Hymenogastraceae

Psilocybe cubensis

Not a plant, though indeed magic.

Bibliography

  1. G. Adam, J. Gartz, S. Lang, and H.-M. Vorbrodt, “Verfahren zur herstellung pilzlicher biomassen,” Feb-1990. https://patents.google.com/patent/DD276420A1/en?inventor=JOCHEN+GARTz&oq=JOCHEN+GARTz.
  2. A. M. Adams et al., “In Vivo Production of Psilocybin in E. Coli,” Metabolic Engineering, vol. 56, pp. 111–119, Dec. 2019. doi: 10.1016/j.ymben.2019.09.009.
  3. S. Agurell, S. Blomkvist, and P. Catalfomo, “Biosynthesis of Psilocybin in Submerged Culture of Psilocybe Cubensis. 1. Incorporation of Labelled Tryptophan and Tryptamine,” Acta Pharmaceutica Suecica, vol. 3, no. 1, pp. 37–44, Feb. 1966.
  4. S. Agurell, J. L. G. Nilsson, S. Liaaen-Jensen, U. Schwieter, and J. Paasivirta, “Biosynthesis of Psilocybin. Part II. Incorporation of Labelled Tryptamine Derivatives.,” Acta Chemica Scandinavica, vol. 22, pp. 1210–1218, 1968. doi: 10.3891/acta.chem.scand.22-1210.
  5. G. Albán-J., M. Ã.-a Terán, M. Ã.-a Robles-U., F. Quinde, and N. Ã. Niveiro, “Psilocybe Cubensis (Agaricales, Basidiomycota) En Ecuador,” Lilloa, vol. 58, no. 2, pp. 86–94, Dec. 2021. doi: 10.30550/j.lil/2021.58.1/2021.09.30.
  6. J. W. Allen, P. O. Box, J. Gartz, F. C. Suarez, P. Sihanonth, and P. Road, “The Occurrence, Cultivation, and Chemistry of Psilocybe Ovoideocystidiata, a New Bluing Species (Agaricales) from Ohio, Pennsy.” Dec-2009. https://www.researchgate.net/publication/235436020.
    Cultivation and analysis of Psilocybe ovoideocystidiata, a new bluing species from Ohio and Bethany West Virginia is presented. Cultivation of this species was demonstrated on hardwood substrate. Analysis of both caps and stems revealed the presence of psilocybin, in most cases psilocin and always low concentrations of baeocystin. Psilocybin, psilocin and baeocystin levels varied in the bluing caps and stems of this new species. The highest concentrations of these alkaloids were found in both naturally grown and cultivated fruiting bodies of Psilocybe ovideocystidiata which, at the present moment is an indigenous species found in Ohio, Pennsylvania and West Virginia.. The relative alkaloidal content of psilocybin, psilocin and baeocystin found in Psilocybe ovoideocystidiata from Ohio was similar to that measured in Psilocybe caerulipes by Leung et al. Recent comparative chemical analysis of both species was unable to be performed due to a denial of specimens through the University of Michigan’s herbarium.
  7. J. W. Allen and M. D. Merlin, “Psychoactive Mushroom Use in Koh Samui and Koh Pha-Ngan, Thailand,” Journal of Ethnopharmacology, vol. 35, no. 3, pp. 205–228, Jan. 1992. doi: 10.1016/0378-8741(92)90020-R.
    This paper presents the results of recent ethnomycological exploration in southern Thailand. Field observations, interviews and collection of fungi specimens were carried out primarily on two islands, Koh Samui and Koh Pha-ngan, situated in the western region of the Gulf of Siam. Some fieldwork was also conducted in the northern Thai province of Chiang Dao and in the southern Thai pro- vince of Surat Thani. During five separate excursions (1989–90), observations were made of occurrence, harvesting, use, and marketing of psychoactive fungi by local Thai natives (males and females, adults and children), foreign tourists, and German im- migrants. The first records of psychoactive Psilocybe subcubensis and Copelandia dung fungi in Thailand are presented in this paper. These fungi exhibited intense bluing reactions when handled, indicating the presence of psilocybin and/or psilocin. Seven collections of Psilocybe cubensis (Earle) Singer and/or Psilocybe subcubensis Guzman and four collections of Copelandia sp. were harvested and sun-dried for herbarium deposit. These fungi are cultivated or occur spontaneously, often appearing in the decomposed manure of domesticated water buffalo (Bubalus bubalis) and at least three different species of cattle (Bos indicus, B. guarus, and B. sundaicus). The psychoactive fungi are cultivated in clandestine plots, both indoors and outdoors, in the uplands and villages on Koh Samui by both Thai natives and some foreigners. The sale of psychoactive fungi directly to tourists and to resort restaurants for use in edible food items such as omelettes and soups is discussed in detail. The preparation and sale of mushroom omelettes adulterated with artificial hallucinogens in some restaurants is also discussed. In addition, the marketing of items such as hand painted T-shirts, post cards, and posters bearing mushroom related motifs in Thailand is described.
  8. N. Anastos, N. W. Barnett, F. M. Pfeffer, and S. W. Lewis, “Investigation into the Temporal Stability of Aqueous Standard Solutions of Psilocin and Psilocybin Using High Performance Liquid Chromatography,” Science & Justice, vol. 46, no. 2, pp. 91–96, Apr. 2006. doi: 10.1016/S1355-0306(06)71579-9.
  9. C. Andersson, J. Kristinsson, N. C. of Ministers, and J. Gry, Occurrence and Use of Hallucinogenic Mushrooms Containing Psilocybin Alkaloids. Nordic Council of Ministers, 2009.
    In some parts of the world mushrooms have had a central role in religious ritual ceremonies. Ethnomycological studies among the Indian tribes of Mexico - the Aztecs and the Chichimecas - revealed the mushrooms to be hallucinogenic. Chemists from a leading Pharmaceutical company took over, isolated and described the mushroom alkaloid psilocybin, that upon dephosphorylation after collection of the mushroom or in the human body, form psilocin that is the active hallucinogenic compound. For a long time psilocybin/psilocin was expected to become a constituent of psychedelic drugs useful for treatment of specific psychoses. As the effect of psilocybin/psilocin resembles that of LSD the isolated compound, as well as mushrooms containing the compound, became popular among recreational users of hallucinogenic drugs in Western America, and from there the habit of using these mushrooms have spread around the world. Psilocybin/psilocin is legally prohibited in many countries which usually treat the compound as a narcotic drug. Some countries also prohibit the use of some or all psilocybin-containing mushrooms. In this respect, the legal situation differs between Nordic countries. Although psilocybin-containing mushrooms are not what Nordic mushroom pickers are trying to find as food or food supplement, there is a risk, admittedly small, that these mushrooms accidentally will be collected. At the present situation, this may be a legal problem in some Nordic countries. This document aims at identifying when this might be the case without going into legal interpretations.
  10. C. Angelini and E. Battistin, “Observatins of Psilocybe cubensis and P. subcubensis.,” Rivista di Micologia, vol. 54, no. 3, pp. 251–261, 2011. https://www.cabdirect.org/cabdirect/abstract/20123040940.
    The finding of Psilocybe cubensis in the Dominican Republic, which probably represents one of the first reports of such a species for that country, has given the opportunity to study specimens growing in their own natural habitat, instead of - as it often happens in the literature - fungi cultivated at home on compost heaps as a recreational drug or growing in the open in temperate countries as a...
  11. A. R. Awan et al., “Convergent Evolution of Psilocybin Biosynthesis by Psychedelic Mushrooms.” bioRxiv, p. 374199, Jul-2018. doi: 10.1101/374199.
    Psilocybin is a psychoactive compound with clinical applications produced by dozens of mushroom species1. There has been a longstanding interest in psilocybin research with regard to treatment for addiction2, depression3, and end-of-life suffering4. However, until recently very little was known about psilocybin biosynthesis and its ecological role. Here we confirm and refine recent findings5 about the genes underpinning psilocybin biosynthesis, discover that there is more than one psilocybin biosynthesis cluster in mushrooms, and we provide the first data directly addressing psilocybin’s ecological role. By analysing independent genome assemblies for the hallucinogenic mushrooms Psilocybe cyanescens and Pluteus salicinus we recapture the recently discovered psilocybin biosynthesis cluster5,6 and show that a transcription factor previously implicated in its regulation is actually not part of the cluster. Further, we show that the mushroom Inocybe corydalina produces psilocybin but does not contain the established biosynthetic cluster, and we present an alternative cluster. Finally, a meta-transcriptome analysis of wild-collected mushrooms provides evidence for intra-mushroom insect gene expression of flies whose larvae grow inside Psilocybe cyanescens. These larvae were successfully reared into adults. Our results show that psilocybin does not confer complete protection against insect mycophagy, and the hypothesis that it is produced as an adaptive defense compound may need to be reconsidered.
  12. E. R. Badham and D. T. Kincaid, “Analysis of Anemotropism in the Mushroom Psilocybe Cubensis,” Canadian Journal of Botany, vol. 62, no. 2, pp. 296–300, Feb. 1984. doi: 10.1139/b84-047.
    The bending of the growing basidiocarp of Psilocybe cubensis into air flow (anemotropism) was investigated. Basidiocarps were placed in a wind tunnel under controlled conditions and anemotropism was evaluated in relation to the physical dimensions of the mushroom and to certain environmental factors. Six lines of evidence are presented demonstrating that the cause of the bending is a greater water loss from the windward side than from the leeward side of the mushroom. Dehydration may result in unequal growth owing to differential production or activity of a hypothetical growth factor or because of a differential in turgor pressure.
  13. E. R. Badham, “The Effect of Light Upon Basidiocarp Initiation in Psilocybe Cubensis,” Mycologia, vol. 72, no. 1, pp. 136–142, Jan. 1980. doi: 10.1080/00275514.1980.12021162.
    Formation of basidiocarp initials in Psilocybe cubensis occurred only when cultures were illuminated. Short durations of light (0.0025 sec of xenon-arc flash) were sufficient for initiation. Light-induced initiation was saturated at a dose of 0.345 × 104 ergs/cm2 at 460 nm. UV and blue wavelengths of 370, 440, and 460 nm were the most effective. Green and red wavelengths greater than 510 nm were ineffective.
  14. E. R. Badham, “Ethnobotany of Psilocybin Mushrooms, Especially Psilocybe Cubensis,” Journal of Ethnopharmacology, vol. 10, no. 2, pp. 249–254, Apr. 1984. doi: 10.1016/0378-8741(84)90007-2.
    Several fungi species collected in the Hawaiian Islands have been reported to be psychoactive. Previous chemical analyses together with the present study indicate that 5 coprophilous and one non-coprophilous species occurring in the islands are now known to contain psychoactive alkaloids. At least some of these species are consumed in the Hawaiian Islands, as well as elsewhere, for non-traditional, recreational purposes. These include Copelandia cyanescens (Berk, et Br.) Singer, Copelandia tropicalis (Ola’h) Singer and Weeks (syn. Panaeolus tropicalis Ola’h), Copelandia anomala Murrill, and Panaeolus subbalteatus (Berk, and Br.) Sacc., which have already been described from the Hawaiian Islands. Three more mind-altering fungi and one non-psychoactive species are reported from this archipelago for the first time. These psychoactive fungi include Copelandia bispora (Malençon et Bertault) Singer and Weeks from O’ahu, Copelandia cambodginiensis (Ola’h et Heim) Singer and Weeks from O’ahu, and Amanita muscaria (L.) Hooker from Kaua’i. Panaeolus goossensiae Beeli identified from O’ahu contains tryptamine compounds; however, the psychoactive alkaloids psilocybin and psilocin were not found in this dung species. The mixed serotonin (5-HT) 1A/2A/2B/2C/6/7 receptor agonist psilocybin dose-dependently induces an altered state of consciousness (ASC) that is characterized by changes in sensory perception, mood, thought, and the sense of self. The psychological effects of psilocybin are primarily mediated by 5-HT2A receptor activation. However, accumulating evidence suggests that 5-HT1A or an interaction between 5-HT1A and 5-HT2A receptors may contribute to the overall effects of psilocybin. Therefore, we used a double-blind, counterbalanced, within-subject design to investigate the modulatory effects of the partial 5-HT1A agonist buspirone (20 mg p.o.) and the non-hallucinogenic 5-HT2A/1A agonist ergotamine (3 mg p.o.) on psilocybin-induced (170 µg/kg p.o.) psychological effects in two groups (n=19, n=17) of healthy human subjects. Psychological effects were assessed using the Altered State of Consciousness (5D-ASC) rating scale. Buspirone significantly reduced the 5D-ASC main scale score for Visionary Restructuralization (VR) (p<0.001), which was mostly driven by a reduction of the VR item cluster scores for elementary and complex visual hallucinations. Further, buspirone also reduced the main scale score for Oceanic Boundlessness (OB) including derealisation and depersonalisation phenomena at a trend level (p=0.062), whereas ergotamine did not show any effects on the psilocybin-induced 5D-ASC main scale scores. The present finding demonstrates that buspirone exerts inhibitory effects on psilocybin-induced effects, presumably via 5-HT1A receptor activation, an interaction between 5-HT1A and 5-HT2A receptors, or both. The data suggest that the modulation of 5-HT1A receptor activity may be a useful target in the treatment of visual hallucinations in different psychiatric and neurological diseases. Plants are repository of novel products of medicinal value that are necessary for production of new drugs with application in the medical and pharmaceutical industries. The recent advances in the application of metabolomics techniques have enhanced the application of analytical techniques for effective quantification useful metabolites. This chapter gives an insight on the principles and classification of primary and secondary plant metabolites and provides detailed information on the biogenesis of carbohydrates, lipids, volatile oils, and resins. Also, this chapter provides information on the medicinal application of these metabolites. The use of psilocybin as treatment for major depressive disorder (MDD) has been examined as a promising alternative to traditional first-line options. We reviewed existing literature to provide a synthesis of the extant neuroimaging observations with psilocybin, and to identify putative therapeutic targets for target engagement studies with psilocybin, and potentially other psychedelics. We assessed neuroimaging observations with psilocybin among participants with MDD and healthy populations. A systematic search was conducted on PubMed, Google Scholar and PsycINFO from database inception to November 17th, 2021. The study quality (i.e., risk of bias) was assessed using the revised Cochrane risk-of-bias tool for randomized trials. A total of ten studies evaluated psilocybin in healthy populations and three studies assessed psilocybin in MDD participants using neuroimaging techniques. Following psilocybin administration, a decrease in amygdala activity and a reduction in depressive symptoms was observed in two studies. Changes in functional connectivity and activation of prefrontal limbic structures, specifically the ventral medial prefrontal cortex and amygdala, was seen in healthy populations. There was high heterogeneity in methodology (e.g., dosing schedule and imaging methods) amongst included studies. Longitudinal studies are needed to further elucidate psilocybin treatment for MDD, its long-term effects and the possibility of sustained therapeutic effects. The incidence of fungal poisoning varies considerably globally and is related to local habits, economic factors and lifestyle. Mushroom poisoning is mostly an accidental result of a mix-up between edible and toxic fungi. However, intentional ingestion of psychotropic (‘magic’) mushrooms has become a problem. Among thousands of mushroom species worldwide, fewer than a hundred are severely toxic. Most fungal toxins cause mild or moderate poisoning, often only gastroenteritis; the ingestion of a few species of extremely poisonous fungi defines the medical dimension of the problem. The most dreaded poisonings are those caused by cytotoxic fungi, for example amatoxins in death cap (Amanita phalloides) and destroying angel (Amanita virosa); both cause severe gastroenteritis and liver damage. Orellanine, occurring in certain Cortinarius spp., can induce severe and persistent kidney damage. Dramatic, but rarely lethal, effects are caused by fungi containing neurotoxins such as muscarine (Clitocybe and Inocybe spp.), psilocybin (Psilocybe and Panaeolus spp. – ‘magic’ mushrooms), isoxazoles (fly agaric, panther cap) and gyromitrin (false morels). Treatment is focused on general symptomatic and supportive care, although antidotes exist for fungi containing muscarine (atropine), gyromitrin (pyridoxine) and amatoxins (silibinin [silibinin is approved e.g. in Sweden and Germany (Legalon SIL D)], penicillin); the benefit of the latter has not yet been fully established. The archaeological, ethno-historical and ethnographic evidence of the use of hallucinogenic substances in Mesoamerica is reviewed. Hallucinogenic cactus, plants and mushrooms were used to induce altered states of consciousness in healing rituals and religious ceremonies. The Maya drank balché (a mixture of honey and extracts of Lonchocarpus) in group ceremonies to achieve intoxication. Ritual enemas and other psychoactive substances were also used to induce states of trance. Olmec, Zapotec, Maya and Aztec used peyote, hallucinogenic mushrooms (teonanacatl: Psilocybe spp.) and the seeds of ololiuhqui (Turbina corymbosa), that contain mescaline, psilocybin and lysergic acid amide, respectively. The skin of the toad Bufo spp. contains bufotoxins with hallucinogenic properties, and was used during the Olmec period. Jimson weed (Datura stramonium), wild tobacco (Nicotiana rustica), water lily (Nymphaea ampla) and Salvia divinorum were used for their psychoactive effects. Mushroom stones dating from 3000 BC have been found in ritual contexts in Mesoamerica. Archaeological evidence of peyote use dates back to over 5000 years. Several chroniclers, mainly Fray Bernardino de Sahagún, described their effects in the sixteenth century. The use of psychoactive substances was common in pre-Columbian Mesoamerican societies. Today, local shamans and healers still use them in ritual ceremonies in Mesoamerica. El continente americano es rico en hongos y plantas psicoactivas, y numero-sas culturas precolombinas mesoamericanas las emplearon con fines mágicos, terapéuticos y religiosos. Se revisan las evidencias arqueológicas, etnohistóricas y etnográficas del uso de sustancias alucinógenas en Mesoamérica. Cactus, plantas y hongos alucinógenos se utilizaron para provocar estados altera-dos del nivel de conciencia en ceremonias rituales y curativas. Los mayas ingerían el balché (hidromiel y extracto de Lonchocarpus) en ceremonias grupales para lograr la embriaguez. También emplearon enemas rituales con sustancias psicoactivas para inducir estados de trance. Olmecas, zapotecas, mayas y aztecas usaron el peyote, los hongos alucinógenos (teonanacatl: Psilocybe spp.) y las semillas de ololiuhqui (Turbina corymbosa), que contienen mescalina, psilocibina y amida del ácido lisérgico, respectivamente. La piel del sapo Bufo spp. contiene bufotoxinas, con propiedades alucinógenas y fue usado desde el periodo olmeca. El toloache (Datura estramonio), el tabaco silvestre (Nicotiana rustica), el lirio de agua (Nymphaea ampla) y la hoja de la pastora (Salvia divinorum) se utilizaron por sus efectos psicotropos. Piedra fún-gicas de 3.000 ãnos de antigüedad se han encontrado en contextos rituales en Mesoamérica. Las evidencias arqueológicas del uso del peyote se remontan a más de 5.000 ãnos. Diversos cronistas, entre ellos Fray Bernardino de Sahagún, relataron sus efectos en el siglo xvi. El empleo de sustancias psicotrópicas fue muy común en las sociedades preco-lombinas mesoamericanas. En la actualidad chamanes y curanderos locales las siguen usando en ceremonias rituales. Psilocybin, the clinically most researched classic psychedelic has recently been tested for its safety and efficacy in a clinical population of treatment resistant depression. The efficacy of psilocybin in clinical depression previously demonstrated in the elecrophysiologic and neuroimaging findings as also in neuropsychological assessments is further validated by the findings of this rigorously conducted randomized trial. Mechanism of action of psilocybin and efficacy in treatment resistant depression are discussed in this paper. Ethical issues of conducting clinical trials with psychedelics are also discussed with particular emphasis on their relative safety and absence of addiction potential. Implications of these issues for conduct of larger trials for establishing risk benefit ratio in treatment resistant depression are further suggested.
  15. E. R. Badham, “The Influence of Humidity Upon Transpiration and Growth in Psilocybe Cubensis,” Mycologia, vol. 77, no. 6, pp. 932–939, Nov. 1985. doi: 10.1080/00275514.1985.12025182.
    The influence of humidity upon individual basidiocarps of of Psilocybe cubensis was studied using an environmentally controlled wind tunnel and a computer program which helped to model growth and development. Regression models were developed which were able to explain 77% of the variation in the transpiration rate and 68% of the variation in growth rate. Transpiration and growth of this mushroom were significantly correlated with the humidity of the air. The fastest growth and the lowest transpiration occurred at the highest humidities. No inhibition of growth was detected at 0 pascals VPD (100% RH). Misting accelerated growth and transpiration while light had no effect. Although humidity was a very important factor influencing transpiration and growth, the size and shape of the mushroom were also important in water relations. The final water content of basidiocarps with thin stipes or those with larger surface area-to-volume ratios was significantly lower that of thickstiped mushrooms or those with small surface area-to-volume ratios even when grown under equal humidity. Growth rates under conditions which promoted the highest levels of hydration of the basidiocarp were rapid (up to estimated 4% increase in dry weight per h).
  16. E. R. Badham, “Modeling Growth, Development, Transpiration and Translocation in the Mushroom Psilocybe Cubensis,” Bulletin of the Torrey Botanical Club, vol. 111, no. 2, pp. 159–164, 1984. doi: 10.2307/2996015.
    A method for studying the growth and development of individual basidiocarps of Psilocybe cubensis was developed Measurements of living mushrooms were put into a computer program which modeled the basidiocarp after common geometric solids Estimates of surface area and volume were determined Both surface area and volume are highly correlated with dry weight These estimates may be useful in measuring growth (non-destructively) and transpiration (water loss per surface area) Translocation of dry matter may also be studied by analysis of derived density values The same measurements can also be used to describe developmental stage (and change in stage) based on the ratio of the cap size to the stipe size
  17. E. R. Badham, “Tropisms in the Mushroom Psilocybe Cubensis,” Mycologia, vol. 74, no. 2, pp. 275–279, Mar. 1982. doi: 10.1080/00275514.1982.12021501.
