ALL Metrics
-
Views
-
Downloads
Get PDF
Get XML
Cite
Export
Track
Research Article
Revised

A whole genome atlas of 81 Psilocybe genomes as a resource for psilocybin production.

[version 2; peer review: 1 not approved]
PUBLISHED 17 Dec 2021
Author details Author details
OPEN PEER REVIEW
REVIEWER STATUS

Abstract

The Psilocybe genus is well known for the synthesis of valuable psychoactive compounds such as Psilocybin, Psilocin, Baeocystin and Aeruginascin. The ubiquity of Psilocybin synthesis in Psilocybe has been attributed to a horizontal gene transfer mechanism of a ~20Kb gene cluster. A recently published highly contiguous reference genome derived from long read single molecule sequencing has underscored interesting variation in this Psilocybin synthesis gene cluster. This reference genome has also enabled the shotgun sequencing of spores from many Psilocybe strains to better catalog the genomic diversity in the Psilocybin synthesis pathway. Here we present the de novo assembly of 81 Psilocybe genomes compared to the P.envy reference genome. Surprisingly, the genomes of Psilocybe galindoi, Psilocybe tampanensis and Psilocybe azurescens lack sequence coverage over the previously described Psilocybin synthesis pathway but do demonstrate amino acid sequence homology to a less contiguous gene cluster and may illuminate the previously proposed evolution of psilocybin synthesis.

Keywords

Psilocybe cubensis, Genome, Single molecule sequencing, Psilocybin

Revised Amendments from Version 1

We have made several corrections based on the reviewers advice but these are best summarized in the response and review section.
Briefly
1)We corrected the lineage concerns raised by the reviewers. 
2)We addressed the contamination concerns with new supplementary figures of Taxon annotated GC plots.
3)We compared 8 samples to the suggested additional genomes (P.serbica and P.cyanescens)
4)We cleaned up the methods section to read more like a journal article
5)We corrected Figure 1 to have 50 samples instead of 49.
6)We included a Figshare to a Newick file and a link to a dockerized pipeline on GitHub for the additional software tools required for the reviewers requests.

See the authors' detailed response to the review by Bryn T M Dentinger and Alexander Bradshaw

Introduction

Psilocybin has recently been awarded breakthrough drug status by the FDA (Tullis 2021). Psilocybin has also been shown to be a safe alternative to traditional selective serotonin re-uptake inhibitors (SSRIs) for the treatment of depression (Griffiths et al. 2016; Ross et al. 2016; Bogenschutz et al. 2018; Carhart-Harris et al. 2021; Davis et al. 2021). SSRIs have also been shown to be effective at reducing SARs-CoV-2 viral load (Schloer et al. 2020; Creeden et al. 2021; Dechaumes et al. 2021; Schloer et al. 2021; Zimniak et al. 2021). One proposed mechanism of the viral interference is hypothesized to be related to the co-expression of ACE2 and DOPA decarboxylase (DDC) which catalyzes the synthesis of serotonin (Attademo and Bernardini 2021). An alternative mechanism of action of Fluvoxamine (SSRI) on SARs-CoV-2 is through agonism of the sigma-1 receptor (Sukhatme et al. 2021). This receptor is known to be regulated by a close chemical relative of psilocybin: N,N-Dimethyltryptamine (Fontanilla et al. 2009). Fluvoxamine is currently being evaluated in several clinical trials for prophylactic and post infection efficacy for SARs-CoV-2.

These are early studies on the use of SSRIs in SARs-CoV-2 and psilocybin is not involved in these studies. Further work is required to solidify this speculative association. Nevertheless, it is possible psilocybin’s serotonin agonism (or sigma-1 receptor activity) will show promise in COViD treatment and prevention in the future, thus understanding the biological mechanism of synthesis is of increasing importance.

