Next Article in Journal
Adjuvant Anti-PD-1 Immunotherapy versus Conventional Therapy for Stage III Melanoma: A Real-World Retrospective Cohort Study
Next Article in Special Issue
Isoxazole/Isoxazoline Skeleton in the Structural Modification of Natural Products: A Review
Previous Article in Journal
Exploring the Neuroprotective Mechanism of Curcumin Inhibition of Intestinal Inflammation against Parkinson’s Disease Based on the Gut-Brain Axis
Previous Article in Special Issue
Antidiabetic Potential of Novel 1,3,5-Trisubstituted-2-Thioxoimidazloidin-4-One Analogues: Insights into α-Glucosidase, α-Amylase, and Antioxidant Activities
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of Synthetic Access to Therapeutic Compounds Extracted from Psilocybe

by
Raphaël Serreau
1,2,
Ammar Amirouche
3,
Amine Benyamina
3 and
Sabine Berteina-Raboin
4,*
1
Unité de Recherche PSYCOMADD, APHP Université Paris Saclay, Hôpital Paul-Brousse, 12 Avenue Paul Vaillant Couturier, 94804 Villejuif, France
2
Addictologie EPSM Georges DAUMEZON, GHT Loiret, 1 Route de Chanteau, 45400 Fleury les Aubrais, France
3
Unité de Recherche PSYCOMADD-Psychiatrie Comorbidités Addictions, APHP Université Paris Saclay, Hôpital Paul-Brousse, 12 Avenue Paul Vaillant Couturier, 94804 Villejuif, France
4
Institut de Chimie Organique et Analytique (ICOA), Université d’Orléans, UMR-CNRS 7311, BP 6759, Rue de Chartres, CEDEX 2, 45067 Orléans, France
*
Author to whom correspondence should be addressed.
Pharmaceuticals 2023, 16(1), 40; https://doi.org/10.3390/ph16010040
Submission received: 25 November 2022 / Revised: 14 December 2022 / Accepted: 23 December 2022 / Published: 28 December 2022
(This article belongs to the Special Issue Heterocyclic Compounds and Their Application in Therapy)

Abstract

:
Psychedelics are used for various pathologies of the central nervous system and are currently the subject of much research, some of which relates to the compounds contained in various Psilocybe-type hallucinogenic mushrooms. It is difficult, however, to obtain and purify sufficient quantities of these compounds from fungi to carry out biological studies, hence the need to develop simple and efficient synthetic routes. We review here the various syntheses used to obtain these molecules, focusing first on the classic historical syntheses, then the use of more recent metallo-catalyzed couplings and finally the known biocatalytic methods for obtaining these molecules. Other access routes are certainly possible and should be the subject of future research given the therapeutic interest of these compounds.

1. Introduction

One of the areas of research on psychedelics that can be used for various pathologies concerns the compounds contained in various Psilocybe-type hallucinogenic mushrooms comprising more than 150 species of which about 80 are referenced as hallucinogens [1]. Some other mushroom species also contain these products [2,3,4]. The first two indole derivatives identified in 1958 and then synthesized in 1959 by Hofmann and co-workers were psilocybin and psilocin [5,6]. This current review was conducted following a search, using SciFinder as a database, on methods of synthesizing psilocin or psilocybin. The psilocybin prodrug is rapidly dephosphorylated in the stomach to generate psilocin, which is the active ingredient; however, the latter exhibits instability, particularly in solution [7,8,9,10]. It was shown by various researchers that the main product isolated from mushrooms of the Psilocybe genus was psilocybin, although psilocin was also present in significant quantities. Recently, however, two studies showed that, for example, for P. cubensis the quantity of the prodrug and the drug were respectively 0.63% and 0.11% of the dry residue in the first study [11] and 1.07% and 0.18% in the second [12]. Overall, the extraction of fresh mushrooms seemed to overestimate the psilocin concentrations in the mushrooms; however, as shown by the results of Hoffmeister [13], the drying process influences the dephosphorylation.
More recently, Sherwood et al. [14] mentioned that research had mainly focused on psilocin and its prodrug psilocybin [13,15,16,17,18,19,20,21,22], whereas other interesting tryptamines can also be found in lower amounts but have unfortunately not generated much research. Fricke and coworkers [23] described the enzymatic pathways of psilocybin and psilocin. They confirmed that the tryptamines possessing a hydroxy at the 4-position or their phosphorylated analogs with various degrees of N-methylation of the aminoethyl chain, known as: 4-hydroxytryptamines, norbaeocystin, baeocystin, aeruginascin and norpsilocin (Figure 1), are also enzymatic substrates.
Only a few rare studies have been carried out on the various analogs present in low quantities in certain species of mushrooms such as Psilocybe azurescens and Psilocybe cyanescens, in particular, baeocystin and its dephosphorylated analog norpsilocin. Some variable effects following recreational use have been reported during the consumption of some species of mushrooms likely dependent on the proportion of the constituents in the species [11,24].
All of these psychedelics belong to the class of psychoactive substances that act through the activation of serotonin 5-HT2A human receptors as a partial agonist [25,26]. Some of them can also act as 5HT3 agonists [27,28]. They generate psychological and physiological effects that are of interest in different pathologies. Phase 2 and 3 clinical trials [29] show favorable results in several CNS indications and the US Food and Drug Administration (FDA) has granted psilocybin Breakthrough Therapy Designation for treatment-resistant depression and the treatment of major depressive disorders [30]. These products, which are particularly interesting in the treatment of depression, need to be obtained under the best conditions in order to be able to carry out the research essential for future clinical use [31]. The number of articles indexed on the subject has grown exponentially over the past 10 years—there have been n = 1520 articles since 1958, more than 50% of which have been published over the past 10 years, showing the dynamics of research on this field.
Other indications include various disorders [32] such as cancer-related existential distress [33], substance addiction [32,34] and obsessive-compulsive disorder [35]. They have also been suggested more recently as a treatment for behavioral disorders and alcohol [36] or tobacco addiction [32,37,38].
Given the increasing prevalence of psychiatric illnesses [39] that can be treated by psychedelics, it seemed of interest to establish an inventory of the synthetic access routes of psilocin and its various derivatives.

