ArticlePDF Available

Phylogenomics of the psychoactive mushroom genus Psilocybe and evolution of the psilocybin biosynthetic gene cluster

Authors:
  • Fundación Fungi

Abstract

Psychoactive mushrooms in the genus Psilocybe have immense cultural value and have been used for centuries in Mesoamerica. Despite the recent surge of interest in these mushrooms due to the psychotherapeutic potential of their natural alkaloid psilocybin, their phylogeny and taxonomy remain substantially incomplete. Moreover, the recent elucidation of the psilocybin biosynthetic gene cluster is known for only five of ~165 species of Psilocybe , four of which belong to only one of two major clades. We set out to improve the phylogeny of Psilocybe using shotgun sequencing of fungarium specimens, from which we obtained 71 metagenomes including from 23 types, and conducting phylogenomic analysis of 2,983 single-copy gene families to generate a fully supported phylogeny. Molecular clock analysis suggests the stem lineage of Psilocybe arose ~67 mya and diversified ~56 mya. We also show that psilocybin biosynthesis first arose in Psilocybe , with 4 to 5 possible horizontal transfers to other mushrooms between 40 and 9 mya. Moreover, predicted orthologs of the psilocybin biosynthetic genes revealed two distinct gene orders within the biosynthetic gene cluster that corresponds to a deep split within the genus, possibly a signature of two independent acquisitions of the cluster within Psilocybe .
PNAS 2024  Vol. 121  No. 3  e2311245121 https://doi.org/10.1073/pnas.2311245121  1 of 9
RESEARCH ARTICLE 
|
Significance
Therapeutic use of psilocybin from
“magic mushrooms” is
revolutionizing mental health
treatment for many conditions,
including depression, PTSD, and
end- of- life care. However,
knowledge of Psilocybe diversity
and its evolutionary history is
substantially incomplete. Our
study presents the most extensive
phylogenomic dataset across
Psilocybe to date, with 23 samples
derived from type specimens.
Using ~3,000 single- copy gene
families, we recovered a robust
and well- supported phylogeny.
Mapping psilocybin biosynthetic
gene orthologs on the phylogeny
revealed two types of gene cluster
order corresponding to a deep
split in the genus. Molecular
dating suggests psilocybin
biosynthesis arose in Psilocybe ~67
mya, concurrent with the K- Pg
mass extinction event. A
signicant advancement in the
understanding of Psilocybe
evolution and psilocybin
biosynthesis is presented.
Author contributions: A.J.B., G.F., and B.T.M.D. designed
research; A.J.B., V.R.- C., and A.R.A. performed research;
V.R.- C., A.R.A., L.G.- D., and B.T.M.D. contributed new
reagents/analytic tools and fungal specimens; A.J.B.,
V.R.- C., and A.R.A. analyzed data; and A.J.B., V.R.- C.,
L.G.- D., and B.T.M.D. wrote the paper.
The authors declare no competing interest.
This article is a PNAS Direct Submission.
Copyright © 2024 the Author(s). Published by PNAS.
This article is distributed under Creative Commons
Attribution- NonCommercial- NoDerivatives License 4.0
(CC BY- NC- ND).
1A.J.B. and V.R.- C. contributed equally to this work.
2To whom correspondence may be addressed. Email:
alexander.j.bradshaw@gmail.com or bryn.dentinger@
gmail.com.
This article contains supporting information online at
https://www.pnas.org/lookup/suppl/doi:10.1073/pnas.
2311245121/- /DCSupplemental.
Published January 9, 2024.
EVOLUTION
Phylogenomics of the psychoactive mushroom genus Psilocybe
and evolution of the psilocybin biosynthetic gene cluster
AlexanderJ.Bradshawa,b,1,2 , VirginiaRamírez- Cruzc,1 , AliR.Awand, GiulianaFurcie, LauraGuzmán- Dávalosf, and BrynT.M.Dentingera,b,2
Edited by Pamela Soltis, University of Florida, Gainesville, FL; received July 20, 2023; accepted November 28, 2023
Psychoactive mushrooms in the genus Psilocybe have immense cultural value and have
been used for centuries in Mesoamerica. Despite the recent surge of interest in these
mushrooms due to the psychotherapeutic potential of their natural alkaloid psilocybin,
their phylogeny and taxonomy remain substantially incomplete. Moreover, the recent
elucidation of the psilocybin biosynthetic gene cluster is known for only five of ~165
species of Psilocybe, four of which belong to only one of two major clades. We set out to
improve the phylogeny of Psilocybe using shotgun sequencing of fungarium specimens,
from which we obtained 71 metagenomes including from 23 types, and conducting
phylogenomic analysis of 2,983 single- copy gene families to generate a fully supported
phylogeny. Molecular clock analysis suggests the stem lineage of Psilocybe arose ~67
mya and diversified ~56 mya. We also show that psilocybin biosynthesis first arose in
Psilocybe, with 4 to 5 possible horizontal transfers to other mushrooms between 40 and
9 mya. Moreover, predicted orthologs of the psilocybin biosynthetic genes revealed two
distinct gene orders within the biosynthetic gene cluster that corresponds to a deep split
within the genus, possibly a signature of two independent acquisitions of the cluster
within Psilocybe.
phylogenomics | psychoactive | mushroom | Psilocybe | evolution
Psilocybin, the psychoactive natural product found in some mushrooms and other Fungi,
is at the forefront of a wave of recent research showing the tremendous potential of psy-
chedelic drugs for a wide range of mental health therapies and understanding human
consciousness (1–3). Although species of Psilocybe (Fr.) P. Kumm. (rst described as
Agaricus semilanceatus Fr.) have been known to the scientic community since the early
19th century (4), Psilocybe mushrooms have been used across Mesoamerica for centuries,
particularly in spiritual ceremonies (5–13). In many of these indigenous cultures, their
use is considered sacramental or devoted to healing practices. It was for this reason that
with the arrival of European colonialism, in most cases, these cultures experienced religious
persecution for their use of these mushrooms (5). is renewed interest in psilocybin has
shined a spotlight on the organisms that produce them, most of which belong to a single
genus of mushroom, Psilocybe. Colloquially known as “magic mushrooms,” their psycho-
active properties were made widely known in 1957 when Robert Gordon Wasson pub-
lished his personal experiences with them in Life magazine (14). However, since Nixons
1971 “war on drugs” campaign in the United States, psilocybin and its analogs have been
listed on Schedule 1 of the Controlled Substances Act, making possession of the com-
pounds without a special license a federal crime. is legal status stied research on
psilocybin and the organisms that produce it for the last 50+ years (11).
New research in the last six years has yielded surprising insights into the biosynthesis
and evolution of psilocybin. In most Fungi, psilocybin is synthesized by four core enzymes
[PsiD (tryptophan decarboxylase), PsiK (kinase), PsiM (methyltransferase), and PsiH
(P450 monooxegynase)] that convert the amino acid tryptophan into psilocybin and whose
genes occur in a biosynthetic gene cluster (BGC) (15). BGCs constitute close proximity
clustering of genes related to the biosynthesis of secondary metabolites, which are com-
monly used for accessory metabolic functions such as chemical defense or metabolism of
variable carbon sources (16). e close proximity allows for ecient and coordinated
expression of the genes in the cluster, which often serve complementary functions in the
production of a specic secondary metabolite (17). In Fungi, secondary metabolite pro-
duction through BGCs is common in the biosynthesis of many important compounds,
such as melanin and penicillin (18). However, BGCs are not strictly limited to Fungi and
can be found across a wide variety of prokaryotes and eukaryotes. For example, the
well- studied lactose operon (“lac operon”) encodes enzymes directly related to the metab-
olism of lactose as a carbon source and is analogous to a BGC in that genes of the operon
are tightly linked but regulated by a single promoter. Similarly, “pathogenicity islands,”
such as SPI- II in Salmonella spp., contain a multitude of genes that are expressed in a
Downloaded from https://www.pnas.org by Bryn Dentinger on January 9, 2024 from IP address 136.36.244.244.
2 of 9   https://doi.org/10.1073/pnas.2311245121 pnas.org
coordinated and systematic manner, akin to BGC expression, and
directly confer the ability to become pathogenic in numerous hosts
(19, 20). In plants, BGCs are involved in chemical defenses against
insect herbivory, such as in the biosynthesis of morphine and
nepetalactone (21, 22). BGCs have also been found in animals,
such as ancient plant- like terpene BGCs in corals that generate
antifeedant chemicals (23). While secondary metabolites are not
necessarily directly related to primary growth and development,
they provide a system in which an organism can favorably alter
interactions within its environment, making them essential to the
evolutionary ecology of all life (16).
e canonical psilocybin BGC is highly conserved across the
handful of species investigated (24–27) and has been reportedly
acquired by mushrooms outside of Psilocybe through horizontal
gene transfer (HGT) (25) and convergent evolution (24).
However, only ve of the ~165 species of Psilocybe (11) have been
investigated at a genomic- level and only 24 have been included
in multi- locus molecular phylogenetic datasets (10, 28). us, the
tempo, mode, and timing of the evolution of the psilocybin BGC
remains largely unknown, impeding our ability to construct a
robust predictive framework for elucidating evolutionary patterns
and the discovery of novel therapeutics.
In this study, we generated genomic- level data from 71 Psilocybe
taxa using fungarium specimens, including 23 types. Type specimens
are the ultimate authority for applying names to other collections
and are therefore essential references for accurately naming and
describing diversity (29). In recent years, generating DNA sequences
from historical specimens (“museomics”) has beneted from advances
in DNA extraction methods, and the increasing accessibility and
cost reduction of amplicon- independent high- throughput sequenc-
ing (30–33). Historical collections represent hundreds of years of
the collective eort and serve as a signicant concentration of rare
and unique samples (34–36). Outside of a handful of common spe-
cies, Psilocybe spp. are not commonly collected, and many specimens
are only represented by a single collection, making research with
them tremendously dicult (11).
Using the metagenomic data derived from our voucher specimens,
we set out to construct a robust, time- calibrated phylogeny. We then
bioinformatically mined the primary psilocybin BGC and mapped
the core genes on the phylogeny to investigate the tempo, mode, and
patterns of its evolution. We also investigated the timing of the origin
and purported horizontal acquisition events of the psilocybin BGC
by combining publicly available data from all known mushrooms
with the psilocybin BGC to gain insight into the evolutionary forces
that gave rise to psilocybin biosynthesis.
Results
Sequencing and Genome Assembly. Metagenomic shotgun
sequencing was performed on 74 samples, including 71 identied
as Psilocybe and three from other taxa known to produce psilocybin.
ese specimens were all derived from museum vouchers ranging
in age from 3 to 74 y, with one sample missing a collection date
(SIAppendix, TableS1). Total reads for samples ranged from 3,909,488
reads (Psilocybe_mexicana_IBUG- 13593, ~11.7x coverage) to
225,748,454 reads (Psilocybe_baeocystis_WTU- F- 011245, ~ 677.2x
coverage). Across all of our samples, we achieved an average of
94,490,694 reads (~283.5x coverage) and a median of 99,070,482
reads (~297.2x coverage). Genome assembly completeness as
measured by the N50 statistic ranged from N50 = 554 (Psilocybe_
tuberosa_WTU- F- 011378) to N50 = 60,042 (Psilocybe_stuntzii_
WTU- F- 011520). e number of contigs ranged from 7,004
(Psilocybe_stuntzii_WTU- F- 011520) to 784,732 (Psilocybe_
caerulescens_var_mazatecorum_SFSU- F- 029971). BUSCO scores
ranged from 30.7% (Psilocybe_fuliginosa_NY- 1901148) to 95.4%
(Psilocybe_stuntzii_WTU- F- 011520). It should be noted that
these assemblies are likely to be highly discontiguous due to low
endogenous DNA content from museum specimens and should be
treated as metagenome assemblies, but that are nonetheless useful
for phylogenomics (37). A complete summary of sequencing results
and genome assembly statistics is available in SIAppendix, TableS2.
Psilocybe Phylogenomic Analysis and Divergence Time of Major
Clades. Multiple sequence alignment statistics for the Psilocybe-
only dataset revealed 87% (2,615) single- copy genes were recovered
for all samples, 11% (339) recovered between 60 and 61 samples,
and less than 1% (29) recovered fewer than 60 (SI Appendix).
Phylogenetic analysis revealed 52/71 were true Psilocybe (Psilocybe
sensu stricto) (incl. 20 types) (Fig.1) and the other 19 were found
to belong to other genera: 14 specimens correspond to Deconica
(W.G. Sm.) P. Karst. (three types, combined in SI Appendix,
Results and Discussion), three form a monophyletic group of un-
certain placement within Strophariaceae, one corresponded to
Kuehneromyces (Psilocybe_laticystis_UBC- F16759, sister to all of
Deconica), and one (Psilocybe_washingtonensis_WTU- F- 055019)
was unable to be accurately assigned to an existing genus and was
therefore removed from further analysis (SIAppendix, Fig.S1 and
TableS1).
Phylogenetic analysis of the concatenated supermatrix and sum-
mary coalescent analysis of individual gene trees yield topologically
identical phylogenies with strong support for most nodes except
for two nodes internal to species clades with multiple representa-
tives (SI Appendix, Figs. S1 and S2). An ancient divergence within
Psilocybe was recovered, corresponding to Clades I & II in
Ramírez- Cruz et al. (10). e phylogenetic relationships of 39
Psilocybe species are reported (Fig. 2, Left).
e stem age of Psilocybe was estimated at 67.61 mya with the
crown age estimated at 56.43 mya (Fig. 2, Left; see also supple-
mentary data on Dryad). e LTT plot does not deviate from the
null hypothesis of linear growth (Fig. 2, Left). If a pattern of expo-
nential growth were present instead, it would suggest that Psilocybe
has undergone an evolutionary radiation or contraction, which is
not supported by our data.
Incorporating Publicly Available Sequences. Across all publicly
available sequences, 40 species were represented, ~24% of the
approximately 165 currently accepted Psilocybe s.s. (11). However,
sequence number and taxonomic representation of molecular
markers varied widely. No species retrieved from the public
databases had all gene regions representing a single voucher
specimen. We retrieved 26 ITS, 32 Ef1a, 30 RPB1, and 32 RPB2
sequences from public databases representing 11, six, nine, and
three species not represented in the genomic dataset, respectively.
