Next Article in Journal
Fluorimetric Detection of Insulin Misfolding by Probes Derived from Functionalized Fluorene Frameworks
Previous Article in Journal
A Brief Review of Recent Theoretical Advances in Fe-Based Catalysts for CO2 Hydrogenation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Research Progress of Bioactive Components in Sanghuangporus spp.

Department of Regenerative Medicine, School of Pharmaceutical Sciences, Jilin University, 1266 Fujin Road, Changchun 130021, China
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(6), 1195; https://doi.org/10.3390/molecules29061195
Submission received: 15 January 2024 / Revised: 15 February 2024 / Accepted: 25 February 2024 / Published: 7 March 2024

Abstract

:
The species in Sanghuangporus are a group of edible mushrooms with a long history of oral use in East Asia as a health-improvement method. They should be classified under the genus Sanghuangporus rather than mistakenly in Phellinus or Inonotus. The major components in this genus consist of polysaccharides, polyphenols, triterpenoids, and flavonoids, all of which exist in the fruiting bodies and mycelia. For extraction, studies have shown methods using hot water, ethanol, DES solvent, and alkaline, followed by purification methods including traditional anion column, Sevag solution, macroporous resin, and magnetic polymers. Proven by modern medical technology, these components possess promising anti-inflammatory, antioxidative, antitumor, and immunoregulation effects; additionally, they have health-improving effects including pulmonary protection, hypoglycemic properties, sleep improvement, gout mitigation, antiaging, neuroprotection, and muscle-strengthening abilities. Several toxicity studies have revealed their safety and recommend a dose of 1 g/kg for mice. As a newly emerged concept, functional food can provide not only life-sustaining nutrients but also some health-improving effects. In conclusion, we substantiate Sanghuang as a functional food by comprehensively presenting information on extraction and purification methods, component medical and structural properties, and nontoxicity, hoping to benefit the development of Sanghuang species as a group of functional food.

1. Introduction

Sanghuangporus spp., also known as Phellinus or Inonotus in some of the literature and some areas worldwide, is a perennial genus of edible mushrooms widely distributed in boreal forests in East Asia, especially in subtropical and tropical areas, holding profound pharmaceutical potential [1,2,3,4,5,6,7,8]. These medicinal mushrooms, as illustrated in Figure 1, have been popular in oral use since ancient times in countries such as China, Korea, and Japan [4,5,9,10,11,12,13,14,15,16,17,18,19]. The first mention could date back to approximately 2000 years ago, as the name ‘Sanger’ appeared in Shen Nong’s Materia Medica for treating gynecologic tumors taking the forms of abdominal lumps [2,20,21,22,23]. Then, the word ‘Sanghuang’ started to appear in a contemporary medical book called ‘Yao Xing Lun’ in the Tang Dynasty as an oral treatment for what people at the time believed to be cervical cancer, indicated by abnormal vaginal bleeding [14,20,21]. Sanghuang was also mentioned in the Compendium of Materia Medica in the Ming Dynasty as a traditional medicinal mushroom taken orally to eliminate the body’s toxins [2,21]. Entering modern times, pharmaceutical research has demonstrated multiple functions of Sanghuangporus spp., including the ability to improve blood circulation, scavenge free radicals, demonstrate anti-inflammatory effects, anti-carcinogenesis, antioxidative abilities, immunomodulatory effects, antidiabetic activities, antimicrobial abilities, and antiaging effects [2,4,9,14,20,21,22,24,25,26,27,28,29]. Some research has also provided evidence of antigout, sleep-improving, neuroprotective, and hepatoprotective activities [3,30,31,32]. Despite their prolonged use in history, research progress on the application of these mushrooms has been slow, mainly due to the following factors. First of all, their classification has been chaotic, with multiple studies presenting their own phylogenetic results. Secondly, few studies have comprehensively gathered key information such as their medicinal properties, extraction, and purification methods. These factors have impeded the scientific use of traditional fungi, preventing its widespread use by more people as a cheap but valuable nutrition source.
The dominant classification for the ‘Sanghuang’ species is as follows: Phylum Basidiomycetes, Class Hymenomycetes, Order Polyporales, Family Polyporaceae, and Genus Sanghuangporus [25,33]. Over the past few years, some studies have shown that the true Sanghuangporus genus contains 15 named species, including the most commonly used three species: S. baumii, S. vaninii, and S. sanghuang. These inhabit Syringa, Morus, and Populus, respectively [25,33]. Despite the fact that not all studies refresh their classifications, this information could still assist future research in providing clearer results with specific species classifications. To comprehensively review information from various sources, this article will include research results that clarify Sanghuangporus and those that still regard Sanghuangporus as Phellinus or Inonotus. Throughout this article, the term ‘Sanghuangporus’ should be comprehended to refer to formal members of the genus Sanghuangporus, and the terms ‘the Sanghuang species’ and ‘Sanghuang’ are used to generally refer to the species mentioned above.
The general process of studying the Sanghuang species often starts from extracting the desired components. To extract them from the Sanghuang species, the fruiting bodies are typically dried and turned into powder, followed by hot water extraction or ethanol extraction, depending on whether the target material comprises water-soluble polysaccharides or other ethanol-soluble components [2,4,34,35,36,37]. These two methods, often assisted by ultrasonic devices, are widely used due to their undeniable advantages: they are cheap and easy to perform [2,13,14,21,32,37,38,39,40,41,42]. But they are also marked by disadvantages, such as long extraction times, low purity or yield, and possible contamination of the environment, which have given rise to some new yet more expensive methods, such as the deep eutectic solvent (DES) method [43]. The mycelia are often acquired through liquid fermentation [15], after which they are used to extract single components, total components, or used as mycelia extracts using the methods mentioned above [3,32,37,44,45,46]. These extracts can be used to study their effects as a whole or further purified using traditional columns, synthetic polymers, or microporous resin, which are discussed in detail in the purification section. Some studies even include structural identification processes for their purified polysaccharides, using methods like infrared (IR) spectrum, methylation analysis, and nuclear magnetic resonance (NMR) spectroscopy [47,48,49]. In the end, the majority of studies focus on the medical properties of these extracts.
Their fruiting bodies and mycelia contain, yet differ in, multiple bioactive components, including polysaccharides, flavonoids, terpenoids, polyphenols, proteins, or saturated fatty acids [5,30,32,47,50]. The fruiting bodies contain mostly water-soluble polysaccharides, whose main monosaccharide composition includes mannose, galactose, glucose, fructose, and xylose [50]. Mycelia, conversely, demonstrate higher levels of polysaccharides, proteins, and other bioactive content than the fruiting bodies [15,50]. The compounds in these two parts provide Sanghuang with many valuable functions, including antitumor activities and free-radical scavenging abilities [6,9,13,16,20,25,26,28,35,37,51,52,53,54]. Polysaccharides are reported to be the major medical component and possess anti-inflammatory, antioxidative, antitumor, diabetes-ameliorating, sleep-improving, and neuroprotective effects, often demonstrated through pathway regulation, free-radical scavenging, and metabolic regulation [3,10,15,32,34,50,55,56,57,58,59]. However, scientists have noticed other components with indispensable medical properties. Hispidin, a polyphenol compound, has been shown by multiple studies to have anticancer and antiviral effects and free-radical scavenging and anti-inflammatory activities [6,21,26,34,35,53,60,61,62,63]. Morin, a flavonoid rarely found in species, possesses significant medical properties like cartilage protection [64,65]. Even some biodegradable fungal pigments could exhibit anticancer, antimicrobial, or antiobesity effects [66].
“Functional foods”, a newly emerged definition in recent years, focuses on a specific kind of food that not only provides life-sustaining nutrition but also improves health or treats diseases [67,68,69,70]. Despite being effective in small doses, most modern medicines have severe side effects, such as organ toxicity or shortening the life span, thus stating the urge to find better supplements or substitutes. Many kinds of mushrooms have now been considered functional foods due to their attractive taste, multiple kinds of beneficial components, and their nutritional value [67,68,69,70]. The secondary metabolites produced by or extracted from them, such as polysaccharides or flavonoids, provide the human body with necessary nutrition and contribute to health improvements, including antioxidant, anticancer, or immunoregulation [71]. Despite the lack of evidence proving Sanghuang is a functional food, its crucial advantages in health improvement and potential for therapeutic advancement have already pointed out the connection [70].
In this review, we summarize the progress in research on the structural and conformational properties of several polysaccharides in hopes of enlightening future research on structure–activity relationships. We conclude and compare different extraction methods and introduce some newly emerging purification methods in detail. Then, we elaborate on the biological activities of Sanghuang components and their possible mechanisms, providing a comprehensive view of how Sanghuangporus might be considered a functional food or be helpful in disease treatment.

2. Phylogenetic Progress

Species in Sanghuangporus have long been classified as Phellinus linteus, Phellinus baumii, Phellinus igniarius, or Inonotus linteus by many researchers from China and adjacent countries [6,7,15,16,17,19,30,40,53,60,72,73,74,75,76,77,78,79]. In 2021, Shen et al. presented concrete proof for such misclassification in a study where they revised the ITS sequence in Genbank and clarified some taxonomical mistakes, such as redetermining I. baumii and I. vaninii as S. baumii and S. vaninii, respectively [80]. These results provide tremendous aid in future research and enlarge the reference base of Sanghuangporus by suggesting that some misclassified species should be Sanghuangporus. Misclassification might originate from the following factors: first, a growing interest of non-taxonomists in the medicinal properties of Sanghuang and some errors in taxonomic revisions; second, translating the traditional names in Chinese or Korean into other languages without checking their classification; third, identifying species based only on their morphological characteristics, which can be misleading because of hybridization or convergent evolution, rather than taking genomic analysis into account, which is unique to each species and thus valuable for classification when building phylogenetic trees based on DNA sequencing [80]. This phenomenon confounded some researchers and caused studies on their pharmacological benefits to be rare, thereby hampering further medical applications [30,33]. Therefore, many studies are now focusing on Sanghuang’s taxonomy and building more accurate phylogenetic networks by drawing phylogenetic trees to reveal how closely species are related [33,52,81,82,83]. A phylogenetic tree is built using the standard method, namely phylogenetic systematics, by grouping different species based on shared characteristics, which should suggest uniformity among trees. Yet, the drawing process is subjective due to different understandings of species relationships, thereby creating different phylogenetic trees on the same topic [83].
Over the past few years, researchers have confirmed a reasonable classification of the ‘Sanghuang’ species: Phylum Basidiomycetes, Class Hymenomycetes, Order Polyporales, Family Polyporaceae, and Genus Sanghuangporus, and that the valid Sanghuangporus genus contains 15 named species, including the most commonly used three species, S. baumii, S. vaninii, and S. sanghuang, which inhabit Syringa, Morus, and Populus, respectively [25,33]. To prove that the Sanghuangporus exists, Han et al. revised 39 sequences from Genbank and argued that many Phellinus species, such as P. linteus, were S. sanghuang and S. baumii [7]. Another group in 2018 conducted a complete mitochondrial genome analysis of S. sanghuang, proving that Phellinus was not the taxonomically correct name [74,75,76]. These results brought up the importance of clarifying the Sanghuang species. In 2019, Zhu et al. dug deeper and studied their biodiversity, divergence, and phylogeny, after which they expanded the Sanghuangporus to four main clades by comparing their morphological similarities, hosts, and living environments [84,85]. Biogeographical estimation showed that the approximate divergence time of the 13 species in their study was in Oligocene in Northeast Asia, possibly caused by global cooling and a gradual replacement of tropical forests by grass [85]. Despite slight variances between results, establishing the Genus Sanghuangporus has been widely accepted, suggesting a clearer future in Sanghuang research.

3. Extraction Methods

3.1. Hot Water Extraction

This traditional method uses hot water as a solvent to boil certain plant parts for hours, usually followed by lyophilization, purification, and adding ethanol or methanol in a few studies for precipitation of crude polysaccharides [1,4,10,11,14,22,45,57,86,87,88,89,90,91,92,93,94]. The phenol sulfuric acid method is then used to measure the content of acquired polysaccharides [90,92,95,96,97,98]. After ethanol precipitation, products mainly contain polysaccharides, whereas the ones lyophilized without precipitation contain other constituents, such as polyphenols [4,24,47,87]. Therefore, ethanol has been most commonly used to assist in water extraction by precipitating polysaccharides from the supernatant [38,55].
Some extracting factors, such as the liquid–solid ratio, the extraction temperature, and the extraction time, can influence the yield of polysaccharides [22]. Yuan et al. stated that the yield of polysaccharides first rose along with the increase in liquid–solid ratio and extraction temperature, and then stopped after extracting for approximately two hours [22]. Researchers use statistical analyses to research on the optimal factor combination, such as designing Box–Behnken tests using the response surface methodology to conduct a cross-analysis between factors [22,39,99,100]. As a well-known for effectively presenting optimized conditions, the response surface methodology can also assist researchers in discovering the extent to which each factor affects the total yield, despite some differences among studies. The results of Yuan et al. suggested that the liquid–solid ratio demonstrated the most decisive influence on the total yield, followed by extraction temperature and extraction time; in contrast, Xu et al., still in consensus with the idea that the liquid–solid ratio has the greatest influence, argued that extraction time demonstrated a more substantial influence than the extraction temperature [22,99].
Assistant methods are used in some studies to make samples more susceptible to extraction or to increase yield. Among them, ultrasonic assistance is most commonly used to process sample powders; during this process, the ultrasound promotes the permeation of solvent molecules into the interstice of tissue cells to ensure that they make full contact with biomolecules [4,36,37,101]. Apart from applying ultrasound, some researchers have applied complex enzymes to assist extraction. Cheng et al. used a composite enzyme including cellulase, pectinase, and protease in a 2:1:1 ratio to extract polysaccharides after soaking samples in ethanol to eliminate triterpenoids [97]. Several other results confirmed that adding composite enzyme could increase polysaccharide yield [98,100]. Some might use microwave combined with ultrasound to assist polysaccharide extraction using hot water [36].

3.2. Ethanol Extraction

Ethanol is often utilized when extracting polyphenols, terpenoids, and flavonoids from the Sanghuang species, often assisted by ultrasound [2,13,14,21,32,37,38,39,40,41,42]. Polyphenols and their derivative compounds, such as hispidin, are often best extracted using ethanol as a solvent [3]. However, ethanol extracts also contain other components, such as sesquiterpenoids and carbonyl compounds’ derivatives, suggesting that compounds in the Sanghuang species might not be exclusive to a few categories [32,101]. Some changes of the solvent might occur in some research papers, such as replacing ethanol with 70% methanol or absolute alcohol [9,25,28].
Researchers also use Box–Behnken tests and multifactor level response surface analysis to acquire their optimal conditions in ethanol extractions [37,102]. Cai et al. used ethanol and ultrasound to extract triterpenoids from S. sanghuang and ranked factors based on their influence on yield: ethanol concentration, liquid–solid ratio, extraction time, and extraction temperature [37]. They added that increasing extraction time would initially increase the yield yet decrease it at some point, which might result from a prolonged ultrasound permeation that destroyed triterpenoid structures and enhanced the dissolution of other substances [37].

3.3. Deep Eutectic Solvent Extraction

Despite the current trend of using ethanol or hot water to extract, these solvents have several disadvantages, such as long extraction time, low purity or yield, and possible contamination to the environment [43]. As a result, a constant search has long been in motion for newer, safer, and more environmentally friendly methods that take less time to finish the extraction while yielding more [43]. DES is a newly emerged extracting method uncovered in 2021 and 2022, which replaces organic solvents like ethanol with a group of environmentally friendly ones acquired by blending two or more solvents or substances in a particular proportion [35,43]. The origin of such a mixed solvent came from the research of Abbott et al., in which they acquired a low-melting eutectic mixture named DES by blending amides and quaternary ammonium salts [43,103]. Compared to old solvents, DES is easy to synthesize, inexpensive, nontoxic, not easy to volatilize, nonflammable, and degradable [35,43].
Similar to the hot water or ethanol extraction methods mentioned above, researchers have utilized some methodologies to ensure the most appropriate conditions. Zheng et al. found an approximate range for each of their experimental factors to yield the best combination by designing single-factor tests [35]. Their highest yield of polyphenols was 1.5 times more than using 60% ethanol, probably due to better viscosity, polarity, and surface tension [35,43]. Zhang et al. added that products extracted with DES showed a better scavenging effect than their ethanol counterparts, pointing out DES’s superiority over ethanol [43]. Therefore, DES methods might draw more attention from researchers in the future.

3.4. Alkaline Extraction Methods

Apart from the more commonly used water extraction and ethanol precipitation methods, some researchers have used alkaline to extract polysaccharides from Sanghuang samples [10,45,104]. Evidence has shown that different solvents could influence the final yield, structure, or even bioactivities of polysaccharides [10]. The alkaline solution could destroy the hydrogen bond in cell walls to release the polysaccharides and improve the extraction rate [45]. Building on previous studies, Wang et al. used 1.25 mol/L NaOH/0.05% NaBH4 to extract polysaccharides from P. linteus mycelia at room temperature for 3 h, followed by neutralization and ethanol precipitation [45]. The polysaccharide content acquired from the phenol–sulfuric acid method was 84.92%. Pei et al. used the same alkaline solution to extract polysaccharides and acquired approximately 90% of polysaccharide content after purification [105]. The higher content compared to the results of Wang et al. might originate from different purification methods. Table 1 concludes the results mentioned above.
Table 1. Conclusion and comparison of the extraction methods.
Table 1. Conclusion and comparison of the extraction methods.
Extraction MethodMaterialProductPurification MethodOptimized ConditionFinal YieldReference
Extraction TimeLiquid–Solid RatioTemperature/Power
Ethanol precipitationLiquid culture broth4 ExopolysaccharidesDEAE-Sepharose Fast Flow column
Sephacryl S-100 HR gel column
---11.77%
55.36%
17.12%
15.75%
[48]
Boiling waterDried fruiting bodyPolysaccharide and Polyphenol----5.51%
23.00%
[4]
Boiling waterFruiting bodyPolysaccharide-2.26 h21.61:1 mL/g99.24 °C9.40%[22]
Boiling waterMyceliumPolysaccharide-2.5 h14:1 mL/g60 °C4.91%[99]
Boiling waterSanghuang powderPolysaccharide-8 h18:1 mL/g90 °C2.12%[96]
Boiling waterMyceliumPolysaccharide-3.5 h45:1 mL/g100 °C3.99%[100]
Boiling waterFruiting body powderPolysaccharideDEAE-Sepharose Fast Flow column
Sephacryl S-400 column
Sephacryl S-200 column
4.35 h26:1 mL/g100 °C-[106]
UltrasonicationMycelium powderPolysaccharide-30 min40:1 mL/g45 °C/120 W10.73%[39]
Triterpenoid-25 min50:1 mL/g45 °C/150 W1.51%
UltrasonicationMycelium powderPolysaccharide-260 s49:1 mL/g464 W13.19%[36]
UltrasonicationDried fruiting bodyPolysaccharide-32.7 min32.5:1 mL/g360 W3.46%[97]
Ethanol extractionMycelia and broth of S. sanghuang8 sesquiterpenoids and 6 polyphenolsSephadex 200 column----[101]
Ethanol extraction and ultrasonicationSanghuang powderFlavonoid and polyphenol-30 min--(10.18 ± 0.85)%
(13.58 ± 1.33)%
(14.62 ± 1.05)%
(15.38 ± 0.76)%
[21]
UltrasonicationS. baumii powderFlavonoid and polyphenolMacroporous membrane30 min---[28]
Ethanol extractionFermented broth of S. lonicericolaPolysaccharopeptideDEAE exchange column24 h--23.00%[9]
Ethanol extractionMycelia of S. sanghuangTriterpenoids-20 min-60 °C13.30%[37]
DES extractionDried fruiting bodies of S. baumiiPhenolics-42 min34:1 mL/mg58 °C12.58%[35]
DES extractionFruiting bodyPolyphenols-21 min260:1 mL/g80 °C(12.45 ± 1.88)%[43]

4. Purification Methods

4.1. Traditional Methods

Multiple studies have demonstrated that chromatography and Sevag reagents are appropriate methods for separating and purifying desired components [9,49,55,101,107]. Sevag reagents have been widely used in purifying fungus polysaccharides as an efficient method to deproteinize or decolorize samples [108,109,110]. For other impurities, such as salts or polysaccharides, researchers have resorted to specific columns based on anion exchange or size exclusion to elute the crude mixture with NaCl or distilled water, which has long been used to purify many compounds from medicinal fungi [5,9,55,90,92,101]. Most studies contained two chromatographic steps using two kinds of columns: diethylaminoethanol (DEAE) chromatography columns (often used first) and Sephacryl columns [5,9,47,48,49,55,92,101,111,112].