2. Results and Discussion

2.1. Original Hofmann Synthesis and Its Several Improvements

The potential synthetic access routes are the same as those for tryptamine syntheses and yet only a few methods have been implemented to obtain these derivatives.
We will see how these psychedelics are obtained from the indole by substitution of position 3 with a view to inserting the aminoethyl chain or via the creation of the 5-membered ring from a suitably functionalized aromatic. The sequences used depend on the starting synthon and the objective of the synthesis, such as, for example, obtaining labeled molecules or not. A third synthetic methodology involves couplings catalyzed by transition metals, either iridium or palladium.
Among the natural tryptamines treated, the target molecules are listed in Figure 1 and differ only by the degree of methylation of the aminoethyl chain of psilocybin and psilocin.
We will also be interested in this review in two analogs: psilacetin (4-acetoxy-N,N-dimethyltryptamines), which is the acetylated analog of psilocybin and can also be hydrolyzed to active psilocin; and the 4-acetoxy-N,N,N-trimethyltryptammonium (4-OAc-TMT) analog of aeruginascin (Figure 2) [40].
The initial methodology, which is also the most well-documented for obtaining psilocybin, consists in inserting the aminoethyl chain in position 3 of an indole core structure.
Hofmann et al. [5,6,41] developed one of the first syntheses of psilocybin for Sandoz, following the process described in Scheme 1. The 4-hydroxyindole protected by a benzyl group 1 was used as starting material and treated with oxalyl chloride followed by dimethylamine to give the expected stable indole 2, which was reduced with lithium aluminum hydride to furnish the desired dimethyl aminoethyl side chain 3. Deprotection of the hydroxyl function using hydrogen with palladium on carbon gave the active psilocin 4. The prodrug 6 was synthesized by phosphorylation with O,O-dibenzylphosphorylchloride leading to 5 which underwent hydrogenation on palladium on carbon.
This first synthesis has since been modified by different chemists who have brought many improvements in the yield and purity of the final compound [42,43].
Nichols et al. [2] developed an improved procedure using a different phosphorylation protocol. They accomplished this step with, what they declared was, the most suitable reagent: tetra-benzyl-pyrophosphate which is a stable crystalline reagent compared to the O,O-dibenzyl phosphoryl chloride used by Hofmann and that led to only 20% yield. With these new conditions, they obtained Psilocybin in 46.9% yield. However, this phosphorylation step was difficult because of the lability of the O,O-dibenzyl ester. Nichols et al. showed that a hydrolytic cleavage of one O-benzyl group occurred at room temperature in the presence of water and the purification of this compound 7 was difficult (Scheme 2).
To overcome these phosphorylation problems, Nichols et al. [42] decided to synthesize the 4-acetoxy-N,N-dimethyltryptamine by hydrogenation of compound 3 with Pd/C in presence of acetic anhydride and sodium acetate in benzene; then, after filtration of the catalyst and some insoluble salts, fumaric acid was added. This prodrug is much easier to synthesize and uses atom-economical reagents. Psilacetin salt has the advantage of being easily crystallizable (Scheme 3). Psilacetin is, moreover, more stable than psilocybin.
Recently, in 2020, Manke et al. [40] reported the first synthesis of a metabolite of aeruginascin, an active compound found in some psilocybes such as Inocybe aeruginascens [13,14,21]. This mushroom was described as a source of hallucinations, causing only a euphoric effect without the dreaded bad trip or dysphoric experience [20] that other mushrooms can generate. Unfortunately, the activity of aeruginascin has not been extensively explored [21]. Aeruginascin can also undergo hydrolysis to lead to its active metabolite 4-HO-TMT (4-hydroxy-N,N,N-trimethyltryptamine). The synthesis of this derivative was achieved from psilacetin in the form of the fumarate salt, which was methylated with an excess of iodomethane in methanol to provide 4-acetoxy-N,N,N-trimethyltryptammonium in 53% yield and then hydrolyzed in the presence of acetic acid in water (Scheme 4).
These compounds (4-AcO-TMT and 4-OH-TMT) were recrystallized from water, for the first time, and characterized by single-crystal diffraction. This easy-to-synthesize active metabolite of aeruginascin showed binding at three receptors: 5-HT1A, 5-HT2A and 5-HT2B and appeared to be as interesting as psilocybin, according to the authors.
Shirota et al. [43] developed the synthesis of psilocin and psilocybin from the 4-hydroxy-indole at the gram scale without the need for chromatographic purification (Scheme 5). They obtained 85% yield and notably showed the spontaneous intramolecular migration of one of the benzyl groups of compound 5 to form the zwitterionic N,O-dibenzyl phosphate derivative 12. This rearrangement product was completely characterized by 2D NMR analyses, HMBC and NOESY correlations [43].
More recently, Sherwood et al. [44] described an improved, practical and scalable synthesis according to the Speeter–Antony tryptamine synthetic pathway [45] previously used. As these conditions did not allow them to obtain the same yields, they adapted Shirota’s conditions to obtain psilocybin with high purity (99.9%) by simple trituration and filtration in various solvents as purification processes. The low solubility of the rearrangement product allowed its isolation by filtration.
Using commercially available 4-acetoxyindole 9, acylation in the presence of oxalyl chloride in ethyl ether requires careful treatment to remove traces of residual oxalyl chloride, which generates the decomposition of the chloro-compound intermediate. Washing this intermediate with heptane makes it possible to obtain a relatively stable compound 10, as a yellow solid, which can be kept for a few days. However, the reaction with dimethylamine can also be launched directly at the end of the treatment in the presence of dimethylamine in solution in THF. This protocol gave the desired product in 80% yield with a purity of 96%, which can be increased to >99% by recrystallization in isopropyl alcohol.
The reduction of compound 11 with LiAlH4 in THF at reflux temperature led to a mixture of compounds (Scheme 6) with the β-hydroxy derivative 13, which can be limited by a prolonged refluxing time, but in this case, decomposition occurred. 2-MeTHF was used because of its higher boiling point and allowed the authors to obtain psilocin 4 in 77% yield and 99.2% HPLC purity after an adapted treatment to remove an excess of LiAlH4. Psilocin was described as unstable and partially broken down into colored derivatives ranging from green to purple over time due to oxidation issues, especially if the psilocin was not obtained pure or not stored in a solid state.
The phosphorylation of psilocin via the rearrangement protocol leading to the zwitterionic intermediate 12 (Scheme 5) was then also implemented. By varying the reaction temperature and the washing solvents, Sherwood et al. [44] managed to obtain 12 pure in 63% yield. The catalytic hydrogenation of this zwitterionic compound with 10% Pd/C at atmospheric pressure in methanol gave rapid access to psilocybin in the form of a purple solid. A study of the most appropriate solvents for the purification of psilocybin by recrystallization was carried out and the product of interest was obtained pure after three successive recrystallizations with a purity of 99.9% determined by HPLC. The first recrystallization was carried out from acetone containing 30% water, the second from water containing 30% acetone and the last from deionized water. Scaling up from 50 g to 3.3 kg of starting material with some solvent modifications generated a slight decrease in yields at each step (Scheme 7). This level of purity has allowed its use in clinical trials.
A second-generation synthesis was also developed on a kilogram-scale production by the same team [46] and provided an alternative solution to the use of very expensive reagents such as TBPP tetra-benzyl pyrophosphate, which was replaced, for the first time, by phosphorus oxychloride. This new method of synthesis is a more atom-efficient process for the production of around one kilogram of psilocybin with an overall yield equivalent to that previously obtained on a smaller scale (Scheme 8).
However, the use of POCl3 instead of TBPP generates several di-phosphorylated intermediates after aqueous quenching and hot recrystallizations from water or methanol can cause the hydrolysis of psilocybin to psilocin. This phosphorylation step remains the most problematic step of the synthesis.