Several mechanisms of psilocybin production have been described in different lineages. Horizontal gene transfer has been thoroughly described by Reynolds et al. and Fricke et al., while convergent evolution has been proposed by Awan (Fricke et al. 2017; Awan 2018; Reynolds et al. 2018). Given the broad geographic distribution of psilocybin-producing mushrooms, these models may not be mutually exclusive.

To address this question, we whole-genome sequenced 81 Psilocybe spp. spore samples to assess the sequence coverage over the previously well-characterized psilocybin synthesis gene cluster and to document the diversity of Psilocybe spores circulating in current marketplaces.

Results

In a whole-genome sequencing survey of 81 Psilocybe spp. genomes we noticed a lack of sequence coverage over the psilocybin gene cluster described by Reynolds et al. for five of the 81 Psilocybin producing genomes (three Psilocybe tampanensis, one Psilocybe galindoi, one Psilocybe azurescens). These genomes also exhibited genome-wide poor read mapping efficiency to the Psilocybe cubensis “P.envy” reference genome, with the exception of the mitochondria and ITS regions. These non-cubensis psilocybin-synthesizing genomes share 99% (945/946), 100% (946/946), and 91% (885/971) ITS sequence identity to a 946 base pair Psilocybe tampanensis sequence reported by Rockefeller et al. (NCBI accession number: MH220315.1). The sequence from Rockefeller et al. was obtained in the wild and verified with photography of morphological features. A longer (1994 bases) ITS accession reported by Wesselink et al. (NCBI accession number: HM035077.1) for Psilocybe mexicana had 99% (1992/1994), 99% (1993/1994), 95% (1918/2022) identity to Psilocybe tampanensis, Psilocybe galindoi, and Psilocybe azurescens assemblies respectively.

The sequences described in this study were derived from spores that can be legally sold for taxonomy purposes but cannot be legally cultured to obtain morphological or chemical verification.

One other genus was labelled as Panaeolus copelandia but delivered two 100% identical ITS sequences to Psilocybe cubensis and Aspergillus fumigatus. This samples was omitted from further analysis. Sequencing and variant calling statistics are displayed in Figure 1 and Supplement Table 1.

64e06e9d-1128-4c4b-8946-fff4a94326f5_figure1a.gif64e06e9d-1128-4c4b-8946-fff4a94326f5_figure1b.gif64e06e9d-1128-4c4b-8946-fff4a94326f5_figure1c.gif64e06e9d-1128-4c4b-8946-fff4a94326f5_figure1d.gif64e06e9d-1128-4c4b-8946-fff4a94326f5_figure1e.gif64e06e9d-1128-4c4b-8946-fff4a94326f5_figure1f.gif64e06e9d-1128-4c4b-8946-fff4a94326f5_figure1g.gif64e06e9d-1128-4c4b-8946-fff4a94326f5_figure1h.gif64e06e9d-1128-4c4b-8946-fff4a94326f5_figure1i.gif

Figure 1.

(Left) Photographs of strains were obtained from the vendors public websites. (Middle Left). Sequence coverage of all contigs larger than 1 kb where coverage is on the Log Y axis and contigs sorted by length (largest to smallest: left to right) are on the X axis. (Middle Right) Assembly statistics from Quast 5.0. (Right Top) Number of Heterozygous and Homozygous SNPs detected with reads mapped to the P. cubensis “P.envy” reference. (Right Bottom) Number of Reads for each run according to Samtools flagstat.

Despite the low read mapping rates for Psilocybe tampanensis, Psilocybe galindoi and Psilocybe azurescens, the de novo assemblies of these genomes have high BUSCO completeness scores (93%) and matching ITS sequences suggesting they are indeed Psilocybe but differ enough at the species level to produce low cross species read mapping rates.

An alternative hypothesis is that these libraries are metagenomic or contaminated with non-Psilocybe spp. fungi but still provide enough sequence coverage of the ITS region due to its high copy number. Contaminated libraries are usually detectable with bimodal sequencing coverage as its rare for the organisms to be equimolar. All five non-P. cubensis species have very uniform sequence coverage across the contigs and high BUSCO completion scores with strong ITS sequence implying clean assemblies. Additionally, contigs with mean coverage also contain tBLASTn alignments to PsiM, PsiK, and PsiH suggesting the mean coverage for the assembly contains genes associated with Psilocybe. We also analyzed these assemblies with blobtools to assess levels of metagenomic contamination (Supplementary Figures 1-4) (Challis et al. 2020).