4.2. Macroporous Resin Methods

The Sanghuang extracts sometimes contain not only proteins or salts but also plant pigments like flavonoids or polyphenols that could not be easily removed with Sevag reagents or columns, thus calling for a more specific and more thorough purification method [86,113,114,115]. With properties like polarity and specific areas, macroporous resin has been developed for the decoloration and purification of crude polysaccharides with a higher efficiency than traditional methods, suggesting the potential for large-scale purification [113,115,116,117]. Evidence has supported the isolation of flavonoids from plant leaves using resins with moderated specific areas [113,115,118,119]. Some studies even pointed out the optimized condition for the resin-purification process. By studying six kinds of macroporous resin on their abilities to absorb Phelligridin LA (PLA) from I. baumii, Wang et al. provided optimized conditions: an extraction pH of 5, a proportion of fermentation broth to resin of 4:1, an extraction temperature of 20 °C, an adsorption time of 150 min, and a concentration of PLA of 6.78 mg/mL [118]. Apart from isolating PLA, the microporous resins they used could also assist in quantifying PLA in different samples based on light absorption at 365 nm [118]. The specific ability of microporous resin in polysaccharide decoloration and purification could be significant when applying them to large-scale use and accelerating the pace of Sanghuang commercialization.

4.3. Molecular Imprinting Technology

Molecular imprinting technology uses synthesized molecular imprinting polymers (MIPs) to bind to a specific template molecule with high affinity by forming specifically three-dimensional binding sites and certain chemical bonds [64,120,121,122]. Magnetic molecularly imprinted polymers (MMIPs) are a newly emerged version of this technology [64,123,124]. Zhang et al. utilized and improved this method after realizing that traditional isolation methods for morin might be poorly specific and inefficient [64]. In their study, they compared multiple ways to isolate morin, among which the synthetic morin magnetic molecularly imprinted polymers (Morin-MMIPs) showed the highest yield. Apart from a high selectivity, their Morin-MMIPs also showed a short purification time, a good absorption ability, and a high absorption efficiency that remained stable even after six cycles [64]. These polymers could purify relatively rare components more efficiently, thereby facilitating more research on fungus component isolation.

5. Conformational Properties of Sanghuang Polysaccharides

Polysaccharides are common medicinal components in fungi whose composition has led to considerable interest from researchers [55,93,125,126]. The more complex the polysaccharide structure is, the more biological activity is associated with its backbone structure [57,58]. Evidence has already shown that more galactose in polysaccharides leads to more potent anti-inflammatory abilities and that a high concentration of (1,3)-β-glucan in polysaccharides could effectively inhibit intestinal inflammation [87,127]. Other studies have shown that interactions between polysaccharides and proteins might be the critical mechanism of antitumor effects and that polysaccharides might prohibit inflammation progress by inhibiting cytokines [127,128]. The enrichment of branched chains is also reported to reduce inflammatory responses [58]. These promising effects make it paramount to detect and elucidate the structure of Sanghuang polysaccharides, since the Sanghuang species could be considered a functional food. However, no consensus has been established on how Sanghuang polysaccharides contribute to each biological effect. Nor have there been any comprehensive studies on structure–effect relationships. To date, the research related to such relationships has often focused on their case. Therefore, the following section lists several new studies and methods elucidating Sanghuang polysaccharide composition, some digging deeper into structure–effect relationships. Figure 2 includes the main processes and methods required for structure identification.
From 2020 to 2022, Cheng et al. studied the conformational properties and effects of cultured mycelia, fruiting bodies, and liquid culture of S. sanghuang [47,48,112]. Building on previous research, they elaborated on the conformational properties of an intracellular polysaccharide named SSIPS1 from the cultured mycelia of S. sanghuang and revealed its potential in blood glucose control [47,89]. After monosaccharide composition detection, methylation analysis, and 1D/2D NMR analysis, they inferred that SSIPS1 had a backbone including structures of 1,4-linked α-D-Glcp and two branches separated from the backbone at the O-6 positions that comprised 1,4-linked α-D-Glcp terminated with α-D-Glcp, 1,4-linked α-D-Glcp, and 1,4-linked β-Galp terminated by α-D-Glcp [47]. They inferred that these structures of SSIPS1 might be related to its inhibitory activities against α-amylase and α-glucosidase, and to the ability to increase glucose metabolism in the HepG2 cell model with insulin resistance [47]. Chain conformation analysis and atomic force microscopy (AFM) revealed that SSIPS1 had a flexible chain and a worm-like structure in an aqueous environment; this was also the case for a novel exopolysaccharide named mannan (SSEPS2) from the liquid broth of S. sanghuang, mentioned in their studies on its antitumor and cell-proliferation-inhibiting effects [47,48]. They discovered, with the methods listed above, that SSEPS2 had a backbone including 1,3-, 1,2,6-, and 1,2-linked α-D-Manp residues and branches composed of α-D-1,6-linked Manp residues and a terminal α-D-Manp residue [48]. However, this study did not elucidate the relationship between structure and antitumor activities. Then, in 2022, they further extended their research to the fruiting bodies of S. sanghuang and acquired a novel mannogalactan (SSPS1) composed of several monosaccharides, such as D-galactose, D-mannose, L-fucose, 3-O-methyl galactose, and D-glucose [112]. Like the two aforementioned polysaccharides, they also discovered the presence of flexible chains in SSPS1 [47,48,112]. But SSPS1 backbone consisted of α (1,6-linked) galactopyronosyl, related to inhibiting tumor cell proliferation [112]. Their study contributed to elaborating on structure–effect relationships and accelerating the process of Sanghuang becoming a functional food. Despite the relatively clear demonstration of these three polysaccharides’ hypoglycemic and antitumor effects, more complex and concrete proof of structure–effect relationships requires future research.
In 2021, Sun et al. isolated a polysaccharide from P. baumii fruiting bodies to study relationships between structures and anti-inflammatory effects [87]. Methylation analysis and one-dimensional NMR analysis revealed that the SHPS-1 polysaccharide consisted of 2.2% arabinose, 15.7% mannose, 49.3% glucose, and 32.8% galactose. SHPS-1 also included a backbone containing 1,3-linked β-D-Glcp and 1,6-linked α-D-Galp residues and branched chains separated at the O-6 of the β-D-Glcp residue. They inferred that the backbone structure and residues might contribute to its promising anti-inflammatory effects [87,128]. SHPS-1 also contained more galactose than other polysaccharides isolated from the Sanghuang species, which is possibly related to better anti-inflammatory activities [87,127]. They surmised that the enrichment of mannose on side chains on SHPS-1 might allow it to stimulate macrophage mannose receptors to inhibit inflammatory bowel disease (IBD) [87]. After comparing their results with previous studies, Sun et al. also pointed out that different culturing substrates of P. baumii might alter the composition of polysaccharides. Their novel study illustrated the structure of a polysaccharide and gave an assumption of the connection between the anti-inflammatory effects of polysaccharides and their structures, which would enlighten future research.
In 2022, Wan et al. successfully extracted four heterogeneous polysaccharides from S. vaninii fruiting bodies, all consisting of 11 different monosaccharides, with glucose and galactose being the main components, comprising 70% of the total [55]. They then focused on one of the polysaccharides, named SVP-1, which consisted of 60.26% glucose, 13.84% mannose, and 11.29% galactose. Methylation revealed the presence of 1,4 and 1,6 connections between SVP-1 glucose groups and between glucose and galactose ones. Fourier transform infrared spectrometry (FTRI) analysis revealed that different eluents might have caused discrepancies between polysaccharides on their monosaccharide composition, molecular weight distribution, functional groups, and uronic acid content among the three acidic polysaccharides, leading to different bioactivities [55]. Wan et al. also confirmed that acidic polysaccharides had promising antitumor effects based on their observation that SVP-1 could inhibit NCI-H460 cell proliferation, corresponding with previous results on other acidic polysaccharides [55,94,129]. However, this result did not show the detailed structure–effect relationship, thus future research is still needed.
Figure 2. The general methods of identifying the structure of bioactive components from Sanghuang [5,48,49,87,130].
Figure 2. The general methods of identifying the structure of bioactive components from Sanghuang [5,48,49,87,130].
Molecules 29 01195 g002
Other evidence on Sanghuang polysaccharides has also shown some similar characteristics. Early in 2019, Ma et al. found a polysaccharide free of uronic acid but abundant in D-mannopyranose and glucose [5]. In the same year, Ran et al. isolated an S. sanghuang polysaccharide with pyran ring structures, which showed promising antiviral effects and healing-promoting effects after being used in silver nanoparticle synthesis [111]. Instead of forming flexible chains in aqueous environments, a polysaccharide from S. vaninii, as He et al. argued, might form a tight spherical conformation with branches in those environments [111,130].
Some studies have already revealed numerous polysaccharide structures and conformation properties from fungi or plants [93,125,126,128]. There is a standard agreement that functional groups or branches are crucial to their medicinal effects, but more detailed relationships between conformational properties and effects still need to be clarified [55]. Researchers have been struggling to identify the core functional groups contributing to the effects from the tremendously complex net of interaction between biomolecules [55,58,131]. Therefore, more evidence could be added to strengthen the ground of the Sanghuang species as functional foods.

6. Medical Properties of the Sanghuang Species

6.1. Anti-Inflammatory Effects

Polysaccharides, flavonoids, polyphenols, and triterpenoids from Sanghuang can affect the inflammation process through several mechanisms [12,18,34,42,87,132,133]. Figure 3 encapsulates several pathways involved in inflammation and where the Sanghuang components might intervene. Many studies have reported an effect on the Phosphatidylinositol-3 kinase/protein kinase B/mammalian target of rapamycin (PI3K/Akt/mTOR) signaling pathway, relating to cell survival [12,13,133,134,135,136,137,138,139]. Akt participates in the signal transduction in lipopolysaccharide (LPS)-induced inflammation by promoting NF-κB activation through IκB kinase (IKK) phosphorylation and regulating downstream activities of mTOR, suggesting that the PI3K/Akt/mTOR pathway could be targeted to reduce proinflammatory cytokine synthesis or NF-κB activation [13,133,135,136,137,140,141,142,143].
Using S. sanghuang mycelia ethanol extracts in an ICR mouse inflammation model induced by LPS; both Lin et al. and Jiang et al. mentioned that 500 mg/kg extracts, injected and orally taken, respectively, showed the most robust effects among all doses on reducing the inflammatory status in the lung and liver by activating the PI3K/Akt/mTOR pathway, reducing the permeation of neutrophils, and decreasing the release of proinflammatory cytokines [13,40,132]. Jiang et al. further stated that orally taking 500 mg/kg of the extract for six consecutive days could prevent paracetamol-induced liver damage compared to their positive control group using N-acetylcysteine that could only exert its greatest effect after the damage has been caused, pointing out the clinical potential of Sanghuang [132]. Instead of using extracts, Huang et al. found that 10 mg/kg of purified hispolon, a polyphenol compound from S. sanghuang, injected intraperitoneally into the ICR mice one hour before LPS induction, showed the strongest protective effect compared to other doses by decreasing the proinflammatory cytokines, decreasing weight loss, and activating the PI3K/Akt/mTOR pathway [133,144,145,146,147]. The purity of their extracts might explain the dosage discrepancies among these three studies.
Besides corroborating the effects on the TLR4/PI3K/Akt/mTOR pathway, some results have mentioned the activation of other pathways. In 2015, Miao et al. reported that 50 mg/kg Phellinus polysaccharides taken orally every day by SD mice with rheumatoid arthritis (RA) might improve the condition of RA by inhibiting the Wnt signaling pathway, based on an observation of a decrease in the expression of β-catenin, C-myc, and ccnd1 [148]. Not many studies have focused on the effects of mushroom components on RA in recent years; therefore, it could be worth researching. In 2017, Lin et al., for the first time, reported that 500 mg/kg of S. sanghuang mycelium extract injected intraperitoneally prior to LPS induction could influence the expression levels of Kruppel-associated box (KRAB)-associated protein 1 (KAP1), thereby regulating NF-κB and Nrf2 pathways and further downregulating the syntheses of proinflammatory cytokines [12]. Combined with a previous study in 2008 on the relationship between KAP1- and STAT3-involved pathways, this result introduced the participation of the STAT pathway and the NF-κB and Nrf2 pathways against the inflammation process [12,18,139,149,150,151,152]. Similarly, in 2021, Sun et al. pointed out that treating RAW 264.7 cells with 250 μg/mL of polysaccharide from P. baumii for 24 h demonstrated anti-inflammatory effects through the STAT-1 pathway instead of the NF-κB pathway because they failed to observe the expression or translocation of factors in the NF-κB pathway, such as IκBα and p65 [87]. In 2021, Li et al. added that 100 μg/mL hispidin 1 h prior to LPS induction showed the strongest anti-inflammatory effects among all the doses by inhibiting the increase in IL-6, NO production, and TNF-α in cells through nuclear factor erythroid 2-related factor 2 (Nrf2) activation that inhibited inflammation-related gene expression [3].
Despite these results, most studies have observed and reported regulation effects on the TLR4/PI3K/AKT/mTOR signaling pathway and the NF-κB pathway activation, strengthening their crucial role in preventing or inhibiting the inflammation process [12,13,18,132,133,139,153]. NF-κB activation would reduce MAPK-related protein phosphorylation, such as ERK and JNK, and influence the activation of the MAPK to regulate cell migration or proliferation, thereby affecting the levels of downstream proinflammatory cytokines and inflammation-related substances [9,12,13,18,41,86,132,139,154,155,156]. The reduced release of cytokines could prevent or stop causing tissue damage or organ failure, such as edema [13,132,133,157,158,159]. As mostly mentioned by multiple studies, Sanghuang products could effectively reduce IL-6, IL-1β, and TNF-α by decreasing their mRNA levels [3,9,12,13,31,34,41,86,160]. But several studies have also introduced other proinflammatory mediator changes, such as NO decrease and IL-10 mRNA level increase [9,18,95,161]. Taken altogether, the anti-inflammatory abilities and protective effects of an overdose of Sanghuang proven by studies in the past ten years provide strong support for their potential use as a functional food.

6.2. Antioxidant Effects

Free radicals and reactive oxygen species (ROS), unstable and detrimental chemicals derived from metabolic pathways in cells with or without specific enzymes, are normally generated through electron transport during respiration or through exposure to exogenous substances, such as tobacco or heavy metals [25,86,162,163,164,165]. Many diseases are related to free radicals and ROS, such as aging, heart diseases, or diabetes mellitus [162,164,165]. The abilities to scavenge free radicals are typically observed and analyzed in vitro by using 2,2-diphenyl-1-picrylhydrazyl (DPPH) and ABTS, creating fade at 517 nm and 734 nm, respectively, with vitamin C (Vc) or butylated hydroxytoluene (BHT) as the positive controls [5,9,25,37,46,96,166]. The color fade can then be observed through spectrophotometry to analyze the radical scavenging abilities of a bioactive component [25,37,46]. Some articles might also add results of scavenging superoxide anions, hydroxyl radicals, or Fe ions, to be more comprehensive [21,37].
Evidence has shown that polysaccharides, polyphenols, flavonoids, and triterpenoids all can scavenge free radicals [21,25,32,35,37,162,167,168,169]. Polyphenols and triterpenoids from the Sanghuang species have long been believed to be major free-radical scavenging components in studies within the 5 years [25,32,35,37,66,162,170,171,172,173]. Cai et al. presented results showing that 18.75 to 350 µg/mL triterpenoid from S. sanghuang could increase free-radical clearance from 10% to 90% [37]. Several other studies confirmed that the antioxidant abilities are dose-dependent and better than those of their positive control group [35,162]. However, it remains controversial how strong Sanghuang polysaccharides’ effect is in scavenge free radicals [35,37,73]. In 2019, Li et al. used 4.0 mg/mL P. igniarius polysaccharides to scavenge ABTS and to demonstrate the superior antioxidant effects of polysaccharides, whereas others argued that they only played minor roles in antioxidation compared to flavonoids [9,46,73,96,106,174]. A new result in 2022, after comparing 15 strains on their performance of scavenging multiple radicals and reducing power, presented results suggesting that flavonoids and ascorbic acids contribute the most to the antioxidative effect; these are followed by polyphenols and triterpenoids, with polysaccharides being the least-effective antioxidant [173]. However, an article might bring an end to this debate by suggesting a combined antioxidative effect contributed by multiple components instead of a single material [175].
Other components can also demonstrate free-radical scavenging effects. Built on previous research on plant and fungal pigments, in 2018, Heo et al. presented results suggesting that 10 mg/mL extracellular pigments from S. baumii possessed higher DPPH and ABTS scavenging abilities compared to other fungal extracts [66,176,177,178,179]. They also gave an original point that some water-soluble polysaccharides from the fruiting bodies or mycelia reported before might be pigments since they had been preprocessed with water several times, so the water-soluble polysaccharides might already be washed away [66,180]. Besides probing into the medical application of extracts, researchers have also been paying attention to fermentation broth or decoction [166,170].
Moreover, some researchers presented results of Sanghuang products affecting antioxidant enzyme activities [5,133,180]. Ma et al. intraperitoneally injected mice with 50, 110, or 170 mg/kg of the exopolysaccharide from S. sanghuang broth and 50 mg/kg of Vitamin E for 35 days and argued that the 170 mg/kg dose group demonstrated comparable antioxidative effects to Vitamin E by showing an observable increase in catalase (CAT), superoxide dismutase (SOD), and Trolox equivalent antioxidant (TEAC) content in mouse serum [5]. Instead of using pure extract, Gu et al. discovered that 12 g/kg of I. sanghuang decoction taken orally for 15 consecutive days showed an increase in the activities of total antioxidant capacity (T-AOC), SOD, and peroxidase (POD), which was nearly comparable to the Vitamin E control group [181]. They further argued that such antioxidant effects might be related to the Nrf2/HO-1 pathways; this has also been mentioned in other studies before and after them [133,156,181,182]. Built on previous studies, Huang et al. argued that hispolon could increase levels of antioxidant proteins, such as SOD, HO-1, and Nrf-2, and that it elevated the activity of the Keap1/Nrf2/HO-1 and LKB1/CaMKK-AMPK axis to prohibit ROS production induced by LPS [133,137,147].
Researchers have recently agreed that different Sanghuang plants exhibit different antioxidant abilities [22,64,73,86,171,183,184,185,186]. Firstly, evidence suggested that, under the same conditions of being wild or cultivated, S. vaninii demonstrated better antioxidant effects and higher total polyphenol and flavonoid content than S. sanghuang, despite others arguing to the contrary [21,184,185,187]. Secondly, different cultivation substrates, methods, and years affect antioxidant abilities [23,183,185,187]. A piece of evidence from 2020 showed that adding substances like salicylic acid could increase the content of flavonoids, thereby increasing antioxidant abilities [185]. Thirdly, the antioxidant abilities within the same species could also differ. In 2020, Wang et al. studied several S. sanghuang samples and reported that the one tagged SH06 exhibited the most robust DPPH scavenging abilities and reducing power compared to others [185]. Guo et al. added that, even in the same plant, intracellular products showed stronger hydroxyl and DPPH scavenging effects than extracellular counterparts [30]. Finally, different solvents might also affect antioxidant abilities [60,73,86,186]. Despite others arguing to the contrary, in 2016, Yang et al. used P. igniarius plants to prove that the water-soluble fractions demonstrated the most potent total antioxidant abilities, with ethanol fractions exhibiting the least potent total antioxidant abilities [188,189].

