Skip to main content

Increasing the production of the bioactive compounds in medicinal mushrooms: an omics perspective

Abstract

Macroscopic fungi, mainly higher basidiomycetes and some ascomycetes, are considered medicinal mushrooms and have long been used in different areas due to their pharmaceutically/nutritionally valuable bioactive compounds. However, the low production of these bioactive metabolites considerably limits the utilization of medicinal mushrooms both in commerce and clinical trials. As a result, many attempts, ranging from conventional methods to novel approaches, have been made to improve their production. The novel strategies include conducting omics investigations, constructing genome-scale metabolic models, and metabolic engineering. So far, genomics and the combined use of different omics studies are the most utilized omics analyses in medicinal mushroom research (both with 31% contribution), while metabolomics (with 4% contribution) is the least. This article is the first attempt for reviewing omics investigations in medicinal mushrooms with the ultimate aim of bioactive compound overproduction. In this regard, the role of these studies and systems biology in elucidating biosynthetic pathways of bioactive compounds and their contribution to metabolic engineering will be highlighted. Also, limitations of omics investigations and strategies for overcoming them will be provided in order to facilitate the overproduction of valuable bioactive metabolites in these valuable organisms.

Background

The application of mushrooms for medicinal purposes has a very long history [1]. Macroscopic fungi, mainly higher Basidiomycetes and some Ascomycetes, are considered medicinal mushrooms and can prevent, alleviate or cure several diseases and balance a healthy diet in the form of powders or extracts [2]. Many higher Basidiomycetes contain high/low molecular weight compounds, such as polysaccharides [3], lectins [4], triterpenes [5], statins, phenolic compounds, and antibiotics, in their fruit bodies, cultured mycelia, and cultured broth [6,7]. According to previous studies, some medicinal properties detected in mushrooms are as follows: antioxidant, antiviral, antifungal, antibacterial [4], antiobesity [8], cardiovascular protective [9], neuroprotective [10], immunomodulating, antitumor [3], hepatoprotective, cholesterol-lowering [11], antidiabetic [12], neuroregenerative, radical scavenging, and detoxicating activities [2,6,13]. For example, G. lucidum, a medicinal mushroom that possesses therapeutic activities such as antitumor, antioxidant, and immunomodulatory effects, is used for postponing aging, improving health, preventing and curing illnesses such as hypertension, gastric cancer, hepatitis, bronchitis as well as minor disorders including insomnia. In fact, it is possible to manufacture several valuable Ganoderma-based products, including soft capsules, injections, tablets, and drinks, by utilizing their spores and basidiocarps [14]. Thus, medicinal mushrooms are important for modern medicine and can be used as a new class of drugs known as “Mushroom Pharmaceuticals” to support a good quality of life and prevent illnesses such as immune system diseases [2,15].

From 1990 to 2020, global mushroom production has raised 13.8-fold to 42.8 million tons [16]. This global industry, which is consisted of edible, medicinal, and wild mushrooms, was approximated to be about $63 billion in 2013, with China being the leading producer of cultivated, edible mushrooms. 54% of this global industry is designated to cultivated, edible mushrooms and was around $34.1 billion in 2015 [17,18]. However, as more increase in edible mushroom consumption is anticipated in upcoming years, annual sales of this component of the world mushroom industry will grow from $34 to $60 billion [19] and their market will reach 24.05 million tons by 2028 [20]. Moreover, the remaining components of the global industry, namely medicinal mushrooms and wild mushrooms, represented 38% ($24 billion) and 8% ($5 billion) of the total value, respectively [17]. 85% of total mushroom production in the world is allocated to five fungal genera, i.e., Lentinula (the main genus) having about 22%, Pleurotus (mainly P. ostreatus, besides P. eryngii, P. djamor, P. pulmonarius, and P. citrinopileatus) with roughly 19% and Auricularia with approximately 17% of the world’s production. Next are Agaricus (mostly A. bisporus and considerably lower A. brasiliensis amounts) and Flammulina, the fourth and fifth most cultivated mushrooms, with 15% and 11% of the total amount, respectively [21]. Some other cultivated mushroom species are G. lucidum, V. volvacea, H. erinaceus, G. frondosa, and T. versicolor, which are desired edible and medicinal species in many regions of the globe [1]. Based on the most recent estimations, the market size of G. lucidum products is worth over US $2.5 billion [14,22]. Furthermore, It was estimated that the annual production of V.volvacea is 330,000 tons in China [23]. Antrodia cinnamomea and Cordyceps militaris are two other examples of medicinal mushrooms. According to estimations, products derived from A. cinnamomea, such as health foods and raw fruiting bodies, have a total market value of more than US$ 100 million annually [24], and the annual sale of C. militaris was evaluated to be about 3 billion RMB in China [25]. Although the current reach for other medicinal mushrooms may not be extensive at a global level, creating awareness about these mushrooms and their benefits will eventually increase their market potential.

Several examples of medicinal mushrooms (mainly those related to our review), their bioactive substances, medicinal properties, and applications are summarized in Table 1. In addition to the utilization of mushrooms as “Mushroom Pharmaceuticals,” they can be used as dietary foods, dietary supplement products, additive and ingredient replacers (such as meat substitutes) [26], cosmeceuticals [2,15], and analgesics [27]. Furthermore, as they possess insecticidal, fungicidal, nematocidal, antiphytoviral, bactericidal, and herbicidal effects, they can be utilized as natural biocontrol agents for plant protection [2,28]. There is an increasing demand for mushrooms due to all of the applications mentioned above as well as the nutritional value and pharmaceutical properties of their bioactive compounds. However, the low production of their bioactive compounds can be a bottleneck for clinical trials and commercial applications [29]. For example, improvements in the production of GA-T (a bioactive substance in G. lucidum) are needed to decrease production expenses and fulfill the demands in large-scale, commercial, and clinical trial fields [30]. Hence, many efforts have been made to increase the production yield of bioactive compounds in medicinal mushrooms via different methods such as optimizing the growth conditions (medium components and cultivation conditions) [31,32], signal transduction induction by inducers [5], and applying heat stress [33].

Table 1 Examples of medicinal mushrooms, their bioactive substances, medicinal properties, and applications

On the other hand, understanding the biosynthetic pathways of bioactive compounds as well as their complex regulation is necessary for achieving improvements in their production [29]. Thus, omics investigations can be novel, powerful, and beneficial tools in this regard. Still, omics approaches have not been adequately exploited for this purpose.

Omic tools, which provide a comprehensive view of cell metabolites, tissues, and organisms, are used to investigate the identification of genes (genomics), mRNA (transcriptomics), metabolites (metabolomics), and protein production (proteomics) under specific environmental conditions or by a particular approach. By utilizing transcriptomic and proteomic methods, it is possible to explain the roles of the fruiting body and vegetative mycelium during the detection of the genes that control the induction or repression of certain metabolic pathways. Moreover, metabolomics studies help determine the metabolites associated with every cellular process and those involved in different culture conditions [34]. To our knowledge, the genome, transcriptome, proteome, and metabolome studies on medicinal mushrooms for increasing the production of pharmaceutical compounds have been rarely reviewed. In fact, up until now 80 articles have conducted omics investigations on medicinal mushrooms with 48.75% of these studies being influential in bioactive compound overproduction. Thus, the present study aims to review for the first time, the omics analyses with the emphasis on improving bioactive substance production. The production of bioactive compounds will be compared before and after exploiting omics-based overproduction strategies and it will be shown that the maximum generated increase can be as high as fourfold. Challenges of omics technologies in medicinal mushroom research and their possible solutions will also be discussed.

Genomics studies on different medicinal mushrooms

Since genome data makes discovering and analyzing the biosynthesis of bioactive metabolites easier in higher fungi, chances for conducting research and developing their metabolic products can be provided by advancements in genome sequencing [129]. Up until now, genomic information of some edible/medicinal mushrooms including, A. bisporus [130], V. volvacea [23], Schizophyllum commune [131], F. velutipes [132], H. erinaceus [133], G. lucidum [134], C. militaris [138], Lignosus rhinocerotis [135], Ganoderma sinense [139], and Sanghuangporus sanghuang [140] has become available and resulted in gaining new insights into various aspects. The results of these genomic analyses are summarized in Table 2. For instance, genome sequencing of the model mushroom S. commune provides deeper knowledge of underlying mechanisms of mushroom formation. This knowledge can be helpful in their bioactive compounds production and their application in industry for achieving enzymes and pharmaceuticals.

Table 2 Summary of the genomics studies on edible/medicinal mushrooms

According to Table 2, genomics investigations have been an effective tool for studying medicinal mushrooms due to their roles in different subjects such as offering a genetic foundation of medicinal effects, improving biological and genetic studies, and elucidating genetic and enzymatic mechanisms in addition to biological characteristics related to different processes. Some of these processes are adaptation, degradation, sexual reproduction and development, sensitivity to different factors, mushroom formation, ethanol and medicinal compounds production, defense, evolutionary origins, and symbiosis. A summary of these applications as well as common techniques employed in genomics studies is demonstrated in Fig. 1. Also, information regarding the main techniques in genomics investigations, their different approaches, advantages, and limitations are provided in Table 3. For instance, FGENESH is the most rapid hidden Markov model-based program for precise ab initio gene structure prediction. When single-gene sequences are studied by this program, about 93% of all coding exon bases, along with 80% of human exons, can be predicted in 1.5 min. However, it is not as accurate as homology-based programs such as Exonerate and DIALIGN [149].

Fig. 1
figure 1

Commonly used techniques in genomics and a summary of the genomics applications in medicinal mushrooms

Table 3 Main techniques in genomics investigations, their different approaches, advantages, and limitations

The genomic studies, which are focused on determining biosynthetic pathways or biosynthetic gene clusters (BGCs) of bioactive metabolites and thus, can be considered more facilitative for increasing the production of these compounds, are discussed below in more detail.

Genomics studies for exploring BGCs

Recent progress in genome sequencing indicates that many putative BGCs are not visible in fungal genomes [157,158]. However, platforms for advanced genome mining, which are beneficial for exploring BGCs of natural bioactive compounds generated from multi-enzyme pathways, can be provided by the existing mushroom genomes [159].

Clearly, genome mining is able to be employed for discovering the biosynthetic genes of formerly acknowledged products as well as new, unfamiliar products by different techniques such as whole-genome comparisons and genome search approaches [159]. A large number of unprecedented fungal metabolic gene clusters determined through genome mining initially seem to be silent (called cryptic or orphan BCGs) and incapable of producing desirable metabolic products. Still, some approaches have become available for the activation of these silent gene clusters via the utilization of different stress types and co-culturing with bacteria [160] or other fungi [161,162]. Subsequently, stimulated gene expression can be further investigated via transcriptomics, proteomics studies, metabolomics [160,162,163,164,165,166,167,168,169,170], and co-expression correlations [171,172]. In addition to genome mining efforts, advancements in bioinformatics software, including antiSMASH, PRISM, and SMURF, have made the understanding of suppression or activation of microbial biosynthetic pathways possible [173].

Chen et al. determined the H. erinaceus gene clusters that participated in bioactive secondary metabolites biosynthesis (e.g., terpenoid and polyketides biosynthesis) by conducting genomic analyses including multiple sequence alignments, phylogenetic investigations, and using software such as antiSMASH [10]. Indeed, the prediction of three gene clusters associated with terpene production and one gene cluster relating to polyketides biosynthesis (PKS) in H. erinaceus resulted in discovering a novel family of diterpene cyclases in this fungus [10,174]. These results can make uncovering and production of valuable secondary metabolites of H. erinaceus and other medicinal mushrooms easier in the future and offer useful data for secondary metabolite exploration in other basidiomycetes. F. filiformis, with the genome length of 35.01 Mb and 10,396 gene models, was predicted to have thirteen putative terpenoid gene clusters, 12 sesquiterpene synthase genes from four different categories, and two type I polyketide synthase gene clusters in its genome. In comparison to its cultivar strain (81 genes), more terpenoid biosynthesis-associated genes were existent in the wild strain (119 genes) [61]. Moreover, the wild strain of F. filiformis has more terpenoid and polyketide synthase gene clusters compared to H. erinaceus. In another study, a distinct network of sesquiterpene synthases and two metabolic gene clusters, which contribute to illudin sesquiterpenoids biosynthesis, were demonstrated by the draft genome sequence of Omphalotus olearius. As a holistic survey of all currently available Basidiomycota genomes became possible through characterizing the sesquiterpene synthases, a prognosticative resource for biosynthesizing terpenoid natural products was presented in these mushrooms [148]. These findings will be a great help in the discovery and biosynthesis of peculiar pharmacologically relevant substances from Basidiomycota.

Genomics studies with the aim of elucidating biosynthetic pathways

Undoubtedly, studying the genome of medicinal mushrooms is effective for promoting research and development in pharmacological and industrial fields [129]. For instance, 16 cytochrome P450 superfamilies, possibly involved in the terpenoid synthesis, were detected by sequencing analysis of the G. lucidum genome via whole-genome shotgun strategy [129,136]. Detection of these superfamilies helped in determining the ganoderic acid synthetic pathway, massively producing triterpenoids, and achieving heterogonous expression through synthetic biotechnology. Moreover, a study on G. lucidum by Liu et al. indicated the genes associated with wood degradation and triterpene biosynthesis by comprehensive annotation of analyzed genes from the genome [137]. Regarding the model medicinal fungus, G.sinense, a comprehensive outline of its secondary metabolism and defense mechanisms was achieved through the investigation of DNA methylation patterns, small RNA transcriptomes, and complete genome sequence [139]. Thus, sequencing analysis, gene annotation, examining small RNA transcriptomes, and patterns of DNA methylation might be suitable techniques in concentrating genomic studies on secondary metabolite biosynthesis in the Ganoderma genus. Small RNA transcriptome analysis has not resulted in the overproduction of bioactive metabolites in G.lucidum yet. However, it was formerly demonstrated that microRNAs (miRNAs) could regulate secondary metabolite biosynthesis in many plants [175,176]. Hence, conducting small RNA profiling for determining miRNAs, studying miRNA-dependent regulation of valuable metabolites, and investigating miRNAs targeting genes associated with biosynthetic pathways can assist us in designing metabolic engineering strategies to improve bioactive substance contents in the desired organism.

As mentioned before (Sec “Background” section), increasing the production of a medicinal compound is not possible without having knowledge of its biosynthetic pathways and regulation. As the genes, pathways, and procedures related to the biosynthesis of the bioactive substances and wood decay by S. sanghuang were unidentified, Shao et al. investigated and reported a 34.5 Mb genome encoding 11,310 predicted genes of S. sanghuang. In this study, homologous genes associated with the biosynthesis of triterpenoids, polysaccharides, and flavonoids were determined. Then, the expression of these genes was investigated throughout four phases of development (10 and 20 days old mycelia, one-year-old fruiting bodies, and fruit bodies with three years of age). Furthermore, 343 transporters and four proteins of the velvet family, which were taking part in modulation, uptake, and redistribution of secondary metabolites, were detected [140]. As a result, genomics analysis can enhance our knowledge about secondary metabolites and their synthesis, which can be helpful for examining the medical applications of bioactive compounds and increasing their production in the future.

Not only the biosynthesis of sesquiterpenoids, antrocamphin, antroquinonol, ergostanes, and triterpenoids but also sexual development was clarified by exploiting genome ontology enrichment and pathway investigations in A. cinnamomea. Moreover, a 32.15-Mb draft genome including 9254 genes was achieved for this mushroom [97]. Also, the genome of H. erinaceus, which is consisted of 9895 genes, is 39.35 Mb and conveys different enzymes and a huge family of cytochrome P450 (CYP) proteins contributing to terpenoid backbones, sesquiterpenes, diterpenoids, and polyketides biosynthesis [10]. As another example, the obtained information from genome sequencing of C. militaris can significantly improve molecular research on the biology, fungal sex, and pathogenicity of this mushroom, uncover its mechanisms of medicinal compound synthesis, and be effective in the commercial production of its fruiting structures. In fact, utilizing the medicinal compounds of this mushroom can be facilitated by exploiting genome sequence data [138]. It is also worth mentioning that throughout the subculture and storage, C. militaris can experience a high frequency of strain degeneration which restricts the large-scale production of its bioactive compounds. In this case, genome-wide analysis of DNA methylation has shed light on the possible degeneration mechanisms of this strain [163] which will be beneficial for facilitating large-scale metabolite production. Regarding DNA methylation analysis, it is possible that the methylome repositories of P. tuoliensis and P. eryngii var. eryngii ease future investigations of epigenetic regulatory mechanisms supporting gene expression throughout the development of mushrooms. Thus, these repositories may have the potential to be considered as a guide for selecting the most suitable lifecycle/developmental phase for overproducing desired metabolites in medicinal mushrooms [164].