    The growth of the mushroom Psilocybe cubensis was studied in a wind tunnel under controlled conditions of wind velocity, humidity, temperature, and light. The basidiocarp stipe grew into the wind up to the time of spore formation. When rotated with the long axis of the stipe perpendicular to the wind, fruitbodies grew upright. When spores began to be formed a negative geotropic curvature of the stipe occurred but no recurvation occurred in a sporeless mutant.
  18. R. G. Benedict, L. E. Brady, and V. E. Tyler, “Occurrence of Psilocin in Psilocybe Baeocystis.,” Journal of Pharmaceutical Sciences, vol. 51, no. 4, 1962. doi: 10.1002/jps.2600510428.
    A report from Univ. Wash., Seattle, on the extraction of psilocin, but not of psilocybin, from dried fruit bodies of P. baeocystis.
  19. S. R. Berlant, “The Entheomycological Origin of Egyptian Crowns and the Esoteric Underpinnings of Egyptian Religion,” Journal of Ethnopharmacology, vol. 102, no. 2, pp. 275–288, Nov. 2005. doi: 10.1016/j.jep.2005.07.028.
    In this paper, I theorize that the Egyptian White and Triple Crowns were originally primordia of the entheogenic Psilocybe (Stropharia) cubensis, which an Egyptian tale known as Cheops and the Magicians allegorically explained grew on barley, and that Osiris was the God of spiritual rebirth because he personified this and other entheogenic mushrooms. I go on to theorize that the plant known commonly as the Eye of Horus, which the Egyptians included in cakes and ales designed to spiritually rebirth the living and the dead, was an entheogenic mushroom cap entirely analogous, if not identical, to Soma. Finally, I explain why so many scholars failed to discern these identities and relationships for so long.
  20. M. W. Beug and J. Bigwood, “Psilocybin and Psilocin Levels in Twenty Species from Seven Genera of Wild Mushrooms in the Pacific Northwest, U.S.A.,” Journal of Ethnopharmacology, vol. 5, no. 3, pp. 271–285, May 1982. doi: 10.1016/0378-8741(82)90013-7.
    The analysis of twenty species from seven genera of Pacific Northwest mushrooms revealed psilocybin (and in some cases psilocin as well) in seven species from three genera. The species found to contain psilocybin (and psilocin) varied from one collection to another by more than a factor of seven in amount present. Total psilocybin and psilocin levels in species known to be in use for recreational and entheogenic purposes varied from 0.1% by dry weight up to a high of nearly 2% by dry weight.
  21. M. W. Beug and J. Bigwood, “Quantitative Analysis of Psilocybin and Psilocin and Psilocybe Baecystis (Singer and Smith) by High-Performance Liquid Chromatography and by Thin-Layer Chromatography,” Journal of Chromatography A, vol. 207, no. 3, pp. 379–385, Mar. 1981. doi: 10.1016/S0021-9673(00)88741-5.
    Rapid quantification of psilocybin and psilocin in extracts of wild mushrooms is accomplished by reversed-phase high-performance liquid chromatography with paire-ion reagents. Nine solvent systems and three solid supports are evaluated for their efficiency is separating psilocybin, psilocin and other components of crud mushroom extracts by thin-layer chromatography.
  22. J. Bigwood and M. W. Beug, “Variation of Psilocybin and Psilocin Levels with Repeated Flushes (Harvests) of Mature Sporocarps of Psilocybe Cubensis (Earle) Singer,” Journal of Ethnopharmacology, vol. 5, no. 3, pp. 287–291, May 1982. doi: 10.1016/0378-8741(82)90014-9.
    Analysis of Psilocybe cubensis (Earle) Singer grown in controlled culture showed that the level of psilocin was generally zero in the first (or sometimes even the second) fruiting of the mushroom from a given culture and that the level reached a maximum by the fourth flush. The level of psilocybin, which was nearly always at least twice the level of psilocin, showed no upward or downward trend as fruiting progressed, but was variable over a factor of four. Samples obtained from outside sources had psilocybin levels varying by over a factor of ten from one collection to the next.
  23. F. Blei, F. Baldeweg, J. Fricke, and D. Hoffmeister, “Biocatalytic Production of Psilocybin and Derivatives in Tryptophan Synthase-Enhanced Reactions,” Chemistry – A European Journal, vol. 24, no. 40, pp. 10028–10031, 2018. doi: 10.1002/chem.201801047.
    Psilocybin (4-phosphoryloxy-N,N-dimethyltryptamine) is the main alkaloid of the fungal genus Psilocybe, the so-called “magic mushrooms.” The pharmaceutical interest in this psychotropic natural product as a future medication to treat depression and anxiety is strongly re-emerging. Here, we present an enhanced enzymatic route of psilocybin production by adding TrpB, the tryptophan synthase of the mushroom Psilocybe cubensis, to the reaction. We capitalized on its substrate flexibility and show psilocybin formation from 4-hydroxyindole and l-serine, which are less cost-intensive substrates, compared to the previous method. Furthermore, we show enzymatic production of 7-phosphoryloxytryptamine (isonorbaeocystin), a non-natural congener of the Psilocybe alkaloid norbaeocystin (4-phosphoryloxytryptamine), and of serotonin (5-hydroxytryptamine) by means of the same in vitro approach.
  24. F. Blei et al., “Simultaneous Production of Psilocybin and a Cocktail of β-Carboline Monoamine Oxidase Inhibitors in ‘Magic’ Mushrooms,” Chemistry – A European Journal, vol. 26, no. 3, pp. 729–734, 2020. doi: 10.1002/chem.201904363.
    The psychotropic effects of Psilocybe “magic” mushrooms are caused by the l-tryptophan-derived alkaloid psilocybin. Despite their significance, the secondary metabolome of these fungi is poorly understood in general. Our analysis of four Psilocybe species identified harmane, harmine, and a range of other l-tryptophan-derived β-carbolines as their natural products, which was confirmed by 1D and 2D NMR spectroscopy. Stable-isotope labeling with 13C11-l-tryptophan verified the β-carbolines as biosynthetic products of these fungi. In addition, MALDI-MS imaging showed that β-carbolines accumulate toward the hyphal apices. As potent inhibitors of monoamine oxidases, β-carbolines are neuroactive compounds and interfere with psilocybin degradation. Therefore, our findings represent an unprecedented scenario of natural product pathways that diverge from the same building block and produce dissimilar compounds, yet contribute directly or indirectly to the same pharmacological effects.
  25. M. J. Bogusz, R.-D. Maier, A. T. Schäfer, and M. Erkens, “Honey with Psilocybe Mushrooms: A Revival of a Very Old Preparation on the Drug Market?,” International Journal of Legal Medicine, vol. 111, no. 3, pp. 147–150, Apr. 1998. doi: 10.1007/s004140050135.
    In 1996 samples of suspicious honey preparations were confiscated at the Dutch-German border. The labels on the 50 ml jars indicated that the honey contained Stropharia cubensis (better known as Psilocybe cubensis). The jars were filled with honey with a ca. 1 cm layer of fine particles on the top. The particles were collected and subjected to microscopic and chemical analysis. By microscopy mushroom tissue (plectenchym) and spores typical for the genus Psilocybe were identified in all samples. The HPLC analysis with atmospheric pressure mass spectrometry and diode array detection revealed psilocine but psilocybine was not found. The quantitative analysis was very difficult due to the matrix problems. A search showed that the honey with Psilocybe can be purchased in Dutch coffee shops without any limitations although psilocine and psilocybine belong to listed substances according to Dutch law.
  26. S. Borner and R. Brenneisen, “Determination of Tryptamine Derivatives in Hallucinogenic Mushrooms Using High-Performance Liquid Chromatography with Photodiode Array Detection,” Journal of Chromatography A, vol. 408, pp. 402–408, Jan. 1987. doi: 10.1016/S0021-9673(01)81831-8.
  27. J. Borovička, M. Oborník, J. Stříbrný, M. E. Noordeloos, L. A. P. Sánchez, and M. Gryndler, “Phylogenetic and Chemical Studies in the Potential Psychotropic Species Complex of Psilocybe Atrobrunnea with Taxonomic and Nomenclatural Notes,” Persoonia - Molecular Phylogeny and Evolution of Fungi, vol. 34, no. 1, pp. 1–9, Jun. 2015. doi: 10.3767/003158515X685283.
    Five Psilocybe species with unresolved systematic position (P. atrobrunnea, P. laetissima, P. medul- losa, P. pelliculosa, and P. silvatica) were investigated using four molecular markers (EF1-α, ITS, LSU, and IGS). Phylogenetic analysis revealed that with the exception of P. laetissima, which is now rightfully classified in the genus Leratiomyces, all investigated species belong to Psilocybe sect. Psilocybe. For the first time, psychotropic compounds psilocin and psilocybin were detected in P. medullosa using gas chromatography-mass spectrometry. On the contrary, neither psilocin, nor psilocybin was detected in P. atrobrunnea and negative results were also obtained from mycelia grown in vitro on tryptamine/tryptophan-amended media. These results strongly suggest that biosynthesis of these alkaloids was lost in P. atrobrunnea. With the exception of minor differences detected in EF1-α marker, all sequences of American and European collections of P. atrobrunnea were identical. On the other hand, a thorough nomenclatural study revealed that the name P. atrobrunnea must be considered dubious; the oldest available candidate name, P. fuscofulva, was therefore adopted. The molecular data suggests that morphologically identical American P. silvatica and European P. medullosa likely represent distinct species; epitypes of both taxa were therefore designated.
  28. A. J. Bradshaw et al., “DNA Authentication and Chemical Analysis of Psilocybe Mushrooms Reveal Widespread Misdeterminations in Fungaria and Inconsistencies in Metabolites,” Applied and Environmental Microbiology, vol. INVALID_SCITE_VALUE, no. INVALID_SCITE_VALUE, pp. e01498–22, Nov. 2022. doi: 10.1128/aem.01498-22.
    The mushroom genus Psilocybe is best known as the core group of psychoactive mushrooms, yet basic information on their diversity, taxonomy, chemistry, and general biology is still largely lacking. In this study, we reexamined 94 Psilocybe fungarium specimens, representing 18 species, by DNA barcoding, evaluated the stability of psilocybin, psilocin, and their related tryptamine alkaloids in 25 specimens across the most commonly vouchered species (Psilocybe cubensis, Psilocybe cyanescens, and Psilocybe semilanceata), and explored the metabolome of cultivated P. cubensis. Our data show that, apart from a few well-known species, the taxonomic accuracy of specimen determinations is largely unreliable, even at the genus level. A substantial quantity of poor-quality and mislabeled sequence data in public repositories, as well as a paucity of sequences derived from types, further exacerbates the problem. Our data also support taxon- and time-dependent decay of psilocybin and psilocin, with some specimens having no detectable quantities of them. We also show that the P. cubensis metabolome possibly contains thousands of uncharacterized compounds, at least some of which may be bioactive. Taken together, our study undermines commonly held assumptions about the accuracy of names and presence of controlled substances in fungarium specimens identified as Psilocybe spp. and reveals that our understanding of the chemical diversity of these mushrooms is largely incomplete. These results have broader implications for regulatory policies pertaining to the storage and sharing of fungarium specimens as well as the use of psychoactive mushrooms for recreation and therapy. IMPORTANCE The therapeutic use of psilocybin, the active ingredient in “magic mushrooms,” is revolutionizing mental health care for a number of conditions, including depression, posttraumatic stress disorder (PTSD), and end-of-life care. This has spotlighted the current state of knowledge of psilocybin, including the organisms that endogenously produce it. However, because of international regulation of psilocybin as a controlled substance (often included on the same list as cocaine and heroin), basic research has lagged far behind. Our study highlights how the poor state of knowledge of even the most fundamental scientific information can impact the use of psilocybin-containing mushrooms for recreational or therapeutic applications and undermines critical assumptions that underpin their regulation by legal authorities. Our study shows that currently available chemical studies are mainly inaccurate, irreproducible, and inconsistent, that there exists a high rate of misidentification in museum collections and public databases rendering even names unreliable, and that the concentration of psilocybin and its tryptamine derivatives in three of the most commonly collected Psilocybe species (P. cubensis, P. cyanescens, and P. semilanceata) is highly variable and unstable in museum specimens spanning multiple decades, and our study generates the first-ever insight into the highly complex and largely uncharacterized metabolomic profile for the most commonly cultivated magic mushroom, P. cubensis.
  29. R. T. Brown et al., “Pharmacokinetics of Escalating Doses of Oral Psilocybin in Healthy Adults,” Clinical Pharmacokinetics, vol. 56, no. 12, pp. 1543–1554, Dec. 2017. doi: 10.1007/s40262-017-0540-6.
    Introduction Psilocybin is a psychedelic tryptamine that has shown promise in recent clinical trials for the treatment of depression and substance use disorders. This open-label study of the pharmacokinetics of psilocybin was performed to describe the pharmacokinetics and safety profile of psilocybin in sequential, escalating oral doses of 0.3, 0.45, and 0.6 mg/kg in 12 healthy adults.
  30. M. D. Buitrago Fajardo, “Estudio químico preliminar y evaluación citotóxica de Psilocybe cubensis var. amazonas,” PhD thesis, Química Farmacéutica, 2022. https://repositorio.unbosque.edu.co/handle/20.500.12495/9293.
    Los usos etnobotánicos reportados para los hongos del género Psilocybe en comunidades de Mesoamérica, Europa y Estados Unidos, se centran en su consumo como producto recreativo. No obstante, en la medicina tradicional oriental se han utilizado en distintas aplicaciones terapéuticas para el tratamiento de dolencias y otros trastornos. En consecuencia, se han identificado más de 180 especies de este género, dentro de las que se destaca la especie Psilocybe cubensis, la cual ha generado un creciente interés en la comunidad científica por su composición química y actividad biológica, principalmente, antimicrobiana y citotóxica. Las setas del género Psilocybe, también conocidas como “setas mágicas”, son especies fúngicas que contienen en su mayoría psilocibina y psilocina, compuestos con propiedades psicodélicas. Por tanto, el presente trabajo de investigación se centró en la determinación del tipo de metabolitos secundarios presentes en el extracto de los cuerpos fructíferos de dos muestras comerciales de Psilocybe cubensis var. amazonas y en la evaluación de su toxicidad y actividad citotóxica, mediante ensayos químicos preliminares y ensayos biológicos in vitro. De este modo, a través de la determinación fitoquímica preliminar se determinó la presencia de alcaloides, aglicones cardiotónicos, esteroles y taninos, así mismo se realizó un estudio de cromatografía en capa delgada con el fin de corroborar los resultados obtenidos en las pruebas de análisis preliminar en tubo de ensayo. Por otra parte, los extractos etanólicos mostraron dosis letales medias (DL50) de 100,00 y 120,23 μg/mL sobre larvas de Artemia salina. Los ensayos de actividad citotóxica realizados con los extractos AM-1 y AM-2 sobre las líneas celulares MCF-7 y MDA-MB-231 presentaron alta actividad citotóxica con IC50 de 85,51; 83,82 μg/mL y 86,50; 78,16 μg/mL respectivamente, con lo que se demuestra que el perfil químico de los hongos psilocibos como Psilocybe cubensis son una alternativa terapéutica promisoria no solo con acción analgésica sino conjuntamente con propiedades anticancerígenas y antioxidantes.
  31. J. F. Casale, “An Aqueous-Organic Extraction Methods for the Isolation and Identification of Psilocin from Hallucinogenic Mushrooms,” Journal of Forensic Sciences, vol. 30, no. 1, pp. 247–250, Jan. 1985.
    A simple aqueous extraction method for the isolation and identification of psilocin from Psilocybe cubensis mushrooms is reported. This method employs a dephosphorylation of the phosphate ester to psilocin, which fascilitates a greater product yield and simplifies identification. Psilocin extracted by this method is sufficiently concentrated and free of cocontaminants to allow identification by infrared spectrscopy and gas chromatography/mass spectrometry.
  32. P. Catalfomo and V. E. \relax J. Tyler, “The Production of Psilocybin in Submerged Culture by Psilocybe Cubensis,” Lloydia, vol. 27, pp. 53–63, 1964.
  33. A. L. Christiansen and K. E. Rasmussen, “Analysis of Indole Alkaloids in Norwegian Psilocybe Semilanceata Using High-Performance Liquid Chromatography and Mass Spectrometry,” Journal of Chromatography A, vol. 244, no. 2, pp. 357–364, Jul. 1982. doi: 10.1016/S0021-9673(00)85700-3.
    High-performance liquid chromatography has been used as separation method before spectroscopic investigations of isolated indole alkaloids in Psylocybe semilanceata mushrooms. By using a volatile buffer the mobile phase was easily removed by freeze-drying, and the residue introduced directly into a mass spectrometer. For final identification of baeocystin, it was necessary to investigate the isolated compound by mass spectrometry, UV spectroscopy and fluorescence detection. The mushrooms were found to contain up to 0.34% baeocystin. The amount of baeocystin was considerably higher in the cap than in the stipe. The content of psilocybin, however, was of the same order of magnitude in both parts of the mushroom. Rather large amounts of psilocybin were also found in dried, old mushrooms, which showed that the mushroom can still be potent after long storage.
  34. A. L. Christiansen, K. E. Rasmussen, and K. Høiland, “The Content of Psilocybin in Norwegian Psilocybe Semilanceata,” Planta Medica, vol. 42, no. 7, pp. 229–235, Jul. 1981. doi: 10.1055/s-2007-971632.
    Thieme E-Books & E-Journals
  35. A. L. Christiansen, K. E. Rasmussen, and F. TØNNESEN, “Determination of Psilocybin in Psilocybe Semilanceata Using High-Performance Liquid Chromatography on a Silica Column,” Journal of Chromatography, vol. 210, pp. 163–167, 1981.
    During the last few years an increasing number of young Norwegians have used Psilocybe semilanceczta (Fr. e-r Seer.) Kummer as a narcotic. The first report on misuse of this mushroom was made in autumn 1977, and knowledge of its hallucinogenic properties has gradually become public through reports in the popular media. The ingestion of P. semilanceata has in some cases resulted in the need for treatment in hospita11,2. Many species of the genus Psilocybe are found in Norway3, but only P. semiIanceata is regarded as hallucinogenic. It occurs on grassy sites in most parts of the country from the middle of August to the middle of October. P. semilanceata is known to contain indole alkaloids, of which psilocybin is considered to be the main constituent. In order to carry out detailed studies of the potency of Norwegian P. semilanceata a quantitative method was required for the assay of psilocybin. Several methods have previously been used for analysis of the hallucinogenic components of Psiiocybe mushrooms. Both paper chromatography4vs and thin-layer chromatography 6*7 have been employed in conjunction with calorimetric reagents as well as with UV spectroscopy. Gas chromatography and gas chromatography-r&&s spcctrometry* have been applied for the analysis of psilocin and psilocybin. Recently two high-performance liquid chromatographic (HPLC) methods have been published ‘*lo White9 separated the three compounds psilocin, psilocybin . and baeocystin on a silica column, and Perkal et al. lo described the quantitation of psilocin and psilocybin by ion-exchange chromatography. We have developed a KPLC method based on a silica column which provides a simple, rapid and accurate quantitation of the psilocybin content of Norwegian P. semilanceata.
  36. V. G. Cortez and G. Coelho, “The Stropharioideae (Strophariaceae, Agaricales) from Santa Maria, Rio Grande Do Sul, Brazil.”
    Collections of Hypholoma subviride, Melanotus proteus, Psilocybe caeruleoannulata, P. coprophila, P. cubensis, P. moellerii, P. pegleriana, Stropharia coronilla, S. rugosoannulata, and S. semiglobata, all placed in subfamily Stropharioideae (Strophariaceae, Agaricales) are discussed and illustrated from southern Brazil. Melanotus proteus, P. moellerii and P. pegleriana are recorded for the first time from the country, and S. rugosoannulata and H. subviride collections represent new records for the state of Rio Grande do Sul.
  37. A. F. Cowan and K. M. Elkins, “Detection and Identification of Psilocybe Cubensis DNA Using a Real-Time Polymerase Chain Reaction High Resolution Melt Assay,” Journal of Forensic Sciences, vol. 63, no. 5, pp. 1500–1505, 2018. doi: 10.1111/1556-4029.13714.
    Psilocybe cubensis, or “magic mushroom,” is the most common species of fungus with psychedelic characteristics. Two primer sets were designed to target Psilocybe DNA using web-based software and NBCI gene sequences. DNA was extracted from eighteen samples, including twelve mushroom species, using the Qiagen DNeasy® Plant Mini Kit. The DNA was amplified by the polymerase chain reaction (PCR) using the primers and a master mix containing either a SYBR® Green I, Radiant™ Green, or LCGreen Plus® intercalating dye; amplicon size was determined using agarose gel electrophoresis. The PCR assays were tested for amplifiability, specificity, reproducibility, robustness, sensitivity, and multiplexing with primers that target marijuana. The observed high resolution melt (HRM) temperatures for primer sets 1 and 7 were 78.85 ± 0.31°C and 73.22 ± 0.61°C, respectively, using SYBR® Green I dye and 81.67 ± 0.06°C and 76.04 ± 0.11°C, respectively, using Radiant™ Green dye.
  38. E. Crocker and N. W. Gauthier, “Don’t Eat Those Wild Mushrooms,” p. 5.
  39. N. D. Davis, G. Morgan-Jones, R. E. Wagener, A. J. Latham, and U. L. Diener, “Psilocybe Cubensis, an Hallucinogenic Mushroom Occurring in Alabama,” Highlights Agric Res, 1974. https://agris.fao.org/agris-search/search.do?recordID=US201303159209.
    AGRICULTURAL SCIENCE AND TECHNOLOGY INFORMATION
  40. R. Demmler, J. Fricke, S. Dörner, M. Gressler, and D. Hoffmeister, “S-Adenosyl-l-Methionine Salvage Impacts Psilocybin Formation in ‘Magic’ Mushrooms,” ChemBioChem, vol. 21, no. 9, pp. 1364–1371, 2020. doi: 10.1002/cbic.201900649.