In combination, these data demonstrate close relatedness of Psilocybe tampanensis and Psilocybe galindoi but distant relatedness to the Psilocybe cubensis “P.envy” reference genome. Given the known psilocybin production in these alternative Psilocybe species, these data also imply an alternative or less clustered gene cluster must exist in these species.

Sequence coverage analysis

Uniform sequence coverage and high read mapping rates were observed over most of the of P. cubensis “P.envy” genome for most of the strains (50 of the 81 are displayed in Figure 2 and Figure 3). A few notable higher coverage exceptions were the ITS region on Scaffold_9 and the mitochondrial genome represented by Scaffold_26. This is not a surprising result given the increased copy number of ITS and mitochondrial sequences in fungal genomes.

64e06e9d-1128-4c4b-8946-fff4a94326f5_figure2.gif

Figure 2. Read mapping efficiency of species specific reads mapped to the P. cubensis “P.envy” reference genome.

Due to size limitations, only 50 genomes are presented. The remaining read mapping results can be seen in the Figure 1.

64e06e9d-1128-4c4b-8946-fff4a94326f5_figure3.gif

Figure 3. Coverage plot of 50 genomes mapped to P. cubensis “P.envy” reference. Coverage (log) on the Y axis and scaffolds 1-32 on the X axis.

Scaffolds labeled with a T have a single ~200 bp terminal telomeric sequence. Scaffolds with TT have telomeric sequence on both ends and are assumed to be tip to tip telomeric chromosomes. Scaffolds are ordered from largest to smallest from left to right. P. tampanensis, P. galidoi and P. azurascens (OR-Coast) has low mapping frequency due to divergence.

A few smaller, repeat-rich contigs (scaffolds 27-31) demonstrated more variable coverage across the Psilocybe genus. The ITS region on Scaffold_9 appears to be a tandem repeat which is collapsed in the reference sequence. Scaffolds 27, 28, 31, and 32 have telomeric ends and are small (171 kb, 117 kb, 69 kb, 33 kb). While these smaller contigs have variable coverage in Psilocybe spp., it is important to understand that they represent less than 2% of the genome. Scaffolds 29 and 30 lack telomeric sequence and are 74 kb and 69 kb respectively. Three scaffolds have telomeric sequences on both ends. (Scaffold_2: 4.3 Mb, Scaffold_3: 4.1 Mb; Scaffold_9: 2.3 Mb) suggesting these are tip-to-tip chromosome assemblies.

The genomes most diverged from the P. cubensis “P.envy” reference genome (P. tampanensis, P. galindoi, P. azurescens) exhibited very low read mapping rates. This is demonstrated by P. azurescens (OR-Coast) in Figure 3. This genome-wide low mapping rate was also observed in the psilocybin production cassette on Scaffold_7 (Figure 4 and Figure 5).

64e06e9d-1128-4c4b-8946-fff4a94326f5_figure4.gif

Figure 4.

Integrated Genome Viewer (IGV) coverage plot over the psilocybin synthesis pathway described by Reynolds et al. Strains Huautla and Mazatapec have complete sequence coverage while P. tampanensis, P. galindoi and P. azurescens are lacking sequencing coverage over most of the genes.

64e06e9d-1128-4c4b-8946-fff4a94326f5_figure5.gif

Figure 5. Top-Heatmap of Illumina un-normalized raw sequence coverage of the psilocybin synthesis pathway across 43 genomes.

Low coverage genomes of the same strain (Costa Rica and Samui) were summed together. Bottom- IGV view of three strains with close homology to the reference genome.