In total, the combined datasets were represented by 67 unique
taxa, 17 of which were only available from public databases
(SIAppendix, Figs.S3–S6).
Misidentication is a common issue in fungaria and Psilocybe
specimens in particular (11, 29). In our study, we identied four
specimens that were consistently incongruent with the other
taxa they clustered with: “Psilocybe acutissima” (GAM00011063),
Psilocybe baeocystis” (WTU- F011245), “Psilocybe quebecensis
(NY1901130), and “Psilocybe silvatica” (VPI- F0003693). ese
specimens were reexamined microscopically and redetermined as
Psilocybe hoogshagenii R. Heim (GAM00011063), Psilocybe cya-
nescens Wakef. (WTU- F01124), and Psilocybe caerulipes (Peck) Sacc.
(NY1901130, VPI- F0003693) (SI Appendix, Table S1).
Downloaded from https://www.pnas.org by Bryn Dentinger on January 9, 2024 from IP address 136.36.244.244.
PNAS 2024  Vol. 121  No. 3  e2311245121 https://doi.org/10.1073/pnas.2311245121  3 of 9
Identication of Psilocybin BGC Genes in Psilocybe. In most
cases, the gene predictions from exonerate matched those from
Augustus/RBB, with disagreements still having correct gene hits.
In a few cases, we found that exonerate identied genes more
accurately and with less ambiguity compared to our RBB method
(SIAppendix, TableS3).
We found variability in the order of the four- core psilocybin-
producing cluster genes within our expanded representation of
Psilocybe species, identifying two distinct patterns. e rst pattern
followed a gene order of PsiD > PsiM > PsiH > PsiK, the canonical
gene order originally reported from Psilocybe cubensis and Psilocybe
serbica M. M. Moser & E. Horak, which was represented entirely
Fig.1. Diversity of Psilocybe sensu stricto. (A) Psilocybe azurescens (image credit, Paul Stamets), (B) Psilocybe cyanescens (image credit Bryn T.M. Dentinger), (C) Psilocybe
semilanceata (image credit Giuliana Furci), (D) Psilocybe zapotecorum (image credit Bryn T.M. Dentinger), (E) Psilocybe cubensis (image credit Oscar Castro- Jauregui),
(F) Psilocybe bohemica (image credit Jan Borovička), (G) Psilocybe yungensis (image credit Virginia Ramírez- Cruz), and (H) Psilocybe mexicana (image credit Oscar Castro- Jauregui).
PsiD PsiKPsiHPsiM
PsiD PsiM PsiK
PsiH PsiH
Psilocybe pleurocystidiosa Isotype(NY-761619)
Psilocybe yungensis(SFSU-F-29944)
Psilocybe heimii Isotype (NY-761610)
Psilocybesemilanceata(SFSU-F-029972)
Psilocybefagicola var. mesocystidiata (NY-761608)
Psilocybe tampanensis(UBC-F10177)
Psilocybecubensisv1.0
Psilocybe weilii Isotype(WTU-F-063525)
Psilocybecolumbiana Isotype (NY-761607)
Psilocybe galindii Isotype (NY-761609)
Psilocybe liniformansvar. americana (NY-1797145)
Psilocybe muliercula(GAM-00011071)
Psilocybezapotecorum(FFCL-689)
Psilocybe pelliculosa (WTU-F-012331)
Psilocybe caerulipes (PUL-00030154)
Psilocybe caerulipes*(VPI-F-0003693)
Psilocybe cyanescens(WTU-F-011306)
Psilocybesamuiensis(WTU-F-055014)
Psilocybe singeri Isotype(NY-761622)
Psilocybesubyungensis Isotype (NY-1197500)
Psilocybe guilartensis Paratype(CFMR-PR-5680)
Psilocybe cyanofibrillosa Isotype(NY-761605)
Psilocybe callosa(NY-1595861)
Psilocybe congolensis Isotype (NY-1652567)
Psilocybeportoricensis(CFMR-PR-4572)
Psilocybe subfimetaria (SFSU-F-029945)
Psilocybe azurescens (UT-M0001775)
Psilocybe hoogshageniivar.convexa Isotype(NY-761612)
Psilocybeaztecorum (SFSU-F-029933)
Psilocybe hoogshagenii* (GAM-00011063)
Psilocybeovoideocystidiata Isotype (XAL-51B)
Psilocybecaerulescens (NY-1920304)
Psilocybeaztecorum var.bonetii(NY-1595856)
Psilocybe cyanescens (NHYD01003446.1)
Psilocybe xalapensis Isotype(NY-761630)
Psilocybe mexicana (IBUG-13593)
Psilocybe bohemica(SFSU-F-029930)
Psilocybe hoogshagenii(SFSU-F-029980)
Psilocybesubcubensis (SFSU-29974)
Psilocybe fimetaria (UBC-F-30923)
Psilocybezapotecorum(GAM-00011076)
Psilocybeangustipleurocystidiata Isotype(NY-761597)
Psilocybesubhoogshagenii(NY-915004)
Psilocybe serbica v1.0
Psilocybe stuntzii (WTU-F-011520)
Psilocybe caerulescens var.mazatecorum (SFSU-F-029971)
Psilocybe cubensis (IBUG-4367)
Psilocybeargentipes(NY-1595850)
Psilocybe cyanescens (KM-189047)
Psilocybewrightii Isotype(NY-761629)
Psilocybearcana Isotype(SFSU-F-000737)
Psilocybe cyanescens*(WTU-F-011245)
PsilocybeazurescensHolotype(WTU-F19095)
Psilocybe hopii Isotype (XAL)
Psilocybemoravica Isotype(SFSU-F-000732)
I
II
Psilocybe caerulipes* (NY-1901130)
A
“C”
“C”
“cubensae”
“mexicanae”
“zapotecorum”
“cordisporae”
D
PsiD
I
II
I
II
PsiH
II
I
Psilocybe angustipleurocystidiata Isotype (NY-761597)|Psi M
Psilocybe xalapensis Isotype (NY-761630)|Psi M
Psilocybe wrightii Isotype (NY-761629)|Psi M
Psilocybe zapatoecorum (GAM00011076)|Psi M
Psilocybe pelliculosa (WTU-F-012331)|Psi M
Psilocybe subcubensis (SFSU-29974)|Psi M
Psilocybe arcana Isotype (SFSU-F-000737)|Psi M
Psilocybe fimetaria (UBC-F30923)|Psi M
Psilocybe columbiana Isotype (NY-761607)|Psi M
Psilocybe weilii Isotype (WTU-F-063525)|Psi M
Psilocybe cyanofibrillosa Isotype (NY-761605)|Psi M
Psilocybe samuiensis (WTU-F-055014)|Psi M
Psilocybe pleurocystidiosa Isotype (NY-761619)|Psi M
Psilocybe argentipes (NY-1595850)|Psi M
Psilocybe aztecorum (SFSU-F-029933)|Psi M
Psilocybe hopii Isotype (XAL) |Psi M
Psilocybe tampanensis (UBC-F10177)|Psi M
Psilocybe
cyanescens
(WTU-F-011245)|Psi M*
Psilocybe semilanceata (SFSU-F-029972)|Psi M
Psilocybe azurescens (UT-M0001775)|Psi M
Psilocybe subhoogshagenii (NY-915004)|Psi M
Psilocybe aztecorum var. bonetii (NY-1595856)|Psi M
Psilocybe callosa (NY-1595861)|Psi M
Psilocybe moravica Isotype (SFSU-F-000732)|Psi M
Psilocybe heimii Isotype (NY-761610)|Psi M
Psilocybe cubensis (IBUG-4367)|Psi M
Psilocybe muliercula (GAM00011071)|Psi M
Psilocybe bohemica (SFSU-F-029930)|Psi M
Psilocybe azurescens Holotype (WTU-F19095)|Psi M
Psilocybe hoogshagenii (SFSU-F-029980)|Psi M
Psilocybe portoricensis (CFMR-PR-4572)|Psi M
Psilocybe fagicola var. mesocystidiata (NY-761608)|Psi M
Psilocybe mexicana (IBUG-13593)|Psi M
Psilocybe subyungensis Isotype NY-1197500|Psi M
Psilocybe caerulipes(VPI-F-0003693)|Psi M*
Psilocybe liniformans var. americana (NY-1797145)|Psi M
Psilocybe yungensis (SFSU-29944)|Psi M
Psilocybe cyanescens (WTU-F-011306)|Psi M
Psilocybe singeri Isotype (NY-761622)|Psi M
Psilocybe caerulipes(NY-1901130)|Psi M*
Psilocybe caerulescens var. mazatecorum (SFSU-F-029971)|Psi M
Psilocybe caerulescens (NY-1920304)|Psi M
Psilocybe subfimetaria (SFSU-F-029945)|Psi M
Psilocybe guilartensis Paratype (CFMR-PR-5680)|Psi M
Psilocybe caerulipes (PUL00030154)|Psi M
Psilocybe galindii Isotype (NY-761609)|Psi M
Psilocybe stuntzii (WTU-F-011520)|Psi M
Psilocybe zapotecorum (FFCL689)|Psi M
Psilocybe congolensis Isotype (NY-1652567)|Psi M
Psilocybe hoogshagenii var. convexa Isotype (NY-761612)|Psi M
Psilocybe ovoideocystidiata Isotype (XAL-51B)|Psi M
Psilocybe hoogshagenii (GAM00011063)|Psi M*
100
80
100
92
78
100
100
100
100
100
100
100
72
100
100
99
100
100
70
36
100
100
100
100
84
100
100
100
100
79
100
100
100
100
100
70
100
100
99
75
100
100
100
100
100
100
100
64
100
63
PsiM
Psilocybe muliercula (GAM00011071)|Psi D
Psilocybe portoricensis (CFMR-PR-4572)|Psi D
Psilocybe hoogshageni (SFSU-F-029980)|Psi D
Psilocybe caerulipes(VPI-F-0003693)|Psi D*
Psilocybe cyanofibrillosa Isotype (NY-761605)|Psi D
Psilocybe azurescens Holotype (WTU-F19095)|Psi D
Psilocybe congolensis Isotype(NY-1652567)|Psi D
Psilocybe heimii Isotype (NY-761610)|Psi D
Psilocybe weilii Isotype (WTU-F-063525)|Psi D
Psilocybe arcana Isotype (SFSU-F-000737)|Psi D
Psilocybe cyanescens (WTU-F-011245)|Psi D*
Psilocybe hoogshagenii (GAM00011063)|Psi D*
Psilocybe aztecorum (SFSU-F-029933)|Psi D
Psilocybe guilartensis Paratype (CFMR-PR-5680)|Psi D
Psilocybe galindii Isotype (NY-761609)|Psi D
Psilocybe stuntzii (WTU-F-011520)|Psi D
Psilocybe azurescens (UT-M0001775)|Psi D
Psilocybe liniformans var. americana (NY-1797145)|Psi D
Psilocybe caerulipes (PUL00030154)|Psi D
Psilocybe angustipleurocystidiata Isotype (NY-761597)|Psi D
Psilocybe caerulipes(NY-1901130)|Psi D*
Psilocybe cubensis (IBUG-4367)|Psi D
Psilocybe cyanescens (WTU-F-011306)|Psi D
Psilocybe ovoideocystidiata Isotype (XAL-51B)|Psi D
Psilocybe mexicana (IBUG-13593)|Psi D
Psilocybe moravica Isotype (SFSU-F-000732)|Psi D
Psilocybe wrightii Isotype (NY-761629)|Psi D
Psilocybe subyungensis Isotype (NY-1197500)|Psi D
Psilocybe hopii Isotype (XAL) |Psi D
Psilocybe fagicola var. mesocystidiata (NY-761608)|Psi D
Psilocybe subcubensis (SFSU-29974)|Psi D
Psilocybe pleurocystidiosa Isotype (NY-761619)|Psi D
Psilocybe yungensis (SFSU-29944) |Psi D
Psilocybe subhoogshagenii (NY-915004)|Psi D
Psilocybe tampanensis UBC-F10177|Psi D
Psilocybe zapotecorum (GAM00011076)|Psi D
Psilocybe hoogshagenii var. convexa Isotype (NY-761612)|Psi D
Psilocybe xalapensis Isotype (NY-761630)|Psi D
Psilocybe aztecorum var. bonetii (NY-1595856)|Psi D
Psilocybe columbiana Isotype (NY-761607)|Psi D
Psilocybe caerulescens var. mazatecorum (SFSU-F-029971)|Psi D
Psilocybe argentipes (NY-1595850)|Psi D
Psilocybe bohemica (SFSU-F-029930)|Psi D
Psilocybe callosa (NY-1595861)|Psi D
Psilocybe fimetaria (UBC-F30923)|Psi D
Psilocybe subfimetaria (SFSU-F-029945)|Psi D
Psilocybe zapotecorum (FFCL689)|Psi D
Psilocybe pelliculosa (WTU-F-012331)|Psi D
Psilocybe singeri Isotype (NY-761622)|Psi D
Psilocybe caerulescens (NY-1920304)|Psi D
Psilocybe samuiensis (WTU-F-055014|Psi D
Psilocybe semilanceata (SFSU-F-029972)|Psi D
100
78
100
100
49
75
99
71
100
100
100
99
100
100
100
100
100
100
72
100
100
100
100
61
100
84
90
100
69
68
100
69
96
89
97
100
100
100
100
42
100
97
100
100
100
100
100
100
96
100
I
II
100
100
78
100
94
100
100
100
55
100
100
54
66
100
100
100
100
64
100
75
100
100
100
100
98
86
98
100
100
96
97
100
100
51
100
100
100
97
100
100
100
100
100
100
100
98
100
100
92
100
100
100
100
100
Psilocybemuliercula(GAM00011071)|Psi H
Psilocybezapotecorum(FFCL689)|Psi H
Psilocybeaztecorumvar. bonetii (NY-1595856)|Psi H
Psilocybe pleurocystidiosa Isotype(NY-761619)|Psi H
Psilocybeaztecorum(SFSU-F-029933)|Psi H
a(IBUG-13593)|Psi H
Psilocybeargentipes(NY-1595850)|Psi H
Psilocybecaerulescens(NY-1920304)|Psi H
Psilocybe
cyanescens
(WTU-F-011245)|Psi H*
Psilocybeangustipleurocystidiata Isotype(NY-761597)|Psi H
10177)|Psi H
is(UBC-F10177)|Psi H
Psilocybe pelliculosa(WTU-F-012331)|Psi H
Psilocybesubhoogshagenii(NY-915004)|Psi H
Psilocybe guilartensis Paratype(CFMR-PR-5680)|Psi H
Psilocybecaerulescensvar.mazatecorum(SFSU-F-029971)|Psi H
Psilocybe heimii Isotype(NY-761610)|Psi H
