ArticlePDF Available

DNA Authentication and Chemical Analysis of Psilocybe Mushrooms Reveal Widespread Taxonomic Misdeterminations and Inconsistencies in Metabolites

Authors:

Figures

Content may be subject to copyright.
DNA Authentication and Chemical Analysis of Psilocybe
Mushrooms Reveal Widespread Misdeterminations in Fungaria
and Inconsistencies in Metabolites
Alexander J. Bradshaw,a,b Talia A. Backman,aVirginia Ramírez-Cruz,dDale L. Forrister,aJaclyn M. Winter,cLaura Guzmán-Dávalos,e
Giuliana Furci,fPaul Stamets,gBryn T. M. Dentingera,b
a
School of Biological Sciences, University of Utah, Salt Lake City, Utah, USA
b
Natural History Museum of Utah, University of Utah, Salt Lake City, Utah, USA
c
Department of Medicinal Chemistry, University of Utah, Salt Lake City, Utah, USA
d
CONACYT-UdeG, Departamento de Botánica y Zoología, Universidad de Guadalajara, Zapopan, Jalisco, Mexico
e
Departamento de Botánica y Zoología, Universidad de Guadalajara, Zapopan, Jalisco, Mexico
f
Fungi Foundation, Brooklyn, New York, USA
g
Fungi Perfecti LLC Laboratories, Shelton, Washington, USA
ABSTRACT The mushroom genus Psilocybe is best known as the core group of psy-
choactive 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 sta-
bility of psilocybin, psilocin, and their related tryptamine alkaloids in 25 specimens
across the most commonly vouchered species (Psilocybe cubensis,Psilocybe cyanes-
cens, 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 reposito-
ries, 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 uncharacter-
ized 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 identied 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 mush-
rooms,is revolutionizing mental health care for a number of conditions, including
depression, posttraumatic stress disorder (PTSD), and end-of-life care. This has spot-
lighted the current state of knowledge of psilocybin, including the organisms that endo-
genously produce it. However, because of international regulation of psilocybin as a con-
trolled 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 scientic information can impact the use of psilocybin-containing
mushrooms for recreational or therapeutic applications and undermines critical assump-
tions 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 misidentication in museum collections and public databases
Editor Irina S. Druzhinina, Royal Botanic
Gardens
Copyright © 2022 Bradshaw et al. This is an
open-access article distributed under the terms
of the Creative Commons Attribution 4.0
International license.
Address correspondence to Alexander J.
Bradshaw, alexander.j.bradshaw@gmail.com, or
Bryn T. M. Dentinger,
bdentinger@nhmu.utah.edu.
The authors declare a conict of interest. Paul
Stamets is the current CEO of Fungi Perfecti
LLC, Co-founder of MycoMedica Life Sciences,
PBC and has patents pending and approved
using psilocybin. All other authors report no
conicting interests in the body or intent of
this work.
Received 31 August 2022
Accepted 6 November 2022
Month YYYY Volume XX Issue XX 10.1128/aem.01498-22 1
GENETICS AND MOLECULAR BIOLOGY
Downloaded from https://journals.asm.org/journal/aem on 06 December 2022 by 155.98.229.65.
rendering even names unreliable, and that the concentration of psilocybin and its trypt-
amine derivatives in three of the most commonly collected Psilocybe species (P. cubensis,
P. cyanescens,andP. semilanceata) is highly variable and unstable in museum specimens
spanning multiple decades, and our study generates the rst-ever insight into the highly
complex and largely uncharacterized metabolomic prole for the most commonly culti-
vated magic mushroom, P. cubensis.
KEYWORDS mycology, Psilocybe, fungaria, metabolomics
Mushrooms belonging to the genus Psilocybe (Fr.) P. Kumm. have become a major
scientic and public curiosity since 1957 when R. Gordon Wasson published a
popular account of his rst-hand experience with psychedelic mushrooms in Life mag-
azine (1). Although species of Psilocybe (the rst of which was originally described as
Agaricus semilanceatus Fr.) have been known to the scientic community since the
early 19th century (2), Wassons article was the rst to make their psychoactive proper-
ties widely known. However, the ceremonial use of the psychoactive mushrooms has
been practiced by the indigenous peoples of Mesoamerica for centuries.
The use of psychoactive mushrooms is reportedly depicted in petroglyphs and murals
found in Siberia, Africa, and Spain, demonstrating a long-held interest in these organisms
across the history of humankind (3). Further, in the 16th century, the Spanish Franciscan
Friar Bernardino de Sahagún was the rst European to document the ceremonial use of
psychoactive mushrooms in Mexico (47).Sahagún,havingspent50yearsthere,began
studying the language and culture of the Aztec, and much of this work was published in
his Historia general de las Cosas de la Nueva España, in which he refers to psychoactive
mushrooms as teonanácatl(later updated to teotlaquilnanácatlas the correct Nahuatl
word), meaning Godsesh,the sacred mushrooms of Mesoamerica. While being rst
recorded by Europeans and Western scientists in the Nahuatl population, ritual usage of
psychoactive mushrooms has occurred in indigenous Mexican cultures including the
Chatins, Chinantecs, Matlazincs, Mazatecs, Mixes, Purepechs, Totonacs, and Zapotecs, with
names for these mushrooms specictoeachculture(7,8).
Despite psychoactive mushrooms being known and used in Mesoamerica for centu-
ries, the publicity from Wassons article spurred widespread use of so-called magic
mushroomsas recreational drugs during the psychedelic era of the 1960s and 1970s,
culminating in the criminalization of psilocybin and its analogs (and by extension the
organisms that produce them) as part of the war on drugspromulgated by Richard
Nixon in 1971 (9). Based on assumptions about the universal presence of these chemi-
cals in all Psilocybe spp., it soon became illegal to possess putative psilocybin-contain-
ing mushrooms in numerous countries, hobbling research on their diversity, biology,
and unique psychoactive properties (10).
Recently, a renewed interest in psilocybin and Psilocybe spp. has emerged due to
increasing evidence that psilocybin is highly effective in treating a variety of mental
health problems, including anxiety and depression (1117). This has led to a resur-
gence in clinical studies, many of which have demonstrated the therapeutic efcacy of
psilocybin (18), as well as basic research that led to the identication of the enzymes
and underlying genes responsible for psilocybin and psilocin biosynthesis (19, 20).
Despite these recent advances, fundamental knowledge of Psilocybe, including its di-
versity, distribution, taxonomy, ecology, and chemistry, remains lacking or in dire need
of update. In fact, even the functional role of psilocybin and psilocin under natural con-
ditions remains a mystery (21, 22).
The genus Psilocybe sensu lato has been estimated to contain between 277 and 326
known species (2325), of which 144 exhibit a characteristic inducible bluingreaction
when damaged. This blue pigment is caused by the enzymatic oligomerization of psi-
locin and indicates the presence of the psychoactive compounds psilocybin and psilo-
cin (2224). It is now accepted that most of the non-bluing species likely belong in the
genus Deconica (W.G. Sm.) P. Karst., and none of them are known to synthesize psilocy-
bin or psilocin (26, 27), and yet the transfer of these species from Psilocybe is largely
DNA and Chemical Analysis of Psilocybe Mushrooms Applied and Environmental Microbiology
Month YYYY Volume XX Issue XX 10.1128/aem.01498-22 2
Downloaded from https://journals.asm.org/journal/aem on 06 December 2022 by 155.98.229.65.
incomplete (2832) (see Supplementary Information 1). In addition to the ongoing
renement of Psilocybe taxonomy, many species are considered rare, with many having
been collected only once, leading to poor specimen representation in fungaria (equal
to herbaria for Fungi; Supplementary Information 1) (3335). This makes identication
difcult without highly detailed and complete taxonomic keys (30, 36, 37) or impossi-
ble because of the dearth of specimens for authentication and poor representation
in DNA sequence databases. This situation is not uncommon in mycology, where
characters useful for eld identication are often insufcient, resulting in rampant
misidentication that cannot be resolved without molecular data (38, 39). On top of
this, ambiguities about legal requirements in several countries for their storage, care,
and sharing exacerbate the problem by further suppressing research, creating a chal-
lenging situation for researchers and legal authorities alike.
Although Psilocybe specimens may contain controlled substances, no evidence
exists on their consistency, and psilocybin has been empirically identied in only ;45
accepted Psilocybe spp., which are only 32% of the ;140 putative bluing species (26,
35, 40, 41). Moreover, even for the species examined, the accuracy and reliability of
these reports are dubious due to a lack of standards for taxonomic identication
(including deposition of vouchers to allow for reexamination), lack of peer review for
some studies, and inconsistency of methods used for chemical analysis (see Table S1 in
the supplemental material) (40, 42). In addition, no data on the preservation of psilocy-
bin/psilocin in fungarium specimens exist, further obfuscating if and to what degree
current regulation of controlled substances applies to them, especially if these com-
pounds actively degrade or have become unmeasurable during storage. This absence
of a common set of methods, the almost complete lack of voucher-based studies, and
the deciency of information on the resilience of the controlled substances under
standard fungarium storage conditions render existing information unreliable and
highly suspect, with the result that there is actually little authoritative information on
the chemistry of most Psilocybe spp.
Psilocybe mushrooms may produce a diversity of metabolites in addition to psilocybin
and psilocin, including compounds such as bioactive
b
-carbolines in Psilocybe cubensis
(Earle) Singer and Psilocybe mexicana R. Heim that may provide monoamine oxidase-inhibi-
ting roles and contribute synergistically to the activity of psilocybin (43). Yet, much of the
metabolome for most mushroom species remains uncharacterized. The metabolomes of a
few Psilocybe species have been investigated recently, and genetic evidence for the biosyn-
thesis of additional bioactive compounds was reported (44). Thus, the potential for other
bioactive compounds within the metabolome of Psilocybe spp. to have synergistic effects
when the mushrooms are consumed as whole organisms is very real, and this information
will be important for potential future implementation of whole-organism-based therapies,
such as those being developed in Oregon, USA (14, 45). A more comprehensive view of
the metabolome of Psilocybe spp. is therefore urgently needed to mitigate any potential
unanticipated bioactivity or adverse side effects of whole-organism consumption, as well
as to provide a baseline for understanding the possible benets of a whole-organism
approach over psilocybin as a sole therapeutic.
In this study, we aimed to establish a baseline of current knowledge of Psilocybe
identity and chemistry in public repositories, including vouchered specimens, DNA
sequence databases, and the published literature. We then set out to provide an
update to this knowledge with new insights on the chemical stability of psilocybin and
its tryptamine intermediates in collections over time and to acquire a birds-eye view
of the metabolome of the most widely cultivated psychedelic mushroom, P. cubensis.
RESULTS
Fungarium specimens and DNA extraction. We curated a database of Psilocybe
museum voucher records utilizing MycoPortal, the most inclusive online aggregator of
fungal voucher specimen records (see online supplemental data) (46). After ltering
out observations (which are mostly not linked to a physical voucher; n= 7,462) and
DNA and Chemical Analysis of Psilocybe Mushrooms Applied and Environmental Microbiology
Month YYYY Volume XX Issue XX 10.1128/aem.01498-22 3
Downloaded from https://journals.asm.org/journal/aem on 06 December 2022 by 155.98.229.65.
non-Psilocybe species under older names, as well as reconciling synonyms, the nal
database consisted of 3,707 records of physical vouchers across 72 institutions found
on MycoPortal. Many institutions either could not or refused to provide loan material
due to the legal ambiguity of Psilocybe specimens. Despite the large number of vouch-
ers present across these institutions, approximately one-third were identied only to
genus level (1,151/3,707). Most of the specimens identied to species level were domi-
nated by a few well-known taxa, i.e., P. cubensis (319/3,707), Psilocybe cyanescens
Wakef. (100/3,707), and Psilocybe semilanceata (Fr.) P. Kumm. (138/3,707). Of the
;3,700 specimens requested, ;1,200 were received from 35 institutions.
A total of 18 species names across 94 specimens were used for DNA barcoding.
Across all samples, DNA extractions yielded 5.5 to 166.1 ng/
m
L with 260:230 values
between 0.22 and 6.52 and with 260:280 values ranging from 1.29 to 2.04 (see Table S2
in the supplemental material). Our best extraction results usually came from larger ma-
terial inputs between 10 and 15 mg.
Phylogenetic analysis. In our analysis of all publicly available sequences identied as
Psilocybe spp. from the National Center for Biotechnology Information (NCBI), we found a
large number of likely incorrect or inconsistent names and sequences that of were very
poor quality or that are highly divergent from mushroom-forming fungi. Four hundred
twenty-three internal transcribed spacer (ITS) sequences in NCBI were labeled as Psilocybe
or Deconica and other labels (when Psilocybe was included as a name in the sequence re-
cord), including one sequence labeled Stropharia rugosoannulata Farl. ex Murrill and one
sequence labeled Uncultured fungus.Thirty-two of 39 sequences (82%) labeled as
Deconica had specic epithets (9 species names), and six did not. Three hundred sixteen of
382 sequences (83%) labeled as Psilocybe had specic epithets (69 species names), and 66
did not. The most abundant Psilocybe species labels were P. cubensis (31 sequences; 10%),
P. cyanescens (25 sequences; 8%), and P. semilanceata (20 sequences; 6%). Fifty-seven
sequences labeled as Psilocybe were attributable to Deconica (Fig. S1).
