PreprintPDF Available

Discovery of the closest free-living relative of the domesticated "magic mushroom" Psilocybe cubensis in Africa

Authors:
Preprints and early-stage research may not have been peer reviewed yet.

Abstract

The "magic mushroom" Psilocybe cubensis is cultivated worldwide for recreational and medicinal uses. Described initially from Cuba in 1904, there has been substantial debate about its origin and diversification. The prevailing view, first proposed by the Psilocybe expert Gaston Guzman in 1983, is that P. cubensis was inadvertently introduced to the Americas when cattle were introduced to the continents from Africa and Europe (~1500 CE), but that its progenitor was endemic to Africa. This hypothesis has never been tested. Here, we report the discovery of the closest wild relative of P. cubensis from sub-Saharan Africa, P. ochraceocentrata nom. prov. Using DNA sequences from type specimens of all known African species of Psilocybe, multi-locus phylogenetic and molecular clock analysis strongly support recognizing the African samples as a new species that last shared a common ancestor with P. cubensis ~1.5 million years ago (~710k - 2.55M years ago 95% HPD). Even at the latest estimated time of divergence, this long predates cattle domestication and the origin of modern humans. Both species are associated with herbivore dung, suggesting this habit likely predisposed P. cubensis to its present specialization on domesticated cattle dung. Ecological niche modeling using bioclimatic variables for global records of these species indicates historical presence across Africa, Asia, and the Americas over the last 3 million years. This discovery sheds light on the wild origins of domesticated P. cubensis and provides new genetic resources for research on psychedelic mushrooms.
Discovery of the closest free-living relative of the domesticated “magic mushroom”
Psilocybe cubensis
in
1
Africa
2
3
Alexander J Bradshaw
1,5
, Cathy Sharp
2
, Breyten Van Der Merwe
3
, Keaton Tremble
4
, Bryn T.M. Dentinger
5
4
5
1
Biology Department, Clark University, Worcester, MA 01610
6
2
Natural History Museum of Zimbabwe, Centenary Park, cnr Park Road &, Leopold Takawira Ave,
7
Bulawayo, Zimbabwe
8
3
Department of Microbiology, Stellenbosch University, Private Bag X1, Stellenbosch, 7600, South Africa
9
4
Department of Biology, Duke University, Durham, NC, USA
10
5
Natural History Museum of Utah and School of Biological Sciences, University of Utah, 301 Wakara
11
Way, Salt Lake City UT 84108
12
13
14
Emails:
15
Abradshaw@clarku.edu
16
mycofreedom@gmail.com,
17
breytenvdmerwe@gmail.com
18
keaton.tremble@duke.edu
19
bryn.dentinger@gmail.com
20
Corresponding Authors:
Alexander J Bradshaw, Bryn T.M. Dentinger
21
22
ORCID:
23
Alexander J Bradshaw: https://orcid.org/0000-0002-6261-621X
24
Cathy Sharp:
https://orcid.org/0009-0003-4985-1543
25
Breyten Van Der Merwe: http://orcid.org/0000-0003-0546-5619
26
Keaton Tremble: https://orcid.org/0000-0002-0788-2830
27
Bryn T.M. Dentinger: https://orcid.org/0000-0001-7965-4389
28
29
Keywords: psychedelic mushroom, psilocybin, diversity, Zimbabwe, Southern Africa
30
31
32
33
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
Abstract:
34
The “magic mushroom”
Psilocybe cubensis
is cultivated worldwide for recreational and medicinal uses.
35
Described initially from Cuba in 1904, there has been substantial debate about its origin and
36
diversification. The prevailing view, first proposed by the
Psilocybe
expert Gastón Guzmán in 1983, is
37
that
P. cubensis
was inadvertently introduced to the Americas when cattle were introduced to the
38
continents from Africa and Europe (~1500 CE), but that its progenitor was endemic to Africa. This
39
hypothesis has never been tested. Here, we report the discovery of the closest wild relative of
P.
40
cubensis
from sub-Saharan Africa,
P. ochraceocentrata
nom. prov. Using DNA sequences from type
41
specimens of all known and accessable African species of
Psilocybe
, multi-locus phylogenetic and
42
molecular clock analysis strongly support recognizing the African samples as a new species that last
43
shared a common ancestor with
P. cubensis
~1.5 million years ago (~710k - 2.55M years ago 95% HPD).
44
Even at the latest estimated time of divergence, this long predates cattle domestication and the origin of
45
modern humans. Both species are a ssociated with herbivore dung, suggesting this habit likely
46
predisposed
P. cubensis
to its present specialization on domesticated cattle dung. Ecological niche
47
modeling using bioclimatic variables for global records of these species indicates historical presence
48
across Africa, Asia, and the Americas over the last 3 million years. This discovery sheds light on the wild
49
origins of domesticated
P. cubensis
and provides new genetic resources for research on psychedelic
50
mushrooms.
51
INTRODUCTION:
52
Psilocybe
cubensis
(Earle) Singer
is the most widely known, collected, and cultivated "magic
53
mushroom" in the world (A. J. Bradshaw et al., 2022).
P.
cubensis
was first described as
Stropharia
54
cubensis
Earle from a cattle-grazed field in Cuba in 1904 and today is globally distributed where it is
55
common in association with domesticated cattle across subtropical and tropical regions of America,
56
Asia, and Australia (T. Froese et al., 2016; Guzman, 2005; McTaggart et al., 2023; Thomas et al., 2002).
P.
57
cubensis
is one of the core species of psychoactive mushrooms used traditionally and
58
contemporaneously for cultural and spiritual ceremonies across Mexico, and has been domesticated
59
with many strains developed by an active social subculture due to its ease of cultivation (Castro Jauregui
60
et al., 2022; Guzmán, 2008; Van Court et al., 2022).
P. cubensis
is also the target of ongoing biochemical
61
and biomedical studies for drug discovery and whole organism therapies for a wide range of psychiatric
62
illnesses (Blei et al., 2020; Brownstien et al., 2024; Fricke et al., 2017; Lerer et al., 2024; Matsushima et
63
al., 2009; Shahar et al., 2024; Zhuk et al., 2015). Yet, despite the cultural, scientific and medical
64
importance of
P. cubensis
, we know little of its specific ecology or evolutionary origins.
65
The prevailing hypothesis, first proposed by Guzmán (1983), is that
P. cubensis
originated in
66
Africa and was transported to the Americas by Spanish colonizers during the 15th and 16th centuries.
67
This hypothesis was largely predicated on the speculation that other species closely related to it remain
68
to be discovered in Africa, a prediction rooted in the observation that the African continent is historically
69
undersampled for
Psilocybe
(Piepenbring et al., 2020; Tsakem et al., 2024)
.
The closest relatives of
P.
70
cubensis
were recently shown to be from both Asia and Africa, but sampling from these and other
71
regions remains vastly incomplete and definitive statements about origins and biogeography have
72
remained untenable (A. J. Bradshaw et al., 2024). Interestingly, despite substantial popular knowledge
73
and conspicuousness of
P. cubensis
in cattl e-grazed pastu res, it has not been officially confirmed from
74
subtropical or tropical Africa, although casual reports of a
P. cubensis
lookalikes have appeared in peer-
75
reviewed literature (Froese et al. 2016) and online databases such as iNaturalist and MycoPortal.
P.
76
cubensis
does occur in Asia, recorded for Thailand (Ma et al. 2014) and India based on an ITS sequence
77
matching
P. cubensis
in GenBank (accession OK165610.1), and multiple records of it occur across South
78
and Southeast Asia in iNaturalist and MycoPortal.
79
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
One obstacle to determining the existence of named species in a given place at a given time i
s
80
the availability of type specimens as references. Molecular identification is particularly sensitive to th
e
81
lack of type reference sequences because sequences in public databases can carry names that have no
t
82
been validated through comparison with them. In the absence of type reference sequences, this ca
n
83
result in a broad misapplication of names resulting in a collective misunderstanding of species
84
identities, ecologies, and distributions. The recent study by Bradshaw et al. (2024) explicitly targete
d
85
type specimens for whole genome sequencing in an effort to remedy this long-standing problem i
n
86
Psilocybe
. While many type specimens were included in the study, it was not exhaustive, and critica
87
types such as
S. cubensis
and species known only from Africa were not included.
88
Recent fieldwork across sub-Saharan Africa between 2013 and 2022 resulted in multipl
e
89
specimens of an unknown
Psilocybe
sp.
that is superficially similar to
P. cubensis
in habit, habitat, an
d
90
general appearance. Upon further comparison of microscopic and molecular characters with the type o
f
91
S. cubensis
and other
P. cubensis
specimens from around the world, the recognition of the Africa
92
specimens as a distinct species is warranted, provisionally named
Psilocybe ochraceocentrata
. A multi
-
93
locus molecular phylogenetic analysis was used to confirm its close relationship to
P. cubensis
, an
d
94
divergence dating, ecological niche modeling, and species distribution modeling were used to predic
t
95
when, in time, these two lineages diverged and where they were likely to be found over the last
3
96
million years.
97
98
MATERIALS AND METHODS:
99
Collections and Sampling
100
Six collections of mushrooms resembling
P. cubensis
were collected across Zimbabwe and South Africa,
101
occurring on or near decomposing herbivore dung. All collections were made on public or private land
102
with the owner's permission. Sporcarp collections were air-dried over low heat and preserved from
103
insect damage using naphthalene. All other voucher details and deposition are reported in Table 1.
104
105
Table 1: Sample and voucher information
106
107
108
Genomic sequencing, assembly, and barcode extraction from Type specimens.
109
s
e
t
n
d
n
l
e
d
f
n
-
d
t
3
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
Type specimens of the Cubensae complex (Table 1) were extracted and sequenced in the same manner
110
as (A. J. Bradshaw et al., 2024). Hymenophore fragments (5 to 15 mg) from dried fungarium samples
111
were homogenized by placing them in 2.0 mL screw-cap tubes containing a single 3.0-mm and 8 × 1.5-
112
mm stainless steel beads and shaking them in a BeadBug™ microtube homogenizer (Sigma-Aldrich,
113
#Z763713) for 120 s at a speed setting of 3,500 rpm. DNA extraction of mechanically homogenized
114
samples was performed using a Phenol-chloroform DNA extraction protocol. Lysis was performed with
115
Monarch® Genomic DNA Purification Kit (NEB, #T3010S) following the manufacturer’s protocol for
116
Tissue Lysis with an overnight incubation at 56 °C , using a volume of 500 ul lysis buffer, 10ul of
117
Proteinase K and increasing the amount of wash buffer to 550 μL during each of the wash steps.