    Psychotropic Psilocybe mushrooms biosynthesize their principal natural product psilocybin in five steps, among them a phosphotransfer and two methyltransfer reactions, which consume one equivalent of 5′-adenosine triphosphate (ATP) and two equivalents of S-adenosyl-l-methionine (SAM). This short but co-substrate-intensive pathway requires nucleoside cofactor salvage to maintain high psilocybin production rates. We characterized the adenosine kinase (AdoK) and S-adenosyl-l-homocysteine (SAH) hydrolase (SahH) of Psilocybe cubensis. Both enzymes are directly or indirectly involved in regenerating SAM. qRT-PCR expression analysis revealed an induced expression of the genes in the fungal primordia and carpophores. A one-pot in vitro reaction with the N-methyltransferase PsiM of the psilocybin pathway demonstrates a concerted action with SahH to facilitate biosynthesis by removal of accumulating SAH.
  41. D. Dhanasekaran et al., “Taxonomic Identification and Bioactive Compounds Characterization of Psilocybe Cubensis DPT1 to Probe Its Antibacterial and Mosquito Larvicidal Competency,” Microbial Pathogenesis, vol. 143, p. 104138, Jun. 2020. doi: 10.1016/j.micpath.2020.104138.
    Mushrooms have an important role in sustainability since they have long been used as valuable food source and traditional medicine around the world. Regrettably, they are among the most rigorously affected populations, along with several plants and animals, due to the destructive activities of mankind. Thus the authentication and conservation of mushroom species are constantly needed to exploit the remarkable potential in them. In this perspective, an attempt has been made to identify and assess the biological attributes of psychedelic mushrooms collected from Kodaikanal, Tamil Nadu, India. The macromorphological features of the psychedelic mushroom DPT1 helped its presumptive identification and the molecular characters depicted by DNA marker revealed its close relationship with the genus Psilocybe. Accordingly, the psychedelic mushroom was identified as Psilocybe cubensis DPT1 and its crude ethyl acetate extract on analysis revealed the occurrence of phytoconstituents like alkaloids, flavonoids, tannins, saponins and carbohydrates. Moreover, it exhibited 80% larvicidal activity against Culex quinquefasciatus mosquito at 800 ppm concentration and an array of antibacterial effects with utmost susceptibility of Proteus vulgaris, and the identification of bioactive compounds by different analytical techniques substantiate that the bioactivities might be due to the presence of phytochemicals. The results of the study indicated that the extract of P. cubensis DPT1 having notable antibacterial and mosquito larvicidal efficacies which could be probed further for the isolation of medicinally important as well as bio-control compounds.
  42. R. J. Dinis-Oliveira, “Metabolism of Psilocybin and Psilocin: Clinical and Forensic Toxicological Relevance,” Drug Metabolism Reviews, vol. 49, no. 1, pp. 84–91, Jan. 2017. doi: 10.1080/03602532.2016.1278228.
    Psilocybin and psilocin are controlled substances in many countries. These are the two main hallucinogenic compounds of the “magic mushrooms” and both act as agonists or partial agonists at 5-hydroxytryptamine (5-HT)2A subtype receptors. During the last few years, psilocybin and psilocin have gained therapeutic relevance but considerable physiological variability between individuals that can influence dose-response and toxicological profile has been reported. This review aims to discuss metabolism of psilocybin and psilocin, by presenting all major and minor psychoactive metabolites. Psilocybin is primarily a pro-drug that is dephosphorylated by alkaline phosphatase to active metabolite psilocin. This last is then further metabolized, psilocin-O-glucuronide being the main urinary metabolite with clinical and forensic relevance in diagnosis.
  43. S. Dörner et al., “Genetic Survey of Psilocybe Natural Products,” ChemBioChem, vol. 23, no. 14, p. e202200249, 2022. doi: 10.1002/cbic.202200249.
    Psilocybe magic mushrooms are best known for their main natural product, psilocybin, and its dephosphorylated congener, the psychedelic metabolite psilocin. Beyond tryptamines, the secondary metabolome of these fungi is poorly understood. The genomes of five species (P. azurescens, P. cubensis, P. cyanescens, P. mexicana, and P. serbica) were browsed to understand more profoundly common and species-specific metabolic capacities. The genomic analyses revealed a much greater and yet unexplored metabolic diversity than evident from parallel chemical analyses. P. cyanescens and P. mexicana were identified as aeruginascin producers. Lumichrome and verpacamide A were also detected as Psilocybe metabolites. The observations concerning the potential secondary metabolome of this fungal genus support pharmacological and toxicological efforts to find a rational basis for yet elusive phenomena, such as paralytic effects, attributed to consumption of some magic mushrooms.
  44. A. Dorr, “MYCOREMEDIATION HANDBOOK,” p. 232.
  45. J. Fricke, F. Blei, and D. Hoffmeister, “Enzymatic Synthesis of Psilocybin,” Angewandte Chemie International Edition, vol. 56, no. 40, pp. 12352–12355, 2017. doi: 10.1002/anie.201705489.
    Psilocybin is the psychotropic tryptamine-derived natural product of Psilocybe carpophores, the so-called “magic mushrooms”. Although its structure has been known for 60 years, the enzymatic basis of its biosynthesis has remained obscure. We characterized four psilocybin biosynthesis enzymes, namely i) PsiD, which represents a new class of fungal l-tryptophan decarboxylases, ii) PsiK, which catalyzes the phosphotransfer step, iii) the methyltransferase PsiM, catalyzing iterative N-methyl transfer as the terminal biosynthetic step, and iv) PsiH, a monooxygenase. In a combined PsiD/PsiK/PsiM reaction, psilocybin was synthesized enzymatically in a step-economic route from 4-hydroxy-l-tryptophan. Given the renewed pharmaceutical interest in psilocybin, our results may lay the foundation for its biotechnological production.
  46. J. Fricke et al., “Enzymatic Route toward 6-Methylated Baeocystin and Psilocybin,” ChemBioChem, vol. 20, no. 22, pp. 2824–2829, 2019. doi: 10.1002/cbic.201900358.
    Psilocybin and its direct precursor baeocystin are indole alkaloids of psychotropic Psilocybe mushrooms. The pharmaceutical interest in psilocybin as a treatment option against depression and anxiety is currently being investigated in advanced clinical trials. Here, we report a biocatalytic route to synthesize 6-methylated psilocybin and baeocystin from 4-hydroxy-6-methyl-l-tryptophan, which was decarboxylated and phosphorylated by the Psilocybe cubensis biosynthesis enzymes PsiD and PsiK. N-Methylation was catalyzed by PsiM. We further present an in silico structural model of PsiM that revealed a well-conserved SAM-binding core along with peripheral nonconserved elements that likely govern substrate preferences.
  47. J. Fricke, F. Blei, and D. Hoffmeister, “Enzymatische Synthese von Psilocybin,” Angewandte Chemie, vol. 129, no. 40, pp. 12524–12527, 2017. doi: 10.1002/ange.201705489.
    Psilocybin ist der psychotrope, vom Tryptamin abgeleitete Naturstoff der Psilocybe-Fruchtkörper, der so genannten “Zauberpilze”. Obwohl dessen Struktur seit 60 Jahren bekannt ist, blieb die enzymatische Grundlage ihrer Biosynthese ungeklärt. Wir charakterisierten vier Psilocybin-Biosyntheseenzyme, nämlich i) PsiD, welches zu einer neuen Klasse pilzlicher l-Tryptophan-Decarboxylasen gehört, ii) PsiK, welches den Phosphotransfer-Schritt katalysiert, iii) die Methyltransferase PsiM, welche einen wiederholten N-Methyltransfer als abschließenden Biosyntheseschritt katalysiert, sowie iv) PsiH, eine Monooxygenase. In einer kombinierten Reaktion mit PsiD, PsiK und PsiM wurde Psilocybin vom 4-Hydroxy-l-tryptophan ausgehend nach einer schrittökonomischen Route synthetisiert. Angesichts des wieder aufkommenden pharmazeutischen Interesses an Psilocybin könnten unsere Ergebnisse die Grundlage für dessen biotechnologische Produktion schaffen.
  48. J. Fricke, C. Lenz, J. Wick, F. Blei, and D. Hoffmeister, “Production Options for Psilocybin: Making of the Magic,” Chemistry – A European Journal, vol. 25, no. 4, pp. 897–903, 2019. doi: 10.1002/chem.201802758.
    The fungal genus Psilocybe and other genera comprise numerous mushroom species that biosynthesize psilocybin (4-phosphoryloxy-N,N-dimethyltryptamine). It represents the prodrug to its dephosphorylated psychotropic analogue, psilocin. The colloquial term “magic mushrooms” for these fungi alludes to their hallucinogenic effects and to their use as recreational drugs. However, clinical trials have recognized psilocybin as a valuable candidate to be developed into a medication against depression and anxiety. We here highlight its recently elucidated biosynthesis, the concurrently developed concept of enzymatic in vitro and heterologous in vivo production, along with previous synthetic routes. The prospect of psilocybin as a promising therapeutic may entail an increased demand, which can be met by biotechnological production. Therefore, we also briefly touch on psilocybin’s therapeutic relevance and pharmacology.
  49. J. Fricke et al., “Scalable Hybrid Synthetic/Biocatalytic Route to Psilocybin,” Chemistry – A European Journal, vol. 26, no. 37, pp. 8281–8285, 2020. doi: 10.1002/chem.202000134.
    Psilocybin, the principal indole alkaloid of Psilocybe mushrooms, is currently undergoing clinical trials as a medication against treatment-resistant depression and major depressive disorder. The psilocybin supply for pharmaceutical purposes is met by synthetic chemistry. We replaced the problematic phosphorylation step during synthesis with the mushroom kinase PsiK. This enzyme was biochemically characterized and used to produce one gram of psilocybin from psilocin within 20 minutes. We also describe a pilot-scale protocol for recombinant PsiK that yielded 150 mg enzyme in active and soluble form. Our work consolidates the simplicity of tryptamine chemistry with the specificity and selectivity of enzymatic catalysis and helps provide access to an important drug at potentially reasonable cost.
  50. T. Froese, G. Guzmán, and L. Guzmán-Dávalos, “On the Origin of the Genus Psilocybe and Its Potential Ritual Use in Ancient Africa and Europe1,” Economic Botany, vol. 70, no. 2, pp. 103–114, Jun. 2016. doi: 10.1007/s12231-016-9342-2.
    On the Origin of the Genus Psilocybe and Its Potential Ritual Use in Ancient Africa and Europe. The role of altered states of consciousness in the production of geometric and figurative art by prehistoric cultures in Africa and Europe has been hotly debated. Helvenston and Bahn have tried to refute the most famous hypothesis, Lewis-Williams’ neuropsychological model, by claiming that appropriate visual hallucinations required the ingestion of LSD, psilocybin, or mescaline, while arguing that none of these compounds were available to the cultures in question. We present here mycological arguments that tell another story. A prehistoric worldwide distribution of the mushroom genus Psilocybe, and therefore of psilocybin, is supported by the existence of endemic species in America, Africa, and Europe, the disjunct distribution of sister species, and the possibility of long-distance spore dispersal. It is more difficult to point to instances of actual prehistoric ritual use in Africa and Europe, but there are a growing number of suggestive findings.
  51. J. Gabriel, K. Švec, D. Kolihová, P. Tlustoš, and J. Száková, “Translocation of Mercury from Substrate to Fruit Bodies of Panellus Stipticus, Psilocybe Cubensis, Schizophyllum Commune and Stropharia Rugosoannulata on Oat Flakes,” Ecotoxicology and Environmental Safety, vol. 125, pp. 184–189, Mar. 2016. doi: 10.1016/j.ecoenv.2015.12.009.
    The cultivation and fructification of 15 saprotrophic and wood-rotting fungal strains were tested on three various semi-natural medium. The formation of fruit bodies was observed for Panellus stipticus, Psilocybe cubensis, Schizophyllum commune and Stropharia rugosoannulata in the frame of 1–2 months. Mercury translocation from the substrate to the fruit bodies was then followed in oat flakes medium. Translocation was followed for treatments of 0, 1.25, 2.5, 5, 10 and 20ppm Hg in the substrate. All four fungi formed fruit bodies in almost all replicates. The fruit body yield varied from 0.5 to 15.3g dry weight. The highest bioconcentration factor (BCF) of 2.99 was found for P. cubensis at 1.25ppm Hg. The BCF decreased with increasing Hg concentration in the substrate: 2.49, 0, 2.38, 1.71 and 1.82 for P. stipticus; 3.00, 2.78, 2.48, 1.81 and 2.15 for P. cubensis; 2.47, 1.81, 1.78, 1.07 and 0.96 for S. commune; and 1.96, 1.84, 1.21, 1.71 and 0.96 for S. rugosoannulata. The Hg contents in the fruit bodies reflected the Hg contents in the substrate; the highest contents in the fruit bodies were found in P. cubensis (43.08±7.36ppm Hg) and P. stipticus (36.42±3.39ppm).
  52. V. Gambaro et al., “DNA-Based Taxonomic Identification of Basidiospores in Hallucinogenic Mushrooms Cultivated in ‘Grow-Kits’ Seized by the Police: LC-UV Quali-Quantitative Determination of Psilocybin and Psilocin,” Journal of Pharmaceutical and Biomedical Analysis, vol. 125, pp. 427–432, Jun. 2016. doi: 10.1016/j.jpba.2016.03.043.
    The taxonomic identification of the biological material contained in the hallucinogenic mushrooms culture media, was carried out using a DNA-based approach, thus highlighting the usefulness of this approach in the forensic identification of illegal samples also when they are present as basidiospores mixed in culture media and spore-bearing fruiting body are not present. This approach is very useful as it allows the unequivocal identification of potentially illicit material before the cultivation and it enables to stop the material to the Customs and to destroy it due to its dangerousness without cultivating the “grow-kits” and without instructing a criminal case. In fact, even if psilocin and psilocybin and the whole mushrooms are illegal in many countries, there is no specific indication in the law about the so called “grow-kits”, containing the spores. To confirm the data obtained by the taxonomic identification, a simple, reliable, efficient LC-UV method, using tryptamine as internal standard, suitable for the forensic quali-quantitative determination of psilocin and psilocybin in hallucinogenic mushroom was optimized, validated and applied to the mushrooms grown after the cultivation of the grow-kits seized by the judicial authority, with the authorization of the Ministry of Health. A cation exchange column was used in a gradient elution mode (Phase A: 50mMK2HPO4; 100mM NaCl pH=3 Phase B: methanol). The developed method was linear over the calibration range with a R2>0.9992 for both the analytes. The detection and quantification limits were respectively 0.01 and 0.1μg/mL for psilocybin and 0.05μg/mL and 0.1μg/mL for psilocin and the intra- and inter-day precision was satisfactory (coefficients of variation <2.0% for both the analytes). The content of psilocybin in the mushrooms grown from the seized “grow-kits” ranged from 1.02 to 7.60mg/g of dry vegetable material, while the content of psilocin from 0.415 to 8.36mg/g.
  53. A. Gandia and A. Adamatzky, “Electrical Spiking of Psilocybin Fungi,” Communicative & Integrative Biology, vol. 15, no. 1, pp. 226–231, Dec. 2022. doi: 10.1080/19420889.2022.2136118.
    Psilocybin fungi, aka “magic” mushrooms, are well known for inducing colorful and visionary states of mind. Such psychoactive properties and the ease of cultivating their basidiocarps within low-tech setups make psilocybin fungi promising pharmacological tools for mental health applications. Understanding of the intrinsic electrical patterns occurring during the mycelial growth can be utilized for better monitoring the physiological states and needs of these species. In this study we aimed to shed light on this matter by characterizing the extra-cellular electrical potential of two popular species of psilocybin fungi: Psilocybe tampanensis and P. cubensis. As in previous experiments with other common edible mushrooms, the undisturbed fungi have shown to generate electric potential spikes and trains of spiking activity. This short analysis provides a proof of intrinsic electrical communication in psilocybin fungi, and further establishes these fungi as a valuable tool for studying fungal electro-physiology.
  54. J. Gartz, “Analysis of Aeruginascin in Fruit Bodies of the Mushroom Inocybe Aeruginascens,” International Journal of Crude Drug Research, vol. 27, no. 3, pp. 141–144, Jan. 1989. doi: 10.3109/13880208909053954.
    Analysis of ten fruit bodies of Inocybe aeruginascens Babos revealed a content of the indole derivative aeruginascin which was in the same order of magnitude as the amounts of psilocybin or baeocystin. There was a correlation between the content of psilocybin, baeocystin and aeruginascin. Aeruginascin seems to modify the pharmacological action of psilocybin to give an always euphoric mood during ingestion of the mushrooms.
  55. J. Gartz and G. K. Moller, “Analysis and Cultivation of Fruit Bodies and Mycelia of Psilocybe Bohemica,” Biochemie und Physiologie der Pflanzen, vol. 184, no. 3, pp. 337–341, Jan. 1989. doi: 10.1016/S0015-3796(89)80023-X.
    The analysis of fruit bodies of Psilocybe bohemica from a single location revealed psilocybin, baeocystin and in some cases psilocin. Psilocybin levels varied from 0.11 % up to 1.34 % by dry weight. The content of baeocystin and psilocybin was highest in the caps of the mushrooms. Psilocybin was also found to be contained in the cultured mycelia of this species. No other alkaloids were detected in the mycelial extracts. A rhizomorphic to closely linear growth of the blueing mycelia was observed on soaked unsterilized cardboard. For the first time, the fruiting of Psilocybe bohemica could be demonstrated.
  56. J. Gartz, “Biotransformation of Tryptamine in Fruiting Mycelia of Psilocybe Cubensis,” Planta Medica, vol. 55, no. 3, pp. 249–250, Jun. 1989. doi: 10.1055/s-2006-961995.
    Thieme E-Books & E-Journals
  57. J. Gartz, “Biotransformation of Tryptamine Derivatives in Mycelial Cultures of Psilocybe,” Journal of Basic Microbiology, vol. 29, no. 6, pp. 347–352, 1989. doi: 10.1002/jobm.3620290608.
    Mycelial cultures of Psilocybe cubensis capable of forming psilocybin and psilocin de novo display a high capacity for hydroxylation of tryptamine derivatives at the 4-position. A specific biotransformation of added synthetic N,N-diethyl-tryptamine was found. Thus high amounts of 4-hydroxy-N,N-diethyltryptamine (up to 3.3%) and a minor quantity of 4-phosphoryloxy-N,N-diethyltryptamine (0.01–0.8%) were isolated from fruiting bodies of Psilocybe cubensis in corresponding experiments. This is the first example of a directed biosynthesis of tryptamine substances by fungi. An effective biotransformation of N-methyltryptamine was also demonstrated with surface cultures of Psilocybe semilanceata. Baeocystin, a possible natural precursor of psilocybin, was detected and quantified in the biomasses. No alkaloids could be found in the culture medium.
  58. J. Gartz and G. Wiedemann, “Discovery of a New Caerulescent Psilocybe Mushroom in Germany: Psilocybe Germanica Sp.Nov.,” Drug Testing and Analysis, vol. 7, no. 9, pp. 853–857, Sep. 2015. doi: 10.1002/dta.1795.
  59. J. Gartz, J. W. Allen, and M. D. Merlin, “Ethnomycology, Biochemistry, and Cultivation of Psilocybe Samuiensis Guzmán, Bandala and Allen, a New Psychoactive Fungus from Koh Samui, Thailand,” Journal of Ethnopharmacology, vol. 43, no. 2, pp. 73–80, Jul. 1994. doi: 10.1016/0378-8741(94)90006-X.
    Several specimens of Psilocybe and Copelandia species in Koh Samui, Thailand were recently collected for herbarium deposit and scientific study. This paper presents an ethnomycological and biochemical study of one of the species; P. samuiensis Guzmán, Bandala and Allen, a new psychoactive gill fungus reported from Thailand. Mycelium for the cultivation of P. samuiensis was obtained on 6% malt agar from the spores of a dried specimen. The growth of P. samuiensis was similar to that of P. tampanensis Guzmán and Pollock, but more rapid than the mycelium of P. semilanceata (Fr.: Sacc.) Kumm. Laboratory analyses indicates that the alkaloid content in cultured fruit bodies of P. samuiensis is of the same order of magnitude as that found in naturally occurring mushrooms of this species. HPLC analyses of both naturally occurring and in vitro cultivated fruit bodies of P. samuiensis revealed high concentrations of psilocybin and psilocin. Small amounts of baeocystin were also detected. Psilocybin levels varied from 0.23% up to 0.90%. The psilocybin content was highest in the caps. Psilocybin was also found in the cultured non-bluing mycelia of P. samuiensis and varied from 0.24% to 0.32% dry weight. The relative alkaloidal content of psilocybin, psilocin, and baeocystin found in P. samuiensis was similar to that measured in many other psychoactive fungi species, but completely different from that found in P. semilanceata.
  60. J. Gartz, “Extraction and Analysis of Indole Derivatives from Fungal Biomass,” Journal of Basic Microbiology, vol. 34, no. 1, pp. 17–22, 1994. doi: 10.1002/jobm.3620340104.
    The occurrence and extraction of indole derivatives in six species from four genera of higher fungi were investigated. By using pure methanol for extraction of the mushrooms analysis revealed the highest concentrations of psilocybin and baeocystin. The psilocin content of the species was higher by using aqueous solutions of alcohols than with methanol alone but was an artificial phenomenon caused by enzymatic destruction of psilocybin. The extraction with dilute acetic acid yielded better results than with the water containing alcohols. The simple one-step procedure with methanol for the quantitative extraction is still the safest method to obtain the genuine alkaloids from fungal biomass.
  61. J. Gartz, G. Adam, and H. M. Vorbrodt, “Growth-Promoting Effect of a Brassinosteroid in Mycelial Cultures of the fungusPsilocybe Cubensis,” Naturwissenschaften, vol. 77, no. 8, pp. 388–389, Aug. 1990. doi: 10.1007/BF01135741.
  62. J. Gartz, “Magic Mushrooms around the World.”
  63. J. Gartz, “Psilocybin in Mycelkulturen von Inocybe Aeruginascens,” Biochemie und Physiologie der Pflanzen, vol. 181, no. 7, pp. 511–517, Jan. 1986. doi: 10.1016/S0015-3796(86)80042-7.