A number of erroneous sequences were encountered in the NCBI database that will
require either removal due to poor sequence quality or artifacts, or updates to the sequence
identity. Some of these (MN510675 to MN510688 and MG969992)hadmanyambiguous
bases and potential errors that introduced excessive numbers of autapomorphies, leading to
exceedingly long branches in the phylogenetic tree. Two sequences labeled as Psilocybe
cubensis (ON415277 and ON415278)areinsteadFusarium spp. (pathogenic fungi distantly
relatedtomushrooms)basedonBLASTresults.Onesequence(
DQ900972)fromanenviron-
mental barcoding study (47) appears to be chimeric, with some portions of the sequence
potentially being random and other parts matching Amoebozoa rather than Fungi based on
BLAST. These sequences contribute nothing to species identication and are likely only to add
more confusion, and we recommend removing them from the public records. One sequence
labeled Psilocybe montana (Pers.) Kumm. (which should be updated to the current name
Deconica montana (Pers.) P. D. Orton; AY129360) clustered with P. cubensis. Sequences labeled
as P. cubensis were recovered in six clades. Some of these were other species of Psilocybe (e.g.,
P. mexicana and P. chuxiongensis T. Ma & K.D. Hyde) whereas others belong to Deconica
(AY129351)orGalerina (AY281023). A specimen from the Universidad de Guadalajara (IBUG)
identied as Psilocybe cubensis that was newly sequenced clustered with sequences labeled as
Stropharia rugosoannulata. Similar to sequences labeled as P. cubensis, sequences labeled as P.
cyanescens were recovered in three clades, two of which correspond to Panaeolus (EU029946)
and Deconica (MF467896). Fifty-ve sequences labeled as Psilocybe (including labels with spe-
cic epithets, no epithets, and uncultured) clustered with sequences labeled as Deconica.
Some phylogenetic relationships between Psilocybe species are recovered for the rst
time. Sequences of P. cubensis were most closely related to a sequence labeled as
Psilocybe natalensis Gartz, D.A. Reid, M.T. Sm. & Eicker (89% rapid bootstrap support [BS]),
together sharing a most recent common ancestor (MRCA) with P. chuxiongensis (100%
BS) (Fig. S1). However, the sequence labeled as P. natalensis in GenBank (OK491080)is
not derived from the type and is of uncertain provenance. The GenBank record refers to
an observation on an online forum (https://mushroomobserver.org/472561) and lists
DNA and Chemical Analysis of Psilocybe Mushrooms Applied and Environmental Microbiology
Month YYYY Volume XX Issue XX 10.1128/aem.01498-22 4
Downloaded from https://journals.asm.org/journal/aem on 06 December 2022 by 155.98.229.65.
the sequence isolation source as Cultivated from a spore printand the country of
origin as South Africa,but no voucher specimen is indicated, and it is unclear if the
data are reproducible or the source can be reexamined. Overall, the inclusion of NCBI
sequences resulted in a large amount of noise in our analysis, which intensied con-
fusion and contributed to a severe lack of clarity in sequence-based identication of
Psilocybe spp.
Due to the large number of misidentications and the noise in our Psilocybe-only
analysis, we performed a second phylogenetic analysis utilizing species hypothesis
(SH) sequences from the family Strophariaceae and the closely related genera
Gymnopilus P. Kumm. and Agrocybe Fayod from UNITE (48). Of the 94 new ITS sequen-
ces generated, ve have no comparable SH in the UNITE fungal ITS database (Table
S2). Only 48 unique SH sequences were found in UNITE: 13 lacked species-level identi-
cation, and two are incorrectly assigned (Fig. 1). Like the NCBI records, the UNITE data-
base also includes outdated taxonomic names such as Psilocybe coprophila (Bull.) P.
Kumm. and Psilocybe subviscida (Peck) Kauffman, both of which belong to the nonpsy-
choactive genus Deconica.
Our phylogenetic analysis revealed Psilocybe sensu stricto to be a monophyletic
group (98% BS) sharing a most recent common ancestor with species of the genus
FIG 1 Phylogenetic tree of Strophariaceae and Psilocybe. (Right) Phylogenetic tree of museum specimen ITS sequences along with all species hypothesis
(SH) sequences for the family Strophariaceae and the genera Agrocybe and Galerina from the UNITE fungal sequence database. Correct identications
within the genus Psilocybe are highlighted in blue, misidentications at the genus level are highlighted in red, and misidentications at the species level
are highlighted in orange. (Left) Visual representation of the three main species used for this study, Psilocybe cubensis (photo provided by F. Landeros) (A),
Psilocybe cyanescens (photo by B. Dentinger) (B), and Psilocybe semilanceata (photo by P. Stamets) (C).
DNA and Chemical Analysis of Psilocybe Mushrooms Applied and Environmental Microbiology
Month YYYY Volume XX Issue XX 10.1128/aem.01498-22 5
Downloaded from https://journals.asm.org/journal/aem on 06 December 2022 by 155.98.229.65.
Agrocybe (Fig. 1), in contrast with previous work that reported other members of
Strophariaceae as related taxa (49). Approximately 31 moderately to highly supported
clades (i.e., species) of Psilocybe are currently represented by ITS sequences. Thirteen of
the 94 specimens newly sequenced and two SH sequences from UNITE that had been
identied as Psilocybe clustered outside the Psilocybe sensu stricto clade. These speci-
mens were phylogenetically placed within clades corresponding to the genera
Agrocybe,Deconica,Galerina Earle, Stropharia (Fr.) Quél., and Tubaria (W.G. Sm.) Gillet.
Additionally, multiple groupings within Psilocybe had incongruent taxonomic names,
which suggests either misidentication or possible synonyms (Fig. 1).
Although taxonomic inconsistencies occurred in every clade, some clades contained more
divergent taxonomic names than others. Groupings within the sister clade to Tubaria included
eight different generic or family-level-only identications [Pachylepyrium Singer, Phaeomarasmius
Scherff., Pholiota (Fr.) P. Kumm., Pleuroammula Singer, Psilocybe,Mythicomyces Redhead & A.H.
Sm., Strophariaceae,andTubaria]. Additionally, the branch belonging to Deconica included two
SH sequences attributed to Psilocybe, six of our vouchered specimens, and eight SH sequences
labeled Strophariaceae species SH sequences (Fig. 1). Agrocybe and Galerina were recovered as
paraphyletic, with a single branch (63% BS) that also included two of our vouchered specimens
labeled as Psilocybe and SH sequences labeled as Pholiota spp.
Tryptamine metabolite proling in voucher specimens. We chose specimens of
the three most commonly vouchered species (P. cubensis,P. cyanescens,andP. semilan-
ceata) for chemical analysis and taxonomically veried each using full-length ITS barcoding
that corresponded to the SH for their respective voucher identication. A single specimen
of P. cubensis was misidentied (IBUG-10258, equals Stropharia rugosoannulata), and one
specimen of P. cyanescens was misidentied (SFSU-F-029946, equals Psilocybe allenii Borov.,
Rockefeller & P.G. Werner) (Table 1 and Table S2). The quantities of all metabolites analyzed
varied substantially between species (Fig. 2). Very few fungarium specimens of P. cubensis
showed detectable levels of psilocybin or psilocin compared to most dried specimens of
TABLE 1 Specimen voucher table of Psilocybe used for chemical analysis
Institution code
a
Accession no. Voucher taxonomic designation Collection yr
WTU WTU-F-054871 Psilocybe cubensis 1970
WTU WTU-F-011258 Psilocybe cubensis 1974
WTU WTU-F-054869 Psilocybe cubensis 1978
IBUG 2924 Psilocybe cubensis 1986
IBUG 13422 Psilocybe cubensis 1990
IBUG 14027 Psilocybe cubensis 1995
IBUG 11267 Psilocybe cubensis 2000
NY 1901145 Psilocybe cubensis 2003
WTU WTU-F-011538 Psilocybe cyanescens 1972
WTU WTU-F-011519 Psilocybe cyanescens 1978
WTU WTU-F-011079 Psilocybe cyanescens 1982
SFSU SFSU-F-029949 Psilocybe cyanescens 1993
WTU WTU-F-011523 Psilocybe cyanescens 1999
SFSU SFSU-F-029967 Psilocybe cyanescens 2007
UBC F24013 Psilocybe cyanescens 2011
UBC F30955 Psilocybe cyanescens 2012
UBC F32228 Psilocybe cyanescens 2015
SFSU SFSU-F-029972 Psilocybe semilanceata 1973
WTU WTU-F-055015 Psilocybe semilanceata 1979
UBC F15293 Psilocybe semilanceata 1999
UBC F17025 Psilocybe semilanceata 2007
UBC F22104 Psilocybe semilanceata 2010
UCSC UCSC-F-00856 Psilocybe semilanceata 2011
UBC F32307 Psilocybe semilanceata 2013
FLAS FLAS-F-63134 Psilocybe semilanceata 2015
a
WTU, University of Washington Herbarium; IBUG, Universidad de Guadalajara; NY, New York Botanical Garden;
SFSU, Harry D. Thiers Herbarium, San Francisco State University; UBC, University of British Columbia Herbarium;
UCSC, University of California, Santa Cruz; FLAS, University of Florida Herbarium.
DNA and Chemical Analysis of Psilocybe Mushrooms Applied and Environmental Microbiology
Month YYYY Volume XX Issue XX 10.1128/aem.01498-22 6
Downloaded from https://journals.asm.org/journal/aem on 06 December 2022 by 155.98.229.65.
P. cyanescens, which had detectable levels of both compounds, albeit reduced, even after
nearly50yearsofstorage(Fig.2).
Concentrations of psilocybin and psilocin, when detectable at all, varied in P. cuben-
sis voucher specimens from 0.55 to 1.9
m
g/mg (dry weight), with only a single sample
having detectable amounts of psilocin at 0.44
m
g/mg (dry weight). Psilocybe cyanescens
specimens contained relatively more and consistently higher alkaloid concentrations,
with psilocybin content ranging from 3.0 to 15.6
m
g/mg (dry weight) and psilocin con-
tent ranging from 0.22 to 5.22
m
g/mg (dry weight). Psilocybe semilanceata specimens
contained psilocybin content ranging from 0.33 to 15.77
m
g/mg (dry weight) and psilo-
cin content ranging from 0.1 to 3.88
m
g/mg (dry weight) (Fig. 2). Specimen age was
poorly correlated with quantity of alkaloids. Alkaloids were low or nearly undetectable
in some specimens of P. cubensis and P. cyanescens ,20 years old but also detectable
at multiple-microgram-per-milligram concentrations in nearly 50-year-old specimens
of P. cyanescens. Older specimens (>20 years) of P. semilanceata had barely detectable
levels of the indole alkaloids.
Psilocybin and psilocin metabolite content in cultivated P. cubensis.Psilocybin
was the major alkaloid recovered from mycelium (6.44
m
g/mg), caps (10.5 to 20
m
g/mg),
and stipes (15.44 to 18.44
m
g/mg) (Table 2). Psilocin was a minor component in all sam-
ples (0.44 to 2.0
m
g/mg). In comparison, psilocybin-to-psilocin ratios ranged from ;5:1
to 47:1 in the caps and ;7.5:1 to 37:1 in the stipes (Table 2). Performing an unpaired t
test, we found no signicant difference in psilocybin or psilocin between whole individ-
ual sporocarps (Psi 5 and Psi 6) from the same reproductive event in a single growth bin
(psilocybin percentage point [pp] = 0.1349 and psilocin pp = 0.4756). Unpaired ttests
also showed no signicant difference in concentrations between the same anatomical
tissues (caps and stipes) or between the two sporocarps Psi 5 and Psi 6 (psilocybin
pp = 0.5081 and psilocin pp = 0.4563) or measured across Psi 4, Psi 5, and Psi 6 (psilocy-
bin pp = 0.8182 and psilocin pp = 0.6898). However, at the full sporocarp level, quantities
Psilocin Psilocybin Tryptophan Baeocystin Norbaeocystin
0
5
10
15
20
50
40302010
Psilocybe cubensis
Compound Concentration (ug/mg)
Specimen Age
0
5
10
15
20
Psilocybe cyanescens
Compound Concentration (ug/mg)
Specimen Age
0
5
10
15
20
Psilocybe semilanceata
Compound Concentration (ug/mg)
Specimen Age
50
40302010 50
40302010
FIG 2 Content of tryptophan, psilocybin, and related alkaloids in fungarium specimens of three species of Psilocybe over 5 decades. Species are Psilocybe
cubensis (left), Psilocybe cyanescens (center), and Psilocybe semilanceata (right). Specimen age (years) is on the xaxis, and concentration of metabolites is on
the yaxis in micrograms per milligrams (dry weight). Points indicate concentrations of individual compounds: tryptophan (green), baeocystin (blue),
norbaeocystin (purple), psilocybin (yellow-green), and psilocin (red).
DNA and Chemical Analysis of Psilocybe Mushrooms Applied and Environmental Microbiology
Month YYYY Volume XX Issue XX 10.1128/aem.01498-22 7
Downloaded from https://journals.asm.org/journal/aem on 06 December 2022 by 155.98.229.65.
of psilocybin and psilocin were statistically different between growth bins (psilocybin
pp = 0.0560 and psilocin pp = 0.0389).