The genetic basis of the therapeutic activities of L. rhinocerotis, a comparative genomics source for polyporoid fungi and a platform for further identification of putative bioactive proteins and pathway enzymes of secondary metabolites is offered by the genome content of this mushroom [135]. By obtaining more information regarding biosynthetic pathways via genomic analyses, more targets for metabolic and pathway engineering can be found, which eventually contribute to rational predictions in the production of desired bioactive compounds.

Hitherto, more insights into the gene clusters or biosynthetic pathways of triterpenoids, ganoderic acids, polysaccharides, flavonoids, sesquiterpenoids, ergostanes, antroquinonol, antrocamphin, and polyketides in medicinal mushrooms have been achieved through genomic studies. Indeed, genomic investigations and genome sequencing programs are considered remarkable resource providers for determining new genes which contribute to the synthesis of bioactive substances (both known and novel substances). Also, more medicinal mushroom genomes will continue to become available [159]. Thus, progress in genome sequencing and genomic studies, genome mining, and bioinformatics, along with the availability of more genomes can greatly assist us in understanding the metabolic functions of desired organisms, which may result in both novel compound identification and improving the production of previously known valuable substances.

Transcriptomics studies on different medicinal mushrooms

The set of all RNA molecules, including mRNA and non-coding RNAs, which are transcribed in one cell or a population of cells, is defined as the transcriptome. In other words, it is the complete transcript set in a specified organism or a particular transcript subset in a specific type of cell. Although genomes of a given cell line are not changeable, external environmental conditions may cause the transcriptome to alter considerably. Because transcriptome includes every cellular mRNA transcript, it reveals the genes actively expressed at any particular moment, excluding mRNA degradation events [129].

In fact, expression profiling, together with advanced next-generation sequencing technology referred to as RNA sequencing (RNA-Seq) technology [177] and bioinformatics infrastructure, is among the most promising procedures for determining responsive genes, their modes of regulation, and related transcription factors in adaptation to certain abiotic and biotic components during a change in metabolism. In other words, in order to perform transcriptomic analysis at the level of nucleotides, high-throughput methods on the basis of DNA microarray technology or RNA-Seq are often used [129]. RNA-Seq allows the easy detection of rare and low-abundance transcripts, single-nucleotide polymorphisms, rare mutations and previously unknown gene isoforms, microbial RNAs, and regulatory micro-RNAs while microarray technology makes the parallel quantification of thousands of genes from various samples possible [178,179]. In addition, using Illumina sequencing technology has paved the way for de novo transcriptome assembly and analyzing gene expression even in species with no full genome data [180].

Transcriptomic analysis has been done in higher fungi [129], including different medicinal mushrooms such as C. militaris [181], G.lucidum [182], V.volvacea [183], P. ostreatus [184], Ophiocordyceps sinensis (Cordyceps sinensis) [185], H. erinaceus [10], F. filiformis [61], A. cinnamomea [97], P. eryngii [186], Termitomyces albuminosus [187], L. edodes [188], and L. rhinocerotis [169]. For instance, genome-wide transcriptome analysis was conducted on different developmental stages of artificially cultivated C. militaris and uncovered 2712 differentially expressed genes between its mycelium and fruiting body [181]. Moreover, as the result of performing developmental transcriptomics on O.sinensis, key pathways and hub genes in the development of this mushroom as well as the gene profile related to its sexual development was better understood, which adds novel data to current models of fruiting body development in edible fungi [189]. Also, Zhu et al. discovered 8906 potential RNA-editing sites in G. lucidum at the genomic level and the genes consisting of RNA-editing sites were functionally categorized by the Kyoto encyclopedia of genes and genomes (KEGG) enrichment and gene ontology analysis. As a result, laccase genes contributing to lignin degradation, key enzymes involved in triterpenoid biosynthesis, and transcription factors were enriched. Furthermore, the influence of transcriptional plasticity on the mushroom development and growth as well as on the adjustment of secondary metabolic biosynthetic pathways was elucidated [190].

Therefore, transcriptome analyses can provide a better understanding of gene expression changes in different developmental stages in medicinal mushrooms. Also, various processes have been clarified through transcriptomics. For instance, regarding P. ostreatus, genome and transcriptome analysis gave insights into the decay process in postharvest mushrooms and indicated the application of high-throughput techniques for establishing models of living organisms exposed to different environmental conditions [184]. In another study, the functional genes of the terpenoid biosynthesis pathway and wood degradation in G. lucidum were demonstrated by analyzing transcriptome through Illumina high-throughput technology [180]. Hence, the obtained transcriptome datasets offer a platform of beneficial public information for future functional genomics studies relating to medicinal mushrooms [188] and can set the stage for choosing the most suitable lifecycle/developmental phase for achieving better and increased production of desired compounds. On the other hand, RNA-Seq along with systems biology tools (such as genome-scale metabolic networks) enables the systematic recognition of reporter metabolites that represent important regions of the metabolic network [191] and hot spots regarding metabolic regulation [192,193]. Thus, these tools can also be advantageous for discovering candidate targets for metabolic engineering purposes. Indeed, by adopting systems approaches, we can initiate experiments toward strain improvement to gain enhanced production of fungal metabolites. Also, this enhancement can be achieved via different routes ranging from maneuvering on cultivation medium to manipulating the cellular metabolic regulation. Some transcriptomic findings related to the biosynthesis of bioactive compounds and the development of their production are discussed below.

Transcriptomics studies focused on cordycepin biosynthesis

The transcriptome of O. sinensis was investigated by Xiang et al. Examining adenosine kinase, 5′-nucleotidase, and adenylate kinase, which are possibly associated with the phosphorylation and dephosphorylation in the biosynthesis of cordycepin, offered valuable data for elucidating the cordycepin biosynthetic pathway. A model for cordycepin synthesis was also achieved [185]. This study offers a transcriptome dataset that can be considered a new resource for discovering genes (such as mating-type genes and genes associated with modulating signal transduction and the level of transcription in fruiting body development) besides examining and illuminating important biosynthetic and developmental pathways not only in O. sinensis but also in other medicinal mushrooms.

Although the metabolic pathways that contribute to the production of cordycepin were acknowledged to be linked to different carbon sources, the cellular regulatory procedures at the systems level were not well described [192]. Therefore, transcriptomic and genome-scale network-driven analyses were performed in C. militaris strain TBRC6039 cultivated on sucrose, glucose, and xylose carbon sources in order to examine the global metabolic response to the biosynthesis of cordycepin. Identification of 2883 DEGs, which were about 17% of the total 16,805 expressed genes, revealed sucrose and glucose-mediated alterations in the transcriptional regulation of central carbon metabolism (CCM). Also, reporter metabolites and main metabolic subnetworks including methionine, adenosine, and cordycepin, were offered via up-regulating cordycepin biosynthetic genes and after exploiting genome-scale metabolic network-driven analysis. These results present valuable data regarding C. militaris for systems-wide cordycepin overproduction [192] and indicate that the applied techniques, transcriptomics combined with genome-scale network-driven investigations, should also be extended to other higher fungi and other bioactive compounds in order to facilitate overproduction. Since C. militaris genome and RNA-sequencing data are available, integrating data for the investigation of cellular metabolism underlying cordycepin production has become possible [194]. Thus, the responsive mechanism of xylose consumption in C. militaris strain TBRC7358, the precursor and energy resources for cell growth and cordycepin production, and a remarkable role of putative alternative pathways for providing cordycepin production precursors on xylose were indicated by DEGs and the reporter metabolites analysis [195]. Enhancement of the cultivation procedure for increasing cordycepin and biomass productivities can be done with the help of the insight gained from this study which sheds light on the molecular mechanism underlying main metabolic pathways in transferring xylose towards cordycepin biosynthesis in C. militaris TBRC7358 [195]. These outcomes indicate that employing transcriptomic studies can clarify both main and alternative metabolic pathways related to the production of medicinal substances. Moreover, based on previous studies, genes related to cordycepin biosynthesis were up-regulated by growing C. militaris in favorable carbon sources. So, cultivating C. militaris strains for growth and cordycepin production relied on favored carbon sources proposing the essentiality of systems design of cultivation medium [196,197].

Another transcriptome analysis was performed on a C. militaris with a two-fold enhancement of cordycepin production caused by adding l-alanine to gain a deeper insight into molecular procedures of l-alanine’s effect on cordycepin biosynthesis. This investigation resulted in the achievement of a metabolic network map from the substrate amino acid to the product cordycepin and it was demonstrated that the Zn2Cys6-type transcription factors contributed to the development of C. militaris fruiting [13] as well as the regulation of its secondary metabolites [198]. This study indicates the plasticity of the cordycepin network, identifies the genes of rate-limiting enzymes in energy production pathways and amino acid conversion, and provides a suitable basis for future improvement of strain breeding and cordycepin yield. Also, these methods can be used for determining the influence of other inducers on metabolite biosynthesis from the molecular point of view.

So far, different tools such as genome-scale metabolic models (GSMMs) and genome-scale network-driven analyses, computer-assisted tools, reporter metabolites analysis, and information gained from other omics investigations have proved to be prominent for transcriptomics studies in cordycepin-producing mushrooms. Combining these tools and integrating their resultant data may generate new strategies for overproducing cordycepin.

Transcriptomics studies focused on the biosynthesis of other valuable bioactive compounds

In order to elucidate the biosynthetic pathway of carotenoids and its related genes, the transcriptomes of C. militaris mycelia grown under dark (CM10_D) and light exposure (CM10_L) conditions were sequenced and compared with each other. Furthermore, according to the KEGG pathway enrichment analysis of DEGs, most DEGs were elevated in “metabolic routes,” “MAPK signaling pathway-yeast,” and “secondary metabolite biosynthesis.” Also, the significant effect of the Cmtns gene in the biosynthesis of carotenoids was demonstrated in this mushroom [199]. Moreover, Yang et al. performed de novo sequencing and transcriptome investigation in the termite mushroom T. albuminosus, and their work resulted in the identification of enzymes related to saponin biosynthesis, including 22 glycosyltransferase and six cytochrome P450s genes [187]. As another example, the first transcriptome re-sequencing examination of L. rhinocerotis was performed by Yap et al., which uncovered the expression of several secondary metabolite biosynthetic routes (especially biosynthesis of terpene) along with putative genes associated with the biosynthesis of sclerotium glucans. Genes that encoded the sugar-binding lectins, cysteine-rich cerato-platanins, and hydrophobins were some of the genes with the highest expression in the sclerotium [169].

Role of comparative transcriptomics in medicinal compound overproduction

Profiling differences in gene expression covering different tissues of H. erinaceus (the monokaryotic mycelium (MK), dikaryotic mycelium (DK), and fruiting body) demonstrated the up-regulation of terpenoid biosynthesis-related genes in mycelia while the gene contributing to polyketides biosynthesis, experienced up-regulation in the fruiting body [10]. A similar study in F. filiformis revealed that contrary to H. erinaceus, a good number of terpenoid biosynthesis genes were up-regulated in the primordium and fruiting body of the wild strain, whereas polyketide synthase genes showed up-regulation in its mycelium. Relatively high transcript levels of UDP-glucose pyrophosphorylase and UDP-glucose dehydrogenase encoding genes, which are associated with the biosynthesis of polysaccharides, were observed in the mycelia as well as fruiting bodies [61]. In another study, DEGs between mycelia and fruiting bodies as well as 242 proteins in the mevalonate pathway, terpenoid pathways, polyketide synthases, and cytochrome P450s which may be related to the biosynthesis of secondary metabolites with therapeutic properties, were identified in A. cinnamomea. Expression enrichment was observed in genes of secondary metabolite routes for tissue-specific substances, such as 14-α-demethylase (CYP51F1) in the fruiting body for transforming lanostane to ergostane triterpenoids, coenzymes Q (COQ) for biosynthesizing antroquinonol in mycelium, and polyketide synthase for antrocamphin production in the fruiting body [97]. Tang et al. exploited RNA-seq technology for analyzing the poly (A) + transcriptome. They generated profiles for comparing the expression of Brown film (BF) and non-Brown film mycelia in order to elucidate the molecular mechanisms in L. edodes during light-induced BF formation. Through de novo assembly, a total of 31,511 contigs was achieved. Moreover, comparative analysis of the expression profiles demonstrated that prospective genes contributing to light-induced BF generation play important parts in fungal photoreception, the production of secondary metabolites, and signal transduction [188]. Henceforth, these findings can offer useful information for molecular breeding, selecting the best tissues/developmental stages with higher potential for producing elevated levels of the desired medicinal compounds, enhancing compound biosynthesis, and improvements in novel compound production through heterologous pathways and metabolic engineering. In addition, they will be advantageous for providing more insights into the mechanisms of gene expression and gene regulation besides further functional and pathway analysis.

In addition to determining DEGs among different tissues and developmental stages in an individual organism, comparative transcriptomics can be used for elucidating processes and gene expression differences among different culture conditions. For instance, G. lucidum goes through differentiation and morphological alterations in liquid static culture. This process, which results in the formation of aerial mycelia and asexual spores with substantial amounts of ganoderic acids, should be studied in order to allow large-scale production of asexual spores and ganoderic acids. Thus, comparative transcriptome analysis via suppression subtractive hybridization (SSH) method incorporated with cDNA array dot blotting was performed for identification of DEGs in liquid static culture contrasted with shaking culture of G. lucidum. Subsequently, 147 unigenes (such as unigenes regarding asexual sporulation and signal transduction) were detected in liquid static culture. Among these 147 unique sequences, protein database matches were identified for 101 (68.7%) expressed sequence tags (ESTs), 88 (59.8% of total) ESTs had considerable similarity to acknowledged proteins, and 13 (8.9% of total) sequences were comparable to hypothetical proteins. However, as there were slight resemblances to the recognized sequences for the remaining 46 ESTs (31.3%), they may demonstrate novel genes [200].

Based on the reviewed transcriptomics studies, it is exemplified that transcriptomic analyses are powerful tools that can be employed for several purposes, including enhancement of understanding about the functions and evolution of fungal genomes and the clarification of the molecular mechanisms of various cellular processes (e.g., mechanisms of gene expression and gene regulation). Furthermore, detection of reporter metabolites, investigation of the transcriptional response of desired organisms in the presence of different factors, and the determination of responsive genes, their modes of regulation, and related transcription factors can be facilitated by exploiting transcriptomic techniques.

Other applications of these techniques include the discovery of the differences in gene expression between various developmental stages and different culture conditions, understanding the changes during the development, and determination of the functional genes, enzymes, and biosynthetic pathways associated with bioactive compounds production. Thus, data obtained from transcriptome studies will be beneficial for investigating functional genomics in medicinal mushrooms, molecular breeding, bioactive compounds overproduction, and improving the synthesis of novel substances via heterologous pathways and metabolic engineering. Common techniques used in transcriptomics studies and a summary of the applications of transcriptome analyses in medicinal mushrooms are provided in Fig. 2.

Fig. 2
figure 2

Commonly used techniques in transcriptomics and a summary of the transcriptomic applications in medicinal mushrooms

Proteomics studies on different medicinal mushrooms

Methodical discovery and quantification of the complete protein set in a biologic system, namely cell, tissue, or organism, performed at a particular moment, are defined as proteomics analysis [129]. Proteome investigations can bring about a myriad of advantages. For instance, these studies are believed to be a suitable strategy for investigating mushroom developmental processes and understanding the roles of enzymes and proteins in prospective cultivation procedures, particularly in mushrooms with challenging cultivation conditions [201]. In addition to being helpful in better understanding the cellular metabolism [129], proteomics supports the identification of the reservoir of minerals and vitamins as well as protein effectors in mushrooms which possibly possess antibiotic, antitumor, antioxidant, antidiabetic, apoptosis, and blood pressure management effects [201]. Also, it is an effective tool for determining quantitative alterations in protein expression of filamentous fungi in reaction to stress exposure [202]. However, identifying all protein spots is not possible via proteomic analysis [129]. Thus, they should be exploited along with other omics studies. Proteomics techniques, including 2-dimensional gel electrophoresis (2-DE) or liquid chromatography coupled with mass spectrometry (LC − MS) (known as standard proteomic approaches) [203], 2DE gel-based proteomics [201], difference gel electrophoresis (DIGE) technology [204], LC-based techniques particularly high-throughput shotgun proteomics [205], gel-free proteomics [206], and iTRAQ labeling technique incorporated with two-dimensional liquid chromatography-tandem mass spectrometry (2D LC − MS/MS) [202], have turned into essential complements to genome and transcriptome techniques in fungal biology [207]. Moreover, 2DE gel-based proteomics is considered the most effective and commonly used technique for investigating fundamental physiological subjects in fungi, especially in edible mushrooms [201].