PsilocybeazurescensHolotype(WTU-F19095)|Psi H
Psilocybeyungensis(SFSU-29944)|Psi H
Psilocybe weilii Isotype(WTU-F-063525)|Psi H
m(GAM00011076)|Psi H
Psilocybestuntzii(WTU-F-011520)|Psi H
Psilocybe liniformansvar.americana(NY-1797145)|Psi H
Psilocybecallosa(NY-1595861)|Psi H
Psilocybecyanescens(WTU-F-011306)|Psi H
Psilocybe ovoideocystidiata Isotype(XAL-51B)|Psi H
Psilocybe hopii Isotype(XAL) |Psi H
Psilocybecongolensis Isotype(NY-1652567)|Psi H
Psilocybe zapotecorum(GAM00011076)|Psi H
scye sage AM00011063)|Psi H*
Psilocybe wrightii Isotype(NY-761629)|Psi H
Psilocybe hoogshagenii(SFSU-F-029980)|Psi H
Psilocybecaerulipes (PUL00030154)|Psi H
Psilocybesemilanceata (SFSU-F-029972)|Psi H
Psilocybecaerulipes(VPI-F-0003693)|Psi H*
13593)|Psi H
Psilocybecolumbiana Isotype(NY-761607)|Psi H
Psilocybexalapensis Isotype(NY-761630)|Psi H
Psilocybe portoricensis(CFMR-PR-4572)|Psi H
-055014)|Psi H
Psilocybeazurescens (UT-M0001775)|Psi H
Psilocybesubfimetaria(SFSU-F-029945)|Psi H
Psilocybecubensis (IBUG-4367)|Psi H
Psilocybesubyungensis Isotype(NY-1197500)|Psi H
Psilocybecyanofibrillosa Isotype (NY-761605)|Psi H
Psilocybefimetaria(UBC-F30923)|Psi H
Psilocybecaerulipes(NY-1901130)|Psi H*
Psilocybecyanescens(WTU-F-011306)|Psi H
Psilocybesubcubensis(SFSU-29974)|Psi H
Psilocybesingeri Isotype(NY-761622)|Psi H
Psilocybe bohemica(SFSU-F-029930)|Psi H
var.convexa Isotype (NY-761612)|Psi H
m(FFCL689)|Psi H
Psilocybeazurescens(UT-M0001775)|Psi H
Psilocybemoravica Isotype(SFSU-F-000732)|Psi H
Psilocybefagicolavar.mesocystidiata(NY-761608)|Psi H
Psilocybea
rcana Isotype(SFSU-F-000737)|Psi H
sotype(NY-71609)|Psi H
Psilocybe caerulescens (NY-1920304)|Psi K
Psilocybe hoogshageni (SFSU-F-029980)|Psi K
Psilocybe singeri Isotype (NY-761622)|Psi K
Psilocybe cyanofibrillosa Isotype (NY-761605)|Psi K
Psilocybe argentipes (NY-1595850)|Psi K
Psilocybe bohemica (SFSU-F-029930)|Psi K
Psilocybe aztecorum var. bonetii (NY-1595856)|Psi K
Psilocybe weilii Isotype (WTU-F-063525)|Psi K
Psilocybe zapotecorum (FFCL689|)Psi K
Psilocybe callosa (NY-1595861)|Psi K
Psilocybe hoogshagenii (GAM00011063)|Psi K*
Psilocybe caerulescens var. mazatecorum (SFSU-F-029971)|Psi K
Psilocybe pelliculosa (WTU-F-012331)|Psi K
Psilocybe samuiensis (WTU-F-055014)|Psi K
Psilocybe portoricensis (CFMR-PR-4572)|Psi K
Psilocybe subyungensis Isotype (NY-1197500)|Psi K
Psilocybe moravica Isotype (SFSU-F-000732)|Psi K
Psilocybe aztecorum (SFSU-F-029933)|Psi K
Psilocybe hoogshagenii var. convexa Isotype (NY-761612)|Psi K
Psilocybe fagicola var. mesocystidiata (NY-761608)|Psi K
Psilocybe hopii Isotype (XAL) |Psi K
Psilocybe fimetaria (UBC-F30923)|Psi K
Psilocybe caerulipes(NY-1901130)|Psi K*
Psilocybe guilartensis Paratype (CFMR-PR-5680)|Psi K
Psilocybe subhoogshagenii (NY-915004)|Psi K
Psilocybe stuntzii (WTU-F-011520)|Psi K
Psilocybe mexicana (IBUG-13593)|Psi K
Psilocybe azurescens Holotype (WTU-F19095)|Psi K
Psilocybe caerulipes(VPI-F-0003693)|Psi K*
Psilocybe arcana Isotype (SFSU-F-000737)|Psi K
Psilocybe cyanescens (WTU-F-011306)|Psi K
Psilocybe congolensis Isotype (NY-1652567)|Psi K
Psilocybe yungensis (SFSU-29944)|Psi K
Psilocybe muliercula (GAM00011071)|Psi K
Psilocybe subfimetaria (SFSU-F-029945)|Psi K
Psilocybe columbiana Isotype (NY-761607)|Psi K
Psilocybe semilanceata (SFSU-F-029972)|Psi K
Psilocybe ovoideocystidiata Isotype (XAL-51B)|Psi K
Psilocybe galindii Isotype (NY-761609)|Psi K
Psilocybe heimii Isotype (NY-761610)|Psi K
Psilocybe angustipleurocystidiata Isotype (NY-761597)|Psi K
Psilocybe subcubensis (SFSU-29974)|Psi K
Psilocybe cyanescens
(WTU-F-011245)|Psi K*
Psilocybe caerulipes (PUL00030154)|Psi K
Psilocybe wrightii Isotype (NY-761629)|Psi K
Psilocybe liniformans var. americana (NY-1797145)|Psi K
Psilocybe tampanensis (UBC-F10177|)Psi K
Psilocybe pleurocystidiosa Isotype (NY-761619)|Psi K
Psilocybe zapotecorum (GAM00011076)|Psi K
Psilocybe xalapensis Isotype NY-761630|Psi K
Psilocybe cubensis (IBUG-4367)|Psi K
Psilocybe azurescens (UT-M0001775)|Psi K
100
47
100
97
100
82
100
73
100
100
80
100
100
94
100
100
100
100
79
100
100
100
100
99
86
82
100
84
100
100
100
56
100
75
100
61
67
100
55
100
100
54
100
100
100
71
96
62
56
90
PsiK
I
II
Agrocybe pediades
Galerina marginata
Gymnopilus chrysopellus
Gymnopilus junonius
outgroup
B
Loremipsum
010203040506070
79
80
99
75
94
96
a
(
I
BU
-13593
)
|P
s
i
H
i
s
(U
B
C
-
F
10
177
)
|
P
s
i
H
y
p
(
)
|
10177
)
|P
s
i
H
1
3593
)
|
P
s
i
H
055014)
|P
s
i
1
0
0
0
0
0
0
P
s
ilo
cy
b
e
a
z
u
resce
n
s
H
o
lo
ty
p
e
(
WT
U
-
F
1
9095
)
|
P
s
i
H
P
s
ilo
cy
0
0
b
e
c
y
a
n
esce
n
s
(
WT
U-
F
-
011306
)
|
P
s
i
H
P
P
s
ilo
il
c
y
b
b
b
e
a
zu
resce
n
s
(U
(U
(U
T
T
T
-
M0001775)
M0001775)
M0001775)
|P
|P
|P
s
i
i
i
H
H
H
1
0
0
1
0
0
P
s
ilo
c
y
b
e
cy
anescens
(
WTU-
F
-011245
)
|P
s
i
H
*
P
s
il
o
cy
b
e
a
zu
resce
n
s
(U
T
-M0001775)|
P
s
i
H
P
s
il
oc
y
b
e
c
ya
n
e
sce
n
s
(
WT
U
-
F
-
011306
)
|P
s
i
H
y
(
)
|
P
s
ilo
cy
b
e
z
a
po
t
e
c
o
r
u
m
(
FF
C
L
6
89
)
|P
s
i
H
P
s
il
o
cy
b
e
z
a
po
t
ec
o
r
u
m
(
G
AM00011076
)
|
P
s
iH
m
(
G
AM00011076)
|P
s
i
H
var
.
c
on
vex
a
I
s
ot
y
p
e
(
N
Y
m
(
F
F
C
L
689
)
|
P
s
i
H
Fig.2.Psilocybe phylogenomic tree and evolution of the psilocybin BCG. Left: Phylogenomic tree of Psilocybe estimated from a gene- partitioned concatenated
supermatrix of 2,983 single- copy gene families estimated in IQ- TREE. Tree is rooted with an outgroup consisting of Hebeloma cylindrosporum and Pholiota alnicola. All
nodes received 100% ultrafast bootstrap support in the concatenated analysis except where indicated. Topology is identical to summary coalescence of individual
gene trees except for the terminal relationships within the P. cyanescens species cluster. Time- transformed branch lengths and divergence dates were estimated
from loci with 100 clock- like, well- supported gene trees in RelTime using two internal calibrations based on the ages for Strophariaceae/Hymenogastraceae and
Gymnopilus, reported by Varga etal. (38) and Ruiz- Dueñas etal. (39). Divergence time 95% HPD condence bars are depicted in gold. Branch labels are percent
nonparametric bootstraps implemented using the within- partition ultrafast bootstrapping option in IQ- TREE. The lineage through time plot is represented by
a purple dashed line. Sequences derived from type specimens are labeled in red. The psilocybin BGC is depicted by arrows to the right of the terminal labels.
Arrows are oriented according to the coding direction for each gene as determined through reciprocal best BLAST hits and color- coded by gene (PsiD = red, PsiK
= green, PsiH = purple, PsiM = gold). Absence of genes indicates that placement in context of the whole cluster could not be reliably determined bioinformatically.
Right: Maximum- likelihood phylograms for each of the four core genes from the psilocybin- producing gene cluster, PsiD (red), PsiK (green), PsiM (gold), and PsiH
(purple). For enhanced readability, trees are midpoint rooted or rooted at the node leading to Clade II (PsiH). Duplicated PsiH genes are highlighted in yellow (full
length) or blue (truncated). The two major clades from the organismal phylogeny are labeled with roman numerals and with shaded boxes in the background.
Numbers on branches are percent nonparametric rapid bootstraps implemented using the within- partition ultrafast bootstrapping option in IQ- TREE.
Downloaded from https://www.pnas.org by Bryn Dentinger on January 9, 2024 from IP address 136.36.244.244.
4 of 9   https://doi.org/10.1073/pnas.2311245121 pnas.org
by Clade II of our dataset. e second was PsiD > PsiK > PsiH >
PsiM, found in Clade I of our organismal phylogeny (Fig. 2,
Right). Across our phylogeny, the original pattern was represented
by 17 Psilocybe specimens, while 35 of our Psilocybe specimens
exhibited the non- canonical pattern.
In addition to gene order and orientation, we investigated
the evolution of each gene separately. Phylogenetic analysis was
performed without using our species constraint tree for each of
the core psilocybin- producing genes, PsiD, PsiK, PsiM, and
PsiH (Fig. 2, Right). We found that PsiD, PsiK, and PsiM were
topologically congruent with the organismal phylogeny, sug-
gesting a vertical inheritance pattern for these genes in Psilocybe.
In contrast, PsiH was largely topologically consistent with the
organismal phylogeny, but multiple taxa in both Clade I and
Clade II were found to have duplicated copies of PsiH that were
recovered as paraphyletic (Fig. 2, Right, and SI Appendix,
Table S3). In all cases, both copies of PsiH are found in tandem
on the same contig (SI Appendix, Table S3). One copy of PsiH
in each of P. zapotecorum, P. mexicana, and P. tampanensis is
truncated, indicating they are not functional. e phylogenetic
topology of the Clade I duplications is consistent with an
ancient duplication followed by gene loss in most taxa, with
relict, incompletely degenerated copies persisting in only a few
lineages.
Psilocybin Homologs Genes Present Outside of Psilocybe and
Ecological Niche Patterns. We also investigated the known
HGT events from Panaeolus (Fr.) Quél., Gymnopilus P. Karst.,
and Pluteus Fr., using publicly available data and specimens
sequenced in this study vouchered as Gymnopilus luteofolius
(Peck) Singer, Pluteus albostipitatus (Dennis) Singer, and Pluteus
salicinus (Pers.) P. Kumm (SI Appendix, TableS1 and Fig.3,
Left). Gene prediction and RBB analysis yielded representation
of the four- core psilocybin- producing genes from the published
genomes of Panaeolus cyanescens and Gymnopilus dilepis (Berk.
& Broome) Singer, and our Gymnopilus luteofolius and Pluteus
salicinus data (SIAppendix, TableS1).
Phylogenetic analysis of the PsiD gene, including those from
our Psilocybe genomic samples, and publicly available sequences,
revealed three branching patterns for our non- Psilocybe taxa. e
rst branch contained Conocybe smithii Watling (= Pholiotina
smithii (Watling) Enderle) as a monophyletic branch sister to our
Psilocybe caerulipes (Peck) Sacc. samples (BS 100%). e second
branch placed Panaeolus cyanescens as a sister group to Psilocybe
cubensis (BS 100%), and the nal contained species of both
Gymnopilus and Pluteus sharing a most recent common ancestor
with all of Psilocybe (BS 100%) (Fig. 3, Left).
After expanding Psilocybe diversity and rening the placement
of known HGT events, we sought to identify whether phyloge-
netic clusters could be attributed to specic ecological niches.