118
Overnight incubation was then followed by a 2-hour incubation with 4ul of RNase A, after which total
119
lysate was placed in homemade Phase Lock gel tubes made using Dow Corning™ High Vacuum Grease
120
(ASIN:
B001UHMNW0)
along with an equal volume of OmniPur® Phenol:Chloroform:Isoamyl Alcohol
121
(25:24:1, TE-saturated, pH 8.0) solution (MilliporeSigma, Calibiochem #D05686) and then mixed by
122
gentle inversion for 15 min using a fixed speed tube rotator. After mixing, tubes were centrifuged at
123
maximum speed (14,000×g) for 10 min; then, the aqueous (top) layer was transferred to a new phase-
124
lock gel tube and the process repeated. DNA precipitation of the aqueous phase was performed by
125
adding 5 M NaCl to a final concentration of 0.3 M and two volumes of room temperature absolute
126
ethanol, inverting the tubes 20× for thorough mixing followed by an overnight incubation at −20 °C. The
127
next day, DNA was pelleted by centrifugation at 14,000×g for 5 min. The DNA pellet was washed twice
128
with freshly prepared, ice-cold 70% ethanol, air-dried for 15 min at room temperature, and then
129
resuspended in 150 μL of Elution Buffer from the Monarch® Genomic DNA kit.
130
131
DNAs were then cleaned using the Zymo Research™ DNA Clean & Concentrator-5 (#D4003) kit, with the
132
5:1 binding buffer protocol to account for the short fragments common of older herbarium samples.
133
Cleaned gDNAs were submitted to the High Throughput Genomics Core at the University of Utah, where
134
sequencing libraries were prepared using the Nextera™ DNA Flex Library Prep (Illumina®, #20018704)
135
and sequenced on a full lane of Illumina® NovaSeq 6000 PE 2 × 150 bp using an S4 flow cell. SRA
136
accession and biosample numbers are provided in Table 1. Raw reads were trimmed using fastP v0.23.4
137
(Chen et al., 2018), assembled using MetaSPades v3.15.5 (Nurk et al., 2017; Prjibelski et al., 2020), and
138
barcodes were extracted using Pathracer v3.16.0.dev (Shlemov & Korobeynikov, 2019) with custom
139
hidden Markov models (A. Bradshaw et al., 2023). Genome assembly stats and resulting BUSCO (Simão
140
et al., 2015) scores are reported in Supplementary Table 2.
141
142
DNA barcoding and sequencing of additional Specimens
143
The Internal transcribed spacer (ITS) DNA barcodes, commonly used for species delineation (Schoch et
144
al., 2012), of specimens “T1,” T2,” and Harding” were amplified using the primers set ITS1 and ITS4
145
(White et al., 1990). All ITS sequences were sequenced using Sanger sequencing and processed in the
146
same manner as Van Der Merwe et al. (2024). gDNA from dried basidiomata was extracted using the
147
Zymo Research™ Quick-DNA Fungal/BacterialMiniprep kits ( #D6005). Approximately 10 mg of fungal
148
tissue was used for extraction, which was carried out according to the manufacturer’s instructions.
149
Successful DNA extraction was visualized on a 1% agarose gel with ethidium bromide. Amplified
150
barcodes were sequenced at the Central Analytical Facility, Stellenbosch University, using an ABI 3730xl
151
DNA analyzer (Thermo Fisher Scientific™, Waltham, Massachusetts) with a 50 cm capillary array and
152
POP-7. All ITS sequences were deposited into NCBI GenBank (Sayers et al., 2019) with accession
153
numbers reported in Table 1.
154
155
Phylogenetic Inference
156
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
Four independent loci were selected for phylogenetic analysis: the nuclear ribosomal internal
157
transcribed spacers (ITS), translation elongation factor 1-alpha (Ef1a), the largest subunit of RNA
158
polymerase II (RPB1), and the second largest subunit of RNA polymerase II (RPB2). In addition to newly
159
generated sequences, selected sequences from NCBI GenBank (Sayers et al., 2019) and the UNITE
160
database sh_general_release_dynamic_04.04.2024 (Abarenkov et al., 2010), including sequences from
161
recently named species from Africa and elsewhere (Canan et al., 2024; Ostunii et al., 2024; Van Der
162
Merwe et al., 2024) and sequences from type specimens (A. Bradshaw et al., 2023; A. J. Bradshaw et al.,
163
2024); Supplementary data, Table 1), were used for phylogenetic analysis. ITS sequences were trimmed
164
at the 5’ and 3’ end conserved motifs 5’- CATTA- and -GACCT-3’ for downstream analysis (Dentinger et
165
al., 2010). Only the coding sequence of the protein-coding genes were used. ITS barcodes were
166
partitioned into ITS1, 5.8S, and ITS2, using ITSx v1.1.3 (Bengtsson-Palme et al., 2013), aligned separately
167
using MAFFT v7.490 with the flags --maxiterate 1000 --localpair for slow but accurate analysis (Katoh,
168
2002). Concatenation of multiple sequence alignments of the three ITS partitions was achieved using
169
SEGUL 0.22.1 (Handika & Esselstyn, 2024). Phylogenetic analyses were performed using IQTREE2 2.3.6
170
with the flags -m MFP+MERGE -bnni -bb 1000, to enable automatic model finder (Kalyaanamoorthy et
171
al., 2017) and 1000 ultrafast bootstrap replicates (Minh et al., 2013). Evolutionary models were
172
automatically determined for ITS1, 5.8S, and ITS2 partitions separately, and for each codon position for
173
EF1a, RPB1, and RPB2. Edges were linked across all partitions. Phylogenetic trees were rendered in
174
FigTree (http://tree.bio.ed.ac.uk/software/figtree/).
175
176
Molecular dating
177
A reduced dataset was constructed using sequences of the three protein-coding loci for all specimens
178
represented by two or more loci. A concatenated matrix was constructed using Segul (Handika &
179
Esselstyn, 2022), and each codon position was given its own partition. A root calibration normally
180
distributed around a mean of 67.6 Ma (SD +- 6) following Bradshaw et al. (2024), and a calibrated Yule
181
speciation model were used to estimate divergence times. Bayesian inference was conducted using
182
BEAST2 v 2.7.7 (Bouckaert et al., 2019) with Markov Chain Monte Carlo (MCMC) sampling to estimate
183
posterior distributions of phylogenetic parameters. The best fitting models of evolution determined with
184
automatic model finder in IQTREE (Nguyen et al., 2015) were used for each partition. A chain length of 1
185
x 10
8
steps was used, sampling every 5000 steps to ensure a thorough parameter space exploration.
186
Convergence and effective sample sizes (ESS) of parameter estimates were assessed using Tracer v1.7.2
187
(Rambaut et al., 2018), with ESS values above 200 considered indicative of sufficient mixing and
188
convergence. The resulting posterior tree distribution was downsampled to every 2000 trees using
189
LogCombiner due to computational limitations and then summarized using TreeAnnotator v1.10
190
(Drummond & Rambaut, 2007) to produce a maximum clade credibility tree, with node heights
191
representing median posterior estimates. The annotated tree was visualized using FigTree as above,
192
with posterior probabilities indicated on nodes to reflect the robustness of inferred relationships and
193
95% highest posterior density (HPD) of estimated divergence times indicated with bars.
194
195
Species Distribution Modeling
196
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
To predict the ranges of
P. cubensis
and
P. ochraceocentrata
, we conducted species distribution
197
modeling (SDM) using GPS coordinates of known collections and the contemporary 19 bioclimatic
198
variables at 2.5M resolution (Fick & Hijmans, 2017) as environmental predictors. In addition, we sought
199
to predict their historical ranges by conducting SDM of
P. cubensis
using four 19 bioclimatic datasets
200
modeled to have occurred during the Current age, Anthropocene (1979 2013),
last interglacial (LIG)
201
~130KYA , Pleistocene, MIS19 (~787 KYA) and the Pliocene ~3.3Mya (Brown et al., 2018; Dolan et al.,
202
2015; Hill, 2015; Karger et al., 2017) at 2.5M resolution accessed from paleoclim.com (Brown et al.,
203
2018). All collection GPS coordinates used for SDM were pulled from MycoPortal (Miller & Bates, 2017)
204
entries, including data from Mushroom Observer (https://mushroomobserver.org/) and iNaturalist
205
(https://www.inaturalist.org). All data points were filtered to remove non-wild collections, including the
206
removal of all entries with specific mentions of samples being cultivated, confiscated by police, or those
207
labeled as known cultivated strains of
P. cubensis
. The locations of all collections and observations were
208
plotted using GPS coordinates over a global map with R packages ggmap v. 4.0.0 (Kahle & Wickham,
209
2013) and sf v1.0-16 (Pebesma, 2018), and then visualized using ggplot2 v3.5.1 (Wickham, 2016).
210
To account for lumping of
P. ochraceocentrata
in prior records of
P. cubensis
, SDM for all
211
environmental datasets was conducted both with and without African collections of
P .cubensis
, using
212
the SDM R package (v1.2-46). We tested six of the most common SDM models ("bioclim",
213
"domain.dismo", "glm", "gam", "rf", and "svm"), using 1000 random points as “absence” points to
214
validate the model (Supplement Data). The best-performing model, according to AUC, COR Deviance,
215
TSS, MCC, and F1 score, was chosen for each environmental dataset.
216
217
Microscopy and Morphological analysis
218
Microscopic analysis on specimens CS-3006, CS-5783, and CS-3309 was performed using a Leitz Wetzlar
219
Orthoplan microscope with a Leitz Wetzlar drawing tube. All spore measurements were taken from
220
thirty spores whose side-view was clearly visible. Spore prints were studied using Melzer’s Reagent, and
221
sections of dried lamellae were studied in Congo Red. Color descriptions were derived from (Rayner,
222
1970), and codes are recorded parenthetically for each specimen. Microscopic analysis of additional
223
specimens T1, T3, and Harding was performed using Methods outlined in Van Der Merwe et al. (2024).
224
The length and width were measured, and the length/width quotient (Q) was calculated and reported in
225
Supplementary Table 1 for all samples.
226
227
RESULTS
228
Molecular systematics
229
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
Independent phylogenetic analysis of ITS, EF1a, and RPB2 all recovered African specimens originally
230
identified as
P. cubensis
,
P.
cf.
cubensis
, and
P.
cf.
natalensis
as a monophyletic group sister to a clade
231
containing sequences from the holotype of
Stropharia cubensis
Earle (=
Psilocybe cubensis
) (Fig. 1) with
232
strong bootstrap support (ITS=97%, EF1a=99%, RPB2=100%). These same African sequences were not
233
reciprocally monophyletic with respect to sequences of
P. cubensis
using RPB1. Sequences of the types
234
of the other known species of
Psilocybe
endemic to Africa were consistently recovered in other clades.
235
Sequences from the holotype of
Stropharia aquamarina
Pegler were placed sister to
Psilocybe
236
wayanadensis
K.A. Thomas, Manim. & Guzmán
and other species from Asia and Australia (BS; ITS=100,
237
EF1a=92, RPB1=88, RPB2=97, Concat=100). Sequences from the holotype of
Psilocybe natalensis
Gartz,
238
D.A. Reid, M.T. Sm. & Eicker
did not match any publicly available sequences and was placed sister to
239
Psilocybe chuxiongensis
T. Ma & K.D. Hyde (typified from Yunnan, China) and
Psilocybe maluti
B. Van der
240
Merwe, Rockefeller & K. Jacobs (typified from South Africa) ( BS; ITS=99, Concat=99). Sequences of the
241
holotype of
Psilocybe jaliscana
Guzmán and the sequences of
Psilocybe subcubensis
Guzmán were
242
nested within a clade of sequences from
P. cubensis
across all loci. Sequences from the holotype of
243
Psilocybe keralensis
K.A. Thomas, Manim. & Guzmán was recovered in Clade I, and its ITS sequence was
244
identical to the ITS from the type specimen of
Psilocybe ingeli
B. Van der Merwe, Rockefeller & K.
245
Jacobs.