6.3. Antitumor Effects

Evidence has proved that polyphenols and polysaccharides possess cell toxicity effects, antiproliferation effects, and anti-angiogenesis effects on multiple cancer cells in vitro [2,4,15,38,42,72,73,131,190,191,192]. Other components, such as a stable lectin SHL24, could also demonstrate an antitumor effect and remain active when facing many sugars and cations, suggesting a potential for pharmacological applications [193].
There has been a debate on the effectiveness of these components’ antitumor activities, among which flavonoids are found to be more effective than polysaccharides and polyphenols [171,172,185]. Some studies have confirmed that polysaccharides participate in antitumor activities indirectly by stimulating an immune response or directly impeding cell proliferation [4,55,129,194,195,196]. Notably, many studies have provided evidence that Sanghuang polysaccharides could deliver antitumor effects with fewer side effects than traditional medicine, such as cyclophosphamide (CTX), allowing them to be taken as a medicinal foods [14,197,198,199]. CTX worked better in inhibiting tumor growth than Sanghuang polysaccharides while being more toxic, causing immunosuppression [36,200]. Surprisingly, evidence from 2015–2021 has shown that Sanghuang polysaccharides could intensify the effects of CTX and simultaneously mitigate its immunosuppression effects, pointing out not only their superiority as functional food capable of impeding tumor progression but also the potential application of being used with chemical medicine [36,199,200].
Mechanisms for stopping tumor progression have been elaborated further by studies in the last ten years. Impeding or preventing cell proliferation is commonly mentioned in research results, and could be achieved through cell cycle arrest and apoptosis regulation [2,4,20,49,190,197,201,202]. In 2022, Guo et al. stated that 300 ug/mL of their 60% ethanol extracts from S. vaninii, basically consisting of polyphenols and flavonoids, demonstrated the most potent anti-proliferation effect by decreasing the cell number in G0/G1 phase by 9.53% and increasing that of G2/M by 9.36% compared to their control group; this effect could be related to inhibiting critical genes in the mTOR signal pathway and relevant protein synthesis [2,72]. Other studies within the 5 years also reported similar changes [20,49,72,86,203,204]. Mechanisms for such arrest might involve regulating expressions of crucial proteins, such as p21, thereby affecting the cyclin-dependent kinase (CDK) complex formation, including the CyclinD-CDK4/6, CyclinE-CDK2, and CyclinA-CDK2 complexes [2,4,72,201,205,206]. On the other hand, studies within the past ten years have proven that Sanghuang components, mostly polysaccharides, could impede cell proliferation by apoptosis, usually involving the Bcl-2 family and the caspase family [4,170,190,201,207,208]. The apoptotic rate of MCF-7 cells after being treated with 0–800 μg/mL S. vaninii polysaccharides rose from 6.1% to 51.8%, in contrast with the rate of living cells dropping from 92.7% to 47.5% compared to the control group; this might be caused by an increased expression of relevant proteins in the caspase family and the mRNA of BAX and p53 [49]. Similar results of downregulated Bcl-2 expression also appeared in several studies in 2021 [4,72,189]. In 2022, He et al. used a dose course of petroleum ether extract of P. vaninii containing flavonoids and discovered their 200 mg/kg group had a similar apoptotic effect compared to that of the CTX positive control, which involved forming an equilibrium between BAX and Bcl-2 and the participation of the death receptor pathway [38]. Besides the mechanisms above, Ca2+ overload could also cause cells to enter the apoptosis phase by facilitating ER stress, contributing to apoptosis and antitumor effects [20].
Suppressing cell migration could also impede tumor progression and growth [49,107]. Despite not being as good as the positive control Doxorubicin hydrochloride, Wan et al. stated that 800 μg/mL S. vaninii polysaccharide could suppress cell migration and leave the scratch area unhealed [49]. Besides polysaccharides, Qiu et al. discovered an antitumor effect of 100 mg/kg S. vaninii Inoscavin A after injecting it intraperitoneally into BALB/c nude mice with HT-29 colon cancer cells, which involved the participation of transmembrane protein smoothened (Smo) receptors and the inhibition of the hedgehog signal pathway [107]. Based on their tumor tissue weight results, they also argue that the antitumor effect of the 100 mg/kg group was weaker than 2 mg/kg of Taxol twice a week [107]. Some studies in the past five years have mentioned that the antimigration effect of Sanghuang is related to the epithelial–mesenchymal transition (EMT) process, causing cells to be less adhesive and more mobile to metastasize to distant regions by inhibiting the matrix metalloproteinase (MMP) from interacting with the extracellular matrix [20,49,72,208,209]. These novel discoveries have shown the possible target of adhesion molecules in cancer treatment.
Other mechanisms also contribute to tumor suppression [49,55]. In 2022, Wan et al. reported that 1000 μg/mL S. vaninii polysaccharides inhibited colony formation of non-small-cell lung cancer cells and rendered apparent cell shrinkage and occasional vacuolar cytoplasm compared to other dose groups [55]. Some studies mentioned increasing the intensity of immune responses as a mechanism for inhibiting tumor progression [36,72,198,203,210,211]. Additionally, anti-angiogenesis is related to tumor suppression, which might result from promoting M2 macrophages to polarize to M1 to inhibit vascular endothelial growth factor (VEGF) formation [4,190,191,197,199]. Based on previous results that VEGF binding to VEGF receptors could activate P13K/AKT/mTOR pathways, Zhao et al. inferred that inhibiting VEGF and suppressing essential protein phosphorylation in this pathway might account for the antitumor effects [197].
Disparities exist in antitumor effects among different extracts. Different polysaccharides show different levels of inhibitory effects on tumors [55,212]. In 2017, Ying et al. first reported that mycelium and fruiting body polysaccharides from P. igniarius, both 1.5 mg/mL, showed promising proliferation inhibiting effects compared to paclitaxel (10 μg/mL, inhibition rate 72.31%) on HepG2 cells (inhibition rates 66.01% and 70.37%, respectively), suggesting that the fruiting body ones were more effective [212]. Different cultivation conditions also affect the antitumor activities of a given plant. Among the oak segment, sawdust of mulberry branch (BS), and artificial media, the Sanghuang plant grown on BS showed more substantial inhibitory effects on cell reproduction than those grown on others, probably due to a boosting effect of the mulberry branch on S. vaninii growth and the accumulation of mycelium components [15,194].

6.4. Immunoregulation Effects

Polysaccharides from the Sanghuang species could exert immunoregulation effects by inhibiting the expression of several factors, such as myeloid differentiation factor 88, tumor necrosis factor receptor-associated factor-6, NF-κB, and P38, and by suppressing the production of oxidative radicals and the secretion of cytokines, such as TNF-α and IL-1 [31,56,191,213]. Sanghuang polysaccharides could also increase the level of IL-2 and reduce epidermal growth factor (EGF) expression [170,197,199,211,214].
Apart from affecting factors or cytokines, evidence has also indicated that Sanghuang products could stimulate the proliferation of T and B cells [14,198,199,213,215]. Early in 2003, Kim et al. showed that a 48 h treatment of 200 μg/mL and 500 μg/mL P. linteus proteoglycan targeted B cells and demonstrated a promising proliferation effect, though slightly weaker than their positive control of 5 μg/mL LPS [56]. In the next year, Lim et al. discovered that taking 2 mg/day of a P. linteus fruiting body extract by oral gavage could regulate T cells by increasing the proportion of splenocytes’ CD4+ and CD8+ T cells [216]. Instead of using a low concentration, in 2019, Wen et al. used 16 mg/kg of both cultured and wild type I. sanghuang chloroform extracts, consisting of polysaccharides and flavones, and orally piped them into mouse stomach for 12 consecutive days and discovered a significantly stronger immuno-regenerative effect than the 4 mg/kg and 8 mg/kg groups by showing more distinctive increase in white blood cells and spleen recovery index and by acting as polyclonal activators of B cells [14]. The higher dosage compared to other studies could be explained by using compound extract or chloroform to extract desired components that could only contain a small amount of polysaccharides, which should be the main component that grants such an immune-regulating effect in Sanghuang. Other evidence showed that the polysaccharides could induce B cells to generate co-stimulatory factors, such as CD80 and CD86 [199,211]. Additionally, polysaccharides could also affect other immune cells, such as peritoneal natural killer (NK) cells, bone marrow cells, and macrophages, thereby increasing the strength of the immune response [86,203]. Taken together, the immunoregulation effect of the Sanghuang species was not limited to the activation of T cells and B cells, which broadened the scope of future research and expanded the possible targets in the commercial application of Sanghuang components.

6.5. Antimicrobial Effects

Antimicrobial effects are only sometimes mentioned in most studies. This review concludes a few studies that have mentioned the microbe-suppression effects of polysaccharides, polyphenols, and flavonoids [78,111]. Cheng et al. in 2019 isolated more than a dozen secondary metabolites, mostly sesquiterpenes, from S. microcystideus and discovered that 50 μg/mL–100 μg/mL was the minimum inhibitory concentration for most of them, manifold lower compared to each positive control group, suggesting a weak antimicrobial effect against an assay of bacteria, such as Staphylococcus aureus and Escherichia coli [78]. In the same year, Zhang et al. treated mice infected with Schistosoma japonicum by consecutively giving 400 mg/kg ethanol-extracted polysaccharides from P. igniarius through oral gavage for 30 days and argued that the fibrosis state and over-expressed TGF-β were ameliorated by both polysaccharide and praziquantel groups through upregulating mRNA expression of Nrf2 and Gsta4 gene, suggesting a potential for a long-term functional food source [217]. The polysaccharide from Zhang’s result could work in vivo and bring multiple systems to work, such as triggering the immune system, and cause a combined effect, whereas the extracts in Cheng’s work could only work alone in vitro, which might explain the discrepancy of antimicrobial effects between these two results.

7. Health-Improving Effects

Over the years, researchers have focused on utilizing medicines with purified materials or artificial compounds to mitigate or reverse some disease symptoms. However, these medicines might bring some side effects [36,200,218]. Recent studies have shed light on the protective or relieving effects of Sanghuang in several diseases and forms of organ dysfunction.

7.1. Pulmonary Protection Effects

Chien et al. has provided up-to-date results in this area [218]. Using a bleomycin-induced mouse model for idiopathic pulmonary fibrosis, they discovered that taking S. Sanghuang orally could significantly decrease fibronectin and collagen expression [218]. Apart from inhibiting fibrosis-related proteins, their results also showed an obvious suppressing effect on inflammation progression, such as inhibiting TNF-α, IL-6, and IL-1β, through regulating key molecules including iNOS, NF-κB, and COX-2 [218]. They further proved the participation of the MAPK pathway by showing a lowering effect on p-ERK, p-JNK, and p-p38 [218]. Moreover, they novelly presented results suggesting that S. sanghuang mitigated autophagy and thereby mitigated the progression of idiopathic pulmonary fibrosis [218].

7.2. Hypoglycemia- and Diabetes-Mitigating Effects

Diabetes mellitus type 2 (T2DM) is a commonly detected disease connected to organ dysfunctions, insulin resistance, and metabolism disorder, posing a significant threat to public health [34,101,215,219,220,221,222]. During the past few decades, using medicinal foods, such as edible fungi and their bioactive components, to replace traditional medicine in tackling chronic diseases like T2DM has been proven beneficial and has, therefore, received significant attention from researchers [222,223]. Recent discoveries have shown that polysaccharides, triterpenoids, and polyphenols from edible fungi possess antidiabetic and hypoglycemic effects by improving insulin resistance [34,101,224].
Mechanisms of improving insulin resistance have been presented in several studies, including the regulating activities of crucial enzymes, scavenging radicals, to increase abilities to resist antioxidative stress, increasing insulin sensitivities by upregulating the expression of peroxisome proliferators-activated receptor-γ (PPAR-γ), and stimulating B cells to secrete insulin [191,215,225]. Cheng et al. discovered that intracellular polysaccharides from S. sanghuang mycelia improved insulin resistance by inhibiting α-glucosidase and α-amylase in vitro and enhancing glucose metabolism in Hep G2 cells [47]. Corroborating the inhibitory effects on these two enzymes, Li et al. discovered that the IC50 for ethanol-extracted polysaccharides of α-glucosidase was lower than the positive control, acarbose (0.43 mg/mL and 0.87 mg/mL, respectively), suggesting a better hypoglycemic effect [186]. Yang et al. in 2021 explained the mechanism whereby the blood-sugar-lowering effect might correlate with the IRS1/PI3K/AKT pathway [34].
Apart from improving insulin resistance, products from the Sanghuang species can also affect the concentration of blood lipids, such as triglycerides and total cholesterol steroid [34,60]. Yang et al. stated that 150 mg/kg polyphenols from P. baumii could lower HDL-C and LDL-C levels comparably to the 100 mg/kg metformin group [34]. Moreover, in 2019, Huang et al. discovered that consecutively treating mice with streptozotocin (STZ)-induced diabetes with 80 mg/kg polysaccharides from P. igniarius for 12 days led to an obvious decrease, similar to their positive control of 4 mg/kg Rosiglitazone, in the fibrosis in kidneys caused by diabetic nephropathy; this could be explained by an upregulated E-cadherin expression, an inhibited α-SMA expression, and a balance between different matrix metalloproteinases [226]. Additionally, a study in 2021 proved that Sanghuang polysaccharides could prevent immune cells from infiltrating into islets, thereby preventing autoimmune diabetes [215].

7.3. Sleep-Improving Effects

Traditional ways to improve sleep with medicine have some side effects that might undermine their benefits, such as increasing susceptibility to suicidal ideation [227]. In the year 2021, this situation has changed, since Li et al. brought a new direction for improving sleep [3]. Li et al. used Sprague Dawley rats as an animal model to point out the value of ethanol extracts of S. sanghuang mycelia (GKSS) enriched with hispidin as a novel food material for sleep improvement [3]. They gave the rats a regular intake of 70 and 150 mg/kg ethanol extracts of GKSS and acquired their data from electroencephalography (EEG) and electromyography (EMG) electrodes implanted in rats’ brains. The time spent in rapid eye movement (REM) sleep and non-REM (NREM) sleep increased, while the waking time significantly decreased. Based on previous research, they focused on inflammatory cytokines and the Nrf2/ARE signal pathway (Figure 3) [3,153,228]. Besides postponing REM sleep by decreasing IL-6 production, they also observed a dose-dependent activation effect of GKSS and hispidin on ARE-luciferase activities, indicating the activation of the Nrf2/ARE signal pathway and proving the sleep-modulating effects of this pathway [3]. The Nrf2 protein is the connection between the molecular clock and metabolism, thereby influencing circadian gene expression or sleep patterns [3,161,229,230]. These results provide support for commercializing the Sanghuang species as functional food for sleep improvement.

7.4. Gout-Mitigating Effects

Gout, a common disease caused by the monosodium urate accumulating in the wrong spots, is usually characterized by high uric acid in the body and related to hyperuricemia [231]. Traditional treatment methods for gout include the used of xanthine oxidase (XOD) inhibitors, such as allopurinol and febuxostat, but they can lead to side effects that could also make patients’ lives difficult, such as kidney failure [231,232]. Therefore, researchers have been trying to find another way to treat gout and have discovered potential in the natural components of plants [232]. In 2021, Guo et al. used 50 mg/kg water extracts of S. vaninii to treat mice for nine days and introduced hyperuricemia on day 6 [30]. The aqueous extract demonstrated better effects compared to colchicine in decreasing uric acid levels by downregulating the expression of the catalyst, xanthine oxidoreductase (XOR); this catalyzed the transition of hypoxanthine into xanthine and led to the production of uric acid [30]. They further pointed out that the water extracts might improve the condition of rats with gout arthritis through demonstrating their anti-inflammatory effects, whereby swelling was ameliorated and the level of uric acid in serum was lowered [30].

7.5. Antiaging Effects

Aging is related to an imbalance caused by the accumulation of oxidative substances and the degeneration of antioxidative defenses [233]. Metformin is a popular antiaging medicine, but it has side effects [234]. Thus, dietary intervention, such as functional food, is a safe and cost-effective method for people to slow their aging [67,235]. Early in 2019, Wang et al. proved that 200 μg/mL ethanol extracts of I. obliquus showed the best antiaging effect in PC12 cell models among other test samples [236]. After using S. vaninii methanol and hot water extracts in vitro that mainly contain phenolic and flavonoid components, Li et al. reported an antiwrinkle ability based on the inhibiting effect on tyrosinase, the self-oxidation of L-3,4-dihydroxyphenylalanine (L-DOPA), elastase, and collagenase; these are critical molecules in the body and arterial aging [162,237,238]. Also, in 2019, Gu et al. stated that a Sanghuang decoction consisting of 2 g/mL crude medicine, mostly polysaccharides, triterpenoids, and flavonoids, could initiate the transcription of gene HO-1, thereby inhibiting oxidative stress and delaying the progression of aging [181].
By treating the Caenorhabditis elegans nematodes with 10, 25, and 50 mg/mL of S. sanghuang extracts to study their effects on lifespan in vivo under the influence of 50 μM 5-fluoro-20-deoxyuridine (FUDR), Dong et al. discovered that, after 14 days, the 25 mg/mL group showed a 26.41% increase in the mean lifespan compared to the control group (20.52 ± 0.34 days and 16.23 ± 0.36 days, respectively) [239]. They also inferred that the antiaging effect was achieved at the mRNA level, such as increased transcriptional regulators of daf-16 and sir-2.1, and that the lifespan was prolonged through the insulin/IGF-1 signaling pathway [239]. Their results, for the first time, revealed the related mechanisms in the antiaging effect of Sanghuang components. This will facilitate the prevalence of the Sanghuang species as a functional food.

7.6. Neuroprotective Effects

Neurodegenerative diseases, such as Parkinson’s disease (PD) and Alzheimer’s disease, have become some of the most challenging diseases and have claimed countless lives over the years [68,233,240]. The past decade saw the potential of the Sanghuang species against neurodegenerative diseases, as they could mitigate the damage to nerve cells caused by free radicals [233,240,241,242]. For instance, an ethyl acetate extract of P. linteus has been proven to be capable of protecting against neuronal cell death, similar to another result showing polysaccharides from I. obliquus protecting HT22 cells in mice with Alzheimer’s disease [243,244]. The Sanghuang species, as a potential functional food, could be of significant assistance in mitigating the development of neurodegenerative diseases.
Parkinson’s disease is marked by a gradual loss of dopaminergic (DA) neurons, causing dysfunction in the neural system and motion abnormalities [15,32,245,246,247,248]. After treating the Zebra fish larvae placed on six six-well plates with 15, 30, and 60 μg/mL of S. vaninii mycelia extract, containing compounds such as carboxylic acid derivatives, cinnamic acid and its derivatives, and MPTP as a neuron toxin to induce dopaminergic neuron (DA) malfunction, Li et al. discovered that the 30 μg/mL group had the best performance in ameliorating the loss of DA neurons and vasculature; the 60 μg/mL group showed the most significant improvement in larva locomotor abilities [32]. Based on previous studies on relationships between oxidative stress and PD, Li confirmed that ameliorating oxidative stress assisted the anti-PD effects, such as inhibiting the mRNA expression of several PD-related genes [32,249]. Their results might bring Sanghuang to the frontier of Parkinson’s disease treatment.

7.7. Effects on Coronavirus Disease 2019 (COVID-19)

The COVID-19 pandemic is caused by severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), which is symptomized by fever, cough, and continued muscle pain [42,250,251,252]. As a recent pandemic, it has quickly been elucidated that SARS-CoV-2 uses spike protein to bind angiotensin-converting enzyme 2 (ACE2) to cross the cell membrane, facilitated by the transmembrane protease serine 2 (TMPRSS2) [42,253,254]. It is strategically reasonable to say that viral receptivity could be reduced by blocking protein binding to impede infection [255]. Chien et al. in 2022 discovered that ethanol extracts of S. sanghuang, containing several phenolic compounds listed in the article, demonstrated significant and dose-dependent inhibiting effects on ACE2 and TMPRSS2 protein expression in HepG2 cells and 293T cells after 24 h [42]. Their study also showed that the inhibiting effect of 100 mg/kg ethanol extracts in the liver and kidney did not demonstrate liver toxicity, renal toxicity, or pulmonary toxicity; these findings represent the advantages of the Sanghuang species as a potential functional food for ameliorating COVID-19 [42].