2.2. Synthesis of Isotopically Labeled Psilocin and Other Psilocin Synthesis

In 1955, Hofmann et al. [47] described a new synthesis of bufotenin or 5-hydroxy N,N-dimethyltryptamine and related 4-oxy-tryptamines according to Scheme 9. In addition to bufotenin, compounds (still unknown in 1955) were also synthesized by analogy. This is how two positional isomers of serotonin, 4- and 6-oxytryptamine, were described based on the serotonin synthesis of Hamlin and Fischer [48]. The N,N-dimethylated 4-oxytryptamine called psilocin was extracted from Psilocybe three years later by the same researchers.
This synthetic method was also used in 1962 by Kalberer et al. [9] using labeled K14CN instead of NaCN (Scheme 10) to obtain the 14C-labeled psilocin 29. The previous steps were similar to those previously. This procedure was improved by Poon et al. in 1985 [49] (Scheme 11). Poon also described one procedure applied to the synthesis of 3H-labeled psilocin (Scheme 12).
Poon used tris (dimethylamino)methane to generate a non-isolated product that afforded semicarbazone by treatment with semicarbazide hydrochloride by a procedure already described by Kruse [50]. After cyclization in 4-benzyloxyindole 19 then formylation and amination, the 4-benzyloxygramine 24 was heated in presence of K14CN at reflux temperature leading to the expected labeled compound after acidification of the solution. Using phosphorus pentachloride, dimethylamine then LiAlH4 as a reductive agent, followed by deprotection of the benzyl group, the 14C-psilocin 29 was synthesized in a shorter time and a better yield, 42% yield from 26 compared to Kalberer’s protocol < 10% yield in their hands (Scheme 11).
The synthesis of 3H-labeled psilocin 34 was obtained by treatment with oxalyl chloride followed by the addition of dimethylamine. The use of labeled lithium aluminum hydride (3H-LiAlH4) followed by debenzylation by hydrogenolysis led to the expected compound in good yields (Scheme 12).
Repke [51,52,53] also investigated the synthesis of these compounds which he pharmacomodulated at the level of the aminoethyl chain (Scheme 13) or by transforming indole into various β-carbolines (Scheme 14). He obtained indole precursors of β-carbolines starting from 2-nitro-6-alkoxy-toluene via the Leimgruber and Batcho method [54]. The tryptamines were synthesized from indole in presence of 1,1-dimethylaminonitroethylene followed by reduction with LiAlH4 then a Pictet–Spengler reaction with glyoxylic or pyruvic acids [55] afforded the corresponding β-carbolines.
In 1998, Somei et al. [56] described a five-step synthesis of psilocin 4 starting from indole-3-carbaldehyde 35 according to the following Scheme 15. In 2002, they applied their strategy to the synthesis of various analogs bearing a formyl or bromine substituent in the 5- or 7-position [57].