Ancestral state reconstruction was performed on the organismal
phylogeny, showing a wood decay ecology as the ancestral condi-
tion in the most recent common ancestor (MRCA) of Clades I
and II, with two independent specializations to coprophilic life-
styles in each (SI Appendix, Fig. S7). Interestingly, our two primary
phylogenetic clades correspond to distinct ecological lifestyles,
with a few notable exceptions (Fig. 3, Left). Clade I, identied
through our PsiD analysis, corresponds almost entirely to the
soil- dwelling saprotrophic lifestyle, except for the Psilocybe caer-
ulipes cluster, which exhibits wood decay ecology (i.e., rotting
wood rather than soil enriched with woody debris). Psilocybe Clade
II, Gymnopilus, and Pluteus were associated primarily with wood
decay, and in the case of Psilocybe cubensis and Panaeolus cyanescens,
a coprophilous lifestyle (Fig. 3, Left).
Evolutionary Origins and Timing of the Psilocybin BGC in
Mushrooms. Divergence dating is consistent with the earliest
possible origin of the psilocybin BGC in Psilocybe approximately
~67 mya with subsequent acquisitions by four other genera
between ~40 (Panaeolus) and ~9 (Gymnopilus) mya (Fig.3, Right
and SIAppendix). e crown age of Psilocybe was estimated at ~57
mya in the phylogenomic dataset, which predates the estimated
origin of psilocybin in the next possible oldest lineage, Panaeolus.
However, the relative ages of the psilocybin- producing Panaeolus
stem and the crown age of Psilocybe in the LSU dataset suggest
the earliest possible origin could be in Panaeolus. Additionally,
two species of Panaeolus and two species of Gymnopilus known
to contain the psilocybin BGC were paraphyletic in the LSU
phylogeny. e most parsimonious interpretation is that the
psilocybin BGC was gained twice independently in each genus.
e rst acquisition in Panaeolus (Pa.) is in Pa. subalteatus ~40
mya with a second acquisition ~16 mya in Pa. cyanescens. e
rst acquisition in Gymnopilus is in G. aeruginosus ~22 mya with
a second in G. junonius ~9 mya. e psilocybin BGC was gained
in Pluteus ~22 mya and in Pholiotina ~11 mya (SIAppendix). To
further test the notion that these acquisitions were due to HGT,
we performed a topological constraint test. e likelihoods of
the best ML trees for PsiD from topologically constrained and
unconstrained searches were −8780.501249 and −8699.391458,
respectively. Tree topology test statistics universally rejected the
constrained topology, corroborating the results of Reynolds etal.
(25) and consistent with expectations from HGT (SIAppendix,
TableS4).
Discussion
Phylogenomics of Psilocybe. Our phylogenomic analysis resulted
in a single, unambiguous and statistically well- supported
phylogeny for 52 specimens of Psilocybe s.s., resolving key nodes
that were previously unsupported (Fig.2, Left). Total topological
congruence inferred from a partitioned supermatrix analysis and
a summary coalescence of individual gene trees demonstrates that
the inferred phylogeny is robust to the methodological approach
and indicates low levels of gene discordance within Psilocybe.
e ancient divergence within Psilocybe rst reported by
Ramírez- Cruz et al. (10) was also recovered here. is early split
coincides with the genic arrangement of the psilocybin BGC
(Fig. 2, Left) but precedes a shift in ecology. Our results suggest
that Psilocybe arose as primarily a wood- decomposing group that
transitioned to soil after the split, with two independent special-
izations on herbivore dung (SI Appendix, Fig. S7). However, addi-
tional critical species, such as the African dung- dwelling
P. natalensis Gartz, D.A. Reid, M.T. Sm. & Eicker, will need to
be included to reconstruct evolution of ecologies more condently
in Psilocybe (further discussed in SI Appendix).
Origin of the Psilocybin BGC.
Reynolds etal. (25) speculated that
the psilocybin BGC may have originated in Fibularhizoctonia
G.C. Adams & Kropp, a genus of anamorphic (asexually
reproducing) Athelia Pers. that has multiple copies of Psi
homologs, albeit not contained in a cluster (40). e Atheliales
are close relatives of the Agaricales with a broad range of ecologies,
including termite symbionts (41), plant pathogens, mycorrhizal
mutualists, and saprobes. Reynolds etal. (25) speculated that
this insect association may have provided the selective force for
the evolution of psilocybin as a modulator of the symbiosis.
Interestingly, Athelia arachnoidea (Berk.) Jülich is mycoparasitic
on lichens common on tree bark in north temperate regions
(42, 43), whereas its anamorph Fibularhizoctonia carotae (Rader)
Downloaded from https://www.pnas.org by Bryn Dentinger on January 9, 2024 from IP address 136.36.244.244.
PNAS 2024  Vol. 121  No. 3  e2311245121 https://doi.org/10.1073/pnas.2311245121  5 of 9
G.C. Adams & Kropp is pathogenic on carrot roots (44). Our
results suggest wood decomposition as the ancestral ecology
of Psilocybe. Both Gymnopilus and Pluteus are also wood
decomposers. e mycoparasitic ecology of the teleomorphic
(sexually reproducing) Athelia arachnoidea on bark is intriguing
because the habitat is shared with wood- dwelling mushrooms
and physical interaction may be one way the psilocybin BGC
could be transferred horizontally. However, the true vector of
these HGTs is unknown.
Tempo, Mode, and Timing of the Evolution of the Psilocybin BGC.
Previous studies have shown that the psilocybin BGC has been
transferred horizontally between several genera of mushrooms
within the order Agaricales (24, 25). Our analysis combined these
data with a much greater representation of Psilocybe, recovering the
PsiD genes in Gymnopilus and Pluteus outside of those in Psilocybe.
In contrast, the copies in Pholiotina and Panaeolus are nested
within the Psilocybe PsiD genes. is topology is consistent with
ancient divergence between the Psilocybe and Gymnopilus/Pluteus
psilocybin BGCs and indicates more recent HGT events between
Psilocybe and both Panaeolus and Pholiotina. Panaeolus cyanescens
appears to have a HGT event with the ancestor of Psilocybe
cubensis/subcubensis. However, the LSU tree indicates that Pa.
subalteatus is relatively older than Pa. cyanescens, and therefore the
possibility remains that an earlier HGT event occurred between
Panaeolus and Psilocybe. Interestingly, Pholiotina smithii appears
to have an HGT event with an ancestor of Clade I, which was
recovered with the highest likelihood of a dung ecology in our
ancestral state reconstruction. Pholiotina smithii and its potentially
conspecic relative Ph. cyanopus both have terrestrial ecologies
in water- saturated and Sphagnum moss- dominated soils and
grasslands, respectively. However, critical taxa are still missing
from the Psilocybe phylogeny, such as the Sphagnum- associated
and non- psilocybin- producing P. fuscofulva that has previously
been shown to occupy an early- diverging position (45), which
may substantially alter the ancestral state reconstructions.
HGT seems to be the most likely explanation for the polyphy-
letic distribution of the psilocybin BGC. Topological constraint
tests of the PsiD phylogeny was performed, and results rejected
the alternative hypothesis that recovered psilocybin- producing
Panaeolus and Pholiotina outside of Psilocybe, consistent with the
hypothesis that the gene cluster was acquired by these two genera
~67 mya
~9 mya
Panaeouls cyanescens
Panaeouls subbalteatus
Panaeolus
Pholiotina/ Conocybe
Pholiotina cyanopus
Psilocybe
Psilocybe atrobrunnea
Gymnopilus
Gymnopilus junonius
Gymnopilus aeruginosus
Pluteus
Pluteus salicinus
~40 mya
~11-22 mya
I
II
0.2
Psilocybe portoricensis (CFMR-PR-4572)|Psi D
Psilocybe zapotecorum (GAM00011076)|Psi D
Psilocybe heimii Isotype (NY-761610)|Psi D
Psilocybe arcana Isotype (SFSU-F-000737)|Psi D
Psilocybe caerulescens (NY-1920304)|Psi D
Psilocybe mexicana (IBUG-13593)|Psi D
Psilocybe cyanofibrillosa Isotype (NY-761605)|Psi D
Psilocybe hoogshagenii (SFSU-F-029980)|Psi D
Psilocybe guilartensis Paratype (CFMR-PR-5680)|Psi D
Psilocybe callosa (NY-1595861)|Psi D
Fibsp1 938295
Psilocybe galindii Isotype (NY-761609)|Psi D
Psilocybe zapotecorum (FFCL689)|Psi D
Psilocybe xalapensis Isotype (NY-761630)|Psi D
Psilocybe pleurocystidiosa Isotype (NY-761619)|Psi D
Psilocybe bohemica (SFSU-F-029930)|Psi D
Psilocybe ovoideocystidiata Isotype (XAL-51B)|Psi D
Psilocybe muliercula (GAM00011071)|Psi D
Gymnopilus dilepis (ASM293838v1)|PsiD
Psilocybe congolensis Isotype (NY-1652567)|Psi D
Gymjun1|Psi D
Psilocybe tampanensis (UBC-F10177)|Psi D
Psilocybe hopii Isotype (XAL) |Psi D
Psilocybe subhoogshagenii (NY-915004)|Psi D
Psilocybe hoogshangenii* (GAM00011063)|Psi D
Psicy01 2263
Psilocybe yungensis (SFSU-29944)|Psi D
Psilocybe azurescens (UT-M0001775)|Psi D
Psilocybe cyanescens (WTU-F-011306)|Psi D
Pluteus albostipitatus (KM-54312)|Psi D
Psilocybe subfimetaria (SFSU-F-029945)|Psi D
Pholiotina smithii 2634DeC
Pluteus salicinus (KM-265162)|Psi D
Psilocybe singeri Isotype (NY-761622)|Psi D
Psilocybe cyanescens (KY984104-1)*|Psi D
Psilocybe caerulipes* (NY-1901130)|Psi D
Psilocybe caerulipes (PUL00030154)|Psi D
Psilocybe columbiana Isotype (NY-761607)|Psi D
Psilocybe liniformans var. americana (NY-1797145)|Psi D
Psilocybe azurescens Holotype (WTU-F19095)|Psi D
Psilocybe cubensis
Pancy01 5060
Psubaeruginosa
Gymlu01 2525
Psilocybe wrightii Isotype (NY-761629)|Psi D
Psilocybe angustipleurocystidiata Isotype (NY-761597)|Psi D
P
s
il
ocy
b
e caeru
l
escens var. mazatecorum
(SFSU
-
F
-
029971)|P
s
i
D
Psilocybe cyanescens* (WTU-F-011245)|Psi D
Psilocybe weilii Isotype (WTU-F-063525)|Psi D
Psilocybe cubensis (KY984101-1)|Psi D
Psilocybe pelliculosa (WTU-F-012331)|Psi D
Psilocybe aztecorum var. bonetii (NY-1595856)|Psi D
Psilocybe samuiensis (WTU-F-055014)|Psi D
Psilocybe stuntzii (WTU-F-011520)|Psi D
Gymnopilus luteofolius (KM-27062)|Psi D
Psilocybe caerulipes* (VPI-F-0003693)|Psi D
Psilocybe cubensis (IBUG-4367)|Psi D
Psilocybe argentipes (NY-1595850)|Psi D
Fibsp1 949465
Psilocybe hoogshagenii var. convexa Isotype (NY-761612)|Psi D
Pholiotina smithii
Psilocybe subyungensis Isotype (NY-1197500)|Psi D
Psilocybe fagicola var. mesocystidiata (NY-761608)|Psi D
Gymch1 1578740
Psilocybe semilanceata (SFSU-F-029972)|Psi D
Psilocybe moravica Isotype (SFSU-F-000732)|Psi D
Psilocybe aztecorum (SFSU-F-029933)|Psi D
Psilocybe subcubensis (SFSU-29974)|Psi D
Panaeolus cyanescens (ASM293835v1)|Psi D
Psilocybe fimetaria (UBC-F30923)|Psi D
Pazurescens
100
34
100
57
100
73
91
83
84
100
100
99
99
79
100
100
85
100
100
81
80
100
100
100
98
59
98
99
100
100
100
65
100
93
61
100
93
60
100
86
38
100
71
75
97
100
100
98
100
97
87
96
48
98
45
96
72
100
99
98
69
76
99
90
100
100
98
100
Soil with
Wood
Soil
wood debris
Dung
Pl
uteu
s
P
lu
P
P
t
eu
t
t
s s
u
u
ali
c
i
i
inu
c
c
s
u
u
Fig.3. PsiD gene tree and LSU gene tree of psilocybin- producing mushrooms and their inactive relatives. Left: Maximum- likelihood phylogram of Agaricomycetes
PsiD homologs. Tree is rooted with Phallomycetidae following Reynolds etal. (25). Numbers on branches are percent nonparametric rapid bootstraps implemented
using the within- partition ultrafast bootstrapping option in IQ- TREE. Icons depicting ecological niche are indicated to the Right of the terminal labels and legend is
depicted in the Upper left. Major clades are denoted with roman numerals and background shading. Right: Maximum- likelihood time tree of all publicly available LSU
sequences for species with the homologous canonical psilocybin BGC and their non- active relatives. Tree is rooted with Calocybe gambosa following the Agaricales
topology in Dentinger etal. (37). Species with the psilocybin BGC are indicated by blue branches. Genera are color- coded in each tree with lines and terminal
labels to corresponding psilocybin- producing taxa outside of Psilocybe in the PsiD tree. Nodes where the earliest possible acquisition of the psilocybin BGC likely
occurred are indicated by blue dots. Time scale is at the bottom with the inferred dates of psilocybin BGC acquisition for each lineage indicated by hashmarks.
Downloaded from https://www.pnas.org by Bryn Dentinger on January 9, 2024 from IP address 136.36.244.244.
6 of 9   https://doi.org/10.1073/pnas.2311245121 pnas.org
through HGT (25) (SI Appendix, Table S4). Additionally, our
phylogenetic analyses are consistent with these prior reports and
show that the psilocybin BGC has been acquired (likely by
HGT) in mushrooms with increasing frequency over the past
~67 mya. In contrast, our molecular dating results provide evi-
dence that the psilocybin BGC has been present in Psilocybe
between ~56 (MRCA of the genus Psilocybe) and ~67 mya
(MRCA of Psilocybe and Panaeolus) (Fig. 3, Right). While our
LSU dataset presented a more recent divergence estimate for the
MRCA of Psilocybe, we only included sequences from genera
containing psilocybin- producing species so these estimates are
likely to be unreliable due to missing taxa. In comparison, diver-
gence dating of our phylogenomic dataset placed the MRCA of
all of Psilocybe at ~57 mya and the MRCA of Psilocybe and
Agrocybe at ~67 mya (SI Appendix).