    Psilocybin was found in the mycelium of Inocybe aeruginascens grown in surface culture. One of the most striking characters of the mycelium is the abundant production of greenish-blue sclerotia. Cultures varied in the tendency which they exhibited to produce sclerotia. The content of the alkaloid was affected by the composition of the substrat. The highest concentration of psilocybin could be detected in tissues grown in nutrient solutions. The carpophores of Inocybe aeruginascens contain more indole derivatives than the mycelia. Psilocybin could not be detected in the nutrient media. The mycelium grows also on a high-molecular carbon source, e.g. wheat flour.
  64. J. Gartz, “Verfahren zur gewinnung von tryptaminderivaten durch kultivierung hoeherer pilze,” Feb-1988. https://patents.google.com/patent/DD254395A1/en?inventor=JOCHEN+GARTz&oq=JOCHEN+GARTz&page=1.
  65. J. Gartz, “Verfahren zur gewinnung von indolalkaloiden aus pilzmycelien und naehrmedium hierfuer,” Apr-1988. https://patents.google.com/patent/DD255749A1/en?inventor=JOCHEN+GARTz&oq=JOCHEN+GARTz&page=1.
  66. J. Gartz, “Verfahren zur extraktion von indolalkaloiden aus pilzmaterial,” Jan-1989. https://patents.google.com/patent/DD264023A1/en?inventor=JOCHEN+GARTz&oq=JOCHEN+GARTz.
  67. J. Gartz, “Verfahren zur gewinnung von 4-substituierten indolverbindungen,” Nov-1989. https://patents.google.com/patent/DD273449A1/en?inventor=JOCHEN+GARTz&oq=JOCHEN+GARTz&page=1.
  68. J. Gartz, “Verfahren zur gewinnung von indolderivaten aus pilzmycelien und naehrmedium hierfuer,” Mar-1989. https://patents.google.com/patent/DD265636A1/en?inventor=JOCHEN+GARTz&oq=JOCHEN+GARTz&page=1.
  69. J. Gartz, “Verfahren zur gewinnung von indolen durch kultivierung von agaricales und naehrboden hierfuer,” Apr-1989. https://patents.google.com/patent/DD266811A1/en?inventor=JOCHEN+GARTz&oq=JOCHEN+GARTz&page=1.
  70. J. Gartz, “Verfahren zur gewinnung von tryptophanderivaten durch kultivierung hoehere pilze,” May-1990. https://patents.google.com/patent/DD278600A1/en?inventor=JOCHEN+GARTz&oq=JOCHEN+GARTz.
  71. J. Gartz, “Verfahren zur gewinnung von baeocystin durch biotransformation von tryptamin,” Feb-1991. https://patents.google.com/patent/DD287053A5/en?inventor=JOCHEN+GARTz&oq=JOCHEN+GARTz.
  72. W. J. Gibbons, M. G. McKinney, P. J. O’Dell, B. A. Bollinger, and J. A. Jones, “Homebrewed Psilocybin: Can New Routes for Pharmaceutical Psilocybin Production Enable Recreational Use?,” Bioengineered, vol. 12, no. 1, pp. 8863–8871, Jan. 2021. doi: 10.1080/21655979.2021.1987090.
    Psilocybin, a drug most commonly recognized as a recreational psychedelic, is quickly gaining attention as a promising therapy for an expanding range of neurological conditions, including depression, anxiety, and addiction. This growing interest has led to many recent advancements in psilocybin synthesis strategies, including multiple in vivo fermentation-based approaches catalyzed by recombinant microorganisms. In this work, we show that psilocybin can be produced in biologically relevant quantities using a recombinant E. coli strain in a homebrew style environment. In less than 2 days, we successfully produced approximately 300 mg/L of psilocybin under simple conditions with easily sourced equipment and supplies. This finding raises the question of how this new technology should be regulated as to not facilitate clandestine biosynthesis efforts, while still enabling advancements in psilocybin synthesis technology for pharmaceutical applications. Here, we present our homebrew results, and suggestions on how to address the regulatory concerns accompanying this new technology.
  73. K. Gotvaldová, J. Borovička, K. Hájková, P. Cihlářová, A. Rockefeller, and M. Kuchař, “Extensive Collection of Psychotropic Mushrooms with Determination of Their Tryptamine Alkaloids,” International Journal of Molecular Sciences, vol. 23, no. 22, p. 14068, Jan. 2022. doi: 10.3390/ijms232214068.
    Since not only psilocybin (PSB) but also PSB-containing mushrooms are used for psychedelic therapy and microdosing, it is necessary to know their concentration variability in wild-grown mushrooms. This article aimed to determine the PSB, psilocin (PS), baeocystin (BA), norbaeocystin (NB), and aeruginascin (AE) concentrations in a large sample set of mushrooms belonging to genera previously reported to contain psychotropic tryptamines. Ultra-high performance liquid chromatography coupled with tandem mass spectrometry was used to quantify tryptamine alkaloids in the mushroom samples. Most mushroom collections were documented by fungarium specimens and/or ITS rDNA/LSU/EF1-α sequencing. Concentrations of five tryptamine alkaloids were determined in a large sample set of 226 fruiting bodies of 82 individual collections from seven mushroom genera. For many mushroom species, concentrations of BA, NB, and AE are reported for the first time. The highest PSB/PS concentrations were found in Psilocybe species, but no tryptamines were detected in the P. fuscofulva and P. fimetaria collections. The tryptamine concentrations in mushrooms are extremely variable, representing a problem for mushroom consumers due to the apparent risk of overdose. The varied cocktail of tryptamines in wild mushrooms could influence the medicinal effect compared to therapy with chemically pure PSB, posing a serious problem for data interpretation.
  74. K. Gotvaldová, K. Hájková, J. Borovička, R. Jurok, P. Cihlářová, and M. Kuchař, “Stability of Psilocybin and Its Four Analogs in the Biomass of the Psychotropic Mushroom Psilocybe Cubensis,” Drug Testing and Analysis, vol. 13, no. 2, pp. 439–446, 2021. doi: 10.1002/dta.2950.
    Psilocybin, psilocin, baeocystin, norbaeocystin, and aeruginascin are tryptamines structurally similar to the neurotransmitter serotonin. Psilocybin and its pharmacologically active metabolite psilocin in particular are known for their psychoactive effects. These substances typically occur in most species of the genus Psilocybe (Fungi, Strophariaceae). Even the sclerotia of some of these fungi known as “magic truffles” are of growing interest in microdosing due to them improving cognitive function studies. In addition to microdosing studies, psilocybin has also been applied in clinical studies, but only its pure form has been administrated so far. Moreover, the determination of tryptamine alkaloids is used in forensic analysis. In this study, freshly cultivated fruit bodies of Psilocybe cubensis were used for monitoring stability (including storage and processing conditions of fruiting bodies). Furthermore, mycelium and the individual parts of the fruiting bodies (caps, stipes, and basidiospores) were also examined. The concentration of tryptamines in final extracts was analyzed using ultra-high-performance liquid chromatography coupled with mass spectrometry. No tryptamines were detected in the basidiospores, and only psilocin was present at 0.47 wt.% in the mycelium. The stipes contained approximately half the amount of tryptamine alkaloids (0.52 wt.%) than the caps (1.03 wt.%); however, these results were not statistically significant, as the concentration of tryptamines in individual fruiting bodies is highly variable. The storage conditions showed that the highest degradation of tryptamines was seen in fresh mushrooms stored at −80°C, and the lowest decay was seen in dried biomass stored in the dark at room temperature.
  75. J. Guerrero-Paredes, J. P. Insuasti, D. A. B. Revelo, and A. P. Soto, “Producción del hongo (Psilocybe cubensis): una revisión Production of the mushroom (Psilocybe cubensis): a review,” vol. 4, 2021.
    Psilocybe cubensis is a medicinal mushroom characterized by the extraction of psilocybin, this drug is used to relieve pain caused by diseases such as cancer, so it is necessary to improve the growing conditions to have a better product, for this research an exhaustive investigation was made about this topic.
  76. G. Guzmán and J. Ott, “Description and Chemical Analysis of a New Species of Hallucinogenic Psilocybe from the Pacific Northwest,” Mycologia, vol. 68, no. 6, pp. 1261–1267, Nov. 1976. doi: 10.1080/00275514.1976.12020019.
  77. G. Guzmán and T. Kasuya, “The Known Species of Psilocybe (Basidiomycotina, Agaricales, Strophariaceae) in Nepal,” Mycoscience, vol. 45, no. 4, pp. 295–297, Aug. 2004. doi: 10.1007/s10267-004-0186-8.
    Psilocybe percevalii, P. pseudobullacea, and P. subcubensis are reported for the first time from Nepal. Of these threes the latter is the only species with neurotropic properties. Previously, only P. montana and P. coprophila have been reported from Nepal. Psilocybe coprophila, as reported from Nepal, probably represents P. pseudobullacea.
  78. G. Guzmán, “New Studies on Hallucinogenic Mushrooms: History, Diversity, and Applications in Psychiatry,” International Journal of Medicinal Mushrooms, vol. 17, no. 11, 2015. doi: 10.1615/IntJMedMushrooms.v17.i11.10.
    This paper is a review of the new studies or new explanations of the hallucinogenic mushrooms, regarding their diversity, history, traditions, and problems in t...
  79. G. Guzmán, “Variation, Distribution, Ethnomycological Data and Relationships of Psilocybe Aztecorum, A Mexican Hallucinogenic Mushroom,” Mycologia, vol. 70, no. 2, pp. 385–396, Mar. 1978. doi: 10.1080/00275514.1978.12020239.
    Based upon data obtained from a study of the type as well as from Heim’s and Singer and Smith’s descriptions and from several fresh specimens collected by the author, Psilocybe aztecorum is here emended. This fungus is reported for the first time from four new localities in the states of Mexico and Tlaxcala. Psilocybe bonetii is reduced to varietal status of P. aztecorum. Relationships with P. baeocystis, P. quebecensis and P. cyanescens are discussed and a possible evolutionary pathway of these species is proposed. The description of P. baeocystis is emended.
  80. J. H. Halpern and H. G. Pope, “Hallucinogens on the Internet: A Vast New Source of Underground Drug Information,” American Journal of Psychiatry, vol. 158, no. 3, pp. 481–483, Mar. 2001. doi: 10.1176/appi.ajp.158.3.481.
    OBJECTIVE: The illicit use of hallucinogens is reemerging in the United States, especially among well-educated adults and teenagers. These same groups are also frequent users of the Internet. The authors sought to characterize the extent of information about hallucinogens available to Internet users. METHOD: Using standard Internet search techniques, the authors located 81 hallucinogen-related sites and categorized the information provided. RESULTS: Internet sites offer thousands of pages of information—albeit of questionable accuracy—on how to obtain, synthesize, extract, identify, and ingest hallucinogens. Much of this information has yet to appear in textbooks. By contrast, the authors found few U.S. government agency sites offering cautionary material about hallucinogen use. CONCLUSIONS: Using the Internet, potential hallucinogen users can bypass traditional channels of medical information and learn in great detail how to obtain and use numerous drugs with unknown hazards.
  81. P. He, X. He, and C. Zhang, “Interactions between Psilocybe Fasciata and Its Companion Fungus Acremonium Strictum,” Ecological Research, vol. 21, no. 3, pp. 387–395, May 2006. doi: 10.1007/s11284-005-0123-0.
    Interactions between Psilocybe fasciata and its companion fungus Acremonium strictum were analysed. The conidia of A. strictum were observed on stipe, flesh and gill of P. fasciata, which suggested that A. strictum is the fungicolous fungi or mycophilic fungi of P. fasciata. The microscopic observations of the interacting hyphae of P. fasciata and A. strictum in dual culture and the negligible activities of chitinase and β-1,3-glucanase in inducing or non-inducing media of the pure and mixed cultures of the two fungi indicated that A. strictum is not the mycoparasite of P. fasciata. In addition, the co-existence, no pigmentation and dew formation in dual culture of both fungi were observed, which implied that the interference competition between the two fungi is weak. The activities of cellobiase, filter paper enzyme, endoglucanase and xylanase of pure and mixed cultures of P. fasciata and A. strictum were the same or similar, which may allow the co-existence of the two fungi. As a consequence of coevolution, the relationship between P. fasciata and A. strictum is commensalism: A. strictum showed no clear benefit to P. fasciata; however, P. fasciata may shelter A. strictum by its psychoactive tryptamines and may be helpful to conidium dispersal of A. strictum. The relationship between P. fasciata and A. strictum is different from that of A. strictum and other fungi.
  82. A. Helbling, W. E. Horner, and S. B. Lehrer, “Comparison of Psilocybe Cubensis Spore and Mycelium Allergens,” Journal of Allergy and Clinical Immunology, vol. 91, no. 5, pp. 1059–1066, May 1993. doi: 10.1016/0091-6749(93)90220-A.
    Background: Basidiospores are an important cause of respiratory allergy in mold-sensitive atopic subjects. Collection of the large amounts of spores required for extract preparation is tedious and difficult. A desirable alternative could be mycelium grown in vitro if it is allergenically similar to spores. Methods: Therefore this study compared the allergen contents of Psilocybe cubensis spore and mycelium extracts by different techniques with the use of pooled sera from subjects who had skin test and RAST results that were positive to P. cubensis spores. Results: Isoelectric focusing immunoprints revealed six common IgE-binding bands at isoelectric points 4.7, 5.0, 5.5, 5.6, 8.7, and 9.3. Two additional bands at isoelectric points 3.9 and 5.7 were detected only in the spore extract. Sodium dodecylsulfate polyacrylamide gel electrophoresis immunoblots exhibited six common IgE-binding bands at 16, 35, 487, 52, 62, and 76 kd, 20 and 40 kd bands were present only in the spore extract. Although RAST and isoelectric focusing inhibition demonstrated that P. cubensis spore and mycelium extracts share many allergens, spores were allergenically more potent than mycelium. Conclusion: The results indicate that mycelium is a useful source of P. cubensis allergen, even though several spore allergens were not detected in mycelium.
  83. A. Helbling, E. Horner, and S. B. Lehrer, “Identification of Psilocybe Cubensis Spore Allergens by Immunoprinting,” International Archives of Allergy and Immunology, vol. 100, no. 3, pp. 263–267, 1993. doi: 10.1159/000236422.
    Previous studies established that Psilocybe cubensis contains potent allergens, and that a significant percentage of atopic subjects were sensitized to P. cubensis spores. The objective of this study was to identify P. cubensis spore allergens using isoelectric focusing (IEF) and sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) immunoprinting. Coomassie blue staining of IEF gels detected approximately 20 bands between pI 3.6 and 9.3. Immunoprints obtained with 15 P. cubensis skin test- and RAST-positive sera revealed 13 IgE-binding bands; the most reactive were at pI 5.0 (80%), 5.6 (87%), 8.7 (80%) and 9.3 (100%). SDS-PAGE resolved 27 proteins ranging from about 13 to 112 kD. SDS-PAGE immunoprints conducted with 11 skin test- and RAST-positive sera demonstrated 18 IgE-binding bands; most sera reacted to 16 (82%), 35 (100%) and 76 kD (91%) allergens. Both electrophoretic procedures demonstrated a single allergen (at pI 9.3 and 35 kD) that reacted with all sera tested. This study corroborates the allergenic significance of P. cubensis spores and identifies the allergens of greatest importance.
  84. A. Hofmann et al., “Psilocybin Und Psilocin, Zwei Psychotrope Wirkstoffe Aus Mexikanischen Rauschpilzen,” Helvetica Chimica Acta, vol. 42, no. 5, pp. 1557–1572, 1959. doi: 10.1002/hlca.19590420518.
    The psychotropically active principles of the Mexican hallucinogenic fungus Psilocybe maxicana HEIM have been isolated and obtained in crystalline form. The two new substances, which have been called psilocybin and psilocin, are present in the fruit bodies, the artificially cultivated mycelium and in the sclerotia. The dried mushroom contains 0.2 to 0.4 per cent psilocybin. Psilocin is present, at the most, in trace amounts only.
  85. A. Horita, “Some Biochemical Studies on Psilocybin and Psilocin,” Armed Services Technical Information Agency, Arlington, Virginia, AD 291 05, Oct. 1962.
  86. W. E. Horner, G. Reese, and S. B. Lehrer, “Identification of the Allergen Psi c 2 from the Basidiomycete Psilocybe Cubensis as a Fungal Cyclophilin,” International Archives of Allergy and Immunology, vol. 107, no. 1-3, pp. 298–300, 1995. doi: 10.1159/000237007.
    Basidiospores are a prevalent and frequent cause of respiratory allergies, yet their allergens remain poorly defined; thus, we have attempted a molecular characterization of representative basidiomycete allergens. A Psilocybe cubensis mycelial cDNA library was immunoscreened with patient serum. A clone was isolated that expressed a 23-kD recombinant allergen as a fusion protein and inhibited a 16-kD band (Psi c 2) in immunoprints of P. cubensis extract, indicating antigenic identity. Sequence (cDNA) analysis of the clone indicates homology with cyclophilin and the deduced amino acid sequence of Psi c 2 showed 78% identity and 4% similarity with the amino acid sequence of Schizosaccharomyces pombe cyclophilin. This recombinant allergen is a useful model for epitope analysis of basidiospore allergens and fungal allergen cross-reactivity, and may provide an improved reagent for basidiospore allergy diagnosis and treatment.
  87. B. Huskins and C. Dockery, “Detection of Psilocybin Mushroom Analogs in Chocolate: Incorporating Current Events into the Undergraduate Teaching Laboratory,” The Chemical Educator, 2009.
    In this experiment, tryptamine is used as a psilocin analog and is dispersed onto a Fisher brand cellulose laboratory sponge to simulate dehydrated mushrooms. The resulting “mushroom” material is ground, molded into chocolate, and presented to student groups for real-world and applied analyses. Students isolate the tryptamine from the chocolate using their knowledge of drug chemistry, solubility, pH, extractions, etc. Qualitative analysis is conducted by comparison to standards (Thin Layer Chromatography or Gas Chromatography) and quantitative analysis is conducted by Gas Chromatography using the method of internal standards.
  88. A. Huxley, “The Doors of Perception,” p. 24.
  89. R. E. Jackson and C. J. Alexopoulos, “Psilocybe Cubensis (Agaricales): A Comparison of Mexican and Texas Types,” The Southwestern Naturalist, vol. 21, no. 2, pp. 227–233, 1976. doi: 10.2307/3669959.
    Two populations of the mushroom Psilocybe cubensis (Earle) Singer were studied. The Houston, Texas, type completed its life cycle about seven weeks before the Palenque, Chiapas type from Mexico in vitro. There were reports and observations of toxic reactions from the Texas type. There were no reports of toxic reactions from the Chiapas type although P. cubensis from Oaxaca has been previously reported slightly toxic.
  90. N. Jensen, “Tryptamines as Ligands and Modulators of the Serotonin 5‑HT2A Receptor and the Isolation of Aeruginascin from the Hallucinogenic Mushroom Inocybe aeruginascens,” PhD thesis, Georg-August-University Göttingen, 2005. doi: 10.53846/goediss-2111.
  91. M. W. Johnson, R. R. Griffiths, P. S. Hendricks, and J. E. Henningfield, “The Abuse Potential of Medical Psilocybin According to the 8 Factors of the Controlled Substances Act,” Neuropharmacology, vol. 142, pp. 143–166, Nov. 2018. doi: 10.1016/j.neuropharm.2018.05.012.
    This review assesses the abuse potential of medically-administered psilocybin, following the structure of the 8 factors of the US Controlled Substances Act (CSA). Research suggests the potential safety and efficacy of psilocybin in treating cancer-related psychiatric distress and substance use disorders, setting the occasion for this review. A more extensive assessment of abuse potential according to an 8-factor analysis would eventually be required to guide appropriate schedule placement. Psilocybin, like other 5-HT2A agonist classic psychedelics, has limited reinforcing effects, supporting marginal, transient non-human self-administration. Nonetheless, mushrooms with variable psilocybin content are used illicitly, with a few lifetime use occasions being normative among users. Potential harms include dangerous behavior in unprepared, unsupervised users, and exacerbation of mental illness in those with or predisposed to psychotic disorders. However, scope of use and associated harms are low compared to prototypical abused drugs, and the medical model addresses these concerns with dose control, patient screening, preparation and follow-up, and session supervision in a medical facility. Conclusions (1) psilocybin has an abuse potential appropriate for CSA scheduling if approved as medicine; (2) psilocybin can provide therapeutic benefits that may support the development of an approvable New Drug Application (NDA) but further studies are required which this review describes; (3) adverse effects of medical psilocybin are manageable when administered according to risk management approaches; and (4) although further study is required, this review suggests that placement in Schedule IV may be appropriate if a psilocybin-containing medicine is approved. This article is part of the Special Issue entitled ‘Psychedelics: New Doors, Altered Perceptions’.
  92. J. Jokiranta, S. Mustola, E. Ohenoja, and M. M. Airaksinen, “Psilocybin in Finnish Psilocybe Semilanceata,” Planta Medica, vol. 50, no. 3, pp. 277–278, Jun. 1984. doi: 10.1055/s-2007-969703.
    Thieme E-Books & E-Journals
  93. Y. K, “Psilocybe Cubensis : A Hallucinogenic Fungus, New to Japan,” Proc Symp Mycol Soc Japan, vol. 41, p. 27, 1997. https://cir.nii.ac.jp/crid/1573950400471929600.
  94. T. Keller, A. Schneider, P. Regenscheit, T. Rucker, J. Jaspers, and W. Kisser, “Analysis of Psilocybin and Psilocin in Psilocybe Subcubensis GUZMA´ N by Ion Mobility Spectrometry and Gas Chromatography–Mass Spectrometry,” Forensic Science International, 1999. doi: 10.1016/S0379-0738(98)00168-6.
    A new method has been developed for the rapid analysis of psilocybin and / or psilocin in fungus material using ion mobility spectrometry. Quantitative analysis was performed by gas chromatography–mass spectrometry after a simple one-step extraction involving homogenization of the dried fruit bodies of fungi in chloroform and derivatization with MSTFA. The proposed methods resulted in rapid procedures useful in analyzing psychotropic fungi for psilocybin and psilocin. © 1999 Elsevier Science Ireland Ltd. All rights reserved.