7.8. Muscle Strengthening Effect

Muscle atrophy is more prevalent among people over 60 years of age than those under 60 years of age due to naturally decreasing muscle mass and strength [256]. The most prevalent treatment for atrophy is neuromuscular electrical stimulation combined with protein [256]. Since members of the Sanghuang species present with promising biological functions, they could be used as daily food to mitigate or prevent muscle atrophy. Early in 2012, Guo et al. reported that mice treated with 400 mg/kg P. linteus polysaccharides through intragastrical administration showed significantly increased swimming time within an hour, increasing by 70.62% compared to those given saline [257]. After increasing the dose to 500 mg/kg, they discovered an increase in hepatic glycogen reserve and a reduction in urea nitrogen level in the blood serum, suggesting that the muscle tissue possessed a stronger tolerance to strenuous exercise and a lower susceptibility to atrophy [257]. To corroborate such a fatigue-relieving effect, Li et al. in 2023 orally administrated 200 mg/kg S. sanghuang ethanol extract, containing 3 mg/g hispidin, for 39 days to mice with fatigue and discovered a significant improvement in swimming time (178.7%), a 34.36% increase in liver glycogen storage, and an 18% decrease in blood urea nitrogen level [256]. They also presented results that the same concentration could regenerate the mass of muscle with atrophy by a ratio of 89.1%, which was related to MYH4 protein expression [256]. These results indicate that Sanghuang products could be potential functional food with supplement nutrients capable of improving muscle health.

8. Toxicity Studies

The effects of bioactive compounds from members of the Sanghuang species make them potential and promising functional foods, only if they are safe for individuals to take in and for cells to absorb. Recent studies have proved their clinical safety [3,24,26,66].
By orally administrating S. vaninii aqueous extract to Sprague Dawley rats to perform toxicity studies, Huo et al. reported that the maximum dose for a 21-day acute toxicity study was 21 g/kg and that the 17-week repeated toxicity test indicated 1.0 g/kg to be a safe dose due to no observed adverse effects [24]. They failed to observe any abnormalities in biochemical blood indexes, and organ weights or shapes were detected during the withdrawal and recuperating period, thereby warranting a safe passage for S. vaninii use in clinical trials [24]. Another result performed a 28-day oral toxicity study on 80 Sprague Dawley rats by giving them 1, 2, and 5 g/kg of GKSS [3]. The body weight and food intake results did not indicate differences between these three groups [3]. Li et al. also stated that, in addition to there being no deaths or noticeable changes in both sexes, S. vaninii hispidin did not cause significant abnormalities in pathology and histopathology. The dosage in these studies was higher than the ones listed in this review, suggesting a higher limit dosage for safe intake of Sanghuang products in mice. Based on some results, the dosage could be pushed to 12 mg/kg [3,26].
Besides in vivo studies to test toxicity on individuals, some researchers have also considered and tested cell toxicity [42]. Li et al. used S. typhimurium test strains to study the mutagenic effects of S. sanghuang mycelia riching hispidin and found that 3 mg/g mycelia did not induce chromosome aberrations [26]. Their results also showed that the mycelia did not induce abnormal reproduction of reticulocytes, nor did they change the proportion of micronucleated reticulocytes [26]. Back in 2014 and 2017, Zeng et al. and Lin et al. provided similar results that showed no signs of induced chromosome aberrations when testing the toxicity and mutagenic effects of P. linteus polysaccharides on bone marrow cells [258,259]. These results indicate the safety of Sanghuang products.

9. Potential to Be Functional Food

Mushrooms are prevalent and crucial in humans’ daily lives; for example, they are significantly valuable for cooking [260]. They can provide multiple kinds or forms of minerals, including but not limited to Vitamin B, potassium, and copper [261,262]. Studies in recent years have proved that many edible mushrooms qualify as supplementary or functional foods or possess tremendous medical potential [260,261,262]. Cardwell et al. revealed that mushrooms have the potential to be a convenient source for Vitamin D that could further facilitate calcium absorbance, despite necessary improvements for increasing Vitamin D concentration [262,263]. In order to be a source of nutrition, the species must contain multiple vital or valuable components and be nontoxic, prevalent, and easy to acquire for as many people as possible to use as a supplement. The progress in the past decade, as discussed above, has shown that the Sanghuang species meet the two criteria. They are prevalent and accessible in stores; promising concentrations of polysaccharides, triterpenoids, flavonoids, polyphenols, and sesquiterpenes are the main nutritional components. Additionally, several studies have proven their nontoxic characteristics [3,24,26,66,184]. Given the evidence from recent years listed above, it is logical to assume that the Sanghuang species can be used in many disease treatments or in the amelioration of dysfunction by regulating many pathways. These results propose a new direction for studying the Sanghuang species as a group of functional food.

10. Conclusions and Perspectives

Recent studies have proved that members of the Sanghuang species possess tremendous medical potential related to some significant biological components, such as triterpenoids, polysaccharides, flavonoids, and polyphenols. This article concludes the medical properties, including regulation effects on inflammation, oxidation, tumors, and immune responses. Health-improving effects are also concluded, including pulmonary protection, hypoglycemic properties, sleep improvement, gout mitigation, antiaging, neuroprotection, and muscle-strengthening abilities. Multiple in vivo studies have confirmed the safety of consuming Sanghuang components. These results demonstrate their potential to be a group of functional foods.
Multiple studies mentioned extracting components with hot water, ethanol, DES solvent, and alkaline. Hot water is the most prevalent method and it is used to extract water-soluble polysaccharides, whereas researchers use ethanol to extract polyphenols or flavonoids. DES solvent is an environmentally friendly and efficient method that has been developed in recent years. After acquiring crude samples, researchers use traditional chromatography and Sevag reagents to remove proteins or salts or rely on MMIPs and macroporous resins to isolate the desired components from crude samples specifically. Different solvents could present extracts with different antioxidant abilities, probably because different materials could dissolve in different solvents. This article also illustrates disparities in Sanghuang bioactive components because of different species and distribution regions. The extent to which these materials function medically differs, which elicits future research to discriminate and elaborate.
As the most prevalent compound, Sanghuang polysaccharides are usually neutral or acidic, including mannose or glucose, and are divided into intracellular and extracellular ones that could form flexible chains in aqueous environments. Studies focusing on structure–effect relationships have found results that suggest that some residues or structures are related to certain biological effects. Flavonoids, triterpenoids, and polyphenols are the primary components in Sanghuang, whose structures have received less attention from researchers and still elude detailed elucidation. These components, as mentioned above, demonstrate the different extents of the effects, such as flavonoids being superior in suppressing tumor growth. They are also produced in different plant parts; therefore, future researchers could isolate certain parts to separate their intended components. Despite recent discoveries that these components could act on several kinds of cells and biomolecules, their exact binding positions and mechanisms of activating pathways still elude understanding. Figure 4 concludes and illustrates the general processes of studying Sanghuang components. Table S1 in the Supplementary Materials presents some results reviewed in this article, including their intended components, extraction methods, the discovered activities, and corresponding signs. These might help future research in studying and consolidating their usage as functional foods.
Numerous studies have focused on the structures, effects, and extraction methods of Sanghuang products. However, little attention has been paid to their phylogeny. Sanghuang has been chaotically and mistakenly classified for years, with species wrongly being assigned to Inonotus and Phellinus. Such a confounding phenomenon has hampered the dissemination of information or underivative results. Despite a growing number of researchers attaching significance to their phylogenic relationships with each other, the results coming from these studies are too diverse. They comprise several versions of phylogenetic trees without compatibility, which can only provide limited assistance to relevant research. This article listed some research focusing on determining where Sanghuangporus fits in evolutionary paths and what species it might include, with methods of genome sequencing and mitochondrial genome sequencing, which might enlighten future research in making further progress in this field.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29061195/s1.

Author Contributions

Conceptualization, M.S.; methodology, J.L. and D.L.; investigation, J.L. and X.Z.; data curation, X.N.; writing—original draft preparation, J.L.; writing—review and editing, M.S. and J.L.; visualization, J.L.; supervision, M.S. and Y.W.; project administration, M.S. and Y.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by [the Science and Technology Department of Jilin Province] grant number [20220204022YY].

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

DPPH2,2-diphenyl-1-picrylhydrazyl
ABTS2,2′-azino-bis
IRinfrared
AFMatomic force microscopy
ITSinternal transcribed spacer
IBDinflammatory bowel disease
FTRIFourier transform infrared spectrometry
DESDeep eutectic solvent
DEAEdiethylaminoethanol
PLAPhelligridin LA
MIPsmolecular imprinting polymers
MMIPsmagnetic molecularly imprinted polymers
VTESVinyltriethoxysilane
Fe3O4 NPsFe3O4 nanoparticles
ROSoxygen species
CATcatalase
SODsuperoxide dismutase
TEACTrolox equivalent antioxidant
T-AOCtotal antioxidant capacity
PODperoxidase
LPSlipopolysaccharide
CTXcyclophosphamide
IKKIκB kinase
KAP1Kruppel-associated box (KRAB)-associated protein 1
RARheumatoid Arthritis
Vcvitamin C
BHTbutylated hydroxytoluene
CDKcyclin-dependent kinase