Proteomic analysis has been performed in different mushrooms such as L. rhinocerotis [208], T. heimii [209], A. bisporus [210], Pleurotus tuber-regium [211], A. cinnamomea [212], G. lucidum [170], P. ostreatus [213], and F. velutipes [214]. For instance, proteomic investigation of antihypertensive proteins was conducted in some edible mushrooms such as A. bisporus [210]. From another perspective, by examining protein expression profiles in different growth and developmental stages, a basis for the evaluation and comparison of these stages is offered in higher fungi. For example, information about biological processes contributing to the development of T. heimii was provided by exploiting the proteomic method of 2D-DIGE for the identification and investigation of the protein profiles of each developmental stage [209]. Moreover, protein fractions of three developmental stages in G.lucidum were analyzed by LC–MS/MS, and expression of a possibly novel highly immunomodulatory protein was indicated [170]. These comparative studies have also been conducted on P. tuber-regium [211] and A. cinnamomea [212].

Hence, both developmental stage assessment and novel mushroom compound identification can be achieved using proteomic techniques. Furthermore, analyzing changes in protein expression between two different mushroom species can be viewed as another application of proteomics that results in uncovering unique properties of individual organisms and eventually will be helpful in the detection of key compounds in their metabolisms. However, proteomic analysis is still in the early and developmental stages in higher fungi and edible mushrooms in comparison to bacterial, plant, and human proteomics investigations as a result of experiment costs and whether complete genome sequences of the mushrooms are available or not [129,201]. Nevertheless, proteomic studies have been executed on these organisms, including Pleurotus species, G. lucidum, and F. velutipes, aiming to improve bioactive metabolite production. These studies will be described below.

Proteomics studies in Pleurotus species

Apparently, the Pleurotus species is considered the most investigated genus of edible mushrooms in the proteomic subject area since it is among the most extensively cultured edible mushrooms [201]. Mycelial growth is limited in the presence of lignin in agro-industrial residues because of the intricate structure of the substrate and complications in using polysaccharides. Thus, investigating lignocellulose-fungi interactions is prominent for becoming aware of the ecology of fungi and optimizing the bioconversion of agro-industrial substrates to biotechnologically important products [34]. Attempts have been made in order to examine the procedure of the lignocellulose-fungi interactions via proteomic studies. For instance, the proteomic profile of P. ostreatus cultivated with different lignocellulose substrates as well as differentially expressed intracellular proteins in these substrates were reported by Xiao et al., which helped in studying the metabolic pathways associated with lignocellulose response in P. ostreatus. Also, 115 proteins were detected and it was demonstrated that enzymes contributing to sugar transformation via different metabolic routes experienced enhancement, and better growth was observed in the presence of xylan and carboxymethylcellulose [213]. In addition to P. ostreatus, these findings can also be useful for other white-rot fungi.

It was previously observed that applying Tween 80 to a submerged fermentation procedure can improve mycelial growth and the production of exopolysaccharides in P. tuber-regium by 51 and 42%, respectively [215]. Thus, a proteomic analysis was performed on this mushroom in order to identify the influence of stimulating agents (Tween 80) on mycelial growth and the production of exopolysaccharides in liquid culture. According to the results, a positive regulation on heat shock proteins (assist in maintaining cell viability under stressful circumstances) as well as on two isoforms of ATP-citrate lyase (can impede the Tricarboxylic acid (TCA) cycle activity and thereby increase exopolysaccharide biosynthesis) was detected. In fact, 32 proteins, which were expressed differentially, were determined by one-dimensional gel electrophoresis, and ATP: citrate lyase isoform 2 was able to increase exopolysaccharide production [216]. In addition to filling the information gap in the underdeveloped field of mushroom proteomics, these findings can explain how stimulatory agents, such as Tween 80, can improve the biosynthesis of beneficial compounds.

Proteomics studies in G. lucidum

Under nitrogen-limiting fermentation conditions, metabolic rearrangements take place due to the induction of growth inhibition via autophagy and imbalances between carbon (C) and N. These rearrangements adjust the division of cells, morphology, and lipids and starch cumulation processes in order to keep cellular structures safe and raise the survival probability. Since nitrogen (N) limitation is a suitable method for increasing ganoderic triterpenoid (GT) accumulation in G. lucidum, Lian et al. analyzed the dynamic adjustment of metabolism reallocation towards GT production in response to N limitation through exploiting iTRAQ-based proteome. Also, they attempted to identify the fundamental molecular mechanisms of the positive effect of N-limiting conditions on achieving high GT concentrations. As a result of applying N-limiting conditions, several changes were observed; (1) cell division ceased possibly due to the occurrence of autophagy, and cells modified their physiological and metabolic activities to compensate for the nutrient limitation; (2) N limitation did not affect cell growth tremendously but caused a considerable increase in GT amounts in the first 20 days. From the 10th day, extended duration of N limitation halted protein contents; (3) biosynthesis of nitrogen-containing substances experienced a decrease; (4) the generation of acetyl-CoA was promoted via metabolic reprogramming of CCM, which may supply GT biosynthesis; (5) in addition to up-regulation of enzymes involved in protein degradation, protein regulation in response to the abiotic stress and oxidation–reduction procedures carried out an important role in retaining cellular homeostasis; (6) while ongoing N limitation raised the mycelial contents of GT, it lowered biomass production of G. lucidum.

The obtained results show that the flux of carbon to GT in N deficient conditions resulted from the intermediary metabolism remodeling in the TCA cycle and glycolysis reactions. G. lucidum may utilize mechanisms such as glycolysis reinforcement and diminishment of other pathways in CCM to increase carbon flux solely toward secondary metabolites. Proteomics-based analyses, which helped in constructing a network of metabolism reallocation toward GT, demonstrated that glycolysis and the TCA cycle produce the carbon skeletons consolidated into GT precursors. Also, a basis for genetic engineering is offered by this study, which can allow the simultaneous synthesis of biomass and GT in G. lucidum [217]. These results may pave the way for establishing networks of metabolism reassignment toward bioactive compounds in other medicinal mushrooms, as well.

Proteomics studies in F. velutipes

Liu et al. applied iTRAQ labeling combined with the 2D LC − MS/MS method for determining the overall chronological alterations in patterns of protein expression and the mechanism of regulation of F. velutipes mycelia in reaction to light and cold stresses. Among the 1046 nonredundant identified proteins, 264 distinctively expressed proteins were related to 176 certain KEGG pathways. Based on comprehensive data analysis, the regulatory network underlying the mycelial light and cold reaction processes of F. velutipes was complicated and multi-dimensional. The reason behind this complexity was that it included different activities like quick energy supply, production of different compounds (lysine, γ-aminobutyric acid, phenylalanine, tyrosine), and calcium signal transduction procedure. Moreover, generating dynein-dependent actin and microtubule cytoskeleton, self-digestion, morphogenesis of organs and tissues, pigment secretion, acclimatization to oxidative stress, and other processes related to stress contribute to this complication [214]. In addition to being helpful for scientifically improving some mushroom cultivation techniques, this information may lead to a deeper understanding of the stress response mechanisms in macro-fungi.

According to the studies mentioned above, proteomics investigations can be utilized for different aims such as analyzing the developmental processes of mushrooms and their associated candidate genes and signaling pathways, examining fundamental physiological subjects, and establishing networks of metabolism reassignment toward bioactive metabolites. Other implementations of proteomic analyses are depicted in Fig. 3. Also, detecting quantitative changes in protein expression of filamentous fungi in response to stress or different factors, explaining the mechanism underlying these responses and their associated metabolic pathways is possible by employing these investigations. Thus, proteomics has become a necessary complement to genome and transcriptome techniques.

Fig. 3
figure 3

Commonly used techniques in proteomics and a summary of the proteomics applications in medicinal mushrooms

Combining transcriptomics and proteomics

Gene and protein expression profiling of medicinal mushrooms have helped in gaining knowledge about the genes and proteins involved in exogenous nutrient bag decomposition in Morchella importuna [218], temperature stress in L. edodes [219], bioactive metabolites in H. erinaceus [220], Cd2+ stress in P. eryngii [186], generation of special odor in S. commune [221], and development of the fruiting body in F. velutipes [222] and D. indusiata [223]. For instance, the study on P. eryngii indicated the coincidence of secondary metabolite production inhibition with the increase in carbohydrate metabolism and the rate of energy [186]. Transcriptomic and proteomic studies were also performed on a dikaryotic strain (DK13 × 3) that were emerged from two monokaryotic P. ostreatus strains (MK13 and MK3). This study offered evidence that growing a dikaryon organism is more advantageous than a monokaryon because the genes contributing to the utilization of macromolecules, cellular material synthesis, ability to withstand stress, and signal transduction had more regulation in the dikaryotic strain compared to MK13 and MK3 strains [224]. Thus, it will be possible to improve the characteristics of the strains and make them more resistant to the environment by selecting monokaryon organisms and doing the crosslink. As a result, the desired improvements will be observable in the formation of the dikaryon. The transcriptomic examinations and transcriptomics combined with proteomic studies on medicinal mushrooms are summarized in Table 4.

Table 4 Summary of transcriptomics studies/transcriptomics combined with proteomics studies on medicinal mushrooms

On the other hand, transcriptomics and proteomics investigations can pave the way for more developmental and medicinal research in mushrooms. For instance, a better understanding of changes during the morphological development of D. indusiata was achieved through de novo transcriptome assembly and shotgun proteomics of its fruiting bodies which resulted in the detection of 4380 proteins. Moreover, annotation and functional analysis of the determined proteins depicted their considerable increase in different activities such as small molecule synthetic and metabolic procedures [223].

High-throughput sequencing analysis was used to achieve transcriptomic and proteomic data with respect to mycelia and fruiting bodies of Agrocybe aegerita. The results of this work, which were helpful in illuminating the polysaccharide and sterol biosynthetic pathways, denoted that the polysaccharide was produced in great amounts in the fruiting bodies [129]. This data can be applied for constructing mushroom cell factories in the future. As another example, even though the genome sequence of T. guangdongense was available, there was not enough information regarding the regulatory networks of its metabolite production routes and sporocarp development. Thus, Wang et al. analyzed the transcriptome and proteome at distinctive developmental phases of T. guangdongense and found 9076 expressed genes as well as 2040 proteins. Also, hub genes were identified by exploiting weighted gene co-expression network analysis (WGCNA). As there was a small correlation between the transcriptomics and proteomics information, post-transcriptional procedures seem important in the development and growth of this mushroom [104]. Also, the down-regulation of terpenoid, polysaccharide, ergosterol, and adenosine production-related proteins was demonstrated during its development.

With respect to G.lucidum, combining De novo transcriptome assembly and proteomic studies under copper stress conditions pointed out genes related to terpenoid production routes and the breakdown of lignocellulose. As a result, it was shown that inducible lignin oxidative enzymes and proteins associated with secondary metabolic routes are highly abundant. Furthermore, through increasing Cu2+ concentrations, lignocellulase secretion in addition to antioxidants production was enhanced and about a fourfold increase was observed in phenolics production [226].

Omics technologies have been effectively utilized for investigating molecular mechanisms in Cordyceps fungi. Transcriptomic and proteomic analyses in artificially cultivated C. militaris have demonstrated the variations in gene expression between its mycelia and sporocarps. 2113 genes showed up-regulation in mycelia while 599 up-regulated genes were identified in sporocarps. Therefore, as it was inferred that the cordycepin metabolism pathway may have a higher activity in the mycelium of C. militaris, it is favorable to use the mycelium of this mushroom for the large-scale production of cordycepin [181].

Moreover, the efficiency of cordycepin can be decreased as the result of in vivo conversion to 3′-deoxyinosine by adenosine deaminase. Since pentostatin is able to impede adenosine deaminase, blending pentostatin with cordycepin can improve this efficiency. Thus, by exploiting transcriptomic and proteomic analyses, Zhao et al. investigated and reported a single gene cluster (consisting of four genes) associated with the production of cordycepin and pentostatin in Cordyceps kyushuensis. This cluster is able to be used for enhancing cordycepin yield and identifying more functional proteins [225]. As these results may also be observable in other Cordyceps fungi, conducting similar investigations on this genus is worth considering for increasing cordycepin production.

Thus, in addition to being an asset to better understanding mushroom development and obtaining strains with improved properties or more resistance to the environment, novel aspects and more data in other areas can be revealed by applying combined omics investigations on macro-fungi, compared to single omics studies. Some of these areas are biosynthetic pathways of bioactive metabolites, changes in the level of amino acids and other nutrients/metabolites, roles of regulatory factors, regulation of expression and cellular processes along with their molecular mechanisms, and the importance of post-transcriptional processes. Therefore, these investigations can eventually be used for increasing the yield of bioactive substances.

Metabolomics studies on different medicinal mushrooms

Since metabolome is dynamic and can be changed every second (similar to transcriptome and proteome), metabolic profiles are able to provide instant photos of the cell's physiological conditions [129]. Indeed, metabolomics is a high-throughput and novel approach [227] that can be applied to higher fungi in order to analyze, both qualitatively and quantitatively, their metabolome existent during a specific period or following induction in a specific condition. Moreover, this approach helps in understanding biological processes [227], determining variation in extrinsic and intrinsic environment perturbation response as well as various phenotypes by exploiting nuclear magnetic resonance (NMR) or combining mass spectrometry (MS) with other chemical analysis systems such as gas chromatography (GC/MS), HPLC (HPLC–MS), and capillary electrophoresis [228]. Metabolomics studies have been executed on Cordyceps bassiana, Phanerochaete chrysosporium, T.versicolor, Dichomitus squalens, P.ostreatus, and D.indusiata. Moreover, metabolite profiles have been exploited for chemotaxonomy [229] and for investigating different developmental phases or growth conditions in higher fungi. For instance, metabolic profiles of mycelia and fruiting bodies of C. bassiana were achieved via multivariate data analysis and H-1 NMR spectroscopy [230]. Also, measuring dynamic multi-parametric metabolic reactions of biological systems to genetic alterations or pathophysiological stimulants in a quantitative way is known as metabonomics. In fact, metabonomics is considered a subset of metabolomics [231] and is described as scientifically analyzing chemical processes including metabolites [129]. However, in order to arrive at more comprehensive conclusions, metabolomics study is regularly combined with other omics technologies such as proteomics and transcriptomics investigations [129]. For instance, metabolomic studies and proteomic investigations of the benzoic acid metabolism were carried out in P. chrysosporium [232].

Ergosterol, along with some of its biosynthetic intermediates, is valuable from an economic point of view, and the products of nearly every stage of ergosterol production are considered drug precursors [233,234]. Wang et al. investigated the differences in genes and metabolites in the ergosterol production route throughout the sporocarp development in F. velutipes by analyzing the transcriptome and metabolome of samples from three developmental phases. In fact, nine cDNA libraries were obtained from mycelia, young fruiting bodies, and mature fruiting bodies and sequenced via Illumina HiSeq 4000 platform. A total of 13 DEGs (six up-regulated and seven down-regulated) were identified throughout the development from mycelium to young sporocarps (T1), whereas solely one DEG (one down-regulated) was detected throughout the development from young sporocarps to mature ones (T2). Exploiting nontargeted metabolomics techniques resulted in the identification of a total of seven metabolites (three increased and four reduced) changed in content in the course of T1, and four metabolites were detected to be different in the period of T2. A combined investigation of the genome-wide connection network demonstrated that the metabolites, which were more probable to be adjusted, were chiefly in the post-squalene pathway part of the ergosterol biosynthetic pathway [235]. These results helped in gaining a deeper knowledge of the metabolic route of ergosterol production in F. velutipes. Therefore, combining metabolomics data with other omics datasets creates a powerful platform for answering many research questions. In addition to common methods and tools in metabolomics research, a summary of the applications of metabolomics investigations in medicinal mushrooms is provided in Fig. 4.