2.3. Metallo-Catalyzed Psilocin Synthesis

Metal-catalyzed reactions require catalytic quantities of metal and can be used with a wide range of functional groups, explaining the widespread use of these transformations, especially in indole synthesis. However, only a few examples have been described for tryptamine synthesis. It was in 1999 that Ogasawara reported a new synthesis of psilocin 4 from N-tert-butoxycarbonyl-3-methoxyaniline 38 prepared from commercially available 3-methoxyaniline [58]. After iodination of this compound to afford 39 in 46% yield, the team was the first to involve a metallo-catalyzed Heck-type coupling with palladium acetate using 2,5-dihydro-2,5-dimethoxyfuran 40 as the double-bond component. After refluxing in trifluoroacetic acid and deprotection of the Boc N-protective group, the authors used the Weinreb conditions [59] to transform the methyl ester 43 into N,N-dimethylamide 44. A reduction of the amide with LiAlH4 followed by deprotection of the methyl alcohol substituent using boron tribromide afforded psilocin 4 in 5.5% overall yield (Scheme 16).
Scammells et al. [60] reported a short preparation of the indole core structure directly via palladium-catalyzed cyclization between a 4-hydroxy ortho iodoaniline and a silylated alkyne bearing the dimethylaminoethyl chain 46. Involving the Larock method [61] where the palladium was inserted on the less hindered site of the alkyne, they were able to generate the desired 3-substituted indole after cyclization 47. This compound bearing the trimethylsilyl group in the 2-position was deprotected in the presence of trifluoroacetic acid. The third step consisted in removing the methyl group on an alcohol substituent using boron tribromide. Using this protocol, psilocin 4 was synthesized in 24% overall yield (Scheme 17). This synthetic methodology requires the prior synthesis of the substrates used. Thus the alkyne 46 is obtained from 3-butyn-1-ol, in 88% total yield, by a procedure already described [62]. Tosylation followed by N,N-dimethylation by nucleophilic substitution followed by lithiation of alkyne, and reaction with trimethylsilyl chloride led to the suitably functionalized alkyne 46. The N-tert-butoxycarbonyl-2-iodo-3-methoxy-aniline 39 was obtained in three steps from 3-methoxy-aniline protected with a tert-butoxycarbonyl ortho-director group followed by a lithiation then consecutive iodination. The expected total regioselectivity of palladium cross-coupling was confirmed by X-ray crystallography.
Psilocin was also obtained on a small scale (only 5.6 mg) from 2-iodo-3-methoxy-aniline 48 and phthalimide-protected 4-aminobutanal 49 in presence of palladium acetate and DABCO (1,4-diazabicyclo [2,2,2]octane) in DMF at 85 °C. Jia et al. [63] investigated the scope and limitations of this palladium coupling with various substrates (Scheme 18). Two nitrogen-protecting groups of 4-aminobutanal were evaluated (Boc and Phth) and synthesized from the commercially available corresponding aminoketal. The method was applied to various natural products. For psilocin synthesis, they used the phthalimido-protected 4-aminobutanal 49 with the 2-iodo-3-methoxy-aniline 48 to lead to the indole core structure 50 whose phthalimido group was deprotected in the presence of hydrazine. The resulting amine 51 was dimethylated, then a final deprotection with boron tribromide provided access to the target molecule 4 in 17% overall yield (Scheme 18).
Another alternative is to start from an indole and introduce an alkyl chain in position 3. This is difficult because of the few electrophilic reagents giving access to the aminoethyl chain, whether protected or not.
Bartlucci et al. [64] reported in 2016 an interesting selective iridium-catalyzed alkylation in position C3 of the indole heterocycle with N-protected ethanolamine via a borrowing hydrogen procedure. Their first tryptamine examples were successfully obtained after 48 h at 150 °C in presence of [Cp*IrCl2]2 as catalyst and Cs2CO3 as base. However, it should be mentioned that some alcohols gave complex mixtures. They described the use of commercially available N,N-dimethylethanolamine 52 to synthesize psilocin, bufotenin and serotonin in two steps from the 4-benzyl-protected indole 19 or 5-benzyl-protected indole followed by hydrogenolysis of benzyl with NH4HCO2 and 10% Pd/C in 61%, 67% and 62% yields, respectively (Scheme 19).
These coupling reactions catalyzed by transition metals are very interesting to obtain small quantities of products for analysis but are more difficult to use on a large scale for production due to the availability and high cost of such catalysts.

2.4. Biocatalytic Route

Given, on the one hand, the increasing demand for these active ingredients for use in clinical trials and, on the other, the difficulties encountered in the known synthetic routes described above, Hoffmeister et al. sought to develop enzymatic access to these drugs. As the kinase gene psiK had been identified in Psilocybe mushrooms, they used it for the biocatalytic synthesis of psilocin in vitro [23,65,66,67] but also in vivo in microbial hosts [68]. In this case, the compounds must be extracted from complex matrices to eliminate cellular dishes with cell lysate or broth. In order to avoid all these problems, Hoefgen et al. and Adams et al. developed a combined synthesis involving the chemical pathway to biosynthesis according to Scheme 20 [68,69]. This new synthetic/biocatalytic route makes it possible to synthesize the target molecule on a large scale. In these conditions, 893 mg of psilocybin was obtained from psilocin in 88.5% yield [65]. This process eliminates steps that may involve heavy metals and the difficult phosphorylation step that requires expensive reagents. They first produced PsiK using E. coli KRXVpJF23 and studied its properties in vitro. Subsequent work confirming the dual biosynthetic and protective role of PsiK, which accepts 4-hydroxytryptamine but also psilocin as substrates, allowed them to use the repair reaction for their study. In Psilocybe mushrooms, an S-adenosyl-L-methionine (SAM)-dependent N-methyltransferase acts twice after the phosphorylation step to complete the biosynthesis of psilocybin from 4-hydroxytryptamine.
These recent results should promote studies on bioprocesses in order to produce this drug but also other drugs at a reasonable cost and in a greener way. Synthesizing in an eco-compatible manner, new, more stable derivatives that retain the same bioavailability would be a major advance in the field.

3. Conclusions

Efficient methods are now available to achieve the synthesis of the psilocybin prodrug, which is usually synthesized because psilocin is described as unstable in solution. However, other prodrugs are possible and easier to synthesize, such as psilacetin, for example. Several studies report the use of other protections and are the subject of recently filed but not yet published patents. Other synthetic methods to access tryptamine analogs have not yet been implemented for these drug candidates due to the lack of commercial accessibility to the starting synthons or due to their difficulty of access; in particular, other metal-catalyzed couplings, although the use of metals in the synthesis route should only be used at the start of the synthesis to avoid the associated toxicity problems. Indeed, some starting building blocks cannot be obtained in a regioselectively pure form, thus requiring purifications that are sometimes difficult and very atom-consuming. The biocatalytic route is a very interesting alternative. However, it is important that further studies be conducted on the accessibility of these compounds. Currently, there is a strong interest in psychedelics. Clinical studies that target various pathologies are underway and should lead to the medical use of some of these substances.

