Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Unusual genome expansion and transcription suppression in ectomycorrhizal Tricholoma matsutake by insertions of transposable elements

  • Byoungnam Min ,

    Contributed equally to this work with: Byoungnam Min, Hyeokjun Yoon

    Roles Data curation, Formal analysis, Investigation, Methodology, Project administration, Resources, Software, Validation, Writing – original draft, Writing – review & editing

    Current Address: Department of Plant and Microbial Biology, University of California, Berkeley, CA, United States of America

    Affiliation Department of Biotechnology, College of Life Sciences and Biotechnology, Korea University, Seoul, Korea

  • Hyeokjun Yoon ,

    Contributed equally to this work with: Byoungnam Min, Hyeokjun Yoon

    Roles Data curation, Formal analysis, Investigation, Methodology, Project administration, Resources, Writing – original draft

    Affiliation School of Life Sciences and Biotechnology, College of Natural Sciences, Kyungpook National University, Daegu, Korea

  • Julius Park,

    Roles Data curation, Formal analysis, Investigation, Methodology, Software, Writing – original draft

    Affiliation Department of Biotechnology, College of Life Sciences and Biotechnology, Korea University, Seoul, Korea

  • Youn-Lee Oh,

    Roles Data curation, Funding acquisition, Investigation, Methodology

    Affiliation Mushroom Research Division, National Institute of Horticulture and Herbal Science (NIHHS), Rural Development Administration (RDA), Eumseong, Korea

  • Won-Sik Kong ,

    Roles Data curation, Formal analysis, Funding acquisition, Investigation, Project administration, Resources, Supervision

    igchoi@korea.ac.kr (IC); wskong@korea.kr (WK); kimjg@knu.ac.kr (JK)

    Affiliation Mushroom Research Division, National Institute of Horticulture and Herbal Science (NIHHS), Rural Development Administration (RDA), Eumseong, Korea

  • Jong-Guk Kim ,

    Roles Data curation, Formal analysis, Funding acquisition, Investigation, Methodology, Project administration, Supervision

    igchoi@korea.ac.kr (IC); wskong@korea.kr (WK); kimjg@knu.ac.kr (JK)

    Affiliation School of Life Sciences and Biotechnology, College of Natural Sciences, Kyungpook National University, Daegu, Korea

  • In-Geol Choi

    Roles Conceptualization, Data curation, Formal analysis, Funding acquisition, Investigation, Methodology, Project administration, Resources, Software, Supervision, Validation, Visualization, Writing – original draft, Writing – review & editing

    igchoi@korea.ac.kr (IC); wskong@korea.kr (WK); kimjg@knu.ac.kr (JK)

    Affiliation Department of Biotechnology, College of Life Sciences and Biotechnology, Korea University, Seoul, Korea

Correction

13 Feb 2020: The PLOS ONE Staff (2020) Correction: Unusual genome expansion and transcription suppression in ectomycorrhizal Tricholoma matsutake by insertions of transposable elements. PLOS ONE 15(2): e0229406. https://doi.org/10.1371/journal.pone.0229406 View correction

Abstract

Genome sequencing of Tricholoma matsutake revealed its unusually large size as 189.0 Mbp, which is a consequence of extraordinarily high transposable element (TE) content. We identified that 702 genes were surrounded by TEs, and 83.2% of these genes were not transcribed at any developmental stage. This observation indicated that the insertion of TEs alters the transcription of the genes neighboring these TEs. Repeat-induced point mutation, such as C to T hypermutation with a bias over “CpG” dinucleotides, was also recognized in this genome, representing a typical defense mechanism against TEs during evolution. Many transcription factor genes were activated in both the primordia and fruiting body stages, which indicates that many regulatory processes are shared during the developmental stages. Small secreted protein genes (<300 aa) were dominantly transcribed in the hyphae, where symbiotic interactions occur with the hosts. Comparative analysis with 37 Agaricomycetes genomes revealed that IstB-like domains (PF01695) were conserved across taxonomically diverse mycorrhizal genomes, where the T. matsutake genome contained four copies of this domain. Three of the IstB-like genes were overexpressed in the hyphae. Similar to other ectomycorrhizal genomes, the CAZyme gene set was reduced in T. matsutake, including losses in the glycoside hydrolase genes. The T. matsutake genome sequence provides insight into the causes and consequences of genome size inflation.

Introduction

Tricholoma matsutake is an ectomycorrhizal (ECM) basidiomycete that establishes a symbiotic relationship with the roots of Pinus densiflora, giving it the name “pine mushroom” [1]. ECM fungi build an aggregated hyphal sheath that encases the whole root tip of its symbiotic partner and mediates the root’s external interactions with the soil [2]. This encasing or root colonization is formed through a hyphal network called the “Hartig net”, which is located inside the root cells–an anatomical pattern that is shared by the majority of ECM fungi [3,4]. The fruiting body of T. matsutake is a highly valued edible mushroom in many countries [1,5]. Unfortunately, attempts to cultivate the fruiting body have been unsuccessful, and the mechanism of mushroom development has not yet been fully understood.

Mushroom formation is proceeded by distinct developmental stages that include the vegetative hyphae stage, the dikaryotic primordia stage, and the mature fruiting body stage [6]. Various genes, including transcriptional factors [7], hydrophobins [8], and light receptors [9], have been suggested as critical genetic factors for fruiting body formation in basidiomycetes. Systematic transcriptomic surveys on fruiting body formation have been carried out for various basidiomycetes [10].

Transposable elements (TEs) play an important role in genome evolution by causing chromosomal rearrangements or by reshaping the regulatory networks [11,12]. Many ECM genomes show a high TE content, leading to comparably larger genome sizes [13], and contain more TEs than their asymbiotic relatives [14]. The effect of the presence of TEs in mushrooms is transcriptional repression, particularly when genes are surrounded by the TEs [15]. In a recent comparative genomic study of two mushroom strains, Pleurotus ostreatus PC15 and PC9, the genes surrounded by transposons in one strain showed strong transcriptional repression, whereas their orthologs in the other strain were normally expressed [15]. Despite the higher TE content in ECM genomes, the transcription tendency of the ECM genes affected by TEs has not been thoroughly examined.

Here, we report the genome sequence of T. matsutake and the transcriptional dynamics over three distinct developmental stages. The most distinct features of the T. matsutake genome were genome expansion by the many TEs and prevailing transcriptional suppression in all developmental stages. In addition, we performed comparative analyses on the T. matsutake and 37 Agaricomycetes genomes to identify potential gene clusters involved in symbiosis.