Metabolomic proling and tissue comparison of cultivated material. We identi-
ed 1,465 distinct spectral features based on their unique mass-to-charge (m/z) ratio and
retention time. Using hierarchical clustering to compare the intensity proles of each of
our P. cubensis samples across the two growing environments, we found that metabolic
proles of caps and stipes were most like one another, with the mycelium being most dif-
ferent from all other samples, regardless of which container they were cultivated in (Fig. 3).
However, in the case of both caps and stipes, Psi 5 and Psi 6 were shown to be more alike,
while Psi 4 formed an outgroup for each tissue type (Fig. 3). We investigated this further by
performing a principal-component analysis (PCA) on the intensity of each feature and par-
tial least-squares discriminant analysis (PLS-DA) for each sample (Fig. S3 and S4). PCA
showed largely overlapping spectral intensities, but the same pattern as that in the hier-
archical clustering was present, with dimension 1 accounting for 15.6% of the variability
and dimension 2 accounting for 63.3% of the variability (Fig. S3). PLS-DA on the full proles
for each sample, rather than the individual spectral features, again showed a pattern of
caps and stipes being most similar but distinct separation between samples Psi 4 (growth
bin 1) and Psi 5 and Psi 6 (growth bin 2), even though all samples were cultivated and har-
vested in the same manner.
In addition to looking at broad-scale patterns of metabolomic proles, we also wished to
investigate if any of the identied spectral features could be annotated to known compounds.
We collected tandem mass spectrometry (MS/MS) data for sample Psi-6-Cap for comparison to
the Global Natural Products Social Molecular Networking database (GNPS; https://gnps.ucsd
.edu/ProteoSAFe/static/gnps-splash.jsp); (50). Using the GNPS spectral library matching the com-
pound database provides an annotation along with a cosine score (MQscore) that indicates
spectral similarity to the annotated compound. An MQscore of 1 suggests an identical match of
a spectral feature to the annotated compound in the database, with decreasing scores suggest-
ing compounds of similar structure as the annotation. Spectral library matching for Psi-6-Cap
revealed 30 features that could be annotated to known compounds (0.02% of the identied
1,465 features) (Table S5). Many of these annotations belonged to compounds involved in cell
membrane formations, sugars, and cholesterol (sn-glycero-3-phosphocholine MQscore = 0.97,
sucrose MQscore = 0.93, and epicoprostanol MQscore = 0.92), which are likely to be found in
high quantities because these proles were derived from freshly preserved specimens. However,
three annotated spectral features were of particular interest as they had structural similarity to
serotonin (MQscore = 0.9292), cocamidopropyl betaine (CAPB) (MQscore = 0.9191), and traza-
done hydrochloride (MQscore = 0.73).
DISCUSSION
Our study shows that fungarium specimens of Psilocybe spp. are regularly misidenti-
ed, little has been done to investigate the stability of the known psychoactive com-
pounds under standard storage conditions, and the chemical complexity of these
organisms has been overlooked by an almost exclusive focus on the principal psycho-
active agents. These deciencies cast a shadow on what little information is available
TABLE 2 Psilocybin and psilocin concentrations in cultivated sporocarps of Psilocybe cubensis
Sample
identier
UT-M catalog
no.
Amt (mg) sampled of
voucher specimen
Psilocybin concn
(mg/mg)
Psilocin concn
(mg/mg) Source (origin of sample)
Psi-mycelium UT-M0001771 7 6.3 0.01 Stock mycelium culture
Psi-4-Cap UT-M0001772 7 10.5 2.11 Single sporocarp from one growth bin
Psi-4-Stipe UT-M0001772 8 15.44 1.9 Single sporocarp from one growth bin
Psi-5-Cap UT-M0001773 7 17.3 0.5 One of two sporocarps from one growth bin
Psi-5-Stipe UT-M0001773 9 16.33 0.4 One of two sporocarps from one growth bin
Psi-6-Cap UT-M0001774 9 19.9 0.44 Two of two sporocarps from one growth bin
Psi-6-Stipe UT-M0001774 7 18.3 1.5 Two of two sporocarps from one growth bin
DNA and Chemical Analysis of Psilocybe Mushrooms Applied and Environmental Microbiology
Month YYYY Volume XX Issue XX 10.1128/aem.01498-22 8
Downloaded from https://journals.asm.org/journal/aem on 06 December 2022 by 155.98.229.65.
on Psilocybe and underscore the increasingly urgent need for comprehensive basic
research on them.
Authentication and diversity of Psilocybe fungarium specimens. Fungi often ex-
hibit cryptic morphology and can be difcult to identify, even for highly trained experts.
Cryptic traits, lack of extensive training, and even clerical errors in databasing can all con-
tribute to misidentication(38,51).Inthissense,itislikelythatspecimensinfungariaof
nearly all taxa suffer from some level of taxonomic uncertainty, but the extent to which
misidentication is a problem has been rarely investigated. A dearth of authentic DNA
barcodes for Psilocybe and other closely related groups further exacerbates the problem.
Without DNA barcodes from type or other authentic material, names can be applied incon-
sistently, leading to different applications of species names to modern collections. This also
makes it difcult to know when a specimen represents a new species, resulting in further
misapplication of named species to novel taxa. We found that close to 13% of the speci-
mens we studied that were cataloged as Psilocybe belonged to other genera, none of
which are known to contain controlled substances. This result challenges the use of names
in collections for meta-analysis or evidence-based decisions on their custody and care.
Many publicly available sequences are incorrectly labeled or are of very poor quality, fur-
ther aggravating the problem for identifying specimens or other samples using a DNA bar-
coding approach. Even some of the most recognizable and frequently collected species (P.
cubensis and P. cyanescens) had sequences that belonged to other related genera and in
two cases to fungi from a different phylum (Ascomycota). Some of these are likely misiden-
tication of the source material, including cultures, highlighting the need for an authorita-
tive, expert-curated reference database of DNA sequences derived from types or other
authentic specimens.
While misidentications in museum collections and public DNA repositories are not
uncommon (38, 39), the ambiguous legal status of Psilocybe specimens makes misidenti-
cations detrimental not only to scientic research but also to properly applying regulatory
policies. Misidentication of putative psilocybin-producing mushroom specimens may lead
to an unnecessary effort and poor use of resources, or even miscarriage of justice in legal
cases.
A review of the currently known species of Psilocybe suggests that 165 belong to
FIG 3 Comparison of chemical proles across cultivated Psilocybe cubensis samples. Three sporocarps
and one mycelium were used for chemical proling. Sporocarps were harvested simultaneously from
the same growth bin (Psi 5 and 6) and from a second growth bin (Psi 4) grown from a single
mycelial culture. Hierarchical clustering of samples (yaxis, left) and spectra (xaxis, top) based on
Euclidean distance similarities of total chemical prole and spectral feature similarity, respectively.
Heatmap shows chemical factor contribution using the intensity values for each identied factor
represented by a color gradient from blue (low contribution) to red (high contribution).
DNA and Chemical Analysis of Psilocybe Mushrooms Applied and Environmental Microbiology
Month YYYY Volume XX Issue XX 10.1128/aem.01498-22 9
Downloaded from https://journals.asm.org/journal/aem on 06 December 2022 by 155.98.229.65.
Psilocybe sensu stricto while 98 belong to Deconica or other genera (see Supplementary
Information 1 in the supplemental material). Fifty-two species that likely belong to
Deconica are currently still classied in Psilocybe. However, these likely include synonyms
and cryptic species that are presently unknown and do not account for undocumented
species that likely exist. Only ve species of Psilocybe sensu stricto are represented by ITS
barcodes derived from types. Critically, sequences from types of the best-known species,
such as Psilocybe cubensis, do not exist. This fundamental information is critical for accu-
rate application of names and should be considered an urgent priority to aid in the iden-
tication of this medically and societally important group.
While our data set had limited representation of the known species of Psilocybe,new
phylogenetic patterns were uncovered. While Ramírez-Cruz et al. (34) reported that the
American species P. cubensis shared a most recent common ancestor (MRCA) with the
southeast Asian species Psilocybe thaiaerugineomaculans Guzmán,Karun.&Ram.-Guill.,our
inclusion of the South African P. natalensis and the more recently described Asian species P.
chuxiongensis revealed that P. cubensis likely shares an MRCA with P. natalensis and not the
Asian species. This is consistent with the African-origin hypothesis for P. cubensis and/or its
progenitor proposed by Guzmán et al. (52) and Froese et al. (53) rather than the Asian-origin
hypothesis suggested by the relationships to Asian species in the work of Ramírez-Cruz
et al. (34) and Ma et al. (54). However, it should be noted that the sequences of P. natalensis
and P. thaiaerugineomaculans are not derived from type specimens (which is also true for P.
cubensis), and so while the geographic implications remain, the taxonomic ones are uncer-
tain. To gain a better understanding of the geographic origins of Psilocybe spp., especially
the enigmatic P. cubensis,neweldwork to expand documentation of Psilocybe in Africa
andAsiashouldbeapriority.
Psilocybin, psilocin, and related metabolites in Psilocybe spp. Decades of incon-
sistency in the identication of Psilocybe and methods for measuring its metabolites render
most previous studies unreliable and unreproducible. In our review of the published litera-
ture on alkaloid chemistry in Psilocybe, there was an almost complete lack of thorough sam-
ple identication, few studies designated vouchers, and literature citations frequently con-
tained incorrect information or opposing statements (see Table S1 in the supplemental
material). In a few instances, reports for some species, such as Psilocybe ovoideocystidiata
Guzmán & Gaines and Psilocybe kumaenorum R. Heim, lacked peer review or were not acces-
sible. Although our focus was specically on literature pertaining to Psilocybe,psilocybinhas
been reported from other species such as Agrocybe praecox (Pers.) Fayod (55, 56), although
there is good reason to believe this may have been a misidentication of Psilocybe subcu-
bensis Guzmán (57). Unfortunately, because no vouchers were deposited, it is impossible to
reexamine the material to determine the true identity. Nonetheless, this dubious report has
persisted through repetition without critical review or scrutiny of the data. A much more rig-
orous set of standards is needed to establish a reliable baseline for information on the
chemistry of psilocybin and related alkaloids in mushrooms.
We chose to quantify psilocybin and related metabolites in fungarium specimens of
Psilocybe to determine their presence and stability over time. One recent study showed
that the method of drying and extent of light exposure can affect psilocybin/psilocin
stability in P. cubensis (58), but this examined storage over only a single year. Our speci-
mens from three species spanned multiple decades, and we found that metabolite
content and stability varied widely and metabolite content was undetectable in most
samples stored for more than 40 years (Fig. 2). However, the sample size for any given
specimen age was small, and information on confounding factors, such as the original
preservation method, was often not available. In one instance, we were able to mea-
sure the chemical content of a specimen of P. cubensis (WTU-F-054869) deposited
around the same time as specimens used for previous chemical analysis (26, 59, 60).
Although there is a strong possibility that this specimen was the same one used in the
original study or was harvested from the same cultivated strain around the same time
(M. Beug, personal communication), no voucher information was available from the
original study and so it was not possible to conrm its identity. Bigwood and Beug (60)
DNA and Chemical Analysis of Psilocybe Mushrooms Applied and Environmental Microbiology
Month YYYY Volume XX Issue XX 10.1128/aem.01498-22 10
Downloaded from https://journals.asm.org/journal/aem on 06 December 2022 by 155.98.229.65.
noted that multiple ushes of P. cubensis contained over 5.0
m
g/mg (dry weight) of psi-
locybin. However, when we measured the WTU-F-054869 voucher, now 42 years old,
we detected less than half of their reported psilocybin content (1.9
m
g/mg) and no psi-
locin (Fig. 2). This provides further evidence that these chemicals are highly unstable in
preserved specimens of P. cubensis, degrading either in a time-dependent manner or
possibly during or soon after the initial preservation. A study designed to specically
test time-dependent degradation with replicates and controls would be needed to
determine the exact causes of the metabolite instability we observed.
Our results also reveal taxon-dependent patterns of metabolite instability. For example,
psilocybin and psilocin in P. cyanescens had a comparably lower rate of degradation over
time than in P. cubensis, with the youngest specimen having 15.6
m
g/mg psilocybin and
5.8
m
g/mg psilocin and the oldest (40-year) specimen having 6.11
m
g/mg psilocybin and
3.66
m
g/mg psilocin. Similarly, in P. semilanceata the youngest sample contained 11.2
m
g/
mg of psilocybin and 3.88
m
g/mg psilocin, while the oldest contained 0.2
m
g/mg psilocybin
and no detectable psilocin. One possible explanation for the generally better chemical pres-
ervation in P. cyanescens and P. semilanceata is that these species have sporocarps that are
comparatively lower in biomass than those of P. cubensis.Thismayresultinamorerapid
desiccation during heat-assisted drying, minimizing the opportunity for spontaneous and/or
enzymatic degradation of psilocybin and psilocin. Another possible explanation may be
from differences in the expression or catalytic activity of the enzymes involved in conversion
of psilocybin and psilocin (22). However, the overall pattern of degradation for the other
metabolites we measured mirrored that of psilocybin and psilocin (Fig. 2), suggesting spon-
taneous rather than enzyme-assisted degradation. Nonetheless, in any given specimen, deg-
radation patterns are likely to differ based on a complex set of factors such as age, preserva-
tion method, and species. Taken together, our results illustrate that the presence and
quantities of psilocybin and psilocin in fungarium specimens are highly unpredictable.