Our molecular dating analyses relied entirely on secondary
calibrations, which are vulnerable to inaccuracies (46). Although
these secondary calibrations were derived from studies that used
fossil calibrations and we applied uniform rather than normal
prior distributions, our absolute dates should still be considered
tentative estimates until primary calibrations are available or cor-
roborating evidence is provided. Although a handful of fossil
Agaricales exist, the closest fossils to Psilocybe are Nidulariaceae,
estimated to be between 45 and 90 mya, but this family is still
very distantly related to Psilocybe (38, 47). An expanded dataset
to enable a single calibration point would be compromised by
additional phylogenetic uncertainty. Moreover, many of the avail-
able fossils cannot be placed phylogenetically with high condence
due to the dearth of reliable morphological characters. However,
regardless of the absolute timing, the presence of the gene cluster
in all Psilocybe s.s. specimens from this study suggests that the
psilocybin BGC is primarily inherited vertically within the genus,
although some patterns suggest that this may not be the only
mechanism.
e evolution and organization of the psilocybin BGC in
Psilocybe exhibit some patterns that are consistent with multiple
gains or losses. Two dierent gene orders within the BGC correlate
with the earliest divergence within Psilocybe (Fig. 2, Left), indicat-
ing an ancient reconstruction of genes within the genus.
Interestingly, the dierence in gene order is consistent with a cir-
cular transmission model (48), in which the core cluster genes
undergo HGT via a circularized intermediate, which linearizes
upon insertion into a new genome, proposed for Pluteus salicinus
in Awan et al. (24). is nding also indicates that transcriptional
regulation of the genes may vary among taxa, resulting in
taxon- specic metabolic proles (49, 50). Due to a high number
of discontiguous assemblies in our dataset, future studies should
place considerable eort on obtaining highly contiguous genomes
from modern collections to better rene the psilocybin BGC
within Psilocybe.
Duplication of PsiH. Upon deeper investigation of PsiH, we
found that it had considerably more sequence variation than any
of the other core genes and has undergone multiple duplications
or multiple losses (Fig.2 and SIAppendix, TableS3). When we
included second- best hit identications to the PsiH phylogeny, we
found that the splitting of clade II became less severe, suggesting
multiple paralogs are present among many species of Psilocybe.
PsiH encodes an indole- 4- monooxygenase that catalyzes the
oxidation of tryptamine at the 4- carbon position of the indole
ring (15), a feature that is critical to the pharmacological properties
of psilocybin/psilocin and that is both rare in nature and dicult
to achieve with traditional synthetic chemistry. is innovation in
the psilocybin BGC could be considered the most important to
the functioning of the pathway due to its specic activity and is
critical for invivo synthesis of psilocybin in heterologous systems
(51, 52). us, our study provides additional PsiH sequences that
may be useful for optimizing production of psilocybin or other
analogs using heterologous expression.
Ecological Role of Psilocybin. Despite its prominence in the
scientic and public imagination, experimental evidence for the
ecological role of psilocybin is largely lacking and speculation reigns
supreme, as thoroughly investigated and reviewed in Meyer and Slot
(2023) (53). Psilocybin is a prodrug that is rapidly converted to
the dephosphorylated psilocin, which mimics serotonin and binds
tightly to serotonin receptors, especially 5- HT
2A
, a pharmacological
action common to psychedelics (54). Serotonin receptors are also
found in high concentration in the human gastrointestinal tract,
where serotonin signaling is involved in a wide variety of functions,
including pain and vomiting (55–57). High- anity binding to
these receptors in mammals, and homologs in distantly related
organisms such as insects and arachnids, produces unnatural and
altered behaviors (58–61). is disorientation may be a direct
deterrent or could render the subject more vulnerable to predation
(or, alternatively, may be a reward). However, in humans, the onset
of symptoms is delayed by 30 min or more, eectively decoupling
the eects from their source. Because of the potentially large eect
of psilocybin on GI processes through interactions with serotonin
receptors in the gut, it is also possible that psilocybin functions as an
emetic or laxative to promote the dispersal of spores before they are
rendered inviable by digestion. In fact, nausea is commonly reported
in human clinical trials (62). However, other than humans, there
are few documented cases of vertebrates consuming psilocybin-
containing mushrooms (consisting entirely of domesticated dogs)
(63). Many Psilocybe mushrooms are uncommon or scarce, and it
is dicult to imagine how animals could learn to recognize them
with infrequent encounters and delayed pharmacological activity,
limiting the selective potential of psilocybin as a psychoactive
defense.
Fungal–insect interactions are ancient and widespread and pro-
vide a more logical hypothesis for development of psilocybin as a
chemical defense in mushrooms (25, 64, 65). To date, this has
been the most commonly asserted hypothesis for the ecological
role of psilocybin, but empirical studies are still lacking.
Furthermore, anecdotes and personal observations conrm that
psilocybin- containing mushrooms regularly have living insect
larvae in them and they can be reared to adults (24). While not
substantive, these anecdotes and initial studies indicate that psil-
ocybin may be ineective against insects or that some insects may
have evolved detoxifying capabilities in response. Moreover, stud-
ies demonstrating that psilocybin alters behavior are absent in
many non- model insects.
An alternative hypothesis is that psilocybin functions indirectly
in an inducible chemical defense system. Lenz et al. (64) eluci-
dated the chemical basis of the blue pigment that forms when
psilocybin- containing mushrooms are damaged as oligo/polymers
of psilocin. e conversion of psilocybin to psilocin and the link-
ing of them into chains is enzymatically controlled. Lenz et al.
(64, 65) pointed out that the psilocin oligo/polymers have chem-
ical properties similar to plant avonoids and polyphenolic tan-
nins, which produce reactive oxygen molecules that can damage
gut tissue. ey proposed the “polymer hypothesis,” where the
psilocin oligo/polymers may be an inducible defense against fun-
givory and psilocybin is simply the artillery kept in reserve for the
true chemical weaponry (64, 65). Intriguingly, the formation of
the blue psilocin oligo/polymers is invariably connected to psilo-
cybin biosynthesis throughout the multiple independent
Downloaded from https://www.pnas.org by Bryn Dentinger on January 9, 2024 from IP address 136.36.244.244.
PNAS 2024  Vol. 121  No. 3  e2311245121 https://doi.org/10.1073/pnas.2311245121  7 of 9
inheritances (24, 25) and convergent evolution (24). e main-
tenance of the enzymatic capacity to induce psilocybin conversion
lends further support to the hypothesis that it is the blue oligo/
polymer that has an ecological function rather than the possibly
accidental pharmacological eects of psilocybin itself.
Materials and Methods
DNA Extraction and Genomic Sequencing. Hymenophore fragments (5 to
15 mg) from dried fungarium samples, derived from a large collection of insti-
tutional loans made to the Natural History Museum of Utah (UMNH) (11) were
homogenized by placing them in 2.0 mL screw- cap tubes containing a single
3.0- mm and 8 × 1.5- mm stainless steel beads and shaking them in a BeadBug™
microtube homogenizer (Sigma- Aldrich, #Z763713) for 120 s at a speed setting
of 3,500 rpm. For samples collected after 1950, DNA extraction was performed
with Monarch® Genomic DNA Purification Kit (NEB, #T3010S) following the man-
ufacturer’s protocol for Tissue Lysis with a 1 h incubation at 56 °C except with
double the volume of lysis buffer and increasing the amount of wash buffer
to 550 µL during each of the wash steps. For samples collected before 1950,
including many of the type specimens (see SIAppendix, TableS1 for holding
institution and collection information), DNA extraction was performed using a
phenol- chloroform DNA extraction protocol. In short, after physical homogeni-
zation, lysis was performed as above, after which total lysate was placed in Phase
Lock Gel™ Light tubes (QuantaBio, #2302820) along with an equal volume of
OmniPur® Phenol:Chloroform:Isoamyl Alcohol (25:24:1, TE- saturated, pH 8.0)
solution (MilliporeSigma, Calibiochem #D05686) and then mixed by gentle
inversion for 15 min using a fixed speed tube rotator. After mixing, tubes were
centrifuged at maximum speed (14,000×g) for 10 min; then, the aqueous (top)
layer was transferred to a new phase- lock gel tube and the process repeated. DNA
precipitation of the aqueous phase was performed by adding 5 M NaCl to a final
concentration of 0.3 M and two volumes of room temperature absolute ethanol,
inverting the tubes 20× for thorough mixing followed by an overnight incubation
at −20 °C. The next day, DNA was pelleted by centrifugation at 14,000×g for
5 min. The DNA pellet was washed twice with freshly prepared, ice- cold 70%
ethanol, air- dried for 15 min at room temperature, and then resuspended in 150
µL of Elution Buffer from the Monarch® Genomic DNA kit. DNAs were submitted to
the High Throughput Genomics Core at the University of Utah, where sequencing
libraries were prepared using the Nextera™ DNA Flex Library Prep (Illumina®,
#20018704) and sequenced on a full lane of Illumina® NovaSeq6000 PE 2 ×
150 bp using an S4 flow cell.
Sequence Processing and Genome Assembly. Sequencing run statistics and
quality metric were visualized for each sample using FastQC version 0.11.9 and
then compared to each other using MultiQC version 1.10 (66). Library complex-
ity was estimated using the “EstimateLibraryComplexity” function of the Picard
toolkit (http://broadinstitute.github.io/picard/). Raw sequencing reads were
trimmed and quality filtered using fastP version 0.20.1 (67) and then assembled
using the paired- end assembly in SPAdes version 3.15.2 (68) with kmer values
alternating every other digit between 21 and 127, inclusive. Genome assembly
stats were quantified using a custom Perl script (https://github.com/hobrien/Perl/
blob/master/ContigStats.pl).
Homolog Extraction and Phylogenetic Analysis. Single- copy gene families
were identified from the Psilocybe serbica genome (Psiser1) Markov Cluster
(MCL) profile (clustering.2497, https://mycocosm.jgi.doe.gov/clm/run/Psiser1-
comparative- qc.2497;SvdlKh?organism=Psiser1) generated by the Joint
Genome Institute’s (JGI) Mycocosm genome portal (69, 70). This profile included
two species of Psilocybe (P. cubensis (Earle) Singer and P. serbica) (15) and six
closely related taxonomic outgroups: Agrocybe pediades (Fr.) Fayod (Agrped1)
(39), Galerina marginata (Batsch) Kühner (Galma1) (71), Gymnopilus chrysopellus
(Berk. & M.A. Curtis) Murrill (Gymch1) (39), G. junonius (Fr.) P.D. Orton (Gymjun1)
(39), Hebeloma cylindrosporum Romagn. (Hebcy2) (72), and Pholiota alnicola
(Fr.) Singer [=Flammula alnicola (Fr.) P. Kumm.)] (Phoaln1) (39). Amino acid
sequences of the resulting 2,983 genes in the cluster from Galma1:v1 (Galerina
marginata) (71) were used as queries for homolog identification in the new
assemblies using the “protein2genome” model in exonerate version 2.4.0 (73).
The protein2genome model allows for amino acid sequences to be aligned to a
genome allowing introns, frameshifts, and exon phase changes in the alignment
to better identify orthologous genes, even when distantly related. The output-
specific flag “- - ryo “%tcs\n”” was also used to extract the sequence information
that occurs in the target genome coding sequence of the protein2genome align-
ment, which was used for further downstream phylogenetic analysis. Nucleotide
sequence output from exonerate for each amino acid query was combined and
aligned using the multiple sequence alignment program MAFFT version 7.475
(74) with the parameters - - maxiterate 1000 - - localpair - - reorder. Concatenation
and multiple sequence alignment summary statistics were performed using the
package AMAS (75). Individual gene trees were estimated with IQ- TREE multi-
core version 2.2.0.3 (76) using ModelFinder (77) and 1,000 ultrafast bootstrap
replicates optimized using the - - bnni flag (78). A summary coalescent tree was
constructed from the individual gene trees using ASTRAL- III (79) after removing
branches with low support (BS < 10%). A phylogenomic tree was also constructed
from a concatenated supermatrix of all genes using a partitioned analysis in IQ-
TREE allowing model selection and model parameter estimation for each gene
partition separately with branch lengths shared among all partitions (option - p),
and branch supports estimated as above.
Divergence Dating. The fast- dating method of the relative rate framework (RRF),
implemented in RelTime and incorporated into the MEGA 11 software package
(80–82), was used to estimate divergence times using the ML topology and branch
lengths estimated from the concatenated supermatrix with Psilocybe s.s. as the in-
group and six selected outgroup taxa (SIAppendix, Fig.S1). We chose this method
as it has been shown to have comparable accuracy to Bayesian- based approaches
but requires far less computational resources (83). However, with 2,983 loci, the
computational burden was excessive even with RelTime. Therefore, to further reduce
computational burden, we selected a subset of 100 loci from the 2,983 genes whose
gene trees were the most clock- like (had the least amount of root- to- tip variation)
and had maximal average bipartition support using the package SortaDate (84) to
use in RelTime. To calibrate the divergence time estimations, the node representing
the most recent common ancestor (MRCA) of Psilocybe plus Agrocybe pediades,
Galerina marginata, and Gymnopilus spp. [=Strophariaceae sensu Varga etal. (38)
and Hymenogastraceae sensu Ruiz- Dueñas et al. (39)] was constrained using a
uniform prior with a minimum of 57 mya [well- supported core shift stem age of
Strophariaceae s.s.; Varga etal. (38)] and maximum of 71 mya [stem age of the
Hymenogastraceae; Ruiz- Dueñas etal. (39)], and fixing the stem age of Gymnopilus
to 54 mya [Varga etal. (38)]. The number of lineages- through- time (LTT) was esti-
mated with the “ltt” function in the R package ape (85) using log- transformed
y- values to account for progressively fewer extinction events toward the present.