Results and discussion

Genomic summary of T. matsutake

Sequencing of the dikaryotic genomic DNA of T. matsutake generated a total length of 189.0 Mbp within 5,255 scaffolds with 111.8× sequencing coverage. We predicted 15,305 gene models using the FunGAP pipeline [16]. The predicted genes were examined for their reliability by RNA-seq, functional domains, and orthologs; thereby, the 14,528 (94.9%) genes were supported by at least one piece of evidence (Fig A in S1 File). A genome completeness test using BUSCO v3.0.2 [17] showed >99% coverage of single-copy orthologs in Basidiomycota (1,323 of 1,335 entries), validating the complete genome assembly and annotation. Because the genome was dikaryotic, we investigated how many genes were allelic by comparing two-member gene families with their relative genomes. As a result, we identified that allele genes were not frequent in the assembled genome because of its lack of two-member gene family expansion (Fig A in S1 File). We also confirmed that there was no contaminated sequence in the final assembly (Fig B in S1 File). K-mer frequency of the genomic DNA reads showed a bimodality, indicating the diploidy (Fig C in S1 File).

As of September 2019, genome sizes of sequenced fungi deposited in the NCBI ranged from 2 Mbp to 2.1 Gbp, with an average of 31.0 Mbp (40.7 Mbp for basidiomycetes), and the T. matsutake genome had a relatively large size (Fig 1). In contrast with the size of the genome, the gene-to-genome ratio was comparatively low (81 genes per Mbp). This indicates the presence of many noncoding DNA regions (e.g., repetitive elements). Data concerning genome assembly and predicted genes are summarized in Table 1. The species tree of T. matsutake with 37 Agaricomycetes is shown in Fig 1.

thumbnail
Fig 1. The genome sequencing of Tricholoma matsutake.

a. Genome size vs. gene number of all available fungal genomes in the NCBI. As of September 2019, 5,415 fungal genome assemblies had been deposited, and 1,618 had gene predictions. We used one genome per genus to draw the plot. Tricholoma matsutake is indicated by the arrow. b. Species tree of Tricholoma matsutake and 37 Agaricomycetes genomes. Only bootstrap values less than 100 are marked. The scale bar that represents the mean number of amino acid substitutions per site is shown. The Aspergillus nidulans genome (GenBank: GCF_000149205.2) was used as an outgroup. The branch to the outgroup was shortened for visualization purposes.

https://doi.org/10.1371/journal.pone.0227923.g001

Repeat elements in the T. matsutake genome

The T. matsutake genome had a high content of repeat elements, which has been suggested as a concurrent feature of ectomycorrhizal genomes [2,13,14]. The repeat elements were estimated to have a total size of 92.4 Mbp, representing 48.9% of the entire 189.0 Mbp genome (Fig 2). The major classes of repeat elements were LTR/Gypsy and LTR/Copia transposable elements, corresponding to 41 (21.9%) and 7 Mbp (3.9%) of the genome, respectively. These two frequent elements have also been observed in other basidiomycetes genomes [14,15]. Within the repeat regions, there were 15,014 complete coding sequences (containing a start codon, a stop codon, and no internal stop codon), which was equivalent to a total size of 10.9 Mbp. These were not included in the final predicted genes or further functional annotation. The majority of these sequences (14,857 of 15,014, 99.0%) were transcriptionally repressed (zero Fragments Per Kilobase of transcript per Million mapped reads (FPKMs) at all developmental stages). It is consistent with the previous report in animals that most inserted TEs are dead-on-arrival, and only a few master genes, inserted at specifically fruitful genomic locations, are transcriptionally active [18]. We presumed that meiotic silencing by the unpaired DNA and the quelling process are the potential repression mechanisms [19] because the genome included the genes responsible for those processes (sad-1, sms-3, and sms-2 for meiotic silencing by unpaired DNA and qde-1, dcl2, and qde2 for the quelling process; Fig 3). Another presumed mechanism is the repeat-induced point mutation (RIP). Despite a previous report that states that RIP does not exist in Agaricomycotina genomes [20], we identified the pattern of CpG hypermutations in the genome, although further studies remain to reveal whether the actual RIP process made this pattern (Fig D in S1 File). Although the genome lacked rid/dim2 responsible for the RIP process [21] (Fig 3), its homolog, masc2, existed with two copies. This pattern is also frequent in other basidiomycetes genomes [22]. Experimental validation remains to be done to reveal the exact function of Masc2 in the control of the RIP process. Intergenic region length distribution indicated that many genes were located in the gene-sparse regions mainly because of the presence of enriched TEs (Fig E in S1 File).

thumbnail
Fig 2. Repeat content in the Tricholoma matsutake and the 37 Agaricomycetes genomes.

RepeatModeler and RepeatMasker (http://www.repeatmasker.org) were used sequentially to predict repeat elements in the genomes.

https://doi.org/10.1371/journal.pone.0227923.g002

thumbnail
Fig 3. Transposable-element-silencing genes over diverse fungal genomes.

The orthologs were inferred using OrthoFinder 1.0.6. Dim-2 and Masc2, Sms-3 and Dcl2, and Sms-2 and Qde2 were further differentiated from the gene trees because they belonged to the same gene families. Reference genes are listed in Table A in S2 File.

https://doi.org/10.1371/journal.pone.0227923.g003

Genes surrounded by transposable elements are transcriptionally repressed

It was previously reported that the transcription of TE-surrounded genes is highly repressed in the fungus Pleurotus ostreatus [15]. There were 702 genes that were found to be surrounded by transposable elements using an ad hoc algorithm, described in the Methods section. Among these TE-surrounded genes, the transcripts of 584 genes (83.2%) were not identified at any developmental stage. This was a much higher percentage than the overall percentage of transcriptionally suppressed genes (34.4%). In an attempt to reveal that these suppressed genes were not pseudogenes or wrongly annotated genes, we investigated their homologs and found that 152 of the suppressed genes (26.0%) had paralogous genes that were not surrounded by TEs and were normally expressed in at least one developmental stage (>1 FPKM). Additionally, 290 of the suppressed genes (49.7%) had orthologous genes in at least five of the Agaricomycetes genomes. Although 89 out of 702 silenced genes were annotated with Pfam, the functional bias of silencing was not identified.