To investigate if the initial method of specimen preservation has a large effect on psilocy-
bin and psilocin degradation (58), we compared psilocybin and psilocin in freshly harvested
specimens that were ashfrozeninliquidN
2
and lyophilized. This preservation method is
thought to optimally preserve naturally occurring concentrations of psilocybin and psilocin
by protecting psilocybin against hydrolytic dephosphorylation (61). Our lyophilized sporo-
carp samples had 10.55 to 19.9
m
g/mg psilocybin and 0.44 to 2.0
m
g/mg psilocin, which is
consistent with previous reports, including the high ratio of psilocybin to psilocin (Table S1).
These quantities are much higher than anything we detected in heat-dried fungarium speci-
mens of P. cubensis, indicating lyophilization may be the best method for preserving psycho-
active alkaloids in this species.
Psilocybin and psilocin concentrations have been shown to vary among mycelium
and parts of the mushroom (21, 62). In an effort to document this variation in our culti-
vated P. cubensis strain, we independently analyzed psilocybin and psilocin concentra-
tions in different samples (mycelium, stipe, and cap) from multiple sporocarps within
the same ush (Psi 5 and Psi 6) and from sporocarps between two growth containers
(Psi 5/Psi 6 versus Psi 4). Mycelium showed one-third the content of psilocybin and ve
to 20 times less psilocin than did the stipe and cap, respectively (Table 2). Although we
found comparatively more alkaloids in the caps than in the stipes, the concentrations
of psilocybin and psilocin at the full sporocarp level were statistically different between
growth bins despite using the same mycelial stock and growth conditions. This sug-
gests that using consistent strain and growth conditions would yield predictably simi-
lar concentrations of both psilocin and psilocybin within a growing environment and
between caps and stipes but that inconspicuous microenvironmental conditions could
be relevant to the metabolomic phenotype (Fig. 2 and Table 2). These results have im-
portant implications for the use of whole sporocarps for therapeutic or recreational
applications, and better replicated studies are needed to elucidate the nuances of this
variation.
Untargeted metabolomics reveals a large diversity of uncharacterized metabo-
lites in P. cubensis.Although the vast majority of chemical studies of Psilocybe have
focused on the principal psychoactive alkaloids and their intermediates, Psilocybe spp.
DNA and Chemical Analysis of Psilocybe Mushrooms Applied and Environmental Microbiology
Month YYYY Volume XX Issue XX 10.1128/aem.01498-22 11
Downloaded from https://journals.asm.org/journal/aem on 06 December 2022 by 155.98.229.65.
almost certainly produce a diversity of other metabolites, at least some of which may
be bioactive (43). Our total metabolomic prole of cultivated P. cubensis detected
1,465 unique spectral features that varied among tissue types and growth replicates
(Fig. 3 and Fig. S3 and S4). Only 0.2% of these could be assigned to known chemicals
using the GNPS reference database. While becoming more common, studies of untar-
geted metabolomics in fungi are relatively new and often focus on Ascomycota rather
than Basidiomycota because of their historical use for antimicrobial and natural prod-
uct discovery (63). The general lack of these studies makes it difcult to judge if our
metabolic prole is typical of Basidiomycota, as some species, such as Trametes versi-
color (L.) Lloyd, have been shown to produce different proles based on the presence
of other fungi (64). Developmental variation is also likely. For example, Ganoderma
sichuanense J.D. Zhao & X.Q. Zhang (equals Ganoderma lingzhi Sheng H. Wu, Y. Cao &
Y.C. Dai) can have as many as 9,000 spectral features that differ depending on the life
cycle stage in which they are investigated (65). While undoubtedly highly variable, the
metabolomic prole of our sporocarps represents a snapshot of mature P. cubensis
mushrooms grown in a conventional medium, which may be useful for future compari-
son and chemical exploration of cultivated material.
Spectral features with similarity to both serotonin and trazodone hydrochloride were
detected in P. cubensis. Serotonin is a neurotransmitter that modulates many neurological
processes (66), while trazodone hydrochloride is a synthetic triazolopyridine derivative with
serotonin-selective reuptake inhibition (SSRI) activity commonly prescribed as an antide-
pressant (67). Serotonin has been reported from Panaeolus (Fr.) Quél., some species of
which also produce psilocybin (68, 69), and many other species of Basidiomycota (70). The
synthetic trazodone hydrochloride is far less likely to be present in the mushrooms since it
is not a natural product. However, motifs and chemical structures with similarity to syn-
thetic compounds have been found to be naturally occurring in other fungi (71), so the
possibility that some fungal natural products could mimic the known function of these syn-
thetic compounds is not unreasonable. Even if these two compounds were incorrectly
identied, these results clearly demonstrate that additional chemicals in the metabolome
of P. cubensis have real potential to contribute to physiological and neurological activity if
consumed.
In addition to endogenous metabolites, the accumulation of chemical compounds
from the growth substrate is a potential source of other physiologically active metabolites.
For example, we detected cocamidopropyl betaine (CAPB), a chemical component of co-
conut (72). CAPB is a common human irritant and allergen (7274),anditspresenceinthe
substrate when cultivating P. cubensis sporocarp may be a cause for concern for some peo-
ple.Whilewecannotabsolutelyconrm that this spectral feature is CAPB, neither can we
rule out that this compound is endogenously produced by P. cubensis,asoursporocarps
were produced using a casing of coconut coir ber, a likely source of this compound. This
would suggest that the accumulation of chemicals and the overall chemical composition
of cultivated sporocarps may be a combination of both endogenous and exogenous sour-
ces. Indeed, mushrooms are known to sequester chemicals from the environment, where
they have been used to remove toxins (75, 76). This is of particular importance for the
potential therapeutic or recreational use of whole mushrooms and requires empirical data
to make informed decisions on policies around their production and regulation (77).
Despite a surge in interest in Psilocybe due to accumulating evidence of the therapeutic
potential of psilocybin, the current state of knowledge of Psilocybe and the psychoactive
compoundsitproducesisstillinitsinfancy.Ourstudyshowsthatmuchmoreresearchis
needed to achieve a more thorough, accurate, and reliable understanding of this enigmatic
and socially important group to develop meaningful policies around its use and care, and
improve its application for human wellbeing.
MATERIALS AND METHODS
Fungarium voucher authentication with DNA barcodes and phylogenetics. (i) Fungarium speci-
mens. We created a curated database of voucher information for all collections identied as the genus
Psilocybe using MycoPortal (46) accessed on 1 June 2020. Records without physical vouchers (i.e.,
DNA and Chemical Analysis of Psilocybe Mushrooms Applied and Environmental Microbiology
Month YYYY Volume XX Issue XX 10.1128/aem.01498-22 12
Downloaded from https://journals.asm.org/journal/aem on 06 December 2022 by 155.98.229.65.
observations only) were removed, and the remaining names were corrected by comparing them to the
most up-to-date list of currently accepted names in Species Fungorum (78). We subsampled specimens
specically for vouchers that were identied to species level and attempted to sample a minimum of
three specimens with the same taxonomic name, totaling 94 specimens, to maximize the diversity of
our data set for accurate species comparison (see Table S1 in the supplemental material).
(ii) DNA extraction and barcode sequencing. Total DNA was extracted from the hymenophore in an
amount ranging from 5 to 15 mg. These fragments were ground to a ne powder by placing them in 2.0-mL
screw-cap tubes containing a single 3.0-mm bead and 8- by 1.5-mm stainless steel beads and shaking them
in a BeadBug microtube homogenizer (catalog no. Z763713; Sigma) for 120 s at speed setting 350. The
groundsamplewasusedasinputtotheMonarchgenomicDNApurication kit (New England Biolabs [NEB]
catalog no. T3010S) following the manufacturers instructions except using twice the volume of lysis buffer
and increasing the amount of wash buffer to 550
m
L during both of the washing steps. These samples were
amplied using the PCR primers ITS-8F (59-AGTCGTAACAAGGTTTCCGTAGGTG-39)andITS-6R(59-TTCCCG
CTTCACTCGCAGT-39), which were specically designed for Agaricomycetes (79), and submitted for Sanger
sequencing on both strands through the Genewiz Corporation as well as the DNA sequencing core at the
University of Utah.
Complementary sequencing chromatograms were then processed by creating a consensus sequence
using Sequencher 5.4 (http://www.genecodes.com) and trimming the 59and 39ends to conserved motifs 59-
CATTA- and -GACCT-39following the method in reference 79 for downstream analysis.
(iii) Phylogenetic analysis. To validate the identities of the voucher specimens, we subjected the ITS
sequences to phylogenetic analysis using two data sets. The rst consisted of all currently available ITS sequen-
ces from the National Center for Biotechnology Information (NCBI) identied as Psilocybe spp. using the query
Psilocybe [ORGN] AND internal transcribed spacer”’ (n= 384). The second included all sequences belonging to
the Species Hypothesis (SH) (generated at the 1.5% similarity threshold) of family Strophariaceae Taxon
Hypothesis (80) TH006803 (244 SH, 26 sequences assigned to Psilocybe) found in the UNITE database (UNITE
classies Psilocybe in Strophariaceae; general release, 10.05.2021) (81). Outgroup SH sequences for the genera
Agrocybe (TH012107) and Gymnopilus (TH012136) were included based on phylogenetic relationships provided
by JGI MycoCosm (https://mycocosm.jgi.doe.gov/mycocosm/species-tree/tree;Y-nUEB?organism=agaricales).
The motivation for using the UNITE SH sequences independent of NCBI sequences is that the UNITE database is
curated to minimize noise by reducing minor sequence variation using representative sequences chosen man-
ually by expert curators or automatically using a standard dynamic similarity threshold. All SHs are assigned a
unique digital object identier (DOI) to allow stable, unambiguous reference across studies, even in the com-
plete absence of meaningful taxonomic names (82).
For Psilocybe, expert verication has been completed for 9/48 SHs, for which a representative
sequence has been manually chosen, whereas the remaining SHs are represented by a sequence auto-
matically chosen by the SH algorithm.
ITS barcode sequences generated in this study were combined with data set 1 and data set 2 for
phylogenetic analysis.
For all data sets, sequences were automatically aligned using the multiple sequence alignment soft-
ware MAFFT v7.475 (83) with the L-INS-i algorithm followed by phylogenetic inference under maximum
likelihood (ML) using IQ-TREE v22.0.3 (84) with the automatic ModelFinder setting (85) and 1,000 ultra-
fast bootstrap replicates (86). The best ML trees were rendered in FigTree v1.4.4 (http://tree.bio.ed.ac.uk/
software/gtree/).
Tryptamine metabolite proling in voucher specimens. (i) Specimen preparation and chemical
extraction. Eight Psilocybe cubensis vouchers (including specimens deposited in the University of
Washington Herbarium [WTU] around the same time as those analyzed in reference 26), six cultivated P.
cubensis samples (three sporocarps split into caps [=pileus] and stipes), nine P. cyanescens samples, and
eight P. semilanceata samples were used for chemical analysis across multiyear time points (P. cubensis =
17 to 50 years, P. cyanescens = 5 to 48 years, and P. semilanceata = 5 to 47 years). Samples were ground
to a ne powder, as described above, prior to solvent extraction as described in the work of Fricke et al.
(20). Briey, homogenized samples were resuspended in a 1:20 ratio of sample (milligrams) to methanol
(microliters) and then sonicated in a Branson 1510 sonicator for 30 min. The sonicated solution was then
passed through a 0.22-
m
mlter, and the ltered solution was diluted to 100
m
g/mL with a methanol-
water (50:50) solution and injected on the ultrahigh-performance liquid chromatograph/mass spectrom-
eter (UHPLC-MS).
(ii) UHPLC-MS method to identify and quantify psilocybin and tryptamine alkaloid intermedi-
ates. Hydrophilic interaction liquid chromatography (HILIC)-mass spectrometry was performed using an
Acquity Arc UHPLC-MS (Waters, Milford, MA, USA) in both positive and negative mode with a Waters
Acquity UPLC ethylene-bridged hybrid (BEH) amide column (1.7
m
m, 2.1 by 100 mm) set to 40°C. The
l
range was set at 200 to 500 nm with a resolution of 1.2 nm and a sampling rate of 10 points/s. The mass
range in positive mode was set at 100 to 800 Da with a gain of 1, probe at 600°C, a capillary voltage of
0.8 kV, and collection at 10 points/s. A ow rate of 0.5 mL/min and a linear solvent gradient of 99.9% sol-
vent A (10 mM ammonium acetate, 95% acetonitrile, 0.014% ammonium hydroxide, pH 9.0) to 30% sol-
vent B (10 mM ammonium acetate with 0.014% ammonium hydroxide, pH 9.0) over 7 min were used.
All samples were resuspended to a concentration of 100
m
g/mL in methanol-water (50%:50%).
(iii) Chemical standard curve creation and calculation. First, we calculated control curves by
measuring the content of six chemical standards for baeocystin, norbaeocystin, norpsilocin, psilocin, psi-
locybin, and tryptophan at three different concentrations in triplicate (Fig. S2 and Table S3). This allowed
us to identify our compounds in spectral data based on retention times and calculate the concentration
of each compound in each sample based on extracted m/z intensities (Table S4). Standard curves were
DNA and Chemical Analysis of Psilocybe Mushrooms Applied and Environmental Microbiology
Month YYYY Volume XX Issue XX 10.1128/aem.01498-22 13
Downloaded from https://journals.asm.org/journal/aem on 06 December 2022 by 155.98.229.65.
measured by running standards at 1.0-, 1.2-, and 1.5-
m
g injections in triplicate and in three replicates.