Fig. 4
figure 4

Commonly utilized approaches in metabolomics and a summary of the metabolomics applications in medicinal mushrooms

Overproduction strategies based on omics data

Utilizing omics data for the design and employment of overproduction strategies have raised the production of some important bioactive compounds in medicinal mushrooms. Results of these investigations are indicated in Table 5. For instance, it was anticipated that Zn2Cys6 transcript factors (mainly CCM_02568 and CCM_01481 genes) might play an important part in improving cordycepin production. Thus, these genes were overexpressed in C. militaris CM10. The overexpressed strains (CM10Tf1/CM10Tf2) were subjected to shake-flask fermentation with L-alanine being added after 5 days and results showed that the highest yield of cordycepin in the fermentation medium (99 mg/L) was about threefold higher compared to the wild type. Moreover, the highest yield of cordycepin in the mycelium of the overexpressed strains was 97 ng/g, which is again 3 times higher than the wild type mycelium [198]. Still, there are reports of higher cordycepin production yields even as high as 8.57 g/L by using non-omics-based strategies [240]. Thus, more attempts should be made for optimizing and boosting omics-based overproduction strategies and approaches. This has happened in the study of Ma et al. Based on a constructed GSMM and omics data of G.lucidum, they had previously found that the yield of extracellular polysaccharides can be enhanced by the addition of l-phenylalanine to the fermentation medium of this mushroom. Optimizing the concentration of l-phenylalanine for the production of extracellular polysaccharides showed that 0.4 g/L of this amino acid results in the maximum production of 0.79 g/L (45.49% increase). However, further optimization regarding the time of L-phenylalanine addition generated more increase in the production of extracellular polysaccharides and their yield was raised from 0.56 to 0.91 g/L by adding 0.4 g/L of the amino acid at 24 h, leading to a considerable increase of 62.50% [241].

Table 5 Results of using omics data for designing overproduction strategies in medicinal mushrooms

Challenges of omics investigations and possible solutions

Based on the studies covered in this review, the statistical contribution of each division of omics studies (i.e., genomics, transcriptomics, etc.) to medicinal mushroom research is demonstrated in a pie chart in Fig. 5. Genomics and integrated omics studies are both considered the most executed omics analyses on medicinal mushrooms. 36% of genomics as well as 57.69% of integrated omics studies produced data that can provide a suitable basis for increasing the production of bioactive compounds in future attempts. According to the pie chart, the next most utilized omics investigation is transcriptomics and 66.6% of the total transcriptomics analyses were allocated to those studies that can facilitate the overproduction of bioactive metabolites. The proteomics studies are in the third rank and 33.3% of these investigations have been performed with the purpose of facilitating the overproduction of the desired metabolites. Finally, metabolomics studies have the least contribution to medicinal mushroom research (4%).

Fig. 5
figure 5

Statistical perspectives of omics investigations in medicinal mushroom research

The number of studies associated with bioactive compound overproduction performed in each omics section is also presented in Fig. 5. The number of studies that have utilized more than one division of omics and performed integrated omics investigations is indicated at the intersections. Therefore, most of the omics studies aiming at improving the production of bioactive metabolites are in the field of combined omics, transcriptomics, and genomics, respectively. However, metabolomics and then proteomics investigations have had the least contribution to the overproduction of bioactive metabolites which is possibly due to the limitations and challenges of omics investigations.

For instance, proteome techniques are not meeting expectations, and reaching the complete proteome has not been accomplished yet. As gel-free proteomic techniques hold promise for future proteomics research of edible mushrooms, offer a broader range of protein coverage (such as membrane protein), and allow in-depth screening of protein synthesis and PTMs, designing future omics studies based on these techniques may be advantageous for achieving more comprehensive proteomic data in medicinal mushrooms. On the other hand, processing and analyzing proteomics data (LC/MS and LC–MS/MS data) is a very complicated multistep process which is the main bottleneck for many larger proteomics inquiries. These limitations can be conquered by effective sample preparation, modern mass spectrometry techniques, and extensive data processing and data analysis [252]. Another challenge is that identification of all protein spots cannot be carried out via proteomic analysis, and advancements in the not fully developed proteomics are dependent on experiment expenses and the availability of whole-genome sequences of mushrooms. Finding strategies for lowering the costs can facilitate and accelerate this development.

As de novo transcriptome assembly and analyzing gene expression, even in species with no full genome data, have been facilitated by Illumina sequencing technology, transcriptomics can be assumed to be less dependent on genomic investigations compared with proteomics. Metabolomics studies also face several challenges such as incomplete coverage of metabolites as well as hurdles and expenses in the experimental application, which may explain why they have been conducted to a lesser degree in medicinal mushrooms compared to other omics studies. For example, there are differences in sampling methods, sample preparation, instrumentation, and data mining between laboratories as well as among scientists in the same laboratory. Also, since no single platform is capable of interpreting the complete metabolome due to their specific analytical limitations, it can be hard to decide on the best platform for conducting metabolomic analyses. Still, the choice of analytical platform, which depends on both the sample and the purpose of the experiment, influences the result of the experiment and data recovery [253]. Different methods, which are frequently utilized in omics studies, are compared in Table 6, and a summary of their advantages and limitations is provided.

Table 6 Advantages and constraints of common techniques in omics studies

At the same time, individual omics investigation technology is faced with obstacles because modulation of cellular activity/metabolism levels has interaction with one another. Hence, it is crucial to use omic technologies integratively to obtain complete data [129]. Thus, although exploiting omics studies in medicinal mushroom research brings about a multitude of benefits, omics technologies are not free of challenges, and since they complement each other, combining omics studies can be beneficial for both achieving improved production of bioactive metabolites and eliminating restrictions.

Conclusions and future perspectives

There is a growing demand for medicinal mushrooms and their bioactive compounds due to nutritional benefits and pharmaceutical applications. Thus, increasing the production of these bioactive substances is essential for minimizing production expenses and meeting large-scale, commercial, and clinical trial demands. One of the methods which have helped in this regard is the exploitation of omics studies. In this review, the statistical contribution of each division of omics studies to medicinal mushroom research was discussed. The obtained omics data can be viewed as tools and prerequisites for systems biology, metabolic engineering, and cell factory construction endeavors. The cell factories obtained based on omics data will then be able to enhance the validness and rationality of synthetic biology and metabolic engineering approaches. This review highlighted that using omics analyses sets the stage for improving the production of bioactive compounds by discovering the functional genes, enzymes, key metabolic compounds, and biosynthetic pathways associated with their biosynthesis. Facilitating strain improvement, identifying more targets and strategies for metabolic and pathway engineering, establishing networks of metabolism reassignment toward bioactive metabolites, and creating powerful platforms for answering subsequent research questions were other assistive roles of omics in medicinal mushroom metabolite overproduction. Also, according to the quantitative data comparisons made among published investigations, it was demonstrated that creating overproduction strategies based on omics data can cause bioactive metabolite production values to experience increase ranging from 1.2 to fourfold. However, exploiting omics technologies and data for designing overproduction strategies in medicinal mushrooms is still far from sufficient.

Combining different levels of omics analyses and developing tools for genetic engineering facilitates the elucidation of the mechanisms of bioactive compound biosynthesis by higher fungi including medicinal mushrooms. This can eventually result in the overproduction and commercialization of the desired medicinal compounds. Moreover, combining omics data provides a comprehensive and systematic outlook, beneficial for the rational design and formulation of future overproduction strategies. Thus, aside from the need for a deeper focus on omics studies and the integration of their resulting data, future attempts must concentrate on improving these investigations and eliminating their limitations through different strategies. For example, combining the obtained data from omics studies with systems biology technologies such as GSMMs can provide better conditions for ideally designing and optimizing the cultivation mediums and increasing the yield of bioactive substances. It is important to mention that integrating proteomics, transcriptomics, and metabolomics data for gaining a better understanding of cellular biology is considered an obstacle in functional genomics and systems biology. Hence, resolving these issues in omics technologies can be noticeably helpful in improving the production of bioactive compounds. Also, as whole-genome sequences of these mushrooms continue to become accessible, we can expect progress in the field of omics studies, especially proteomics, in the future.

Availability of data and materials

All data are included in the manuscript and Additional information, and further queries about sharing data can be directed to the corresponding author.

Abbreviations

BGC:

Biosynthetic gene cluster

KEGG:

Kyoto encyclopedia of genes and genomes

CCM:

Central carbon metabolism

TCA:

Tricarboxylic acid

GT:

Ganoderic triterpenoid

NMR:

Nuclear magnetic resonance

GSMMs:

Genome-scale metabolic models

ESTs:

Expressed sequence tags

WGCNA:

Weighted gene co-expression network analysis

References

  1. Zervakis GI, Koutrotsios G. Solid-state fermentation of plant residues and agro-industrial wastes for the production of medicinal mushrooms. In: Agrawal DC, Tsay H-S, Shyur L-F, Wu Y-C, Wang S-Y, editors. Medicinal plants and fungi: recent advances in research and development. Singapore: Springer; 2017. p. 365–96.

    Chapter  Google Scholar 

  2. Gargano ML, Van Griensven LJLD, Omoanghe S, Lindequist U, Venturella G, Wasser SP, et al. Medicinal mushrooms : valuable biological resources of high exploitation potential. Plant Biosyst-An Int J Deal Asp Plant Biol. 2017;151(3):530–47. https://doi.org/10.1080/11263504.2017.1301590.

    Article  Google Scholar 

  3. Ma Z, Ye C, Deng W, Xu M, Wang Q, Liu G, et al. Reconstruction and analysis of a genome-scale metabolic model of Ganoderma lucidum for improved extracellular polysaccharide production. Front Microbiol. 2018;9:3076. https://doi.org/10.3389/fmicb.2018.03076.

    Article  Google Scholar 

  4. Rathore H, Prasad S, Sharma S. Mushroom nutraceuticals for improved nutrition and better human health: a review. PharmaNutrition. 2017;5(2):35–46.

    Article  Google Scholar 

  5. Ye L, Liu S, Xie F, Zhao L, Wu X. Enhanced production of polysaccharides and triterpenoids in Ganoderma lucidum fruit bodies on induction with signal transduction during the fruiting stage. PLoS ONE. 2018;13(4):e0196287.

    Article  Google Scholar 

  6. Wasser SP. Medicinal mushroom science: current perspectives, advances, evidences, and challenges. Biomed J. 2014;37(6):345–56.

    Article  Google Scholar 

  7. De Silva DD, Rapior S, Sudarman E, Stadler M, Xu J, Aisyah Alias S, et al. Bioactive metabolites from macrofungi: ethnopharmacology, biological activities and chemistry. Fungal Divers. 2013;62(1):1–40. https://doi.org/10.1007/s13225-013-0265-2.

    Article  Google Scholar 

  8. Chang C-J, Lin C-S, Lu C-C, Martel J, Ko Y-F, Ojcius DM, et al. Ganoderma lucidum reduces obesity in mice by modulating the composition of the gut microbiota. Nat Commun. 2015;6(1):7489. https://doi.org/10.1038/ncomms8489.

    Article  CAS  Google Scholar 

  9. Wang J, Wen X, Zhang Y, Zou P, Cheng L, Gan R, et al. Quantitative proteomic and metabolomic analysis of Dictyophora indusiata fruiting bodies during post-harvest morphological development. Food Chem. 2021;339(Feb 2020):127884.

    Article  CAS  Google Scholar 

  10. Chen J, Zeng X, Yang YL, Xing YM, Zhang Q, Li JM, et al. Genomic and transcriptomic analyses reveal differential regulation of diverse terpenoid and polyketides secondary metabolites in Hericium erinaceus. Sci Rep. 2017;7(Feb):10151.

    Article  Google Scholar 

  11. Fei Y, Li N, Zhang DH, Xu JW. Increased production of ganoderic acids by overexpression of homologous farnesyl diphosphate synthase and kinetic modeling of ganoderic acid production in Ganoderma lucidum. Microb Cell Fact. 2019;18(1):115.

    Article  CAS  Google Scholar 

  12. Zhang Y, Hu T, Zhou H, Zhang Y, Jin G, Yang Y. Antidiabetic effect of polysaccharides from Pleurotus ostreatus in streptozotocin-induced diabetic rats. Int J Biol Macromol. 2016;83:126–32.

    Article  CAS  Google Scholar 

  13. Ma G, Yang W, Zhao L, Pei F, Fang D, Hu Q. A critical review on the health promoting effects of mushrooms nutraceuticals. Food Sci Hum Wellness. 2018;7(2):125–33.

    Article  Google Scholar 

  14. Liu SR, Zhang WR. Hyperproduction of exopolysaccharides by submerged mycelial culture of Ganoderma lucidum using a solid seed grown in fine-powder of wheat bran and in vitro evaluation of the antioxidant activity of the exopolysaccharides produced. Food Sci Biotechnol. 2018;27(4):1129–36. https://doi.org/10.1007/s10068-018-0343-z.

    Article  CAS  Google Scholar 

  15. Chang ST, Wasser SP. Current and future research trends in agricultural and biomedical applications of medicinal mushrooms and mushroom products (review). Int J Med Mushrooms. 2018;20(12):1121–33.

    Article  Google Scholar 

  16. FAOSTAT. Food and Agriculture Organization of the United Nations Statistic. 2022. Database: http://www.fao.org/faostat/en/#data. Accessed 9 Nov 2022.

  17. Royse DJ, Baars J, Tan Q. Current overview of mushroom production in the world. Edible Med Mushrooms. 2017;2010:5–13.

    Article  Google Scholar 

  18. Niazi AR, Ghafoor A. Different ways to exploit mushrooms: a review. All Life. 2021;14(1):450–60. https://doi.org/10.1080/26895293.2021.1919570.

    Article  CAS  Google Scholar 

  19. Grimm D, Wösten HAB. Mushroom cultivation in the circular economy. Appl Microbiol Biotechnol. 2018;102(18):7795–803.

    Article  CAS  Google Scholar 

  20. Fortune Business Insights. Mushroom market to reach 24.05 million tonnes by 2028; growing health consciousness to boost nutritional food consumption and propel market growth. 2022. https://www.fortunebusinessinsights.com/press-release/mushroom-market-9301. Accessed 9 Nov 2022.

  21. Royse DJ. A Global perspective on the high five: agaricus, pleurotus, lentinula, auricularia & flammulina. In: Proceedings of 8th International Conference on Mushroom Biology and Mushroom Products (ICMBMP8), New Delhi, India, 19–22 November 2014 Vol. I & II. Solan: ICAR-Directorate of Mushroom Research; 2014. p. 1–6.

  22. Bishop KS, Kao CHJ, Xu Y, Glucina MP, Paterson RRM, Ferguson LR. From 2000years of Ganoderma lucidum to recent developments in nutraceuticals. Phytochemistry. 2015;114:56–65.

    Article  CAS  Google Scholar 

  23. Bao D, Gong M, Zheng H, Chen M, Zhang L, Wang H, et al. Sequencing and comparative analysis of the straw mushroom (Volvariella volvacea) genome. PLoS ONE. 2013;8(3):e58294.

    Article  CAS  Google Scholar 

  24. Wang H-C, Chu F-H, Chien S-C, Liao J-W, Hsieh H-W, Li W-H, et al. Establishment of the metabolite profile for an Antrodia cinnamomea health food product and investigation of its chemoprevention activity. J Agric Food Chem. 2013;61(36):8556–64.

    Article  CAS  Google Scholar 

  25. Dong C, Guo S, Wang W, Liu X. Cordyceps industry in China. Mycology. 2015;6(2):121–9.

    Article  Google Scholar 

  26. Rangel-Vargas E, Rodriguez JA, Domínguez R, Lorenzo JM, Sosa ME, Andrés SC, et al. Edible mushrooms as a natural source of food ingredient/additive replacer. Foods. 2021;10(11):2687.

    Article  CAS  Google Scholar 

  27. Ganeshpurkar A, Rai G. Experimental evaluation of analgesic and anti-inflammatory potential of Oyster mushroom Pleurotus florida. Indian J Pharmacol. 2013;45(1):66–70.

    Article  Google Scholar 

  28. Barseghyan GS, Barazani A, Wasser SP. Chapter 8-Medicinal mushrooms with anti-phytopathogenic and insecticidal properties. San Diego: Academic Press; 2016. p. 137–53.

    Google Scholar 

  29. Xiao H, Zhong J-J. Production of useful terpenoids by higher-fungus cell factory and synthetic biology approaches. Trends Biotechnol. 2016;34(3):242–55.

    Article  CAS  Google Scholar 

  30. Li HJ, Zhang DH, Han LL, Yu X, Zhao P, Li T, et al. Further improvement in ganoderic acid production in static liquid culture of Ganoderma lucidum by integrating nitrogen limitation and calcium ion addition. Bioprocess Biosyst Eng. 2016;39(1):75–80.

    Article  CAS  Google Scholar 

  31. Mao X-B, Eksriwong T, Chauvatcharin S, Zhong J-J. Optimization of carbon source/nitrogen ratio for cordycepin production by submerged cultivation of medicinal mushroom Cordyceps militaris. Process Biochem. 2005;1(40):1667–72.