Incorporating Publicly Available Sequences. Four traditional phylogenetic
loci for Fungi (ITS, Ef1a, RPB1, and RPB2) were bioinformatically extracted from the
genome assemblies using PathRacer (86). PathRacer uses hidden Markov models
(HMMs) to query de Bruijn assembly graphs for compatible paths, independent
of assembled contiguous sequences, minimizing the limitations of discontiguous
assemblies for sequence mining. HMMs were built using nucleotide sequences
from 10,859 complete ITS sequences obtained from the NCBI RefSeq Targeted
Loci Project (87), and 487 partial EF1a exons, 639 partial RPB1 exons, and 937
partial RPB2 exons generated by AFTOL (88) and downloaded from GenBank,
and aligned automatically with the L- INS- i algorithm in MAFFT version 7.475
(74). HMM profiles for each gene were generated with hmmer version 3.1b2
(hmmer.org) using the multiple sequence alignments as inputs. ITS sequences
were manually trimmed from edge sequence outputs. Sequences of EF1a,
RPB1, and RPB2 were extracted from edge sequence outputs with exonerate
using the flag - protein2genome using query protein sequences from Psilocybe
cyanescens Wakef. in GenBank (GenBank accessions: Ef1a=ADI71893.1;
RPB1=AHB18799.1; RPB2=AHH34099.1). All Psilocybe sequences correspond-
ing to the same loci (ITS, EF1a, RPB1, and RPB2) were retrieved from NCBI as
well as all Psilocybe species hypotheses (SH) ITS sequences from the curated
fungal database UNITE (89). Each dataset from NCBI was checked for overlap
of different genes for a single vouchered specimen using fungarium accession
numbers or collections IDs in the GenBank records. The loci extracted from the
genomes generated in this study were combined with the publicly available data
in single locus matrices and then automatically aligned with the L- INS- i algorithm
in MAFFT. Phylogenetic trees for each gene as well as a concatenated supermatrix
of the four traditional phylogenetic loci were constructed using IQ- TREE with the
same parameters as above, including gene partitions in the concatenated matrix
to allow for separate models and model parameters to be estimated per gene,
Downloaded from https://www.pnas.org by Bryn Dentinger on January 9, 2024 from IP address 136.36.244.244.
8 of 9   https://doi.org/10.1073/pnas.2311245121 pnas.org
and with the addition of a backbone constraint using the rooted phylogenomic
topology implemented with the - g flag.
Identication of Psilocybin BGC Genes in Psilocybe.
We used gene prediction
and reciprocal- best BLAST (RBB) to identify homologs of gene from the psilocy-
bin BGC in our genome assemblies. Due to the diploid nature of our genomes,
we first used the software package Redundans version 0.14a (90) to phase our
assemblies before gene prediction. Psilocybin gene cluster detection was per-
formed using the Augustus version 3.4.0 gene prediction software (91) with the
flags –singlestrand=true to predict genes independently on each strand with the
coprinus_cinereus training set. Following gene prediction, the most likely psilo-
cybin BGC for each species was chosen as follows: Using the protein sequences
for PsiD (tryptophan decarboxylase), PsiK (kinase), PsiM (methyltransferase), and
PsiH (P450) based on the publicly available chromosomal level assembly for
P. cubensis (92), we used BLASTP with default parameters to search for homologs
in the predicted translated proteins for each species. The top three hits for each
gene were then used together to find putative clusters by examining the gene
numbers assigned by augustus. Putative “clusters” were identified as any subset
of the 21 genes where the predicted gene numbers of any gene within the subset
was at most two genes away from one other gene in the subset. This set of com-
putationally generated, putative psilocybin BGCs was then manually curated by
querying the top three hits against the Psi gene sequences from the chromosomal
assembly of P. cubensis using BLASTn to measure their similarities. The largest
subset with high similarities to the Psi genes was chosen as the cluster candidate
for each Psilocybe species (Fig.2). To corroborate our gene predictions, we utilized
exonerate with the same Psi protein queries. Identifications were compared to
gene prediction output to determine the order and orientation of each gene in the
cluster based on the strand position (+/- ) of the start codon for each gene (Fig.2).
Individual gene multiple sequence alignments and maximum likelihood trees
were estimated for each psi gene using MAFFT and IQ- TREE as above.
Global PsiD Phylogenetic Analysis, Topological Constraint Testing, and
Ancestral State Reconstruction. PsiD nucleotide sequences extracted from
all primary psilocybin BGC- containing mushrooms with exonerate were tran-
scribed utilizing codon- aware multiple sequence alignment through HYPHY
version 2.5.36 (93) following the tutorial outlined at https://github.com/veg/
hyphy- analyses/blob/master/codon- msa/README.md. Extracted PsiD amino acid
sequences were combined with the PsiD amino acid sequences published in
Reynolds etal. (25), undergoing multiple sequence alignment and phylogenetic
analysis under the same parameters used for individual gene trees. Ecological
lifestyle was determined from original species descriptions and monographs
when available, or directly from collection notes (SIAppendix, TableS1).
To test the hypothesis that psilocybin- producing Panaeolus (Fr.) Quél. and
Pholiotina Fayod PsiD are derived within Psilocybe PsiD, consistent with HGT, topo-
logical constraint testing was performed by comparing the best ML tree constrained
to recover a monophyletic Psilocybe in IQ- TREE version 1.6.12 with the uncon-
strained PsiD tree using bootstrap proportion (BP), weighted Kishino–Hasegawa
test (94), weighted Shimodaira–Hasegawa test (95), expected likelihood weights
(96) with 1,000 RELL replicates, and the approximately unbiased (AU) test (97). An
unrooted topological constraint tree was rendered in Mesquite version 3.2 (98).
Ancestral state reconstruction was performed on our concatenated supermatrix
phylogenetic tree using RASP 4 (4.3 build 20220517 x64) (99). Each sample was
coded for their respective ecological niches (soil, wood, dung, and soil with wood
debris) and run under default settings using the MCMC Bayesian option of the
Multistate Reconstruction in BayesTraits model (100) (SIAppendix). MCMC output
was analyzed using Tracer version 1.7.2 (101), which displayed a unimodal distribution
supporting convergence with an effective samples size (ESS) of 450 (SIAppendix).
Tempo, Mode, and Timing of the Primary Psilocybin BGC.
In order to assess
the relative timing of psilocybin biosynthesis across mushrooms, a dataset was
constructed of all publicly available sequences (194 individuals across 147 spe-
cies) of the ribosomal large subunit DNA region (28S; downloaded from NCBI
on 14 August 2022) for all mushroom genera with species containing the pri-
mary psilocybin BGC: Conocybe Fayod /Pholiotina, Gymnopilus, Panaeolus Fr.,
Pluteus, and Psilocybe. A sequence of Calocybe gambosa (Fr.) Donk was used as
an outgroup. Due to the poor overlap between taxa for other gene regions, the
28S region had the best species representation that could be aligned across all
genera. Multiple sequence alignment was performed using the L- INS- i algorithm
in MAFFT. Phylogenetic analysis was performed under maximum likelihood using
IQ- TREE with automatic model selection and up to 1,000 nonparametric rapid
bootstraps. Divergence dating of the LSU gene tree was performed with RelTime
in the same manner as above but with fixed node ages for Gymnopilus p.p. (54
mya), Panaeolus (60 mya), and Conocybe (45 mya) based on Varga etal. (38).
Data, Materials, and Software Availability. Raw short- read sequences
for this project have been deposited in the Sequence Read Archive (SRA) and
assigned the bioproject number PRJNA904752 and SRA and Biosample acces-
sion numbers are reported in SIAppendix, TableS1. Raw tree files, assemblies,
gene prediction files, multiple sequence alignments, and Psilocybe constraint
tree are available through Dryad (https://datadryad.org/stash/landing/show?id=
doi%3A10.5061%2Fdryad.tmpg4f52s) (102). Any code or specific script requests
should be sent to the corresponding author.
ACKNOWLEDGMENTS. We acknowledge the Natural History Museum of Utah for
its commitment to collaborative Science and the Genomics Core Facility, a part of the
Health Sciences Cores at The University of Utah, for their input and high- quality work.
Additionally, we acknowledge the numerous institutions that provided specimens
for destructive sampling, many of which were rare and irreplaceable. Further, we wish
to recognize the dedicated and hard work performed by Isabelle Galland and Talia
A. Backman in their help processing many of these samples for DNA sequencing.
Further, we wish to thank David Scott Flocken for his insight and thought- provoking
conversations on the work generated here. Additionally, we would like to thank
Jan Borovička and Oscar Castro- Jauregui for providing high- quality images and
graciously allowing us to use them for publication. We would also like to thank Paul
Stamets who provided images and feedback on this publication, and Fungi Perfecti
LLC who provided funding which helped facilitate this work. Additionally, we would
like to thank Dr. Jason Slot for providing data from previous publications as well as
intriguing conversions about Psilocybin evolution. V.R.- C. and L.G.- D. would like
to thank the University of Guadalajara and CONAHCYT for their support of their
research. This work was supported by a grant from NSF (DEB #2114785) and partially
from a generous donation to the NHMU from Fungi Perfecti LLC.
Author aliations: aSchool of Biological Sciences, University of Utah, Salt Lake City, UT 84112;
bNatural History Museum of Utah, Collections and Research, University of Utah, Salt Lake
City, UT 84108; cConsejo Nacional de Humanidades, Ciencias y Tecnologías (CONAHCYT),
Departamento de Botánica y Zoología, Universidad de Guadalajara, Zapopan 45147,
Mexico; dGenomics Innovation Unit, Guy’s and St.Thomas’ NHS Foundation Trust, St Thomas’
Hospital, London SE1 7EH, United Kingdom; eFungi Foundation, Brooklyn, NY 11216; and
fDepartamento de Botánica y Zoología, Universidad de Guadalajara, Zapopan 45147, Mexico
1. R. L. Carhart- Harris et al., The entropic brain: A theory of conscious states informed by neuroimaging
research with psychedelic drugs. Front. Hum. Neurosci. 8, 20 (2014).
2. M. W. Johnson, R. R. Griffiths, Potential therapeutic effects of Psilocybin. Neurotherapeutics 14,
734–740 (2017).
3. M. W. Johnson, P. S. Hendricks, F. S. Barrett, R. R. Griffiths, Classic psychedelics: An integrative review
of epidemiology, therapeutics, mystical experience, and brain network function. Pharmacol. Therap.
197, 83–102 (2019).
4. E. Fries, Observationes Mycologicae (Havniae: Sumptibus G. Bonnieri, 2021), pp. 1815- 1818.
5. R. Heim, R. G. Wasson, Les Champignons Hallucinogènes du Mexique: Ètudes Ethnologiques,
Taxinomiques, Biologiques, Physiologiques et Chimiques (Museum National d’Histoire Naturelle, 1959).
6. G. Guzmán Huerta, Nueva localidad de importancia etnomicológica de los hongos neurotrópicos
mexicanos (Necaxa, Puebla) Ciencia (México) 20, 85–88 (1960).
7. N. Echevarria, B. Arochi, M. Sabina, A. Henestrosa, A. Estrada, María Sabina: Mujer Espíritu [Woman
Spirit] (CONACULTA–IMCINE, 2006).
8. G. Guzmán, Hallucinogenic mushrooms in Mexico: An overview. Econ. Bot. 62, 404–412
(2008).
9. J. Kristinsson, Occurrence and use of hallucinogenic mushrooms containing psilocybin alkaloids (Nordic
Council of Ministers, 2008).
10. V. Ramírez- Cruz et al., Phylogenetic inference and trait evolution of the psychedelic mushroom
genus Psilocybe sensu lato (Agaricales). Botany 91, 573–591 (2013).
11. A. J. Bradshaw et al., DNA authentication and chemical analysis of Psilocybe mushrooms reveal
widespread misdeterminations in Fungaria and inconsistencies in metabolites. Appl. Environ.
Microbiol. 88, e01498- 22 (2022).
12. O. S. Castro Jauregui, V. Ramírez- Cruz, A. J. Bradshaw, A. Cortés- Pérez, L. Guzmán- Dávalos, Los
hongos sagrados del género Psilocybe en Jalisco. Nubes y Ciencia 12, 14–21 (2022).
13. R. C. Van Court et al., Diversity, biology, and history of psilocybin- containing fungi: Suggestions for
research and technological development. Fungal Biol. 126, 308–319 (2022).
14. R. G. Wasson, Seeking the magic mushroom. Life 49, 109–120 (1957).
Downloaded from https://www.pnas.org by Bryn Dentinger on January 9, 2024 from IP address 136.36.244.244.
PNAS 2024  Vol. 121  No. 3  e2311245121 https://doi.org/10.1073/pnas.2311245121  9 of 9
15. J. Fricke, F. Blei, D. Hoffmeister, Enzymatic synthesis of psilocybin. Angew. Chem. Int. Ed. 56,
12352–12355 (2017).
16. J. C. Slot, E. Gluck- Thaler, Metabolic gene clusters, fungal diversity, and the generation of accessory
functions. Curr. Opin. Genet. Dev. 58–59, 17–24 (2019).
17. N. P. Keller, Translating biosynthetic gene clusters into fungal armor and weaponry. Nat. Chem. Biol.
11, 671–677 (2015).
18. N. P. Keller, Fungal secondary metabolism: Regulation, function and drug discovery. Nat. Rev.
Microbiol. 17, 167–180 (2019).
19. S. R. Waterman, D. W. Holden, Functions and effectors of the Salmonella pathogenicity island 2 type
III secretion system. Cell Microbiol. 5, 501–511 (2003).
20. A. E. Osbourn, B. Field, Operons. Cell. Mol. Life Sci. 66, 3755–3775 (2009).
21. D. DellaPenna, S. E. O’Connor, Plant gene clusters and opiates. Science 336, 1648–1649 (2012).
22. B. R. Lichman et al., The evolutionary origins of the cat attractant nepetalactone in catnip. Sci. Adv. 6,
eaba0721 (2020).
23. I. Burkhardt, T. De Rond, P.Y.- T. Chen, B. S. Moore, Ancient plant- like terpene biosynthesis in corals.
Nat. Chem. Biol. 18, 664–669 (2022).