Author Contributions

Conceptualization, R.S. and S.B.-R.; methodology, R.S. and S.B.-R.; validation, R.S., A.A., A.B. and S.B.-R.; formal analysis, R.S. and S.B.-R.; investigation, R.S., A.A., A.B. and S.B.-R.; writing—original draft preparation, S.B.-R.; writing—review and editing, R.S., A.A., A.B. and S.B.-R.; supervision, R.S. and S.B.-R.; project administration, R.S., A.A., A.B. and S.B.-R. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gartz, J. Further investigations on psychoactive mushrooms of the genera Psilocybe, Gymnopilus and Conocybe. Ann. Musei Civ. Rovereto 1992, 7, 265–274. [Google Scholar]
  2. Stijve, T.; Kuyper, T.W. Occurrence of psilocybin in various higher fungi from several European countries. Planta Med. 1985, 51, 385–387. [Google Scholar] [CrossRef] [PubMed]
  3. Semerdzieva, M.; Wurst, M.; Koza, T.; Gartz, J. Psilocybin in fruiting-bodies of Inocybe aeruginascens. Planta Med. 1986, 52, 83–85. [Google Scholar] [CrossRef]
  4. Gartz, J. Detection of tryptamines derivatives in fungi of the genera Gerronema, Hygrocybe, Psathyrella and Inocybe. Biochem. Physiol. Pflanz. 1986, 181, 275–278. [Google Scholar] [CrossRef]
  5. Hofmann, A.; Frey, A.; Ott, H.; Petrzilka, T.; Troxler, F. The structure and synthesis of psilocybin. Experientia 1958, 14, 397–399. [Google Scholar] [CrossRef] [PubMed]
  6. Hofmann, A.; Troxler, F. Identification of psilocin. Experientia 1959, 15, 101–102. [Google Scholar] [CrossRef] [PubMed]
  7. Horita, A.; Weber, L.J. The enzymatic dephosphorylation and oxidation of psilocybin and psilocin by mammalian tissue homogenates. Biochem. Pharmacol. 1961, 7, 47–54. [Google Scholar] [CrossRef]
  8. Horita, A.; Weber, L.J. Dephosphorylation of psilocybin in the intact mouse. Toxicol. Appl. Pharmacol. 1962, 4, 730–737. [Google Scholar] [CrossRef]
  9. Kalberer, F.; Kreis, W.; Rutschmann, J. The fate of psilocin in the rat. Biochem. Pharmacol. 1962, 11, 261–269. [Google Scholar] [CrossRef]
  10. Hasler, F.; Bourquin, D.; Brenneisen, R.; Bär, T.; Vollenweider, F.X. Determination of psilocin and 4-hydroxyindole-3-acetic acid in plasma by HPLC-ECD and pharmacokinetic profiles of oral and intravenous psilocybin in man. Pharm. Acta Helv. 1997, 72, 175–184. [Google Scholar] [CrossRef]
  11. Gartz, J. Extraction and analysis of indole derivatives from fungal biomass. J. Basic Microbiol. 1994, 34, 17–22. [Google Scholar] [CrossRef] [PubMed]
  12. Borner, S.; Brenneisen, R. Determination of tryptamine derivatives in hallucinogenic mushrooms using high-performance liquid chromatography with photodiode array detection. J. Chromatograph. 1987, 408, 402–408. [Google Scholar] [CrossRef] [PubMed]
  13. Lenz, C.; Wick, J.; Hoffmeister, D. Identification of ω-N-Methyl-4-hydroxytryptamine (Norpsilocin) as a Psilocybe Natural Product. J. Nat. Prod. 2017, 80, 2835–2838. [Google Scholar] [CrossRef] [PubMed]
  14. Sherwood, A.M.; Halberstadt, A.L.; Klein, A.K.; McCorvy, J.D.; Kaylo, K.W.; Kargbo, R.B.; Meisenheimer, P. Synthesis and biological evaluation of tryptamines found in hallucinogenic mushrooms: Norbaeocystin, baeocystin, norpsilocin and aeruginascin. J. Nat. Prod. 2020, 83, 461–467. [Google Scholar] [CrossRef]
  15. Repke, D.B.; Leslie, D.T.; Guzman, G. Baeocystin in Psilocybe, Conocybe and Panaeolus. Lloyda 1977, 40, 566–578. [Google Scholar]
  16. Repke, D.B.; Leslie, D.T. Baeocystin in Psilocybe semilanceata. J. Pharm. Sci. 1977, 66, 113–114. [Google Scholar] [CrossRef]
  17. Leung, A.; Paul, A. Baeocystin and norbaeocystin: New analogs of psilocybin from Psilocybe baeocystis. J. Pharm. Sci. 1968, 57, 1667–1671. [Google Scholar] [CrossRef]
  18. Stijve, T.; Klan, J.; Kuyper, T.W. Occurrence of psilocybin and baeocystin in the genus Inocybe. Persoonia 1985, 12, 469–473. [Google Scholar]
  19. Gartz, J. Occurrence of psilocybin, psilocin and baeocystin in Gymnopilus purpuratus. Persoonia 1989, 14, 19–22. [Google Scholar]
  20. Gartz, J. Analysis of aeruginascin in fruit bodies of the mushroom Inocybe aeruginascens. Int. J. Crude Drug Res. 1989, 27, 141–144. [Google Scholar] [CrossRef]
  21. Jensen, N.; Gartz, J.; Laatsch, H. Aeruginascin, a trimethylammonium analogue of psilocybin from the hallucinogenic mushroom Inocybe aeruginascens. Planta Med. 2006, 72, 665–666. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Gartz, J. Variation der Alkaloidmengen in Fruchtkörpern von Inocybe aeruginascens. Planta Med. 1987, 53, 539–541. [Google Scholar] [CrossRef] [PubMed]
  23. Fricke, J.; Blei, F.; Hoffmeister, D. Enzymatic synthesis of psilocybin. Angew. Chem. Int. Ed. 2017, 56, 12352–12355. [Google Scholar] [CrossRef] [PubMed]