Transcriptomic dynamics in the hyphae, primordia, and fruiting body developmental stages

We compared transcriptomic changes between the hyphae, primordia, and fruiting body developmental stages (Fig 4). Of 15,305 predicted genes, 10,046 (65.6%) genes were transcribed in at least one condition (>1 FPKM), and 5,259 genes (34.4%) were not observed in any developmental stage. The majority of the unexpressed genes (4,976 genes, 94.6%) were annotated as hypothetical proteins that lacked known functional domains. On the contrary, 355 genes were constantly expressed during development, where the genes belonged to the top 1000 highest FPKM genes in all developmental stages. These were mostly housekeeping genes, including ribosomal proteins, heat shock proteins, cytochrome, transporters, and ATP synthases. In the hyphae, primordia, and fruiting body developmental stages, 2382, 765, and 884 genes were overexpressed over the other two stages, respectively (Data A in S3 File).

thumbnail
Fig 4. Three developmental stages of Tricholoma matsutake: Hyphae, primordia, and fruiting body.

a. The Venn diagram depicts the number of expressed genes (>1 FPKM) across the three developmental stages. b. Upregulated and downregulated genes during development. Gene functional categorization was carried out using Gene Ontology Slim http://www.geneontology.org.

https://doi.org/10.1371/journal.pone.0227923.g004

The transition from hyphae to primordia (H-to-P transition) upregulated 2248 genes and downregulated 3195 genes, and the transition from primordia to fruiting body (P-to-F transition) revealed an upregulation of 1754 genes and a downregulation of 1971 genes (Fig 4). In the H-to-P transition, the gene function related to the signal transduction, the GTPase activity, and the nucleic acid binding transcription factor activity were enriched, whereas the genes related to the ribosome, the mitochondrion, and the cofactor metabolic process were downregulated. On the other hand, in the P-to-F transition, the enriched functional categories were chromosome, phosphatase activity, and DNA metabolic process, whereas ribosome, RNA binding, and intracellular genes were suppressed (P < 0.01, estimated by Fisher’s exact test).

Trima_09940 was the most expressed gene in the fruiting body stage (39,299 FPKM). The translated protein had a length of 158 aa and a signal sequence for secretion. In addition, this gene had a diedel domain (PF13164), which is related to the insect immune response [23]. The homologs based on sequence similarity were found in Piloderma croceum (ectomycorrhizal basidiomycete), Sistotremastrum niveocremeum (saprotrophic basidiomycete), and Fusarium mangiferae (plant pathogenic ascomycete). The homolog in Drosophila was also identified with 45.8% identity. Although the biological or molecular function of this gene was unclear, it is thought that it may play an essential role in fruiting body formation.

Transcriptional regulators related to fruiting body formation: Transcription factors, light receptors, and hydrophobins

We identified 370 transcription factor genes using a DNA-binding domain search. These included fst4, fst3, hom1, hom2, bri1, gat1, and c2h2 homologs that play critical roles in mushroom formation [7,24] (Table B in S2 File). Although the five genes (fst4, fst3, hom1, hom2, and gat1) were overexpressed at the primordia and fruiting body developmental stages, as reported in previous studies, the expression levels of bri1 and c2h2 were not significantly changed over the three stages. Among 190 differentially expressed genes of the transcription factor (fold change > 2), 53 genes (27.9%) were overexpressed at both primordia and fruiting body developmental stages (Fig 5). This indicates that these two stages share many regulatory processes that are not shared with the hyphae stage. The transcription factors overexpressed at the primordia and fruiting body stages were classified as helix-turn-helix, basic helix-loop-helix/leucine zipper, and β-scaffold factors with minor groove contacts.

thumbnail
Fig 5. Transcription factor expressions at the three developmental stages.

Differentially expressed genes were determined based on logFC (>1 or <−1), calculated by IsoEM2 and IsoDE2. When two conditions were more expressed than the other but there was no difference between them (−1 < logFC < 1), we assigned this gene as being co-overexpressed in these two conditions. Row-wise Z-scores of fragments per kilobase of transcript per million mapped reads were used.

https://doi.org/10.1371/journal.pone.0227923.g005

Blue-light receptor complex WC1/2 is necessary for mushroom development because its deletion prevents mushroom formation [9]. T. matsutake harbored the blue-light receptor complex WC1/2 (encoded by Trima_13733 and Trima_03536). Although wc1 gene expression was higher in the fruiting body stage, the level of the wc2 gene was enriched in the hyphae and fruiting body (Data A in S3 File) developmental stages.

Hydrophobins have multiple biological roles that include fruiting body formation and host–fungus interaction [25]. A total of eight hydrophobin genes were annotated, which was a relatively small number compared with the other 37 Agaricomycetes genomes (Fig F in S1 File), and they ranged from 0 to 130 (average, 20). All of the hydrophobin genes were differentially expressed in the three developmental stages: four were only overexpressed at the fruiting body stage, two at the hyphae stage, and one at the primordia stage (Fig F in S1 File). A hydrophobin gene (Trima_02415) was overexpressed at both the hyphae and primordia stages. Trima_15224 had the fifth highest FPKM value in the fruiting body stage among all predicted genes. This hydrophobin gene might be involved in fruiting body formation.

Small secreted protein genes are dominantly expressed at hyphae

The small secreted protein genes of ectomycorrhizal fungi are important in symbiotic development [26]. The number of small secreted protein genes in Agaricomycetes genomes ranged from 196 to 1,053, but the ectomycorrhizal genomes had fewer small secreted protein genes, ranging from 289 to 576 (Fig G in S1 File). Among the 445 predicted small secreted protein genes, 251 genes (56.4%) were differentially expressed with at least one comparison, and 96 (21.6%) genes were not expressed (zero FPKM values) in all stages (Fig H in S1 File). Many of the differentially expressed small secreted protein genes were overexpressed at the hyphae stage (87 genes). There were 82 cysteine-rich small secreted protein genes (>3% of cysteine of translated protein sequence), such as fungal specific cysteine-rich protein (PF05730), calcium-binding protein (PF12192), peptidase inhibitor (PF03995), and various carbohydrate-binding modules. Some cysteine-rich and calcium-binding domains are involved in fungal pathogenesis [27,28].

IstB-like domain is conserved in mycorrhizal genomes

To examine the ectomycorrhizae-specific functional domains, we performed the enrichment test over the T. matsutake and 37 Agaricomycetes genomes. Among those 21 Pfam domains enriched in the T. matsutake genome (P < 0.01, estimated by Fisher’s exact test), the IstB-like ATP-binding domain (Pfam: PF01695), which is a putative transposase [29], was highly conserved over taxonomically diverse mycorrhizal species (Fig I in S1 File). Interestingly, in the gene tree, the IstB-like genes from Basidiomycota, including T. matsutake, are closer to the ones from Mucoromycota than those from Ascomycota. This might indicate that the gene transfer between Basidiomycota and Mucoromycota occurred during evolution. It might also indicate that the gene transfer between mycorrhizae from different groups occurred during evolution. Although this domain is frequent in bacteria (http://pfam.xfam.org/family/IstB_IS21), fungal IstB-like domains were found to have low sequence similarity with the bacterial domains (<6% matched length coverages). There was lack of evidence for bacterial sequence insertion, such as the HTH-like domain in the protein sequences or the repeats on the flanks. Additionally, their homologs were conserved across the fungal kingdom. Three were overexpressed in the hyphae stage, and one was in the fruiting body stage.