24. A. R. Awan et al., Convergent evolution of psilocybin biosynthesis by psychedelic mushrooms.
bioRxiv [Preprint] (2018). https://doi.org/10.1101/374199 (Accessed 15 November 2023).
25. H. T. Reynolds et al., Horizontal gene cluster transfer increased hallucinogenic mushroom diversity.
Evol. Lett. 2, 88–101 (2018).
26. K. McKernan et al., A whole genome atlas of 81 Psilocybe genomes as a resource for psilocybin
production. F1000Res 10, 961 (2021).
27. S. Dörner et al., Genetic survey of Psilocybe natural products. ChemBioChem 23, e202200249 (2022).
28. J. Borovička, M. E. Noordeloos, M. Gryndler, M. Oborník, Molecular phylogeny of Psilocybe
cyanescens complex in Europe, with reference to the position of the secotioid Weraroa novae-
zelandiae. Mycol. Progr. 10, 149–155 (2011).
29. C. Andrew, J. Diez, T. Y. James, H. Kauserud, Fungarium specimens: A largely untapped source in
global change biology and beyond. Philos. Trans. R. Soc. B 374, 20170392 (2019).
30. K. C. Rowe et al., Museum genomics: Low- cost and high- accuracy genetic data from historical
specimens. Mol. Ecol. Resources 11, 1082–1092 (2011).
31. K. Bi et al., Unlocking the vault: Next- generation museum population genomics. Mol. Ecol. 22,
6018–6032 (2013).
32. M. Staats et al., Genomic treasure troves: Complete genome sequencing of herbarium and insect
museum specimens. PLoS ONE 8, e69189 (2013).
33. G. Besnard et al., From museums to genomics: Old herbarium specimens shed light on a C3 to C4
transition. J. Exp. Bot. 65, 6711–6721 (2014).
34. H. N. Poinar et al., Molecular coproscopy: Dung and diet of the extinct ground sloth Nothrotheriops
shastensis. Science 281, 402–406 (1998).
35. A. P. Boast et al., Coprolites reveal ecological interactions lost with the extinction of New Zealand
birds. Proc. Natl. Acad. Sci. U.S.A. 115, 1546–1551 (2018).
36. S. M. Latorre et al., Museum phylogenomics of extinct Oryctes beetles from the Mascarene Islands.
bioRxiv [Preprint] (2021). https://doi.org/10.1101/2020.02.19.954339 (Accessed 15 November 2023).
37. B. T. M. Dentinger et al., Tales from the crypt: Genome mining from Fungarium specimens improves
resolution of the mushroom tree of life. Biol. J. Linn. Soc. 117, 11–32 (2016).
38. T. Varga et al., Megaphylogeny resolves global patterns of mushroom evolution. Nat. Ecol. Evol. 3,
668–678 (2019).
39. F. J. Ruiz- Dueñas et al., Genomic analysis enlightens agaricales lifestyle evolution and increasing
peroxidase diversity. Mol. Biol. Evol. 38, 1428–1446 (2021).
40. Z. Konkel, K. Scott, J. C. Slot, Draft genome sequence of the termite- associated “cuckoo fungus”,
Athelia (Fibularhizoctonia) sp. TMB Strain TB5. Microbiol. Resour. Announc. 10, e01230- 20 (2021).
41. K. Matsuura, Distribution of termite egg- mimicking fungi (“termite balls”) in Reticulitermes spp.
(Isoptera: Rhinotermitidae) nests in Japan and the United States. Appl. Entomol. Zool. 40, 53–61 (2005).
42. L. Arvidsson, Athelia arachnoidea (Berk.) Jül. and its influence on epiphytic cryptogams in urban
areas. Göteborgs Svampklubbs Arsskrift 1976, 4–10 (1975).
43. L. Arvidsson, Svampangrepp på lavar—En orsak till lavöken. Svensk Botanisk Tidskrift 72, 285–292 (1979).
44. G. C. Adams, B. R. Kropp, Athelia arachnoidea, the sexual state of Rhizoctonia carotae, a pathogen of
carrot in cold storage. Mycologia 88, 459–472 (1996).
45. J. Borovička et al., Phylogenetic and chemical studies in the potential psychotropic species complex
of Psilocybe atrobrunnea with taxonomic and nomenclatural notes. Persoonia 34, 1–9 (2015).
46. J. J. Schenk, Consequences of secondary calibrations on divergence time estimates. PLoS ONE 11,
e0148228 (2016).
47. M. Sánchez- García et al., Fruiting body form, not nutritional mode, is the major driver of
diversification in mushroom- forming fungi. Proc. Natl. Acad. Sci. U.S.A. 117, 32528–32534 (2020).
48. C. M. Johnson, A. D. Grossman, Integrative and conjugative elements (ICEs): What they do and how
they work. Annu. Rev. Genet. 49, 577–601 (2015).
49. D. Hoffmeister, N. P. Keller, Natural products of filamentous fungi: Enzymes, genes, and their
regulation. Nat. Prod. Rep. 24, 393–416 (2007).
50. W. Yin, N. P. Keller, Transcriptional regulatory elements in fungal secondary metabolism. J.
Microbiol. 49, 329–339 (2011).
51. A. M. Adams et al., In vivo production of psilocybin in E. coli. Metab. Eng. 56, 111–119 (2019).
52. N. Milne et al., Metabolic engineering of Saccharomyces cerevisiae for the de novo production of
psilocybin and related tryptamine derivatives. Metab. Eng. 60, 25–36 (2020).
53. M. Meyer, J. Slot, The evolution and ecology of psilocybin in nature. Fungal Genet. Biol. 167,
103812 (2023).
54. A. M. Araújo, F. Carvalho, M. de L. Bastos, P. Guedes de Pinho, M. Carvalho, The hallucinogenic world
of tryptamines: An updated review. Arch. Toxicol. 89, 1151–1173 (2015).
55. J. C. Bornstein, Serotonin in the gut: What does it do? Front. Neurosci. 6, 16 (2012).
56. G. M. Mawe, J. M. Hoffman, Serotonin signalling in the gut–functions, dysfunctions and therapeutic
targets. Nat. Rev. Gastroenterol. Hepatol. 10, 473–486 (2013).
57. A. S. French et al., The role of serotonin in feeding and gut contractions in the honeybee. J. Insect.
Physiol. 61, 8–15 (2014).
58. A. Christiansen, R. Baum, P. N. Witt, Changes in spider webs brought about by mescaline, psilocybin
and an increase in body weight. J. Pharmacol. Exp. Therap. 136, 31–37 (1962).
59. C. D. Nichols, J. Ronesi, W. Pratt, E. Sanders- Bush, Hallucinogens and Drosophila: Linking serotonin
receptor activation to behavior. Neuroscience 115, 979–984 (2002).
60. T. Passie, J. Seifert, U. Schneider, H. M. Emrich, The pharmacology of psilocybin. Addict. Biol. 7,
357–364 (2002).
61. G. R. Boyce et al., Psychoactive plant- and mushroom- associated alkaloids from two behavior
modifying cicada pathogens. Fungal Ecol. 41, 147–164 (2019).
62. R. L. Carhart- Harris et al., Psilocybin with psychological support for treatment- resistant depression:
An open- label feasibility study. Lancet Psychiatry 3, 619–627 (2016).
63. M. Beug, M. Shaw, K. Cochran, Thirty plus years of mushroom poisoning: Summary of the
approximately 2,000 reports in the NAMA Case Registry. McIlvainea 16, 47–68 (2006).
64. C. Lenz et al., Injury- triggered blueing reactions of Psilocybe “magic” mushrooms. Angew. Chem.
Int. Ed. 59, 1450–1454 (2020).
65. C. Lenz, S. Dörner, A. Sherwood, D. Hoffmeister, Structure elucidation and spectroscopic analysis of
chromophores produced by oxidative psilocin dimerization. Chem. Eur. J. 27, 12166–12171 (2021).
66. P. Ewels, M. Magnusson, S. Lundin, M. Käller, MultiQC: Summarize analysis results for multiple tools
and samples in a single report. Bioinformatics 32, 3047–3048 (2016).
67. S. Chen, Y. Zhou, Y. Chen, J. Gu, fastp: An ultra- fast all- in- one FASTQ preprocessor. Bioinformatics 34,
i884–i890 (2018).
68. A. Bankevich et al., SPAdes: A new genome assembly algorithm and its applications to single- cell
sequencing. J. Comput. Biol. 19, 455–477 (2012).
69. I. V. Grigoriev et al., The genome portal of the Department of Energy Joint Genome Institute. Nucleic
Acids Res. 40, D26–D32 (2012).
70. I. V. Grigoriev et al., MycoCosm portal: Gearing up for 1000 fungal genomes. Nucleic Acids Res. 42,
D699–D704 (2014).
71. R. Riley et al., Extensive sampling of basidiomycete genomes demonstrates inadequacy of the
white- rot/brown- rot paradigm for wood decay fungi. Proc. Natl. Acad. Sci. U.S.A. 111, 9923–9928
(2014).
72. Mycorrhizal Genomics Initiative Consortium et al., Convergent losses of decay mechanisms and
rapid turnover of symbiosis genes in mycorrhizal mutualists. Nat. Genet. 47, 410–415 (2015).
73. G. S. C. Slater, E. Birney, Automated generation of heuristics for biological sequence comparison.
BMC Bioinf. 6, 31 (2005).
74. K. Katoh, MAFFT: A novel method for rapid multiple sequence alignment based on fast Fourier
transform. Nucleic Acids Res. 30, 3059–3066 (2002).
75. M. L. Borowiec, AMAS: A fast tool for alignment manipulation and computing of summary statistics.
PeerJ 4, e1660 (2016).
76. L.- T. Nguyen, H. A. Schmidt, A. von Haeseler, B. Q. Minh, IQ- TREE: A fast and effective stochastic
algorithm for estimating maximum- likelihood phylogenies. Mol. Biol. Evol. 32, 268–274 (2015).
77. S. Kalyaanamoorthy, B. Q. Minh, T. K. F. Wong, A. von Haeseler, L. S. Jermiin, ModelFinder: Fast
model selection for accurate phylogenetic estimates. Nat. Methods 14, 587–589 (2017).
78. D. T. Hoang, O. Chernomor, A. von Haeseler, B. Q. Minh, L. S. Vinh, UFBoot2: Improving the ultrafast
bootstrap approximation. Mol. Biol. Evol. 35, 518–522 (2018).
79. C. Zhang, M. Rabiee, E. Sayyari, S. Mirarab, ASTRAL- III: Polynomial time species tree reconstruction
from partially resolved gene trees. BMC Bioinf. 19, 153 (2018).
80. F. U. Battistuzzi, Q. Tao, L. Jones, K. Tamura, S. Kumar, RelTime relaxes the strict molecular clock
throughout the phylogeny. Genome Biol. Evol. 10, 1631–1636 (2018).
81. Q. Tao, K. Tamura, B. Mello, S. Kumar, Reliable confidence intervals for RelTime estimates of
evolutionary divergence times. Mol. Biol. Evol. 37, 280–290 (2020).
82. K. Tamura, G. Stecher, S. Kumar, MEGA11: Molecular evolutionary genetics analysis version 11.
Mol. Biol. Evol. 38, 3022–3027 (2021).
83. F. P. Costa, C. G. Schrago, B. Mello, Assessing the relative performance of fast molecular dating
methods for phylogenomic data. BMC Genomics 23, 798 (2022).
84. S. A. Smith, J. W. Brown, J. F. Walker, So many genes, so little time: A practical approach to
divergence- time estimation in the genomic era. PLoS ONE 13, e0197433 (2018).
85. E. Paradis, J. Claude, K. Strimmer, APE: Analyses of phylogenetics and evolution in R language.
Bioinformatics 20, 289–290 (2004).
86. A. Shlemov, A. Korobeynikov, "PathRacer: Racing profile HMM paths on assembly graph" in
Algorithms for Computational Biology, Lecture Notes in Computer Science, I. Holmes, C. Martín- Vide,
M. A. Vega- Rodríguez, Eds. (Springer International Publishing, 2019), pp. 80–94.
87. N. A. O’Leary et al., Reference sequence (RefSeq) database at NCBI: Current status, taxonomic
expansion, and functional annotation. Nucleic Acids Res. 44, D733–D745 (2016).
88. F. Lutzoni et al., Assembling the fungal tree of life: Progress, classification, and evolution of
subcellular traits. Am. J. Bot. 91, 1446–1480 (2004).
89. K. Abarenkov et al., The UNITE database for molecular identification of fungi—Recent updates and
future perspectives. New Phytol. 186, 281–285 (2010).
90. L. P. Pryszcz, T. Gabaldón, Redundans: An assembly pipeline for highly heterozygous genomes.
Nucleic Acids Res. 44, e113 (2016).
91. M. Stanke et al., AUGUSTUS: Abinitio prediction of alternative transcripts. Nucleic Acids Res. 34,
W435–W439 (2006).
92. K. McKernan et al., A draft reference assembly of the Psilocybe cubensis genome. F1000Res 10, 281
(2021).
93. S. L. K. Pond, S. D. W. Frost, S. V. Muse, HyPhy: Hypothesis testing using phylogenies. Bioinformatics
21, 676–679 (2005).
94. H. Kishino, M. Hasegawa, Evaluation of the maximum likelihood estimate of the evolutionary tree
topologies from DNA sequence data, and the branching order in hominoidea. J. Mol. Evol. 29,
170–179 (1989).
95. H. Shimodaira, M. Hasegawa, Multiple comparisons of log- likelihoods with applications to
phylogenetic inference. Mol. Biol. Evol. 16, 1114–1116 (1999).
96. K. Strimmer, A. Rambaut, Inferring confidence sets of possibly misspecified gene trees. Proc. R. Soc.
Lond. B 269, 137–142 (2002).
97. H. Shimodaira, An approximately unbiased test of phylogenetic tree selection. Syst. Biol. 51,
492–508 (2002).
98. W. P. Maddison, D. R. Maddison, Mesquite: A modular system for evolutionary analysis (2019).
99. Y. Yu, C. Blair, X. He, RASP 4: Ancestral state reconstruction tool for multiple genes and characters.
Mol. Biol. Evol. 37, 604–606 (2020).