  24. Gartz, J. Magic Mushrooms Around the World: A Scientific Journey Across Cultures and Time; LIS Publishers: Los Angeles, CA, USA, 1996. [Google Scholar]
  25. Geiger, H.A.; Wurst, M.G.; Daniels, R.N. DARK Classics in Chemical Neuroscience: Psilocybin. ACS Chem. Neurosci. 2018, 9, 2438–2447. [Google Scholar] [CrossRef] [PubMed]
  26. Dinis-Oliveira, R.J. Metabolism of psilocybin and psilocin: Clinical and forensic toxicological relevance. Drug Metab. Rev. 2017, 49, 84–91. [Google Scholar] [CrossRef]
  27. Pereira, N.A.; Marins, J.C.; Moussatché, H. Some pharmacological studies on bufotenine and bufotenidine. Rev. Bras. Biol. 1963, 23, 211–222. [Google Scholar]
  28. Glennon, R.; Peroutka, S.; Dukat, M. Binding characteristics of a quaternary amine analog of serotonin 5-HT Q. In Serotonin: Molecular Biology, Receptors and Functional Effects; Fozard, J.R., Saxena, P.R., Eds.; Birkhaüser: Basel, Switzerland, 1991; pp. 186–191. [Google Scholar]
  29. Bogenschutz, M.P.; Ross, S.; Bhatt, S.; Baron, T.; Forcehimes, A.A.; Laska, E.; Mennenga, S.E.; O’Donnell, K.; Owens, L.T.; Podrebarac, S.; et al. Percentage of Heavy Drinking Days Following Psilocybin-Assisted Psychotherapy vs. Placebo in the Treatment of Adult Patients with Alcohol Use Disorder: A Randomized Clinical Trial. JAMA Psychiatry 2022, 79, 953–962. [Google Scholar] [CrossRef]
  30. Feltman, R. The FDA Is Fast Tracking a Second Psilocybin Drug to Treat Depression; Popular Sciences. 2019. Available online: https://popsci.com/story/health/psilocybin-magic-mushroom-fda-breakthrough-depression/ (accessed on 26 November 1999).
  31. Nichols, D.E.; Johnson, M.W.; Nichols, C.D. Psychedelics as Medicines: An Emerging New Paradigm. Clin. Pharmacol. Ther. 2017, 101, 209–219. [Google Scholar] [CrossRef]
  32. Johnson, M.W.; Griffiths, R.R. Potential Therapeutic Effects of Psilocybin. Neurotherapeutics 2017, 14, 734–740. [Google Scholar] [CrossRef] [Green Version]
  33. Carhart-Harris, R.L.; Goodwin, G.M. The therapeutic potential of psychedelic drugs: Past, present, and future. Neuropsychopharmacology 2017, 42, 2105–2113. [Google Scholar] [CrossRef] [Green Version]
  34. Summergrad, P. Psilocybin in end of life care: Implications for further research. J. Psychopharmacol. 2016, 30, 1203–1204. [Google Scholar] [CrossRef] [PubMed]
  35. Moreno, F.A.; Wiegand, C.B.; Taitano, E.K.; Delgado, P.L. Safety, tolerability, and efficacy of psilocybin in 9 patients with obsessive-compulsive disorder. J. Clin. Psychiatry 2006, 67, 1735–1740. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Bogenschutz, M.P.; Forcehimes, A.A.; Pommy, J.A.; Wilcox, C.E.; Barbosa, P.; Strassman, R.J. Psilocybin-assisted treatment for alcohol dependence: A proof-of-concept study. J. Psychopharmacol. 2015, 29, 289–299. [Google Scholar] [CrossRef] [PubMed]
  37. Johnson, M.W.; Garcia-Romeu, A.; Cosimano, M.P.; Griffiths, R.R. Pilot study of the 5-HT2AR agonist psilocybin in the treatment of tobacco addiction. J. Psychopharmacol. 2014, 28, 983–992. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Garcia-Romeu, A.; Griffiths, R.; Johnson, M. Psilocybin-occasioned mystical experiences in the treatment of tobacco addiction. Curr. Drug Abuse Rev. 2015, 7, 157–164. [Google Scholar] [CrossRef] [PubMed]
  39. Giordana, J.Y.; Porteaux, C. Mental Health of elderly people: The prevalence and representations of psychiatric disorders. L’éncephale 2010, 36, 59–64. [Google Scholar] [CrossRef] [PubMed]
  40. Chadeayne, A.R.; Pham, D.N.K.; Reid, B.G.; Golen, J.A.; Manke, D.R. Active metabolite of aeruginascin (4-hydroxy-N,N,N-trimethylamine): Synthesis, structure and serotoninergic binding affinity. ACS Omega 2020, 5, 16940–16943. [Google Scholar] [CrossRef]
  41. Troxler, F.; Seemann, F.; Hofmann, L.A. Abwandlungsprodukte von Psilocybin und Psilocin. 2. Mitteilung über synthetische Indolverbindungen. Helv. Chim. Acta 1959, 42, 2073–2103. [Google Scholar] [CrossRef]
  42. Nichols, D.; Frescas, S. Improvements to the Synthesis of Psilocybin and a Facile Method for Preparing the O-Acetyl Prodrug of Psilocin. Synthesis 1999, 6, 935–938. [Google Scholar] [CrossRef] [Green Version]
  43. Shirota, O.; Hakamata, W.; Goda, Y. Concise Large-Scale Synthesis of Psilocin and Psilocybin, Principal Hallucino- genic Constituents of “Magic Mushroom”. J. Nat. Prod. 2003, 66, 885–887. [Google Scholar] [CrossRef] [Green Version]
  44. Sherwood, A.M.; Meisenheimer, P.; Tarpley, G.; Kargbo, R.B. An Improved, Practical, and Scalable Five-Step Synthesis of Psilocybin. Synthesis 2020, 52, 688–694. [Google Scholar] [CrossRef]
  45. Speeter, M.E.; Anthony, W.C. The action of oxalyl chloride on indoles: A new approach to tryptamines. J. Am. Chem. Soc. 1954, 76, 6208–6210. [Google Scholar] [CrossRef]