We examined other outstanding functional domains that were enriched in T. matsutake: cadmium resistance transporter (PF03596), neprosin (PF03080); and carbohydrate-binding domain (PF10645, CBM52 family) (Data B in S3 File). These were mainly distributed over ascomycete genomes, and only a few basidiomycete genomes possessed them (Figs J–L in S1 File). Neprosin, a peptidase that cleaves C-terminal to proline residues under highly acidic conditions [30], is usually found in plant genomes, although few bacterial genomes have it. Among fungi, only the Tricholoma and Laccaria genomes contained neprosin (PF03080) domain. The gene tree of this domain showed the grouping between basidiomycetes and several actinobacteria genes (Fig K in S1 File). This implies that the source of this gene could be the gene transfer from actinobacteria. The CBM52 module is required for septum localization in Schizosaccharomyces pombe binding to β-1,3-glucan [31]. This domain (PF10645) was also contained in the Gymnopus luxurians and Sistotremastrum suecicum genomes with six and one copies, respectively. The cadmium resistance transporter (PF03596) and CBM52 gene trees were mostly consistent with their species phylogeny. Therefore, the origin of this gene family could be the common ancestor of fungi, and deletion events have occurred to make the current gene tree. Further research to verify their exact function remains to be carried out. The functional domains in the T. matsutake and other genomes are summarized in Data B in S3 File.

CAZymes are reduced in the genome and differentially expressed during development

The T. matsutake genome had a reduced number of carbohydrate-active enzymes (CAZymes) compared with other Agaricomycetes genomes. A total of 394 predicted CAZyme genes included 143 glycoside hydrolases (GHs), 33 carbohydrate-binding modules (CBMs), 90 glycosyl transferases (GTs), nine polysaccharide lyases (PLs), 59 carbohydrate esterases (CEs), and 60 auxiliary activities (AAs). Many CAZymes are involved in plant cell wall degradation. For example, some AAs, such as manganese peroxidases, versatile peroxidases, and lignin peroxidases, degrade lignin, and some GHs and CEs degrade cellulose and hemicellulose. CBMs often attach to other CAZyme domains to help them bind to target substrates. The genome had one of the lowest numbers of total CAZymes compared with the other 38 Agaricomycetes. Interestingly, ectomycorrhizal fungi, including T. matsutake, showed similar CAZyme profiles, as shown in Fig 6, except for Piloderma croceum. It has been reported that ectomycorrhizal basidiomycetes have lost major gene families, such as plant-cell-wall-degrading enzymes [13], which includes almost all GH families, especially the GH6 family that enzymatically degrades crystalline cellulose [13,32]. This lack of GH6 family was also observed in the T. matsutake genome. The CAZymes in the T. matsutake and other genomes are summarized in Data C in S3 File. Interestingly, we identified two CAZyme submodules that are uniquely found in the T. matsutake genome, and other ectomycorrhizal genomes lacked CBM16 and CBM52. CBM16 is common in bacteria and binds to glucomannan and kappa-carrageenan [33]. The Trima_13517 with this module has a chitin biosynthesis protein CHS5 domain (PF16892), implying its role in cell wall biosynthesis. CBM52 binds to β-1,3-glucan [31] and is often associated with GH81 (endo-β-1,3-glucanase), although the Trima_00904 with the CBM52 lacked any other known functional domains. Further investigation is needed to shed light on their biological functions.

thumbnail
Fig 6. CAZyme genes in Tricholoma matsutake and 37 Agaricomycetes.

All available CAZyme modules were counted. Scaled values based on row Z-scores were used to fill each cell.

https://doi.org/10.1371/journal.pone.0227923.g006

Different numbers of CAZymes were expressed during development, and 57 at the hyphae stage, 44 at the primordia stage, and 47 at the fruiting body stage were more transcribed than in other conditions (Fig 7). Four of nine PL families were overexpressed in the fruiting body stage. Several of the GH subfamilies, such as GH5, GH17, and GH20, were only, or mostly, expressed in the fruiting body stage (4 of 15 for GH5, 2 of 2 for GH17, and 2 of 2 for GH20). GH5 has a role in the degradation of lignocellulose [32]. Interestingly, 15 of 44 (34%) primordia-activated genes were auxiliary activities (AAs), and 11 of 57 (19%) and 5 of 47 (11%) were overexpressed at the hyphae and fruiting body stages, respectively. The AA families activated at the primordia stage included AA1, AA3, AA7, and AA9. Although many AAs are involved in lignin degradation [34], their functions in primordia development are unknown.

thumbnail
Fig 7. Differentially expressed CAZymes at each of the developmental stages.

The pie graphs depict the number of CAZyme families for each specific expression type. Abbreviations for CAZyme families are as follows: glycoside hydrolase (GH), carbohydrate-binding module (CBM), glycosyl transferase (GT), polysaccharide lyase (PL), carbohydrate esterase (CE), and auxiliary activity (AA). Row-wise Z-scores of fragments per kilobase of transcript per million mapped reads were used.

https://doi.org/10.1371/journal.pone.0227923.g007

Methods and materials

Strain and culture conditions

Tricholoma matsutake KMCC04578 at the primordia and fruiting body phases were harvested from Gachang, located near Daegu, South Korea. The dikaryotic mycelia were isolated from the gills of the fruiting bodies and cultured in potato-dextrose broth (PDB; 4 g/L potato peptone, 20 g/L glucose, and pH 5.6 ± 0.2) for 30 days at 25°C.

Isolation of genomic DNA and total RNA

The genomic DNA was extracted from the mycelium using a cetyl trimethyl ammonium bromide (CTAB)-based fungal DNA isolation protocol [35]. The total RNA was extracted from T. matsutake (mycelium, primordium, and stipe of the fruiting body) using an RNeasy mini isolation kit (Qiagen, Valencia, CA, USA). The samples were ground to a fine powder using a mortar and pestle under liquid nitrogen. The resulting samples were homogenized with 15 mL of buffer RLT containing β-mercaptoethanol. After centrifugation for 10 min at 3,000 g, the upper phase was mixed with 15 mL of 70% EtOH, and the total RNA was isolated using an RNA Binding Spin Column under centrifugation for 5 min at 3,000 g. After two wash steps, the total RNA was extracted using DEPC-treated water. The RNA samples (A260/A280 > 1.8) were collected and subjected to further experiments.