The baseline integrations were recorded for each, and the standard curves were generated using the
line of best t from the standard measurements.
Cultivation of P. cubensis.Specimen cultivation and processing were carried out as follows. Specimens
of P. cubensis were cultivated from spore syringes acquired from Spore Works (sporeworks.com; Golden
Teacher spore syringe microscopy kit, SKU PSMICsy-090) inoculated on 2% malt agar plates with 50
m
g/mL
ampicillin antibiotic. Isolated sections of contiguous mycelium from initial spore plates were subcultured by
removing ;0.2cmoftissuefromtheedgeofthegrowingmyceliumwithaame-sterilized scalpel to
another plate of 2% malt agar and ampicillin with a thin sheet of sterilized cellophane plastic placed on the
surface to facilitate the removal of mycelium for chemical processing. A single culture was used for chemical
analysis as well as cultivation of sporocarps. Agar plugs from this culture were placed into spawn growth me-
dium and incubated at room temperature without a consistent light cycle for ;2to3weeksuntilmedium
was entirely colonized by mycelium based on visual inspection. Spawn growth was achieved using 1-L wide-
mouth Mason jars (Amazon, ASIN: B07HGG3DD1) containing one bag of Bens Original Ready Rice whole-
grain brown rice (Amazon, ASIN: B00G9U46DW), topped with custom lids that included a 0.22-
m
mlter for
sterile airow (Amazon, ASIN: B08M5XPFTP), with autoclaving for 40 min on a liquid cycle. To induce sporo-
carps, fully colonized spawn was mixed aseptically with one brick of coconut coir (Amazon, ASIN:
B01M8FUUSJ), 300 g of vermiculite (Amazon, ASIN: B001693Y3Y), and enough tap water (anecdotally thought
to be better than deionized [DI] H
2
O) to ll half a container with a moist but not muddy consistency within
an 18-qt plastic bin (Amazon, ASIN: B01C3BGTAS).
After 2 to 3 weeks, the lid was left ajar for increased airow to promote sporocarp production (;1week).
Mycelium and sporocarps underwent lyophilization prior to chemical analysis. Viable culture plugs were
retained for long-term storage at room temperature in 50 mL of distilled water (87), and lyophilized voucher
samples used for chemistry are deposited in the Garret Herbarium at the Natural History Museum of Utah
(UT-M) under the accession numbers listed in Table 2.
UPLC-MS of cultivated Psilocybe cubensis samples. We applied a UPLC-MS-based untargeted
metabolomics pipeline to identify shared and unique metabolites across three cultivated P. cubensis
sporocarps and vegetative mycelium. Tissue samples were lyophilized, ground, and extracted in 80:20
(vol/vol) methanol-water, producing 2 mL of retained supernatant from 100 mg (62.5 mg) of sample for
spectroscopic analysis. Small molecules (detector range of 50 to 2,000 Da) from the extraction were ana-
lyzed using UHPLC (Waters Acquity I-Class, 2.1- by 100-mm BEH amide columns) and mass spectrometry
(Waters Xevo G2 quadrupole time of ight [QToF]) (UPLC-MS) in positive and negative ionization mode.
Additionally, MS/MS spectra were acquired for sample Psi-6-Cap by running data-dependent acquisition
mode (DDA), whereby MS/MS data were collected for all metabolites that were ionized above a set
threshold (total ion current [TIC] of 5,000).
Data processing and spectral feature analysis. Raw data from the UPLC-MS analysis were processed
using the R package XCMS (88) for peak detection, peak alignment, and peak ltering. We used the following
parameters: peak detection method centWave[ppm = 15, peakwidth = c(0.2, 5), snthresh = 5, prelter =
c(1,500)]; peak grouping method density(bw = 3); and retention time correction method peakGroups,in
which retention time standards containing eight known compounds as a reference sample and integrate-
areas-of-missing-peaks method FillChromPeaksParamwere used. We then used the R package Camera (89)
to assign the various coeluting features (unique m/z-retention time combinations) derived from one com-
pound into pc groups. The parameters used were peak grouping after retention time groupFWHM
(perfwhm, 0.7; sigma = 6), verify grouping groupCorr,annotateisotopesndIsotopes, and annotate adducts
ndAdducts(polarity = positive).
Chemical prole comparison was performed by using the intensity values for each identied factor across
all cultivated samples. Intensity values were processed and normalized by the total using the decostand com-
mand from the R package Vegan (90). Normalized values were then used for PCA and PLS-DA, which were
calculated and visualized using the R package mixOmics (91). To investigate the similarity of the chemical
proles, Euclidean distances were calculated using the base R function dist. These calculations were then
assigned hierarchical clustering using the base R function hclust. These two measurements were then visual-
ized together using the aheatmap function of the R package NMF (92).
GNPS annotation to known compounds. The identication of compounds from metabolic studies is
often one of the most difcult parts of chemical proling,asthousandsofcompoundsandtheirderivatives
are often produced but most lack standard references. In an attempt to make annotation and exploration of
metabolomic data sets more accessible, GNPS was created as an initiative to make a standardized and acces-
sible platform to compare spectral data across numerous organisms and chemical databases, as well as pro-
viding a platform that would allow older data to be reanalyzed as databases become better over time (50,
93). GNPS utilizes a molecular networking approach that groups sets of spectra from related molecules to-
gether to allow for comparison to known compounds even if the specic spectrum being compared has
never been identied. In this regard, annotations with high MQscores may have structure and function highly
similar to those of the known compounds, while those with lower scores may be similar in structure but
have novel or altered function compared to the known compound.
A molecular network was created using the online workow (https://ccms-ucsd.github.io/GNPSDocumentation/)
on the GNPS website (http://gnps.ucsd.edu). The data were ltered by removing all MS/MS fragment ions
within 617 Da of the precursor m/z.MS/MSspectrawerewindowltered by choosing only the top 6 frag-
ment ions in the 650-Da window throughout the spectrum. The precursor ion mass tolerance was set to
2.0 Da, and the MS/MS fragment ion tolerance was set to 0.5 Da. A network was then created where edges
were lteredtohaveacosinescoreabove0.7andmorethan6matchedpeaks.Further,edgesbetweentwo
nodes were kept in the network only if each of the nodes appeared in the others respective top 10 most
DNA and Chemical Analysis of Psilocybe Mushrooms Applied and Environmental Microbiology
Month YYYY Volume XX Issue XX 10.1128/aem.01498-22 14
Downloaded from https://journals.asm.org/journal/aem on 06 December 2022 by 155.98.229.65.
similar nodes. Finally, the maximum size of a molecular family was set to 100, and the lowest-scoring edges
were removed from them until the molecular family size was below this threshold. The spectra in the net-
work were then searched against the GNPS spectral libraries. The library spectra were ltered in the same
manner as the input data. All matches kept between network spectra and library spectra were required to
have a score above 0.7 and at least 6 matched peaks.
Data availability. All edited ITS sequence data for all specimens were submitted to the National
Center for Biotechnology Information with NCBI accession numbers reported in Table S2. The collection
database, multiple sequence alignments, and trees are available in Figshare project number 129332. The
whole metabolomic spectra have been submitted as a MASSIVE data set available at https://massive
.ucsd.edu/ProteoSAFe/private-dataset.jsp?task=1b798991ec6b40d48b49b74e971e2ecf.
SUPPLEMENTAL MATERIAL
Supplemental material is available online only.
SUPPLEMENTAL FILE 1, PDF le, 0.5 MB.
SUPPLEMENTAL FILE 2, XLSX le, 0.02 MB.
SUPPLEMENTAL FILE 3, XLSX le, 0.02 MB.
SUPPLEMENTAL FILE 4, XLSX le, 0.04 MB.
SUPPLEMENTAL FILE 5, XLSX le, 0.01 MB.
SUPPLEMENTAL FILE 6, XLSX le, 0.02 MB.
ACKNOWLEDGMENTS
We acknowledge the Natural History Museum of Utah for their commitment to
collaborative science as well as the DNA Sequencing Core Facility, the University of Utah,
where DNA barcode sequencing was performed. In addition, we thank the institutions who
provided samples for this study and specically thank the University of Washington
Herbarium (WTU), Universidad de Guadalajara (IBUG), the New York Botanical Garden (NY),
the Harry D. Thiers Herbarium, San Francisco State University (SFSU), the University of British
Columbia Herbarium (UBC), and the University of Florida Herbarium (FLAS), who allowed us
to perform destructive sampling for chemistry in addition to DNA barcoding. We also thank
the Fungi Foundation for supporting this work in its genesis and with Giuliana Furcistime.
Further, we acknowledge the dedicated and hard work performed by Isabelle Galland, who
helped to process many of these samples for DNA sequencing. Finally, we acknowledge the
Drug Enforcement Agency (DEA), for providing the permission to perform this work under
license RW0535508.
Paul Stamets is the current CEO of Fungi Perfecti LLC, Co-founder of MycoMedica Life
Sciences, PBC and has patents pending and approved using psilocybin. All other authors
report no conicting interests in the body or intent of this work.
Alexander J. Bradshaw developed the conceptual framework and performed initial
experimental details and planning and outreach to collection institutions, as well as databasing
and curation of samples, DNA extraction and sequencing, phylogenetic analysis of samples,
initial chemical work, cultivation of specimens, data analysis, and preparation and editing of
the manuscript. Talia A. Backman performed DNA extraction of samples, the chemistry analysis
of voucher samples, the creation of chemical standard curves, and the preparation and editing
of the manuscript. Virginia Ramírez-Cruz provided expertise and advice on Psilocybe taxonomy,
a comprehensive review of the current status of rare specimens in collection institutions, and a
compilation of all previous chemical work done in Psilocybe and aided in the preparation and
editing of the manuscript. Jaclyn M. Winter (DEA license RW0535508) provided the mass
spectrometry equipment and expertise in developing the HILIC-MS methodology for analyzing
and quantifying the tryptamine alkaloids and generating chemical standard curves, as well as
the preparation and editing of the manuscript. Dale L. Forrister performed the untargeted
metabolomic analysis of samples as well as the preparation and editing of the manuscript.
Laura Guzmán-Dávalos contributed with outreach to collection institutions and comments,
additions, and editing of the manuscript. Bryn T. M. Dentinger provided expertise to oversee all
DNA sequencing and phylogenetic work as well as contributing to the original experimental
design of this work in addition to the preparation and editing of the manuscript. Giuliana Furci
contributed with outreach to collection institutions and comments and additions to the
manuscript. Paul Stamets contributed comments and additions to the manuscript, as well as
high-resolution images for specimens.
DNA and Chemical Analysis of Psilocybe Mushrooms Applied and Environmental Microbiology
Month YYYY Volume XX Issue XX 10.1128/aem.01498-22 15
Downloaded from https://journals.asm.org/journal/aem on 06 December 2022 by 155.98.229.65.
This work was generously supported by a donation to the NHMU from Fungi Perfecti
LLC.
REFERENCES
1. Wasson RG. 1957. Seeking the magic mushroom. Life 49(19):100102, 109120.
2. Fries E. 1818. Observationes mycologicae. Havniae: sumptibus G. Bon-
nieri. http://archive.org/details/observationesmyc1181frie. Accessed 30
December 2021.
3. Guzmán G. 2016. Las relaciones de los hongos sagrados con el hombre a
través del tiempo. An Antropol 50:134147. https://doi.org/10.1016/j
.antro.2015.10.005.
4. Guzmán G. 2012. Nuevas observaciones taxonómicas y etnomicológicas
en Psilocybe s.s. (Fungi, Basidiomycota, Agaricomycetidae, Agaricales,
Strophariaceae) de México, África y España. Acta Bot Mex 100:79106.
5. Guzmán Huerta G. 1960. Nueva localidad de importancia etnomicológica
de los hongos neurotrópicos mexicanos (Necaxa, Pue., México). Cien Mex
20:8588.
6. Nichols DE. 2020. Psilocybin: from ancient magic to modern medicine. J
Antibiot (Tokyo) 73:679686. https://doi.org/10.1038/s41429-020-0311-8.
7. Guzmán G. 2008. Hallucinogenic mushrooms in Mexico: an overview.
Econ Bot 62:404412. https://doi.org/10.1007/s12231-008-9033-8.
8. Guzmán G. 1997. Los nombres de los hongos y lo relacionado con ellos
en América Latina: introducción a la etnomicobiota y micología aplicada
de la region: sinonimia vulgar y cientíca. Instituto de Ecología, Xalapa,
Veracruz, México.
9. Csete J, Kamarulzaman A, Kazatchkine M, Altice F, Balicki M, Buxton J,
Cepeda J, Comfort M, Goosby E, Goulão J, Hart C, Kerr T, Lajous AM, Lewis
S, Martin N, Mejía D, Camacho A, Mathieson D, Obot I, Ogunrombi A,
Sherman S, Stone J, Vallath N, Vickerman P, Zábranský T, Beyrer C. 2016.
Public health and international drug policy. Lancet 387:14271480.
https://doi.org/10.1016/S0140-6736(16)00619-X.