References

  1. Wang, H.; Ma, J.X.; Zhou, M.; Si, J.; Cui, B.K. Current advances and potential trends of the polysaccharides derived from medicinal mushrooms sanghuang. Front. Microbiol. 2022, 13, 965934. [Google Scholar] [CrossRef]
  2. Guo, S.; Duan, W.; Wang, Y.; Chen, L.; Yang, C.; Gu, X.; Xue, Q.; Li, R.; Zhang, Z. Component Analysis and Anti-Colorectal Cancer Mechanism via AKT/mTOR Signalling Pathway of Sanghuangporus vaninii Extracts. Molecules 2022, 27, 1153. [Google Scholar] [CrossRef]
  3. Li, I.C.; Chang, F.C.; Kuo, C.C.; Chu, H.T.; Li, T.J.; Chen, C.C. Pilot Study: Nutritional and Preclinical Safety Investigation of Fermented Hispidin-Enriched Sanghuangporus sanghuang Mycelia: A Promising Functional Food Material to Improve Sleep. Front. Nutr. 2021, 8, 788965. [Google Scholar] [CrossRef]
  4. Yu, T.; Zhong, S.; Sun, Y.; Sun, H.; Chen, W.; Li, Y.; Zhu, J.; Lu, L.; Huo, J. Aqueous extracts of Sanghuangporus vaninii induce S-phase arrest and apoptosis in human melanoma A375 cells. Oncol. Lett. 2021, 22, 628. [Google Scholar] [CrossRef]
  5. Ma, X.K.; She, X.; Peterson, E.C.; Wang, Y.Z.; Zheng, P.; Ma, H.; Zhang, K.; Liang, J. A newly characterized exopolysaccharide from Sanghuangporus sanghuang. J. Microbiol. 2019, 57, 812–820. [Google Scholar] [CrossRef] [PubMed]
  6. Zhou, Q.; Wang, J.; Jiang, H.; Wang, G.; Wang, Y. Deep sequencing of the Sanghuangporus vaninii transcriptome reveals dynamic landscapes of candidate genes involved in the biosynthesis of active compounds. Arch. Microbiol. 2021, 203, 2315–2324. [Google Scholar] [CrossRef] [PubMed]
  7. Han, J.G.; Hyun, M.W.; Kim, C.S.; Jo, J.W.; Cho, J.H.; Lee, K.H.; Kong, W.S.; Han, S.K.; Oh, J.; Sung, G.H. Species identity of Phellinus linteus (sanghuang) extensively used as a medicinal mushroom in Korea. J. Microbiol. 2016, 54, 290–295. [Google Scholar] [CrossRef]
  8. Song, T.; Xu, F.; Shen, Y.; Fan, L.; Cai, W. The complete mitochondrial genome of Sanghuangporus vaninii Zhehuang-1 (Hymenochaetales, Basidiomycota). Mitochondrial DNA B Resour. 2021, 6, 1096–1097. [Google Scholar] [CrossRef] [PubMed]
  9. Zuo, K.; Tang, K.; Liang, Y.; Xu, Y.; Sheng, K.; Kong, X.; Wang, J.; Zhu, F.; Zha, X.; Wang, Y. Purification and antioxidant and anti-Inflammatory activity of extracellular polysaccharopeptide from sanghuang mushroom, Sanghuangporus lonicericola. J. Sci. Food Agric. 2021, 101, 1009–1020. [Google Scholar] [CrossRef]
  10. Kozarski, M.; Klaus, A.; Niksic, M.; Jakovljevic, D.; Helsper, J.P.F.G.; Van Griensven, L.J.L.D. Antioxidative and immunomodulating activities of polysaccharide extracts of the medicinal mushrooms Agaricus bisporus, Agaricus brasiliensis, Ganoderma lucidum and Phellinus linteus. Food Chem. 2011, 129, 1667–1675. [Google Scholar] [CrossRef]
  11. Li, T.; Yang, Y.; Liu, Y.; Zhou, S.; Yan, M.Q.; Wu, D.; Zhang, J.; Tang, C. Physicochemical characteristics and biological activities of polysaccharide fractions from Phellinus baumii cultured with different methods. Int. J. Biol. Macromol. 2015, 81, 1082–1088. [Google Scholar] [CrossRef]
  12. Lin, W.C.; Deng, J.S.; Huang, S.S.; Wu, S.H.; Chen, C.C.; Lin, W.R.; Lin, H.Y.; Huang, G.J. Anti-Inflammatory Activity of Sanghuangporus sanghuang Mycelium. Int. J. Mol. Sci. 2017, 18, 347. [Google Scholar] [CrossRef] [PubMed]
  13. Lin, W.-C.; Deng, J.-S.; Huang, S.-S.; Lin, W.-R.; Wu, S.-H.; Lin, H.-Y.; Huang, G.-J. Anti-inflammatory activity of Sanghuangporus sanghuang by suppressing the TLR4-mediated PI3K/AKT/mTOR/IKKβ signaling pathway. RSC Adv. 2017, 7, 21234–21251. [Google Scholar] [CrossRef]
  14. Wen, Y.; Wan, Y.Z.; Qiao, C.X.; Xu, X.F.; Wang, J.; Shen, Y. Immunoregenerative effects of the bionically cultured Sanghuang mushrooms (Inonotus sanghuagn) on the immunodeficient mice. J. Ethnopharmacol. 2019, 245, 112047. [Google Scholar] [CrossRef] [PubMed]
  15. Huo, J.; Sun, Y.; Pan, M.; Ma, H.; Lin, T.; Lv, Z.; Li, Y.; Zhong, S. Non-targeted metabonomics and transcriptomics revealed the mechanism of mulberry branch extracts promoting the growth of Sanghuangporus vaninii mycelium. Front. Microbiol. 2022, 13, 1024987. [Google Scholar] [CrossRef] [PubMed]
  16. Liu, Z.; Liu, R.; Tong, X.; Zou, L. New Insights into Methyl Jasmonate Regulation of Triterpenoid Biosynthesis in Medicinal Fungal Species Sanghuangporus baumii (Pilat) L.W. Zhou & Y.C. Dai. J. Fungi 2022, 8, 889. [Google Scholar] [CrossRef]
  17. Ma, Y.; Gao, W.; Zhang, F.; Zhu, X.; Kong, W.; Niu, S.; Gao, K.; Yang, H. Community composition and trophic mode diversity of fungi associated with fruiting body of medicinal Sanghuangporus vaninii. BMC Microbiol. 2022, 22, 251. [Google Scholar] [CrossRef] [PubMed]
  18. Zhang, M.; Xie, Y.; Su, X.; Liu, K.; Zhang, Y.; Pang, W.; Wang, J. Inonotus sanghuang Polyphenols Attenuate Inflammatory Response Via Modulating the Crosstalk Between Macrophages and Adipocytes. Front. Immunol. 2019, 10, 286. [Google Scholar] [CrossRef] [PubMed]
  19. Tang, S.; Jin, L.; Lei, P.; Shao, C.; Wu, S.; Yang, Y.; He, Y.; Ren, R.; Xu, J. Whole-genome assembly and analysis of a medicinal fungus: Inonotus hispidus. Front. Microbiol. 2022, 13, 967135. [Google Scholar] [CrossRef]
  20. He, P.Y.; Hou, Y.H.; Yang, Y.; Li, N. The anticancer effect of extract of medicinal mushroom Sanghuangprous vaninii against human cervical cancer cell via endoplasmic reticulum stress-mitochondrial apoptotic pathway. J. Ethnopharmacol. 2021, 279, 114345. [Google Scholar] [CrossRef]
  21. Lyu, G.; Song, T.; Cai, W.; Zhang, Z. Comparative study of chemical components and antioxidant activities of wild Sanghuangporus sanghuang and Sanghuangporus vaninii. Mycosystema 2021, 40, 1833–1843. [Google Scholar] [CrossRef]
  22. Yuan, M.; Fu, T.; Xu, C.; Chen, A.; Shao, Y. The Optimum Extracting Technology and Antioxidant Activity of Polysaccharides from Fruiting Body of Phellinus igniarius. Food Ind. 2021, 42, 103–108. [Google Scholar]
  23. Li, X.; Xie, Y.; Wang, H.; Li, J.; Liu, J.; Hu, L.; Wang, S. Comparison of chemical constituents and antioxidant activities of Sanghuangporus vaninii in different substrates. J. Food Saf. Qual. 2021, 12, 9183–9188. [Google Scholar] [CrossRef]
  24. Huo, J.; Sun, Y.; Zhong, S.; Li, Y.; Yang, R.; Xia, L.; Wang, J.; Zhang, M.; Zhu, J. Safety evaluation of aqueous extracts of Sanghuangporus vaninii fruiting body in Sprague-Dawley rats. Food Sci. Nutr. 2020, 8, 5107–5113. [Google Scholar] [CrossRef] [PubMed]
  25. Song, T.; Zhang, Z.; Jin, Q.; Feng, W.; Shen, Y.; Fan, L.; Cai, W. Nutrient profiles, functional compositions, and antioxidant activities of seven types of grain fermented with Sanghuangporus sanghuang fungus. J. Food Sci. Technol. 2021, 58, 4091–4101. [Google Scholar] [CrossRef] [PubMed]
  26. Li, I.C.; Chen, C.C.; Sheu, S.J.; Huang, I.H.; Chen, C.C. Optimized production and safety evaluation of hispidin-enriched Sanghuangporus sanghuang mycelia. Food Sci. Nutr. 2020, 8, 1864–1873. [Google Scholar] [CrossRef]
  27. Xia, Y.; Yang, C.; Liu, X.; Wang, G.; Xiong, Z.; Song, X.; Yang, Y.; Zhang, H.; Ai, L. Enhancement of triterpene production via in situ extractive fermentation of Sanghuangporus vaninii YC-1. Biotechnol. Appl. Biochem. 2021, 69, 2561–2572. [Google Scholar] [CrossRef] [PubMed]
  28. Shen, Q.H.; Huang, Q.; Xie, J.Y.; Wang, K.; Qian, Z.M.; Li, D.Q. A rapid analysis of antioxidants in Sanghuangporus baumii by online extraction-HPLC-ABTS. RSC Adv. 2021, 11, 25646–25652. [Google Scholar] [CrossRef]
  29. Su, X.; Liu, K.; Xie, Y.; Zhang, M.; Wu, X.; Zhang, Y.; Wang, J. Mushroom Inonotus sanghuang alleviates experimental pulmonary fibrosis: Implications for therapy of pulmonary fibrosis. Biomed. Pharmacother. 2021, 133, 110919. [Google Scholar] [CrossRef]
  30. Guo, Q.; Zhao, L.; Zhu, Y.; Wu, J.; Hao, C.; Song, S.; Shi, W. Optimization of culture medium for Sanghuangporus vaninii and a study on its therapeutic effects on gout. Biomed. Pharmacother. 2021, 135, 111194. [Google Scholar] [CrossRef]
  31. Liu, S.; Mo, J.; Pan, D.; Liu, G. Research Progress on Pharmacological Actions and Extraction Methods of Polysaccharide from Phellinus igniarius. Biotechnol. Bull. 2018, 34, 63–67. [Google Scholar] [CrossRef]
  32. Li, X.; Gao, D.; Paudel, Y.N.; Li, X.; Zheng, M.; Liu, G.; Ma, Y.; Chu, L.; He, F.; Jin, M. Anti-Parkinson’s Disease Activity of Sanghuangprous vaninii Extracts in the MPTP-Induced Zebrafish Model. ACS Chem. Neurosci. 2022, 13, 330–339. [Google Scholar] [CrossRef]
  33. Jiang, J.H.; Wu, S.H.; Zhou, L.W. The First Whole Genome Sequencing of Sanghuangporus sanghuang Provides Insights into Its Medicinal Application and Evolution. J. Fungi 2021, 7, 787. [Google Scholar] [CrossRef]
  34. Yang, K.; Zhang, S.; Geng, Y.; Tian, B.; Cai, M.; Guan, R.; Li, Y.; Ye, B.; Sun, P. Anti-Inflammatory Properties In Vitro and Hypoglycaemic Effects of Phenolics from Cultivated Fruit Body of Phellinus baumii in Type 2 Diabetic Mice. Molecules 2021, 26, 2285. [Google Scholar] [CrossRef] [PubMed]
  35. Zheng, N.; Ming, Y.; Chu, J.; Yang, S.; Wu, G.; Li, W.; Zhang, R.; Cheng, X. Optimization of Extraction Process and the Antioxidant Activity of Phenolics from Sanghuangporus baumii. Molecules 2021, 26, 3850. [Google Scholar] [CrossRef]
  36. Zhang, Y.; Xu, J.; Yang, J.; He, K.; Li, Y.; Duan, Y. Progresses on the effects of Phellinus linteus on tumors. Jiangsu Sci. Technol. Inf. 2021, 38, 60–64. [Google Scholar]
  37. Cai, C.; Ma, J.; Han, C.; Jin, Y.; Zhao, G.; He, X. Extraction and antioxidant activity of total triterpenoids in the mycelium of a medicinal fungus, Sanghuangporus sanghuang. Sci. Rep. 2019, 9, 7418. [Google Scholar] [CrossRef] [PubMed]
  38. He, S.; Bao, H.; Wei, Y.; Liu, Y.; Liu, J. Antitumor effect and mechanism of different extracts of cultivated Phellinus vaninii on H22 tumor bearing mice. Chin. J. Biotech. 2022, 38, 1025–1038. [Google Scholar] [CrossRef]
  39. Guo, D.; Ma, H.; Cao, Y.; Wang, W.; Wang, Y.; Wang, W. Ultrasonic wave–assisted Extraction of Active Substances from Phellinus igniarius. J. Chin. Inst. Food Sci. Technol. 2015, 15, 87–92. [Google Scholar] [CrossRef]
  40. Lin, W.-C.; Deng, J.-S.; Huang, S.-S.; Wu, S.-H.; Lin, H.-Y.; Huang, G.-J. Evaluation of antioxidant, anti-inflammatory and anti-proliferative activities of ethanol extracts from different varieties of Sanghuang species. RSC Adv. 2017, 7, 7780–7788. [Google Scholar] [CrossRef]
  41. Song, M.; Park, H.J. 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, 311–318. [Google Scholar] [CrossRef] [PubMed]
  42. Chien, L.H.; Deng, J.S.; Jiang, W.P.; Chen, C.C.; Chou, Y.N.; Lin, J.G.; Huang, G.J. Study on the potential of Sanghuangporus sanghuang and its components as COVID-19 spike protein receptor binding domain inhibitors. Biomed. Pharmacother. 2022, 153, 113434. [Google Scholar] [CrossRef] [PubMed]
  43. Zhang, Y.; Zhang, Z.; Song, T.; Chen, C.; Cai, W.; Lyu, G. Optimization of extraction process and examination of antioxidant activities of polyphenols extracted from Sanghuangporus vaninii by using deep eutectic solvent. Mycosystema 2022, 41, 1694–1703. [Google Scholar]
  44. Yuan, W.; Zhou, R.; Sun, M.; Yang, H.; Lu, T.; Zhao, Y.; Zheng, W. Effects of nitric oxide on the accumulation of polyphenols and triterpenoids in the submerged cultures of Sanghuangporus vaninii. Mycosystema 2023, 42, 932–938. [Google Scholar]
  45. Wang, K.; Ding, Z.; Pei, J.; Yan, J. Antioxidant Activities of Polysaccharides from Phellinus linteus Mycelia by Alkaline Extraction. Sci. Technol. Food Ind. 2020, 41, 289–294. [Google Scholar] [CrossRef]
  46. Li, R.; Wang, Y.; Xia, J.; Luo, D.; Wang, T. Optimization of Extraction Process of Polysaccharide of Phellinus igniarius Mycelium and Analysis of Its Antioxidant Activity in vitro. Chin. Agric. Sci. Bull. 2019, 35, 143–150. [Google Scholar]
  47. Cheng, J.; Song, J.; Wei, H.; Wang, Y.; Huang, X.; Liu, Y.; Lu, N.; He, L.; Lv, G.; Ding, H.; et al. Structural characterization and hypoglycemic activity of an intracellular polysaccharide from Sanghuangporus sanghuang mycelia. Int. J. Biol. Macromol. 2020, 164, 3305–3314. [Google Scholar] [CrossRef]
  48. Cheng, J.; Song, J.; Liu, Y.; Lu, N.; Wang, Y.; Hu, C.; He, L.; Wei, H.; Lv, G.; Yang, S.; et al. Conformational properties and biological activities of alpha-D-mannan from Sanghuangporus sanghuang in liquid culture. Int. J. Biol. Macromol. 2020, 164, 3568–3579. [Google Scholar] [CrossRef]
  49. Wan, X.; Jin, X.; Xie, M.; Liu, J.; Gontcharov, A.A.; Wang, H.; Lv, R.; Liu, D.; Wang, Q.; Li, Y. Characterization of a polysaccharide from Sanghuangporus vaninii and its antitumor regulation via activation of the p53 signaling pathway in breast cancer MCF-7 cells. Int. J. Biol. Macromol. 2020, 163, 865–877. [Google Scholar] [CrossRef]
  50. Chen, H.; Tian, T.; Miao, H.; Zhao, Y.Y. Traditional uses, fermentation, phytochemistry and pharmacology of Phellinus linteus: A review. Fitoterapia 2016, 113, 6–26. [Google Scholar] [CrossRef]
  51. Wang, S.; Liu, Z.; Wang, X.; Liu, R.; Zou, L. Mushrooms Do Produce Flavonoids: Metabolite Profiling and Transcriptome Analysis of Flavonoid Synthesis in the Medicinal Mushroom Sanghuangporus baumii. J. Fungi 2022, 8, 582. [Google Scholar] [CrossRef]
  52. Duan, Y.; Han, H.; Qi, J.; Gao, J.M.; Xu, Z.; Wang, P.; Zhang, J.; Liu, C. Genome sequencing of Inonotus obliquus reveals insights into candidate genes involved in secondary metabolite biosynthesis. BMC Genom. 2022, 23, 314. [Google Scholar] [CrossRef] [PubMed]
  53. Wang, T.; Wang, G.; Zhang, G.; Hou, R.; Zhou, L.; Tian, X. Systematic analysis of the lysine malonylome in Sanghuangporus sanghuang. BMC Genom. 2021, 22, 840. [Google Scholar] [CrossRef] [PubMed]
  54. Wang, X.; Wang, S.; Xu, X.; Sun, J.; Ma, Y.; Liu, Z.; Sun, T.; Zou, L. Molecular cloning, characterization, and heterologous expression of an acetyl-CoA acetyl transferase gene from Sanghuangporus baumii. Protein Expr. Purif. 2020, 170, 105592. [Google Scholar] [CrossRef] [PubMed]
  55. Wan, X.; Jin, X.; Wu, X.; Yang, X.; Lin, D.; Li, C.; Fu, Y.; Liu, Y.; Liu, X.; Lv, J.; et al. Structural characterisation and antitumor activity against non-small cell lung cancer of polysaccharides from Sanghuangporus vaninii. Carbohydr. Polym. 2022, 276, 118798. [Google Scholar] [CrossRef] [PubMed]
  56. Kim, G.-Y.; Park, S.-K.; Lee, M.-K.; Lee, S.-H.; Oh, Y.-H.; Kwak, J.-Y.; Yoon, S.; Lee, J.-D.; Park, Y.-M. Proteoglycan isolated from Phellinus linteus activates murine B lymphocytes via protein kinase C and protein tyrosine kinase. Int. Immunopharmacol. 2003, 3, 1281–1292. [Google Scholar] [CrossRef] [PubMed]
  57. Lu, J.; He, R.; Sun, P.; Zhang, F.; Linhardt, R.J.; Zhang, A. Molecular mechanisms of bioactive polysaccharides from Ganoderma lucidum (Lingzhi), a review. Int. J. Biol. Macromol. 2020, 150, 765–774. [Google Scholar] [CrossRef] [PubMed]
  58. Qu, J.; Huang, P.; Zhang, L.; Qiu, Y.; Qi, H.; Leng, A.; Shang, D. Hepatoprotective effect of plant polysaccharides from natural resources: A review of the mechanisms and structure-activity relationship. Int. J. Biol. Macromol. 2020, 161, 24–34. [Google Scholar] [CrossRef] [PubMed]
  59. Shon, Y.H.; Nam, K.S. Inhibition of cytochrome P450 isozymes in rat liver microsomes by polysaccharides derived from Phellinus linteus. Biotechnol. Lett. 2003, 25, 167–172. [Google Scholar] [CrossRef] [PubMed]
  60. Yin, F.; Xiang, R.; Jing, L.; Chen, M.; Guo, W.; Zhang, C.; Jiang, F. Research Progresses of the Hypolipidemic and Antioxidant Effects of Phellinus fungi. J. Chin. Med. Mater. 2018, 41, 2721–2725. [Google Scholar] [CrossRef]
  61. Guo, J.; Liu, X.; Li, Y.; Ji, H.; Liu, C.; Zhou, L.; Huang, Y.; Bai, C.; Jiang, Z.; Wu, X. Screening for proteins related to the biosynthesis of hispidin and its derivatives in Phellinus igniarius using iTRAQ proteomic analysis. BMC Microbiol. 2021, 21, 81. [Google Scholar] [CrossRef]
  62. Jang, J.S.; Lee, J.S.; Lee, J.H.; Kwon, D.S.; Lee, K.E.; Lee, S.Y.; Hong, E.K. Hispidin produced from Phellinus linteus protects pancreatic beta-cells from damage by hydrogen peroxide. Arch. Pharm. Res. 2010, 33, 853–861. [Google Scholar] [CrossRef]
  63. Park, I.-H.; Chung, S.-K.; Lee, K.-B.; Yoo, Y.-C.; Kim, S.-K.; Kim, G.-S.; Song, K.-S. An antioxidant hispidin from the mycelial cultures of Phellinus linteus. Arch. Pharmacal Res. 2004, 27, 615–618. [Google Scholar] [CrossRef] [PubMed]
  64. Zhang, W.; Luo, Y.; Xie, Z.; Kong, C.; Na, Z. Extraction and detection of morin from Sanghuangporus lonicericola by magnetic molecularly imprinted polymers coupled with HPLC analysis. J. Food Sci. 2022, 87, 1575–1585. [Google Scholar] [CrossRef]
  65. Rajput, S.A.; Wang, X.Q.; Yan, H.C. Morin hydrate: A comprehensive review on novel natural dietary bioactive compound with versatile biological and pharmacological potential. Biomed. Pharmacother. 2021, 138, 111511. [Google Scholar] [CrossRef] [PubMed]
  66. Heo, Y.M.; Kim, K.; Kwon, S.L.; Na, J.; Lee, H.; Jang, S.; Kim, C.H.; Jung, J.; Kim, J.J. Investigation of Filamentous Fungi Producing Safe, Functional Water-Soluble Pigments. Mycobiology 2018, 46, 269–277. [Google Scholar] [CrossRef]