Genome sequencing and genome assembly

Three sequencing libraries were generated for T. matsutake: two Illumina paired-end libraries (500 and 250 bp insert sizes) and an Illumina mate pair library (5 kbp insert size) (Table C in S2 File). Raw reads were quality-controlled by trimming the low-quality bases (<30 Phred quality score) and removing short reads after trimming (<50 bp for MiSeq reads and <30 bp for HiSeq reads) using Trim Galore 0.4.4 (https://www.bioinformatics.babraham.ac.uk/projects/trim_galore/). The mitochondrial genomic reads were removed by aligning all of the reads against the reported T. matsutake mitogenome sequence (GenBank ID: JX985789.1) [36] using Bowtie2 (-k 1—very-sensitive—end-to-end) [37]. As a result, 1.1% of the total reads were removed.

ALLPATHS [38] was used for the assembly using the three Illumina libraries with PLOIDY = 2 option. We identified and removed the four scaffolds derived from vector contamination obtained by the BLASTn search against the UniVec database (https://www.ncbi.nlm.nih.gov/tools/vecscreen/univec/). One scaffold of human DNA contamination was also removed. Additional sequence contaminations were examined by drawing a scatterplot of GC contents and sequence coverages using Blobology [39] with alignment of all scaffolds against the NCBI nt database. Thus, we corroborated that there was no additional sequence contamination in the final assembly.

Gene prediction

FunGAP [16] was used to predict the protein-coding genes in the assembly. The genome assembly and RNA-seq reads from the hyphae stage were inputted into the program. A Laccaria_bicolor gene model was selected for Augustus inside the FunGAP. This generated 17,018 preliminary predicted genes. We manually removed 1,707 transposable element genes, such as retrotransposon gag protein (Pfam: PF03732) and reverse transcriptase (Pfam: PF07727) based on Pfam annotation by InterProScan 5.25–64 [40]. Targeted Pfam domains are listed in Table D in S2 File.

The remaining 15,305 genes were examined for their reliability by RNA-seq reads alignment, functional domain annotation, and ortholog search against relatives. RNA-seq reads from the three developmental stages (hyphae, primordia, and fruiting body) were aligned into the genome assembly using HISAT2 [41], and FPKM values were calculated using IsoEM2 [42]. Only >1 FPKM genes were considered as RNA-seq-supported genes. Pfam domains were annotated by InterProScan 5.25–64, and the genes containing at least one Pfam domain were considered as functional domain-supported genes. An ortholog search was performed by OrthoFinder 1.0.6 [43] using T. matsutake and 37 Agaricomycetes genomes. When a gene belonged to a gene family that contains members from more than five genomes, we considered that the gene was supported by the ortholog search. To check for genome completeness, we used BUSCO v3.0.2 [17], in which the basidiomycota_odb9 database was used. Because the assembled genome was dikaryotic, we estimated how many genes were allelic by comparing the numbers of two-member gene families with the five Agaricales genomes. This was obtained by parsing the OrthoFinder output.

Comparative analysis

We chose 37 Agaricomycetes genomes for comparative analyses (Table E in S2 File). In the NCBI database, as at the time of writing, there are 59 Agaricomycetes genome assemblies with predicted genes. We eliminated three incomplete genomes based on BUSCO calculations (<90% completeness). We also sampled two genomes from each order, excluding Agaricales (to which T. matsutake belongs), to reduce computing time. This yielded 38 genomes as the final targets for comparative analysis. A species tree was built using RAxML 8.1.3 [44] from the concatenated single-copy orthologs obtained by OrthoFinder 1.0.6 [43]. We used -f a -x 12345 -p 12345 -# 100 -m PROTGAMMAWAG options for RAxML. Mafft 7.273 [45] and Gblocks 0.91b [46] were used to align the concatenated sequences and extract the conserved regions.

RNA-sequencing

Illumina RNA sequencing generated 108, 133, and 127 million RNA-seq reads from the hyphae, primordia, and fruiting body stages, respectively (Table F in S2 File). Trim Galore 0.4.4 (https://www.bioinformatics.babraham.ac.uk/projects/trim_galore/) was used for adapter removal, low-quality base trimming (<20 Phred score), and short-reads filtering (<40 bp). The RNA-seq reads were aligned to the genome by HISAT 2.0.5 [41]. Owing to a single replicate of the RNA-sequencing libraries, differentially expressed genes were estimated by IsoEM2 and IsoDE2, performing a bootstrapping-based approach using an accurate expectation-maximization algorithm [42]. We used the—auto-fragment-distrib option for IsoEM2 and the -pval 0.05 (desired P value) option for IsoDE2. These programs generate a logarithm of fold change (logFC) between two conditions for each gene. When the logFC value of a certain gene is less than −1 or greater than 1, we considered that gene to be significantly differentially expressed. We classified all of the genes into seven expression patterns: hyphae-enriched, primordia-enriched, fruiting-body-enriched, hyphae- and primordia-enriched, primordia- and fruiting-body-enriched, hyphae- and fruiting-body-enriched, and not significantly different among the samples. When two conditions were more expressed than the other but there was no difference between them, we assigned this gene as being co-overexpressed in those two conditions.

To obtain upregulated or downregulated genes during development, we accounted for the differentially expressed genes in the hyphae–primordia comparison and the primordia–fruiting body comparison. The differentially expressed genes were functionally classified based on Gene Ontology terms. First, Gene Ontology terms were assigned by running InterProScan 5.25–64 [40] against the PfamA database with the—goterms option. Second, we assigned each Gene Ontology term to a high-level Gene Ontology term (GO slim) by running owltools (http://code.g.,oogle.com/p/owltools/) with the—map2slim and —subset goslim_generic options. Finally, Fisher’s exact enrichment test was carried out on each GO slim using Python scipy.stats.fisher_exact function (https://www.scipy.org/).

Transcription factor annotation

Transcription factor genes were predicted based on the Pfam domain annotation. The transcription factor Pfam domains were obtained from the DBD database (http://www.transcriptionfactor.org) [47] in addition to the three functional domains: ARID/BRIGHT DNA-binding domain (PF01388) and fungal specific transcription factor domains (PF04082 and PF11951). The known transcription factor homologs were identified by a BLASTp search, where the best hit was selected. Their orthology relations were validated using OrthoFinder 1.0.6 [43].

Functional domain annotation

InterProScan 5.25–64 [40] predicted the functional domains from the protein sequences of T. matsutake and 38 Agaricomycetes genomes with Pfam 31.0 [48]. The enriched and depleted functions were estimated by a Fisher’s exact test with the scipy.stats.fisher_exact function of Scipy Python module (https://www.scipy.org/). Four selected functions (PF01695, PF03596, PF03080, and PF10645) were BLASTp-searched against the NCBI nr database, and the 50 top-hit sequences were used to build the gene trees. Mafft 7.273 [45] was used for multiple genome sequence alignment with the—maxiterate 1000—localpair options. FastTree 2.1.10 [49] built the trees with default options.