10. Jones NS, Comparin JH. 2020. Interpol review of controlled substances
20162019. Forensic Sci Int Synerg 2:608669. https://doi.org/10.1016/j
.fsisyn.2020.01.019.
11. Belser AB, Agin-Liebes G, Swift TC, Terrana S, Devenot N, Friedman HL,
Guss J, Bossis A, Ross S. 2017. Patient experiences of psilocybin-assisted
psychotherapy: an interpretative phenomenological analysis. J Humanist
Psychol 57:354388. https://doi.org/10.1177/0022167817706884.
12. Bird CIV, Modlin NL, Rucker JJH. 2021. Psilocybin and MDMA for the treat-
ment of trauma-related psychopathology. Int Rev Psychiatry 33:229249.
https://doi.org/10.1080/09540261.2021.1919062.
13. Kessler RC, Sampson NA, Berglund P, Gruber MJ, Al-Hamzawi A, Andrade L,
BuntingB,DemyttenaereK,FlorescuS,deGirolamoG,GurejeO,HeY,HuC,
Huang Y, Karam E, Kovess-Masfety V, Lee S, Levinson D, Medina Mora ME,
Moskalewicz J, Nakamura Y, Navarro-Mateu F, Browne MAO, Piazza M,
Posada-Villa J, Slade T, ten Have M, Torres Y, Vilagut G, Xavier M, Zarkov Z,
Shahly V, Wilcox MA. 2015. Anxious and non-anxious major depressive disor-
der in the World Health Organization World Mental Health Surveys. Epide-
miol Psychiatr Sci 24:210226. https://doi.org/10.1017/S2045796015000189.
14. Lowe H, Toyang N, Steele B, Valentine H, Grant J, Ali A, Ngwa W, Gordon
L. 2021. The therapeutic potential of psilocybin. Molecules 26:2948.
https://doi.org/10.3390/molecules26102948.
15. Polito V, Stevenson RJ. 2019. A systematic study of microdosing psychedelics.
PLoS One 14:e0211023. https://doi.org/10.1371/journal.pone.0211023.
16. Rootman JM, Kryskow P, Harvey K, Stamets P, Santos-Brault E, Kuypers
KPC, Polito V, Bourzat F, Walsh Z. 2021. Adults who microdose psyche-
delics report health related motivations and lower levels of anxiety and
depression compared to non-microdosers. Sci Rep 11:22479. https://doi
.org/10.1038/s41598-021-01811-4.
17. Vahratian A, Blumberg SJ, Terlizzi EP, Schiller JS. 2021. Symptoms of anxi-
ety or depressive disorder and use of mental health care among adults
during the COVID-19 pandemic United States, August 2020February
2021. MMWR Morb Mortal Wkly Rep 70:490494. https://doi.org/10.15585/
mmwr.mm7013e2.
18. Gukasyan N, Davis AK, Barrett FS, Cosimano MP, Sepeda ND, Johnson MW,
Grifths RR. 2022. Efcacy and safety of psilocybin-assisted treatment for
major depressive disorder: prospective 12-month follow-up. J Psychophar-
macol 36:151158. https://doi.org/10.1177/02698811211073759.
19. Reynolds HT, Vijayakumar V, Gluck-Thaler E, Korotkin HB, Matheny PB,
Slot JC. 2018. Horizontal gene cluster transfer increased hallucinogenic
mushroom diversity. Evol Lett 2:88101. https://doi.org/10.1002/evl3.42.
20. Fricke J, Blei F, Hoffmeister D. 2017. Enzymatic synthesis of psilocybin.
Angew Chem Int Ed Engl 56:1235212355. https://doi.org/10.1002/anie
.201705489.
21. Awan AR, Winter JM, Turner D, Shaw WM, Suz LM, Bradshaw AJ, Ellis T,
Dentinger BTM. 2018. Convergent evolution of psilocybin biosynthesis by
psychedelic mushrooms. bioRxiv. 374199. https://doi.org/10.1101/374199.
22. Lenz C, Wick J, Braga D, García-Altares M, Lackner G, Hertweck C, Gressler
M, Hoffmeister D. 2020. Injury-triggered blueing reactions of Psilocybe
magicmushrooms. Angew Chem 132:14661470. https://doi.org/10
.1002/ange.201910175.
23. Guzmán G. 2005. Species diversity of the genus Psilocybe (Basidiomyco-
tina, Agaricales, Strophariaceae) in the world mycobiota, with special
attention to hallucinogenic properties. Int J Med Mushrooms 7:305332.
https://doi.org/10.1615/IntJMedMushr.v7.i12.280.
24. He M-Q, Zhao R-L, Hyde KD, Begerow D, Kemler M, Yurkov A, McKenzie
EHC, Raspé O, Kakishima M, Sánchez-Ramírez S, Vellinga EC, Halling R,
Papp V, Zmitrovich IV, Buyck B, Ertz D, Wijayawardene NN, Cui B-K,
Schoutteten N, Liu X-Z, Li T-H, Yao Y-J, Zhu X-Y, Liu A-Q, Li G-J, Zhang
M-Z, Ling Z-L, Cao B, Antonín V, Boekhout T, da Silva BDB, De Crop E,
Decock C, Dima B, Dutta AK, Fell JW, Geml J, Ghobad-Nejhad M, Giachini
AJ, Gibertoni TB, Gorjón SP, Haelewaters D, He S-H, Hodkinson BP, Horak
E, Hoshino T, Justo A, Lim YW, Menolli N, Meši
c A, Moncalvo J-M, Mueller
GM, Nagy LG, Nilsson RH, Noordeloos M, Nuytinck J, Orihara T,
Ratchadawan C, Rajchenberg M, Silva-Filho AGS, Sulzbacher MA, Tkal
cec
Z, Valenzuela R, Verbeken A, Vizzini A, Wartchow F, Wei T-Z, Weiß M, Zhao
C-L, Kirk PM. 2019. Notes, outline and divergence times of Basidiomycota.
Fungal Divers 99:105367. https://doi.org/10.1007/s13225-019-00435-4.
25. Ainsworth GC. 2008. Ainsworth & Bisbys dictionary of the fungi, 10th ed.
CABI, Wallingford, Oxon, United Kingdom.
26. Beug MW, Bigwood J. 1982. Psilocybin and psilocin levels in twenty species
from seven genera of wild mushrooms in the Pacic Northwest, U.S.A. J Eth-
nopharmacol 5:271285. https://doi.org/10.1016/0378-8741(82)90013-7.
27. Marcano V, Méndez AM, Castellano F, Salazar FJ, Martinez L. 1994. Occur-
rence of psilocybin and psilocin in Psilocybe pseudobullacea (Petch) Pegler
from the Venezuelan Andes. J Ethnopharmacol 43:157159. https://doi.org/
10.1016/0378-8741(94)90013-2.
28. Ramírez-Cruz V, Guzmán G, Guzmán-Dávalos L. 2012. New combinations in
the genus Deconica (Fungi, Basidiomycota, Agaricales). Sydowia 64:217219.
29. Noordeloos M. 2009. Genus Deconica (W. G. Sm.) P. Karst., in Europe - new
combinations. Österr Z Pilzkd 18:207210.
30. Noordeloos ME. 2011. Strophariaceae sl. Fungi Europaei 13. Edizioni Can-
dusso, Alassio, Italy.
31. Ramírez-Cruz V, da Silva PS, Villalobos-Arámbula AR, Matheny PB, Noordeloos
M,MorgadoL,daSilveriaRMB,Guzmán-DávalosL.2020.Twonewspeciesof
Deconica (Agaricales, Basidiomycota) from Australia and Mexico. Mycol Prog
19:13171328. https://doi.org/10.1007/s11557-020-01629-w.
32.VanCourtRC,WisemanMS,MeyerKW,BallhornDJ,AmsesKR,SlotJC,
Dentinger BTM, Garibay-Orijel R, Uehling JK. 2022. Diversity, biology, and his-
tory of psilocybin-containing fungi: suggestions for research and technologi-
cal development. Fungal Biol 126:308319. https://doi.org/10.1016/j.funbio
.2022.01.003.
33. Guzmán G, Allen JW, Sihanonth P. 2006. Distribution of the hallucinogenic
mushroom Psilocybe antioquensis Guzmán et al. (Agaricomycetideae) in
Colombia, Mexico, and Cambodia. Int J Med Mushrooms 8:8589. https://doi
.org/10.1615/IntJMedMushr.v8.i1.110.
34. Ramírez-Cruz V, Guzmán G, Villalobos-Arámbula AR, Rodríguez A,
Matheny PB, Sánchez-García M, Guzmán-Dávalos L. 2013. Phylogenetic
inference and trait evolution of the psychedelic mushroom genus Psilo-
cybe sensu lato (Agaricales). Botany 91:573591. https://doi.org/10.1139/
cjb-2013-0070.
35. Stamets PE. 1996. Psilocybin mushrooms of the world: an identication
guide. Ten Speed, Berkeley, CA.
36. Guzmán G. 1983. The genus Psilocybe: a systematic revision of the known
species including the history, distribution, and chemistry of the hallucino-
genic species. J Cramer, Vaduz, Liechtenstein.
37. Singer R, Smith AH. 1958. Mycological investigations on Teonanacatl, the
Mexican hallucinogenic mushroom. Part II. A taxonomic monograph of
DNA and Chemical Analysis of Psilocybe Mushrooms Applied and Environmental Microbiology
Month YYYY Volume XX Issue XX 10.1128/aem.01498-22 16
Downloaded from https://journals.asm.org/journal/aem on 06 December 2022 by 155.98.229.65.
Psilocybe, section Caerulescentes. Mycologia 50:262303. https://doi.org/
10.1080/00275514.1958.12024726.
38. Andrew C, Diez J, James TY, Kauserud H. 2019. Fungarium specimens: a
largely untapped source in global change biology and beyond. Philos
Trans R Soc B 374:20170392. https://doi.org/10.1098/rstb.2017.0392.
39. Hofstetter V, Buyck B, Eyssartier G, Schnee S, Gindro K. 2019. The unbearable
lightness of sequenced-based identication. Fungal Divers 96:243284.
https://doi.org/10.1007/s13225-019-00428-3.
40. Allen J. 2011. A chemical referral and reference guide to the known spe-
cies of psilocin and/or psilocybin-containing mushrooms and their pub-
lished analysis and bluing reactions: an updated and revised list. Ethno-
mycol J Sacred Mushroom Stud 2011:152183.
41. TylšF, Pálení
cek T, Horá
cek J. 2014. Psilocybin summary of knowledge
and new perspectives. Eur Neuropsychopharmacol 24:342356. https://
doi.org/10.1016/j.euroneuro.2013.12.006.
42. Andersson C, Kristinsson J, Gry J. 2008. Occurrence and use of hallucino-
genic mushrooms containing psilocybin alkaloids. Nordic Council of Min-
isters, Copenhagen, Denmark.
43. Blei F, Dörner S, Fricke J, Baldeweg F, Trottmann F, Komor A, Meyer F,
Hertweck C, Hoffmeister D. 2020. Simultaneous productionof psilocybin and
acocktailof
b
-carboline monoamine oxidase inhibitors in magicmush-
rooms. Chemistry 26:729734. https://doi.org/10.1002/chem.201904363.
44. Dörner S, Rogge K, Fricke J, Schäfer T, Wurlitzer JM, Gressler M, Pham
DNK, Manke DR, Chadeayne AR, Hoffmeister D. 2022. Genetic survey of
Psilocybe natural products. Chembiochem 23:e202200249. https://doi
.org/10.1002/cbic.202200249.
45. Basky G. 2021. Policy in focus: is psilocybin the next cannabis? CMAJ 193:
E1741E1742. https://doi.org/10.1503/cmaj.1095974.
46. Miller AN, Bates ST. 2017. The Mycology Collections Portal (MyCoPortal).
IMA Fungus 8:A65A66. https://doi.org/10.1007/BF03449464.
47. Bonito G, Isikhuemhen OS, Vilgalys R. 2010. Identication of fungi associ-
ated with municipal compost using DNA-based techniques. Bioresour
Technol 101:10211027. https://doi.org/10.1016/j.biortech.2009.08.109.
48. Kõljalg U, Nilsson RH, Abarenkov K, Tedersoo L, Taylor AF, Bahram M,
Bates ST, Bruns TD, Bengtsson-Palme J, Callaghan TM, Douglas B,
Drenkhan T, Eberhardt U, Dueñas M, Grebenc T, Grifth GW, Hartmann M,
Kirk PM, Kohout P, Larsson E, Lindahl BD, Lücking R, Martín MP, Matheny
PB, Nguyen NH, Niskanen T, Oja J, Peay KG, Peintner U, Peterson M,
Põldmaa K, Saag L, Saar I, Schüßler A, Scott JA, Senés C, Smith ME, Suija A,
Taylor DL, Telleria MT, Weiss M, Larsson KH. 2013. Towards a unied para-
digm for sequence-based identication of fungi. Mol Ecol 22:52715277.
https://doi.org/10.1111/mec.12481.