246
247
Molecular dating and geological timeline
248
Due to the close relationship between
P. ochraceocentrata
and
P. cubensis
, we performed molecular
249
dating using Bayesian inference with BEAST of a concatenated dataset of EF1a, RPB1, and RPB2 (Figure
250
1). Phylogenetic reconstruction generated a tree with highly supported nodes (posterior probability
251
>95%). Only a minority of nodes received less than a 0.95 posterior probability and most of these were
252
near the tips among closely related species and species complexes. Molecular dating places the MRCA
253
of
P. ochraceocentrata
and
P. cubensis
at ~1.56 million years ago (MYA)(0.71-2.55 95% HPD). This
254
estimated divergence date corresponds to the Pleistocene epoch (2.5 MYA - 11.7 KYA) following the
255
mass emergence of grass biomes in warm climates with the evolution of the C
4
photosynthetic pathway
256
(8 to 3 MYA) (Edwards et al., 2010).
257
258
259
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
260
Figure 1.
Phylogeny, photo documentation, and geographical distribution of
Psilocybe ochraceocentrata
.
261
LEFT: Maximum clade credibility chronogram of
Psilocybe
spp. resulting from Bayesian divergence
262
analysis of three loci (EF1a, RPB1, RPB2). All nodes received >95% posterior probability except for
263
internal nodes with posterior probabilities indicated. Shaded bars at nodes represent 95%HPD. Shaded
264
bars have been removed from nodes where descendants represent multiple individuals of a species for
265
improved readability. The mean divergence time for the common ancestor of
P. cubensis
and
P.
266
ochraceocentrata
is indicated by a red arrow. Scale bar at the bottom represents the estimated time
267
using a root calibration normally distributed around a mean of 67.6 Ma (SD +- 6) following Bradshaw et
268
al. (2024). Geologic epoch time scale is approximate. Colored vertical bars represent noteworthy global
269
events that illustrate correlation of major divergences within
Psilocybe
: the K-Pg event ~65 MYA, the
270
Eocene Epoch Climate Optimum (EECO), the Terminal Eocene Event (TEE), the Mid-Miocene Climate
271
Optimum (MMCO), and the emergence of C4 grass biomes (C4). Right: in situ photos of
Psilocybe
272
cubensis
(BD1406, top),
P. ochraceocentrata
(CS3006, middle), and records of
P. cubensis
(gold squares)
273
from Mycoportal, iNatrualist, and Mushroom Observer with African records reinterpreted as
P.
274
ochraceocentratta
(Purple squares).
275
276
Ecological niche modeling and species distribution of
P. cubensis
277
.
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
Due to the close relatedness of
P. ochraceocentrata
and
P. cubensis
, we chose to investigate the
278
theoretical species distribution of
P. cubensis
through time using ecological niche modeling (ENM) and
279
species distribution modeling (SDM). To do so, we used publicly available data from MycoPortal, which
280
includes data from iNaturalist and Mushroom Observer. Occurrence data was filtered for only those
281
with georeferences and removed any entries that suggested occurrences were acquired from non-
282
natural sources, such as cultivation. After filtering, we found 1001 occurrences, 12 of which were from
283
the African continent (Supplementary data). Due to the inability to authenticate specimens from Africa
284
as
P. cubensis
, and sparse literature reports of its distribution across Africa and India (Guzmán, 2014;
285
Thomas & Manimohan, 2003) , we performed ENM and SNM in two ways. The first analysis included all
286
geopoints (minus confirmed
P. ochraceocentrata
specimens), assuming that
P. cubensis
specimens from
287
Africa were correctly identified (Figure 2). The second analysis removed African specimens, assuming
288
that these specimens are
P. ochraceocentrata
(Supplementary Figure 5)
.
ENM and SDM were done using
289
multiple geological datasets, ranging from the Pliocene (~3MYA) to the Modern day (Figure 2,
290
Supplementary data).
291
Both data sets exhibited high similarity with ENM and SDM across time, predicting similar
292
species distributions (Figure 2, Supplementary Figure 5). The most extensive predicted ranges occurred
293
across the Southern United States, Central America, and South America (ENM >0.8, SDM =1), with a
294
presence across Southern Africa, Southeast Asia, and Australasia, albeit as a more restricted range
295
(ENM<0.5, SNM=1) (Figure 2). Across time, ENM and SDM indicated reduced distribution across Africa,
296
while the Americas, South East Asia, Australasia, and Europe exhibited an increased presence and
297
distribution up to the Last Interglacial (LIG) ~130 thousand years ago (KYA). Post LIG into contemporary
298
times, ENM and SDM predict distributions more in line with the Pleistocene, MIS19 (~787 KYA) and the
299
Pliocene (~3MYA), but with a slightly increased predicted presence in Africa and distribution across
300
Southeast Asia and Australasia remaining similar to the LIG (Figure 2).
301
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
302
Figure 2.
Ecological niche modeling (ENM) and Species distribution modeling (SDM) of
Psilocybe cubensi
s
303
and
P. ochraceocentrata
through time. All occurrence data for
P. cubensis
from public sources (likely
304
including
P. ochraceocentrata
misidentified as
P. cubensis
in Africa) across three time scales since the
305
mid-Pleistocene.
Top Row:
Contemporary (Modern day),
Middle Row:
Pleistocene, MIS19 (~787 KYA) ,
306
Bottom Row:
Pliocene (~3MYA).
Left:
ENM with distribution likelihood indicated as a heat gradient from
307
purple (0%) to yellow (100%).
Right:
SDM with predicted species presence as present (1, yellow) or
308
absent (0, purple).
309
310
Taxonomy
311
Psilocybe ochraceocentrata
C. Sharp, A. Bradshaw, B. Dentinger & B. van der Merwe
nov. sp.
312
Index Fungorum Registration #: IF 902894
313
Holotype Deposition: Natural History Museum of Zimbabwe, BUL8013
314
GenBank Accession: PQ315824 (ITS), PQ317397( EF1a), PQ317404 (RPB1), PQ317412(RPB2)
315
SRA Accession:
316
317
s
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
Etymology:
318
Pileus with a yellow-ochre center.
319
320
Diagnosis:
321
Typus: Zimbabwe, Matabeleland South Province, The Farmhouse, Kezi Road, Matobo Hills. QDS 2028A4.
322
In thick leaf litter in mixed deciduous woodland on granitic sand. 14 Jan 2013. Collectors C. Sharp & R.
323
Aldridge. Holotype: CS3006 (Field ID, Private Fungarium of C. Sharp), Holotype split: Natural History
324
Museum of Zimbabwe, BUL 8013.
325
326
General field description:
327
Fruiting body
medium-sized, up to 105 mm tall and growing in tight clusters; a pale fungus that bruise
s
328
blue-green.
Pileus
65-75 mm diam.; first cream-colored, then vinaceous-grey (115) with center ochreou
s
329
(44) to fulvous (43); pale yellow towards margin; convex then planate, often undulating; surface finel
y
330
and radially silky with a dull sheen.
Flesh
bright cream-coloured, firm to pithy.
Margin
first down-curved
;
331
edge smooth often with radial cracking.
Lamellae
adnate; first pale then greyish-sepia (106) to brown
-
332
vinaceous (84) to very dark sepia and almost black near margin; face of gill speckled as spores ripen; t
o
333
12 mm deep, thin, papery and fragile; edge thin, smooth or finely scalloped, pale; sparsely t
o
334
moderately crowded with lamellulae, 6-8/cm.
Stipe
central; up to 95 mm long x 9-10 mm; cream
-
335
coloured, bluing where handled; cylindrical, often twisted; apex minutely tufted (x10), streake
d
336
longitudinally, dull sheen, base often with white silky hairs.
Flesh
fibrous in walls and center hollow.
Rin
g
337
median to higher; white first then with dark spores; membranous, striate, very fragile and soon clingin
g
338
to stipe and disintegrating.
Mycelium
is white to cream-coloured, forming a thick, compact mat amongs
t
339
leaf litter.
Bruising
instantly to blue-green (94) where handled or damaged.
Odour
of typical mushroom
340
Spore-print colour
dark vinaceous-grey (116) to purplish-grey (128).
341
342
343
Figure 3:
344
s
s
y
;
-
o
o
-
d
g
g
t
.
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
(A)
Psilocybe ochraceocentrata
. Holotype collection (CS-3006). Photo credit: C.Sharp,(B) Older
345
fruiting body of
Psilocybe ochraceocentrata
; color change possibly due to water logging,
346
(C)
Psilocybe ochraceocentrata
found on dung; (D) Cluster of
Psilocybe ochraceocentrata
(South
347
Africa). Photo credit: Talan Moult, (E) Young sporocarp of
Psilocybe ochraceocentrata
(Harding).
348
Photo credit: Talan Moult.
349
350
Habit and habitat:
351
Miombo woodland, mixed deciduous woodland, all on granitic sand; CS3309 was found growing on old,
352
decomposed herbivore dung of unknown origin.
353
354
Macroscopic Description:
355
Fruiting body
in a tight cluster, to 105 mm tall.
Pileus
to 65 mm diameter; first cream-colored to pale
356
vinaceous-grey (115) with ochreous (44) or fulvous (43) center; first convex then planate; surface finely
357
streaked radially with dull sheen.
Flesh
cream-coloured, firm.
Margin
straw-yellow (40), down-curved,
358
smooth; some radial cracking.
Lamellae
greyish-sepia (106) to brown vinaceous (84), face speckled with
359
ripening spores; adnate; thin, papery, fragile; edge finely scalloped and pale; with lamellulae.
Stipe
to 95
360
mm long x 9 (apex) 10 (mid) 10 (base) mm; cylindrical, twisted or not; apex minutely tufted (x10
361
lens), longitudinally streaked to silky-hairy at the base.
Flesh
hollow with fibrous walls.
Ring
just above
362
median, white, striate, membranous to clinging, and very fragile.
Bruising
immediately to glaucous sky-
363
blue (93) to glaucous blue-green (94).
Odour
typical mushroom.
Spore-print
dark vinaceous-grey (116)
364
to purplish-grey (128).
Chemical reactions:
not recorded
365
366
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
Microscopic description:
367
Figure 4:
368
369
Microscopic illustrations and microscopic feature morphology of
Psilocybe ochraceocentrata
: (A)Basidi
a
370
from Holotype specimen CS-3006; (B): Basidiospores from Holotype specimen CS-3006; (C
371
Cheilocystidia from specimen CS-5783. Microscopic images are derived from specimen
P
372
ochraceocentrata
"Harding.
373
Spores derived from Holotype CS-3006,:
Ranges are given parenthetically with mean values underlined
;
374
ellipsoid, a few lenticulars; thick-walled, smooth-walled with germ-pore; (11)11.8(12.5) x (6.5)7.6(8.0
375
µm; Q = (1.44)1.55(1.71).
Basidia
elongate or clavate and variable in size, 22-38 x 8-12.5 µm; fou
r
376
sterigmata, 4-5µm long with rounded or acute apex.