  67. Duan, H.; Pan, J.; Guo, M.; Li, J.; Yu, L.; Fan, L. Dietary strategies with anti-aging potential: Dietary patterns and supplements. Food Res. Int. 2022, 158, 111501. [Google Scholar] [CrossRef] [PubMed]
  68. Atlante, A.; Amadoro, G.; Bobba, A.; Latina, V. Functional Foods: An Approach to Modulate Molecular Mechanisms of Alzheimer’s Disease. Cells 2020, 9, 2347. [Google Scholar] [CrossRef] [PubMed]
  69. Serafini, M.; Peluso, I. Functional Foods for Health: The Interrelated Antioxidant and Anti-Inflammatory Role of Fruits, Vegetables, Herbs, Spices and Cocoa in Humans. Curr. Pharm. Des. 2017, 22, 6701–6715. [Google Scholar] [CrossRef] [PubMed]
  70. Chaturvedi, V.K.; Agarwal, S.; Gupta, K.K.; Ramteke, P.W.; Singh, M.P. Medicinal mushroom: Boon for therapeutic applications. 3 Biotech 2018, 8, 334. [Google Scholar] [CrossRef]
  71. Hussain, J.; Ali, L.; Khan, A.L.; Rehman, N.U.; Jabeen, F.; Kim, J.S.; Al-Harrasi, A. Isolation and bioactivities of the flavonoids morin and morin-3-O-β-D-glucopyranoside from Acridocarpus orientalis—A wild Arabian medicinal plant. Molecules 2014, 19, 17763–17772. [Google Scholar] [CrossRef]
  72. Liu, M.; Zeng, J. Research Progress on Chemical Constituents of Phellinus and Its Antitumor Mechanism. Chin. J. Exp. Tradit. Med. Formulae 2023, 29, 1–9. [Google Scholar] [CrossRef]
  73. Wang, W.; Song, J.; Yan, J.; Lu, N.; Fu, L.; Yuan, W.; Zhou, Z. Effects of harvest time on content of active components and antioxidant activities of fruiting bodies of Sanghuangporus cultivated with substitute materials. Mycosystema 2020, 39, 2369–2379. [Google Scholar] [CrossRef]
  74. Han, J.G.; Oh, J.; Jo, J.W.; Kim, C.S.; Kwag, Y.N.; Han, S.K.; Sung, G.H. The complete mitochondrial genome of Sanghuangporus sanghuang (Hymenochaetaceae, Basidiomycota). Mitochondrial DNA B Resour. 2018, 3, 456–457. [Google Scholar] [CrossRef]
  75. Lim, Y.W.; Lee, J.S.; Jung, H.S. Type studies on Phellinus baumii and Phellinus linteus. Mycotaxon 2003, 85, 201–210. [Google Scholar]
  76. Zhou, L.-W.; Vlasák, J.; Decock, C.; Assefa, A.; Stenlid, J.; Abate, D.; Wu, S.-H.; Dai, Y.-C. Global diversity and taxonomy of the Inonotus linteus complex (Hymenochaetales, Basidiomycota): Sanghuangporus gen. nov., Tropicoporus excentrodendri and T. guanacastensis gen. et spp. nov., and 17 new combinations. Fungal Divers. 2015, 77, 335–347. [Google Scholar] [CrossRef]
  77. Ma, Y.J.; Gao, W.Q.; Zhu, X.T.; Kong, W.B.; Zhang, F.; Yang, H.Q. Identification and profiling of the community structure and potential function of bacteria from the fruiting bodies of Sanghuangporus vaninii. Arch. Microbiol. 2022, 204, 564. [Google Scholar] [CrossRef] [PubMed]
  78. Cheng, T.; Chepkirui, C.; Decock, C.; Matasyoh, J.C.; Stadler, M. Sesquiterpenes from an Eastern African Medicinal Mushroom Belonging to the Genus Sanghuangporus. J. Nat. Prod. 2019, 82, 1283–1291. [Google Scholar] [CrossRef] [PubMed]
  79. Agnestisia, R.; Ono, A.; Nakamura, L.; Chino, R.; Nodera, K.; Aiso-Sanada, H.; Nezu, I.; Ishiguri, F.; Suzuki, T.; Yokota, S. The complete mitochondrial genome sequence of the medicinal fungus Inonotus obliquus (Hymenochaetaceae, Basidiomycota). Mitochondrial DNA B Resour. 2019, 4, 3504–3506. [Google Scholar] [CrossRef] [PubMed]
  80. Shen, S.; Liu, S.L.; Jiang, J.H.; Zhou, L.W. Addressing widespread misidentifications of traditional medicinal mushrooms in Sanghuangporus (Basidiomycota) through ITS barcoding and designation of reference sequences. IMA Fungus 2021, 12, 10. [Google Scholar] [CrossRef] [PubMed]
  81. Wu, S.H.; Chang, C.C.; Wei, C.L.; Jiang, G.Z.; Cui, B.K. Sanghuangporus toxicodendri sp. nov. (Hymenochaetales, Basidiomycota) from China. MycoKeys 2019, 57, 101–111. [Google Scholar] [CrossRef]
  82. Francis, A.; Huber, K.T.; Moulton, V. Tree-Based Unrooted Phylogenetic Networks. Bull. Math. Biol. 2018, 80, 404–416. [Google Scholar] [CrossRef]
  83. McLennan, D.A. How to Read a Phylogenetic Tree. Evol. Educ. Outreach 2010, 3, 506–519. [Google Scholar] [CrossRef]
  84. Tian, X.-M.; Yu, H.-Y.; Zhou, L.-W.; Decock, C.; Vlasák, J.; Dai, Y.-C. Phylogeny and taxonomy of the Inonotus linteus complex. Fungal Divers. 2012, 58, 159–169. [Google Scholar] [CrossRef]
  85. Zhu, L.; Song, J.; Zhou, J.L.; Si, J.; Cui, B.K. Species Diversity, Phylogeny, Divergence Time, and Biogeography of the Genus Sanghuangporus (Basidiomycota). Front. Microbiol. 2019, 10, 812. [Google Scholar] [CrossRef]
  86. Li, Y.; Zhang, Q.; Chen, X.; Zhang, A. Research progress in purification, bio-activity and product development of polysaccharides from Sanghuangporus. Edible Med. Mushrooms 2022, 30, 26–31+61. [Google Scholar]
  87. Sun, Y.; Huo, J.; Zhong, S.; Zhu, J.; Li, Y.; Li, X. Chemical structure and anti-inflammatory activity of a branched polysaccharide isolated from Phellinus baumii. Carbohydr. Polym. 2021, 268, 118214. [Google Scholar] [CrossRef] [PubMed]
  88. Yan, J.-K.; Wang, Y.-Y.; Ma, H.-L.; Wang, Z.-B.; Pei, J.-J. Structural characteristics and antioxidant activity in vivo of a polysaccharide isolated from Phellinus linteus mycelia. J. Taiwan Inst. Chem. Eng. 2016, 65, 110–117. [Google Scholar] [CrossRef]
  89. Wei, C.Y.; Li, W.Q.; Shao, S.S.; He, L.; Cheng, J.; Han, S.; Liu, Y. Structure and chain conformation of a neutral intracellular heteropolysaccharide from mycelium of Paecilomyces cicadae. Carbohydr. Polym. 2016, 136, 728–737. [Google Scholar] [CrossRef] [PubMed]
  90. Yuan, Q.; Zhao, L.; Li, Z.; Harqin, C.; Peng, Y.; Liu, J. Physicochemical analysis, structural elucidation and bioactivities of a high-molecular-weight polysaccharide from Phellinus igniarius mycelia. Int. J. Biol. Macromol. 2018, 120, 1855–1864. [Google Scholar] [CrossRef] [PubMed]
  91. Galinari, E.; Sabry, D.A.; Sassaki, G.L.; Macedo, G.R.; Passos, F.M.L.; Mantovani, H.C.; Rocha, H.A.O. Chemical structure, antiproliferative and antioxidant activities of a cell wall alpha-d-mannan from yeast Kluyveromyces marxianus. Carbohydr. Polym. 2017, 157, 1298–1305. [Google Scholar] [CrossRef] [PubMed]
  92. Li, S.C.; Yang, X.M.; Ma, H.L.; Yan, J.K.; Guo, D.Z. Purification, characterization and antitumor activity of polysaccharides extracted from Phellinus igniarius mycelia. Carbohydr. Polym. 2015, 133, 24–30. [Google Scholar] [CrossRef] [PubMed]
  93. Chen, S.; Huang, H.; Huang, G. Extraction, derivatization and antioxidant activity of cucumber polysaccharide. Int. J. Biol. Macromol. 2019, 140, 1047–1053. [Google Scholar] [CrossRef]
  94. Liu, B.; Shang, Z.Z.; Li, Q.M.; Zha, X.Q.; Wu, D.L.; Yu, N.J.; Han, L.; Peng, D.Y.; Luo, J.P. Structural features and anti-gastric cancer activity of polysaccharides from stem, root, leaf and flower of cultivated Dendrobium huoshanense. Int. J. Biol. Macromol. 2020, 143, 651–664. [Google Scholar] [CrossRef]
  95. Guo, S.; Rausch, W.D. Compare of antioxidant and anti-inflammatory activity of different Phellinus. Chin. J. Pharmacovigil. 2017, 14, 394–398. [Google Scholar]
  96. Yan, H.; Jin, Q.; Li, P.; Yu, T.; Han, C.; Cui, C. Extraction and antioxidation of polysaccharide from Phellinus linteus. Food Sci. Technol. 2013, 38, 201–205. [Google Scholar] [CrossRef]
  97. Cheng, J.W.; He, L.; Wei, H.L.; Song, J.L.; Hu, C.J.; Zou, J.Q.; Li, H.B.; Jiang, Y.H. Ultrasonic-assisted Enzymatic Method Extraction of Polysaccharide from Inonotus sanghuang. J. Zhejiang For. Sci. Tech. 2017, 37, 33–38. [Google Scholar]
  98. Gong, L.; Liu, G.; Bai, J.; Liu, D.; He, Y. Optimization of Extraction of Phellinus igniarius and Antioxidant and Antibacterial ActivitieS of Lyophilized Powder. Food Res. Dev. 2021, 42, 143–149. [Google Scholar]
  99. Xu, Y.; Wang, J.; Yin, X.; Jin, L.; Liu, J.; Zhang, C. Optimization of extraction process of Phellinus linteus polysaccharide. J. Med. Sci. Yanbian Univ. 2021, 44, 242–244. [Google Scholar] [CrossRef]
  100. Shi, Z.; Bao, H. Progresses on extraction methods of polysaccharides of Phellinus fungi. North. Hortic. 2017, 01, 191–195. [Google Scholar]
  101. Zhang, J.-J.; Chen, B.-S.; Dai, H.-Q.; Ren, J.-W.; Zhou, L.-W.; Wu, S.-H.; Liu, H.-W. Sesquiterpenes and polyphenols with glucose-uptake stimulatory and antioxidant activities from the medicinal mushroom Sanghuangporus sanghuang. Chin. J. Nat. Med. 2021, 19, 693–699. [Google Scholar] [CrossRef]
  102. Li, Z.; Lang, K.; Zhang, Z.; Wang, Y.; Wang, Q. Research on Process of Ultrasonic Assisted Extraction of Phellinus igniarius Polysaccharide. Agric. Sci. Technol. Equip. 2018, 02, 42–43. [Google Scholar] [CrossRef]
  103. Abbott, A.P.; Capper, G.; Davies, D.L.; Rasheed, R.K.; Tambyrajah, V. Novel solvent properties of choline chloride/urea mixtures. Chem. Commun. 2003, 39, 70–71. [Google Scholar] [CrossRef]
  104. Jia, D.; Qin, F.; Liu, B.; Tang, Y.; Chen, W. Effect of dietary supplementation of Phellinus linteus polysaccharide on lipopolysaccharide induced production performance serum biochemistry, immunity and antioxidant capacity of laying hens. Feed Res. 2022, 45, 46–50. [Google Scholar] [CrossRef]
  105. Pei, J.J.; Wang, Z.B.; Ma, H.L.; Yan, J.K. Structural features and antitumor activity of a novel polysaccharide from alkaline extract of Phellinus linteus mycelia. Carbohydr. Polym. 2015, 115, 472–477. [Google Scholar] [CrossRef]
  106. Cui, S.; Zeng, P.; Lang, M.; Xie, H.; Chen, Y.; Li, C.; Shen, Z.; Li, Y.; Shi, L. Study on the Antioxidant Effects of the Polysaccharide from Phellinus baumii in Vitro. Bull. Seric. 2019, 50, 12–14+22. [Google Scholar]
  107. Qiu, P.; Liu, J.; Zhao, L.; Zhang, P.; Wang, W.; Shou, D.; Ji, J.; Li, C.; Chai, K.; Dong, Y. Inoscavin A, a pyrone compound isolated from a Sanghuangporus vaninii extract, inhibits colon cancer cell growth and induces cell apoptosis via the hedgehog signaling pathway. Phytomedicine 2022, 96, 153852. [Google Scholar] [CrossRef] [PubMed]
  108. Zhang, S.P.; Du, X.X.; Zhou, Q.S.; Jie, M.; Hui, L. Study on Deproteinization and Decoloration in Fucoidan. Adv. Mater. Res. 2011, 354–355, 62–66. [Google Scholar] [CrossRef]
  109. Li, X.; Zhao, R.; Zhou, H.L.; Wu, D.H. Deproteinization of Polysaccharide from the Stigma Maydis by Sevag Method. Adv. Mater. Res. 2011, 340, 416–420. [Google Scholar] [CrossRef]
  110. Wang, X.; Yuan, Y.; Wang, K.; Zhang, D.; Yang, Z.; Xu, P. Deproteinization of gellan gum produced by Sphingomonas paucimobilis ATCC 31461. J. Biotechnol. 2007, 128, 403–407. [Google Scholar] [CrossRef] [PubMed]
  111. Ran, L.; Zou, Y.; Cheng, J.; Lu, F. Silver nanoparticles in situ synthesized by polysaccharides from Sanghuangporus sanghuang and composites with chitosan to prepare scaffolds for the regeneration of infected full-thickness skin defects. Int. J. Biol. Macromol. 2019, 125, 392–403. [Google Scholar] [CrossRef]
  112. Cheng, J.; Song, J.; Wang, Y.; Wei, H.; He, L.; Liu, Y.; Ding, H.; Huang, Q.; Hu, C.; Huang, X.; et al. Conformation and anticancer activity of a novel mannogalactan from the fruiting bodies of Sanghuangporus sanghuang on HepG2 cells. Food Res. Int. 2022, 156, 111336. [Google Scholar] [CrossRef]
  113. Liang, L.; Liu, G.; Yu, G.; Song, Y.; Li, Q. Simultaneous decoloration and purification of crude oligosaccharides from pumpkin (Cucurbita moschata Duch) by macroporous adsorbent resin. Food Chem. 2018, 277, 744–752. [Google Scholar] [CrossRef]
  114. Li, J.; Shi, S.; Tu, M.; Via, B.; Sun, F.F.; Adhikari, S. Detoxification of Organosolv-Pretreated Pine Prehydrolysates with Anion Resin and Cysteine for Butanol Fermentation. Appl. Biochem. Biotechnol. 2018, 186, 662–680. [Google Scholar] [CrossRef] [PubMed]
  115. Liu, J.; Luo, J.; Sun, Y.; Ye, H.; Lu, Z.; Zeng, X. A simple method for the simultaneous decoloration and deproteinization of crude levan extract from Paenibacillus polymyxa EJS-3 by macroporous resin. Bioresour. Technol. 2010, 101, 6077–6083. [Google Scholar] [CrossRef] [PubMed]
  116. Li, J.; Chase, H.A. Development of adsorptive (non-ionic) macroporous resins and their uses in the purification of pharmacologically-active natural products from plant sources. Nat. Prod. Rep. 2010, 27, 1493–1510. [Google Scholar] [CrossRef] [PubMed]
  117. Song, M.; Jiao, P.; Qin, T.; Jiang, K.; Zhou, J.; Zhuang, W.; Chen, Y.; Liu, D.; Zhu, C.; Chen, X.; et al. Recovery of lactic acid from the pretreated fermentation broth based on a novel hyper-cross-linked meso-micropore resin: Modeling. Bioresour. Technol. 2017, 241, 593–602. [Google Scholar] [CrossRef] [PubMed]
  118. Wang, T.; Sun, S.; Liang, C.; Li, H.; Liu, A.; Zhu, H. Effective isolation of antioxidant Phelligridin LA from the fermentation broth of Inonotus baumii by macroporous resin. Bioprocess. Biosyst. Eng. 2020, 43, 2095–2106. [Google Scholar] [CrossRef] [PubMed]
  119. Ren, J.; Zheng, Y.; Lin, Z.; Han, X.; Liao, W. Macroporous resin purification and characterization of flavonoids from Platycladus orientalis (L.) Franco and their effects on macrophage inflammatory response. Food Funct. 2017, 8, 86–95. [Google Scholar] [CrossRef] [PubMed]
  120. Mulder, H.A.; Halquist, M.S. Growing Trends in the Efficient and Selective Extraction of Compounds in Complex Matrices Using Molecularly Imprinted Polymers and Their Relevance to Toxicological Analysis. J. Anal. Toxicol. 2021, 45, 312–321. [Google Scholar] [CrossRef] [PubMed]
  121. Cao, J.; Shen, C.; Wang, X.; Zhu, Y.; Bao, S.; Wu, X.; Fu, Y. A porous cellulose-based molecular imprinted polymer for specific recognition and enrichment of resveratrol. Carbohydr. Polym. 2021, 251, 117026. [Google Scholar] [CrossRef] [PubMed]
  122. Bereli, N.; Cimen, D.; Huseynli, S.; Denizli, A. Detection of amoxicillin residues in egg extract with a molecularly imprinted polymer on gold microchip using surface plasmon resonance and quartz crystal microbalance methods. J. Food Sci. 2020, 85, 4152–4160. [Google Scholar] [CrossRef] [PubMed]
  123. Chen, M.J.; Yang, H.L.; Si, Y.M.; Tang, Q.; Chow, C.F.; Gong, C.B. Photoresponsive Surface Molecularly Imprinted Polymers for the Detection of Profenofos in Tomato and Mangosteen. Front. Chem. 2020, 8, 583036. [Google Scholar] [CrossRef] [PubMed]
  124. Wu, C.; He, J.; Chen, N.; Li, Y.; Yuan, L.; Zhao, D.; He, L.; Gu, K.; Zhang, S. Synthesis of cobalt-based magnetic nanoporous carbon core-shell molecularly imprinted polymers for the solid-phase extraction of phthalate plasticizers in edible oil. Anal. Bioanal. Chem. 2018, 410, 6943–6954. [Google Scholar] [CrossRef] [PubMed]
  125. Zan, X.; Cui, F.; Li, Y.; Yang, Y.; Wu, D.; Sun, W.; Ping, L. Hericium erinaceus polysaccharide-protein HEG-5 inhibits SGC-7901 cell growth via cell cycle arrest and apoptosis. Int. J. Biol. Macromol. 2015, 76, 242–253. [Google Scholar] [CrossRef] [PubMed]
  126. Chen, P.; Liu, H.P.; Ji, H.H.; Sun, N.X.; Feng, Y.Y. A cold-water soluble polysaccharide isolated from Grifola frondosa induces the apoptosis of HepG2 cells through mitochondrial passway. Int. J. Biol. Macromol. 2019, 125, 1232–1241. [Google Scholar] [CrossRef] [PubMed]
  127. Chen, W.; Zhu, X.; Ma, J.; Zhang, M.; Wu, H. Structural Elucidation of a Novel Pectin-Polysaccharide from the Petal of Saussurea laniceps and the Mechanism of its Anti-HBV Activity. Carbohydr. Polym. 2019, 223, 115077. [Google Scholar] [CrossRef]
  128. Nie, Y.; Lin, Q.; Luo, F. Effects of Non-Starch Polysaccharides on Inflammatory Bowel Disease. Int. J. Mol. Sci. 2017, 18, 1372. [Google Scholar] [CrossRef]
  129. Li, X.L.; Wang, Z.H.; Zhao, Y.X.; Luo, S.J.; Zhang, D.W.; Xiao, S.X.; Peng, Z.H. Isolation and antitumor activities of acidic polysaccharide from Gynostemma pentaphyllum Makino. Carbohydr. Polym. 2012, 89, 942–947. [Google Scholar] [CrossRef]
  130. He, P.; Wu, S.; Zheng, K.; Xu, C. Molecular Structure and Antioxidant Activity of Polysaccharides from Phellinus vaninii. J. Food Sci. Biotechnol. 2015, 37, 1673–1689. [Google Scholar] [CrossRef]
  131. Qi, W.; Zhou, X.; Wang, J.; Zhang, K.; Zhou, Y.; Chen, S.; Nie, S.; Xie, M. Cordyceps sinensis polysaccharide inhibits colon cancer cells growth by inducing apoptosis and autophagy flux blockage via mTOR signaling. Carbohydr. Polym. 2020, 237, 116113. [Google Scholar] [CrossRef] [PubMed]
  132. Jiang, W.P.; Deng, J.S.; Huang, S.S.; Wu, S.H.; Chen, C.C.; Liao, J.C.; Chen, H.Y.; Lin, H.Y.; Huang, G.J. Sanghuangporus sanghuang Mycelium Prevents Paracetamol-Induced Hepatotoxicity through Regulating the MAPK/NF-kappaB, Keap1/Nrf2/HO-1, TLR4/PI3K/Akt, and CaMKKbeta/LKB1/AMPK Pathways and Suppressing Oxidative Stress and Inflammation. Antioxidants 2021, 10, 897. [Google Scholar] [CrossRef] [PubMed]
  133. Huang, C.Y.; Deng, J.S.; Huang, W.C.; Jiang, W.P.; Huang, G.J. Attenuation of Lipopolysaccharide-Induced Acute Lung Injury by Hispolon in Mice, Through Regulating the TLR4/PI3K/Akt/mTOR and Keap1/Nrf2/HO-1 Pathways, and Suppressing Oxidative Stress-Mediated ER Stress-Induced Apoptosis and Autophagy. Nutrients 2020, 12, 1742. [Google Scholar] [CrossRef] [PubMed]
  134. Wang, Y.; Liu, Y.; Zhang, X.Y.; Xu, L.H.; Ouyang, D.Y.; Liu, K.P.; Pan, H.; He, J.; He, X.H. Ginsenoside Rg1 regulates innate immune responses in macrophages through differentially modulating the NF-kappaB and PI3K/Akt/mTOR pathways. Int. Immunopharmacol. 2014, 23, 77–84. [Google Scholar] [CrossRef] [PubMed]