  95. H. Klima, I. Lucic, and K. W. Kratky, “Ultraweak Photon Emission of Psilocybe Cubensis Mycelium Tissue: Comparison of Tissue Treated with Acoustic Waves and Non-Treated Tissue,” International Journal of Modelling, Identification and Control, vol. 5, no. 3, pp. 210–213, Jan. 2008. doi: 10.1504/IJMIC.2008.023123.
    All kinds of living matter (cell, organ tissue or organism) emit certain amount of light quantum that can be measured as ultraweak photon emission, also called biophoton emission. In this paper we investigate the photon emission of mycelium cultures of Psilocybe cubensis that were treated with acoustic waves of 194.71 Hz for 13 days compared to non-treated control samples.
  96. Y. Koike, K. Wada, G. Kusano, S. Nozoe, and K. Yokoyama, “Isolation of Psilocybin From Psilocybe Argentipes and Its Determination in Specimens of Some Mushrooms,” Journal of Natural Products, vol. 44, no. 3, pp. 362–365, May 1981. doi: 10.1021/np50015a023.
  97. Y. Koike, K. Wada, G. Kusano, S. Nozoe, and K. Yokoyama, “Isolation of Psilocybin From Psilocybe Argentipes and Its Determination in Specimens of Some Mushrooms,” ACS Publications. American Chemical Society, Jul-2004. doi: 10.1021/np50015a023.
  98. R. Kysilka and M. Wurst, “A Novel Extraction Procedure for Psilocybin and Psilocin Determination in Mushroom Samples,” Planta Medica, vol. 56, no. 3, pp. 327–328, Jun. 1990. doi: 10.1055/s-2006-960970.
    Thieme E-Books & E-Journals
  99. T. Laussmann and S. Meier-Giebing, “Forensic Analysis of Hallucinogenic Mushrooms and Khat (Catha edulisForsk) Using Cation-Exchange Liquid Chromatography,” Forensic Science International, vol. 195, no. 1, pp. 160–164, Feb. 2010. doi: 10.1016/j.forsciint.2009.12.013.
    Hallucinogenic mushrooms (e.g. Psilocybe and Panaeolus species) as well as leaves and young shoots of the khat tree (Catha edulisForsk) are illicit drugs in many countries. The exact concentration of the hallucinogenic alkaloids psilocin and psilocybin in mushrooms and the sympathomimetic alkaloids cathinone and cathine in khat is usually essential for jurisdiction. Facing an increasing number of mushroom and khat seizures by German customs authorities, a convenient comprehensive quantitative HPLC method based on cation-exchange liquid chromatography for these rather “exotic” drugs has been developed which avoids time-consuming multi-step sample preparation or chemical derivatization procedures. Using this method a number of different hallucinogenic fungi species and products that are mainly distributed via the internet have been analysed (dried and fresh Psilocybe cubensisSinger as well as P. cubensis collected from “grow boxes”, Panaeolus cyanescensBerkeleyandBroome and so-called “philosopher stones” (sclerotia of Psilocybe species)). Highest total amounts of psilocin have been detected in dried P. cyanescens reaching up to 3.00±0.24mg per 100mg. The distribution of khat alkaloids in different parts of the khat shoots has been studied. High concentrations of cathinone have not only been detected in leaves but also in green parts and barks of stalks. Additionally, the sample treatment for fresh mushroom and khat samples has been optimised. Highest amounts of alkaloids were found when fresh material was freeze-dried.
  100. R. E. Lee, “A Technique for the Rapid Isolation and Identification of Psilocin from Psilocin/Psilocybin-Containing Mushrooms,” Journal of Forensic Sciences, vol. 30, no. 3, pp. 931–941, Jul. 1985. doi: 10.1520/JFS11028J.
    A method has been developed for the rapid isolation and identification of psilocin from psilocin/psilocybin-containing mushrooms. Based on the difference in the solubility properties in butyl chloride of psilocin and other constituents present in psilocin/psilocybin-containing mushrooms, psilocin was easily separated in pure form.
  101. C. Lenz, J. Wick, and D. Hoffmeister, “Identification of ω-N-Methyl-4-Hydroxytryptamine (Norpsilocin) as a Psilocybe Natural Product,” Journal of Natural Products, vol. 80, no. 10, pp. 2835–2838, Oct. 2017. doi: 10.1021/acs.jnatprod.7b00407.
    We report the identification of ω-N-methyl-4-hydroxytryptamine (norpsilocin, 1) from the carpophores of the hallucinogenic mushroom Psilocybe cubensis. The structure was elucidated by 1D and 2D NMR spectroscopy and high-resolution mass spectrometry. Norpsilocin has not previously been reported as a natural product. It likely represents the actual psychotropic agent liberated from its 4-phosphate ester derivative, the known natural product baeocystin. We further present a simple and artifact-free extraction method that prevents dephosphorylation and therefore helps reflect the naturally occurring metabolic profile of Psilocybe mushrooms in subsequent analyses.
  102. C. Lenz et al., “Injury-Triggered Blueing Reactions of Psilocybe ‘Magic’ Mushrooms,” Angewandte Chemie, vol. 132, no. 4, pp. 1466–1470, 2020. doi: 10.1002/ange.201910175.
    Upon injury, psychotropic psilocybin-producing mushrooms instantly develop an intense blue color, the chemical basis and mode of formation of which has remained elusive. We report two enzymes from Psilocybe cubensis that carry out a two-step cascade to prepare psilocybin for oxidative oligomerization that leads to blue products. The phosphatase PsiP removes the 4-O-phosphate group to yield psilocin, while PsiL oxidizes its 4-hydroxy group. The PsiL reaction was monitored by in situ 13C NMR spectroscopy, which indicated that oxidative coupling of psilocyl residues occurs primarily via C-5. MS and IR spectroscopy indicated the formation of a heterogeneous mixture of preferentially psilocyl 3- to 13-mers and suggest multiple oligomerization routes, depending on oxidative power and substrate concentration. The results also imply that phosphate ester of psilocybin serves a reversible protective function.
  103. C. Lenz, A. Sherwood, R. Kargbo, and D. Hoffmeister, “Taking Different Roads: L-Tryptophan as the Origin of Psilocybe Natural Products,” ChemPlusChem, vol. 86, no. 1, pp. 28–35, 2021. doi: 10.1002/cplu.202000581.
    Psychotropic fungi of the genus Psilocybe, colloquially referred to as „magic mushrooms”, are best known for their l-tryptophan-derived major natural product, psilocybin. Yet, recent research has revealed a more diverse secondary metabolism that originates from this amino acid. In this minireview, the focus is laid on l-tryptophan and the various Psilocybe natural products and their metabolic routes are highlighted. Psilocybin and its congeners, the heterogeneous blue-colored psilocyl oligomers, alongside β-carbolines and N,N-dimethyl-l-tryptophan, are presented as well as current knowledge on their biosynthesis is provided. The multidisciplinary character of natural product research is demonstrated, and pharmacological, medicinal, ecological, biochemical, and evolutionary aspects are included.
  104. A. Y. Leung and A. G. Paul, “Baeocystin and Norbaeocystin: New Analogs of Psilocybin from Psilocybe Baeocystis,” Journal of Pharmaceutical Sciences, vol. 57, no. 10, pp. 1667–1671, Oct. 1968. doi: 10.1002/jps.2600571007.
    Two new 4-phosphoryloxytryptamine derivatives have been isolated from Psilocybe baeocystis grown in submerged cultures. Using ultraviolet, infrared, and mass spectral analyses the structures have been determined. Both are analogs of the psychotomimetic, psilocybin, and have been named baeocystin (monomethyl analog) and norbaeocystin (demethyl analog).
  105. A. Y. Leung, A. H. Smith, and A. G. Paul, “Production of Psilocybin in Psilocybe Baeocystis Saprophytic Culture,” Journal of Pharmaceutical Sciences, vol. 54, no. 11, pp. 1576–1579, 1965. doi: 10.1002/jps.2600541104.
    The carpophores of six species of mushrooms were analyzed by TLC for indole derivatives. Three species, Psilocybe baeocystis, P. caerulipes, and P. stricttipes, were found to contain psilocybin, and traces of psilocin were found to occur in the first two species. Neither compound could be detected in P. atrobrunnea, Stropharia aeruginosa, or S. semiglobata. Psilocybin has been isolated from P. baeocystis grown in submerged culture and identified by TLC, melting point, and ultraviolet and infrared spectra. Thin-layer chromatographic data for 36 indole derivatives in five solvent systems are included.
  106. W. G. Levine, “Formation of Blue Oxidation Product from Psilocybin,” Nature, vol. 215, pp. 1292–1293, 1967. https://www.nature.com/articles/2151292a0.
  107. H.-C. Lin, R. T. Hewage, Y.-C. Lu, and Y.-H. Chooi, “Biosynthesis of Bioactive Natural Products from Basidiomycota,” Organic & Biomolecular Chemistry, vol. 17, no. 5, pp. 1027–1036, Jan. 2019. doi: 10.1039/C8OB02774A.
    The Basidiomycota, also called club fungi, comprise a diverse group of fungi. Basidiomycota are strongly related to ecosystem functioning along with human life. These fungi display a wide range of bioactivities, and some are known to produce of deadly toxins or hallucinogens. Some Basidiomycota have be used as medicinal mushrooms for thousands of years. Recently, the biosyntheses of several classes of natural products from Basidiomycota have been reported. Here, we review recent studies on the biosynthetic pathways and enzymes of bioactive natural products from Basidiomycota fungi, with a focus on terpenoids, alkaloids, ribosomally synthesized and post-translationally modified peptides (RiPPs), and polyketides.
  108. LulzSwag_Technician, “Psilocybe Cubensis Strain List,” r/PsilocybinMushrooms. Aug-2022. www.reddit.com/r/PsilocybinMushrooms/comments/wounoh/psilocybe_cubensis_strain_list/.
    This took me FOREVER to edit. I found a list online of a bunch of strains and what they were crossed with so I went through and cleaned it up and then added more. I know it’s not even close to complete but if you want to comment to add to the list feel free. This is the most complete list I’ve seen so far and I’m going to try to keep it updated.I’ve tried to include the NAME of the strain, what it’s crossed with, where/who discovered it, and whatever other info available. I’m sure I’ve made mistakes but hopefully this will help people know what they’re working with at least.
  109. T. Ma, Y. Feng, X.-F. Lin, S. C. Karunarathna, W.-F. Ding, and K. D. Hyde, “Psilocybe Chuxiongensis, a New Bluing Species from Subtropical China,” Phytotaxa, vol. 156, no. 4, p. 211, Jan. 2014. doi: 10.11646/phytotaxa.156.4.3.
    A new bluing species of Psilocybe in sect. Caerulescentes is described from subtropical China. It is closely related to P. cubensis but can be differentiated by the lack of an annulus and the buff-yellow to yellowish brown, hemispheric to hemispheric-convex pileus without an umbo or papilla. Phylogenetic analyses of ITS, nrLSU and combined rpb2-tef1-α datasets using maximum parsimony and Bayesian inference also indicate its uniqueness. The relationship with P. cubensis is well-supported by molecular data with high support values in all three datasets. Psilocybe chuxiongensis sp. nov. is presented here with a description, photographs, and line drawings.
  110. V. Marcano, A. M. Méndez, F. Castellano, F. J. Salazar, and L. Martinez, “Occurrence of Psilocybin and Psilocin in Psilocybe Pseudobullacea (Petch) Pegler from the Venezuelan Andes,” Journal of Ethnopharmacology, vol. 43, no. 2, pp. 157–159, Jul. 1994. doi: 10.1016/0378-8741(94)90013-2.
    Using thin-layer Chromatographic and spectroscopic (UV) methods two Psilocybe species from the Venezuelan Andes were analysed for the hallucinogens psilocybin and psilocin. These species are: P. montana (Pers. ex Fr.) Kumm. and P. pseudobullacea (Petch) Pegler. Both hallucinogens were found in P. pseudobullacea, white P. montana was found to be exempt of these compounds.
  111. P. Margot and R. Watling, “Studies in Australian Agarics and Boletes: II. Further Studies in Psilocybe,” Transactions of the British Mycological Society, vol. 76, no. 3, pp. 485–489, Jun. 1981. doi: 10.1016/S0007-1536(81)80077-0.
    The presence of psychotomimetic indolalkylamines in three Australian species of Psilocybe is reported for the first time. Several European species of Strophariaceae are also analysed and the results compared with those from the Australian collections. Taxonomic significance of these results is discussed, particularly in respect to Australian collections of Psilocybe cubensis.
  112. T. Maruyama, O. Shirota, N. Kawahara, K. Yokoyama, Y. Makino, and Y. Goda, “Discrimination of Psychoactive Fungi (Commonly Called ‘Magic Mushrooms’) Based on the DNA Sequence of the Internal Transcribed Spacer Region.,” Journal of the Food Hygienic Society of Japan (Shokuhin Eiseigaku Zasshi), vol. 44, no. 1, pp. 44–48, 2003. doi: 10.3358/shokueishi.44.44.
  113. Y. Matsushima, F. Eguchi, T. Kikukawa, and T. Matsuda, “Historical Overview of Psychoactive Mushrooms,” Inflammation and Regeneration, vol. 29, no. 1, pp. 47–58, 2009. doi: 10.2492/inflammregen.29.47.
    Humans have used psychoactive mushrooms for medical, recreational, religious and ritual purposes since pre-history. Previous studies have clarified that psychoactive mushrooms produce psychoactive agents such as psilocybin, psilocin, ibotenic acid, and muscimol. However, the status of psychoactive mushrooms in most countries as illegal hallucinogens has prevented full investigation of their biochemical properties. Recent studies have shown that many psychoactive agents pass through the blood-brain barrier and act on neurotransmitter receptors. Psilocybin and psilocin are 5-HT1A and 5-HT2A/C receptor agonists, respectively, while ibotenic acid is a glutamic acid receptor agonist and muscimol is a GABAA receptor agonist. A new psychoactive agent, aeruginascin, has also been isolated from psychoactive mushrooms, and it is expected that more useful compounds will be discovered as the technology of component analysis advances. In addition, it has been shown that psilocybin and psilocin have high therapeutic efficiency for obsessive-compulsive disorder, which is a difficult-to-treat nervous disease. The increase of nervous diseases in modern society has thus given new importance to psychoactive mushrooms. In this review, we summarize the history of the use of psychoactive mushrooms, from pre-history to the modern age, describe their classification and distribution, survey previous studies, and discuss their therapeutic effects for difficult-to-treat nervous disease. The utilization and distribution of psychoactive mushrooms in Japan is given special attention, as there are few articles on this subject.
  114. K. McKernan, L. T. Kane, S. Crawford, C.-S. Chin, A. Trippe, and S. McLaughlin, “A Draft Reference Assembly of the Psilocybe Cubensis Genome,” F1000Research, vol. 10, p. 281, Jun. 2021. doi: 10.12688/f1000research.51613.2.
    We describe the use of high-fidelity single molecule sequencing to assemble the genome of the psychoactive Psilocybe cubensis mushroom. The genome is 46.6Mb, 46% GC, and in 32 contigs with an N50 of 3.3Mb. The BUSCO completeness scores are 97.6% with 1.2% duplicates. The Psilocybin synthesis cluster exists in a single 3.2Mb contig. The dataset is available from NCBI BioProject with accessions PRJNA687911 and PRJNA700437.
  115. K. McKernan, L. Kane, Y. Helbert, L. Zhang, N. Houde, and S. McLaughlin, “A Whole Genome Atlas of 81 Psilocybe Genomes as a Resource for Psilocybin Production.,” no. 10:961. F1000Research, Dec-2021. doi: 10.12688/f1000research.55301.2.
    The Psilocybe genus is well known for the synthesis of valuable psychoactive compounds such as Psilocybin, Psilocin, Baeocystin and Aeruginascin. The ubiquity of Psilocybin synthesis in Psilocybe has been attributed to a horizontal gene transfer mechanism of a ~20Kb gene cluster. A recently published highly contiguous reference genome derived from long read single molecule sequencing has underscored interesting variation in this Psilocybin synthesis gene cluster. This reference genome has also enabled the shotgun sequencing of spores from many Psilocybe strains to better catalog the genomic diversity in the Psilocybin synthesis pathway. Here we present the de novo assembly of 81 Psilocybe genomes compared to the P.envy reference genome. Surprisingly, the genomes of Psilocybe galindoi , Psilocybe tampanensis and Psilocybe azurescens lack sequence coverage over the previously described Psilocybin synthesis pathway but do demonstrate amino acid sequence homology to a less contiguous gene cluster and may illuminate the previously proposed evolution of psilocybin synthesis.
  116. M. G. Moldavan, A. A. Grodzinskaya, E. F. Solomko, M. L. Lomberh, S. P. Wasser, and V. M. Storozhuk, “The Effect of Psilocybe Cubensis Extract on Hippocampal Neurons in Vitro,” 2001.
  117. H. Morris, “Blood Spore: The Experimental Odyssey of a Curious Superstimulant,” Harper’s Magazine, vol. July 2013, Jul. 2013. https://harpers.org/archive/2013/07/blood-spore/.
    Of murder and mushrooms
  118. P. Munguia, “Spatial Ecology of Psilocybe (Fr.) P. Kumm. (Agaricomycetideae) Species in Two Mexican Regions,” International Journal of Medicinal Mushrooms, vol. 8, no. 4, 2006. doi: 10.1615/IntJMedMushr.v8.i4.80.
    Species of the genus Psilocybe have received much attention in the past due to their hallucinogenic compounds. Their diversity and distribution patterns,...
  119. L. G. Nagy et al., “Lessons on Fruiting Body Morphogenesis from Genomes and Transcriptomes of Agaricomycetes.” bioRxiv, p. 2021.12.09.471732, Apr-2022. doi: 10.1101/2021.12.09.471732.
    Fruiting bodies of mushroom-forming fungi (Agaricomycetes) are among the most complex structures produced by fungi. Unlike vegetative hyphae, fruiting bodies grow determinately and follow a genetically encoded developmental program that orchestrates tissue differentiation, growth and sexual sporulation. In spite of more than a century of research, our understanding of the molecular details of fruiting body morphogenesis is limited and a general synthesis on the genetics of this complex process is lacking. In this paper, we aim to comprehensively identify conserved genes related to fruiting body morphogenesis and distill novel functional hypotheses for functionally poorly characterized genes. As a result of this analysis, we report 921 conserved developmentally expressed gene families, only a few dozens of which have previously been reported in fruiting body development. Based on literature data, conserved expression patterns and functional annotations, we provide informed hypotheses on the potential role of these gene families in fruiting body development, yielding the most complete description of molecular processes in fruiting body morphogenesis to date. We discuss genes related to the initiation of fruiting, differentiation, growth, cell surface and cell wall, defense, transcriptional regulation as well as signal transduction. Based on these data we derive a general model of fruiting body development, which includes an early, proliferative phase that is mostly concerned with laying out the mushroom body plan (via cell division and differentiation), and a second phase of growth via cell expansion as well as meiotic events and sporulation. Altogether, our discussions cover 1480 genes of Coprinopsis cinerea, and their orthologs in Agaricus bisporus, Cyclocybe aegerita, Armillaria ostoyae, Auriculariopsis ampla, Laccaria bicolor, Lentinula edodes, Lentinus tigrinus, Mycena kentingensis, Phanerochaete chrysosporium, Pleurotus ostreatus, and Schizophyllum commune, providing functional hypotheses for ∼10% of genes in the genomes of these species. Although experimental evidence for the role of these genes will need to be established in the future, our data provide a roadmap for guiding functional analyses of fruiting related genes in the Agaricomycetes. We anticipate that the gene compendium presented here, combined with developments in functional genomics approaches will contribute to uncovering the genetic bases of one of the most spectacular multicellular developmental processes in fungi.
  120. S. Ncube, A. Poliwoda, E. Cukrowska, P. Wieczorek, and L. Chimuka, “Comparative Study of Different Column Types for the Separation of Polar Basic Hallucinogenic Alkaloids,” South African Journal of Chemistry, vol. 69, pp. 189–195, 2016. doi: 10.17159/0379-4350/2016/v69a23.
  121. J. M. Neal, R. G. Benedict, and L. R. Brady, “Interrelationship of Phosphate Nutrition, Nitrogen Metabolism, and Accumulation of Key Secondary Metabolites in Saprophytic Cultures of Psilocybe Cubensis, Psilocybe Cyanescens, and Panaeolus Campanulatus,” Journal of Pharmaceutical Sciences, vol. 57, no. 10, pp. 1661–1667, Oct. 1968. doi: 10.1002/jps.2600571006.
    The three basidiomycetes were grown on rotary shakers in four nutrient media containing various amounts of phosphate to determine the relative effect of this nutrient on the trichloroacetic acid-soluble and -insoluble nitrogen metabolites and to detect possible correlation between the fungal free amino acid pool (soluble nitrogen) and accumulation of characteristic hydroxytryptamine derivatives. The species were selected to represent different patterns of metabolism in fruiting bodies and vegetative mycelia. Vegetative mycelium of P. cubensis was characterized by a relatively high soluble nitrogen component and by the capacity to accumulate psilocybin and psilocin under selected conditions. The closely related P. cyanescens was less responsive to variations in phosphate nutrient and lacked the capacity to accumulate appreciable quantities of soluble nitrogenous compounds or detectable quantities of key tryptamines. P. campanulatus grown in phosphate-rich media appeared to have an adequate free amino acid pool and to excrete some exocellular nitrogen metabolites during longer incubation periods; no 5-hydroxytryptamine derivatives were detected in cultures of this fungus.