Despite the low read mapping rates to P. cubensis “P.envy”, P. tampanensis assembled into 23 kb N50 genome with a 93% BUSCO completeness score and a 99% identical match to Rockefeller et al. ITS sequence for P. tampanensis. These non-cubensis strains were also mapped against previously deposited Psilocybe genomes in NCBI and JGI (P. cyanescens from Reynolds et al. and P. serbica from Fricke et al.). One can see equally low read mapping rates for P. tampanensis and P. galindoi but P. azurescens has slightly higher mapping rates with these genomes (Table 1).

Table 1. Read mapping rates of P. cubensis (Malabar) and non-cubensis strains to the genome references of P. cubensis (P’envy), P. serbica, and P.cyanescens.

NCBI SRA accessionStrain_nameP-envyPsiser1 serbicaPsicy2 cyanescens
SRX11099695Tampa-1-SW11.3%11.7%8.3%
SRX11099696Tampa-2-SW11.4%11.6%8.1%
SRX11099697Tampa-3-SW10.2%11.0%8.0%
SRX11099711OR-Coast-3-SW (Azurescens)13.0%21.0%63.0%
SRX11343020P-Mexicana-Galindoi-1-Mush11.4%9.6%7.3%
SRX11099683Malabar-1-PS (Cubensis)91.1%14.0%14.3%
SRX11099682Malabar-2-PS (Cubensis)92.5%13.9%14.1%
SRX11099680Malabar-3-SW (Cubensis)92.3%13.6%14.0%

These low read mapping rates are reflective of a highly diverged species at the nucleotide level but to exhaust the exploration of psilocybin production, amino acid level homology was further explored.

We utilized tBLASTn to identify alternative psilocybin-producing enzymes in the reference P. cubensis “P.envy” genome. PsiK exists in the gene cluster on P.envy scaffold_7 but also has a close homolog on P.envy scaffold_1. The copy in the gene cluster on P.envy scaffold_7 has a unique SnpEFF high impact variant at p.Arg173Gly that only exists in P. cubensis “P.envy” (Figure 6B). The copy on P.envy scaffold_1 is missing the 5′ exon in P. tampanensis, P. galindoi and P. azurenscens, and is significantly diverged. The most closely related P.envy scaffold_7 PsiK amino acid sequences are seen in a multiple sequence alignment seen in Figure 6A. Further cloning and expression is required to confirm if this additional copy provides any pathway redundancy. Of interest, these alternative alignments are only partially clustered with PsiM and often PsiM is located in the middle of 80 kb contigs in the absence of other pathway related genes (Table 2). This demonstrates a non-clustered psilocybin synthesis pathway in the non-P. cubensis strains.

64e06e9d-1128-4c4b-8946-fff4a94326f5_figure6a.gif

Figure 6A. Multiple sequence alignment using CLUSTALW of the most closely related PsiK genes in P. tampanensis, P. galindoi and P. azurenscens.

The R173G mutation is also unique to P. cubensis “P.envy” when considering non-cubensis species.

64e06e9d-1128-4c4b-8946-fff4a94326f5_figure6b.gif64e06e9d-1128-4c4b-8946-fff4a94326f5_figure6c.gif64e06e9d-1128-4c4b-8946-fff4a94326f5_figure6d.gif

Figure 6B. Variation map of PsiK, PsiD, PsiM, PsiH, PsiR -Scaffold_7.

Grey is no alignment, Blue is reference allele, Aqua is low impact variant and Red represent High Impact SNPs as according to SNPeff.

Table 2.

tBLASTn search of P. tampanensis, P. galindoi, P.cubensis and P. azurenscens for PsiK, PsiH, PsiD and PsiM reveals a fragmented psilocybin synthesis cluster with PsiM most often being on large but independent contigs disrupting the contiguity of the psilocybin synthesis cluster.