49. Varga T, Krizsán K, Földi C, Dima B, Sánchez-García M, Sánchez-Ramírez S,
Szöllosi GJ, Szarkándi JG, Papp V, Albert L, Andreopoulos W, Angelini C,
Antonín V, Barry KW, Bougher NL, Buchanan P, Buyck B, Bense V,
Catcheside P, Chovatia M, Cooper J, Dämon W, Desjardin D, Finy P, Geml
J, Haridas S, Hughes K, Justo A, Karasi
nski D, Kautmanova I, Kiss B,
Kocsubé S, Kotiranta H, LaButti KM, Lechner BE, Liimatainen K, Lipzen A,
Lukács Z, Mihaltcheva S, Morgado LN, Niskanen T, Noordeloos ME, Ohm
RA, Ortiz-Santana B, Ovrebo C, Rácz N, Riley R, Savchenko A, Shiryaev A,
Soop K, Spirin V, Szebenyi C, Tomšovský M, Tulloss RE, Uehling J,
Grigoriev IV, Vágvölgyi C, Papp T, Martin FM, Miettinen O, Hibbett DS,
Nagy LG. 2019. Megaphylogeny resolves global patterns of mushroom
evolution. Nat Ecol Evol 3:668678. https://doi.org/10.1038/s41559-019
-0834-1.
50. Wang M, Carver JJ, Phelan VV, Sanchez LM, Garg N, Peng Y, Nguyen DD,
Watrous J, Kapono CA, Luzzatto-Knaan T, Porto C, Bouslimani A, Melnik
AV, Meehan MJ, Liu W-T, Crüsemann M, Boudreau PD, Esquenazi E,
Sandoval-Calderón M, Kersten RD, Pace LA, Quinn RA, Duncan KR, Hsu
C-C, Floros DJ, Gavilan RG, Kleigrewe K, Northen T, Dutton RJ, Parrot D,
Carlson EE, Aigle B, Michelsen CF, Jelsbak L, Sohlenkamp C, Pevzner P,
Edlund A, McLean J, Piel J, Murphy BT, Gerwick L, Liaw C-C, Yang Y-L,
Humpf H-U, Maansson M, Keyzers RA, Sims AC, Johnson AR, Sidebottom
AM, Sedio BE, Klitgaard A, Larson CB, Boya PCA, Torres-Mendoza D,
Gonzalez DJ, Silva DB, Marques LM, Demarque DP, Pociute E, ONeill EC,
Briand E, Helfrich EJN, Granatosky EA, Glukhov E, Ryffel F, Houson H,
Mohimani H, Kharbush JJ, Zeng Y, Vorholt JA, Kurita KL, Charusanti P,
McPhail KL, Nielsen KF, Vuong L, Elfeki M, Traxler MF, Engene N, Koyama
N, Vining OB, Baric R, Silva RR, Mascuch SJ, Tomasi S, Jenkins S, Macherla
V, Hoffman T, Agarwal V, Williams PG, Dai J, Neupane R, Gurr J, Rodríguez
AMC, Lamsa A, Zhang C, Dorrestein K, Duggan BM, Almaliti J, Allard P-M,
Phapale P, Nothias L-F, Alexandrov T, Litaudon M, Wolfender J-L, Kyle JE,
Metz TO, Peryea T, Nguyen D-T, VanLeer D, Shinn P, Jadhav A, Müller R,
Waters KM, Shi W, Liu X, Zhang L, Knight R, Jensen PR, Palsson BØ,
Pogliano K, Linington RG, Gutiérrez M, Lopes NP, Gerwick WH, Moore BS,
Dorrestein PC, Bandeira N. 2016. Sharing and community curation of mass
spectrometry data with Global Natural Products Social Molecular Network-
ing. Nat Biotechnol 34:828837. https://doi.org/10.1038/nbt.3597.
51. McLaughlin DJ, Hibbett DS, Lutzoni F, Spatafora JW, Vilgalys R. 2009. The
search for the fungal tree of life. Trends Microbiol 17:488497. https://doi
.org/10.1016/j.tim.2009.08.001.
52. Guzmán G, Allen J, Gartz J. 1998. A worldwide geographical distribution
of the neurotropic fungi, an analysis and discussion. Ann Mus Civ Rover-
eto 14:189280.
53. Froese T, Guzmán G, Guzmán-Dávalos L. 2016. On the origin of the genus
Psilocybe and its potential ritual use in ancient Africa and Europe. Econ
Bot 70:103114. https://doi.org/10.1007/s12231-016-9342-2.
54. Ma T, Feng Y, Lin XF, Karunarathna SC, Ding WF, Hyde KD. 2014. Psilocybe
chuxiongensis, a new bluing species from subtropical China. Phytotaxa
156:211. https://doi.org/10.11646/phytotaxa.156.4.3.
55. Koike Y, Wada K, Kusano G, Nozoe S, Yokoyama K. 1981. Isolation of psilocy-
bin from Psilocybe argentipes and its determination in specimens of some
mushrooms. J Nat Prod 44:362365. https://doi.org/10.1021/np50015a023.
56. Mahmood ZA. 2013. Bioactive alkaloids from fungi: psilocybin, p 523552. In
Ramawat KG, Mérillon J-M (ed), Natural products. Springer, Berlin, Germany.
57. Keller T, Schneider A, Regenscheit P, Dirnhofer R, Rücker T, Jaspers J,
Kisser W. 1999. Analysis of psilocybin and psilocin in Psilocybe subcubensis
Guzmán by ion mobility spectrometry and gas chromatographymass
spectrometry. Forensic Sci Int 99:93105. https://doi.org/10.1016/S0379
-0738(98)00168-6.
58. Gotvaldová K, Hájková K, Borovi
cka J, Jurok R, Cihlá
rová P, Kucha
rM.
2021. Stability of psilocybin and its four analogs in the biomass of the
psychotropic mushroom Psilocybe cubensis. Drug Test Anal 13:439446.
https://doi.org/10.1002/dta.2950.
59. Beug MW, Bigwood J. 1981. Quantitative analysis of psilocybin and psilo-
cin and Psilocybe baeocystis (Singer and Smith) by high-performance liq-
uid chromatography and by thin-layer chromatography. J Chromatogr A
207:379385. https://doi.org/10.1016/S0021-9673(00)88741-5.
60. Bigwood J, Beug MW. 1982. Variation of psilocybin and psilocin levels with
repeated ushes (harvests) of mature sporocarps of Psilocybe cubensis
(Earle) Singer. J Ethnopharmacol 5:287291. https://doi.org/10.1016/0378
-8741(82)90014-9.
61. Lenz C, Wick J, Hoffmeister D. 2017. Identication of
v
-N-methyl-4-
hydroxytryptamine (norpsilocin) as a Psilocybe natural product. J Nat
Prod 80:28352838. https://doi.org/10.1021/acs.jnatprod.7b00407.
62. Tsujikawa K, Kanamori T, Iwata Y, Ohmae Y, Sugita R, Inoue H, Kishi T.
2003. Morphological and chemical analysis of magic mushrooms in Ja-
pan. Forensic Sci Int 138:8590. https://doi.org/10.1016/j.forsciint.2003.08
.009.
63. Alberti F, Kaleem S, Weaver JA. 2020. Recent developments of tools for
genome and metabolome studies in basidiomycete fungi and their appli-
cation to natural product research. Biol Open 9:bio056010. https://doi
.org/10.1242/bio.056010.
64. Yao L, Zhu L-P, Xu X-Y, Tan L-L, Sadilek M, Fan H, Hu B, Shen X-T, Yang J,
Qiao B, Yang S. 2016. Discovery of novel xylosides in co-culture of basidio-
mycetes Trametes versicolor and Ganoderma applanatum by integrated
metabolomics and bioinformatics. Sci Rep 6:33237. https://doi.org/10
.1038/srep33237.
65. Satria D, Tamrakar S, Suhara H, Kaneko S, Shimizu K. 2019. Mass spectrom-
etry-based untargeted metabolomics and
a
-glucosidase inhibitory activ-
ity of lingzhi (Ganoderma lingzhi) during the developmental stages. Mole-
cules 24:2044. https://doi.org/10.3390/molecules24112044.
66. Berger M, Gray JA, Roth BL. 2009. The expanded biology of serotonin. Annu
Rev Med 60:355366. https://doi.org/10.1146/annurev.med.60.042307.110802.
67. Fagiolini A, Comandini A, Catena DellOsso M, Kasper S. 2012. Rediscover-
ing trazodone for the treatment of major depressive disorder. CNS Drugs
26:10331049. https://doi.org/10.1007/s40263-012-0010-5.
68. Ott J, Guzmán G. 1976. Detection of psilocybin in species of Psilocybe,
Panaeolus and Psathyrella. Lloydia 39:258260.
69. Wier JK, Tyler VE. 1963. Quantitative determination of serotonin in Panaeolus
species. J Pharm Sci 52:419422. https://doi.org/10.1002/jps.2600520504.
70. Muszy
nska B, Sułkowska-Ziaja K, Ekiert H. 2011. Indole compounds in
fruiting bodies of some edible Basidiomycota species. Food Chem 125:
13061308. https://doi.org/10.1016/j.foodchem.2010.10.056.
71. Takahashi A, Nunozawa T, Endo T, Nozoe S. 1992. Isolation of 1-beta-D-
arabinofuranosylcytosine from the mushroom Xerocomus nigromaculatus
Hongo. Chem Pharm Bull (Tokyo) 40:13131314. https://doi.org/10.1248/
cpb.40.1313.
DNA and Chemical Analysis of Psilocybe Mushrooms Applied and Environmental Microbiology
Month YYYY Volume XX Issue XX 10.1128/aem.01498-22 17
Downloaded from https://journals.asm.org/journal/aem on 06 December 2022 by 155.98.229.65.
72. Shaffer KK, Jaimes JP, Hordinsky MK, Zielke GR, Warshaw EM. 2006. Aller-
genicity and cross-reactivity of coconut oil derivatives: a double-blind
randomized controlled pilot study. Dermatitis 17:7176.
73. de Groot AC, van der Walle HB, Weyland JW. 1995. Contact allergy to
cocamidopropyl betaine. Contact Dermatitis 33:419422. https://doi.org/
10.1111/j.1600-0536.1995.tb02078.x.
74. Suuronen K, Pesonen M, Aalto-Korte K. 2012. Occupational contact
allergy to cocamidopropyl betaine and its impurities. Contact Dermatitis
66:286292. https://doi.org/10.1111/j.1600-0536.2011.02036.x.
75. Akhtar N, Mannan MA. 2020. Mycoremediation: expunging environmen-
tal pollutants. Biotechnol Rep (Amst) 26:e00452. https://doi.org/10.1016/j
.btre.2020.e00452.
76. Singh M, Srivastava PK, Verma PC, Kharwar RN, Singh N, Tripathi RD. 2015.
Soil fungi for mycoremediation of arsenic pollution in agriculture soils. J
Appl Microbiol 119:12781290. https://doi.org/10.1111/jam.12948.
77. Aday JS, Bloesch EK, Davoli CC. 2020. 2019: a year of expansion in psyche-
delic research, industry, and deregulation. Drug Sci Policy Law 6:
205032452097448. https://doi.org/10.1177/2050324520974484.
78. Kirk PM. 2019. Species Fungorum in the catalogue of life. The Catalogue
of Life Partnership, Leiden, The Netherlands.
79. Dentinger BTM, Margaritescu S, Moncalvo J. 2010. Rapid and reliable
high-throughput methods of DNA extraction for use in barcoding and
molecular systematics of mushrooms. Mol Ecol Resour 10:628633.
https://doi.org/10.1111/j.1755-0998.2009.02825.x.
80. Kõljalg U, Nilsson HR, Schigel D, Tedersoo L, Larsson K-H, May TW, Taylor
AFS, Jeppesen TS, Frøslev TG, Lindahl BD, Põldmaa K, Saar I, Suija A,
Savchenko A, Yatsiuk I, Adojaan K, Ivanov F, Piirmann T, Pöhönen R, Zirk
A, Abarenkov K. 2020. The taxon hypothesis paradigmon the unambig-
uous detection and communication of taxa. Microorganisms 8:1910.
https://doi.org/10.3390/microorganisms8121910.
81. Grigoriev IV, Nikitin R, Haridas S, Kuo A, Ohm R, Otillar R, Riley R, Salamov
A, Zhao X, Korzeniewski F, Smirnova T, Nordberg H, Dubchak I, Shabalov I.
2014. MycoCosm portal: gearing up for 1000 fungal genomes. Nucleic
Acids Res 42:D699D704. https://doi.org/10.1093/nar/gkt1183.
82. Nilsson RH, Larsson K-H, Taylor AFS, Bengtsson-Palme J, Jeppesen TS,
Schigel D, Kennedy P, Picard K, Glöckner FO, Tedersoo L, Saar I, Kõljalg U,
Abarenkov K. 2019. The UNITE database for molecular identication of
fungi: handling dark taxa and parallel taxonomic classications. Nucleic
Acids Res 47:D259D264. https://doi.org/10.1093/nar/gky1022.
83. Katoh K, Misawa K, Kuma K-I, Miyata T. 2002. MAFFT: a novel method for
rapid multiple sequence alignment based on fast Fourier transform.
Nucleic Acids Res 30:30593066. https://doi.org/10.1093/nar/gkf436.
84. Nguyen L-T, Schmidt HA, von Haeseler A, Minh BQ. 2015. IQ-TREE: a fast and
effective stochastic algorithm for estimating maximum-likelihood phyloge-
nies. Mol Biol Evol 32:268274. https://doi.org/10.1093/molbev/msu300.