  135. Keppler-Noreuil, K.M.; Parker, V.E.; Darling, T.N.; Martinez-Agosto, J.A. Somatic overgrowth disorders of the PI3K/AKT/mTOR pathway & therapeutic strategies. Am. J. Med. Genet. C Semin. Med. Genet. 2016, 172, 402–421. [Google Scholar] [CrossRef]
  136. Thimmulappa, R.K.; Lee, H.; Rangasamy, T.; Reddy, S.P.; Yamamoto, M.; Kensler, T.W.; Biswal, S. Nrf2 is a critical regulator of the innate immune response and survival during experimental sepsis. J. Clin. Investig. 2006, 116, 984–995. [Google Scholar] [CrossRef]
  137. Song, J.; Zhang, W.; Wang, J.; Yang, H.; Zhao, X.; Zhou, Q.; Wang, H.; Li, L.; Du, G. Activation of Nrf2 signaling by salvianolic acid C attenuates NF-kappaB mediated inflammatory response both in vivo and in vitro. Int. Immunopharmacol. 2018, 63, 299–310. [Google Scholar] [CrossRef]
  138. Wang, Z.; Hao, W.; Hu, J.; Mi, X.; Han, Y.; Ren, S.; Jiang, S.; Wang, Y.; Li, X.; Li, W. Maltol Improves APAP-Induced Hepatotoxicity by Inhibiting Oxidative Stress and Inflammation Response via NF-kappaB and PI3K/Akt Signal Pathways. Antioxidants 2019, 8, 395. [Google Scholar] [CrossRef]
  139. Tsuruma, R.; Ohbayashi, N.; Kamitani, S.; Ikeda, O.; Sato, N.; Muromoto, R.; Sekine, Y.; Oritani, K.; Matsuda, T. Physical and functional interactions between STAT3 and KAP1. Oncogene 2008, 27, 3054–3059. [Google Scholar] [CrossRef]
  140. Yang, N.; Li, C.; Tian, G.; Zhu, M.; Bu, W.; Chen, J.; Hou, X.; Di, L.; Jia, X.; Dong, Z.; et al. Organic acid component from Taraxacum mongolicum Hand.-Mazz alleviates inflammatory injury in lipopolysaccharide-induced acute tracheobronchitis of ICR mice through TLR4/NF-kappaB signaling pathway. Int. Immunopharmacol. 2016, 34, 92–100. [Google Scholar] [CrossRef]
  141. Kim, J.H.; Choo, Y.Y.; Tae, N.; Min, B.S.; Lee, J.H. The anti-inflammatory effect of 3-deoxysappanchalcone is mediated by inducing heme oxygenase-1 via activating the AKT/mTOR pathway in murine macrophages. Int. Immunopharmacol. 2014, 22, 420–426. [Google Scholar] [CrossRef]
  142. Ma, C.; Zhu, L.; Wang, J.; He, H.; Chang, X.; Gao, J.; Shumin, W.; Yan, T. Anti-inflammatory effects of water extract of Taraxacum mongolicum hand.-Mazz on lipopolysaccharide-induced inflammation in acute lung injury by suppressing PI3K/Akt/mTOR signaling pathway. J. Ethnopharmacol. 2015, 168, 349–355. [Google Scholar] [CrossRef]
  143. Wang, L.; Gui, Y.S.; Tian, X.L.; Cai, B.Q.; Wang, D.T.; Zhang, D.; Zhao, H.; Xu, K.F. Inactivation of mammalian target of rapamycin (mTOR) by rapamycin in a murine model of lipopolysaccharide-induced acute lung injury. Chin. Med. J. 2011, 124, 3112–3117. [Google Scholar]
  144. Huang, G.J.; Deng, J.S.; Chiu, C.S.; Liao, J.C.; Hsieh, W.T.; Sheu, M.J.; Wu, C.H. Hispolon Protects against Acute Liver Damage in the Rat by Inhibiting Lipid Peroxidation, Proinflammatory Cytokine, and Oxidative Stress and Downregulating the Expressions of iNOS, COX-2, and MMP-9. Evid. Based Complement. Altern. Med. 2012, 2012, 480714. [Google Scholar] [CrossRef]
  145. Huang, G.J.; Deng, J.S.; Huang, S.S.; Hu, M.L. Hispolon induces apoptosis and cell cycle arrest of human hepatocellular carcinoma Hep3B cells by modulating ERK phosphorylation. J. Agric. Food Chem. 2011, 59, 7104–7113. [Google Scholar] [CrossRef] [PubMed]
  146. Huang, G.-J.; Yang, C.-M.; Chang, Y.-S.; Amagaya, S.; Wang, H.-C.; Hou, W.-C.; Huang, S.-S.; Hu, M.-L. Hispolon Suppresses SK-Hep1 Human Hepatoma Cell Metastasis by Inhibiting Matrix Metalloproteinase-2/9 and Urokinase-Plasminogen Activator through the PI3K/Akt and ERK Signaling Pathways. J. Agric. Food Chem. 2010, 58, 9468–9475. [Google Scholar] [CrossRef] [PubMed]
  147. Hsieh, Y.H.; Deng, J.S.; Chang, Y.S.; Huang, G.J. Ginsenoside Rh2 Ameliorates Lipopolysaccharide-Induced Acute Lung Injury by Regulating the TLR4/PI3K/Akt/mTOR, Raf-1/MEK/ERK, and Keap1/Nrf2/HO-1 Signaling Pathways in Mice. Nutrients 2018, 10, 1208. [Google Scholar] [CrossRef] [PubMed]
  148. Mu, C.; Zhou, G.; Qin, H.; Chen, J.; Li, C. Effect Of Phellinus Polysaccharide On Canonical Wnt Signaling In Synovium In Rats With Rheumatoid Arthritis. J. Sichuan Univ. (Med. Sci. Ed.) 2015, 46, 376–379. [Google Scholar] [CrossRef]
  149. Iyengar, S.; Farnham, P.J. KAP1 protein: An enigmatic master regulator of the genome. J. Biol. Chem. 2011, 286, 26267–26276. [Google Scholar] [CrossRef] [PubMed]
  150. Kamitani, S.; Togi, S.; Ikeda, O.; Nakasuji, M.; Sakauchi, A.; Sekine, Y.; Muromoto, R.; Oritani, K.; Matsuda, T. Kruppel-associated box-associated protein 1 negatively regulates TNF-alpha-induced NF-kappaB transcriptional activity by influencing the interactions among STAT3, p300, and NF-kappaB/p65. J. Immunol. 2011, 187, 2476–2483. [Google Scholar] [CrossRef] [PubMed]
  151. Gegotek, A.; Niklinski, J.; Charkiewicz, R.; Bielawska, K.; Kozlowski, M.; Skrzydlewska, E. Relationships between level of lipid peroxidation products and expression of Nrf2 and its activators/inhibitors in non-small cell lung cancer tissue. Free Radic. Biol. Med. 2014, 75 (Suppl. 1), S31. [Google Scholar] [CrossRef]
  152. Maruyama, A.; Nishikawa, K.; Kawatani, Y.; Mimura, J.; Hosoya, T.; Harada, N.; Yamamato, M.; Itoh, K. The novel Nrf2-interacting factor KAP1 regulates susceptibility to oxidative stress by promoting the Nrf2-mediated cytoprotective response. Biochem. J. 2011, 436, 387–397. [Google Scholar] [CrossRef]
  153. Shao, H.J.; Jeong, J.B.; Kim, K.J.; Lee, S.H. Anti-inflammatory activity of mushroom-derived hispidin through blocking of NF-kappaB activation. J. Sci. Food Agric. 2015, 95, 2482–2486. [Google Scholar] [CrossRef]
  154. Li, J.; Yuan, J.; Zhang, Y.; Shui, B.; Fan, Y.; Ding, X. Therapeutic Effects of Phellinus igniarius Mycelium Polysaccharide on Collagen-induced Arthritis Rats. J. Zhejiang Chin. Med. Univ. 2014, 38, 526–530. [Google Scholar] [CrossRef]
  155. Noh, J.R.; Kim, Y.H.; Hwang, J.H.; Gang, G.T.; Kim, K.S.; Lee, I.K.; Yun, B.S.; Lee, C.H. Davallialactone protects against acetaminophen overdose-induced liver injuries in mice. Food Chem. Toxicol. 2013, 58, 14–21. [Google Scholar] [CrossRef] [PubMed]
  156. Hsieh, Y.H.; Deng, J.S.; Pan, H.P.; Liao, J.C.; Huang, S.S.; Huang, G.J. Sclareol ameliorate lipopolysaccharide-induced acute lung injury through inhibition of MAPK and induction of HO-1 signaling. Int. Immunopharmacol. 2017, 44, 16–25. [Google Scholar] [CrossRef] [PubMed]
  157. Fitzgerald, K.A.; Kagan, J.C. Toll-like Receptors and the Control of Immunity. Cell 2020, 180, 1044–1066. [Google Scholar] [CrossRef]
  158. Katsanos, K.H.; Papadakis, K.A. Inflammatory Bowel Disease: Updates on Molecular Targets for Biologics. Gut Liver 2017, 11, 455–463. [Google Scholar] [CrossRef] [PubMed]
  159. Park, J.S.; Choi, J.; Hwang, S.H.; Kim, J.K.; Kim, E.K.; Lee, S.Y.; Lee, B.I.; Park, S.H.; Cho, M.L. Cottonseed Oil Protects Against Intestinal Inflammation in Dextran Sodium Sulfate-Induced Inflammatory Bowel Disease. J. Med. Food 2019, 22, 672–679. [Google Scholar] [CrossRef] [PubMed]
  160. Lee, J.-B.; Shin, Y.-O.; Bae, J.-S.; Min, Y.-K.; Yang, H.-M.; Seo, H.-S.; Kwon, D.-K.; Kang, J.-Y.; Song, Y.-J. Effect of Phellinus linteus extract supplementation on cortisol and related cytokines in young male adults. Food Sci. Biotechnol. 2010, 19, 671–675. [Google Scholar] [CrossRef]
  161. Zhang, Q.; Su, G.; Zhao, T.; Wang, S.; Sun, B.; Zheng, L.; Zhao, M. The memory improving effects of round scad (Decapterus maruadsi) hydrolysates on sleep deprivation-induced memory deficits in rats via antioxidant and neurotrophic pathways. Food Funct. 2019, 10, 7733–7744. [Google Scholar] [CrossRef] [PubMed]
  162. Im, K.H.; Baek, S.A.; Choi, J.; Lee, T.S. Antioxidant, Anti-Melanogenic and Anti-Wrinkle Effects of Phellinus vaninii. Mycobiology 2019, 47, 494–505. [Google Scholar] [CrossRef] [PubMed]
  163. Bailly, C.; Corbineau, F.; van Doorn, W.G. Free radical scavenging and senescence in Iris tepals. Plant Physiol. Biochem. 2001, 39, 649–656. [Google Scholar] [CrossRef]
  164. Valko, M.; Leibfritz, D.; Moncol, J.; Cronin, M.T.D.; Mazur, M.; Telser, J. Free radicals and antioxidants in normal physiological functions and human disease. Int. J. Biochem. Cell Biol. 2007, 39, 44–84. [Google Scholar] [CrossRef] [PubMed]
  165. Aseervatham, G.S.B.; Sivasudha, T.; Jeyadevi, R.; Arul Ananth, D. Environmental factors and unhealthy lifestyle influence oxidative stress in humans—An overview. Environ. Sci. Pollut. Res. 2013, 20, 4356–4369. [Google Scholar] [CrossRef] [PubMed]
  166. Lyu, G.; Wang, X.; Song, T.; Zhang, Z. Comparison of Antioxidant Activity of Liquid Fermentation Products of Two Wild Phellinus Strain. Edible Med. Mushrooms 2018, 26, 175–177+183. [Google Scholar]
  167. Ververidis, F.; Trantas, E.; Douglas, C.; Vollmer, G.; Kretzschmar, G.; Panopoulos, N. Biotechnology of flavonoids and other phenylpropanoid-derived natural products. Part I: Chemical diversity, impacts on plant biology and human health. Biotechnol. J. 2007, 2, 1214–1234. [Google Scholar] [CrossRef] [PubMed]
  168. Zhu, T.; Kim, S.-H.; Chen, C.-Y. A Medicinal Mushroom: Phellinus linteus. Curr. Med. Chem. 2008, 15, 1330–1335. [Google Scholar] [CrossRef]
  169. Mustafa, R.A.; Hamid, A.A.; Mohamed, S.; Bakar, F.A. Total Phenolic Compounds, Flavonoids, and Radical Scavenging Activity of 21 Selected Tropical Plants. J. Food Sci. 2010, 75, C28–C35. [Google Scholar] [CrossRef]
  170. Wu, X.; Wang, C.; Bao, H. Antitumor Effect of Water Extracts from Fruiting Bodies of Different “Sanghuang” Fungi. Biotechnol. Bull. 2018, 34, 138–143. [Google Scholar] [CrossRef]
  171. Wang, W.; Lu, N.; Yan, J.; Fu, L.; Song, J.; Yuan, W.; Zhou, Z. Effects of growth age on nutrition, content of active components and antioxidant activities of fruiting bodies of Sanghuangporus baumii cultivated on oak segment. Mycosystema 2021, 40, 668–680. [Google Scholar] [CrossRef]
  172. Liu, F.; Pang, D.; Zou, Y.; Liao, S.; Xiao, G. The Total Flavone Content and in Vitro Antioxidant Activity of Phellinus sp. Edible Fungi China 2014, 33, 47–49+56. [Google Scholar] [CrossRef]
  173. Wang, H.; Ma, J.X.; Wu, D.M.; Gao, N.; Si, J.; Cui, B.K. Identifying Bioactive Ingredients and Antioxidant Activities of Wild Sanghuangporus Species of Medicinal Fungi. J. Fungi 2023, 9, 242. [Google Scholar] [CrossRef] [PubMed]
  174. Shao, Y.; Guo, H.; Zhang, J.; Liu, H.; Wang, K.; Zuo, S.; Xu, P.; Xia, Z.; Zhou, Q.; Zhang, H.; et al. The Genome of the Medicinal Macrofungus Sanghuang Provides Insights Into the Synthesis of Diverse Secondary Metabolites. Front. Microbiol. 2019, 10, 3035. [Google Scholar] [CrossRef]
  175. Chen, W.; Zhang, J.; Wu, D.; Zhang, Z.; Li, W.; Wang, W.; Wang, K.; Yang, Y. HPLC fingerprint and spectrum effect relationship of antioxidant activities of Sanghuangporus baumii. Mycosystema 2021, 40, 2364–2375. [Google Scholar] [CrossRef]
  176. Mapari, S.A.S.; Nielsen, K.F.; Larsen, T.O.; Frisvad, J.C.; Meyer, A.S.; Thrane, U. Exploring fungal biodiversity for the production of water-soluble pigments as potential natural food colorants. Curr. Opin. Biotechnol. 2005, 16, 231–238. [Google Scholar] [CrossRef] [PubMed]
  177. Fox, E.M.; Howlett, B.J. Secondary metabolism: Regulation and role in fungal biology. Curr. Opin. Microbiol. 2008, 11, 481–487. [Google Scholar] [CrossRef]
  178. Nirlane da Costa Souza, P.; Luiza Bim Grigoletto, T.; Alberto Beraldo de Moraes, L.; Abreu, L.M.; Henrique Souza Guimarães, L.; Santos, C.; Ribeiro Galvão, L.; Gomes Cardoso, P. Production and chemical characterization of pigments in filamentous fungi. Microbiology 2016, 162, 12–22. [Google Scholar] [CrossRef]
  179. Teixeira, M.F.S.; Martins, M.S.; da Silva, J.C.; Kirsch, L.S.; Fernandes, O.C.C.; Carneiro, A.L.B.; de Conti, R.; Durán, N. Amazonian biodiversity: Pigments from Aspergillus and Penicillium-characterizations, antibacterial activities and their Toxicities. Curr. Trends Biotechnol. Pharm. 2012, 6, 300–311. [Google Scholar]
  180. Luo, J.; Liu, J.; Sun, Y.; Ye, H.; Zhou, C.; Zeng, X. Medium optimization, preliminary characterization and antioxidant activity in vivo of mycelial polysaccharide from Phellinus baumii Pilát. Carbohydr. Polym. 2010, 81, 533–540. [Google Scholar] [CrossRef]
  181. Gu, L.; Cai, W.; Jin, Q.; Zhang, Z.; Wang, J.; Zhang, X. Effect of Sanghuang Decoction on Oxidative Damage Induced by D-galactose in Mice. Chin. J. Mod. Appl. Pharm. 2019, 36, 2144–2148. [Google Scholar] [CrossRef]
  182. Zhang, W.; Wang, G.; Zhou, S. Protective Effects of Isoliquiritigenin on LPS-Induced Acute Lung Injury by Activating PPAR-γ. Inflammation 2018, 41, 1290–1296. [Google Scholar] [CrossRef]
  183. Li, X.; He, F.; Ma, Y.; Liu, G.; Chu, L.; Zheng, M. The Growth Characteristics and Antioxidant Activity of Different Sanghuang Strain in Fermentation Broth. Edible Fungi China 2021, 40, 57–61+66. [Google Scholar] [CrossRef]
  184. Fu, L.; Lu, N.; Yan, J.; Wang, W.; Song, J.; Yuan, W.; Zhou, Z. Analyses and evaluation of nutrition, active component and antioxidant activities of fruiting bodies of three species of Sanghuangporus. Mycosystema 2021, 40, 2148–2158. [Google Scholar] [CrossRef]
  185. Wang, R.; Xu, Z.; Li, J.; Wang, J.; Bian, Y.; Chen, Z. Selection of a High Antioxidant Capacity Sanghuangporus sanghuang Strain and Elicitation of It with Phytohormones for Increased Intracellular Flavonoids. J. Edible Mushrooms 2020, 27, 61–67. [Google Scholar] [CrossRef]
  186. Li, Y.; Yin, C.; Fan, X.; Shi, D.; Yao, F.; Li, C.; Liu, Y.; Cheng, Y.; Gao, H. In Vitro Anti-oxidant, Hypoglycemic, and Hypouricemic Activities of Sanghuangporus vaninii Extracts. Mod. Food Sci. Technol. 2022, 38, 71–80. [Google Scholar] [CrossRef]
  187. Lyu, G.; Song, T.; Zhang, Y.; Cai, W.; Zhang, Z. Comparative study on antioxidant activities of fermentation products of Sanghuangporus sanghuang and S. vaninii based on UPLC-triple-TOF-MS. Mycosystema 2023, 42, 939–948. [Google Scholar]
  188. Yang, L.; Yang, H.; Que, S.; Ban, L. Comparative Analysis of the Antioxidant Activity of the Different Components from Phellinus igniarius. Food Res. Dev. 2016, 37, 208–210. [Google Scholar]
  189. Gong, R.; Tian, X.; Ou, X.; Li, S.; Ma, L.; He, X.; Miao, T. Effects of Phellinus linteus Ploysaccharide on Growth Performance, Hematological Parameters and Immunological Indexes of Growing-laying Hens. Chin. J. Vet. Med. 2021, 57, 86–90. [Google Scholar]
  190. Wu, J.; Wang, Z.; Yuan, X.; Shan, Q.; Wang, X. Mechanism of inhibitory effect of Phelinus linteus polysaccharide on glioma cell proliferation and migration through PI3K/Akt signaling pathway. Chin. J. Cancer Prev. Treat. 2020, 27, 840–847. [Google Scholar] [CrossRef]
  191. Lang, M.; Zeng, P.; Liu, M.; Xie, H.; Li, X.; Shi, L. Research Progress on Polysaccharide from Phellinus baumii. Bull. Seric. 2017, 48, 14–18. [Google Scholar]
  192. Fotsis, T.; Pepper, M.S.; Aktas, E.; Breit, S.; Rasku, S.; Adlercreutz, H.; Wähälä, K.; Montesano, R.; Schweigerer, L. Flavonoids, dietary-derived inhibitors of cell proliferation and in vitro angiogenesis. Cancer Res. 1997, 57, 2916–2921. [Google Scholar] [PubMed]
  193. Chen, J.; Sun, J.; Lyu, F.; Gao, G.; Zhao, M.; Zhang, K. Isolation and bioactive characters of a novel lectin SHL24 from the liquid fermentation of Phellinus baumii. Microbiol. China 2018, 45, 420–427. [Google Scholar] [CrossRef]
  194. He, P.Y.; Yang, Y.; Di, L.; Li, J.L.; Li, N. A comparative study on in vitro antitumor activities of the medicinal fungus Sanghuangporus baumii cultivated in different substrates. Mycosystema 2020, 39, 1400–1409. [Google Scholar] [CrossRef]
  195. Liu, X.; Liu, F.; Zhao, S.; Guo, B.; Ling, P.; Han, G.; Cui, Z. Purification of an acidic polysaccharide from Suaeda salsa plant and its anti-tumor activity by activating mitochondrial pathway in MCF-7 cells. Carbohydr. Polym. 2019, 215, 99–107. [Google Scholar] [CrossRef]
  196. Meng, X.; Liang, H.; Luo, L. Antitumor polysaccharides from mushrooms: A review on the structural characteristics, antitumor mechanisms and immunomodulating activities. Carbohydr. Res. 2016, 424, 30–41. [Google Scholar] [CrossRef] [PubMed]
  197. Zhao, H.; Zhou, J.; Qi, Y. The effects of Phellinus igniarius Polysaccharides on regulating PI3K/AKT/mTOR Pathway to Inhibit Abdominal Tumor Bearing Tumor of Liver Cancer and to Reduce the Toxicity and Increase the Efficacy of Chemotherapy. Chin. J. Clin. Pharmacol. Ther. 2020, 25, 401–407. [Google Scholar]
  198. Wei, J.; Li, J.; Wen, C.; Fan, Y.; Ding, X. Therapeutic Effect of Polysaccharides from Phellinus igniarius on Liver Cancer in H22 Bearing Mice. J. Chin. Med. Mater. 2016, 39, 2868–2870. [Google Scholar] [CrossRef]
  199. Lyu, F.; Zhang, N.; Yu, S. Research progress on the anti-tumor mechanism of Sanghuang polysaccharides. Shanghai J. Tradit. Chin. Med. 2015, 49, 87–89. [Google Scholar] [CrossRef]
  200. Zhao, J.; Wang, X.; Yu, Z. Influence of Phellinus igniarius Polysaccharides on Growth and Peripheral Blood Cells of Rats Treated by Cyclophosphamide. Chin. J. Mod. Appl. Pharm. 2017, 34, 832–835. [Google Scholar] [CrossRef]