    Article  Google Scholar 

  32. Kang C, Wen T-C, Kang J-C, Meng Z-B, Li G-R, Hyde KD. Optimization of large-scale culture conditions for the production of cordycepin with Cordyceps militaris by liquid static culture. ScientificWorldJournal. 2014;2014:510627.

    Article  Google Scholar 

  33. Tan X, Sun J, Xu Z, Li H, Hu J, Ning H, et al. Effect of heat stress on production and in-vitro antioxidant activity of polysaccharides in Ganoderma lucidum. Bioprocess Biosyst Eng. 2018;41(1):135–41. https://doi.org/10.1007/s00449-017-1850-7.

    Article  CAS  Google Scholar 

  34. Téllez-Téllez M, Díaz-Godínez G. Omic tools to study enzyme production from fungi in the Pleurotus genus. BioResources. 2019;14(1):2420–57.

    Article  Google Scholar 

  35. Bobovčák M, Kuniaková R, Gabriž J, Majtán J. Effect of pleuran (β-glucan from Pleurotus ostreatus) supplementation on cellular immune response after intensive exercise in elite athletes. Appl Physiol Nutr Metab Physiol Appl Nutr Metab. 2010;35(6):755–62.

    Article  Google Scholar 

  36. Jesenak M, Majtan J, Rennerova Z, Kyselovic J, Banovcin P, Hrubisko M. Immunomodulatory effect of pleuran (β-glucan from Pleurotus ostreatus) in children with recurrent respiratory tract infections. Int Immunopharmacol. 2013;15(2):395–9.

    Article  CAS  Google Scholar 

  37. Enman J, Hodge D, Berglund KA, Rova U. Production of the bioactive compound eritadenine by submerged cultivation of shiitake (Lentinus edodes) mycelia. J Agric Food Chem. 2008;56(8):2609–12. https://doi.org/10.1021/jf800091a.

    Article  CAS  Google Scholar 

  38. Fujii T, Maeda H, Suzuki F, Ishida N. Isolation and characterization of a new antitumor polysaccharide, KS-2, extracted from culture mycelia of Lentinus edodes. J Antibiot. 1978;31(11):1079–90.

    Article  CAS  Google Scholar 

  39. Minato K, Mizuno M, Terai H, Tsuchida H. Autolysis of lentinan, an antitumor polysaccharide, during storage of Lentinus edodes, shiitake mushroom. J Agric Food Chem. 1999;47(4):1530–2. https://doi.org/10.1021/jf981022w.

    Article  CAS  Google Scholar 

  40. Shimada Y, Morita T, Sugiyama K. Eritadenine-induced alterations of plasma lipoprotein lipid concentrations and phosphatidylcholine molecular species profile in rats fed cholesterol-free and cholesterol-enriched diets. Biosci Biotechnol Biochem. 2003;67(5):996–1006. https://doi.org/10.1271/bbb.67.996.

    Article  CAS  Google Scholar 

  41. Casaril K, Kasuya M, Vanetti M. Antimicrobial activity and mineral composition of shiitake mushrooms cultivated on agricultural waste. Braz Arch Biol Technol. 2011;54(5):991–1002.

    Article  CAS  Google Scholar 

  42. Gu Y-H, Belury MA. Selective induction of apoptosis in murine skin carcinoma cells (CH72) by an ethanol extract of Lentinula edodes. Cancer Lett. 2005;220(1):21–8.

    Article  CAS  Google Scholar 

  43. Handayani D, Chen J, Meyer BJ, Huang XF. Dietary shiitake mushroom (Lentinus edodes) prevents fat deposition and lowers triglyceride in rats fed a high-fat diet. J Obes. 2011;2011:258051.

    Article  CAS  Google Scholar 

  44. Ngai PHK, Ng TB. Lentin, a novel and potent antifungal protein from shitake mushroom with inhibitory effects on activity of human immunodeficiency virus-1 reverse transcriptase and proliferation of leukemia cells. Life Sci. 2003;73(26):3363–74.

    Article  CAS  Google Scholar 

  45. Ishikawa NK, Fukushi Y, Yamaji K, Tahara S, Takahashi K. Antimicrobial cuparene-type sesquiterpenes, enokipodins C and D, from a mycelial culture of Flammulina velutipes. J Nat Prod. 2001;64(7):932–4.

    Article  CAS  Google Scholar 

  46. Wu C-M, Wu T-Y, Kao S-S, Ko J-L, Jinn T-R. Expression and purification of a recombinant Fip-fve protein from Flammulina velutipes in baculovirus-infected insect cells. J Appl Microbiol. 2008;104(5):1354–62.

    Article  CAS  Google Scholar 

  47. Ikekawa T, Maruyama H, Miyano T, Okura A, Sawasaki Y, Naito K, et al. Proflamin, a new antitumor agent: preparation, physicochemical properties and antitumor activity. Jpn J Cancer Res. 1985;76(2):142–8.

    CAS  Google Scholar 

  48. Gong F, Chen Y, Gong M, Zhang C. Crystallization and some characterization of flammulin purified from the fruit bodies of Flammulina velutipes. Bioresour Technol. 1998;64(2):153–6.

    Article  CAS  Google Scholar 

  49. Hsieh KY, Hsu CI, Lin JY, Tsai CC, Lin RH. Oral administration of an edible-mushroom-derived protein inhibits the development of food-allergic reactions in mice. Clin Exp allergy J Br Soc Allergy Clin Immunol. 2003;33(11):1595–602.

    Article  CAS  Google Scholar 

  50. Bao HND, Ushio H, Ohshima T. Antioxidative activity and antidiscoloration efficacy of ergothioneine in mushroom (Flammulina velutipes) extract added to beef and fish meats. J Agric Food Chem. 2008;56(21):10032–40. https://doi.org/10.1021/jf8017063.

    Article  CAS  Google Scholar 

  51. Bao HND, Ushio H, Ohshima T. Antioxidative activities of mushroom (Flammulina velutipes) extract added to bigeye tuna meat: dose-dependent efficacy and comparison with other biological antioxidants. J Food Sci. 2009;74(2):C162–9.

    Article  CAS  Google Scholar 

  52. Lin L, Cui F, Zhang J, Gao X, Zhou M, Xu N, et al. Antioxidative and renoprotective effects of residue polysaccharides from Flammulina velutipes. Carbohydr Polym. 2016;146:388–95.

    Article  CAS  Google Scholar 

  53. Su A, Yang W, Zhao L, Pei F, Yuan B, Zhong L, et al. Flammulina velutipes polysaccharides improve scopolamine-induced learning and memory impairment in mice by modulating gut microbiota composition. Food Funct. 2018;9(3):1424–32.

    Article  CAS  Google Scholar 

  54. Zhang T, Ye J, Xue C, Wang Y, Liao W, Mao L, et al. Structural characteristics and bioactive properties of a novel polysaccharide from Flammulina velutipes. Carbohydr Polym. 2018;197:147–56.

    Article  CAS  Google Scholar 

  55. Hu Q, Yu J, Yang W, Kimatu BM, Fang Y, Ma N, et al. Identification of flavonoids from Flammulina velutipes and its neuroprotective effect on pheochromocytoma-12 cells. Food Chem. 2016;204:274–82.

    Article  CAS  Google Scholar 

  56. Wang Y, Bao L, Yang X, Li L, Li S, Gao H, et al. Bioactive sesquiterpenoids from the solid culture of the edible mushroom Flammulina velutipes growing on cooked rice. Food Chem. 2012;132(3):1346–53.

    Article  CAS  Google Scholar 

  57. Li H-P, Yang W-J, Qu S-X, Pei F, Luo X, Mariga AM, et al. Variation of volatile terpenes in the edible fungi mycelia Flammulina velutipes and communications in fungus-mite interactions. Food Res Int. 2018;103:150–5.

    Article  CAS  Google Scholar 

  58. Rahman MA, Abdullah N, Aminudin N. Antioxidative effects and inhibition of human low density lipoprotein oxidation in vitro of polyphenolic compounds in Flammulina velutipes (Golden Needle Mushroom). Oxid Med Cell Longev. 2015;2015:403023.

    Article  Google Scholar 

  59. Tang C, Hoo PC-X, Tan LT-H, Pusparajah P, Khan TM, Lee L-H, et al. Golden needle mushroom: a culinary medicine with evidenced-based biological activities and health promoting properties. Front Pharmacol. 2016;7:474.

    Article  Google Scholar 

  60. Kasprzycka A, Lalak-Kańczugowska J, Tys J. Flammulina velutipes treatment of non-sterile tall wheat grass for enhancing biodegradability and methane production. Bioresour Technol. 2018;263:660–4.

    Article  CAS  Google Scholar 

  61. Chen J, Li J, Tang Y, Ma K, Li B, Zeng X, et al. Genome-wide analysis and prediction of genes involved in the biosynthesis of polysaccharides and bioactive secondary metabolites in high-temperature-tolerant wild Flammulina filiformis. BMC Genomics. 2020;21:719.

    Article  CAS  Google Scholar 

  62. Kawagishi H, Masui A, Tokuyama S, Nakamura T. Erinacines J and K from the mycelia of Hericium erinaceum. Tetrahedron. 2006;1(62):8463–6.

    Article  Google Scholar 

  63. Lee EW, Shizuki K, Hosokawa S, Suzuki M, Suganuma H, Inakuma T, et al. Two novel diterpenoids, erinacines H and I from the mycelia of Hericium erinaceum. Biosci Biotechnol Biochem. 2000;64(11):2402–5.

    Article  CAS  Google Scholar 

  64. Kenmoku H, Shimai T, Toyomasu T, Kato N, Sassa T, Erinacine Q. a new erinacine from Hericium erinaceum, and its biosynthetic route to erinacine C in the basidiomycete. Biosci Biotechnol Biochem. 2002;66(3):571–5.

    Article  CAS  Google Scholar 

  65. Kawagishi H, Shirai R, Sakamoto H, Yoshida S, Ojima F, Ishiguro Y. Erinapyrones A and B from the cultured mycelia of Hericium erinaceum. Chem Lett. 1992;21(12):2475–6. https://doi.org/10.1246/cl.1992.2475.

    Article  Google Scholar 

  66. Lu Q-Q, Tian J-M, Wei J, Gao J-M. Bioactive metabolites from the mycelia of the basidiomycete Hericium erinaceum. Nat Prod Res. 2014;28(16):1288–92. https://doi.org/10.1080/14786419.2014.898145.

    Article  CAS  Google Scholar 

  67. Kobayashi S, Tamanoi H, Hasegawa Y, Segawa Y, Masuyama A. Divergent synthesis of bioactive resorcinols isolated from the fruiting bodies of Hericium erinaceum: total syntheses of hericenones A, B, and I, hericenols B-D, and erinacerins A and B. J Org Chem. 2014;79(11):5227–38. https://doi.org/10.1021/jo500795z.

    Article  CAS  Google Scholar 

  68. Wang K, Bao L, Qi Q, Zhao F, Ma K, Pei Y, et al. Erinacerins C-L, isoindolin-1-ones with α-glucosidase Inhibitory activity from cultures of the medicinal mushroom Hericium erinaceus. J Nat Prod. 2015;78(1):146–54. https://doi.org/10.1021/np5004388.

    Article  CAS  Google Scholar 

  69. Li JL, Lu L, Dai CC, Chen K, Qiu JY. A comparative study on sterols of ethanol extract and water extract from Hericium erinaceus. Zhongguo Zhong yao za zhi Zhongguo zhongyao zazhi China J Chinese Mater Med. 2001;26(12):831–4.

    CAS  Google Scholar 

  70. Li W, Zhou W, Kim E-J, Shim SH, Kang HK, Kim YH. Isolation and identification of aromatic compounds in Lion’s Mane Mushroom and their anticancer activities. Food Chem. 2015;170:336–42.

    Article  CAS  Google Scholar 

  71. Friedman M. Chemistry, nutrition, and health-promoting properties of Hericium erinaceus (Lion’s Mane) mushroom fruiting bodies and mycelia and their bioactive compounds. J Agric Food Chem. 2015;63(32):7108–23.

    Article  CAS  Google Scholar 

  72. Tsai-Teng T, Chin-Chu C, Li-Ya L, Wan-Ping C, Chung-Kuang L, Chien-Chang S, et al. Erinacine A-enriched Hericium erinaceus mycelium ameliorates Alzheimer’s disease-related pathologies in APPswe/PS1dE9 transgenic mice. J Biomed Sci. 2016;23(1):49.

    Article  Google Scholar 

  73. Ma B-J, Shen J-W, Yu H-Y, Ruan Y, Wu T-T, Zhao X. Hericenones and erinacines: stimulators of nerve growth factor (NGF) biosynthesis in Hericium erinaceus. Mycology. 2010;1(2):92–8. https://doi.org/10.1080/21501201003735556.

    Article  CAS  Google Scholar 

  74. Ma B-J, Yu H-Y, Shen J-W, Ruan Y, Zhao X, Zhou H, et al. Cytotoxic aromatic compounds from Hericium erinaceum. J Antibiot. 2010;63(12):713–5. https://doi.org/10.1038/ja.2010.112.

    Article  CAS  Google Scholar 

  75. Ueda K, Tsujimori M, Kodani S, Chiba A, Kubo M, Masuno K, et al. An endoplasmic reticulum (ER) stress-suppressive compound and its analogues from the mushroom Hericium erinaceum. Bioorg Med Chem. 2008;16(21):9467–70.

    Article  CAS  Google Scholar 

  76. Kawagishi H, Mori H, Uno A, Kimura A, Chiba S. A sialic acid-binding lectin from the mushroom Hericium erinaceum. FEBS Lett. 1994;340(1–2):56–8.

    Article  CAS  Google Scholar 

  77. Kawagishi H, Shimada A, Shirai R, Okamoto K, Ojima F, Sakamoto H, et al. Erinacines A, B and C, strong stimulators of nerve growth factor (NGF)-synthesis, from the mycelia of Hericium erinaceum. Tetrahedron Lett. 1994;35(10):1569–72.

    Article  CAS  Google Scholar 

  78. Gong M, An J, Lü H-Z, Wu C-F, Li Y-J, Cheng J-Q, et al. Effects of denaturation and amino acid modification on fluorescence spectrum and hemagglutinating activity of Hericium erinaceum Lectin. Acta Biochim Biophys Sin. 2004;36(5):343–50.

    Article  CAS  Google Scholar 

  79. Liao W, Luo Z, Liu D, Ning Z, Yang J, Ren J. Structure characterization of a novel polysaccharide from Dictyophora indusiata and its macrophage immunomodulatory activities. J Agric Food Chem. 2015;63(2):535–44. https://doi.org/10.1021/jf504677r.

    Article  CAS  Google Scholar 

  80. Liao W, Yu Z, Lin Z, Lei Z, Ning Z, Regenstein JM, et al. Biofunctionalization of selenium nanoparticle with Dictyophora Indusiata polysaccharide and its antiproliferative activity through death-receptor and mitochondria-mediated apoptotic pathways. Sci Rep. 2015;21(5):18629.

    Article  Google Scholar 

  81. Hua Y, Yang B, Tang J, Ma Z, Gao Q, Zhao M. Structural analysis of water-soluble polysaccharides in the fruiting body of Dictyophora indusiata and their in vivo antioxidant activities. Carbohydr Polym. 2012;87(1):343–7.

    Article  CAS  Google Scholar 

  82. Wang W, Liu H, Zhang Y, Feng Y, Yuan F, Song X, et al. Antihyperlipidemic and hepatoprotective properties of alkali- and enzyme-extractable polysaccharides by Dictyophora indusiata. Sci Rep. 2019;9(1):14266. https://doi.org/10.1038/s41598-019-50717-9.

    Article  CAS  Google Scholar 

  83. Wang Y, Lai L, Teng L, Li Y, Cheng J, Chen J, et al. Mechanism of the anti-inflammatory activity by a polysaccharide from Dictyophora indusiata in lipopolysaccharide-stimulated macrophages. Int J Biol Macromol. 2019;126:1158–66.

    Article  CAS  Google Scholar 

  84. Liu X, Chen Y, Wu L, Wu X, Huang Y, Liu B. Optimization of polysaccharides extraction from Dictyophora indusiata and determination of its antioxidant activity. Int J Biol Macromol. 2017;103:175–81.

    Article  CAS  Google Scholar 

  85. Huang M, Chen X, Tian H, Sun B, Chen H. Isolation and identification of antibiotic albaflavenone from Dictyophora indusiata (Vent: Pers.) Fischer. J Chem Res. 2011;35(11):659–60. https://doi.org/10.3184/174751911X13202334527264.