377
Notes from additionally examined samples:
378
[1]
Spores
can be highly variable in size; note CS3309 has much smaller spores with a few in the extrem
e
379
range (Supplementary Table 1)
380
a
)
.
;
)
r
e
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
[2]
Spores
shape ellipsoid, some collections having more lens-shaped spores than others; thick-walled,
381
smooth-walled with apical germ-pore.
Basidia
elongate or clavate; 22-30 x 8-11 µm; four sterigmata, 4-
382
5µm long with rounded or acute apex.
Cystidia
no pleurocystidia were observed; despite numerous
383
sections, no cheilocystidia were observed in the Type collection but very clear in CS3309 and CS5783;
384
capitate in shape, thin-walled, 21-26 x 9-13µm across the widest part.
Hyphae
clamp connections
385
observed in the Type.
386
[3] CS-2680. Mashonaland Central Province, Mukuvisi Woodlands, Harare. QDS 1731C3. 18 Jan 2012.
387
Collector Daniel Nyamajiwa. The spore range is a good match even if a bit smaller, e.g.
388
(10)10.8(11.5)[13] x [6.5](7)7.4(8) µm but shape slightly different, Q = (1.33)1.4 5 (1.57) (Supplementary
389
Table 1).
390
[4] Even though the ochre-yellow center is diagnostic, one noticeable morphological difference is that
391
the green-blue pigments turn black and remain dark while the specimen is fresh. This may be due to the
392
older fruiting bodies or because they contain more water in a very wet season. All signs of green or black
393
coloration are absent in the dried samples (Figure 2, B).
394
Synonomizations
395
396
Psilocybe cubensis
(Earle) Singer, Sydowia 2(1-6): 37 (1948)
397
398
Basionym:
399
Stropharia cubensis
Earle, Inf. an. Estac. Cent. agr. Cuba 1: 240 (1906)
400
401
Synonyms:
402
=
Naematoloma caerulescens
Pat., Bull. Soc. mycol. Fr. 23(2): 78 (1907)
403
c
Hypholoma caerulescens
(Pat.) Sacc. & Trotter, Syll. fung. (Abellini) 21: 212 (1912)
404
c
Psilocybe cubensis
var.
caerulescens
(Pat.) Singer & A.H. Sm., Mycologia 50(2): 269 (1958)
405
=
Stropharia cyanescens
Murrill, Mycologia 33(3): 279 (1941)
406
c
cubensis
var.
cyanescens
(Murrill) Singer & A.H. Sm., Mycologia 50(2): 269 (1958)
407
=
Psilocybe jaliscana
Guzmán, Docums Mycol. 29(no. 116): 46 (2000)
408
409
DISCUSSION:
410
Psilocybe
has become a world-renowned genus of mushrooms, primarily due to their
411
psychoactivity. Currently,
Psilocybe
mushrooms have been targeted as a source of natural products for
412
use in a global mental health crisis (Carhart-Harris et al., 2017; Daniel & Haberman, 2017; Gandy et al.,
413
2020; Johnson & Griffiths, 2017). However, due to the issue of legality in their collection,
414
characterization of their diversity and studies of their biology have been severely stifled. Despite the
415
~160 species of true
Psilocybe
described around the world, most were described in the Americas
416
(Borovička et al., 2011; Guzmán, 2014; Guzman et al., 2013; Johnston & Buchanan, 1995; Ma et al.,
417
2014; Picker & Rickards, 1970). In contrast, only seven species of
Psilocybe
(including
P.
418
ochraceocentrata)
have been typified from Africa, ranging from the cedar forests of Northern Africa
419
(
Psilocybe mairei)
to the grassland of South Africa (
Psilocybe maluti
) (Guzmán, 2014; Van Der Merwe et
420
al., 2024); many more still undocumented species are predicted.
421
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
The subject of psychoactive mushrooms, in particular
Psilocybe
, and their proposed medical
422
benefits is rapidly gaining interest globally, and no less so in Zimbabwe. Recreational use in Zimbabwe is
423
generally isolated and often dependent on the availability of imported products, likely the frequently
424
cultivated
P. cubensis
(Musshoff et al., 2000). However, historical indigenous knowledge of these fungi is
425
lacking outside of Mesoamerica, making traditional-use claims challenging to validate. However, there
426
have been suggestions that access to the information has been restricted due to a sense of protected
427
information among traditional healers (T. Froese et al., 2016; Guzmán, 2014; Van Der Merwe et al.,
428
2024). This issue is likely a consequence of the historical lack of mycological studies across Africa, which
429
remains one of the most understudied geographic locals for fungal diversity (Antonelli et al., 2024; Crous
430
et al., 2006). Consequently, the people of Zimbabwe's use of
Psilocybe
for ceremonial or medicinal
431
purposes is unknown.
432
Psilocybe
has been previously grouped into four major sections, including the “Cordisporae”,
433
“Mexicanae,” “Zapotecorum,” and “Cubensae” sections. Of these sections, the Cubensae has had little
434
documentation of novel species. The Cubensae section includes species found across Central America,
435
Southeast Asia, India, and Africa, includ ing the most iconic species
P. cubensis
(A. J. Bradshaw et al.,
436
2024; Guzmán, 1983, 1995; Ramírez-Cruz et al., 2013). However, the abundance of diversity within the
437
section remains uncertain, primarily due to a limited representation of type specimens. Type specimens
438
serve as the authoritative description of a species, and their representation is essential to characterize
439
new species definitively. In particular, some specimens of the Cubensae complex, such as
Psilocybe
440
jaliscana
(typified from Mexico) and
Psilocybe aquamarina
(typified from Kenya), were thought to be
441
synonymous with
P. cubensis
and
P. subcubensis
, respectively (Guzman 2014, and personal
442
Communication, Virginia Ramírez-Cruz and Alonso Cortés-Pérez). While the case for
Psilocybe jaliscana
443
being synonymous with
P. cubensis
is true, the same is not true for
P. aquamarina
, which is reported
444
here as genuinely novel.
445
Further, a new issue arises when publicly deposited data with type specimens is validated. The
446
commercially sold “Natal Super Strength (NSS)” (OK491080.1) strain of
P. natalensis
(typified from
447
KwaZulu-Natal) does not ma tch the type specimen of
P. natalensis
. Instead, four of the five publicly
448
deposited sequences cluster with
P. ochraceocentrata
, indicating misidentification. This could lead to
449
future regulatory issues and confusion over the species identity of commercially sold
Psilocybe
strains.
450
Misidentifications at these levels illustrate the difficulty of identifying many species of
Psilocybe
, which
451
often exhibit plastic morphology, necessitating type specimen validation for taxonomic accuracy and
452
stability. This phenomenon is not unique to the Cubensae complex, as it has also been shown to occur
453
across the genus multiple times in commonly collected species (Awan et al., 2018; A. J. Bradshaw et al.,
454
2022, 2024). Without the ability to accurately describe what species are being consumed medically or
455
being commercialized, we jeopardize future work in understanding the species-specific secondary
456
metabolism of
Psilocybe.
With this work, we explicitly targeted type specimens, allowing for direct and
457
accurate comparison of future specimens
of
Psilocybe
from around the world. Notably, we provide the
458
molecular data from the holotype of
Stropharia cubensis
Earle, confirming the identity of the majority of
459
commercially available strains of “magic mushrooms” as
P. cubensis
for the first time
.
460
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
Psilocybe cubensis
is a globally distributed species whose origin is debated. Our ecological
461
modeling suggests that
P. cubensis
could have been found historically but discontinuosly across the
462
Americas, Southeast Asia, Southern and Central Africa, and Australasia between 0.71 and 2.55MYA
463
(Figure 2, Supplementary data). While there is an extensive predicted range, it has been shown that
P.
464
cubensis
from Australia has relatively low genetic diversity, suggesting recent arrival with domesticated
465
cattle to the continent (McTaggart et al., 2023). There are also no authenticated specimens of
P.
466
cubensis
from Africa and only two sequences of it from India (iNat144581118, OK165610.1). While it is
467
possible that
P. cubensis
may occur naturally in Africa,
P. ochraceocentrata
has a strikingly
468
morphological and genetic similarity and shares its almost exclusively ruminant coprophilic habit, which
469
is likely to lead to misidentification. Our divergence analysis suggests that their MRCA likely originated
470
alongside the large herbivores, possibly during the expansion of the C4 grasslands in East Africa 1.8-1.2
471
MYA (Cerling, 1992; Cerling et al., 1988). Coincidentally, this is also the period when
Homo erectus
472
became the dominant hominin in East Africa and the first to spread from Africa through Eurasia via the
473
Levantine corridor alongside large herbivores, including bovids (Antón & Swisher, Iii, 2004; Belmaker,
474
2010; Dennell & Roebroeks, 2005; Zhu et al., 2018). These major migration events present a possible
475
avenue for dispersal of the MRCA of
P. cubensis
and
P. ochraceocentrata
from Africa and their
476
subsequent divergence in Asia and Africa after aridification lead to loss of habitat in the intervening
477
region.
478
P. cubensis
was typified from Cuba in 1904 and is regularly found across the Americas, where it
479
is associated with herbivore dung. However, cattle did not reach the Americas until 1493 (Sluyter, 2023)
480
during the European colonization of Mesoamerica. While the origins and modern domestication of
481
cattle are still debated, there is no question that their evolutionary origins reside in the Old World as
482
opposed to the New World. Domestication of large grazing cattle (
Bos taurus
) and Asian zebu (
Bos
483
indicus
) occurred between ~8-10KYA. The MRCA of both of these domesticated animals, aurochs (
Bos
484
primigenius
), occupied a range from northern Africa to both coasts of Eurasia (Hanotte et al., 2002; Pitt
485
et al., 2019; Zeuner, 1963), overlapping with estimated historic species distribution of
P. cubensis.
486
Precolonial presence of
P. cubensis
in the Americas is not known, but its estimated divergence
487
from
P. ochraceocentrata
and ecological niche modeling suggest it may have existed there before the
488
arrival of Europeans. Bison (
Bison
spp.) are bovids that dominated grasslands in the Americas
489
throughout the Pleistocene and may have facilitated the precolonial arrival and spread of
P. cubensis
.
490
Bison are thought to have arrived from Asia to the Americas in multiple waves, with the first occurring
491
~195–135KYA during the LIG and then the secondc45–21KYA (D. Froese et al., 2017). After their arrival,
492
bison became one of the dominant herbivores in the Americas and were widely distributed across North
493
America and Central America. Bison are known ecological engineers, expanding and maintaining
494
grassland ecosystems wherever they roam (Gates et al., 2010; Ripple et al., 2015), making them
495
important vectors for coprophilic fungi, potentially including
P. cubensis
. The predicted ranges that we
496
see in the ENM and SDM results overlap with reported range of bison across the Americas, including
497
their co-occurrence across the coastlines of Alaska and the Pacific Northwest, and the peninsula of
498
Florida during the LIG (Supplementary data). The occurrence of
P. cubensis
in the Antilles archipelago is
499
most likely a recent development since no large herbivore existed that could have supported it except
500
for giant ground sloths (
Megalocnus
spp.), although this seems unlikely given these were probably
501
browsers with grasses making up a limited portion of their diets (Dantas et al., 2023). While we cannot
502
confirm the Americas are an origin of
P. cubensis
, we also cannot rule it out with the presently available
503
data. However, the inferred divergence of
P. cubensis
and
P. ochraceocentrata
millions of years ago
504
rather than hundreds of years makes the hypothesis that
P. cubensis
was brought to the Americas from
505
its ancestral home in Africa, where it then would have to have gone extinct, less parsimonious.