CAZyme annotation

Carbohydrate-related enzymes were predicted using three different tools: dbCAN HMMs 5.0 [50], a database that uses HMM profiles of known CAZyme sequences; BLASTp, a tool for searching the protein sequences against the CAZyme sequence database; and Pfam 31.0, domains annotated with CAZyme entries. All three tools were run and integrated to make a final CAZyme prediction. We assigned a gene as a CAZyme when more than two of the programs gave the same prediction on a gene.

Small secreted protein gene prediction

Referring to previous works [5153], we combined four extracellular protein prediction programs: SignalP, WoLF PSORT, TargetP, and ProtComp. SignalP 4.1 [54] was run with the default option, and “signal peptide = Y” and “Networks-used = SignalP-noTM” tags were used to obtain the signal peptide-containing protein sequences. WoLF PSORT [55] was used with “OrganismType = fungi,” and the most voted localization was used for each protein. TargetP 1.1 [56] was used with -N option for using non-plant networks, and “Loc = S” was used to obtain the secreted proteins. ProtComp v9 (http://www.softberry.com/berry.phtml) was used with -NODB -NOOL options, and “Integral Prediction of protein location” was used for assigning the protein locations. Four programs predicted 780, 1344, 2173, and 823 proteins as secreted, respectively (Fig G in S1 File). Only 78 proteins were predicted as secreted by all programs. The genes predicted by at least three programs (589 proteins) were considered as preliminary secreted protein. We excluded transmembrane, endoplasmic reticulum, and glycophosphatidylinositol-anchored proteins from the candidates with TMHMM, PS_SCAN, and GPI-SOM programs. TMHMM 2.0 [57] was used with default options, and when a transmembrane helix was located within 70 aa of the N-terminal and other helixes were not identified, we considered this as a non-transmembrane protein, with reference to previous work [58]. PS_SCAN 1.86 [59] was used to scan the endoplasmic-reticulum-targeting proteins (PROSITE: PS00014). GPI-SOM 1.5 [60] was used with default options. This resulted in 1788 transmembrane proteins, 16 endoplasmic-reticulum-targeting proteins, and 1718 glycophosphatidylinositol-anchored proteins. By the removal of these proteins, we obtained 455 proteins as secreted proteins in the T. matsutake genome.

Repeat elements analysis

RepeatModeler and RepeatMasker (http://www.repeatmasker.org) were used sequentially to predict repeat elements in the genomes. RepeatModeler produced 2375 consensus repeat sequences with an average of 951 bp, including 289 LTR, 130 DNA, and 39 LINE elements (Table G in S2 File). Classified repeat sequences produced by RepeatModeler were used as a library for RepeatMasker.

The protein-coding sequences within repeat elements were predicted by running Braker1 [61] on the unmasked assembly. The evidence of repeat-induced point mutations was calculated by following the Amselem et al.’s (2015) method [22]. Briefly, repeat sequences were extracted into a FASTA file using rmOut2Fasta.pl script within the RepeatMasker package. We split the sequences so that one FASTA file would contain one repeat family sequence. C-to-T hypermutation of specific dinucleotides was calculated using Mafft-7.273 [45] and RIPCAL 2.0 [62]. Finally, the dinucleotide biases were calculated by counting the repeat elements with >2 transition/transversion ratio and where more than one-third of the sequences had a dinucleotide hypermutations bias. We considered genes as TE-surrounded when a gene had repeat elements at both upstream and downstream within a distance of 1000 bp. Only >400 bp repeat elements were accounted for because there were so many short fragments (139,210 elements).

The genes responsible for genome defense against TEs were identified using gene family and gene tree analyses. We targeted three mechanisms, including repeat-induced point mutation, meiotic silencing by unpaired DNA, and quelling, and nine reference genes related to these mechanisms were used to find their orthologs in the proteome of T. matsutake and the other genomes (Table A in S2 File). We identified the gene families of BLASTp top hits against the reference genes. The genes of Dim-2 and Masc2, Sms-3 and Dcl2, and Sms-2 and Qde2 belonged to the same gene family. Therefore, we additionally constructed gene trees to distinguish them, and Mafft-7.273 [45] and FastTree [49] were used to build the gene trees.

Conclusion

This study aimed to explore the genome composition of ectomycorrhizal Tricholoma matsutake. The repetitive insertions of the transposable elements made this species harbor a remarkably large genome size. These inserted TEs are occasionally involved in transcriptional suppression of the nearby protein-encoding genes. We identified the evidence of genome defense against TEs by C to T hypermutation with a bias over “CpG” dinucleotides. Developmental transcriptomic dynamics revealed that many transcriptional factors are expressed in the primordia and fruiting body stages, and small secreted proteins are expressed in the hyphae stage. The genome contained less carbohydrate-active enzymes than other ectomycorrhizal fungi. These results will help in understanding how Tricholoma matsutake has developed and maintained its lifestyle.

Data availability

This Whole Genome Shotgun project has been deposited at DDBJ/ENA/GenBank under the accession number PKSN00000000. The version described in this paper is version PKSN02000000. The MycoBank ID of this species is 307044. The data used in this study are available at dx.doi.org/10.6084/m9.figshare.11301098. The authors declare that all other data supporting the findings of this study are available within the article and its Supplementary Information files or are available from the corresponding authors upon request.

Supporting information

Acknowledgments

The authors acknowledge paying open access publication fees from the School of Life Sciences and Biotechnology for the BK21Plus Program, Korea University.