    Article  CAS  Google Scholar 

  86. Kawagishi H, Ishiyama D, Mori H, Sakamoto H, Ishiguro Y, Furukawa S, et al. Dictyophorines A and B, two stimulators of NGF-synthesis from the mushroom Dictyophora indusiata. Phytochemistry. 1997;45(6):1203–5.

    Article  CAS  Google Scholar 

  87. Sharma VK, Choi J, Sharma N, Choi M, Seo S-Y. In vitro anti-tyrosinase activity of 5-(hydroxymethyl)-2-furfural isolated from Dictyophora indusiata. Phytother Res. 2004;18(10):841–4.

    Article  CAS  Google Scholar 

  88. Lee I-K, Yun B-S, Han G, Cho D-H, Kim Y-H, Yoo I-D. Dictyoquinazols A, B, and C, new neuroprotective compounds from the mushroom Dictyophora indusiata. J Nat Prod. 2002;65(12):1769–72.

    Article  CAS  Google Scholar 

  89. Ishiyama D, Fukushi Y, Ohnishi-Kameyama M, Nagata T, Mori H, Inakuma T, et al. Monoterpene-alcohols from a mushroom Dictyophora indusiata. Phytochemistry. 1999;50(6):1053–6.

    Article  CAS  Google Scholar 

  90. Lee I-H, Huang R-L, Chen C-T, Chen H-C, Hsu W-C, Lu M-K. Antrodia camphorata polysaccharides exhibit anti-hepatitis B virus effects. FEMS Microbiol Lett. 2002;209(1):63–7.

    Article  CAS  Google Scholar 

  91. Yu Y-L, Chen I-H, Shen K-Y, Huang R-Y, Wang W-R, Chou C-J, et al. A triterpenoid methyl antcinate K isolated from Antrodia cinnamomea promotes dendritic cell activation and Th2 differentiation. Eur J Immunol. 2009;39(9):2482–91.

    Article  CAS  Google Scholar 

  92. Hsieh Y-H, Chu F-H, Wang Y-S, Chien S-C, Chang S-T, Shaw J-F, et al. Antrocamphin A, an anti-inflammatory principal from the fruiting body of Taiwanofungus camphoratus, and its mechanisms. J Agric Food Chem. 2010;58(5):3153–8. https://doi.org/10.1021/jf903638p.

    Article  CAS  Google Scholar 

  93. Huang N-K, Cheng J-J, Lai W-L, Lu M-K. Antrodia camphorata prevents rat pheochromocytoma cells from serum deprivation-induced apoptosis. FEMS Microbiol Lett. 2005;244(1):213–9.

    Article  CAS  Google Scholar 

  94. Huang C-H, Chang Y-Y, Liu C-W, Kang W-Y, Lin Y-L, Chang H-C, et al. Fruiting body of niuchangchih (Antrodia camphorata) protects livers against chronic alcohol consumption damage. J Agric Food Chem. 2010;58(6):3859–66. https://doi.org/10.1021/jf100530c.

    Article  CAS  Google Scholar 

  95. Huang G-J, Huang S-S, Lin S-S, Shao Y-Y, Chen C-C, Hou W-C, et al. Analgesic effects and the mechanisms of anti-inflammation of ergostatrien-3beta-ol from Antrodia camphorata submerged whole broth in mice. J Agric Food Chem. 2010;58(12):7445–52.

    Article  CAS  Google Scholar 

  96. Geethangili M, Tzeng Y-M. Review of pharmacological effects of Antrodia camphorata and its bioactive compounds. Evid Based Complement Alternat Med. 2011;2011:212641.

    Article  Google Scholar 

  97. Lu MJ, Fan W, Wang W, Chen T, Tang Y, Chu F. Genomic and transcriptomic analyses of the medicinal fungus Antrodia cinnamomea for its metabolite biosynthesis and sexual development. Proc Natl Acad Sci USA. 2014;111(44):E4743–52.

    Article  Google Scholar 

  98. Kumar KJS, Wang S-Y. Antioxidant properties of antrodia cinnamomea: an extremely rare and coveted medicinal mushroom endemic to Taiwan. In: Agrawal DC, Tsay H-S, Shyur L-F, Wu Y-C, Wang S-Y, editors. Medicinal plants and fungi: recent advances in research and development. Singapore: Springer; 2017. p. 135–64. https://doi.org/10.1007/978-981-10-5978-0_6.

    Chapter  Google Scholar 

  99. Cui JD. Biotechnological production and applications of Cordyceps militaris, a valued traditional Chinese medicine. Crit Rev Biotechnol. 2015;35(4):475–84.

    Article  Google Scholar 

  100. Ahn YJ, Park SJ, Lee SG, Shin SC, Choi DH. Cordycepin: selective growth inhibitor derived from liquid culture of Cordyceps militaris against Clostridium spp. J Agric Food Chem. 2000;48(7):2744–8.

    Article  CAS  Google Scholar 

  101. Zhou X, Meyer CU, Schmidtke P, Zepp F. Effect of cordycepin on interleukin-10 production of human peripheral blood mononuclear cells. Eur J Pharmacol. 2002;453(2–3):309–17.

    Article  CAS  Google Scholar 

  102. Nakamura K, Shinozuka K, Yoshikawa N. Anticancer and antimetastatic effects of cordycepin, an active component of Cordyceps sinensis. J Pharmacol Sci. 2015;127(1):53–6.

    Article  CAS  Google Scholar 

  103. Cha J-Y, Ahn H-Y, Cho Y-S, Je J-Y. Protective effect of cordycepin-enriched Cordyceps militaris on alcoholic hepatotoxicity in Sprague-Dawley rats. Food Chem Toxicol Int J Publ Br Ind Biol Res Assoc. 2013;60:52–7.

    Article  CAS  Google Scholar 

  104. Wang G, Li M, Zhang C, Cheng H, Gao Y, Deng W, et al. Transcriptome and proteome analyses reveal the regulatory networks and metabolite biosynthesis pathways during the development of Tolypocladium guangdongense. Comput Struct Biotechnol J. 2020;18:2081–94. https://doi.org/10.1016/j.csbj.2020.07.014.

    Article  CAS  Google Scholar 

  105. Zhang C, Huang H, Deng W, Li T. Genome-wide analysis of the Zn(II)2Cys6 zinc cluster-encoding gene family in Tolypocladium guangdongense and its light-induced expression. Genes. 2019;10(3):179.

    Article  CAS  Google Scholar 

  106. Mizuno T, Wang G, Zhang J, Kawagishi H, Nishitoba T, Li J. Reishi, Ganoderma lucidum and Ganoderma tsugae: bioactive substances and medicinal effects. Food Rev Int. 1995;11(1):151–66. https://doi.org/10.1080/87559129509541025.

    Article  CAS  Google Scholar 

  107. Paterson RRM. Ganoderma—a therapeutic fungal biofactory. Phytochemistry. 2006;67(18):1985–2001.

    Article  CAS  Google Scholar 

  108. Sone Y, Okuda R, Wada N, Kishida E, Misaki A. Structures and antitumor activities of the polysaccharides isolated from fruiting body and the growing culture of mycelium of Ganoderma lucidum. Agric Biol Chem. 1985;49(9):2641–53. https://doi.org/10.1080/00021369.1985.10867134.

    Article  CAS  Google Scholar 

  109. Upadhyay M, Shrivastava B, Jain A, Kidwai M, Kumar S, Gomes J, et al. Production of ganoderic acid by Ganoderma lucidum RCKB-2010 and its therapeutic potential. Ann Microbiol. 2014;64(2):839–46. https://doi.org/10.1007/s13213-013-0723-9.

    Article  CAS  Google Scholar 

  110. Xu J-W, Zhao W, Zhong J-J. Biotechnological production and application of ganoderic acids. Appl Microbiol Biotechnol. 2010;87(2):457–66.

    Article  CAS  Google Scholar 

  111. You B-J, Lee H-Z, Chung K-R, Lee M-H, Huang M-J, Tien N, et al. Enhanced production of ganoderic acids and cytotoxicity of Ganoderma lucidum using solid-medium culture. Biosci Biotechnol Biochem. 2012;76(8):1529–34.

    Article  CAS  Google Scholar 

  112. You BJ, Lee MH, Tien N, Lee MS, Hsieh HC, Tseng LH, et al. A novel approach to enhancing ganoderic acid production by Ganoderma lucidum using apoptosis induction. PLoS ONE. 2013;8(1):2–8.

    Article  Google Scholar 

  113. Zhou X-W, Su K-Q, Zhang Y-M. Applied modern biotechnology for cultivation of Ganoderma and development of their products. Appl Microbiol Biotechnol. 2012;93(3):941–63.

    Article  CAS  Google Scholar 

  114. Xu J-W, Xu Y-N, Zhong J-J. Production of individual ganoderic acids and expression of biosynthetic genes in liquid static and shaking cultures of Ganoderma lucidum. Appl Microbiol Biotechnol. 2010;85(4):941–8.

    Article  CAS  Google Scholar 

  115. Akihisa T, Nakamura Y, Tagata M, Tokuda H, Yasukawa K, Uchiyama E, et al. Anti-inflammatory and anti-tumor-promoting effects of triterpene acids and sterols from the fungus Ganoderma lucidum. Chem Biodivers. 2007;4(2):224–31.

    Article  CAS  Google Scholar 

  116. Berovič M, Habijanič J, Zore I, Wraber B, Hodžar D, Boh B, et al. Submerged cultivation of Ganoderma lucidum biomass and immunostimulatory effects of fungal polysaccharides. J Biotechnol. 2003;103(1):77–86.

    Article  Google Scholar 

  117. Feng J, Feng N, Tang QJ, Liu YF, Yang Y, Liu F, et al. Optimization of Ganoderma lucidum polysaccharides fermentation process for large-scale production. Appl Biochem Biotechnol. 2019;189(3):972–86.

    Article  CAS  Google Scholar 

  118. Hsieh C, Yang FC. Reusing soy residue for the solid-state fermentation of Ganoderma lucidum. Bioresour Technol. 2004;91(1):105–9.

    Article  CAS  Google Scholar 

  119. Jong SC, Birmingham JM. Medicinal benefits of the mushroom Ganoderma. Cambridge: Academic Press; 1992. p. 101–34.

    Google Scholar 

  120. Ma C, Feng M, Zhai X, Hu M, You L, Luo W, et al. Optimization for the extraction of polysaccharides from Ganoderma lucidum and their antioxidant and antiproliferative activities. J Taiwan Inst Chem Eng. 2013;44(6):886–94.

    Article  CAS  Google Scholar 

  121. Min BS, Gao JJ, Nakamura N, Hattori M. Triterpenes from the spores of Ganoderma lucidum and their cytotoxicity against meth-A and LLC tumor cells. Chem Pharm Bull. 2000;48(7):1026–33.

    Article  CAS  Google Scholar 

  122. Huo J, Zhong S, Du X, Cao Y, Wang W, Sun Y, et al. Whole-genome sequence of Phellinus gilvus (mulberry Sanghuang ) reveals its unique medicinal values. J Adv Res. 2020;24:325–35. https://doi.org/10.1016/j.jare.2020.04.011.

    Article  CAS  Google Scholar 

  123. Chen W, Tan H, Liu Q, Zheng X, Zhang H, Liu Y, et al. A review: the bioactivities and pharmacological applications of Phellinus linteus. Molecules. 2019;24(10):1888.

    Article  CAS  Google Scholar 

  124. Suabjakyong P, Saiki R, Van Griensven LJLD, Higashi K, Nishimura K, Igarashi K, et al. Polyphenol extract from Phellinus igniarius protects against acrolein toxicity in vitro and provides protection in a mouse stroke model. PLoS ONE. 2015;10(3):e0122733.

    Article  Google Scholar 

  125. Chang ST, Buswell JA. Mushroom nutriceuticals. World J Microbiol Biotechnol. 1996;12(5):473–6. https://doi.org/10.1007/BF00419460.

    Article  CAS  Google Scholar 

  126. Hung PV, Nhi NNY. Nutritional composition and antioxidant capacity of several edible mushrooms grown in the Southern Vietnam. Int Food Res J. 2012;19(2):611–5.

    CAS  Google Scholar 

  127. Cheung LM, Cheung PCK, Ooi VEC. Antioxidant activity and total phenolics of edible mushroom extracts. Food Chem. 2003;81(2):249–55.

    Article  CAS  Google Scholar 

  128. Joseph JA, Shukitt-Hale B, Denisova NA, Bielinski D, Martin A, McEwen JJ, et al. Reversals of age-related declines in neuronal signal transduction, cognitive, and motor behavioral deficits with blueberry, spinach, or strawberry dietary supplementation. J Neurosci. 1999;19(18):8114–21.

    Article  CAS  Google Scholar 

  129. Qin H, Xu J, Xiao J, Tang Y, Xiao H. Cell factories of higher fungi for useful metabolite production. Adv Biochem Eng Biotechnol. 2015;155:199–235.

    Google Scholar 

  130. Morin E, Kohler A, Baker AR, Foulongne-Oriol M, Lombard V, Nagye LG, et al. Genome sequence of the button mushroom Agaricus bisporus reveals mechanisms governing adaptation to a humic-rich ecological niche. Proc Natl Acad Sci. 2012;109(43):17501–6.

    Article  CAS  Google Scholar 

  131. Ohm RA, de Jong JF, Lugones LG, Aerts A, Kothe E, Stajich JE, et al. Genome sequence of the model mushroom Schizophyllum commune. Nat Biotechnol. 2010;28(9):957–63. https://doi.org/10.1038/nbt.1643.

    Article  CAS  Google Scholar 

  132. Park Y-J, Baek JH, Lee S, Kim C, Rhee H, Kim H, et al. Whole genome and global gene expression analyses of the model mushroom Flammulina velutipes reveal a high capacity for lignocellulose degradation. PLoS ONE. 2014;9(4):e93560.

    Article  Google Scholar 

  133. Gong W, Wang Y, Xie C, Zhou Y, Zhu Z, Peng Y. Whole genome sequence of an edible and medicinal mushroom, Hericium erinaceus (Basidiomycota, Fungi). Genomics. 2020;112(3):2393–9.

    Article  CAS  Google Scholar 

  134. Chen S, Xu J, Liu C, Zhu Y, Nelson DR, Zhou S, et al. Genome sequence of the model medicinal mushroom Ganoderma lucidum. Nat Commun. 2012;3(1):913. https://doi.org/10.1038/ncomms1923.

    Article  CAS  Google Scholar 

  135. Yap H-YY, Chooi Y-H, Firdaus-Raih M, Fung S-Y, Ng S-T, Tan C-S, et al. The genome of the tiger milk mushroom, Lignosus rhinocerotis, provides insights into the genetic basis of its medicinal properties. BMC Genomics. 2014;15(1):635.

    Article  Google Scholar 

  136. Shi L, Ren A, Mu D, Zhao M. Current progress in the study on biosynthesis and regulation of ganoderic acids. Appl Microbiol Biotechnol. 2010;88(6):1243–51.

    Article  CAS  Google Scholar 

  137. Liu D, Gong J, Dai W, Kang X, Huang Z, Zhang H-M, et al. The genome of Ganderma lucidum provide insights into triterpenes biosynthesis and wood degradation. PLoS ONE. 2012;7(5):e36146.

    Article  CAS  Google Scholar 

  138. Zheng P, Xia Y, Xiao G, Xiong C, Hu X, Zhang S, et al. Genome sequence of the insect pathogenic fungus Cordyceps militaris, a valued traditional Chinese medicine. Genome Biol. 2011;12(11):R116.

    Article  CAS  Google Scholar 

  139. Zhu Y, Xu J, Sun C, Zhou S, Xu H, Nelson DR, et al. Chromosome-level genome map provides insights into diverse defense mechanisms in the medicinal fungus Ganoderma sinense. Sci Rep. 2015;5:11087.

    Article  CAS  Google Scholar 

  140. Shao Y, Guo H, Zhang J, Liu H, Wang K, Zuo S, et al. The genome of the medicinal macrofungus Sanghuang provides insights into the synthesis of diverse secondary metabolites. Front Microbiol. 2020;10:3035. https://doi.org/10.3389/fmicb.2019.03035.

    Article  Google Scholar 

  141. Martin F, Kohler A, Murat C, Balestrini R, Coutinho PM, Jaillon O, et al. Périgord black truffle genome uncovers evolutionary origins and mechanisms of symbiosis. Nature. 2010;464(7291):1033–8. https://doi.org/10.1038/nature08867.

    Article  CAS  Google Scholar 

  142. Nagy LG, Riley R, Tritt A, Adam C, Daum C, Floudas D, et al. Comparative genomics of early-diverging mushroom-forming fungi provides insights into the origins of lignocellulose decay capabilities. Mol Biol Evol. 2016;33(4):959–70.

