enhancement: A CONSORT-randomized, double-blind, placebo-controlled pilot trial. Medicine. 2022;101(40):30829. https://doi.org/10.1097/md.0000000000030829.

    Article  Google Scholar 

  253. Ryu HY, Lee H, Kang JH. A pilot clinical study of the efficacy and safety of Phellinus Linteus (Sanghuang) extract treatment for knee osteoarthritis. J Acupuncture Res. 2022;39(2):115–21. https://doi.org/10.13045/jar.2022.00010.

    Article  Google Scholar 

  254. Sung HL, Ho KH, Chang MK, Woo JL. Potential impact of Phellinus linteus on adherence to adjuvant treatment after curative resection of pancreatic ductal adenocarcinoma: outcomes of a propensity score-matched analysis. Integr Cancer Ther. 2018;18:1534735418816825. https://doi.org/10.1177/1534735418816825.

    Article  Google Scholar 

  255. Yang HL, Kumar KJS, Hseu YC. Multiple molecular targets of Antrodia camphorata: a suitable candidate for ovarian cancer chemoprevention. In: Aft R, editor. Targeting new pathways and cell death in breast cancer. InTech. 2012. https://doi.org/10.5772/21789.

  256. Huang J, Yadav VK, Srivastava P, Wu AT, Huynh TT, Wei PL, Huang CYF, Huang TH. Antrodia cinnamomea enhances chemo-sensitivity of 5-FU and suppresses colon tumorigenesis and cancer stemness via up-regulation of tumor suppressor miR-142-3p. Biomolecules. 2019;9(8):306. https://doi.org/10.3390/biom9080306.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  257. Kim YR. Immunomodulatory activity of the water extract from medicinal mushroom Inonotus obliquus. Mycobiology. 2005;33(3):158–62. https://doi.org/10.4489/myco.2005.33.3.158.

    Article  PubMed  PubMed Central  Google Scholar 

  258. Hu Y, Teng C, Yu S, Wang X, Liang J, Bai X, Dong L, Song T, Yu M, Qu J. Inonotus obliquus polysaccharide regulates gut microbiota of chronic pancreatitis in mice. AMB Express. 2017;7:1–11. https://doi.org/10.1186/s13568-017-0341-1.

    Article  CAS  Google Scholar 

  259. Xin X, Qu J, Veeraraghavan VP, Mohan SK, Gu K. Assessment of the gastroprotective effect of the chaga medicinal mushroom, Inonotus obliquus (Agaricomycetes), against the gastric mucosal ulceration induced by ethanol in experimental rats. Int J Med Mushrooms. 2019;21(8):805–16. https://doi.org/10.1615/intjmedmushrooms.2019031154.

    Article  PubMed  Google Scholar 

  260. Clinicaltrials.gov. U.S. National Library of Medicine. https://clinicaltrials.gov/ (nd). Accessed 17 Aug 2023.

  261. Kim GY, Kim SH, Hwang SY, Kim HY, Park YM, Park SK, Lee MK, Lee SH, Lee TH, Lee JD. Oral administration of proteoglycan isolated from Phellinus linteus in the prevention and treatment of collagen-induced arthritis in mice. Biol Pharm Bull. 2003;26(6):823–31. https://doi.org/10.1248/bpb.26.823.

    Article  PubMed  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

This research did not receive any specific grant from funding agencies in the public, commercial, or not-for-profit sectors.

Author information

Authors and Affiliations

Authors

Contributions

PT, TK, XT, SM and AC contributed to the conceptualization and framework of the manuscript. PT, TK, XT, and SM wrote the manuscript with input from CY, YY, LL, SY, and AC. All authors contributed to manuscript revision, read, and approved the final submitted version.

Corresponding author

Correspondence to Adeline Yoke Yin Chia.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

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

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Tee, P.Y.E., Krishnan, T., Cheong, X.T. et al. A review on the cultivation, bioactive compounds, health-promoting factors and clinical trials of medicinal mushrooms Taiwanofungus camphoratus, Inonotus obliquus and Tropicoporus linteus. Fungal Biol Biotechnol 11, 7 (2024). https://doi.org/10.1186/s40694-024-00176-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40694-024-00176-3

Keywords