References

  1. 1. Park H, Ka K-H. Spore Dispersion of Tricholoma matsutake at a Pinus densiflora Stand in Korea. Mycobiology. 2010;38: 203–205. pmid:23956655
  2. 2. Kuo A, Kohler A, Martin FM, Grigoriev IV. Expanding genomics of mycorrhizal symbiosis. Front Microbiol. 2014;5: 582. pmid:25408690
  3. 3. Molloy S. ECM fungi and all that JAZz. Nat Rev Microbiol. 2014;12: 459. Available: https://doi.org/10.1038/nrmicro3305 pmid:24931037
  4. 4. Martin F, Kohler A, Murat C, Veneault-Fourrey C, Hibbett DS. Unearthing the roots of ectomycorrhizal symbioses. Nat Rev Microbiol. 2016;14: 760–773. pmid:27795567
  5. 5. Trudell SA, Xu J, Saar I, Justo A, Cifuentes J. North American matsutake: names clarified and a new species described. Mycologia. 2017;109: 379–390. pmid:28609221
  6. 6. Miles PG, Chang ST. Mushrooms: Cultivation, Nutritional Value, Medicinal Effect, and Environmental Impact. CRC Press; 2004. Available: https://books.google.com/books?id=XO4EGzpp1M0C
  7. 7. Ohm RA, de Jong JF, de Bekker C, Wosten HAB, Lugones LG. Transcription factor genes of Schizophyllum commune involved in regulation of mushroom formation. Mol Microbiol. 2011;81: 1433–1445. pmid:21815946
  8. 8. Wessels JGH, De Vries OMH, Asgeirsdottir SA, Schuren FHJ. Hydrophobin Genes Involved in Formation of Aerial Hyphae and Fruit Bodies in Schizophyllum. Plant Cell. 1991;3: 793–799. pmid:12324614
  9. 9. Ohm RA, Aerts D, Wosten HAB, Lugones LG. The blue light receptor complex WC-1/2 of Schizophyllum commune is involved in mushroom formation and protection against phototoxicity. Environ Microbiol. 2013;15: 943–955. pmid:22998561
  10. 10. Nowrousian M. Fungal Genomics. Fungal Genomics, 2nd Edition. 2014. pp. 149–172. https://doi.org/10.1007/978-3-642-45218-5_7
  11. 11. Kidwell MG, Lisch DR. Perspective: transposable elements, parasitic DNA, and genome evolution. Evolution. 2001;55: 1–24. pmid:11263730
  12. 12. Feschotte C. Transposable elements and the evolution of regulatory networks. Nat Rev Genet. 2008;9: 397–405. pmid:18368054
  13. 13. Peter M, Kohler A, Ohm RA, Kuo A, Krützmann J, Morin E, et al. Ectomycorrhizal ecology is imprinted in the genome of the dominant symbiotic fungus Cenococcum geophilum. Nat Commun. 2016;7: 12662. pmid:27601008
  14. 14. Hess J, Skrede I, Wolfe BE, LaButti K, Ohm RA, Grigoriev IV, et al. Transposable element dynamics among asymbiotic and ectomycorrhizal Amanita fungi. Genome Biol Evol. 2014;6: 1564–1578. pmid:24923322
  15. 15. Castanera R, López-Varas L, Borgognone A, LaButti K, Lapidus A, Schmutz J, et al. Transposable Elements versus the Fungal Genome: Impact on Whole-Genome Architecture and Transcriptional Profiles. PLOS Genet. 2016;12: e1006108. Available: https://doi.org/10.1371/journal.pgen.1006108 pmid:27294409
  16. 16. Min B, Grigoriev IV, Choi I-G. FunGAP: Fungal Genome Annotation Pipeline using evidence-based gene model evaluation. Bioinformatics. 2017;33: 2936–2937. pmid:28582481
  17. 17. Simao FA, Waterhouse RM, Ioannidis P, Kriventseva EV, Zdobnov EM. BUSCO: assessing genome assembly and annotation completeness with single-copy orthologs. Bioinformatics. 2015;31: 3210–3212. pmid:26059717
  18. 18. Ludwig A, Rozhdestvensky TS, Kuryshev VY, Schmitz J, Brosius J. An unusual primate locus that attracted two independent Alu insertions and facilitates their transcription. J Mol Biol. 2005;350: 200–214. pmid:15922354
  19. 19. Nakayashiki H. RNA silencing in fungi: Mechanisms and applications. FEBS Lett. 2005;579: 5950–5957. pmid:16137680
  20. 20. Horns F, Petit E, Yockteng R, Hood ME. Patterns of Repeat-Induced Point Mutation in Transposable Elements of Basidiomycete Fungi. Genome Biol Evol. 2012;4: 240–247. pmid:22250128
  21. 21. Freitag M, Williams RL, Kothe GO, Selker EU. A cytosine methyltransferase homologue is essential for repeat-induced point mutation in Neurospora crassa. Proc Natl Acad Sci. 2002;99: 8802–8807. pmid:12072568
  22. 22. Amselem J, Lebrun M-H, Quesneville H. Whole genome comparative analysis of transposable elements provides new insight into mechanisms of their inactivation in fungal genomes. BMC Genomics. 2015;16: 141. pmid:25766680
  23. 23. Coste F, Kemp C, Bobezeau V, Hetru C, Kellenberger C, Imler J-L, et al. Crystal Structure of Diedel, a Marker of the Immune Response of Drosophila melanogaster. PLoS One. 2012;7: 1–8. pmid:22442689
  24. 24. 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: 957–963. pmid:20622885
  25. 25. Bayry J, Aimanianda V, Guijarro JI, Sunde M, Latgé J-P. Hydrophobins—Unique Fungal Proteins. PLOS Pathog. 2012;8: e1002700. Available: https://doi.org/10.1371/journal.ppat.1002700 pmid:22693445
  26. 26. Martin F, Aerts A, Ahrén D, Brun A, Danchin EGJ, Duchaussoy F, et al. The genome of Laccaria bicolor provides insights into mycorrhizal symbiosis. Nature. 2008;452: 88–92. pmid:18322534
  27. 27. Kulkarni RD, Kelkar HS, Dean RA. An eight-cysteine-containing CFEM domain unique to a group of fungal membrane proteins. Trends Biochem Sci. 2003;28: 118–121. pmid:12633989
  28. 28. Beck MR, Dekoster GT, Cistola DP, Goldman WE. NMR structure of a fungal virulence factor reveals structural homology with mammalian saposin B. Mol Microbiol. 2009;72: 344–353. pmid:19298372
  29. 29. Yeo CC, Poh CL. Characterization of IS1474, an insertion sequence of the IS21 family isolated from Pseudomonas alcaligenes NCIB 9867. FEMS Microbiol Lett. 1997;149: 257–263. pmid:9141667
  30. 30. Schrader CU, Lee L, Rey M, Sarpe V, Man P, Sharma S, et al. Neprosin, a Selective Prolyl Endoprotease for Bottom-up Proteomics and Histone Mapping. Mol Cell Proteomics. 2017;16: 1162–1171. pmid:28404794
  31. 31. Martin-Cuadrado AB, Encinar del Dedo J, de Medina-Redondo M, Fontaine T, del Rey F, Latge JP, et al. The Schizosaccharomyces pombe endo-1,3-beta-glucanase Eng1 contains a novel carbohydrate binding module required for septum localization. Mol Microbiol. 2008;69: 188–200. pmid:18466295