    Article  CAS  Google Scholar 

  143. Chen L, Gong Y, Cai Y, Liu W, Zhou Y, Xiao Y, et al. Genome sequence of the edible cultivated mushroom Lentinula edodes (Shiitake) reveals insights into lignocellulose degradation. PLoS ONE. 2016;11(8):e0160336.

    Article  Google Scholar 

  144. Yang R-H, Li Y, Wáng Y, Wan J-N, Zhou C-L, Wāng Y, et al. The genome of Pleurotus eryngii provides insights into the mechanisms of wood decay. J Biotechnol. 2016;239:65–7.

    Article  CAS  Google Scholar 

  145. Stajich JE, Wilke SK, Ahrén D, Au CH, Birren BW, Borodovsky M, et al. Insights into evolution of multicellular fungi from the assembled chromosomes of the mushroom Coprinopsis cinerea (Coprinus cinereus). Proc Natl Acad Sci USA. 2010;107(26):11889–94.

    Article  CAS  Google Scholar 

  146. Floudas D, Held BW, Riley R, Nagy LG, Koehler G, Ransdell AS, et al. Evolution of novel wood decay mechanisms in Agaricales revealed by the genome sequences of Fistulina hepatica and Cylindrobasidium torrendii. Fungal Genet Biol. 2015;76:78–92.

    Article  CAS  Google Scholar 

  147. Mercière M, Laybats A, Carasco-Lacombe C, Tan JS, Klopp C, Durand-Gasselin T, et al. Identification and development of new polymorphic microsatellite markers using genome assembly for Ganoderma boninense, causal agent of oil palm basal stem rot disease. Mycol Prog. 2015;14(11):103. https://doi.org/10.1007/s11557-015-1123-2.

    Article  Google Scholar 

  148. Wawrzyn GT, Quin MB, Choudhary S, López-Gallego F, Schmidt-Dannert C. Draft genome of Omphalotus olearius provides a predictive framework for sesquiterpenoid natural product biosynthesis in Basidiomycota. Chem Biol. 2012;19(6):772–83.

    Article  CAS  Google Scholar 

  149. Ejigu GF, Jung J. Review on the computational genome annotation of sequences obtained by next-generation sequencing. Biology. 2020;9(9):295.

    Article  CAS  Google Scholar 

  150. Lee N, Hwang S, Kim J, Cho S, Palsson B, Cho B-K. Mini review: Genome mining approaches for the identification of secondary metabolite biosynthetic gene clusters in Streptomyces. Comput Struct Biotechnol J. 2020;18:1548–56.

    Article  CAS  Google Scholar 

  151. Staden R. A strategy of DNA sequencing employing computer programs. Nucleic Acids Res. 1979;6(7):2601–10.

    Article  CAS  Google Scholar 

  152. Consortium IHGS. Finishing the euchromatic sequence of the human genome. Nature. 2004;431(7011):931–45. https://doi.org/10.1038/nature03001.

    Article  CAS  Google Scholar 

  153. Fuentes-Pardo AP, Ruzzante DE. Whole-genome sequencing approaches for conservation biology: advantages, limitations and practical recommendations. Mol Ecol. 2017;26(20):5369–406.

    Article  CAS  Google Scholar 

  154. Waterston RH, Lander ES, Sulston JE. On the sequencing of the human genome. Proc Natl Acad Sci USA. 2002;99(6):3712–6.

    Article  CAS  Google Scholar 

  155. Clark DP, Pazdernik NJ. Chapter 9-Genomics & systems biology. Boston: Academic Press; 2013. p. 248–72.

    Google Scholar 

  156. Verma S, Gazara RK. Chapter 3-Next-generation sequencing: an expedition from workstation to clinical applications. In: Raza K, Dey NBT-TB in H and M, editors. Translational Bioinformatics in Healthcare and Medicine. Cambridge: Academic Press; 2021. p. 29–47.

    Chapter  Google Scholar 

  157. Doroghazi JR, Albright JC, Goering AW, Ju K-S, Haines RR, Tchalukov KA, et al. A roadmap for natural product discovery based on large-scale genomics and metabolomics. Nat Chem Biol. 2014;10(11):963–8. https://doi.org/10.1038/nchembio.1659.

    Article  CAS  Google Scholar 

  158. Min B, Kim S, Oh Y-L, Kong W-S, Park H, Cho H, et al. Genomic discovery of the hypsin gene and biosynthetic pathways for terpenoids in Hypsizygus marmoreus. BMC Genomics. 2018;19(1):789. https://doi.org/10.1186/s12864-018-5159-y.

    Article  CAS  Google Scholar 

  159. Kües U, Badalyan SM. Making use of genomic information to explore the biotechnological potential of medicinal mushrooms. In: Agrawal DC, Tsay H-S, Shyur L-F, Wu Y-C, Wang S-Y, editors. Medicinal plants and fungi: recent advances in research and development. Singapore: Springer; 2017. p. 397–458. https://doi.org/10.1007/978-981-10-5978-0_13.

    Chapter  Google Scholar 

  160. Tauber JP, Schroeckh V, Shelest E, Brakhage AA, Hoffmeister D. Bacteria induce pigment formation in the basidiomycete Serpula lacrymans. Environ Microbiol. 2016;18(12):5218–27.

    Article  CAS  Google Scholar 

  161. Zheng W, Zhao Y, Zheng X, Liu Y, Pan S, Dai Y, et al. Production of antioxidant and antitumor metabolites by submerged cultures of Inonotus obliquus cocultured with Phellinus punctatus. Appl Microbiol Biotechnol. 2011;89(1):157–67.

    Article  CAS  Google Scholar 

  162. Yao L, Zhu L-P, Xu X-Y, Tan L-L, Sadilek M, Fan H, et al. Discovery of novel xylosides in co-culture of basidiomycetes Trametes versicolor and Ganoderma applanatum by integrated metabolomics and bioinformatics. Sci Rep. 2016;6(1):33237. https://doi.org/10.1038/srep33237.

    Article  CAS  Google Scholar 

  163. Xin X, Yin J, Zhang B, Li Z, Zhao S, Gui Z. Genome-wide analysis of DNA methylation in subcultured Cordyceps militaris. Arch Microbiol. 2019;201(3):369–75.

    Article  CAS  Google Scholar 

  164. Wen J, Zhang Z, Gong L, Xun H, Li J, Qi B, et al. Transcriptome changes during major developmental transitions accompanied with little alteration of DNA methylome in two pleurotus species. Genes. 2019;10(6):465.

    Article  CAS  Google Scholar 

  165. Kuan Y-C, Wu Y-J, Hung C-L, Sheu F. Trametes versicolor protein YZP activates regulatory B lymphocytes - gene identification through de novo assembly and function analysis in a murine acute colitis model. PLoS ONE. 2013;8(9):e72422.

    Article  CAS  Google Scholar 

  166. Martinez OF, Agbale CM, Nomiyama F, Franco OL. Deciphering bioactive peptides and their action mechanisms through proteomics. Expert Rev Proteomics. 2016;13(11):1007–16.

    Article  CAS  Google Scholar 

  167. Plaza DF, Lin C-W, van der Velden NSJ, Aebi M, Künzler M. Comparative transcriptomics of the model mushroom Coprinopsis cinerea reveals tissue-specific armories and a conserved circuitry for sexual development. BMC Genomics. 2014;15(1):492. https://doi.org/10.1186/1471-2164-15-492.

    Article  CAS  Google Scholar 

  168. Ren A, Li M-J, Shi L, Mu D-S, Jiang A-L, Han Q, et al. Profiling and quantifying differential gene transcription provide insights into ganoderic acid biosynthesis in Ganoderma lucidum in response to methyl jasmonate. PLoS ONE. 2013;8(6):e65027.

    Article  CAS  Google Scholar 

  169. Yap H-YY, Chooi Y-H, Fung S-Y, Ng S-T, Tan C-S, Tan N-H. Transcriptome analysis revealed highly expressed genes encoding secondary metabolite pathways and small cysteine-rich proteins in the sclerotium of Lignosus rhinocerotis. PLoS ONE. 2015;10(11):e0143549.

    Article  Google Scholar 

  170. Yu G-J, Yin Y-L, Yu W-H, Liu W, Jin Y-X, Shrestha A, et al. Proteome exploration to provide a resource for the investigation of Ganoderma lucidum. PLoS ONE. 2015;10(3):e0119439–e0119439.

    Article  Google Scholar 

  171. Umemura M, Koike H, Nagano N, Ishii T, Kawano J, Yamane N, et al. MIDDAS-M: motif-independent de novo detection of secondary metabolite gene clusters through the integration of genome sequencing and transcriptome data. PLoS ONE. 2013;8(12):e84028.

    Article  Google Scholar 

  172. Umemura M, Koike H, Machida M. Motif-independent de novo detection of secondary metabolite gene clusters-toward identification from filamentous fungi. Front Microbiol. 2015;5(6):371.

    Google Scholar 

  173. Baral B, Akhgari A, Metsä-Ketelä M. Activation of microbial secondary metabolic pathways: avenues and challenges. Synth Syst Biotechnol. 2018;3(3):163–78.

    Article  Google Scholar 

  174. Yang Y-L, Zhang S, Ma K, Xu Y, Tao Q, Chen Y, et al. Discovery and characterization of a new family of diterpene cyclases in bacteria and fungi. Angew Chemi Int Ed Engl. 2017;56(17):4749–52.

    Article  CAS  Google Scholar 

  175. Bulgakov VP, Avramenko TV. New opportunities for the regulation of secondary metabolism in plants: focus on microRNAs. Biotechnol Lett. 2015;37(9):1719–27.

    Article  CAS  Google Scholar 

  176. Kajal M, Singh K. Small RNA profiling for identification of miRNAs involved in regulation of saponins biosynthesis in Chlorophytum borivilianum. BMC Plant Biol. 2017;17(1):265. https://doi.org/10.1186/s12870-017-1214-0.

    Article  CAS  Google Scholar 

  177. Haas BJ, Papanicolaou A, Yassour M, Grabherr M, Blood PD, Bowden J, et al. De novo transcript sequence reconstruction from RNA-seq using the Trinity platform for reference generation and analysis. Nat Protoc. 2013;8(8):1494–512.

    Article  CAS  Google Scholar 

  178. Wang Z, Gerstein M, Snyder M. RNA-Seq: a revolutionary tool for transcriptomics. Nat Rev Genet. 2009;10(1):57–63.

    Article  CAS  Google Scholar 

  179. Li J, Hou R, Niu X, Liu R, Wang Q, Wang C, et al. Comparison of microarray and RNA-Seq analysis of mRNA expression in dermal mesenchymal stem cells. Biotechnol Lett. 2016;38(1):33–41.

    Article  CAS  Google Scholar 

  180. Yu G-J, Wang M, Huang J, Yin Y-L, Chen Y-J, Jiang S, et al. Deep insight into the Ganoderma lucidum by comprehensive analysis of its transcriptome. PLoS ONE. 2012;7(8):e44031.

    Article  CAS  Google Scholar 

  181. Yin Y, Yu G, Chen Y, Jiang S, Wang M, Jin Y, et al. Genome-wide transcriptome and proteome analysis on different developmental stages of Cordyceps militaris. PLoS ONE. 2012;7(12):e51853.

    Article  CAS  Google Scholar 

  182. Li J, Zhang J, Chen H, Chen X, Lan J, Liu C. Complete mitochondrial genome of the medicinal mushroom Ganoderma lucidum. PLoS ONE. 2013;8(8):e72038.

    Article  CAS  Google Scholar 

  183. Tao Y, van Peer AF, Chen B, Chen Z, Zhu J, Deng Y, et al. Gene expression profiling reveals large regulatory switches between succeeding stipe stages in Volvariella volvacea. PLoS ONE. 2014;9(5):e97789.

    Article  Google Scholar 

  184. Ramírez L, Oguiza JA, Pérez G, Lavín JL, Omarini A, Santoyo F, et al. Genomics and transcriptomics characterization of genes expressed during postharvest at 4°C by the edible basidiomycete Pleurotus ostreatus. Int Microbiol Off J Spanish Soc Microbiol. 2011;14(2):111–20.

    Google Scholar 

  185. Xiang L, Li Y, Zhu Y, Luo H, Li C, Xu X, et al. Transcriptome analysis of the Ophiocordyceps sinensis fruiting body reveals putative genes involved in fruiting body development and cordycepin biosynthesis. Genomics. 2014;103(1):154–9.

    Article  CAS  Google Scholar 

  186. Li Q, Huang W, Xiong C, Zhao J. Transcriptome analysis reveals the role of nitric oxide in Pleurotus eryngii responses to Cd(2+) stress. Chemosphere. 2018;201:294–302.

    Article  CAS  Google Scholar 

  187. Yang F, Xu B, Zhao S, Li J, Yang Y, Tang X, et al. De novo sequencing and analysis of the termite mushroom (Termitomyces albuminosus) transcriptome to discover putative genes involved in bioactive component biosynthesis. J Biosci Bioeng. 2012;114(2):228–31.

    Article  CAS  Google Scholar 

  188. Tang L-H, Jian H-H, Song C-Y, Bao D-P, Shang X-D, Wu D-Q, et al. Transcriptome analysis of candidate genes and signaling pathways associated with light-induced brown film formation in Lentinula edodes. Appl Microbiol Biotechnol. 2013;97(11):4977–89.

    Article  CAS  Google Scholar 

  189. Li X, Wang F, Liu Q, Li Q, Qian Z, Zhang X, et al. Developmental transcriptomics of Chinese cordyceps reveals gene regulatory network and expression profiles of sexual development-related genes. BMC Genomics. 2019;20(1):337.

    Article  Google Scholar 

  190. Zhu Y, Luo H, Zhang X, Song J, Sun C, Ji A, et al. Abundant and selective RNA-editing events in the medicinal mushroom Ganoderma lucidum. Genetics. 2014;196(4):1047–57.

    Article  CAS  Google Scholar 

  191. Mardinoglu A, Agren R, Kampf C, Asplund A, Uhlen M, Nielsen J. Genome-scale metabolic modelling of hepatocytes reveals serine deficiency in patients with non-alcoholic fatty liver disease. Nat Commun. 2014;5(1):3083. https://doi.org/10.1038/ncomms4083.

    Article  CAS  Google Scholar 

  192. Raethong N, Laoteng K, Vongsangnak W. Uncovering global metabolic response to cordycepin production in Cordyceps militaris through transcriptome and genome-scale network-driven analysis. Sci Rep. 2018;8(1):9250.

    Article  Google Scholar 

  193. Patil KR, Nielsen J. Uncovering transcriptional regulation of metabolism by using metabolic network topology. Proc Natl Acad Sci USA. 2005;102(8):2685–9.

    Article  CAS  Google Scholar 

  194. Vongsangnak W, Raethong N, Mujchariyakul W, Nguyen NN, Leong HW, Laoteng K. Genome-scale metabolic network of Cordyceps militaris useful for comparative analysis of entomopathogenic fungi. Gene. 2017;626:132–9.

    Article  CAS  Google Scholar 

  195. Wongsa B, Raethong N, Chumnanpuen P, Wong-ekkabut J, Laoteng K, Vongsangnak W. Alternative metabolic routes in channeling xylose to cordycepin production of Cordyceps militaris identified by comparative transcriptome analysis. Genomics. 2020;112(1):629–36.

    Article  CAS  Google Scholar 

  196. Nielsen J, Keasling JD. Engineering cellular metabolism. Cell. 2016;164(6):1185–97.

    Article  CAS  Google Scholar 

  197. O’Brien EJ, Monk JM, Palsson BO. Using genome-scale models to predict biological capabilities. Cell. 2015;161(5):971–87.

    Article  Google Scholar 

  198. Chen B, Wei T, Xue L, Zheng Q, Ye Z, Lin J. Transcriptome analysis reveals the flexibility of cordycepin network in cordyceps militaris activated by L-alanine addition. Front Microbiol. 2020;11:577.

    Article  Google Scholar 

  199. Lou H-W, Zhao Y, Tang H-B, Ye Z-W, Wei T, Lin J-F, et al. Transcriptome analysis of Cordyceps militaris reveals genes associated with carotenoid synthesis and identification of the function of the cmtns gene. Front Microbiol. 2019;10:2105.

    Article  Google Scholar 

  200. Xu J-W, Zhao W, Xu Y-N, Zhong J-J. Isolation and analysis of differentially expressed genes during asexual sporulation in liquid static culture of Ganoderma lucidum by suppression subtractive hybridization. Mol Biol Rep. 2012;39(4):3603–10.