85. Kalyaanamoorthy S, Minh BQ, Wong TKF, von Haeseler A, Jermiin LS.
2017. ModelFinder: fast model selection for accurate phylogenetic esti-
mates. Nat Methods 14:587589. https://doi.org/10.1038/nmeth.4285.
86. Minh BQ, Nguyen MAT, von Haeseler A. 2013. Ultrafast approximation for
phylogenetic bootstrap. Mol Biol Evol 30:11881195. https://doi.org/10
.1093/molbev/mst024.
87. Nakasone KK, Peterson SW, Jong S-C. 2004. Preservation and distribution
of fungal cultures, p 3747. In Foster M, Bills G (ed), Biodiversity of fungi:
inventory and monitoring methods. Elsevier, New York, NY.
88. Smith CA, Want EJ, OMaille G, Abagyan R, Siuzdak G. 2006. XCMS: proc-
essing mass spectrometry data for metabolite proling using nonlinear
peak alignment, matching, and identication. Anal Chem 78:779787.
https://doi.org/10.1021/ac051437y.
89. Kuhl C, Tautenhahn R, Böttcher C, Larson TR, Neumann S. 2012. CAMERA:
an integrated strategy for compound spectra extraction and annotation
of liquid chromatography/mass spectrometry data sets. Anal Chem 84:
283289. https://doi.org/10.1021/ac202450g.
90. Dixon P. 2003. VEGAN, a package of R functions for community ecology. J
Veg Sci 14:927930. https://doi.org/10.1111/j.1654-1103.2003.tb02228.x.
91. Rohart F, Gautier B, Singh A, Cao K-A. 2017. mixOmics: an R package
for omics feature selection and multiple data integration. PLoS Comput
Biol 13:e1005752. https://doi.org/10.1371/journal.pcbi.1005752.
92. Gaujoux R, Seoighe C. 2010. A exible R package for nonnegative matrix fac-
torization. BMC Bioinformatics 11:367. https://doi.org/10.1186/1471-2105-11
-367.
93. AronAT,GentryEC,McPhailKL,NothiasL-F,Nothias-EspositoM,Bouslimani
A, Petras D, Gauglitz JM, Sikora N, Vargas F, van der Hooft JJJ, Ernst M, Kang
KB,AcevesCM,Caraballo-RodríguezAM,KoesterI,WeldonKC,BertrandS,
RoullierC,SunK,TehanRM,BoyaPCA,ChristianMH,GutiérrezM,UlloaAM,
Tejeda Mora JA, Mojica-Flores R, Lakey-Beitia J, Vásquez-Chaves V, Zhang Y,
CalderónAI,TaylerN,KeyzersRA,TugizimanaF,NdlovuN,AksenovAA,
Jarmusch AK, Schmid R, Truman AW, Bandeira N, Wang M, Dorrestein PC.
2020. Reproducible molecular networking of untargeted mass spectrometry
data using GNPS. Nat Protoc 15:19541991. https://doi.org/10.1038/s41596
-020-0317-5.
DNA and Chemical Analysis of Psilocybe Mushrooms Applied and Environmental Microbiology
Month YYYY Volume XX Issue XX 10.1128/aem.01498-22 18
Downloaded from https://journals.asm.org/journal/aem on 06 December 2022 by 155.98.229.65.
... If participants feel it necessary, they should also be able to provide alternative or updated identifications (this being an expanded form of taxonomic referencing; (Thomer et al., 2012)). Several taxonomic groups of museum natural history collections are known to have high levels of misidentification (Bradshaw et al., 2022;Moran, 1983;Nekola et al., 2019), and morphological or ecological validation is particularly important for groups such as fungi given the known unreliability of their published DNA sequences (Bridge et al., 2003). The use of natural history societies and amateur naturalists to not only provide but also monitor and validate contemporary biological records is a well-developed practice (Turnhout et al., 2016). ...
Article
Amateur naturalists and natural history societies are abundant throughout the UK, and indeed throughout the globe, and have intimate relationships with the museum sector. Importantly, amateur naturalists and natural history societies often possess expertise in fields underserved by museums, such as taxonomy or the local history of natural history. The amateur naturalist, in the context of the museum, is simultaneously volunteer and expert and this dichotomy can lead to tension in the naturalist‐museum professional relationship. The rise of mass‐participation citizen science and digitization projects are likely to further enflame this tension as naturalists can feel marginalized as their identifier of “volunteer” subsumes that of “expert.” This paper recommends best practice when working with amateur naturalists and natural history societies. It examines the different value systems and priorities broadly held by museum professionals and naturalists, with a particular focus on digitization. This paper also suggests how the museum field can work toward more equitable knowledge building and sharing practices when working with naturalists and other expert‐volunteers. In particular, it puts forward the concept of messy databases as a way to meaningfully engage with volunteer‐experts.
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
Background: Preliminary data suggest that psilocybin-assisted treatment produces substantial and rapid antidepressant effects in patients with major depressive disorder (MDD), but little is known about long-term outcomes. Aims: This study sought to examine the efficacy and safety of psilocybin through 12 months in participants with moderate to severe MDD who received psilocybin. Methods: This randomized, waiting-list controlled study enrolled 27 patients aged 21–75 with moderate to severe unipolar depression (GRID-Hamilton Depression Rating Scale (GRID-HAMD)⩾17). Participants were randomized to an immediate or delayed (8weeks) treatment condition in which they received two doses of psilocybin with supportive psychotherapy. Twenty-four participants completed both psilocybin sessions and were followed through 12months following their second dose. Results: All 24 participants attended all follow-up visits through the 12-month timepoint. Large decreases from baseline in GRID-HAMD scores were observed at 1-, 3-, 6-, and 12-month follow-up (Cohen d=2.3, 2.0, 2.6, and 2.4, respectively). Treatment response (⩾50% reduction in GRID-HAMD score from baseline) and remission were 75% and 58%, respectively, at 12months. There were no serious adverse events judged to be related to psilocybin in the long-term follow-up period, and no participants reported psilocybin use outside of the context of the study. Participant ratings of personal meaning, spiritual experience, and mystical experience after sessions predicted increased well-being at 12months, but did not predict improvement in depression. Conclusions: These findings demonstrate that the substantial antidepressant effects of psilocybin-assisted therapy may be durable at least through 12 months following acute intervention in some patients.
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
Full-text available
The use of psychedelic substances at sub-sensorium ‘ microdoses’, has gained popular academic interest for reported positive effects on wellness and cognition. The present study describes microdosing practices, motivations and mental health among a sample of self-selected microdosers ( n = 4050) and non-microdosers ( n = 4653) via a mobile application. Psilocybin was the most commonly used microdose substances in our sample (85%) and we identified diverse microdose practices with regard to dosage, frequency, and the practice of stacking which involves combining psilocybin with non-psychedelic substances such as Lion’s Mane mushrooms, chocolate, and niacin. Microdosers were generally similar to non-microdosing controls with regard to demographics, but were more likely to report a history of mental health concerns. Among individuals reporting mental health concerns, microdosers exhibited lower levels of depression, anxiety, and stress across gender. Health and wellness-related motives were the most prominent motives across microdosers in general, and were more prominent among females and among individuals who reported mental health concerns. Our results indicate health and wellness motives and perceived mental health benefits among microdosers, and highlight the need for further research into the mental health consequences of microdosing including studies with rigorous longitudinal designs.
Article
Full-text available
This review examines the role of trauma in psychiatric morbidity and analogous psychoneurobiological changes. Trauma is a necessary criterion for Post-Traumatic Stress Disorder (PTSD), however, trauma history is highly correlated with a variety of psychiatric conditions. Some evidence suggests that Major Depressive Disorder (MDD) is the most common psychiatric condition that arises following trauma. Approximately 50% of PTSD cases present with co-morbid MDD. Overlapping symptomatology and neurobiology between these conditions underlie the debate over whether these phenomena result from problematic nosology or whether comorbid MDD + PTSD is a distinct phenotype of trauma-related psychopathology. Regardless, similar treatment approaches have been employed historically, with varying success. The drug-assisted psychotherapy treatment model, which combines pharmacological and psychotherapeutic approaches, is currently being trialled as a novel treatment approach in psychiatry. Both psilocybin- and 3,4-Methylenedioxymethamphetamine (MDMA)-assisted psychotherapy have received Food and Drug Administration ‘breakthrough therapy’ designation for the treatment of resistant MDD and PTSD, respectively. This paper reviews the therapeutic rationale of both psilocybin and MDMA for treating both trauma-related MDD and PTSD.
Article
Full-text available
The psychedelic effects of some plants and fungi have been known and deliberately exploited by humans for thousands of years. Fungi, particularly mushrooms, are the principal source of naturally occurring psychedelics. The mushroom extract, psilocybin has historically been used as a psychedelic agent for religious and spiritual ceremonies, as well as a therapeutic option for neuropsychiatric conditions. Psychedelic use was largely associated with the “hippie” counterculture movement, which, in turn, resulted in a growing, and still lingering, negative stigmatization for psychedelics. As a result, in 1970, the U.S. government rescheduled psychedelics as Schedule 1 drugs, ultimately ending scientific research on psychedelics. This prohibition on psychedelic drug research significantly delayed advances in medical knowledge on the therapeutic uses of agents such as psilocybin. A 2004 pilot study from the University of California, Los Angeles, exploring the potential of psilocybin treatment in patients with advanced-stage cancer managed to reignite interest and significantly renewed efforts in psilocybin research, heralding a new age in exploration for psychedelic therapy. Since then, significant advances have been made in characterizing the chemical properties of psilocybin as well as its therapeutic uses. This review will explore the potential of psilocybin in the treatment of neuropsychiatry-related conditions, examining recent advances as well as current research. This is not a systematic review.
Article
Full-text available
The spread of disease and increase in deaths during large outbreaks of transmissible diseases is often associated with fear and grief (1). Social restrictions, limits on operating nonessential businesses, and other measures to reduce pandemic-related mortality and morbidity can lead to isolation and unemployment or underemployment, further increasing the risk for mental health problems (2). To rapidly monitor changes in mental health status and access to care during the COVID-19 pandemic, CDC partnered with the U.S. Census Bureau to conduct the Household Pulse Survey (HPS). This report describes trends in the percentage of adults with symptoms of an anxiety disorder or a depressive disorder and those who sought mental health services. During August 19, 2020-February 1, 2021, the percentage of adults with symptoms of an anxiety or a depressive disorder during the past 7 days increased significantly (from 36.4% to 41.5%), as did the percentage reporting that they needed but did not receive mental health counseling or therapy during the past 4 weeks (from 9.2% to 11.7%). Increases were largest among adults aged 18-29 years and among those with less than a high school education. HPS data can be used in near real time to evaluate the impact of strategies that address mental health status and care of adults during the COVID-19 pandemic and to guide interventions for groups that are disproportionately affected.
Article
Full-text available
Here, we describe the taxon hypothesis (TH) paradigm, which covers the construction, identification, and communication of taxa as datasets. Defining taxa as datasets of individuals and their traits will make taxon identification and most importantly communication of taxa precise and reproducible. This will allow datasets with standardized and atomized traits to be used digitally in identification pipelines and communicated through persistent identifiers. Such datasets are particularly useful in the context of formally undescribed or even physically undiscovered species if data such as sequences from samples of environmental DNA (eDNA) are available. Implementing the TH paradigm will to some extent remove the impediment to hastily discover and formally describe all extant species in that the TH paradigm allows discovery and communication of new species and other taxa also in the absence of formal descriptions. The TH datasets can be connected to a taxonomic backbone providing access to the vast information associated with the tree of life. In parallel to the description of the TH paradigm, we demonstrate how it is implemented in the UNITE digital taxon communication system. UNITE TH datasets include rich data on individuals and their rDNA ITS sequences. These datasets are equipped with digital object identifiers (DOI) that serve to fix their identity in our communication. All datasets are also connected to a GBIF taxonomic backbone. Researchers processing their eDNA samples using UNITE datasets will, thus, be able to publish their findings as taxon occurrences in the GBIF data portal. UNITE species hypothesis (species level THs) datasets are increasingly utilized in taxon identification pipelines and even formally undescribed species can be identified and communicated by using UNITE. The TH paradigm seeks to achieve unambiguous, unique, and traceable communication of taxa and their properties at any level of the tree of life. It offers a rapid way to discover and communicate undescribed species in identification pipelines and data portals before they are lost to the sixth mass extinction.
Article
Two new species of Deconica with chrysocystidia are described from Australia and Mexico. Deconica milvispora, from a warm temperate rainforest in Queensland, Australia, is characterized by rhomboid basidiospores with nodulose angles not observed in any member of the genus until now. Deconica esperancensis, from a cloud forest in the south of Mexico, is characterized by subrhomboid basidiospores and short cheilocystidia. Full descriptions, illustrations, and comparisons with similar species are provided for both species. Bayesian and maximum likelihood analyses were performed employing sequences of nuc ITS rDNA, including seven new sequences from historical type collections of D. aequatoriae, D. alpestris, D. semiinconspicua, Deconica submaritima, D. thailandensis, D. umbrina, and Psilocybe subrunneocystidiata (= Deconica neorhombispora). As expected, the results showed the phylogenetic position of the new species described here within Deconica sect. Chrysocystidiatae, a supported monophyletic group. This section comprises many species, the majority with chrysocystidia, ten of them newly discovered but unnamed taxa. At this time, two taxa are described. Additionally, discussion of the traits found in the four major clades of Deconica is presented.