  122. L. G. Nicholas and K. Ogame, Psilocybin Mushroom Handbook: Easy Indoor & Outdoor Cultivation. Ed Rosenthal, 2006.
    Here is a practical step-by-step guide to cultivating four species of psilocybin-containing mushrooms, indoors and outside. Anyone with a clean kitchen, some basic equipment, and a closet shelf or shady flowerbed will be able to grow a bumper crop. This Handbook also includes an introduction to mushroom biology, a guide for supplies, and advice on discreetly integrating psychedelic mushrooms into outdoor gardens. Hand-drawn illustrations and full-color and black-&-white photographs provide the reader with steps in the cultivation process and exact identification of desired species. The four species detailed include two species that have previously had very little coverage: Psilocybe mexicana (a tiny mushroom used for millennia by indigenous Mexican shamans) and Psilocybe azurescens (a newly described species native to the Pacific Northwest and easily grown outdoors on woodchips). This innovative book also offers a wealth of information about the use of psilocybin-containing mushrooms in both traditional and modern contexts. Contributing ethnobotanist Kathleen Harrison highlights the history, ritual and mythology of sacred Psilocybe mushrooms used in indigenous shamanic settings. The book’s authors offer insights into how these principles might be put into practice by the modern voyager, to provide, safe, healing and fruitful journeys.
  123. “ACKNOWLEDGMENTS,” p. 139.
  124. “Ask Dr. Shulgin Online: 4-Hydroxy-5-Methoxy-N,N-Dimethyltryptamine, Psilocybe Mushrooms, Psilocin.” . http://www.cognitiveliberty.org/ccle1/shulgin/blg/2005/12/4-hydroxy-5-methoxy-nn_07.html.
  125. “The Azurescens Foundation,” p. 20.
  126. “CULTURAL STUDIES ON THE MUSHROOM PSILOCYBE CUBENSIS - ProQuest.” . https://www.proquest.com/openview/abf2dca474c27264a2affad4925db83b/1?pq-origsite=gscholar&cbl=18750&diss=y.
    Explore millions of resources from scholarly journals, books, newspapers, videos and more, on the ProQuest Platform.
  127. “Extraction Method for the Isolation and Identification of Psilocin - [Www.Rhodium.Ws].” . https://www.designer-drug.com/pte/12.162.180.114/dcd/chemistry/psilocin.extraction.html.
  128. “FBI Report.” . https://www.seanet.com/ rwmcpherson/forensic.htm.
  129. “Get to Know the Beta Carbolines – Newcomers to the Science of Magic Mushrooms,” Psychedelic Science Review. Jul-2020. https://psychedelicreview.com/get-to-know-the-beta-carbolines-newcomers-to-the-science-of-magic-mushrooms/.
    Understanding the pharmacology of magic mushrooms is not just about tryptamines anymore.
  130. The Hallucinogenic Mushrooms: Diversity, Traditions, Use and Abuse with Special Reference to the Genus Psilocybe. CRC Press, 2019, pp. 256–277. doi: 10.1201/9780429061653-11.
    The traditions, uses and abuses, and studies of hallucinogenic mushrooms, mostly species of Psilocybe, are reviewed and critically analyzed. Amanita muscaria seems to be the oldest hallucinogenic mushroom used by man, although the first hallucinogenic substance, LSD, was isolated from ergot, Claviceps purpurea. Amanita muscaria is still used in North Eastern Siberia and by some North American Indians. In the past, some Mexican Indians, as well as Guatemalan Indians possibly used A. muscaria. Psilocybe has more than 150 hallucinogenic species throughout the world, but they are used in traditional ways only in Mexico and New Guinea. Some evidence suggests that a primitive tribe in the Sahara used Psilocybe in religions ceremonies centuries before Christ. New ethnomycological observations in Mexico are also described.
  131. “HPLC/DAD Analysis of Psilocin and Psilocybin in Psilocybe Cubensis.” . https://air.unimi.it/handle/2434/146135.
  132. “The Mushroom Entheogen - the Measure of the Mushroom.” . http://en.psilosophy.info/the_mushroom_entheogen_the_measure_of_the_mushroom.html.
  133. “New Aspects of the Occurrence Chemistry and Cultivation of European Hallucinogenic Mushrooms.” . http://en.psilosophy.info/new_aspects_of_the_occurrence_chemistry_and_cultivation_of_european_hallucinogenic_mushrooms.html.
  134. “PhillyGoldenTeacher - YouTube.” . https://www.youtube.com/.
    Enjoy the videos and music you love, upload original content, and share it all with friends, family, and the world on YouTube.
  135. “THE PRODUCTION OF PSILOCYBIN IN SUBMERGED CULTURE BY PSILOCYBE CUBENSIS - ProQuest.” . https://www.proquest.com/openview/70c4594353a221173629819fcfcdeb40/1.pdf?pq-origsite=gscholar&cbl=18750&diss=y.
    Explore millions of resources from scholarly journals, books, newspapers, videos and more, on the ProQuest Platform.
  136. “Shroom Gummy Bears 🐻 - The Psychedelic Experience - Shroomery Message Board.” . https://www.shroomery.org/forums/showflat.php/Number/26936656/fpart/all.
  137. “Species Diversity of the Genus Psilocybe (Basidiomycotina, Agaricales, Strophariaceae) in the World Mycobiota, with Special Attention to Hallucinogenic Properties,” International Journal of Medicinal Mushrooms, vol. 7, no. 1 & 2, 2005. doi: 10.1615/IntJMedMushr.v7.i12.280.
    An exhaustive world revision of all names considered in the genus Psilocybe s.l. is presented, of which the hallucinogenic species were treated with spec...
  138. “Swim’s Psychedelic Cook Book: Mescaline, DMT and Harmalas,” p. 198.
  139. “Toxic and Hallucinogenic Mushroom Identification via DNA-Based Macroarray - ProQuest.” . https://www.proquest.com/openview/7cc70eab4c70ee759bc6f9e4cabb980e/1?pq-origsite=gscholar&cbl=18750.
    Explore millions of resources from scholarly journals, books, newspapers, videos and more, on the ProQuest Platform.
  140. K. G. Nugent and B. J. Saville, “Forensic Analysis of Hallucinogenic Fungi: A DNA-Based Approach,” Forensic Science International, vol. 140, no. 2, pp. 147–157, Mar. 2004. doi: 10.1016/j.forsciint.2003.11.022.
    Hallucinogenic fungi synthesize two controlled substances, psilocin and psilocybin. Possession of the fungal species that contain these compounds is a criminal offence in North America. Some related species that are morphologically similar, do not contain the controlled substances. Therefore, unambiguous identification of fungi to the species level is critical in determining if a mushroom is illegal. We investigate a phylogenetic approach for the identification of species that contain the psychoactive compounds. We analyzed 35 North American specimens representing seven different genera of hallucinogenic and non-hallucinogenic mushrooms. We amplified and sequenced the internal transcribed spacer region of the rDNA (ITS-1) and a 5′ portion of the nuclear large ribosomal subunit of rRNA (nLSU rRNA or 28S). ITS-1 locus sequence data was highly variable and produced a phylogenetic resolution that was not consistent with morphological identifications. In contrast, the nLSU rRNA data clustered isolates from the same species and separated hallucinogen containing and non-hallucinogen containing isolates into distinct clades. With this information, we propose an approach that combines the specificity of PCR detection and the resolving power of phylogenetic analysis to efficiently and unambiguously identify hallucinogenic fungal specimens for legal purposes.
  141. E. Ohenoja, J. Jokiranta, T. Mäkinen, A. Kaikkonen, and M. M. Airaksinen, “The Occurrence of Psilocybin and Psilocin in Finnish Fungi,” ACS Publications. American Chemical Society, Jul-2004. doi: 10.1021/np50052a030.
  142. O. T. Oss, O. N. Oeric, and I. the Obscure, Psilocybin, Magic Mushroom Grower’s Guide: A Handbook for Psilocybin Enthusiasts. Berkeley, Calif: And/Or Press, 1976.
    Pp. 63; 8 full page color plates, 54 text figures (black-and-white photos and line-drawings). Color pictorial stiff wrappers, sm 8vo. Psilocybin is one of the most active, and least toxic of all psychedelics. It is very well be the most perfect psychedelic! No complex chemicals, special equipment or knowledge of chemistry is needed to grow Psilocybin mushrooms at home. Given the spores, all that is needed to grow them is a little bit of grain, some chalk, a pressure cooker and a few mason jars. Signature of Nancy Weber on the title page. No other ownership marks.
  143. O. T. Oss and O. N. Oeric, Psilocybin: Magic Mushroom Grower’s Guide : A Handbook for Psilocybin Enthusiasts. Ed Rosenthal, 1992.
    In the 1970s, two of the most influential thinkers of the psychedelic era gathered what was then known about psilocybin botany and culture and presented it in Psilocybin: Magic Mushroom Grower’s Guide. Writing under pseudonyms, the McKenna brothers provided simple, reliable, and productive methods for magic mushroom propagation, including black-and-white photographs that showed the techniques of the time.The development of more modern cultivation techniques does not eclipse the cultural contributions of this book. Philosophical asides, whimsical illustrations evoking the mystical nature of mushrooms, and speculations about the relationship of these organisms to humankind provide a lasting legacy. Truly the classic manual on home cultivation, the wisdom of Psilocybin: Magic Mushroom Grower’s Guide continues to inspire new students of psycho-mycology–and refreshes psychedelic memories for others.
  144. J. OTT and \relax G. U. Z. M. A. N. G, “Detection of Psilocybin in Species of Psilocybe, Panaeolus and Psathyrella.,” 1976.
  145. T. Passie, J. Seifert, U. Schneider, and H. M. Emrich, “The Pharmacology of Psilocybin,” Addiction Biology, vol. 7, no. 4, pp. 357–364, Oct. 2002. doi: 10.1080/1355621021000005937.
    Psilocybin (4-phosphoryloxy-N,N-dimethyltryptamin e) is the major psychoactive alkaloid of some species of mushrooms distributed worldwide.These mushrooms represent a growing problem regarding hallucinogenic drug abuse. Despite its experimental medical use in the 1960s, only very few pharmacological data about psilocybin were known until recently. Because of its still growing capacity for abuse and the widely dispersed data this review presents all the available pharmacological data about psilocybin.
  146. H. Nevalainen, L. Kautto, and J. Te’o, “Methods for Isolation and Cultivation of Filamentous Fungi,” in Environmental Microbiology, vol. 1096, I. T. Paulsen and A. J. Holmes, Eds. Totowa, NJ: Humana Press, 2014, pp. 3–16. doi: 10.1007/978-1-62703-712-9_1.
    Filamentous fungi are important organisms for basic discovery, industry, and human health. Their natural growth environments are extremely variable, a fact reflected by the numerous methods developed for their isolation and cultivation. Fungal culture in the laboratory is usually carried out on agar plates, shake flasks, and bench top fermenters starting with an inoculum that typically features fungal spores. Here we discuss the most popular methods for the isolation and cultivation of filamentous fungi for various purposes with the emphasis on enzyme production and molecular microbiology.
  147. J. Pechwang et al., “Biotransformation of Ent-Kaur-16-En-19-Oic Acid by Psilocybe Cubensis,” Natural Product Research, vol. 24, no. 10, pp. 905–914, Jun. 2010. doi: 10.1080/14786410802420739.
    Biotranformation of ent-kaur-16-en-19-oic acid (1) using Psilocybe cubensis resulted in hydroxylated products. After two days of incubation, ent-16β,17-dihydroxy-kauran-19-oic acid (2) was isolated. After further incubation for nine days, two novel metabolites, ent-12α,16β,17-trihydroxy-kauran-19-oic acid (3) and ent-11α,16β,17-trihydroxy-kauran-19-oic acid (4), were obtained. The metabolites were identified by spectroscopic methods and X-ray crystallography. Compounds 1–4 were evaluated for their cytotoxic properties against the human leukaemia K562 cell line; only compound 1 showed moderate activity.
  148. M. Pellegrini, M. C. Rotolo, E. Marchei, R. Pacifici, F. Saggio, and S. Pichini, “Magic Truffles or Philosopher’s Stones: A Legal Way to Sell Psilocybin?,” Drug Testing and Analysis, vol. 5, no. 3, pp. 182–185, 2013. doi: 10.1002/dta.1400.
    “Magic mushrooms” is the most common name given to hallucinogenic fungi containing the psychoactive alkaloids psilocybin and psilocin. In recent years, fungis’ sclerotia, commonly called “magic truffles” have become a form of supply of psychoactive Psilocybe alkaloids since Psilocybe sclerotia are not specifically included in the laws banning the sale, the purchase and the use of such substances and mushrooms containing them. A liquid chromatography –tandem mass spectrometry (LC-MS/MS) method was developed for the rapid determination of psilocybin and psilocin in Psilocybe sclerotia. Following a simple step extraction with methanol, the alkaloids were separated on a reversed-phase column using a gradient of 0.1% formic acid – acetonitrile s a mobile phase at a flow rate of 0.2 mL/min.. Separated analytes were detected by electrospray ionization tandem mass spectrometry in the positive ion mode using multiple reaction monitoring. The developed method was linear over the calibration range for all two substances under investigation, with a r2 > 0.99. The detection and quantification limits were 0.3 µg and 1 µg per 100 mg truffles, for both psilocin and psilocybin and the intra- and inter-day coefficients of variation were always better than 15%. Using this method, the presence of only psilocybin was demonstrated in examined Psilocybe sclerotia. The content of psilocybin was found to vary over a concentration range of 59.3 to 167.8 µg per 100 mg of fresh sclerotia. Copyright © 2012 John Wiley & Sons, Ltd.
  149. M. Perkal, G. L. Blackman, A. L. Ottrey, and L. K. Turner, “Determination of Hallucinogenic Components of Psilocybe Mushrooms Using High-Performance Liquid Chromatography,” Journal of Chromatography A, vol. 196, no. 1, pp. 180–184, Aug. 1980. doi: 10.1016/S0021-9673(00)80375-1.
    The retention behavior of primary, secondary and tertiary amines was studied using normal-phase-HPLC on silica, diol, and cyano stationary phases. Several classes of amines, including benzylamines, anilines, ephedrines, tryptamines, and azatryptamines were chromatographed using mixtures of hexane and ethoxynonafluorobutane with methylene chloride and methanol. Peak tailing, diminished selectivity and low plate count were minimized by the addition of volatile amines to the mobile phase. The optimal additive was n-propylamine at 0.1% concentration. On diol columns, the elution order of free primary, N-N-methyl, and N,N-dimethylamines was predictable, while the elution order of primary and secondary amines on cyano columns varied depending on the alcohol modifier concentration. The feasibility of preparative normal-phase chromatography was demonstrated by the separation of a mixture of primary, secondary and tertiary amines obtained by direct methylation of norephedrine. The procedures described may provide a practical alternative to traditional methods of analysis and purification of potential drug candidates. The determination of psilocybin was carried out by reversed-phase liquid chromatography (HPLC) with fluorescence (FL) detection. Psilocybin was labeled with 5-dimethylaminonaphthalene-1-[N-(2-aminoethyl)]sulfonamide (DNS-ED) at 60 °C for 4 h in the presence of 1-ethyl-3-(3-dimethylaminopropyl)-carbodiimide (EDC) as the activation reagent. The resulting derivative was separated on a Mightysil RP-18 GP column (150 mm × 4.6 mm, i.d. 3 μm) with the mixture of 50 mM ammonium acetate (AcONH4) and CH3CN, and detected at 539 nm (excitation at 321 nm). The structure of the derivative was identified by HPLC–ESI-MS. A good linear relation of the calibration curve of psilocybin was observed under the proposed conditions for labeling, separation and detection. The quantification limit was 4.4 ng in 1 mg dried mushroom. The proposed procedure was successfully used for the determination of psilocybin in real samples. The contents of psilocybin in six magic mushrooms by the proposed HPLC–FL method were less than 20.0 ng in 1 mg dried samples. A new method has been developed for the rapid analysis of psilocybin and/or psilocin in fungus material using ion mobility spectrometry. Quantitative analysis was performed by gas chromatography–mass spectrometry after a simple one-step extraction involving homogenization of the dried fruit bodies of fungi in chloroform and derivatization with MSTFA. The proposed methods resulted in rapid procedures useful in analyzing psychotropic fungi for psilocybin and psilocin. In order to investigate the pharmacokinetic properties of psilocybin (PY), the main psychoactive compound of Psilocybe mushrooms, high performance liquid chromatographic procedures with column-switching coupled with electrochemical detection (HPLC-ECD) for reliable quantitative determination of the PY metabolites psilocin (PI) and 4-hydroxyindole-3-acetic acid (4HIAA) in human plasma were established. Sample work-up includes protection of the highly unstable phenolic analytes with ascorbic acid, freeze-drying and in-vitro microdialysis. The data of two controlled clinical studies with healthy volunteers are presented. The subjects (N = 6 for both studies) received single oral PY doses of 0.224 ± 0.02 mg/kg b.wt. (10–20 mg) and intravenous doses of 1 mg PY, respectively. Peak plasma levels of PI after oral administration of PY were measured after 105 ± 37 min showing an average concentration of 8.2 ± 2.8 ng PI/ml plasma. 4HIAA peak concentrations of 150 ± 61 ng/ml plasma were found 113 ± 41 min after ingestion of PY. After intravenous administration, a mean PI maximum plasma concentration of 12.9 ± 5.6 ng/ml plasma was found 1.9 ± 1.0 min after injection. The maximum plasma levels appearing within a very short period indicate a rapid dephosphorylation of PY also when administered systemically. 4HIAA was not detected after 1 mg of intravenous PY. Estimates for the absolute bioavailability of PI after oral administration of PY were 52.7 ± 20% (N = 3). An efficient analytical and isolation method was elaborated for biologically active tryptamines using a computer-aided liquid chromatographic-gas chromatographic system. The separation method includes a new efficient extraction procedure, optimization programmed for high-performance liquid chromatographic separation, identification by diode-array detection and a spectrometric and electrochemical assay. The identification of indole alkaloids was confirmed by thin-layer and gas chromatography and mass spectrometry. The method was used for analysis and isolation of psychotropic substances in extracts from the fruit bodies of hallucinogenic fungi of genera Psilocybe, Inocybe and Amanita and in mycelial extracts from the species Psilocybe bohemica.
  150. J. Picker and R. W. Rickards, “The Occurrence of the Psychotomimetic Agent Psilocybin in an Australian Agaric, Psilocybe Subaeruginosa,” Australian Journal of Chemistry, vol. 23, no. 4, pp. 853–855, 1970. doi: 10.1071/ch9700853.
    Australian Journal of Chemistry - an International Journal for Chemical Science publishes research papers from all fields of chemical science.
  151. L. Plotnik, G. Gibbs, and T. Graham, “Psilocybin Conspectus: Status, Production Methods, and Considerations,” International Journal of Medicinal Mushrooms, vol. 24, no. 1, 2022. doi: 10.1615/IntJMedMushrooms.2021041921.
    Psilocybin is a psychoactive alkaloid that is produced naturally by approximately 200 species of mushrooms. The potential medical use of this molecule for the t...
  152. H. Rafati, H. Riahi, and A. Mohammadi, “Enhancement of Indole Alkaloids Produced by Psilocybe Cubensis (Earle) Singer (Agaricomycetideae) in Controlled Harvesting Light Conditions,” International Journal of Medicinal Mushrooms, vol. 11, no. 4, pp. 419–426, 2009. doi: 10.1615/IntJMedMushr.v11.i4.80.
    Indole alkaloids of Psilocybe cubensis have been reported to be the potential candidates for drug discovery in central nervous system (CNS) disorders. In...
  153. Z. A. Mahmood, “Bioactive Alkaloids from Fungi: Psilocybin,” in Natural Products, K. G. Ramawat and J.-M. Mérillon, Eds. Berlin, Heidelberg: Springer Berlin Heidelberg, 2013, pp. 523–552. doi: 10.1007/978-3-642-22144-6_19.
  154. S. A. Redhead, J.-M. Moncalvo, R. Vilgalys, P. B. Matheny, L. Guzmán-Dávalos, and G. Guzmán, “(1757) Proposal to Conserve the Name Psilocybe (Basidiomycota) with a Conserved Type,” Taxon, vol. 56, no. 1, pp. 255–257, 2007. https://www.jstor.org/stable/25065762.
  155. D. B. Repke, D. T. Leslie, and G. Guzmán, “Baeocystin in Psilocybe, Conocybe and Panaeolus,” Lloydia, vol. 40, no. 6, pp. 566–578, Nov. 1977.
    Sixty collections of ten species referred to three families of the Agaricales have been analyzed for the presence of baeocystin by thin-layer chromatography. Baeocystin was detected in collections of Psilocybe, Conocybe, and Panaeolus from the U.S.A., Canada, Mexico, and Peru. Laboratory cultivated fruitbodies of Psilocybe cubensis, P. semilanceata, and P. cyanescens were also studied. Intra-species variation in the presence of decay rate of baeocystin, psilocybin and psilocin are discussed in terms of age and storage factors. In addition, evidence is presented to support the presence of 4-hydroxytryptamine in collections of P. baeocystis and P. cyanescens. The possible significance of baeocystin and 4-hydroxytryptamine in the biosynthesis of psilocybin in these organisms is discussed.
  156. D. B. Repke, D. T. Leslie, D. M. Mandell, and N. G. Kish, “GLC-Mass Spectral Analysis of Psilocin and Psilocybin,” Journal of Pharmaceutical Sciences, vol. 66, no. 5, pp. 743–744, 1977. doi: 10.1002/jps.2600660539.
    With the combined technique of GLC-mass spectrometry, psilocin and psilocybin, two hallucinogenic indoles, were analyzed as their trimethylsilyl derivatives. The method was applied to these two components in an extract of Psilocybe cubensis (Earle) Sing.
  157. A. Roberts, “An Overview of Decriminalization Efforts in Regard to Psychedelic Plants in the United States, 2019-2020,” no. 3944724. Rochester, NY, Oct-2021. doi: 10.2139/ssrn.3944724.
    This paper examines the recent developments made in psychedelic-related drug policy in the United States. The paper gives an overview of the decriminalization efforts made at the state and local levels. The paper also looks at the historical, cultural, political, and public health factors that have shaped psychedelic policy throughout American history and into the current day. Lastly, the paper shares some concerns about discrimination and unequal access present in psychedelic-assisted psychotherapy.
  158. J. C. Rodriguez Martinez, “The Tolimas and the Mushroom: Mycolatry in Pre-Hispanic Colombia,” in Biology, Cultivation and Applications of Mushrooms, A. Arya and K. Rusevska, Eds. Singapore: Springer, 2022, pp. 487–496. doi: 10.1007/978-981-16-6257-7_17.
    Mushrooms have made a significant role in the ancient civilizations of Greece, India, Russia, and Mesoamerica among other countries. The use of entheogens to reach religious ecstasy and communicate with deities was common in the ancient indigenous tribes of pre-Hispanic America and it’s a practice that still continues to this day. It seems a mushroom cult was practiced by the members of Tolima indigenous cultures in pre-Hispanic ColombiaColombia. It is a common belief that mushroom cults entered ColombiaColombia from the north, from Mexico to Panama and to ColombiaColombia. The origins of the cults coming from the south also should not be discarded. And of course the cult originating in Colombian territories is a possibility. On the basis of pre-Hispanic objects displayed in the Museo del Oro, located in Bogotá, ColombiaColombia, certain facts, stories of past lives, and thoughts of many people who love these objects and had studied them in the past are analyzed. They are objects that gave life to a collection by sharing stories of their past, strengthening ties with the communities and contexts to which they belong, and, above all, have been creating and generating a unique ancient knowledge from the cultures of the past to our own present.
  159. S. E. Rodriguez-Cruz, “Analysis and Characterization of Psilocybin and Psilocin Using Liquid Chromatography - Electrospray Ionization Mass Spectrometry (LC-ESI-MS) with Collision-Induced-Dissociation (CID) and Source-Induced-Dissociation (SID),” Microgram Journal, vol. 3, no. 3-4, pp. 175–182, Jul. 2005.
    The rapid analysis of psilocybin and psilocin using liquid chromatography electrospray ionization mass spectrometry (LC-ESI-MS) is presented. Full-scan MS experiments provide molecular weight information, but little fragmentation. Similarly, collision-induced-dissociation (CID) experiments generate only a limited number of fragments. However, source-induced-dissociation (SID) experiments result in more extensive fragmentation. The combined results from these complementary techniques allows for the more complete characterization of psilocin and the thermally-labile psilocybin.
  160. H. Roger, H. Albert, B. Artur, K. Hans, and C. Roger, “Obtaining Psilocybin and Psilocin from Fungal Material,” May-1965. https://patents.google.com/patent/US3183172A/en.
  161. D. G. Ruch and J. J. Motta, “Ultrastructure and Cytochemistry of Dormant Basidiospores of Psilocybe Cubensis,” Mycologia, vol. 79, no. 3, pp. 387–398, May 1987. doi: 10.1080/00275514.1987.12025395.
    The ultrastructure of dormant basidiospores of Psilocybe cubensis is described. The spore wall is characterized by three distinct layers and a germ pore. A pore cap is described for the first time in a species of Psilocybe. The protoplast is surrounded by a well defined plasma membrane with many distinct invaginations. Internally, large numbers of nonmembrane-bound lipids occur at both ends of the spore. Two nuclei are typically present and the nuclear envelope has many nuclear pores. Mitochondria with only a few, but well delineated, plate-like cristae are present. There is scant endoplasmic reticulum. Ribosomes occur regularly attached to the ER and outer mitochondrial membrane, as well as being densely packed throughout the cytoplasm. Variously sized vacuoles were demonstrated cytochemically to contain acid phosphatase. Microbody-like organelles are described but cytochemical tests to determine if these organelles are functionally similar to those of higher plants were unsuccessful. Cell-free extracts of basidiospores exhibit catalase and isocitrate lyase activity but no malate synthase activity.
  162. S. Russell, The Essential Guide to Cultivating Mushrooms: Simple and Advanced Techniques for Growing Shiitake, Oyster, Lion’s Mane, and Maitake Mushrooms at Home. North Adams, MA: Storey Publishing, 2014.
  163. K. Saito, T. Toyo’oka, M. Kato, T. Fukushima, O. Shirota, and Y. Goda, “Determination of Psilocybin in Hallucinogenic Mushrooms by Reversed-Phase Liquid Chromatography with Fluorescence Detection,” Talanta, vol. 66, no. 3, pp. 562–568, Apr. 2005. doi: 10.1016/j.talanta.2004.11.031.
    The determination of psilocybin was carried out by reversed-phase liquid chromatography (HPLC) with fluorescence (FL) detection. Psilocybin was labeled with 5-dimethylaminonaphthalene-1-[N-(2-aminoethyl)]sulfonamide (DNS-ED) at 60°C for 4h in the presence of 1-ethyl-3-(3-dimethylaminopropyl)-carbodiimide (EDC) as the activation reagent. The resulting derivative was separated on a Mightysil RP-18 GP column (150mm×4.6mm, i.d. 3μm) with the mixture of 50mM ammonium acetate (AcONH4) and CH3CN, and detected at 539nm (excitation at 321nm). The structure of the derivative was identified by HPLC–ESI-MS. A good linear relation of the calibration curve of psilocybin was observed under the proposed conditions for labeling, separation and detection. The quantification limit was 4.4ng in 1mg dried mushroom. The proposed procedure was successfully used for the determination of psilocybin in real samples. The contents of psilocybin in six magic mushrooms by the proposed HPLC–FL method were less than 20.0ng in 1mg dried samples.
  164. S. G. Saupe, “Occurrence of Psilocybin/Psilocin in Pluteus Salicinus (Pluteaceae),” Mycologia, vol. 73, no. 4, pp. 781–784, Jul. 1981. doi: 10.1080/00275514.1981.12021406.
  165. T. Schäfer, K. Kramer, S. Werten, B. Rupp, and D. Hoffmeister, “Characterization of the Gateway Decarboxylase for Psilocybin Biosynthesis,” ChemBioChem, vol. n/a, no. n/a, p. e202200551, 2022. doi: 10.1002/cbic.202200551.
    The l-tryptophan decarboxylase PsiD catalyzes the initial step of the metabolic cascade to psilocybin, the major indoleethylamine natural product of the “magic” mushrooms and a candidate drug against major depressive disorder. Unlike numerous pyridoxal phosphate (PLP)-dependent decarboxylases for natural product biosyntheses, PsiD is PLP-independent and resembles type II phosphatidylserine decarboxylases. Here, we report on the in vitro biochemical characterization of Psilocybe cubensis PsiD along with in silico modeling of the PsiD structure. A non-canonical serine protease triad for autocatalytic cleavage of the pro-protein was predicted and experimentally verified by site-directed mutagenesis.
  166. A. M. Sherwood et al., “Synthesis and Biological Evaluation of Tryptamines Found in Hallucinogenic Mushrooms: Norbaeocystin, Baeocystin, Norpsilocin, and Aeruginascin,” Journal of Natural Products, vol. 83, no. 2, pp. 461–467, Feb. 2020. doi: 10.1021/acs.jnatprod.9b01061.
    A general synthetic method was developed to access known tryptamine natural products present in psilocybin-producing mushrooms. In vitro and in vivo experiments were then conducted to inform speculations on the psychoactive properties, or lack thereof, of the natural products. In animal models, psychedelic activity by baeocystin alone was not evident using the mouse head twitch response assay, despite its putative dephosphorylated metabolite, norpsilocin, possessing potent agonist activity at the 5-HT2A receptor.
  167. O. Shirota, W. Hakamata, and Y. Goda, “Concise Large-Scale Synthesis of Psilocin and Psilocybin, Principal Hallucinogenic Constituents of ‘Magic Mushroom,’” Journal of Natural Products, vol. 66, no. 6, pp. 885–887, Jun. 2003. doi: 10.1021/np030059u.
  168. A. Sinclair, “High Times in Ancient Egypt: The Use and Abuse of Psychoactive Plant Identifications in Alternative Egyptology,” 14.04.21. https://hcommons.org/deposits/item/hc:41627/.
    Text to a presentation on the misrepresentation of ancient Egyptian psychoactive consumption in academic publications and public media that was given by me at the Alternative Egyptology Symposium, hosted by the Allard Pierson Museum, Amsterdam, 14-04-2021. There is an academic paper in preparation.
  169. J. Solano, L. Anabalón, S. Figueroa, C. Lizama, L. C. Reyes, and D. Gangitano, “Psychedelic Fungus (Psilocybe Sp.) Authentication in a Case of Illegal Drug Traffic: Sporological, Molecular Analysis and Identification of the Psychoactive Substance,” Science & Justice, vol. 59, no. 1, pp. 102–108, Jan. 2019. doi: 10.1016/j.scijus.2018.08.005.
    In nature, there are >200 species of fungi with hallucinogenic properties. These fungi are classified as Psilocybe, Gymnopilus, and Panaeolus which contain active principles with hallucinogenic properties such as ibotenic acid, psilocybin, psilocin, or baeocystin. In Chile, fungi seizures are mainly of mature specimens or spores. However, clandestine laboratories have been found that process fungus samples at the mycelium stage. In this transient stage of growth (mycelium), traditional taxonomic identification is not feasible, making it necessary to develop a new method of study. Currently, DNA analysis is the only reliable method that can be used as an identification tool for the purposes of supporting evidence, due to the high variability of DNA between species. One way to identify the species of a distinctive DNA fragment is to study PCR products analyzed by real time PCR and sequencing. One of the most popular sequencing methods of forensic interest at the generic and intra-generic levels in plants is internal transcribed spacer (ITS). With real time PCR it is possible to distinguish PCR products by differential analysis of their melting temperature (Tm) curves. This paper describes morphological, chemical, and genetic analysis of mycelia of psychedelic fungi collected from a clandestine laboratory. The fungus species were identified using scanning electron microscopy (SEM), mass spectrometry, HRM analysis, and ITS sequencing. The sporological studies showed a generally smooth surface and oval shape, with maximum length 10.1 μm and width 6.4 μm. The alkaloid Psilocyn was identified by mass spectrometry, while HRM analysis and ITS sequencing identified the species as Psilocybe cubensis. A genetic match was confirmed between the HRM curves obtained from the mycelia (evidence) and biological tissue extracted from the fruiting bodies. Mycelia recovered from the evidence and fruiting bodies (control) were genetically indistinguishable.
  170. S. R. Sommano et al., “Novel Perspective of Medicinal Mushroom Cultivations: A Review Case for ‘Magic’ Mushrooms,” Agronomy, vol. 12, no. 12, p. 3185, Dec. 2022. doi: 10.3390/agronomy12123185.
    Fruiting bodies, mycelia, or spores in the form of extracts or powder of various medicinal mushrooms are used to prevent, treat, or cure a range of ailments and balance a healthy diet. Medicinal mushrooms are found in several genera of fungi and their fruit bodies, cultured mycelia, and cultured broth contains phytochemical constituents such as triterpenes, lectins, steroids, phenols, polyphenols, lactones, statins, alkaloids, and antibiotics. Edible mushrooms are considered functional foods that can be used as supplements for complementary and alternative medicines where the markets are growing rapidly. Several species of edible mushrooms possess therapeutic potential and functional characteristics. The psilocybin-containing types, sometimes known as magic mushrooms, have been utilized for generations by indigenous communities due to their hallucinogenic, medicinal, and mind-manifestation properties. Recent clinical research also convinces that these psychedelics have the potential to treat addiction, depression, anxiety, and other mental health concerns. This has escalated the demand for the natural products derived from the mushrooms of these sources, yet the agronomic aspect and biotechnology approaches to produce the active ingredients are not collectively documented. The objectives of this review article are to examine the general type and variation of therapeutic mushrooms, especially those belonging to the Psilocybe. The biotechnology approach for cultivation and the production of secondary metabolites is also appraised. The ultimate purposes are to provide guidance for farmers and companies to pursue sustainable ways to produce natural products for the development of functional food and pharmaceuticals and to support the alteration of the stigmatic drug concerns around psychedelic mushrooms.
  171. P. Stamets and J. S. Chilton, The Mushroom Cultivator: A Practical Guide to Growing Mushrooms at Home. Olympia, Wash. : Seattle, Wa: Agarikon Press ; Western distribution by Homestead Book Co, 1983.
  172. K. Stebelska, “Chapter 84 - Assays for Detection of Fungal Hallucinogens Such as Psilocybin and Psilocin,” in Neuropathology of Drug Addictions and Substance Misuse, V. R. Preedy, Ed. San Diego: Academic Press, 2016, pp. 909–926. doi: 10.1016/B978-0-12-800212-4.00084-4.
    Methods applicable to the detection of hallucinogenic indole compounds, namely psilocybin, psilocin, and related molecules, that have been developed so far for purposes of mycological research, toxicological studies, or forensic investigations are described in this chapter. The most current literature as well as some interesting older publications are cited, providing a historical overview of the subject. Methods of qualitative and quantitative analysis are presented. Additionally, preliminary purification steps and sample treatment before the exact quantitative determination are proposed. The methods described utilized various analytical techniques for the measurement of psilocybin/psilocin content in the analyzed material, including capillary electrophoresis, chemiluminescence detection systems, high-performance liquid chromatography with electrochemical, ultraviolet, or fluorescence detection systems, gas chromatography, and liquid chromatography coupled with mass spectrometry. The methods developed for simultaneous determination of several commonly abused psychoactive substances along with psilocin/psilocybin are also mentioned.
  173. K. Stebelska, “Fungal Hallucinogens Psilocin, Ibotenic Acid, and Muscimol: Analytical Methods and Biologic Activities,” Therapeutic Drug Monitoring, vol. 35, no. 4, pp. 420–442, Aug. 2013. doi: 10.1097/FTD.0b013e31828741a5.
    Psychoactive drugs of fungal origin, psilocin, ibotenic acid, and muscimol among them have been proposed for recreational use and popularized since the 1960s, XX century. Despite their well-documented neurotoxicity, they reached reputation of being safe and nonaddictive. Scientific efforts to find any medical application for these hallucinogens in psychiatry, psychotherapy, and even for religious rituals support are highly controversial. Even if they show any healing potential, their usage in psychotherapy is in some cases inadequate and may additionally harm seriously suffering patients. Hallucinogens are thought to reduce cognitive functions. However, in case of indolealkylamines, such as psilocin, some recent findings suggest their ability to improve perception and mental skills, what would motivate the consumption of “magic mushrooms.” The present article offers an opportunity to find out what are the main symptoms of intoxication with mushrooms containing psilocybin/psilocin, muscimol, and ibotenic acid. The progress in analytical methods for detection of them in fungal material, food, and body fluids is reviewed. Findings on the mechanisms of their biologic activity are summarized. Additionally, therapeutic potential of these fungal psychoactive compounds and health risk associated with their abuse are discussed.
  174. S. I. Stein, G. L. Closs, and N. W. Gabel, “Observations on Psychoneurophysiologically Significant Mushrooms,” Mycopathologia, vol. 11, no. 3, pp. 205–216, Sep. 1959. doi: 10.1007/BF02063078.
    The extracts ofP. venenosus, P. sphrinctrinus, Psil. mexicana, Psil. cubensis, andPsil. caerulescens have been compared by paper chromatography. It was found that neither Panaeoli contain psilocybin, the main constituent of the threePsilocybe species. The extract fromP. venenosus has been chromatographed, and two of the three major constituents were purified and obtained crystalline. One of these compounds could be shown to possess the same chromophore as psilocybin, a 4-oxygenated indole system, and seems most likely to be the active compound of the mushroom.
  175. T. Stijve and T. W. Kuyper, “Occurrence of Psilocybin in Various Higher Fungi from Several European Countries,” Planta Medica, vol. 51, no. 5, pp. 385–387, Oct. 1985. doi: 10.1055/s-2007-969526.
    Thieme E-Books & E-Journals
  176. T. Stijve, “Psilocin, Psilocybin, Serotonin and Urea in Panaeolus Cyanescens from Various Origin,” PERSOONIA, vol. 15, no. 1, pp. 117–121, 1992.
    The occurrenceof tryptamine derivatives and urea in Panaeolus cyanescens, also known as Copelandia cyanescens, from Australia, Hawaii and Thailand was investigated. All 70 collections contained psilocin, serotonin and urea. Those from Hawaii were also relatively rich in psilocybin, whereas the species from Australia and Thailand were virtually exempt of this compound. Many collections also contained detectable amounts of precursors as tryptophan, tryptamine and baeocystin, but 5-hydroxytryptophan — widely encountered in many other Panaeoloideae — was found to be absent in all samples. The role of these 4- and 5-hydroxylated tryptamine derivatives in the metabolism of the fungus and their possible chemotaxonomic significance is briefly discussed. Volunteers ingesting samples of Panaeolus cyanescens reported a stronger psychotropic effect than that experienced with the same amount ofPsilocybe semilanceata.
  177. T. Stijve, “Worldwide Occurrence of Psychoactive Mushrooms - an Update.,” Czech Mycology, vol. 48, no. 1, pp. 11–19, May 1995. doi: 10.33585/cmy.48103.
  178. R. Strassman, DMT: The Spirit Molecule: A Doctor’s Revolutionary Research into the Biology of Near-Death and Mystical Experiences. Rochester, Vt: Park Street Press, 2001.
  179. D. Strauss, S. Ghosh, Z. Murray, and M. Gryzenhout, “Psilocybin Containing Mushrooms: A Rapidly Developing Biotechnology Industry in the Psychiatry, Biomedical and Nutraceutical Fields,” 3 Biotech, vol. 12, no. 12, p. 339, Nov. 2022. doi: 10.1007/s13205-022-03355-4.
    Humans have collected and used hallucinogenic mushrooms for ethnic medicinal, recreational, and religious purposes since before recorded history. Currently, the use of these mushrooms is illegal in most countries, but where their use is legal they are applied as self medication. Psilocybin and psilocin, two psychoactive alkaloids, are naturally synthesized by hallucinogenic mushrooms. The chemical structure of these compounds are similar to the neurotransmitter serotonin. Activation of this system by psilocybin and psilocin may produce temporary changes in the brain that induce hallucinations and feelings of euphoria. Adjustment of the serotonin system in this way can moderate symptoms of related mental disorders. This review summarizes relevant and current information regarding the discovery of hallucinogenic mushrooms and their contained psychoactive compounds, the events that lead to their criminalization and decriminilization, and the state of knowledge of psilocybin, psilocin, and derivatives. Last, research on the psychoactive properties of these mushrooms is placed in perspective to possible applications for human dysfunctions.
  180. M. P. Torrens-Spence, C.-T. Liu, T. Pluskal, Y. K. Chung, and J.-K. Weng, “Monoamine Biosynthesis via a Noncanonical Calcium-Activatable Aromatic Amino Acid Decarboxylase in Psilocybin Mushroom,” ACS Chemical Biology, vol. 13, no. 12, pp. 3343–3353, Dec. 2018. doi: 10.1021/acschembio.8b00821.
    Aromatic l-amino acid decarboxylases (AAADs) are a phylogenetically diverse group of enzymes responsible for the decarboxylation of aromatic amino acid substrates into their corresponding aromatic arylalkylamines. AAADs have been extensively studied in mammals and plants as they catalyze the first step in the production of neurotransmitters and bioactive phytochemicals, respectively. Unlike mammals and plants, the hallucinogenic psilocybin mushroom Psilocybe cubensis reportedly employs an unrelated phosphatidylserine-decarboxylase-like enzyme to catalyze l-tryptophan decarboxylation, the first step in psilocybin biosynthesis. To explore the origin of this chemistry in psilocybin mushroom, we generated the first de novo transcriptomes of P. cubensis and investigated several putative l-tryptophan-decarboxylase-like enzymes. We report the biochemical characterization of a noncanonical AAAD from P. cubensis (PcncAAAD) that exhibits substrate permissiveness toward l-phenylalanine, l-tyrosine, and l-tryptophan, as well as chloro-tryptophan derivatives. The crystal structure of PcncAAAD revealed the presence of a unique C-terminal appendage domain featuring a novel double-β-barrel fold. This domain is required for PcncAAAD activity and regulates catalytic rate and thermal stability through calcium binding. PcncAAAD likely plays a role in psilocybin production in P. cubensis and offers a new tool for metabolic engineering of aromatic-amino-acid-derived natural products.
  181. J. M. Trappe, “The Hallucinogenic and Nonhallucinogenic Species of the Genus Psilocybe Fayod (Basidiomycotina) in Washington State, USA: New Records and a New Species,” International Journal of Medicinal Mushrooms, vol. 7, no. 4, 2005. doi: 10.1615/IntJMedMushr.v7.i4.80.
    Twenty-four species of Psilocybe are known from Washington State, USA, of which 20 have been previously reported. Psilocybe pratensis, P. semistriata,...
  182. F. Troxler, F. Seemann, and A. Hofmann, “Abwandlungsprodukte von Psilocybin Und Psilocin. 2. Mitteilung Über Synthetische Indolverbindungen,” Helvetica Chimica Acta, vol. 42, no. 6, pp. 2073–2103, 1959. doi: 10.1002/hlca.19590420638.
    Various modifications were made in the molecular structure of the natural psychotropic substances psilocybin (I) and psilocin (11) to obtain an insight into the relationship between structure and psychotropic action of this group of substances. A description is given of the synthesis and properties of the position isomers of I and II with a phosphoryloxy or hydroxy group in position 5, 6 or 7, of the four isomeric hydroxygramines, and of a series of further derivatives of 4-hydroxy-indole in which the structure of II was systematically modified, i.e. psilocin derivatives with other substitution at the ω-nitrogen; derivatives of II substituted at the indole nitrogen; psilocin derivatives with one additional methylene group in the side-chain or with a methyl-substituted or hydroxylated side-chain; phosphoric acid esters of some derivatives of II; esters of II with organic carbonic and sulfonic acids, with methylcarbaminic and with sulfuric acid; position isomers of psilocin with the dimethylaminoethyl side-chain in position 1 or 2.
  183. K. Tsujikawa et al., “Morphological and Chemical Analysis of Magic Mushrooms in Japan,” Forensic Science International, vol. 138, no. 1, pp. 85–90, Dec. 2003. doi: 10.1016/j.forsciint.2003.08.009.