    Article  CAS  Google Scholar 

  201. Al-Obaidi JR. Proteomics of edible mushrooms: a mini-review. Electrophoresis. 2016;37(10):1257–63.

    Article  CAS  Google Scholar 

  202. Kim Y, Nandakumar MP, Marten MR. Proteomics of filamentous fungi. Trends Biotechnol. 2007;25(9):395–400.

    Article  CAS  Google Scholar 

  203. Patterson SD, Aebersold RH. Proteomics: the first decade and beyond. Nat Genet. 2003;33(3):311–23. https://doi.org/10.1038/ng1106.

    Article  CAS  Google Scholar 

  204. Marouga R, David S, Hawkins E. The development of the DIGE system: 2D fluorescence difference gel analysis technology. Anal Bioanal Chem. 2005;382(3):669–78.

    Article  CAS  Google Scholar 

  205. Thelen JJ, Miernyk JA. The proteomic future: where mass spectrometry should be taking us. Biochem J. 2012;444(2):169–81.

    Article  CAS  Google Scholar 

  206. Bianco L, Perrotta G. Methodologies and perspectives of proteomics applied to filamentous fungi: from sample preparation to secretome analysis. Int J Mol Sci. 2015;16(3):5803–29.

    Article  CAS  Google Scholar 

  207. de Oliveira JMPF, de Graaff LH. Proteomics of industrial fungi: trends and insights for biotechnology. Appl Microbiol Biotechnol. 2011;89(2):225–37.

    Article  Google Scholar 

  208. Yap H-YY, Fung S-Y, Ng S-T, Tan C-S, Tan N-H. Genome-based proteomic analysis of Lignosus rhinocerotis (Cooke) Ryvarden sclerotium. Int J Med Sci. 2015;12(1):23–31.

    Article  CAS  Google Scholar 

  209. Rahmad N, Al-Obaidi JR, Nor Rashid NM, Zean NB, Mohd Yusoff MHY, Shaharuddin NS, et al. Comparative proteomic analysis of different developmental stages of the edible mushroom Termitomyces heimii. Biol Res. 2014;47(1):30.

    Article  Google Scholar 

  210. Lau C-C, Abdullah N, Shuib AS, Aminudin N. Proteomic analysis of antihypertensive proteins in edible mushrooms. J Agric Food Chem. 2012;60(50):12341–8.

    Article  CAS  Google Scholar 

  211. Chen L, Zhang B-B, Cheung PCK. Comparative proteomic analysis of mushroom cell wall proteins among the different developmental stages of Pleurotus tuber-regium. J Agric Food Chem. 2012;60(24):6173–82.

    Article  CAS  Google Scholar 

  212. Lin Y-L, Wen T-N, Chang S-T, Chu F-H. Proteomic analysis of differently cultured endemic medicinal mushroom Antrodia cinnamomea TT Chang et WN Chou from Taiwan. Int J Med Mushrooms. 2011;13(5):473–81.

    Article  CAS  Google Scholar 

  213. Xiao Q, Ma F, Li Y, Yu H, Li C, Zhang X. Differential proteomic profiles of Pleurotus ostreatus in response to lignocellulosic components provide insights into divergent adaptive mechanisms. Front Microbiol. 2017;8:480. https://doi.org/10.3389/fmicb.2017.00480.

    Article  Google Scholar 

  214. Liu J, Chang M, Meng J, Feng C, Wang Y. A comparative proteome approach reveals metabolic changes associated with Flammulina velutipes mycelia in response to cold and light stress. J Agric Food Chem. 2018;66:3716–25.

    Article  CAS  Google Scholar 

  215. Zhang B-B, Cheung PCK. Use of stimulatory agents to enhance the production of bioactive exopolysaccharide from Pleurotus tuber-regium by submerged fermentation. J Agric Food Chem. 2011;59(4):1210–6. https://doi.org/10.1021/jf104425w.

    Article  CAS  Google Scholar 

  216. Zhang B-B, Chen L, Cheung PCK. Proteomic insights into the stimulatory effect of Tween 80 on mycelial growth and exopolysaccharide production of an edible mushroom Pleurotus tuber-regium. Biotechnol Lett. 2012;34(10):1863–7.

    Article  Google Scholar 

  217. Lian D, Li L, Liu X, Zhong X, Wang H, Zhou S, et al. Time-scale dynamics of proteome predicts the central carbon metabolism involved in triterpenoid accumulation responsive to nitrogen limitation in Ganoderma lucidum. Fungal Biol. 2021;125(4):294–304.

    Article  CAS  Google Scholar 

  218. Tan H, Kohler A, Miao R, Liu T, Zhang Q, Zhang B, et al. Multi-omic analyses of exogenous nutrient bag decomposition by the black morel Morchella importuna reveal sustained carbon acquisition and transferring. Environ Microbiol. 2019;21(10):3909–26. https://doi.org/10.1111/1462-2920.14741.

    Article  CAS  Google Scholar 

  219. Wang G-Z, Ma C-J, Luo Y, Zhou S-S, Zhou Y, Ma X-L, et al. Proteome and transcriptome reveal involvement of heat shock proteins and indoleacetic acid metabolism process in Lentinula edodes thermotolerance. Cell Physiol Biochem. 2018;50(5):1617–37.

    Article  CAS  Google Scholar 

  220. Zeng X, Ling H, Yang J, Chen J, Guo S. Proteome analysis provides insight into the regulation of bioactive metabolites in Hericium erinaceus. Gene. 2018;666:108–15.

    Article  CAS  Google Scholar 

  221. Freihorst D, Brunsch M, Wirth S, Krause K, Kniemeyer O, Linde J, et al. Smelling the difference: transcriptome, proteome and volatilome changes after mating. Fungal Genet Biol. 2018;112:2–11.

    Article  CAS  Google Scholar 

  222. Liu J, Chang M, Meng J, Feng C, Zhao H, Zhang M. Comparative proteome reveals metabolic changes during the fruiting process in Flammulina velutipes. J Agric Food Chem. 2017;65(24):5091–100.

    Article  CAS  Google Scholar 

  223. Wang J, Wen X, Yang B, Liu D, Li X, Geng F. De novo transcriptome and proteome analysis of Dictyophora indusiata fruiting bodies provides insights into the changes during morphological development. Int J Biol Macromol. 2019;146:875–86. https://doi.org/10.1016/j.ijbiomac.2019.09.210.

    Article  CAS  Google Scholar 

  224. Liu T, Li H, Ding Y, Qi Y, Gao Y, Song A, et al. Genome-wide gene expression patterns in dikaryon of the basidiomycete fungus Pleurotus ostreatus. Brazilian J Microbiol. 2017;48(2):380–90.

    Article  CAS  Google Scholar 

  225. Zhao X, Zhang G, Li C, Ling J. Cordycepin and pentostatin biosynthesis gene identified through transcriptome and proteomics analysis of Cordyceps kyushuensis Kob. Microbiol Res. 2019;218:12–21.

    Article  CAS  Google Scholar 

  226. Jain KK, Kumar A, Shankar A, Pandey D, Chaudhary B, Sharma KK. De novo transcriptome assembly and protein profiling of copper-induced lignocellulolytic fungus Ganoderma lucidum MDU-7 reveals genes involved in lignocellulose degradation and terpenoid biosynthetic pathways. Genomics. 2020;112(1):184–98.

    Article  CAS  Google Scholar 

  227. Xu Y-J, Luo F, Gao Q, Shang Y, Wang C. Metabolomics reveals insect metabolic responses associated with fungal infection. Anal Bioanal Chem. 2015;407(16):4815–21. https://doi.org/10.1007/s00216-015-8648-8.

    Article  CAS  Google Scholar 

  228. Ito Y, Hirasawa T, Shimizu H. Metabolic engineering of Saccharomyces cerevisiae to improve succinic acid production based on metabolic profiling. Biosci Biotechnol Biochem. 2014;78(1):151–9. https://doi.org/10.1080/09168451.2014.877816.

    Article  CAS  Google Scholar 

  229. Heinke R, Schöne P, Arnold N, Wessjohann L, Schmidt J. Metabolite profiling and fingerprinting of Suillus species (basidiomycetes) by electrospray mass spectrometry. Eur J Mass Spectrom. 2014;20(1):85–97. https://doi.org/10.1255/ejms.1235.

    Article  Google Scholar 

  230. Park SJ, Hyun S-H, Suh HW, Lee S-Y, Sung G-H, Kim SH, et al. Biochemical characterization of cultivated Cordyceps bassiana mycelia and fruiting bodies by 1H nuclear magnetic resonance spectroscopy. Metabolomics. 2013;9(1):236–46. https://doi.org/10.1007/s11306-012-0442-4.

    Article  CAS  Google Scholar 

  231. Ramsden JJ. Metabolomics and metabonomics. In: Ramsden JJ, editor. Bioinformatics computational biology. London: Springer; 2009.

    Google Scholar 

  232. Matsuzaki F, Shimizu M, Wariishi H. Proteomic and metabolomic analyses of the white-rot fungus phanerochaete chrysosporium exposed to exogenous benzoic acid. J Proteome Res. 2008;7(6):2342–50. https://doi.org/10.1021/pr700617s.

    Article  CAS  Google Scholar 

  233. Rhee Y-H, Jeong S-J, Lee H-J, Lee H-J, Koh W, Jung JH, et al. Inhibition of STAT3 signaling and induction of SHP1 mediate antiangiogenic and antitumor activities of ergosterol peroxide in U266 multiple myeloma cells. BMC Cancer. 2012;20(12):28.

    Article  Google Scholar 

  234. Pluchino LA, Liu AK-Y, Wang H-CR. Reactive oxygen species-mediated breast cell carcinogenesis enhanced by multiple carcinogens and intervened by dietary ergosterol and mimosine. Free Radic Biol Med. 2015;80:12–26.

    Article  CAS  Google Scholar 

  235. Wang R, Ma P, Li C, Xiao L, Liang Z, Dong J. Combining transcriptomics and metabolomics to reveal the underlying molecular mechanism of ergosterol biosynthesis during the fruiting process of Flammulina velutipes. BMC Genomics. 2019;20:999.

    Article  Google Scholar 

  236. Raethong N, Wang H, Nielsen J, Vongsangnak W. Optimizing cultivation of Cordyceps militaris for fast growth and cordycepin overproduction using rational design of synthetic media. Comput Struct Biotechnol J. 2020;18:1–8.

    Article  CAS  Google Scholar 

  237. Zhou J, Ji S, Ren M, He Y, Jing X, Xu J. Enhanced accumulation of individual ganoderic acids in a submerged culture of Ganoderma lucidum by the overexpression of squalene synthase gene. Biochem Eng J. 2014;90:178–83. https://doi.org/10.1016/j.bej.2014.06.008.

    Article  CAS  Google Scholar 

  238. Xu J-W, Ji S-L, Li H-J, Zhou J-S, Duan Y-Q, Dang L-Z, et al. Increased polysaccharide production and biosynthetic gene expressions in a submerged culture of Ganoderma lucidum by the overexpression of the homologous α-phosphoglucomutase gene. Bioprocess Biosyst Eng. 2015;38(2):399–405.

    Article  CAS  Google Scholar 

  239. Ma Z, Xu M, Wang Q, Wang F, Zheng H, Gu Z, et al. Development of an efficient strategy to improve extracellular polysaccharide production of Ganoderma lucidum using L-phenylalanine as an enhancer. Front Microbiol. 2019;10(Oct):2306.

    Article  Google Scholar 

  240. Das SK, Masuda M, Hatashita M, Sakurai A, Sakakibara M. A new approach for improving cordycepin productivity in surface liquid culture of Cordyceps militaris using high-energy ion beam irradiation. Lett Appl Microbiol. 2008;47(6):534–8.

    Article  CAS  Google Scholar 

  241. Ma Z, Xu M, Wang Q, Wang F, Zheng H, Gu Z, et al. Development of an Efficient Strategy to Improve Extracellular Polysaccharide Production of Ganoderma lucidum Using L-Phenylalanine as an Enhancer. Front Microbiol. 2019;10(October):2306.

    Article  Google Scholar 

  242. Albarano L, Esposito R, Ruocco N, Costantini M. Genome mining as new challenge in natural products discovery. Mar Drugs. 2020;18(4):199.

    Article  CAS  Google Scholar 

  243. Wilhelm BT, Landry J-R. RNA-Seq-quantitative measurement of expression through massively parallel RNA-sequencing. Methods. 2009;48(3):249–57.

    Article  CAS  Google Scholar 

  244. Zhao S, Fung-Leung W-P, Bittner A, Ngo K, Liu X. Comparison of RNA-Seq and microarray in transcriptome profiling of activated T cells. PLoS ONE. 2014;9(1):e78644.

    Article  Google Scholar 

  245. Wang C, Gong B, Bushel PR, Thierry-Mieg J, Thierry-Mieg D, Xu J, et al. The concordance between RNA-seq and microarray data depends on chemical treatment and transcript abundance. Nat Biotechnol. 2014;32(9):926–32.

    Article  CAS  Google Scholar 

  246. Liu Y, Morley M, Brandimarto J, Hannenhalli S, Hu Y, Ashley EA, et al. RNA-Seq identifies novel myocardial gene expression signatures of heart failure. Genomics. 2015;105(2):83–9.

    Article  CAS  Google Scholar 

  247. Whitley SK, Horne WT, Kolls JK. Research techniques made simple: methodology and clinical applications of RNA sequencing. J Invest Dermatol. 2016;136(8):e77-82.

    Article  CAS  Google Scholar 

  248. Russo G, Zegar C, Giordano A. Advantages and limitations of microarray technology in human cancer. Oncogene. 2003;22(42):6497–507. https://doi.org/10.1038/sj.onc.1206865.

    Article  CAS  Google Scholar 

  249. Joshi K, Patil D. Chapter 9 - Proteomics. In: Patwardhan B, Chaguturu RBT-IA in DD, editors. Innovative Approaches in Drug Discovery. Boston: Academic Press; 2017. p. 273–94.

    Book  Google Scholar 

  250. Baggerman G, Vierstraete E, De Loof A, Schoofs L. Gel-based versus gel-free proteomics: a review. Comb Chem High Throughput Screen. 2005;8(8):669–77.

    Article  CAS  Google Scholar 

  251. Granlund I, Hall M, Schröder WP. Difference gel electrophoresis (DIGE). eLS. 2009. https://doi.org/10.1002/9780470015902.a0021881. (Major Reference Works).

    Article  Google Scholar 

  252. Drabik A, Bodzoń-Kułakowska A, Silberring J. Gel Electrophoresis. In: Ciborowski P, Silberring JBT-PP editors. Proteomic Profiling and Analytical Chemistry: The Crossroads, vol.2. Boston: Elsevier; 2016. p. 115–43

    Chapter  Google Scholar 

  253. Meleady P. Two-dimensional gel electrophoresis and 2D-DIGE. Methods Mol Biol. 2018;1664:3–14.

    Article  CAS  Google Scholar 

  254. Beretov J, Wasinger VC, Graham PH, Millar EK, Kearsley JH, Li Y. Proteomics for breast cancer urine biomarkers. Adv Clin Chem. 2014;63:123–67.

    Article  CAS  Google Scholar 

  255. Xie F, Liu T, Qian W-J, Petyuk VA, Smith RD. Liquid chromatography-mass spectrometry-based quantitative proteomics. J Biol Chem. 2011;286(29):25443–9.

    Article  CAS  Google Scholar 

  256. Emwas A-H, Roy R, McKay RT, Tenori L, Saccenti E, Gowda GAN, et al. NMR spectroscopy for metabolomics research. Metabolites. 2019;9(7):123.

    Article  CAS  Google Scholar 

  257. Chandramouli K, Qian P-Y. Proteomics: challenges, techniques and possibilities to overcome biological sample complexity. Hum Genomics Proteomics. 2009;8(2009):239204.

    Google Scholar 

  258. Johnson CH, Gonzalez FJ. Challenges and opportunities of metabolomics. J Cell Physiol. 2012;227(8):2975–81.

    Article  CAS  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

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

Author information

Authors and Affiliations

Authors

Contributions

NA and HN wrote the manuscript. HM and HN supervised the project and revised the manuscript. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Hoda Nouri or Hamid Moghimi.

Ethics declarations

Ethics approval and consent to participate

Ethics approval not applicable as this article does not describe any studies involving human participants or animals.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher's Note

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

Rights and permissions

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

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Arshadi, N., Nouri, H. & Moghimi, H. Increasing the production of the bioactive compounds in medicinal mushrooms: an omics perspective. Microb Cell Fact 22, 11 (2023). https://doi.org/10.1186/s12934-022-02013-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12934-022-02013-x

Keywords