NCBI NameJAIFHE01.1JAILYQ01.1JAILYP01.1JAIFHD01.1SRX10250988
P. azurescensP. tampanensis_1P. tampanensis_2P. mexicanaP. cubensis (GT)
PsiK_ContigJAIFHE010009952.1JAILYQ010013513.1JAILYP010003074.1JAIFHD010015065.1K141_42154
Contig Size4,2549,3228,8364,01676,612
PsiH_ContigJAIFHE010002467.1JAILYQ010013513.1JAILYP010003074.1JAIFHD010015065.1K141_42154
Contig Size40,6819,3228,8364,01676,612
PsiD_ContigJAIFHE010002467.1JAILYQ010012987.1JAILYP010013169.1JAIFHD010017873.1K141_42154
Contig Size40,68144,43544,4353,52876,612
PsiM_ContigJAIFHE010002467.1JAILYQ010005515.1JAILYP010014511.1JAIFHD010002461.1K141_42154
Contig Size40,68181,91182,50788076,612

Interestingly, the previously reported P. cyanescens and P. serbica have these genes clustered in the assembly. The 2 Mb N50 P. serbica assembly also has homologs of the pathway on two scaffolds (Scaffold_17 and Scaffold_135). Neither of these genomes have raw reads in the NCBI SRA to support further analysis.

We constructed a phylogenetic tree (Figure 7) of the various genomes. Many of the genomes share strain names but were acquired from different spore banks (PS versus SW versus ITW versus Mush). These phylogenetically cluster together as expected. Many samples were sequenced multiple times as biological replicates and are represented in the sample nomenclature as Name_replica_vendor.

64e06e9d-1128-4c4b-8946-fff4a94326f5_figure7.gif

Figure 7. Phylogenetic tree of 81 psilocybe strains displayed using iTOL (https://itol.embl.de).

Panaeolus Copelandia samples has an P. cubensis ITS sequence is likely a mislabeled sample.

Conclusions

Psilocybin synthesis appears to have evolved both a conserved ~20 kb gene cluster seen in many P. cubensis, P. cyanescens, and P. serbica fungi but also a less clustered pathway in regards to a non-contiguous PsiM gene in P. tampanensis, P. galindoi and P. azurescens that still needs further characterization and scrutiny. As these assemblies mature, these fragmented contigs may be scaffolded into proximity but it is unlikely the gene cluster remains confined to 20 kb. Alternative psilocybin production has also been suggested in the cicada-infecting Massospora spp. fungi (Boyce et al. 2019). Taken together, these data underscore the need for further exploration of psilocybin genomics for alternative or redundant synthesis pathways. Given the divergence of the other psilocybin-producing mushrooms, simply mapping reads from other Psilocybe species to the P. cubensis “P.envy” reference genome can be misleading. Until more complete references exist for the P. tampanensis, P. azurescens and P. galindoi genomes, searching for conserved amino acid sequences more tolerant to synonymous DNA mutations will be required. Due to the legal status of culturing Psilocybe, verifiable biobanks do not exist. As a result, these data are reliant on commercial nomenclature systems currently in place and may be further validated as legalization progresses. The genetic variation in the species is substantial and these genomes will help to further correlate genotype to chemotype associations for this nascent field.

Methods

DNA isolation

Spores were obtained from four vendors (sporeworks.com, Premiumspores.com, Mushrooms.com and InoculateTheWorld.com). Spore preparations utilized a modified DNA isolation procedure described in McKernan et al. (McKernan 2021). Briefly 1.4 mL of spores was centrifuged, decanted and resuspended in 200 μL of ddH2O. 25 μL of a thaumatin-like protein was added and incubated at 37°C (Medicinal Genomics part #420206) for 30 minutes. 12.5 μL of MGC lysis buffer was added and incubated at 65°C for 30 minutes with 9 steel beads. Vortexing was performed every 7 minutes. Lysed sample were micro-centrifuged and 200 μL of supernatant was aspirated and added to 250 μL of Medicinal Genomics (MGC) binding buffer (MGC part# 420001) for magnetic bead isolation. The samples were incubated with the MGC magnetic bead mixture for 10 minutes, magnetically separated and washed two times with 70% ethanol. The beads were dried at 37°C for 5 minutes to remove excess ethanol and eluted with 25 μL of ddH2O.