506
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
The recent characteriz ation of novel species of
Psilocybe
from Africa (Van Der Merwe et al.,
507
2024) indicates that more diversity remains to be described there. The research landscape of
Psilocybe
508
has been heavily stigmatized due to the near-global government restrictions on possession of the
509
psychoactive compounds psilocybin and psilocin. As a consequence of this long-standing impediment,
510
research on
Psilocybe
has a rich history of cooperation between citizens and professional scientists. The
511
study presented here was made possible by the collections and observations of numerous citizen
512
scientists, analysis with modern phylogenetic analysis and ecological modeling, and validation against
513
type specimens maintained in museum collections. Together, these resources provided a robust,
514
collaborative framework that enabled the discovery of
P. ochraceocentrata
and further refined our
515
understanding of the possible geographic origins of one of the world’s most infamous mushrooms,
P.
516
cubensis
. More investment and less regulation on the collecting of fungi, including species and
517
specimens predicted but not proven to contain controlled substances, is the only way to accelerate
518
scientific discovery to gain a more complete understanding of biodiversity and its importance to human
519
well-being before it is lost.
520
521
ACKNOWLEDGEMENTS:
522
The Authors would like to acknowledge Virginia Ramírez-Cruz and Laura Guzmán-Dávalos for assistance
523
in identifying difficult-to-find specimens. Sariah VanderVeur and Toma Ipsen for assistance in sample
524
preparation and help with technical lab duties. We would also like to thank the holding institutions who
525
provided material, much of which is rare and irreplaceable: Instituto de Ecología (INECOL, XAL),
526
University of Leipzig (LZ), Universidad de Guadalajara (IBUG), and Royal Botanic Gardens KEW (K). We
527
would also like to thank Talan Moult for documenting and collecting the South African specimens of
528
Psilocybe ochraceocentrata
used in this study
.
Dr. David Minter is thanked for his thoughts and
529
information on correct Latin designation.
530
531
Data availability:
532
All extracted DNA barcodes have been deposited on NCBI GenBank with corresponding accession
533
numbers reported in Table 1. All raw genomic sequencing data has been deposited in the Short Read
534
Archive (SRA) under Bioproject number PRJNA1159811; Biosample accession numbers are reported in
535
Table 1. All Type specimens derived molecular data has also been provided for RefSeq designation and
536
curation. Any code or specific script requests should be sent to the corresponding author.
537
Supplementary data, including microscopic features, genome assembly statistics, sequence alignments,
538
raw phylogenetic trees, and ecological and species distribution modeling outputs, can be downloaded
539
from the DRYAD data repository at https://doi.org/10.5061/dryad.5x69p8df2.
540
541
Conflict of interest:
542
The authors declare that there is no conflict of interest.
543
544
List of Supplementary Figures:
545
SF1 ML tree of ITS
546
SF2 ML tree of EF1a
547
SF3 ML tree of RPB1
548
SF4 ML tree of RPB2
549
SF5 ENM and SDM without African occurrence data
550
551
List of Supplementary Tables:
552
ST1 Microscopic characteristics and measurements
553
ST2 Genome assembly and BUSCO statistics
554
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
555
Supplementary data:
556
All occurrence data pulled from MycoPortal, including filtered sets used for ENM and SDM
557
558
References:
559
560 Abarenkov, K., Henrik Nilsson, R., Larsson, K., Alexander, I. J., Eberhardt, U., Erland, S., 561
Høiland, K., Kjøller, R., Larsson, E., Pennanen, T., Sen, R., Taylor, A. F. S., Tedersoo, 562
L., Ursing, B. M., Vrålstad, T., Liimatainen, K., Peintner, U., & Kõljalg, U. (2010). The 563
UNITE database for molecular identification of fungi – recent updates and future 564
perspectives. New Phytologist, 186(2), 281–285. https://doi.org/10.1111/j.1469-565
8137.2009.03160.x
566
Antón, S. C., & Swisher, Iii, C. C. (2004). Early Dispersals of Homo from Africa. Annual Review 567
of Anthropology, 33(1), 271–296. 568
https://doi.org/10.1146/annurev.anthro.33.070203.144024
569
Antonelli, A., Teisher, J. K., Smith, R. J., Ainsworth, A. M., Furci, G., Gaya, E., Gonçalves, S. C., 570
Hawksworth, D. L., Larridon, I., Sessa, E. B., Simões, A. R. G., Suz, L. M., Acedo, C., 571
Aghayeva, D. N., Agorini, A. A., Al Harthy, L. S., Bacon, K. L., Chávez
Hernández, M. 572
G., Colli
Silva, M., … Williams, C. (2024). The 2030 Declaration on Scientific Plant and 573
Fungal Collecting. PLANTS, PEOPLE, PLANET, ppp3.10569. 574
https://doi.org/10.1002/ppp3.10569
575
Awan, A. R., Winter, J. M., Turner, D., Shaw, W. M., Suz, L. M., Bradshaw, A. J., Ellis, T., & 576
Dentinger, B. T. M. (2018). Convergent evolution of psilocybin biosynthesis by 577
psychedelic mushrooms. bioRxiv, 374199. https://doi.org/10.1101/374199
578
Belmaker, M. (2010). Early Pleistocene Faunal Connections Between Africa and Eurasia: An 579
Ecological Perspective. In J. G. Fleagle, J. J. Shea, F. E. Grine, A. L. Baden, & R. E. 580
Leakey (Eds.), Out of Africa I (pp. 183–205). Springer Netherlands. 581
https://doi.org/10.1007/978-90-481-9036-2_12
582
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
Bengtsson
Palme, J., Ryberg, M., Hartmann, M., Branco, S., Wang, Z., Godhe, A., De Wit, P., 583
Sánchez
García, M., Ebersberger, I., De Sousa, F., Amend, A., Jumpponen, A., 584
Unterseher, M., Kristiansson, E., Abarenkov, K., Bertrand, Y. J. K., Sanli, K., Eriksson, 585
K. M., Vik, U., … Nilsson, R. H. (2013). Improved software detection and extraction of 586
ITS1 and
ITS
2 from ribosomal
ITS
sequences of fungi and other eukaryotes for analysis 587
of environmental sequencing data. Methods in Ecology and Evolution, 4(10), 914–919. 588
https://doi.org/10.1111/2041-210X.12073
589
Blei, F., Dörner, S., Fricke, J., Baldeweg, F., Trottmann, F., Komor, A., Meyer, F., Hertweck, C., 590
& Hoffmeister, D. (2020). Simultaneous Production of Psilocybin and a Cocktail of 591
β
Carboline Monoamine Oxidase Inhibitors in “Magic” Mushrooms. Chemistry – A 592
European Journal, 26(3), 729–734. https://doi.org/10.1002/chem.201904363
593
Borovi
č
ka, J., Noordeloos, M. E., Gryndler, M., & Oborník, M. (2011). Molecular phylogeny of 594
Psilocybe cyanescens complex in Europe, with reference to the position of the secotioid 595
Weraroa novae-zelandiae. Mycological Progress, 10(2), 149–155. 596
https://doi.org/10.1007/s11557-010-0684-3
597
Bouckaert, R., Vaughan, T. G., Barido-Sottani, J., Duchêne, S., Fourment, M., Gavryushkina, 598
A., Heled, J., Jones, G., Kühnert, D., De Maio, N., Matschiner, M., Mendes, F. K., Müller, 599
N. F., Ogilvie, H. A., Du Plessis, L., Popinga, A., Rambaut, A., Rasmussen, D., Siveroni, 600
I., … Drummond, A. J. (2019). BEAST 2.5: An advanced software platform for Bayesian 601
evolutionary analysis. PLOS Computational Biology, 15(4), e1006650. 602
https://doi.org/10.1371/journal.pcbi.1006650
603
Bradshaw, A. J., Backman, T. A., Ramírez-Cruz, V., Forrister, D. L., Winter, J. M., Guzmán-604
Dávalos, L., Furci, G., Stamets, P., & Dentinger, B. T. M. (2022). DNA Authentication 605
and Chemical Analysis of Psilocybe Mushrooms Reveal Widespread Misdeterminations 606
in Fungaria and Inconsistencies in Metabolites. Applied and Environmental Microbiology, 607
88(24), e01498-22. https://doi.org/10.1128/aem.01498-22
608
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
Bradshaw, A. J., Ramírez-Cruz, V., Awan, A. R., Furci, G., Guzmán-Dávalos, L., & Dentinger, B. 609
T. M. (2024). Phylogenomics of the psychoactive mushroom genus Psilocybe and 610
evolution of the psilocybin biosynthetic gene cluster. Proceedings of the National 611
Academy of Sciences, 121(3), e2311245121. https://doi.org/10.1073/pnas.2311245121
612
Bradshaw, A., Ramírez-Cruz, V., Awan, A., Furci, G., Guzmán-Dávalos, L., & Dentinger, B. 613
(2023). Phylogenomics of the psychoactive mushroom genus Psilocybe and evolution of 614
the psilocybin biosynthetic gene cluster (Version 15, p. 9767543543 bytes) [Dataset]. 615
Dryad. https://doi.org/10.5061/DRYAD.TMPG4F52S
616
Brown, J. L., Hill, D. J., Dolan, A. M., Carnaval, A. C., & Haywood, A. M. (2018). PaleoClim, high 617
spatial resolution paleoclimate surfaces for global land areas. Scientific Data, 5(1), 618
180254. https://doi.org/10.1038/sdata.2018.254
619
Brownstien, M., Lazar, M., Botvinnik, A., Shevakh, C., Blakolmer, K., Lerer, L., Lifschytz, T., & 620
Lerer, B. (2024). Striking long-term beneficial effects of single dose psilocybin and 621
psychedelic mushroom extract in the SAPAP3 rodent model of OCD-like excessive self-622
grooming. Molecular Psychiatry. https://doi.org/10.1038/s41380-024-02786-0
623
Canan, K., Ostunii, S., Rockefeller, A., & Birkebak, J. (2024). Psilocybe caeruleorhiza: A new, 624
cold weather fruiting species of psilocybin containing mushroom from the midwest in 625
section Aztecorum. McIlvainea, 33. https://namyco.org/publications/mcilvainea-journal-626
of-american-amateur-mycology/
627
Carhart-Harris, R. L., Roseman, L., Bolstridge, M., Demetriou, L., Pannekoek, J. N., Wall, M. B., 628
Tanner, M., Kaelen, M., McGonigle, J., Murphy, K., Leech, R., Curran, H. V., & Nutt, D. 629
J. (2017). Psilocybin for treatment-resistant depression: fMRI-measured brain 630
mechanisms. Scientific Reports, 7(1), 13187. https://doi.org/10.1038/s41598-017-13282-631
7
632
Castro Jauregui, O. S., Ramírez-Cruz, V., Bradshaw, A. J., Cortés-Pérez, A., & Guzmán-633
Dávalos, L. (2022). Los hongos sagrados del género Psilocybe en Jalisco. Nubes y 634
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
Ciencia, 12, 14–21.
635
Cerling, T. E. (1992). Development of grasslands and savannas in East Africa during the 636
Neogene. Palaeogeography, Palaeoclimatology, Palaeoecology, 97(3), 241–247. 637
https://doi.org/10.1016/0031-0182(92)90211-M
638
Cerling, T. E., Bowman, J. R., & O’Neil, J. R. (1988). An isotopic study of a fluvial-lacustrine 639
sequence: The Plio-Pleistocene koobi fora sequence, East Africa. Palaeogeography, 640
Palaeoclimatology, Palaeoecology, 63(4), 335–356. https://doi.org/10.1016/0031-641
0182(88)90104-6
642
Chen, S., Zhou, Y., Chen, Y., & Gu, J. (2018). fastp: An ultra-fast all-in-one FASTQ 643
preprocessor. Bioinformatics, 34(17), i884–i890. 644
https://doi.org/10.1093/bioinformatics/bty560
645
Crous, P. W., Rong, I. H., Wood, A., Lee, S., Glen, H., Botha, W., Slippers, B., De Beer, W. Z., 646
Wingfield, M. J., & Hawksworth, D. L. (2006). How many species of fungi are there at the 647
tip of Africa? Studies in Mycology, 55, 13–33. https://doi.org/10.3114/sim.55.1.13
648
Daniel, J., & Haberman, M. (2017). Clinical potential of psilocybin as a treatment for mental 649
health conditions. Mental Health Clinician, 7(1), 24–28. 650
https://doi.org/10.9740/mhc.2017.01.024
651
Dantas, M. A. T., Campbell, S. C., & McDonald, H. G. (2023). Paleoecological inferences about 652
the Late Quaternary giant sloths. Journal of Mammalian Evolution, 30(4), 891–905. 653
https://doi.org/10.1007/s10914-023-09681-5
654
Dennell, R., & Roebroeks, W. (2005). An Asian perspective on early human dispersal from 655
Africa. Nature, 438(7071), 1099–1104. https://doi.org/10.1038/nature04259
656
Dentinger, B. T. M., Margaritescu, S., & Moncalvo, J. (2010). Rapid and reliable 657
high
throughput methods of DNA extraction for use in barcoding and molecular 658
systematics of mushrooms. Molecular Ecology Resources, 10(4), 628–633. 659
https://doi.org/10.1111/j.1755-0998.2009.02825.x
660
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
Dolan, A. M., Haywood, A. M., Hunter, S. J., Tindall, J. C., Dowsett, H. J., Hill, D. J., & 661
Pickering, S. J. (2015). Modelling the enigmatic Late Pliocene Glacial Event—Marine 662
Isotope Stage M2. Global and Planetary Change, 128, 47–60. 663
https://doi.org/10.1016/j.gloplacha.2015.02.001
664
Drummond, A. J., & Rambaut, A. (2007). BEAST: Bayesian evolutionary analysis by sampling 665
trees. BMC Evolutionary Biology, 7(1), 214. https://doi.org/10.1186/1471-2148-7-214
666
Edwards, E. J., Osborne, C. P., Strömberg, C. A. E., Smith, S. A., C Grasses Consortium, Bond, 667
W. J., Christin, P.-A., Cousins, A. B., Duvall, M. R., Fox, D. L., Freckleton, R. P., 668
Ghannoum, O., Hartwell, J., Huang, Y., Janis, C. M., Keeley, J. E., Kellogg, E. A., 669
Knapp, A. K., Leakey, A. D. B., … Tipple, B. (2010). The Origins of C 4 Grasslands: 670
Integrating Evolutionary and Ecosystem Science. Science, 328(5978), 587–591. 671
https://doi.org/10.1126/science.1177216
672
Fick, S. E., & Hijmans, R. J. (2017). WorldClim 2: New 1
km spatial resolution climate surfaces 673
for global land areas. International Journal of Climatology, 37(12), 4302–4315. 674
https://doi.org/10.1002/joc.5086
675
Fricke, J., Blei, F., & Hoffmeister, D. (2017). Enzymatic Synthesis of Psilocybin. Angewandte 676
Chemie International Edition, 56(40), 12352–12355. 677
https://doi.org/10.1002/anie.201705489
678
Froese, D., Stiller, M., Heintzman, P. D., Reyes, A. V., Zazula, G. D., Soares, A. E. R., Meyer, 679
M., Hall, E., Jensen, B. J. L., Arnold, L. J., MacPhee, R. D. E., & Shapiro, B. (2017). 680
Fossil and genomic evidence constrains the timing of bison arrival in North America. 681
Proceedings of the National Academy of Sciences, 114(13), 3457–3462. 682
https://doi.org/10.1073/pnas.1620754114
683
Froese, T., Guzmán, G., & Guzmán-Dávalos, L. (2016). On the Origin of the Genus Psilocybe 684
and Its Potential Ritual Use in Ancient Africa and Europe1. Economic Botany, 70(2), 685
103–114. https://doi.org/10.1007/s12231-016-9342-2
686
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
Gandy, S., Forstmann, M., Carhart-Harris, R. L., Timmermann, C., Luke, D., & Watts, R. (2020). 687
The potential synergistic effects between psychedelic administration and nature contact 688
for the improvement of mental health. Health Psychology Open, 7(2), 689
205510292097812. https://doi.org/10.1177/2055102920978123
690
Gates, C. C., Freese, C. H., Gogan, P. J., & Kotzman, M. (2010). American bison: Status survey 691
and conservation guidelines 2010. IUCN. 692
https://books.google.com/books?hl=en&lr=&id=koUrGx-693
i2ucC&oi=fnd&pg=PR11&dq=C.+C.+Gates,+C.+H.+Freese,+P.+J.+Gogan,+M.+Kotzma694
n,+American+Bison:+Status+Survey+and+Conservation+Guidelines+2010+(IUCN,+Gla695
nd,+Switzerland,+2010).&ots=VYEfbm1jA_&sig=FAszmMQVic5vjaMvgjnjC7MK2EE
696
Guzmán, G. (1983). The genus Psilocybe: A systematic revision of the known species including 697
the history, distribution, and chemistry of the hallucinogenic species. J. Cramer.
698
Guzmán, G. (1995). Supplement to the monograph of the genus Psilocybe.
699
Guzman, G. (2005). Species Diversity of the Genus Psilocybe (Basidiomycotina, Agaricales, 700
Strophariaceae) in the World Mycobiota, with Special Attention to Hallucinogenic 701
Properties. International Journal of Medicinal Mushrooms, 7(1–2), 305–332. 702
https://doi.org/10.1615/IntJMedMushr.v7.i12.280
703
Guzmán, G. (2008). Hallucinogenic Mushrooms in Mexico: An Overview. Economic Botany, 704
62(3), 404–412. https://doi.org/10.1007/s12231-008-9033-8
705
Guzmán, G. (2014). Psilocybe s. Str. (Agaricales, Strophariaceae) in Africa with description of a 706
new species from the Congo. Sydowia An International Journal of Mycology, 66, 43–53. 707
https://doi.org/10.12905/0380.sydowia66(1)2014-0043
708
Guzman, G., Cortes-Perez, A., & Ramirez-Guillen, F. (2013). The Japanese Hallucinogenic 709
Mushrooms Psilocybe and a New Synonym of P. subcaerulipes with Three Asiatic 710
Species Belong to Section Zapotecorum (Higher Basidiomycetes). International Journal 711
of Medicinal Mushrooms, 15(6), 607–615. 712
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
https://doi.org/10.1615/IntJMedMushr.v15.i6.90
713
Handika, H., & Esselstyn, J. (2022). SEGUL: An ultrafast, memory-efficient alignment 714
manipulation and summary tool for phylogenomics [Preprint]. Preprints. 715
https://doi.org/10.22541/au.165167823.30911834/v1
716
Handika, H., & Esselstyn, J. A. (2024). SEGUL: Ultrafast, memory
efficient and mobile
friendly 717
software for manipulating and summarizing phylogenomic datasets. Molecular Ecology 718
Resources, e13964. https://doi.org/10.1111/1755-0998.13964
719
Hanotte, O., Bradley, D. G., Ochieng, J. W., Verjee, Y., Hill, E. W., & Rege, J. E. O. (2002). 720
African Pastoralism: Genetic Imprints of Origins and Migrations. Science, 296(5566), 721
336–339. https://doi.org/10.1126/science.1069878
722
Hill, D. J. (2015). The non-analogue nature of Pliocene temperature gradients. Earth and 723
Planetary Science Letters, 425, 232–241. https://doi.org/10.1016/j.epsl.2015.05.044
724
Johnson, M. W., & Griffiths, R. R. (2017). Potential Therapeutic Effects of Psilocybin. 725
Neurotherapeutics, 14(3), 734–740. https://doi.org/10.1007/s13311-017-0542-y
726
Johnston, P. }, & Buchanan, P. K. (1995). The genus Psilocybe (Agaricales) in New Zealand. 727
New Zealand Journal of Botany, 33(3), 379–388. 728
https://doi.org/10.1080/0028825X.1995.10412964
729
Kahle, D., & Wickham, H. (2013). ggmap: Spatial Visualization with ggplot2. The R Journal, 730
5(1), 144. https://doi.org/10.32614/RJ-2013-014
731
Kalyaanamoorthy, S., Minh, B. Q., Wong, T. K. F., von Haeseler, A., & Jermiin, L. S. (2017). 732
ModelFinder: Fast model selection for accurate phylogenetic estimates. Nature Methods, 733
14(6), 587–589. https://doi.org/10.1038/nmeth.4285
734
Karger, D. N., Conrad, O., Böhner, J., Kawohl, T., Kreft, H., Soria-Auza, R. W., Zimmermann, N. 735
E., Linder, H. P., & Kessler, M. (2017). Climatologies at high resolution for the earth’s 736
land surface areas. Scientific Data, 4(1), 170122. https://doi.org/10.1038/sdata.2017.122
737
Katoh, K. (2002). MAFFT: A novel method for rapid multiple sequence alignment based on fast 738
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
Fourier transform. Nucleic Acids Research, 30(14), 3059–3066. 739
https://doi.org/10.1093/nar/gkf436
740
Lerer, E., Botvinnik, A., Shahar, O., Grad, M., Blakolmer, K., Shomron, N., Lotan, A., Lerer, B., 741
& Lifschytz, T. (2024). Effects of psilocybin, psychedelic mushroom extract and 5-742
hydroxytryptophan on brain immediate early gene expression: Interaction with 743
serotonergic receptor modulators. Frontiers in Pharmacology, 15, 1391412. 744
https://doi.org/10.3389/fphar.2024.1391412
745
Ma, T., Feng, Y., Lin, X.-F., Karunarathna, S. C., Ding, W.-F., & Hyde, K. D. (2014). Psilocybe 746
chuxiongensis, a new bluing species from subtropical China. Phytotaxa, 156(4), 211. 747
https://doi.org/10.11646/phytotaxa.156.4.3
748
Matsushima, Y., Shirota, O., Kikura-Hanajiri, R., Goda, Y., & Eguchi, F. (2009). Effects of 749
Psilocybe argentipes on Marble-Burying Behavior in Mice. Bioscience, Biotechnology, 750
and Biochemistry, 73(8), 1866–1868. https://doi.org/10.1271/bbb.90095
751
McTaggart, A. R., McLaughlin, S., Slot, J. C., McKernan, K., Appleyard, C., Bartlett, T. L., 752
Weinert, M., Barlow, C., Warne, L. N., Shuey, L. S., Drenth, A., & James, T. Y. (2023). 753
Domestication through clandestine cultivation constrained genetic diversity in magic 754
mushrooms relative to naturalized populations. Current Biology, 33(23), 5147-5159.e7. 755
https://doi.org/10.1016/j.cub.2023.10.059
756
Miller, A. N., & Bates, S. T. (2017). The Mycology Collections Portal (MyCoPortal). IMA Fungus, 757
8(2), A65–A66. https://doi.org/10.1007/BF03449464
758
Minh, B. Q., Nguyen, M. A. T., & von Haeseler, A. (2013). Ultrafast Approximation for 759
Phylogenetic Bootstrap. Molecular Biology and Evolution, 30(5), 1188–1195. 760
https://doi.org/10.1093/molbev/mst024
761
Musshoff, F., Madea, B., & Beike, J. (2000). Hallucinogenic mushrooms on the German 762
market—Simple instructions for examination and identification. Forensic Science 763
International, 113(1–3), 389–395. https://doi.org/10.1016/S0379-0738(00)00211-5
764
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
Nguyen, L.-T., Schmidt, H. A., von Haeseler, A., & Minh, B. Q. (2015). IQ-TREE: A Fast and 765
Effective Stochastic Algorithm for Estimating Maximum-Likelihood Phylogenies. 766
Molecular Biology and Evolution, 32(1), 268–274. 767
https://doi.org/10.1093/molbev/msu300
768
Nurk, S., Meleshko, D., Korobeynikov, A., & Pevzner, P. A. (2017). metaSPAdes: A new 769
versatile metagenomic assembler. Genome Research, 27(5), 824–834. 770
https://doi.org/10.1101/gr.213959.116
771
Ostunii, S., Rockefeller, A., Jacobs, J., & Birkebak, J. (2024). Psilocybe niveotropicalis: A new 772
species of psilocybin containing mushroom from South Florida. McIlvainea, 33. 773
https://namyco.org/publications/mcilvainea-journal-of-american-amateur-mycology/
774
Pebesma, E. (2018). Simple Features for R: Standardized Support for Spatial Vector Data. The 775
R Journal, 10(1), 439. https://doi.org/10.32614/RJ-2018-009
776
Picker, J., & Rickards, R. (1970). The occurrence of the psychotomimetic agent psilocybin in an 777
Australian agaric, Psilocybe subaeruginosa. Australian Journal of Chemistry, 23(4), 853. 778
https://doi.org/10.1071/CH9700853
779
Piepenbring, M., Maciá-Vicente, J. G., Codjia, J. E. I., Glatthorn, C., Kirk, P., Meswaet, Y., 780
Minter, D., Olou, B. A., Reschke, K., Schmidt, M., & Yorou, N. S. (2020). Mapping 781
mycological ignorance – checklists and diversity patterns of fungi known for West Africa. 782
IMA Fungus, 11(1), 13. https://doi.org/10.1186/s43008-020-00034-y
783
Pitt, D., Sevane, N., Nicolazzi, E. L., MacHugh, D. E., Park, S. D. E., Colli, L., Martinez, R., 784
Bruford, M. W., & Orozco
terWengel, P. (2019). Domestication of cattle: Two or three 785
events? Evolutionary Applications, 12(1), 123–136. https://doi.org/10.1111/eva.12674
786
Prjibelski, A., Antipov, D., Meleshko, D., Lapidus, A., & Korobeynikov, A. (2020). Using SPAdes 787
De Novo Assembler. Current Protocols in Bioinformatics, 70(1), e102. 788
https://doi.org/10.1002/cpbi.102
789
Rambaut, A., Drummond, A. J., Xie, D., Baele, G., & Suchard, M. A. (2018). Posterior 790
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
Summarization in Bayesian Phylogenetics Using Tracer 1.7. Systematic Biology, 67(5), 791
901–904. https://doi.org/10.1093/sysbio/syy032
792
Ramírez-Cruz, V., Guzmán, G., Villalobos-Arámbula, A. R., Rodríguez, A., Matheny, P. B., 793
Sánchez-García, M., & Guzmán-Dávalos, L. (2013). Phylogenetic inference and trait 794
evolution of the psychedelic mushroom genus Psilocybe sensu lato (Agaricales). Botany, 795
91(9), 573–591. https://doi.org/10.1139/cjb-2013-0070
796
Rayner, R. W. (1970). A mycological colour chart. Commonwealth Mycological Institute
; British 797
Mycological Society.
798
Ripple, W. J., Newsome, T. M., Wolf, C., Dirzo, R., Everatt, K. T., Galetti, M., Hayward, M. W., 799
Kerley, G. I. H., Levi, T., Lindsey, P. A., Macdonald, D. W., Malhi, Y., Painter, L. E., 800
Sandom, C. J., Terborgh, J., & Van Valkenburgh, B. (2015). Collapse of the world’s 801
largest herbivores. Science Advances, 1(4), e1400103. 802
https://doi.org/10.1126/sciadv.1400103
803
Sayers, E. W., Cavanaugh, M., Clark, K., Ostell, J., Pruitt, K. D., & Karsch-Mizrachi, I. (2019). 804
GenBank. Nucleic Acids Research, gkz956. https://doi.org/10.1093/nar/gkz956
805
Schoch, C. L., Seifert, K. A., Huhndorf, S., Robert, V., Spouge, J. L., Levesque, C. A., Chen, W., 806
Fungal Barcoding Consortium, Fungal Barcoding Consortium Author List, Bolchacova, 807
E., Voigt, K., Crous, P. W., Miller, A. N., Wingfield, M. J., Aime, M. C., An, K.-D., Bai, F.-808
Y., Barreto, R. W., Begerow, D., … Schindel, D. (2012). Nuclear ribosomal internal 809
transcribed spacer (ITS) region as a universal DNA barcode marker for Fungi. 810
Proceedings of the National Academy of Sciences, 109(16), 6241–6246. 811
https://doi.org/10.1073/pnas.1117018109
812
Shahar, O., Botvinnik, A., Shwartz, A., Lerer, E., Golding, P., Buko, A., Hamid, E., Kahn, D., 813
Guralnick, M., Blakolmer, K., Wolf, G., Lotan, A., Lerer, L., Lerer, B., & Lifschytz, T. 814
(2024). Effect of chemically synthesized psilocybin and psychedelic mushroom extract 815
on molecular and metabolic profiles in mouse brain. Molecular Psychiatry, 29(7), 2059–816
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
2073. https://doi.org/10.1038/s41380-024-02477-w
817
Shlemov, A., & Korobeynikov, A. (2019). PathRacer: Racing profile HMM paths on assembly 818
graph. https://doi.org/10.1101/562579
819
Simão, F. A., Waterhouse, R. M., Ioannidis, P., Kriventseva, E. V., & Zdobnov, E. M. (2015). 820
BUSCO: Assessing genome assembly and annotation completeness with single-copy 821
orthologs. Bioinformatics, 31(19), 3210–3212. 822
https://doi.org/10.1093/bioinformatics/btv351
823
Sluyter, A. (2023). Cattle in Latin American History. In A. Sluyter, Oxford Research 824
Encyclopedia of Latin American History. Oxford University Press. 825
https://doi.org/10.1093/acrefore/9780199366439.013.1153
826
Thomas, K., & Manimohan, P. (2003). The genus Agrocybe in Kerala State, India. Mycotaxon. 827
https://www.semanticscholar.org/paper/The-genus-Agrocybe-in-Kerala-State%2C-India.-828
Thomas-Manimohan/919ec55a40a6313691253205ee5104b88a3dc96e
829
Thomas, K., Manimohan, P., Guzmán, G., Tapia, F., & Ramirez-Guillen, F. (2002). The genus 830
Psilocybe in Kerala State, India. Mycotaxon -Ithaca Ny-, 83, 195–207.
831
Tsakem, B., Tchamgoue, J., Kinge, R. T., Tiani, G. L. M., Teponno, R. B., & Kouam, S. F. 832
(2024). Diversity of African fungi, chemical constituents and biological activities. 833
Fitoterapia, 178, 106154. https://doi.org/10.1016/j.fitote.2024.106154
834
Van Court, R. C., Wiseman, M. S., Meyer, K. W., Ballhorn, D. J., Amses, K. R., Slot, J. C., 835
Dentinger, B. T. M., Garibay-Orijel, R., & Uehling, J. K. (2022). Diversity, biology, and 836
history of psilocybin-containing fungi: Suggestions for research and technological 837
development. Fungal Biology, 126(4), 308–319. 838
https://doi.org/10.1016/j.funbio.2022.01.003
839
Van Der Merwe, B., Rockefeller, A., Kilian, A., Clark, C., Sethathi, M., Moult, T., & Jacobs, K. 840
(2024). A description of two novel Psilocybe species from southern Africa and some 841
notes on African traditional hallucinogenic mushroom use. Mycologia, 1–14. 842
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
https://doi.org/10.1080/00275514.2024.2363137
843
White, T., Bruns, T., Lee, S., Taylor, J., Innis, M., Gelfand, D., & Sninsky, J. (1990). 844
Amplification and Direct Sequencing of Fungal Ribosomal RNA Genes for 845
Phylogenetics. In Pcr Protocols: A Guide to Methods and Applications, (Vol. 31, pp. 846
315–322).
847
Wickham, H. (2016). ggplot2: Elegant Graphics for Data Analysis (2nd ed. 2016). Springer 848
International Publishing
: Imprint: Springer. https://doi.org/10.1007/978-3-319-24277-4
849
Zeuner, F. E. (1963). A History of Domesticated Animals. Hutchinson.
850
Zhu, Z., Dennell, R., Huang, W., Wu, Y., Qiu, S., Yang, S., Rao, Z., Hou, Y., Xie, J., Han, J., & 851
Ouyang, T. (2018). Hominin occupation of the Chinese Loess Plateau since about 2.1 852
million years ago. Nature, 559(7715), 608–612. https://doi.org/10.1038/s41586-018-853
0299-4
854
Zhuk, O., Jasicka-Misiak, I., Poliwoda, A., Kazakova, A., Godovan, V., Halama, M., & 855
Wieczorek, P. (2015). Research on Acute Toxicity and the Behavioral Effects of 856
Methanolic Extract from Psilocybin Mushrooms and Psilocin in Mice. Toxins, 7(4), 1018–857
1029. https://doi.org/10.3390/toxins7041018
858
859
.CC-BY-NC 4.0 International licenseavailable under a
(which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made
The copyright holder for this preprintthis version posted December 7, 2024. ; https://doi.org/10.1101/2024.12.03.626483doi: bioRxiv preprint
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
Obsessive compulsive disorder (OCD) is a highly prevalent disorder that causes serious disability. Available treatments leave 40% or more of people with OCD significantly symptomatic. There is an urgent need for novel therapeutic approaches. Mice that carry a homozygous deletion of the SAPAP3 gene (SAPAP3 KO) manifest a phenotype of excessive self-grooming, tic-like head-body twitches and anxiety. These behaviors closely resemble pathological self-grooming behaviors observed in humans in conditions that overlap with OCD. Following a preliminary report that the tryptaminergic psychedelic, psilocybin, may reduce symptoms in patients with OCD, we undertook a randomized controlled trial of psilocybin in 50 SAPAP3 KO mice (28 male, 22 female). Mice that fulfilled inclusion criteria were randomly assigned to a single intraperitoneal injection of psilocybin (4.4 mg/kg), psychedelic mushroom extract (encompassing the same psilocybin dose) or vehicle control and were evaluated after 2, 12, and 21 days by a rater blind to treatment allocation for grooming characteristics, head-body twitches, anxiety, and other behavioral features. Mice treated with vehicle (n = 18) manifested a 118.71 ± 95.96% increase in total self-grooming (the primary outcome measure) over the 21-day observation period. In contrast, total self-grooming decreased by 14.60 ± 17.90% in mice treated with psilocybin (n = 16) and by 19.20 ± 20.05% in mice treated with psychedelic mushroom extract (n = 16) (p = 0.001 for effect of time; p = 0.0001 for time × treatment interaction). Five mice were dropped from the vehicle group because they developed skin lesions; 4 from the psilocybin group and none from the psychedelic mushroom extract group. Secondary outcome measures such as head-body twitches and anxiety all showed a significant improvement over 21 days. Notably, in mice that responded to psilocybin (n = 12) and psychedelic mushroom extract (n = 13), the beneficial effect of a single treatment persisted up to 7 weeks. Mice initially treated with vehicle and non-responsive, showed a clear and lasting therapeutic response when treated with a single dose of psilocybin or psychedelic mushroom extract and followed for a further 3 weeks. While equivalent to psilocybin in overall effect on self-grooming, psychedelic mushroom extract showed superior effects in alleviating head-body twitches and anxiety. These findings strongly justify clinical trials of psilocybin in the treatment of OCD and further studies aimed at elucidating mechanisms that underlie the long-term effects to alleviate excessive self-grooming observed in this study. Prepared with BioRender (https://www.biorender.com/).
Article
Full-text available
Phylogenetic studies now routinely require manipulating and summarizing thousands of data files. For most of these tasks, currently available software requires considerable computing resources and substantial knowledge of command‐line applications. We develop an ultrafast and memory‐efficient software, SEGUL, that performs common phylogenomic dataset manipulations and calculates statistics summarizing essential data features. Our software is available as standalone command‐line interface (CLI) and graphical user interface (GUI) applications, and as a library for Rust, R and Python, with possible support of other languages. The CLI and library versions run native on Windows, Linux and macOS, including Apple ARM Macs. The GUI version extends support to include mobile iOS, iPadOS and Android operating systems. SEGUL leverages the high performance of the Rust programming language to offer fast execution times and low memory footprints regardless of dataset size and platform choice. The inclusion of a GUI minimizes bioinformatics barriers to phylogenomics while SEGUL's efficiency reduces economic barriers by allowing analysis on inexpensive hardware. Our support for mobile operating systems further enables teaching phylogenomics where access to computing power is limited.
Article
Full-text available
Background: Immediate early genes (IEGs) are rapidly activated and initiate diverse cellular processes including neuroplasticity. We report the effect of psilocybin (PSIL), PSIL-containing psychedelic mushroom extract (PME) and 5-hydroxytryptophan (5-HTP) on expression of the IEGs, cfos, egr1, and egr2 in mouse somatosensory cortex (SSC). Methods: In our initial experiment, male C57Bl/6j mice were injected with PSIL 4.4 mg/kg or 5-HTP 200 mg/kg, alone or immediately preceded by serotonergic receptor modulators. IEG mRNA expression 1 hour later was determined by real time qPCR. In a replication study a group of mice treated with PME was added. Results: In our initial experiment, PSIL but not 5-HTP significantly increased expression of all three IEGs. No correlation was observed between the head twitch response (HTR) induced by PSIL and its effect on the IEGs. The serotonergic receptor modulators did not significantly alter PSIL-induced IEG expression, with the exception of the 5-HT2C antagonist (RS102221), which significantly enhanced PSIL-induced egr2 expression. 5-HTP did not affect IEG expression. In our replication experiment, PSIL and PME upregulated levels of egr1 and cfos while the upregulation of egr2 was not significant. Conclusions: We have shown that PSIL and PME but not 5-HTP (at a dose sufficient to induce HTR), induced a significant increase in cfos and egr1 expression in mouse SSC. Our findings suggest that egr1 and cfos expression may be associated with psychedelic effects.
Article
Full-text available
Psychoactive mushrooms in the genus Psilocybe have immense cultural value and have been used for centuries in Mesoamerica. Despite the recent surge of interest in these mushrooms due to the psychotherapeutic potential of their natural alkaloid psilocybin, their phylogeny and taxonomy remain substantially incomplete. Moreover, the recent elucidation of the psilocybin biosynthetic gene cluster is known for only five of ~165 species of Psilocybe , four of which belong to only one of two major clades. We set out to improve the phylogeny of Psilocybe using shotgun sequencing of fungarium specimens, from which we obtained 71 metagenomes including from 23 types, and conducting phylogenomic analysis of 2,983 single-copy gene families to generate a fully supported phylogeny. Molecular clock analysis suggests the stem lineage of Psilocybe arose ~67 mya and diversified ~56 mya. We also show that psilocybin biosynthesis first arose in Psilocybe , with 4 to 5 possible horizontal transfers to other mushrooms between 40 and 9 mya. Moreover, predicted orthologs of the psilocybin biosynthetic genes revealed two distinct gene orders within the biosynthetic gene cluster that corresponds to a deep split within the genus, possibly a signature of two independent acquisitions of the cluster within Psilocybe .
Article
Full-text available
Therapeutic psychedelic administration and contact with nature have been associated with the same psychological mechanisms: decreased rumination and negative affect, enhanced psychological connectedness and mindfulness-related capacities, and heightened states of awe and transcendent experiences, all processes linked to improvements in mental health amongst clinical and healthy populations. Nature-based settings can have inherently psychologically soothing properties which may complement all stages of psychedelic therapy (mainly preparation and integration) whilst potentiating increases in nature relatedness, with associated psychological benefits. Maximising enhancement of nature relatedness through therapeutic psychedelic administration may constitute an independent and complementary pathway towards improvements in mental health that can be elicited by psychedelics.
Article
Full-text available
Abstract Scientific information about biodiversity distribution is indispensable for nature conservation and sustainable management of natural resources. For several groups of animals and plants, such data are available, but for fungi, especially in tropical regions like West Africa, they are mostly missing. Here, information for West African countries about species diversity of fungi and fungus-like organisms (other organisms traditionally studied by mycologists) is compiled from literature and analysed in its historical context for the first time. More than 16,000 records of fungi representing 4843 species and infraspecific taxa were found in 860 publications relating to West Africa. Records from the Global Biodiversity Information Facility (GBIF) database (2395 species), and that of the former International Mycological Institute fungal reference collection (IMI) (2526 species) were also considered. The compilation based on literature is more comprehensive than the GBIF and IMI data, although they include 914 and 679 species names, respectively, which are not present in the checklist based on literature. According to data available in literature, knowledge on fungal richness ranges from 19 species (Guinea Bissau) to 1595 (Sierra Leone). In estimating existing species diversity, richness estimators and the Hawksworth 6:1 fungus to plant species ratio were used. Based on the Hawksworth ratio, known fungal diversity in West Africa represents 11.4% of the expected diversity. For six West African countries, however, known fungal species diversity is less than 2%. Incomplete knowledge of fungal diversity is also evident by species accumulation curves not reaching saturation, by 45.3% of the fungal species in the checklist being cited only once for West Africa, and by 66.5% of the fungal species in the checklist reported only for a single country. The documentation of different systematic groups of fungi is very heterogeneous because historically investigations have been sporadic. Recent opportunistic sampling activities in Benin showed that it is not difficult to find specimens representing new country records. Investigation of fungi in West Africa started just over two centuries ago and it is still in an early pioneer phase. To promote proper exploration, the present checklist is provided as a tool to facilitate fungal identification in this region and to aid conceptualisation and justification of future research projects. Documentation of fungal diversity is urgently needed because natural habitats are being lost on a large scale through altered land use and climate change.
Article
Full-text available
High-resolution, easily accessible paleoclimate data are essential for environmental, evolutionary, and ecological studies. The availability of bioclimatic layers derived from climatic simulations representing conditions of the Late Pleistocene and Holocene has revolutionized the study of species responses to Late Quaternary climate change. Yet, integrative studies of the impacts of climate change in the Early Pleistocene and Pliocene – periods in which recent speciation events are known to concentrate – have been hindered by the limited availability of downloadable, user-friendly climatic descriptors. Here we present PaleoClim, a free database of downscaled paleoclimate outputs at 2.5-minute resolution (~5 km at equator) that includes surface temperature and precipitation estimates from snapshot-style climate model simulations using HadCM3, a version of the UK Met Office Hadley Centre General Circulation Model. As of now, the database contains climatic data for three key time periods spanning from 3.3 to 0.787 million years ago: the Marine Isotope Stage 19 (MIS19) in the Pleistocene (~787 ka), the mid-Pliocene Warm Period (~3.264–3.025 Ma), and MIS M2 in the Late Pliocene (~3.3 Ma).
Article
Full-text available
Motivation Quality control and preprocessing of FASTQ files are essential to providing clean data for downstream analysis. Traditionally, a different tool is used for each operation, such as quality control, adapter trimming and quality filtering. These tools are often insufficiently fast as most are developed using high-level programming languages (e.g. Python and Java) and provide limited multi-threading support. Reading and loading data multiple times also renders preprocessing slow and I/O inefficient. Results We developed fastp as an ultra-fast FASTQ preprocessor with useful quality control and data-filtering features. It can perform quality control, adapter trimming, quality filtering, per-read quality pruning and many other operations with a single scan of the FASTQ data. This tool is developed in C++ and has multi-threading support. Based on our evaluation, fastp is 2–5 times faster than other FASTQ preprocessing tools such as Trimmomatic or Cutadapt despite performing far more operations than similar tools. Availability and implementation The open-source code and corresponding instructions are available at https://github.com/OpenGene/fastp.
Article
GenBank® (www.ncbi.nlm.nih.gov/genbank/) is a comprehensive, public database that contains over 6.25 trillion base pairs from over 1.6 billion nucleotide sequences for 450 000 formally described species. Daily data exchange with the European Nucleotide Archive (ENA) and the DNA Data Bank of Japan (DDBJ) ensures worldwide coverage. Recent updates include a new version of Genome Workbench that supports GenBank submissions, new submission wizards for viral genomes, enhancements to BankIt and improved handling of taxonomy for sequences from pathogens.
Article
Simple features are a standardized way of encoding spatial vector data (points, lines, polygons) in computers. The sf package implements simple features in R, and has roughly the same capacity for spatial vector data as packages sp, rgeos, and rgdal. We describe the need for this package, its place in the R package ecosystem, and its potential to connect R to other computer systems. We illustrate this with examples of its use.