  201. Zhong, S.; Li, Y.G.; Ji, D.F.; Lin, T.B.; Lv, Z.Q. Protocatechualdehyde Induces S-Phase Arrest and Apoptosis by Stimulating the p27(KIP1)-Cyclin A/D1-CDK2 and Mitochondrial Apoptotic Pathways in HT-29 Cells. Molecules 2016, 21, 934. [Google Scholar] [CrossRef] [PubMed]
  202. Zhang, M.; Chen, H.; Huang, J.; Li, Z.; Zhu, C.; Zhang, S. Effect of lycium barbarum polysaccharide on human hepatoma QGY7703 cells: Inhibition of proliferation and induction of apoptosis. Life Sci. 2005, 76, 2115–2124. [Google Scholar] [CrossRef] [PubMed]
  203. Wang, L.; Yao, L.; Jin, Y.; Jin, C.; Dong, Y.; Shou, D.; Wang, Y. Activation Effect of Human TLR4 Signaling Pathway by Polysaccharide from Phellinus igniarius. Chin. J. Mod. Appl. Pharm. 2019, 36, 1178–1182. [Google Scholar] [CrossRef]
  204. Huang, L.; An, C.; Liu, D.; Pan, Y. Effects of liquid fermented flavonoids of Phellinus igniarus on antitumor activity and cell cycle. J. Shandong Univ. Technol. (Nat. Sci. Ed.) 2019, 33, 65–68+73. [Google Scholar] [CrossRef]
  205. Al Bitar, S.; Gali-Muhtasib, H. The Role of the Cyclin Dependent Kinase Inhibitor p21(cip1/waf1) in Targeting Cancer: Molecular Mechanisms and Novel Therapeutics. Cancers 2019, 11, 1475. [Google Scholar] [CrossRef] [PubMed]
  206. Moussa, R.S.; Park, K.C.; Kovacevic, Z.; Richardson, D.R. Ironing out the role of the cyclin-dependent kinase inhibitor, p21 in cancer: Novel iron chelating agents to target p21 expression and activity. Free Radic. Biol. Med. 2019, 133, 276–294. [Google Scholar] [CrossRef] [PubMed]
  207. Hennessy, E.J. Selective inhibitors of Bcl-2 and Bcl-xL: Balancing antitumor activity with on-target toxicity. Bioorg. Med. Chem. Lett. 2016, 26, 2105–2114. [Google Scholar] [CrossRef]
  208. Liao, K.F.; Chiu, T.L.; Chang, S.F.; Wang, M.J.; Chiu, S.C. Hispolon Induces Apoptosis, Suppresses Migration and Invasion of Glioblastoma Cells and Inhibits GBM Xenograft Tumor Growth In Vivo. Molecules 2021, 26, 4497. [Google Scholar] [CrossRef]
  209. Chen, D.; Li, D.; Xu, X.-B.; Qiu, S.; Luo, S.; Qiu, E.; Rong, Z.; Zhang, J.; Zheng, D. Galangin inhibits epithelial-mesenchymal transition and angiogenesis by downregulating CD44 in glioma. J. Cancer 2019, 10, 4499–4508. [Google Scholar] [CrossRef]
  210. Jung, G.H.; Kang, J.H. Efficacy of Phellinus linteus (sanghuang) extract for improving immune functions: Study protocol for a randomized, double-blinded, placebo-controlled pilot trial. Medicine 2020, 99, e18829. [Google Scholar] [CrossRef]
  211. Li, D.; Liu, H.; Su, B.; Liu, Z. Study on the Effect of Polysaccharides from Phellinus baumii on Tumor Microenvironment. J. Taishan Med. Coll. 2020, 41, 318–320. [Google Scholar]
  212. Ying, R.; Wu, C.; Huang, M.; Wang, Y. Anti-tumor activity of polysaccharides from Phellinus igniarius fruiting body and mycelium. China Food Addit. 2017, 38, 57–61. [Google Scholar]
  213. Gao, W.; Zhang, N.; Yu, S. Research progress in anti-tumor effects and mechanism of Phellinus fungi. China J. Chin. Mater. Medica 2014, 39, 4165–4168. [Google Scholar]
  214. Li, T.; Yang, Y.; Liu, Y.; Bao, D. Research progress in structural characteristics, preparation methods, and biological activities of polysaccharides from Phellinus fungi. J. Edible Mushrooms 2013, 20, 65–70. [Google Scholar] [CrossRef]
  215. Ma, C.; Zhang, Z.; Sun, S.; Zhong, Y. Research Progress on Fermentation Culture, Chemical Compnents and Bioactivity of Phellinus linteus. Edible Fungi China 2021, 40, 14–18. [Google Scholar] [CrossRef]
  216. Lim, B.O.; Yamada, K.; Cho, B.-G.; Jeon, T.; Hwang, S.-G.; Park, T.; Kang, S.A.; Park, D.K. Comparative Study on the Modulation of IgE and Cytokine Production by Phellinus linteus Grown on Germinated Brown Rice, Phellinus linteus and Germinated Brown Rice in Murine Splenocytes. Biosci. Biotechnol. Biochem. 2004, 68, 2391–2394. [Google Scholar] [CrossRef] [PubMed]
  217. Zhang, J.N.; Qu, H.Y.; Zhang, J.M.; Feng, J.M.; Song, W.J.; Yuan, F.H. Polysaccharide from Phellinus igniarius alleviates oxidative stress and hepatic fibrosis in Schistosoma japonicum-infected mice. Chin. J. Schisto Control 2019, 31, 615–621. [Google Scholar] [CrossRef]
  218. Chien, L.H.; Deng, J.S.; Jiang, W.P.; Chou, Y.N.; Lin, J.G.; Huang, G.J. Evaluation of lung protection of Sanghuangporus sanghuang through TLR4/NF-kappaB/MAPK, keap1/Nrf2/HO-1, CaMKK/AMPK/Sirt1, and TGF-beta/SMAD3 signaling pathways mediating apoptosis and autophagy. Biomed. Pharmacother. 2023, 165, 115080. [Google Scholar] [CrossRef] [PubMed]
  219. Cao, X.; La, X.; Zhang, B.; Wang, Z.; Li, Y.; Bo, Y.; Chang, H.; Gao, X.; Tian, C.; Wu, C.; et al. Sanghuang Tongxie Formula Ameliorates Insulin Resistance in Drosophila Through Regulating PI3K/Akt Signaling. Front. Pharmacol. 2022, 13, 874180. [Google Scholar] [CrossRef] [PubMed]
  220. Magkos, F.; Hjorth, M.F.; Astrup, A. Diet and exercise in the prevention and treatment of type 2 diabetes mellitus. Nat. Rev. Endocrinol. 2020, 16, 545–555. [Google Scholar] [CrossRef] [PubMed]
  221. Zhu, Y.; Sidell, M.A.; Arterburn, D.; Daley, M.F.; Desai, J.; Fitzpatrick, S.L.; Horberg, M.A.; Koebnick, C.; McCormick, E.; Oshiro, C.; et al. Racial/Ethnic Disparities in the Prevalence of Diabetes and Prediabetes by BMI: Patient Outcomes Research To Advance Learning (PORTAL) Multisite Cohort of Adults in the U.S. Diabetes Care 2019, 42, 2211–2219. [Google Scholar] [CrossRef]
  222. Mirmiran, P.; Bahadoran, Z.; Azizi, F. Functional foods-based diet as a novel dietary approach for management of type 2 diabetes and its complications: A review. World J. Diabetes 2014, 5, 267–281. [Google Scholar] [CrossRef]
  223. Ballali, S.; Lanciai, F. Functional food and diabetes: A natural way in diabetes prevention? Int. J. Food Sci. Nutr. 2012, 63 (Suppl. 1), 51–61. [Google Scholar] [CrossRef]
  224. Khursheed, R.; Singh, S.K.; Wadhwa, S.; Gulati, M.; Awasthi, A. Therapeutic potential of mushrooms in diabetes mellitus: Role of polysaccharides. Int. J. Biol. Macromol. 2020, 164, 1194–1205. [Google Scholar] [CrossRef]
  225. Shi, D.; Cui, Q.; Qi, X.; Cui, C.; Chen, Z. Effect of Phellinus igniarius polysaccharide on type I diabetes mice model. Agric. Sci. J. Yanbian Univ. 2017, 39, 33–38. [Google Scholar] [CrossRef]
  226. Huang, Q.; Lin, P.; Wang, M.; Chen, H.; Zheng, D.; Huang, Q.; Zhang, S.; Shi, Z. Phellinus igniarius polysaccharide attenuates renal fibrosis in diabetic nephropathy mice based on P311/TGF- β1/Snail1 Pathway. Pharmacol. Clin. Tradit. Chin. Med. 2019, 35, 28–33. [Google Scholar] [CrossRef]
  227. Littlewood, D.L.; Russell, K. Is there a role for sleep medicine in suicide prevention? Sleep Med. 2020, 66, 262–263. [Google Scholar] [CrossRef] [PubMed]
  228. Godos, J.; Ferri, R.; Caraci, F.; Cosentino, F.I.I.; Castellano, S.; Shivappa, N.; Hebert, J.R.; Galvano, F.; Grosso, G. Dietary Inflammatory Index and Sleep Quality in Southern Italian Adults. Nutrients 2019, 11, 1324. [Google Scholar] [CrossRef] [PubMed]
  229. Wible, R.S.; Ramanathan, C.; Sutter, C.H.; Olesen, K.M.; Kensler, T.W.; Liu, A.C.; Sutter, T.R. NRF2 regulates core and stabilizing circadian clock loops, coupling redox and timekeeping in Mus musculus. Elife 2018, 7, e31656. [Google Scholar] [CrossRef] [PubMed]
  230. Frigato, E.; Benedusi, M.; Guiotto, A.; Bertolucci, C.; Valacchi, G. Circadian Clock and OxInflammation: Functional Crosstalk in Cutaneous Homeostasis. Oxid. Med. Cell Longev. 2020, 2020, 2309437. [Google Scholar] [CrossRef]
  231. Liu, Z.-Q.; Sun, X.; Zhang, T.; Zhang, L.-L.; Wu, C.-J. Phytochemicals in traditional Chinese medicine can treat gout by regulating intestinal flora through inactivating NLRP3 and inhibiting XOD activity. J. Pharm. Pharmacol. 2022, 74, 919–929. [Google Scholar] [CrossRef]
  232. Chi, X.; Zhang, H.; Zhang, S.; Ma, K.A.-O. Chinese herbal medicine for gout: A review of the clinical evidence and pharmacological mechanisms. Chin. Med. 2020, 15, 17. [Google Scholar] [CrossRef]
  233. Mattson, M.P.; Arumugam, T.V. Hallmarks of Brain Aging: Adaptive and Pathological Modification by Metabolic States. Cell Metab. 2018, 27, 1176–1199. [Google Scholar] [CrossRef] [PubMed]
  234. Soukas, A.A.; Hao, H.; Wu, L. Metformin as Anti-Aging Therapy: Is It for Everyone? Trends Endocrinol. Metab. 2019, 30, 745–755. [Google Scholar] [CrossRef]
  235. Okoro, N.O.; Odiba, A.S.; Osadebe, P.O.; Omeje, E.O.; Liao, G.; Fang, W.A.-O.; Jin, C.A.-O.; Wang, B.A.-O. Bioactive Phytochemicals with Anti-Aging and Lifespan Extending Potentials in Caenorhabditis elegans. Molecules 2021, 26, 7323. [Google Scholar] [CrossRef]
  236. Wang, W.; Zhang, J.; Yang, Y.; Yan, M.; Wu, D.; Wu, N.; Zhang, H. A Comparative Study on Antioxidant and Anti-aging Activities of Ten Edible/Medicinal Fungi. Nat. Prod. Res. Dev. 2013, 25, 1027–1032. [Google Scholar] [CrossRef]
  237. Papaemmanouil, C.D.; Peña-García, J.; Banegas-Luna, A.A.-O.; Kostagianni, A.D.; Gerothanassis, I.A.-O.; Pérez-Sánchez, H.A.-O.; Tzakos, A.A.-O. ANTIAGE-DB: A Database and Server for the Prediction of Anti-Aging Compounds Targeting Elastase, Hyaluronidase, and Tyrosinase. Antioxidants 2022, 11, 2268. [Google Scholar] [CrossRef] [PubMed]
  238. Ooyama, T.; Sakamato, H. Elastase in the prevention of arterial aging and the treatment of atherosclerosis. In Ciba Foundation Symposium 192-The Molecular Biology and Pathology of Elastic Tissues: The Molecular Biology and Pathology of Elastic Tissues: Ciba Foundation Symposium; John Wiley & Sons, Ltd.: Chichester, UK, 2007. [Google Scholar]
  239. Dong, Z.; Wang, Y.; Hao, C.; Cheng, Y.; Guo, X.; He, Y.; Shi, Y.; Wang, S.; Li, Y.; Shi, W. Sanghuangporus sanghuang extract extended the lifespan and healthspan of Caenorhabditis elegans via DAF-16/SIR-2.1. Front. Pharmacol. 2023, 14, 1136897. [Google Scholar] [CrossRef]
  240. Liu, K.; Qi, C.; Liu, Y.; Huai, Y.; Hu, H.; Xiao, X.; Wang, J. Phenolic composition and neuroprotective effects of the ethyl-acetate fraction from Inonotus sanghuang against H2O2-induced apoptotic cell death of primary cortical neuronal cells. Food Sci. Biotechnol. 2022, 31, 1213–1223. [Google Scholar] [CrossRef] [PubMed]
  241. Jayasena, T.; Poljak, A.; Smythe, G.; Braidy, N.; Münch, G.; Sachdev, P. The role of polyphenols in the modulation of sirtuins and other pathways involved in Alzheimer’s disease. Ageing Res. Rev. 2013, 12, 867–883. [Google Scholar] [CrossRef]
  242. Phan, C.-W.; David, P.; Naidu, M.; Wong, K.-H.; Sabaratnam, V. Therapeutic potential of culinary-medicinal mushrooms for the management of neurodegenerative diseases: Diversity, metabolite, and mechanism. Crit. Rev. Biotechnol. 2015, 35, 355–368. [Google Scholar] [CrossRef]
  243. Choi, D.J.; Cho, S.; Seo, J.Y.; Lee, H.B.; Park, Y.I. Neuroprotective effects of the Phellinus linteus ethyl acetate extract against H2O2-induced apoptotic cell death of SK-N-MC cells. Nutr. Res. 2016, 36, 31–43. [Google Scholar] [CrossRef]
  244. 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, 769–778. [Google Scholar] [CrossRef]
  245. Kitamura, Y.; Sakanashi, M.; Ozawa, A.; Saeki, Y.; Nakamura, A.; Hara, Y.; Saeki, K.I.; Arimoto-Kobayashi, S. Protective effect of Actinidia arguta in MPTP-induced Parkinson’s disease model mice. Biochem. Biophys. Res. Commun. 2021, 555, 154–159. [Google Scholar] [CrossRef]
  246. Rezaei, O.; Nateghinia, S.; Estiar, M.A.; Taheri, M.; Ghafouri-Fard, S. Assessment of the role of non-coding RNAs in the pathophysiology of Parkinson’s disease. Eur. J. Pharmacol. 2021, 896, 173914. [Google Scholar] [CrossRef]
  247. Rai, S.N.; Singh, P.; Varshney, R.; Chaturvedi, V.K.; Vamanu, E.; Singh, M.P.; Singh, B.K. Promising drug targets and associated therapeutic interventions in Parkinson’s disease. Neural Regen. Res. 2021, 16, 1730–1739. [Google Scholar] [CrossRef]
  248. Rai, S.N.; Chaturvedi, V.K.; Singh, P.; Singh, B.K.; Singh, M.P. Mucuna pruriens in Parkinson’s and in some other diseases: Recent advancement and future prospective. 3 Biotech 2020, 10, 522. [Google Scholar] [CrossRef] [PubMed]
  249. Hemmati-Dinarvand, M.; Saedi, S.; Valilo, M.; Kalantary-Charvadeh, A.; Alizadeh Sani, M.; Kargar, R.; Safari, H.; Samadi, N. Oxidative stress and Parkinson’s disease: Conflict of oxidant-antioxidant systems. Neurosci. Lett. 2019, 709, 134296. [Google Scholar] [CrossRef] [PubMed]
  250. Fiolet, T.; Kherabi, Y.; MacDonald, C.J.; Ghosn, J.; Peiffer-Smadja, N. Comparing COVID-19 vaccines for their characteristics, efficacy and effectiveness against SARS-CoV-2 and variants of concern: A narrative review. Clin. Microbiol. Infect. 2022, 28, 202–221. [Google Scholar] [CrossRef] [PubMed]
  251. Long, B.; Carius, B.M.; Chavez, S.; Liang, S.Y.; Brady, W.J.; Koyfman, A.; Gottlieb, M. Clinical update on COVID-19 for the emergency clinician: Presentation and evaluation. Am. J. Emerg. Med. 2022, 54, 46–57. [Google Scholar] [CrossRef] [PubMed]
  252. Peeling, R.W.; Heymann, D.L.; Teo, Y.Y.; Garcia, P.J. Diagnostics for COVID-19: Moving from pandemic response to control. Lancet 2021, 399, 757–768. [Google Scholar] [CrossRef]
  253. Hoffmann, M.; Kleine-Weber, H.; Schroeder, S.; Kruger, N.; Herrler, T.; Erichsen, S.; Schiergens, T.S.; Herrler, G.; Wu, N.H.; Nitsche, A.; et al. SARS-CoV-2 Cell Entry Depends on ACE2 and TMPRSS2 and Is Blocked by a Clinically Proven Protease Inhibitor. Cell 2020, 181, 271–280.e8. [Google Scholar] [CrossRef]
  254. Glowacka, I.; Bertram, S.; Muller, M.A.; Allen, P.; Soilleux, E.; Pfefferle, S.; Steffen, I.; Tsegaye, T.S.; He, Y.; Gnirss, K.; et al. Evidence that TMPRSS2 activates the severe acute respiratory syndrome coronavirus spike protein for membrane fusion and reduces viral control by the humoral immune response. J. Virol. 2011, 85, 4122–4134. [Google Scholar] [CrossRef] [PubMed]
  255. Jia, H. Pulmonary Angiotensin-Converting Enzyme 2 (ACE2) and Inflammatory Lung Disease. Shock 2016, 46, 239–248. [Google Scholar] [CrossRef] [PubMed]
  256. Li, I.C.; Lu, T.Y.; Lin, T.W.; Chen, A.Y.; Chu, H.T.; Chen, Y.L.; Li, T.J.; Chen, C.C. Hispidin-enriched Sanghuangporus sanghuang mycelia SS-MN4 ameliorate disuse atrophy while improving muscle endurance. J. Cachexia Sarcopenia Muscle 2023, 14, 2226–2238. [Google Scholar] [CrossRef] [PubMed]
  257. Guo, J.; Zhao, S.; Ma, S.; Liang, Y. Effect of Phellinus Linzeus Polysaccharide on Anti-fatigue and Enhancing Anoxia Endurance Function. Edible Fungi China 2012, 31, 42–46. [Google Scholar] [CrossRef]
  258. Zeng, H. Effect of the Phellinus linteus Polysaccharides on Chromosome and Micronucleus of Murine Bone Marrow. Strait Pharm. J. 2014, 26, 16–18. [Google Scholar]
  259. Lin, H.; Lin, Y. Study on the PoIysaccharides from Phellinus linteus on Sperm AbnormaIity and Chromosome Aberration in TesticIe CeII of Mice. Strait Pharm. J. 2017, 29, 23–25. [Google Scholar]
  260. Hussain, S.; Sher, H.; Ullah, Z.; Elshikh, M.S.; Al Farraj, D.A.; Ali, A.; Abbasi, A.M. Traditional Uses of Wild Edible Mushrooms among the Local Communities of Swat, Pakistan. Foods 2023, 12, 1705. [Google Scholar] [CrossRef]
  261. Cardwell, G.; Bornman, J.F.; James, A.P.; Black, L.J. A Review of Mushrooms as a Potential Source of Dietary Vitamin D. Nutrients 2018, 10, 1498. [Google Scholar] [CrossRef]
  262. Jo Feeney, M.; Miller, A.M.; Roupas, P. Mushrooms-Biologically Distinct and Nutritionally Unique: Exploring a “Third Food Kingdom”. Nutr. Today 2014, 49, 301–307. [Google Scholar] [CrossRef] [PubMed]
  263. Khazai, N.; Judd, S.E.; Tangpricha, V. Calcium and vitamin D: Skeletal and extraskeletal health. Curr. Rheumatol. Rep. 2008, 10, 110–117. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (A) The growing state of the Sanghuang species; (BD) the dried state of the Sanghuang species.
Figure 1. (A) The growing state of the Sanghuang species; (BD) the dried state of the Sanghuang species.
Molecules 29 01195 g001
Figure 3. The pathways involved in the anti-inflammatory effects and sleep-improving effects of the Sanghuang species. Red arrows stand for the possible intervention spots of Sanghuang components.
Figure 3. The pathways involved in the anti-inflammatory effects and sleep-improving effects of the Sanghuang species. Red arrows stand for the possible intervention spots of Sanghuang components.
Molecules 29 01195 g003
Figure 4. The general process of studying bioactive components from Sanghuang.
Figure 4. The general process of studying bioactive components from Sanghuang.
Molecules 29 01195 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lu, J.; Su, M.; Zhou, X.; Li, D.; Niu, X.; Wang, Y. Research Progress of Bioactive Components in Sanghuangporus spp. Molecules 2024, 29, 1195. https://doi.org/10.3390/molecules29061195

AMA Style

Lu J, Su M, Zhou X, Li D, Niu X, Wang Y. Research Progress of Bioactive Components in Sanghuangporus spp. Molecules. 2024; 29(6):1195. https://doi.org/10.3390/molecules29061195

Chicago/Turabian Style

Lu, Jungu, Manman Su, Xuan Zhou, Deming Li, Xinhui Niu, and Yi Wang. 2024. "Research Progress of Bioactive Components in Sanghuangporus spp." Molecules 29, no. 6: 1195. https://doi.org/10.3390/molecules29061195

Article Metrics

Back to TopTop