Library construction for whole genome sequencing

Fragmentation

Genomic DNA (gDNA) was quantified with a Qubit (Thermo Fisher Scientific) and normalized to reflect 4–8 ng/μL in 13 μL of TE buffer. Libraries were generated using enzymatic fragmentation with the NEB Ultra II kits (NEB part # E7103). Briefly, 3.5 μL of 5× NEB fragmentation buffer and 1 μL of Ultra II fragmentation enzyme mix are added to 13 μL of DNA. This reaction was tip-mixed 10 times, vortexed, and quickly centrifuged. Fragmentation was performed in a BioRad CFX96 thermocycler at 3.5 minutes at 37°C, 30 minutes at 65°C. The reaction was kept on ice until ready for adaptor ligation.

Adaptor ligation

The master mix for ligation was prepared on ice using 0.75 μL of Agilent SureSelect Adaptor Oligo Mix, 0.5 μL of ddH2O, 15 μL of New England Biolabs (NEB) Ultra II Ligation Master Mix, 0.5 μL of NEB Ligation enhancer for a total reaction volume of 16.75 μL.

Ligation was performed by the addition of 16.75 μL of ligation master mix to the 17.5 μL Fragmentation/End Prep DNA reaction mixture. This mixture was incubated for 15 minutes at 20°C. Excess adaptors and adaptor dimers were purified with AMPure XP beads (Beckman Coulter #A63881) using 16 μl (approximately 0.45×) of resuspended AMPure XP beads added to the ligation reactions. Samples were tip-mixed 10 times and incubated for 5 minutes at 25°C. The PCR plate was magnetically separated (Medicinal Genomics #420202). After the solution is clear (about 5 minutes), carefully remove and discard the superna, and the supernatant discarded. The magnetic beads were washed by adding 200 μL of 70% ethanol to the PCR plate while the plate remained in the magnetic stand. Ethanol wash was repeated once for a total of 2 washes. The beads are dried at room temperature for ~7 minutes while the PCR plate is on the magnetic stand with the lid open. DNA was eluted from the beads with 10 μL of H2O, and 9 μL of purified DNA is added to a fresh well for further amplification.

PCR amplification

12.5 μL 2× NEBNext Q5 Hot Start Master Mix (New England Biolabs #M0492S) is added to 9 μL ligated DNA from previous step. 3.5 μL NEB 8-bp index primer/universal primer is added to DNA barcode the sample. PCR was performed using the cycling program 98°C for 30 seconds as an initial denaturization step. 6 cycles of denaturization, annealing and extension were performed cycling between 98°C for 10 seconds and 65°C for 75 seconds. A final 65°C for 5 minutes was performed with a final 4°C forever step.

PCR reaction cleanup

Add 15 μL of resuspended AMPure XP beads to the PCR reactions (~25 μL). Mix well by tip mixing 10 times. Incubate the mixture for 5 minutes at room temperature. The plate was magnetically separated for 5 minutes and the supernatant discarded. The plate was washed twice with 200 μL of 70% ethanol while the plate remained on the magnetic stand. The ethanol wash was incubated at room temperature for 30 seconds, and the supernatant carefully removed. Beads were dried at room temperature for 7 minutes while the PCR plate remained on the magnetic stand with the lid open. Beads were eluted 15 μL of nuclease free H2O and transfered into a fresh well.

Sample quality control and sequencing

Libraries were evaluated on an Agilent Tape Station prior to pooling for Illumina sequencing. Sequencing was performed by GeneWiz, Cambridge MA. A total of 690 million paired reads (2×150 bp) were generated, averaging over 12 million read pairs per sample and a total sequence of 207 Gb.