  32. 32. Zhao Z, Liu H, Wang C, Xu J-R. Comparative analysis of fungal genomes reveals different plant cell wall degrading capacity in fungi. BMC Genomics. 2013;14: 274. pmid:23617724
  33. 33. Bae B, Ohene-Adjei S, Kocherginskaya S, Mackie RI, Spies MA, Cann IKO, et al. Molecular basis for the selectivity and specificity of ligand recognition by the family 16 carbohydrate-binding modules from Thermoanaerobacterium polysaccharolyticum ManA. J Biol Chem. 2008;283: 12415–12425. pmid:18025086
  34. 34. Levasseur A, Drula E, Lombard V, Coutinho PM, Henrissat B. Expansion of the enzymatic repertoire of the CAZy database to integrate auxiliary redox enzymes. Biotechnol Biofuels. 2013;6: 41. pmid:23514094
  35. 35. Fulton TM, Chunwongse J, Tanksley SD. Microprep protocol for extraction of DNA from tomato and other herbaceous plants. Plant Mol Biol Report. 1995;13: 207–209.
  36. 36. Yoon H, Kong W-S, Kim YJ, Kim J-G. Complete mitochondrial genome of the ectomycorrhizal fungus Tricholoma matsutake. Mitochondrial DNA Part A, DNA mapping, Seq Anal. 2016;27: 3855–3857. pmid:25208172
  37. 37. Langmead B, Salzberg SL. Fast gapped-read alignment with Bowtie 2. Nat Methods. 2012;9: 357. Available: https://doi.org/10.1038/nmeth.1923 pmid:22388286
  38. 38. Butler J, MacCallum I, Kleber M, Shlyakhter IA, Belmonte MK, Lander ES, et al. ALLPATHS: de novo assembly of whole-genome shotgun microreads. Genome Res. 2008;18: 810–820. pmid:18340039
  39. 39. Kumar S, Jones M, Koutsovoulos G, Clarke M, Blaxter M. Blobology: exploring raw genome data for contaminants, symbionts and parasites using taxon-annotated GC-coverage plots. Front Genet. 2013;4: 237. pmid:24348509
  40. 40. Jones P, Binns D, Chang H-Y, Fraser M, Li W, McAnulla C, et al. InterProScan 5: genome-scale protein function classification. Bioinformatics. 2014;30: 1236–1240. pmid:24451626
  41. 41. Pertea M, Kim D, Pertea GM, Leek JT, Salzberg SL. Transcript-level expression analysis of RNA-seq experiments with HISAT, StringTie and Ballgown. Nat Protoc. 2016;11: 1650–1667. pmid:27560171
  42. 42. Mandric I, Temate-Tiagueu Y, Shcheglova T, Al Seesi S, Zelikovsky A, Mandoiu II. Fast bootstrapping-based estimation of confidence intervals of expression levels and differential expression from RNA-Seq data. Bioinformatics. 2017;33: 3302–3304. pmid:28605502
  43. 43. Emms DM, Kelly S. OrthoFinder: solving fundamental biases in whole genome comparisons dramatically improves orthogroup inference accuracy. Genome Biol. 2015;16: 157. pmid:26243257
  44. 44. Stamatakis A. RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics. 2014;30: 1312–1313. pmid:24451623
  45. 45. Katoh K, Standley DM. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol Biol Evol. 2013;30: 772–780. pmid:23329690
  46. 46. Talavera G, Castresana J. Improvement of phylogenies after removing divergent and ambiguously aligned blocks from protein sequence alignments. Syst Biol. 2007;56: 564–577. pmid:17654362
  47. 47. Kummerfeld SK, Teichmann SA. DBD: a transcription factor prediction database. Nucleic Acids Res. 2006;34: D74–81. pmid:16381970
  48. 48. Finn RD, Coggill P, Eberhardt RY, Eddy SR, Mistry J, Mitchell AL, et al. The Pfam protein families database: towards a more sustainable future. Nucleic Acids Res. 2016;44: D279–85. pmid:26673716
  49. 49. Price MN, Dehal PS, Arkin AP. FastTree 2–Approximately Maximum-Likelihood Trees for Large Alignments. PLoS One. 2010;5: e9490. Available: https://doi.org/10.1371/journal.pone.0009490 pmid:20224823
  50. 50. Yin Y, Mao X, Yang J, Chen X, Mao F, Xu Y. dbCAN: a web resource for automated carbohydrate-active enzyme annotation. Nucleic Acids Res. 2012;40: W445–51. pmid:22645317
  51. 51. Kim K-T, Jeon J, Choi J, Cheong K, Song H, Choi G, et al. Kingdom-Wide Analysis of Fungal Small Secreted Proteins (SSPs) Reveals their Potential Role in Host Association. Front Plant Sci. 2016;7: 186. pmid:26925088
  52. 52. Lum G, Min XJ. FunSecKB: the Fungal Secretome KnowledgeBase. Database. 2011;2011. pmid:21300622
  53. 53. Garcia K, Ané J-M. Comparative Analysis of Secretomes from Ectomycorrhizal Fungi with an Emphasis on Small-Secreted Proteins. Front Microbiol. 2016;7: 1734. pmid:27853454
  54. 54. Petersen TN, Brunak S, von Heijne G, Nielsen H. SignalP 4.0: discriminating signal peptides from transmembrane regions. Nat Methods. 2011;8: 785. Available: https://doi.org/10.1038/nmeth.1701 pmid:21959131
  55. 55. Horton P, Park K-J, Obayashi T, Fujita N, Harada H, Adams-Collier CJ, et al. WoLF PSORT: protein localization predictor. Nucleic Acids Res. 2007;35: W585–7. pmid:17517783
  56. 56. Emanuelsson O, Brunak S, von Heijne G, Nielsen H. Locating proteins in the cell using TargetP, SignalP and related tools. Nat Protoc. 2007;2: 953–971. pmid:17446895
  57. 57. Krogh A, Larsson B, von Heijne G, Sonnhammer EL. Predicting transmembrane protein topology with a hidden Markov model: application to complete genomes. J Mol Biol. 2001;305: 567–580. pmid:11152613
  58. 58. Min XJ. Evaluation of computational methods for secreted protein prediction in different eukaryotes. J Proteomics Bioinform. 2010;3: 143–147.
  59. 59. de Castro E, Sigrist CJA, Gattiker A, Bulliard V, Langendijk-Genevaux PS, Gasteiger E, et al. ScanProsite: detection of PROSITE signature matches and ProRule-associated functional and structural residues in proteins. Nucleic Acids Res. 2006;34: W362–5. pmid:16845026
  60. 60. Fankhauser N, Maser P. Identification of GPI anchor attachment signals by a Kohonen self-organizing map. Bioinformatics. 2005;21: 1846–1852. pmid:15691858
  61. 61. Hoff KJ, Lange S, Lomsadze A, Borodovsky M, Stanke M. BRAKER1: Unsupervised RNA-Seq-Based Genome Annotation with GeneMark-ET and AUGUSTUS. Bioinformatics. 2016;32: 767–769. pmid:26559507
  62. 62. Hane JK, Oliver RP. RIPCAL: a tool for alignment-based analysis of repeat-induced point mutations in fungal genomic sequences. BMC Bioinformatics. 2008;9: 478. pmid:19014496