100. A. Meade, M. Pagel, "Ancestral state reconstruction using BayesTraits" in Environmental Microbial
Evolution, Methods in Molecular Biology, H. Luo, Ed. (Springer, US, 2022), pp. 255–266.
101. A. Rambaut, A. J. Drummond, D. Xie, G. Baele, M. A. Suchard, Posterior summarization in Bayesian
phylogenetics using Tracer 1.7. Syst. Biol. 67, 901–904 (2018).
102. A. Bradshaw et al., Phylogenomics of the psychoactive mushroom genus Psilocybe and evolution
of the psilocybin biosynthetic gene cluster [Dataset]. Dryad. https://datadryad.org/stash/landing/
show?id=doi:10.5061/dryad.tmpg4f52s. Deposited 6 November 2023.
Downloaded from https://www.pnas.org by Bryn Dentinger on January 9, 2024 from IP address 136.36.244.244.
... Estos hongos, a lo largo de la historia, han ocupado un lugar destacado en las tradiciones chamánicas debido a su capacidad para provocar profundos cambios en los estados de consciencia (Swanson, 2018). El estudio específico de especies dentro del género Psilocybe ha permitido adentrarse en los mecanismos bioquímicos y psicológicos detrás de las alteraciones en la consciencia humana (Bradshaw et al., 2024). La presencia de estos compuestos psicoactivos ha sido identificada en especies como P. cubensis, P. semilanceata y otras, evidenciando su potencial para incidir en la percepción y la cognición humanas, generando estados alterados de consciencia que trascienden los límites convencionales de la experiencia humana. ...
Article
Full-text available
Esta revisión bibliográfica explora el papel de los hongos alucinógenos, específicamente el género Psilocybe de la familia Hymenogastraceae, en el desarrollo evolutivo de la conciencia humana. A través de un enfoque multidisciplinario que abarca biología, etnobotánica y neurociencia, se examina la influencia de compuestos como la psilocibina y la psilocina en los consumidores, revelando su potencial para desencadenar efectos neurológicos y psicológicos significativos. Se explora el impacto en la conciencia humana, señalando cambios perceptuales, cognitivos y emocionales. Se analiza el desarrollo evolutivo y antropológico de los seres humanos, explorando cómo la disponibilidad de alimentos, el uso de la psilocibina y las adaptaciones fisiológicas han podido influir en el éxito reproductivo y la supervivencia a lo largo del tiempo.
... Psilocybin, a natural tryptamine compound found in psychedelic mushrooms (Psilocybe, Conocybe/Pholiotina, Gymnopilus, Panaeolus, etc.) (Bradshaw et al., 2024), shares a similar structure and mode of action as serotonin. Despite being a Schedule I drug in the US, there is renewed interest in its therapeutic use owing to its relatively low harm profile compared to similar substances such as ketamine. ...
Article
Full-text available
With the increasing sustainability challenges, synthetic biology is offering new possibilities for addressing the emerging problems through the cultivation and fermentation of mushrooms. In this perspective, we aim to provide an overview on the research and applications mushroom synthetic biology, emphasizing the need for increased attention and inclusion of this rapidly advancing field in future mushroom technology over China and other countries. By leveraging synthetic biology, mushrooms are expected to play a more versatile role in various area, including traditional fields like circular economy, human wellness and pharmaceutics, as well as emerging fields like vegan meat, mushroom-based materials and pollution abatement. We are confident that these efforts using synthetic biology strategies have the potential to strengthen our capacity to effectively address sustainable challenges, leading to the development of a more sustainable social economy and ecology.
Preprint
Psilocybe zapotecorum is a strongly blue-bruising psilocybin mushroom used by indigenous groups in southeastern Mexico and beyond. While this species has a rich history of ceremonial use, research into its chemistry and genetics have been limited. Herein, we detail mushroom morphology and report on cultivation parameters, chemical profile, and the full genome sequence of P. zapotecorum . First, growth and cloning methods are detailed that are simple, and reproducible. In combination with high resolution microscopic analysis, the strain was barcoded, confirming species-level identification. Full genome sequencing reveals the architecture of the psilocybin gene cluster in P. zapotecorum, and can serve as a reference genome for Psilocybe Clade I. Characterization of the tryptamine profile revealed a psilocybin concentration of 17.9±1.7 mg/g, with a range of 10.6-25.7 mg/g (n=7), and similar tryptamines (psilocin, baeocystin, norbaeocystin, norpsilocin, aeruginascin, 4-HO-tryptamine, and tryptamine) in lesser concentrations for a combined tryptamine concentration of 22.5±3.2 mg/g. These results show P. zapotecorum to be a potent – and variable – Psilocybe mushroom. Chemical profiling, genetic analysis, and cultivation assist in demystifying these mushrooms. As clinical studies with psilocybin gain traction, understanding the diversity of psilocybin mushrooms will assure that psilocybin therapy does not become synonymous with psilocybin mushrooms.
Preprint
Full-text available
Psychoactive mushrooms in the genus Psilocybe have immense cultural value and have been used for centuries in Mesoamerica. Despite a recent surge in interest in these mushrooms due to emerging evidence that psilocybin, the main psychoactive compound, is a promising therapeutic for a variety of mental illnesses, their phylogeny and taxonomy remain substantially incomplete. Moreover, the recent elucidation of the psilocybin biosynthetic gene cluster is known for only five species of Psilocybe, four of which belong to only one of two major clades. We set out to improve the phylogeny for Psilocybe using shotgun sequencing of 71 fungarium specimens, including 23 types, and conducting phylogenomic analysis using 2,983 single-copy gene families to generate a fully supported phylogeny. Molecular clock analysis suggests the stem lineage arose ~66 mya and diversified ~53 mya. We also show that psilocybin biosynthesis first arose in Psilocybe, with 4-5 possible horizontal transfers to other mushrooms between 40 and 22 mya. Moreover, predicted orthologs of the psilocybin biosynthetic genes revealed two distinct gene orders within the cluster that corresponds to a deep split within the genus, possibly consistent with the independent acquisition of the cluster. This novel insight may predict differences in chemistry between the two major clades of the genus, providing further resources for the development of novel therapeutics.
Article
Full-text available
Advances in genome sequencing techniques produced a significant growth of phylogenomic datasets. This massive amount of data represents a computational challenge for molecular dating with Bayesian approaches. Rapid molecular dating methods have been proposed over the last few decades to overcome these issues. However, a comparative evaluation of their relative performance on empirical data sets is lacking. We analyzed 23 empirical phylogenomic datasets to investigate the performance of two commonly employed fast dating methodologies: penalized likelihood (PL), implemented in treePL, and the relative rate framework (RRF), implemented in RelTime. They were compared to Bayesian analyses using the closest possible substitution models and calibration settings. We found that RRF was computationally faster and generally provided node age estimates statistically equivalent to Bayesian divergence times. PL time estimates consistently exhibited low levels of uncertainty. Overall, to approximate Bayesian approaches, RelTime is an efficient method with significantly lower computational demand, being more than 100 times faster than treePL. Thus, to alleviate the computational burden of Bayesian divergence time inference in the era of massive genomic data, molecular dating can be facilitated using the RRF, allowing evolutionary hypotheses to be tested more quickly and efficiently.
Article
Full-text available
The mushroom genus Psilocybe is best known as the core group of psychoactive mushrooms, yet basic information on their diversity, taxonomy, chemistry, and general biology is still largely lacking. In this study, we reexamined 94 Psilocybe fungarium specimens, representing 18 species, by DNA barcoding, evaluated the stability of psilocybin, psilocin, and their related tryptamine alkaloids in 25 specimens across the most commonly vouchered species (Psilocybe cubensis, Psilocybe cyanescens, and Psilocybe semilanceata), and explored the metabolome of cultivated P. cubensis. Our data show that, apart from a few well-known species, the taxonomic accuracy of specimen determinations is largely unreliable, even at the genus level. A substantial quantity of poor-quality and mislabeled sequence data in public repositories, as well as a paucity of sequences derived from types, further exacerbates the problem. Our data also support taxon- and time-dependent decay of psilocybin and psilocin, with some specimens having no detectable quantities of them. We also show that the P. cubensis metabolome possibly contains thousands of uncharacterized compounds, at least some of which may be bioactive. Taken together, our study undermines commonly held assumptions about the accuracy of names and presence of controlled substances in fungarium specimens identified as Psilocybe spp. and reveals that our understanding of the chemical diversity of these mushrooms is largely incomplete. These results have broader implications for regulatory policies pertaining to the storage and sharing of fungarium specimens as well as the use of psychoactive mushrooms for recreation and therapy. IMPORTANCE The therapeutic use of psilocybin, the active ingredient in "magic mushrooms," is revolutionizing mental health care for a number of conditions, including depression, posttraumatic stress disorder (PTSD), and end-of-life care. This has spotlighted the current state of knowledge of psilocybin, including the organisms that endogenously produce it. However, because of international regulation of psilocybin as a controlled substance (often included on the same list as cocaine and heroin), basic research has lagged far behind. Our study highlights how the poor state of knowledge of even the most fundamental scientific information can impact the use of psilocybin-containing mushrooms for recreational or therapeutic applications and undermines critical assumptions that underpin their regulation by legal authorities. Our study shows that currently available chemical studies are mainly inaccurate, irreproducible, and inconsistent, that there exists a high rate of misidentification in museum collections and public databases rendering even names unreliable, and that the concentration of psilocybin and its tryptamine derivatives in three of the most commonly collected Psilocybe species (P. cubensis, P. cyanescens, and P. semilanceata) is highly variable and unstable in museum specimens spanning multiple decades, and our study generates the first-ever insight into the highly complex and largely uncharacterized metabolomic profile for the most commonly cultivated magic mushroom, P. cubensis.
Article
Full-text available
Octocorals are major contributors of terpenoid chemical diversity in the ocean. Natural products from other sessile marine animals are primarily biosynthesized by symbiotic microbes rather than by the host. Here, we challenge this long-standing paradigm by describing a monophyletic lineage of animal-encoded terpene cyclases (TCs) ubiquitous in octocorals. We characterized 15 TC enzymes from nine genera, several of which produce precursors of iconic coral-specific terpenoids, such as pseudopterosin, lophotoxin and eleutherobin. X-ray crystallography revealed that coral TCs share conserved active site residues and structural features with bacterial TCs. The identification of coral TCs enabled the targeted identification of the enzyme that constructs the coral-exclusive capnellane scaffold. Several TC genes are colocalized with genes that encode enzymes known to modify terpenes. This work presents an example of biosynthetic capacity in the kingdom Animalia that rivals the chemical complexity generated by plants, unlocking the biotechnological potential of octocorals for biomedical applications. Marine natural products are often made by symbiotic microbes, but octocorals encode their own terpene cyclases that share features with bacterial versions and produce terpenoid precursors, including the coral-exclusive capnellane scaffold.
Article
Full-text available
Psilocybe magic mushrooms are best known for their main natural product, psilocybin, and its dephosphorylated congener, the psychedelic metabolite psilocin. Beyond tryptamines, the secondary metabolome of these fungi is poorly understood. The genomes of five species (P. azurescens, P. cubensis, P. cyanescens, P. mexicana, and P. serbica) were browsed to understand more profoundly common and species‐specific metabolic capacities. The genomic analyses revealed a much greater and yet unexplored metabolic diversity than evident from parallel chemical analyses. P. cyanescens and P. mexicana were identified as aeruginascin producers. Lumichrome and verpacamide A were also detected as Psilocybe metabolites. The observations concerning the potential secondary metabolome of this fungal genus support pharmacological and toxicological efforts to find a rational basis for yet elusive phenomena, such as paralytic effects, attributed to consumption of some magic mushrooms.
Article
Full-text available
Therapeutic use of psilocybin has become a focus of recent international research, with preliminary data showing promise to address a range of treatment-resistant mental health conditions. However, use of psilocybin as a healing entheogen has a long history through traditional consumption of mushrooms from the genus Psilocybe. The forthcoming adoption of new psilocybin-assisted therapeutic practices necessitates identification of preferred sources of psilocybin; consequently, comprehensive understanding of psilocybin-containing fungi is fundamental to consumer safety. Here we examine psilocybin producing fungi, discuss their biology, diversity, and ethnomycological uses. We also review recent work focused on elucidation of psilocybin biosynthetic production pathways, especially those from the genus Psilocybe, and their evolutionary history. Current research on psilocybin therapies is discussed, and recommendations for necessary future mycological research are outlined.
Article
Fungi produce diverse metabolites that can have antimicrobial, antifungal, antifeedant, or psychoactive properties. Among these metabolites are the tryptamine-derived compounds psilocybin, its precursors, and natural derivatives (collectively referred to as psiloids), which have played significant roles in human society and culture. The high allocation of nitrogen to psiloids in mushrooms, along with evidence of convergent evolution and horizontal transfer of psilocybin genes, suggest they provide a selective benefit to some fungi. However, no precise ecological roles of psilocybin have been experimentally determined. The structural and functional similarities of psiloids to serotonin, an essential neurotransmitter in animals, suggest that they may enhance the fitness of fungi through interference with serotonergic processes. However, other ecological mechanisms of psiloids have been proposed. Here, we review the literature pertinent to psilocybin ecology and propose potential adaptive advantages psiloids may confer to fungi.
Chapter
The fossil record is the best evidence of the characteristics of extinct species, but only a narrow range of traits fossilize or survive the fossilization process. Lacking fossil or other evidence about the past, ancestral states can be reconstructed. Three pieces of information are combined when reconstructing ancestral states: extant or known trait values (data); the evolutionary history, linking the species of interest (phylogeny); and the evolutionary model of trait change. These reconstructed ancestral states can be interpreted as our best guess as to the route evolution took, given the distribution of the trait across species, the relationships among them, and our model of evolution. Because the information we use to reconstruct the past is often not known without error, uncertainty about their true values should be accounted for when reconstructing ancestral states. In this chapter we describe how ancestral states can be reconstructed using a Bayesian framework implemented in the software BayesTraits to account for uncertainty in the phylogenetic tree and the model of evolution.
Article
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 cassette. A recently published highly contiguous reference genome derived from long read single molecule sequencing has underscored interesting variation in this Psilocybin synthesis gene cassette. 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 genomes 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 an alternative pathway and may illuminate previously proposed convergent evolution of Psilocybin synthesis.