DNA assembly

Reads were assembled with MegaHit v.1.2.9 (https://academic.oup.com/bioinformatics/article/31/10/1674/177884) (Li et al. 2015; Li et al. 2016). The Nextflow mapping and assembly pipeline is published on GitHub. Quast 5.0 (http://quast.sourceforge.net) was used to calculate the assembly quality statistics (Gurevich et al. 2013). Sequencing data is deposited in NCBI under Project ID PRJNA700437 and PRJNA687911.

Read mapping parameters

Reads were mapped to the P. cubensis “P.envy” reference and to their own assemblies to generate BAM files and coverage statistics using bwa-mem (version 0.7.17-r1188).

bwa mem -R "@RG\\tID:$id_run\\tPU:$id_run\\tSM:$id_run\\tLB:$id_run\\tPL:illumina" \
       -t $cpu -M $fasta_ref ${fq[0]} ${fq[1]}|\
   samtools view -hu - \
       |sambamba sort --tmpdir=. /dev/stdin -o ${id_sample}.bam

A workflow for this is deposited in github. https://github.com/mclaugsf/mgc-public/blob/master/AOAC_TYM_ERV/bwa-pair.nf

Variant analysis

Illumina whole-genome shotgun data (McKernan et al. NCBI Project: PRJNA687911) was mapped to the P. cubensis “P.envy” HiFi reference assembly using bwa-mem (version 0.7.17-r1188), samtools (version 1.8), sorted with sambamba (version 0.6.7) and variants were identified using bcftools (v1.10.2) (Danecek et al. 2021).

bcftools mpileup -Ou -f $(Hashemi et al.) ${bam}|\
    bcftools call -m -Ov|\
    bgzip -c > ${sample}-gvcf.vcf.gz

The VCF files were merged using bcftools and left-aligned to split up multi-allelic entries

bcftools merge *gvcf.vcf.gz --output-type v > merged.gvcf.gz
bcftools norm -f/NGS/genewiz-june1st-
10spores/ref/fungus/Psilocybe_cubensis_Envy_scaffolds.fa -m-any merged-norm.vcf.gz

The annotation from the funannotate pipeline was converted from gff3 format into SnpEff (v4.3t. 2017-11-24) database as described here (https://pcingola.github.io/SnpEff/se_buildingdb/)

Phylogenetic analysis

Phylogentic analysis utilized APE (v5.3.) and SNPRelate (v. 1.20.1) (Zheng et al. 2012; Paradis and Schliep 2019).

Coverage analysis

Coverage analysis was performed using BAM files mapped to the P. cubensis “P.envy” genome reference using bwa-mem as described above (Li 2018). BAM files were analyzed using pileup.sh from the BBmap version 4/30/2020 (http://bib.irb.hr/datoteka/773708.Josip_Maric_diplomski.pdf). The Heatmap was generated using Heatmap.2 from the gplots version 3.1.1 in R Studio (https://www.rdocumentation.org/packages/gplots/versions/3.1.1/topics/heatmap.2).

Blobtools

A dockerized pipeline for blobtools was constructed and deposited in github. https://github.com/mclaugsf/mgc-public/tree/master/fungus/blobtools

Data availability

Underlying data

Zenodo: Underlying data for ‘A whole genome atlas of 81 Psilocybe spp. genomes as a resource for psilocybin production’. https://doi.org/10.5281/zenodo.5062843 (McKernan et al. 2021).

The project contains the following underlying data:

Accession number: NCBI, Project ID PRJNA700437 and PRJNA687911

Root URL: https://www.ncbi.nlm.nih.gov/search/all/?term=PRJNA700437

Accession number URL: https://www.ncbi.nlm.nih.gov/search/all/?term=PRJNA687911

Newick file for Figure 7: https://figshare.com/articles/dataset/For_F1000_merged-norm-with-strain-names_newick/17108450

Supplementary Figures 1-4: https://figshare.com/articles/figure/Supplementary_Figure_1-4_F1000_McKernan/17131481

Data are available under the terms of the Creative Commons Attribution 4.0 International license (CC-BY 4.0).

Comments on this article Comments (0)

Version 2
VERSION 2 PUBLISHED 23 Sep 2021
Comment
Author details Author details
Competing interests
Grant information
Copyright
Download
 
Export To
metrics
Views Downloads
F1000Research - -
PubMed Central
Data from PMC are received and updated monthly.
- -
Citations
CITE
how to cite this article
McKernan K, Kane L, Helbert Y et al. A whole genome atlas of 81 Psilocybe genomes as a resource for psilocybin production. [version 2; peer review: 1 not approved] F1000Research 2021, 10:961 (https://doi.org/10.12688/f1000research.55301.2)
NOTE: it is important to ensure the information in square brackets after the title is included in all citations of this article.
track
receive updates on this article
Track an article to receive email alerts on any updates to this article.

Open Peer Review

Current Reviewer Status: ?
Key to Reviewer Statuses VIEW
ApprovedThe paper is scientifically sound in its current form and only minor, if any, improvements are suggested
Approved with reservations A number of small changes, sometimes more significant revisions are required to address specific details and improve the papers academic merit.
Not approvedFundamental flaws in the paper seriously undermine the findings and conclusions
Version 1
VERSION 1
PUBLISHED 23 Sep 2021
Views
102
Cite
Reviewer Report 19 Oct 2021
Bryn T M Dentinger, Natural History Museum of Utah & School of Biological Sciences,, University of Utah, Salt Lake City, UT, USA 
Alexander Bradshaw, Natural History Museum of Utah & School of Biological Sciences, University of Utah, Salt Lake City, UT, USA 
Not Approved
VIEWS 102
The authors present genomic data from 81 samples of putative Psilocybe spp. The source of the samples used for sequencing is problematic in several ways, not the least being clearly documented misidentification and contamination. This reality limits the accuracy of ... Continue reading
CITE
CITE
HOW TO CITE THIS REPORT
Dentinger BTM and Bradshaw A. Reviewer Report For: A whole genome atlas of 81 Psilocybe genomes as a resource for psilocybin production. [version 2; peer review: 1 not approved]. F1000Research 2021, 10:961 (https://doi.org/10.5256/f1000research.58864.r95528)
NOTE: it is important to ensure the information in square brackets after the title is included in all citations of this article.
  • Author Response 17 Dec 2021
    Kevin McKernan, Research and Development, Medicinal Genomics, Beverly, 01915, USA
    17 Dec 2021
    Author Response
    Version 2 Per recommendation of the Moderator/Editor.

    Response to Dentinger and Bradshaw

    We appreciate the many helpful comments from Dentinger and Bradshaw. Many of them are on point ... Continue reading
COMMENTS ON THIS REPORT
  • Author Response 17 Dec 2021
    Kevin McKernan, Research and Development, Medicinal Genomics, Beverly, 01915, USA
    17 Dec 2021
    Author Response
    Version 2 Per recommendation of the Moderator/Editor.

    Response to Dentinger and Bradshaw

    We appreciate the many helpful comments from Dentinger and Bradshaw. Many of them are on point ... Continue reading

Comments on this article Comments (0)

Version 2
VERSION 2 PUBLISHED 23 Sep 2021
Comment
Alongside their report, reviewers assign a status to the article:
Approved - the paper is scientifically sound in its current form and only minor, if any, improvements are suggested
Approved with reservations - A number of small changes, sometimes more significant revisions are required to address specific details and improve the papers academic merit.
Not approved - fundamental flaws in the paper seriously undermine the findings and conclusions
Sign In
If you've forgotten your password, please enter your email address below and we'll send you instructions on how to reset your password.

The email address should be the one you originally registered with F1000.

Email address not valid, please try again

You registered with F1000 via Google, so we cannot reset your password.

To sign in, please click here.

If you still need help with your Google account password, please click here.

You registered with F1000 via Facebook, so we cannot reset your password.

To sign in, please click here.

If you still need help with your Facebook account password, please click here.

Code not correct, please try again
Email us for further assistance.
Server error, please try again.