  46. Kargbo, R.B.; Sherwood, A.M.; Walker, A.; Cozzi, N.V.; Dagger, R.E.; Sable, J.; O’Hern, K.; Kaylo, K.; Patterson, T.; Tarpley, G.; et al. Direct phosphorylation of psilocin enables optimized cGMP kilogram-scale manufacture od psilocybin. ACS Omega 2020, 5, 16959–16966. [Google Scholar] [CrossRef]
  47. Stoll, A.; Troxler, F.; Peyer, J.; Hofmann, A. Eine neue synthese von Bufotenin und verwandten Oxy-tryptaminen. Helv. Chim. Acta 1955, 38, 1452–1472. [Google Scholar] [CrossRef]
  48. Hamlin, K.E.; Fischer, F.E. The synthesis of 5-hydroxytryptamine. J. Am. Chem. Soc. 1951, 73, 5007–5008. [Google Scholar] [CrossRef]
  49. Poon, G.; Chui, Y.C.; Law, F.C.P. Synthesis of psilocin labelled with 14C and 3H. J. Label. Compd. Radiopharm. 1985, 23, 167–174. [Google Scholar] [CrossRef]
  50. Kruse, L.I. Synthesis of 4-Substituted Indoles from o-Nitrotoluenes. Heterocycles 1981, 16, 1119–1124. [Google Scholar] [CrossRef]
  51. Repke, D.B.; Ferguson, W.J.; Bates, D.K. Psilocin Analogs. I. Synthesis of 3[2-(dialkylamino(ethyl]- and 3[2-(cycloalkylamino(ethyl]-indol-4-ols. J. Heterocyclic. Chem. 1977, 14, 71–74. [Google Scholar] [CrossRef]
  52. Repke, D.B.; Ferguson, W.J.; Bates, D.K. Psilocin Analogs. II. Synthesis of 3[2-(dialkylamino(ethyl]-, 3[2-(N-methyl-N-alkylamino(ethyl]- and 3[2-(cycloalkylamino(ethyl]-indol-4-ols. J. Heterocyclic. Chem. 1981, 18, 175–179. [Google Scholar] [CrossRef]
  53. Repke, D.B.; Ferguson, W.J. Psilocin Analogs. III. Synthesis of 5-Methoxy- and 5-Hydroxy-1,2,3,4-tetrahydro-9H-pyrido[3,4-b]indoles. J. Heterocyclic. Chem. 1982, 19, 845–848. [Google Scholar] [CrossRef]
  54. Leimgruber, W.; Batcho, A.D. Batcho-Leimgruber indole synthesis. In Name Reactions; Springer: Berlin/Heidelberg, Germany, 2006; pp. 36–38. [Google Scholar] [CrossRef]
  55. Spenser, I.D. A synthesis of Harmaline. Can. J. Chem. 1959, 37, 1851–1858. [Google Scholar] [CrossRef]
  56. Yamada, F.; Tamura, M.; Somei, M. A five-step synthesis of psilocin from indole-3-carbaldehyde. Heterocycles 1998, 49, 451–457. [Google Scholar] [CrossRef] [Green Version]
  57. Yamada, F.; Tamura, M.; Hasegawa, A.; Somei, M. Synthetic Studies of Psilocin Analogs Having Either a Formyl Group or Bromine Atom at the 5- or 7-Position. Chem. Pharm. Bull. 2002, 50, 92–99. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Sakagami, H.; Ogasawara, K. A New Synthesis of Psilocin. Heterocycles 1999, 51, 1131–1135. [Google Scholar] [CrossRef]
  59. Basha, A.; Lipton, M.; Weinreb, S.M. A mild, general method for conversion of esters to amides. Tetrahedron Lett. 1977, 18, 4171–4172. [Google Scholar] [CrossRef]
  60. Gathergood, N.; Scammells, P.J. Preparation of 4-hydroxytryptamine scaffold via palladium-catalyzed cyclisation: Apractical and versatile synthesis of psilocin. Org. Lett. 2003, 5, 921–923. [Google Scholar] [CrossRef]
  61. Larock, R.C.; Yum, E.K. Synthesis of indoles via palladium-catalyzed heteroannulation of internal alkynes. J. Am. Chem. Soc. 1991, 113, 6689–6690. [Google Scholar] [CrossRef]
  62. Smith, A.L. Traceless Solid Phase Synthesis of Indole Derivatives. British UK Patent Applications GB2328941A, 3 October 1999. [Google Scholar]
  63. Hu, C.; Qin, H.; Cui, C.; Jia, Y. Palladium-catalyzed synthesis of tryptamines and tryptamine homologues: Synthesis of psilocin. Tetrahedron 2009, 65, 9075–9080. [Google Scholar] [CrossRef]
  64. Bartolucci, S.; Mari, M.; Di Gregorio, G.; Piersanti, G. Observations concerning the synthesis of tryptamine homologues and branched tryptamine derivatives via the borrowing hydrogen process: Synthesis of psilocin, bufotenin and serotonin. Tetrahedron 2016, 72, 2233–2238. [Google Scholar] [CrossRef] [Green Version]
  65. Fricke, J.; Kargbo, R.; Regestein, L.; Lenz, C.; Peschel, G.; Rosenbaum, M.A.; Sherwood, A.; Hoffmeister, D. Scalable hybrid synthetic/biocatalytic route to psilocybin. Chem. Eur. J. 2020, 26, 8281–8285. [Google Scholar] [CrossRef]
  66. Fricke, J.; Lenz, C.; Wick, J.; Blei, D.; Hoffmeister, D. Production options for Psilocybin: Making of the magic. Chem. Eur. J 2019, 25, 897–903. [Google Scholar] [CrossRef] [PubMed]
  67. Blei, D.; Baldeweg, F.; Fricke, J.; Hoffmeister, D. Biocatalytic production of psilocybin and derivatives in tryptophan synthase-enhanced reactions. Chem. Eur. J. 2018, 24, 10028–10031. [Google Scholar] [CrossRef] [PubMed]
  68. Hoefgen, S.; Lin, J.; Fricke, J.; Stroe, M.; Mattern, D.J.; Kufs, J.E.; Hortschansky, P.; Brakhage, A.A.; Hoffmeister, D.; Valiante, V. Facile assembly and fluorescence-based screening method for heterologous expression of biosynthetic pathways in fungi. Metab. Eng. 2018, 48, 44–51. [Google Scholar] [CrossRef] [PubMed]