    Morphological and toxicological analyses were performed on hallucinogenic mushrooms that are currently circulated in Japan. Scanning electron microscope (SEM) indicated a three-dimensional microstructures in the mushrooms. The complementary use of SEM with an optical microscope was effective for observing characteristic tissues, such as basidiomycetes, spores, cystidia and basidia. Hallucinogenic alkaloids were extracted with methanol and determined by high performance liquid chromatography (HPLC) with a UV detector set at 220nm. The psilocin/psilocybin contents in Psilocybe cubensis were in the range of 0.14–0.42%/0.37–1.30% in the whole mushroom (0.17–0.78%/0.44–1.35% in the cap and 0.09–0.30%/0.05–1.27% in the stem), respectively. The hallucinogenic alkaloids in Copelandia were 0.43–0.76%/0.08–0.22% in the whole mushroom (0.64–0.74%/0.02–0.22% in the cap and 0.31–0.78%/0.01–0.39% in the stem). It thus appears that P. cubensis is psilocybin-rich, whereas Copelandia is psilocin-rich.
  184. F. Tylš, T. Páleníček, and J. Horáček, “Psilocybin – Summary of Knowledge and New Perspectives,” European Neuropsychopharmacology, vol. 24, no. 3, pp. 342–356, Mar. 2014. doi: 10.1016/j.euroneuro.2013.12.006.
    Psilocybin, a psychoactive alkaloid contained in hallucinogenic mushrooms, is nowadays given a lot of attention in the scientific community as a research tool for modeling psychosis as well as due to its potential therapeutic effects. However, it is also a very popular and frequently abused natural hallucinogen. This review summarizes all the past and recent knowledge on psilocybin. It briefly deals with its history, discusses the pharmacokinetics and pharmacodynamics, and compares its action in humans and animals. It attempts to describe the mechanism of psychedelic effects and objectify its action using modern imaging and psychometric methods. Finally, it describes its therapeutic and abuse potential.
  185. S. E. Unger and R. G. Cooks, “Application of Mass Spectrometry/Mass Spectrometry (MS/MS) to the Identification of Natural Products in Psilocybe Cyanescens,” Analytical Letters, vol. 12, no. 11, pp. 1157–1167, Jan. 1979. doi: 10.1080/00032717908067906.
    Temperature profiling of MIKE spectra allows identification of protonated psilocin in the presence of other compounds with the same mass. The use of multiple ionizing methods to give ions (M+H)+, (M+ML4)+, and M is shown to assist in determining the molecular weights of new natural products. Structural information is obtained from the MIKE spectra. All these determinations can be made on the intact mushroom or its simple alcoholic extracts.
  186. R. C. Van Court et al., “Diversity, Biology, and History of Psilocybin-Containing Fungi: Suggestions for Research and Technological Development,” Fungal Biology, vol. 126, no. 4, pp. 308–319, Apr. 2022. doi: 10.1016/j.funbio.2022.01.003.
    Therapeutic use of psilocybin has become a focus of recent international research, with preliminary data showing promise to address a range of treatment-resistant mental health conditions. However, use of psilocybin as a healing entheogen has a long history through traditional consumption of mushrooms from the genus Psilocybe. The forthcoming adoption of new psilocybin-assisted therapeutic practices necessitates identification of preferred sources of psilocybin; consequently, comprehensive understanding of psilocybin-containing fungi is fundamental to consumer safety. Here we examine psilocybin producing fungi, discuss their biology, diversity, and ethnomycological uses. We also review recent work focused on elucidation of psilocybin biosynthetic production pathways, especially those from the genus Psilocybe, and their evolutionary history. Current research on psilocybin therapies is discussed, and recommendations for necessary future mycological research are outlined.
  187. R. Vanhaelen-Fastré and M. Vanhaelen, “Qualitative and Quantitative Determinations of Hallucinogenic Components of Psilocybe Mushrooms by Reversed-Phase High-Performance Liquid Chromatography,” Journal of Chromatography A, vol. 312, pp. 467–472, Jan. 1984. doi: 10.1016/S0021-9673(01)92800-6.
    A simple, rapid and sensitive method for the determination of psilocin and psilocybin is described. This is the first report on the determination of psilocin and psilocybin using flow injection analysis with acidic potassium permanganate and tris(2,2′-bipyridyl)ruthenium(II) chemiluminescence. The limits of detection (signal-to-noise ratio = 3) are 9 × 10−10 M and 3 × 10−10 M for psilocin and psilocybin, respectively. A concise synthetic route for psilocin in three steps from readily available starting materials is also described. The structures were elucidated on the basis of spectroscopic data. A new method has been developed for the rapid analysis of psilocybin and/or psilocin in fungus material using ion mobility spectrometry. Quantitative analysis was performed by gas chromatography–mass spectrometry after a simple one-step extraction involving homogenization of the dried fruit bodies of fungi in chloroform and derivatization with MSTFA. The proposed methods resulted in rapid procedures useful in analyzing psychotropic fungi for psilocybin and psilocin. Two modifications of the HPLC–ED method with respect to extraction procedure used have been developed for psilocin, the active metabolite of psilocybin, in human plasma using either liquid–liquid extraction (LLE) or automated on-line solid-phase extraction (on-line SPE). Each type of the sample preparation required a different HPLC system followed by electrochemical detection at 650 to 675 mV. The limit of quantitation of both modifications was 10 ng/ml psilocin. There was no significant difference observable between the LLE and the on-line SPE in terms of method standard deviation (LLE 1.82%, on-line SPE 1.13%) and the analytical results. However, the advantages of on-line SPE in addition to different selectivity were less manual effort, smaller plasma volumes of 400 μl (LLE 2 ml) and a recovery of psilocin in human plasma of nearly 100% (LLE 88%). In contrast to a previous procedure both methods were rapid, simple and reliable and yielded high plasma recoveries. They were used successfully in the quantitation of psilocin in plasma samples obtained from healthy volunteers after p.o. administration of 0.2 mg psilocybin per kg body mass. Plasma concentration curves and pharmacokinetic parameters were calculated. In order to investigate the pharmacokinetic properties of psilocybin (PY), the main psychoactive compound of Psilocybe mushrooms, high performance liquid chromatographic procedures with column-switching coupled with electrochemical detection (HPLC-ECD) for reliable quantitative determination of the PY metabolites psilocin (PI) and 4-hydroxyindole-3-acetic acid (4HIAA) in human plasma were established. Sample work-up includes protection of the highly unstable phenolic analytes with ascorbic acid, freeze-drying and in-vitro microdialysis. The data of two controlled clinical studies with healthy volunteers are presented. The subjects (N = 6 for both studies) received single oral PY doses of 0.224 ± 0.02 mg/kg b.wt. (10–20 mg) and intravenous doses of 1 mg PY, respectively. Peak plasma levels of PI after oral administration of PY were measured after 105 ± 37 min showing an average concentration of 8.2 ± 2.8 ng PI/ml plasma. 4HIAA peak concentrations of 150 ± 61 ng/ml plasma were found 113 ± 41 min after ingestion of PY. After intravenous administration, a mean PI maximum plasma concentration of 12.9 ± 5.6 ng/ml plasma was found 1.9 ± 1.0 min after injection. The maximum plasma levels appearing within a very short period indicate a rapid dephosphorylation of PY also when administered systemically. 4HIAA was not detected after 1 mg of intravenous PY. Estimates for the absolute bioavailability of PI after oral administration of PY were 52.7 ± 20% (N = 3).
  188. W.-wei Wang, “Aspects of Secondary Metabolism in Basidiomecetes: I. Biological and Biochemical Studies on Psilocybe Cubensis II. A Survey of Phenol-O-Methyltransferase in Species of Lentinus and Lentinellus,” Master's thesis, University of British Columbia, Vancouver, Canada, 1977.
    I. Psilocybe cubensis was cultured successfully in two media. Medium A was devised by Catalfomo and Tyler and Medium B was a modification of a medium which has been used for ergot alkaloid production by Claviceps purpurea. Only when the fungus was kept on Sabouraud agar plates.did it subsequently produce psilocybin when transferred to liquid media. A quantitative time-course study of psilocybin production in the two media was carried out. Maximal production appeared on the fifth day. The activities of an acid phosphatase, acting on psilocybin, were measured from mycelia grown in the two media. Enzyme activity from the A culture was very high and a blue color caused by oxidation of psilocin formed in five minutes. The effect of adding L-tryptophan on alkaloid production as well as 14 the fate of tryptophan-C was also investigated. Tryptophan stimulated significantly psilocybin production in the very beginning in the B medium. The degradation of tryptophan was different in the two media. It was converted to kynurenine and anthranilic acid in A medium and to tryptamine in tryptophan added B medium (B’ medium). Radioactive D,L-tryptophan side chain labeled, gave labeled psilocin and psilocybin. Potassium deficiency decreased psilocybin production while a potassium supplement had no effect. The fungus did not produce polyacetylenic compounds in the medium but ergosterol was detected as a major acetate derived metabolite when the fungus was kept on MYP agar plates and transferred subsequently to liquid media. Psilocin has very slight antibiotic activity against Candida albicans whereas psilocybin has none. II. Eight species of Lentinus and Lentinellus. were investigated for the occurrence of a phenol-O-methyltransferase. Only Lentinus lepideus and Lentinus pbnderbsus showed enzyme activity in both light and dark conditions. The specificity of the enzyme for a number of substrates was also examined. Of six compounds tested, methyl p-coumarate, methyl caffeate and methyl ferulate.served-as substrates. The products of enzymic activity were identified-by radioautography.
  189. R. G. Wasson, “The Hallucinogenic Mushrooms of Mexico and Psilocybin: A Bibliography (Second Printing, with Corrections and Addenda),” Botanical Museum Leaflets, Harvard University, vol. 20, no. 2a, pp. 25–73c, 1963. https://www.jstor.org/stable/41762224.
  190. T. Watanabe, Pictorial Atlas of Soil and Seed Fungi: Morphologies of Cultured Fungi and Key to Species, 2Nd ed. Boca Raton, Fla: CRC Press, 2002.
  191. E. Watkins-Dulaney, S. Straathof, and F. Arnold, “Tryptophan Synthase: Biocatalyst Extraordinaire,” ChemBioChem, vol. 22, no. 1, pp. 5–16, 2021. doi: 10.1002/cbic.202000379.
    Tryptophan synthase (TrpS) has emerged as a paragon of noncanonical amino acid (ncAA) synthesis and is an ideal biocatalyst for synthetic and biological applications. TrpS catalyzes an irreversible, C−C bond-forming reaction between indole and serine to make l-tryptophan; native TrpS complexes possess fairly broad specificity for indole analogues, but are difficult to engineer to extend substrate scope or to confer other useful properties due to allosteric constraints and their heterodimeric structure. Directed evolution freed the catalytically relevant TrpS β-subunit (TrpB) from allosteric regulation by its TrpA partner and has enabled dramatic expansion of the enzyme’s substrate scope. This review examines the long and storied career of TrpS from the perspective of its application in ncAA synthesis and biocatalytic cascades.
  192. A. T. Weil, “The Use of Psychoactive Mushrooms in the Pacific Northwest: An Ethnopharmacologic Report,” Botanical Museum Leaflets, Harvard University, vol. 25, no. 5, pp. 131–149, 1977. doi: 10.5962/p.168622.
  193. P. C. White, “Analysis of Extracts from Psilocybe Semilanceata Mushrooms by High-Pressure Liquid Chromatography,” Journal of Chromatography A, vol. 169, pp. 453–456, Feb. 1979. doi: 10.1016/0021-9673(75)85080-1.
    The retention behavior of primary, secondary and tertiary amines was studied using normal-phase-HPLC on silica, diol, and cyano stationary phases. Several classes of amines, including benzylamines, anilines, ephedrines, tryptamines, and azatryptamines were chromatographed using mixtures of hexane and ethoxynonafluorobutane with methylene chloride and methanol. Peak tailing, diminished selectivity and low plate count were minimized by the addition of volatile amines to the mobile phase. The optimal additive was n-propylamine at 0.1% concentration. On diol columns, the elution order of free primary, N-N-methyl, and N,N-dimethylamines was predictable, while the elution order of primary and secondary amines on cyano columns varied depending on the alcohol modifier concentration. The feasibility of preparative normal-phase chromatography was demonstrated by the separation of a mixture of primary, secondary and tertiary amines obtained by direct methylation of norephedrine. The procedures described may provide a practical alternative to traditional methods of analysis and purification of potential drug candidates. In order to investigate the pharmacokinetic properties of psilocybin (PY), the main psychoactive compound of Psilocybe mushrooms, high performance liquid chromatographic procedures with column-switching coupled with electrochemical detection (HPLC-ECD) for reliable quantitative determination of the PY metabolites psilocin (PI) and 4-hydroxyindole-3-acetic acid (4HIAA) in human plasma were established. Sample work-up includes protection of the highly unstable phenolic analytes with ascorbic acid, freeze-drying and in-vitro microdialysis. The data of two controlled clinical studies with healthy volunteers are presented. The subjects (N = 6 for both studies) received single oral PY doses of 0.224 ± 0.02 mg/kg b.wt. (10–20 mg) and intravenous doses of 1 mg PY, respectively. Peak plasma levels of PI after oral administration of PY were measured after 105 ± 37 min showing an average concentration of 8.2 ± 2.8 ng PI/ml plasma. 4HIAA peak concentrations of 150 ± 61 ng/ml plasma were found 113 ± 41 min after ingestion of PY. After intravenous administration, a mean PI maximum plasma concentration of 12.9 ± 5.6 ng/ml plasma was found 1.9 ± 1.0 min after injection. The maximum plasma levels appearing within a very short period indicate a rapid dephosphorylation of PY also when administered systemically. 4HIAA was not detected after 1 mg of intravenous PY. Estimates for the absolute bioavailability of PI after oral administration of PY were 52.7 ± 20% (N = 3). An efficient analytical and isolation method was elaborated for biologically active tryptamines using a computer-aided liquid chromatographic-gas chromatographic system. The separation method includes a new efficient extraction procedure, optimization programmed for high-performance liquid chromatographic separation, identification by diode-array detection and a spectrometric and electrochemical assay. The identification of indole alkaloids was confirmed by thin-layer and gas chromatography and mass spectrometry. The method was used for analysis and isolation of psychotropic substances in extracts from the fruit bodies of hallucinogenic fungi of genera Psilocybe, Inocybe and Amanita and in mycelial extracts from the species Psilocybe bohemica.
  194. P. P. Wieczorek, D. Witkowska, I. Jasicka-Misiak, A. Poliwoda, M. Oterman, and K. Zielińska, “Chapter 5 - Bioactive Alkaloids of Hallucinogenic Mushrooms,” in Studies in Natural Products Chemistry, vol. 46, Atta-ur-Rahman, Ed. Elsevier, 2015, pp. 133–168. doi: 10.1016/B978-0-444-63462-7.00005-1.
    The aim of our chapter is to review recent developments in a group of medicinally important natural products–alkaloids, with reference to the structure–activity studies in respect of certain diseases. Alkaloids covered by our review come from mushrooms called “hallucinogenic.” Hallucinogenic compounds have been chemically identified in mushrooms belonging to various genera, e.g., Agrocybe, Amanita, Conocybe, Galerina, Gymnopilus, Hypholoma, Inocybe, Panaeolus, Psilocybe, Pholiotina, Pluteus, and Weraroa [J.W. Allen, Ethnomycol. J. Sacred Mushroom Stud. 9 (2012) 130–175]. One of the largest classes of alkaloids is indole alkaloids. Indoles are probably the most widely distributed heterocyclic compounds in nature having medicinal importance [K.N. Kumar et al., Molecules 18 (2013) 6620–6662]. Two of simple indole alkaloids: psilocin (3-[2 (dimethylamino)ethyl]-4-indolol) and psilocybin ([3-(2-dimethylaminoethyl)-1H-indol-4-yl] dihydrogen phosphate) are present in most psychedelic mushrooms. Psilocin is a serotonin agonist – psilocybin/psilocin caused effects are thought to be mediated mainly by activation of 5-HT2A receptor. Ligands for the 5-HT2A receptor may be extremely useful tools for future cognitive neuroscience research [D.E. Nichols, Pharmacol. Ther. 101 (2004) 131–181]. They are also other analogs of psilocybin: baeocystin, norbaeocystin, bufotenin, and aeruginascin that are found in hallucinogenic mushrooms. Bufotenin occur also in some animal species (genus Bufo) and plants. Some of the mushroom belonging to genera Gymnopilus and Pholiota were shown to possess bisnoryangonin and hispidin, alkaloids with antimicrobial [K. Shinto et al., J. Home Econ. Jpn. 58 (9) (2007) 563–568] and antioxidant [Lee In-K et al., Mycobiology, 36 (1) (2008) 55–59] activity. Also psychoactive species of genus Amanita, contain the alkaloids (muscimol, ibotenic acid, and muscazone) that react with neurotransmitter receptors in the central nervous system. Isoxazoles, to which these alkaloids belong, have been found to inhibit voltage-gated sodium channels to control pain, enable the construction of tetracycline antibiotic derivatives, and as treatments for depression. The information within this review is intended to serve as a reference tool to better enable future research into important and fascinating area of pharmacognostic science as well as other parts of medical science.
  195. M. Wurst, M. Semerdžieva, and J. Vokoun, “Analysis of Psychotropic Compounds in Fungi of the Genus Psilocybe by Reversed-Phase High-Performance Liquid Chromatography,” Journal of Chromatography A, vol. 286, pp. 229–235, Mar. 1984. doi: 10.1016/S0021-9673(01)99190-3.
    High-performance liquid chromatography (HPLC) was used for the analysis of the minor consistuents psilocybin and psilocin in fungi of the genus Psilocybe. The separation and determination of these compounds was carried out on a stationary phase of LiChrosorb RP-18. The analytical column (A) and semipreparative column (B) were eluted isocratically with water—ethanol—acetic acid (79.2:20:0.8) at flow-rates of 20 ml/h (A) and 180 ml/h (B). The compounds were detected with a UV detector at 267 nm and a fluorometric detector (excitation, 280 nm; emission, 360 nm). The UV detection limit of psilocybin was 20–40 ng (267 nm) and several ng could be detected fluorometrically. The identity of the compounds was verified by HPLC and thin-layer chromatography and by mass spectrometry and UV spectroscopy. The compounds were determined by means of a direct calibration method and by means of the method of internal normalization. The standard deviation of the determination was ±3.4% (relative). The above methods were used to determine these compounds in extracts of fruit bodies of two species of the genus Psilocybe growing at various places in Czechoslovakia, and found to contain 0.25–1.15% of psilocybin and 0.02–0.16% psilocin per dry mass.
  196. M. Wurst, R. Kysilka, and T. Koza, “Analysis and Isolation of Indole Alkaloids of Fungi by High-Performance Liquid Chromatography,” Journal of Chromatography A, vol. 593, no. 1-2, pp. 201–208, Feb. 1992. doi: 10.1016/0021-9673(92)80287-5.
    An efficient analytical and isolation method was elaborated for biologically active tryptamines using a computer-aided liquid chromatographic-gas chromatographic system. The separation method includes a new efficient extraction procedure, optimization programme for high-performance liquid chromatographic separation, identification by diode-array detection and a spectrometric and electrochemical assay. The identification of indole alkaloids was confirmed by thin-layer and gas chromatography and mass spectrometry. The method was used for analysis and isolation of psychotropic substances in extracts from the fruit bodies of hallucinogenic fungi of genera Psilocybe, Inocybe and Amanita and in mycelial extracts from the species Psilocybe bohemica.
  197. M. Wurst, R. Kysilka, and M. Flieger, “Psychoactive Tryptamines from Basidiomycetes,” Folia Microbiologica, vol. 47, no. 1, pp. 3–27, Feb. 2002. doi: 10.1007/BF02818560.
    The review lists natural sources,i.e. strains and species of fungi producing predominantly psychoactive tryptamines (indolealkylamines), their chemical structure and properties, toxic effects on the man and psychic symptoms of intoxication. It describes the biosynthesis and production of some tryptamines by the mycelial culture ofPsilocybe bohemicaŠebek, a survey of methods for their analysis and isolation. It evaluates the worldwide use and abuse of psychoactive fungi as sources of drugs in general and in the Czechia in particular during the last two and a half decades.
  198. X. Zhang et al., “Multi-Locus Identification of Psilocybe Cubensis by High-Resolution Melting (HRM),” Forensic Sciences Research, vol. 7, no. 3, pp. 490–497, Jul. 2022. doi: 10.1080/20961790.2021.1875580.
    Hallucinogenic mushroom is a kind of toxic strain containing psychoactive tryptamine substances such as psilocybin, psilocin and ibotenic acid, etc. The mushrooms containing hallucinogenic components are various, widely distributed and lack of standard to define, which made a great challenge to identification. Traditional identification methods, such as morphology and toxicology analysis, showed shortcomings in old or processed samples, while the DNA-based identification of hallucinogenic mushrooms would allow to identify these samples due to the stability of DNA. In this paper, four primer sets are designed to target Psilocybe cubensis DNA for increasing resolution of present identification method, and the target markers include largest subunit of RNA polymerase II (marked as PC-R1), psilocybin-related phosphotransferase gene (marked as PC-PT), glyceraldehyde 3-phosphate dehydrogenase (marked as PC-3) and translation EF1α (marked as PC-EF). Real-time PCR with high-resolution melting (HRM) assay were used for the differentiation of the fragments amplified by these primer sets, which were tested for specificity, reproducibility, sensitivity, mixture analysis and multiplex PCR. It was shown that the melting temperatures of PC-R1, PC-PT, PC-3 and PC-EF of P. cubensis were (87.93 ± 0.12) °C, (82.21 ± 0.14) °C, (79.72 ± 0.12) °C and (80.11 ± 0.19) °C in our kinds of independent experiments. Significant HRM characteristic can be shown with a low concentration of 62.5 pg/µL DNA sample, and P. cubensis could be detected in mixtures with Homo sapiens or Cannabis sativa. In summary, the method of HRM analysis can quickly and specifically distinguish P. cubensis from other species, which could be utilized for forensic science, medical diagnosis and drug trafficking cases. Supplemental data for this article are available online at https://doi.org/10.1080/20961790.2021.1875580.