  69. Adams, A.M.; Kaplan, N.A.; Wei, Z.; Brinton, J.D.; Monnier, C.S.; Enacopol, A.L.; Ramelot, T.A.; Jones, J.A. In vivo production of psilocybin in E. coli. Metab. Eng. 2019, 56, 111–119. [Google Scholar] [CrossRef]
Figure 1. Tryptamines present in genus Psilocybe mushrooms.
Figure 1. Tryptamines present in genus Psilocybe mushrooms.
Pharmaceuticals 16 00040 g001
Figure 2. Aeruginascin analog (4-AcO-TMT).
Figure 2. Aeruginascin analog (4-AcO-TMT).
Pharmaceuticals 16 00040 g002
Scheme 1. Synthesis of psilocybin by Hofmann et al. in 1958 and 1959 [5,6,41].
Scheme 1. Synthesis of psilocybin by Hofmann et al. in 1958 and 1959 [5,6,41].
Pharmaceuticals 16 00040 sch001
Scheme 2. Improved synthesis of psilocybin by Nichols et al. in 1999 [42].
Scheme 2. Improved synthesis of psilocybin by Nichols et al. in 1999 [42].
Pharmaceuticals 16 00040 sch002
Scheme 3. Synthesis of psilacetin (O-acetyl)-prodrug by Nichols et al. in 1999 [42].
Scheme 3. Synthesis of psilacetin (O-acetyl)-prodrug by Nichols et al. in 1999 [42].
Pharmaceuticals 16 00040 sch003
Scheme 4. Synthesis of aeruginascin by Manke et al. [40].
Scheme 4. Synthesis of aeruginascin by Manke et al. [40].
Pharmaceuticals 16 00040 sch004
Scheme 5. Synthesis of psilocybin by Shirota et al. [43].
Scheme 5. Synthesis of psilocybin by Shirota et al. [43].
Pharmaceuticals 16 00040 sch005
Scheme 6. Reduction of ketoamide 11 via β-hydroxy intermediate 13.
Scheme 6. Reduction of ketoamide 11 via β-hydroxy intermediate 13.
Pharmaceuticals 16 00040 sch006
Scheme 7. Results of the scale-up for Sherwood’s psilocybin synthesis [44,46].
Scheme 7. Results of the scale-up for Sherwood’s psilocybin synthesis [44,46].
Pharmaceuticals 16 00040 sch007
Scheme 8. Results using POCl3 in the second generation of Kargbo’s psilocybin synthesis [46].
Scheme 8. Results using POCl3 in the second generation of Kargbo’s psilocybin synthesis [46].
Pharmaceuticals 16 00040 sch008
Scheme 9. 4-oxy-tryptamine synthesis by Hofmann [47].
Scheme 9. 4-oxy-tryptamine synthesis by Hofmann [47].
Pharmaceuticals 16 00040 sch009
Scheme 10. 14c-labeled psilocin synthesis by Kalberer et al. [9].
Scheme 10. 14c-labeled psilocin synthesis by Kalberer et al. [9].
Pharmaceuticals 16 00040 sch010
Scheme 11. Improved 14C-labeled psilocin synthesis by Poon et al. [49].
Scheme 11. Improved 14C-labeled psilocin synthesis by Poon et al. [49].
Pharmaceuticals 16 00040 sch011
Scheme 12. 3H-labeled psilocin synthesis by Poon et al. [49].
Scheme 12. 3H-labeled psilocin synthesis by Poon et al. [49].
Pharmaceuticals 16 00040 sch012
Scheme 13. 3-[2-(diethylamino)ethyl)]indol-4-ol synthesis by Repke et al. [51].
Scheme 13. 3-[2-(diethylamino)ethyl)]indol-4-ol synthesis by Repke et al. [51].
Pharmaceuticals 16 00040 sch013
Scheme 14. β-Carbolines synthesis by Repke et al. [52,53].
Scheme 14. β-Carbolines synthesis by Repke et al. [52,53].
Pharmaceuticals 16 00040 sch014
Scheme 15. Psilocin synthesis by Somei et al. [56].
Scheme 15. Psilocin synthesis by Somei et al. [56].
Pharmaceuticals 16 00040 sch015
Scheme 16. A new synthesis of psilocin by Ogasawara et al. [59].
Scheme 16. A new synthesis of psilocin by Ogasawara et al. [59].
Pharmaceuticals 16 00040 sch016
Scheme 17. Concise synthesis of psilocin by palladium-catalyzed cyclization [58].
Scheme 17. Concise synthesis of psilocin by palladium-catalyzed cyclization [58].
Pharmaceuticals 16 00040 sch017
Scheme 18. Synthesis of psilocin by palladium coupling [63].
Scheme 18. Synthesis of psilocin by palladium coupling [63].
Pharmaceuticals 16 00040 sch018
Scheme 19. Interesting direct iridium-catalyzed synthesis of psilocin [64].
Scheme 19. Interesting direct iridium-catalyzed synthesis of psilocin [64].
Pharmaceuticals 16 00040 sch019
Scheme 20. Synthetic/biocatalytic pathway to psilocin [65].
Scheme 20. Synthetic/biocatalytic pathway to psilocin [65].
Pharmaceuticals 16 00040 sch020
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Serreau, R.; Amirouche, A.; Benyamina, A.; Berteina-Raboin, S. A Review of Synthetic Access to Therapeutic Compounds Extracted from Psilocybe. Pharmaceuticals 2023, 16, 40. https://doi.org/10.3390/ph16010040

AMA Style

Serreau R, Amirouche A, Benyamina A, Berteina-Raboin S. A Review of Synthetic Access to Therapeutic Compounds Extracted from Psilocybe. Pharmaceuticals. 2023; 16(1):40. https://doi.org/10.3390/ph16010040

Chicago/Turabian Style

Serreau, Raphaël, Ammar Amirouche, Amine Benyamina, and Sabine Berteina-Raboin. 2023. "A Review of Synthetic Access to Therapeutic Compounds Extracted from Psilocybe" Pharmaceuticals 16, no. 1: 40. https://doi.org/10.3390/ph16010040

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop