A pilot study of cerebral metabolism and serotonin 5-HT2A receptor occupancy in rats treated with the psychedelic tryptamine DMT in conjunction with the MAO inhibitor harmine

Klemens Egger, Frederik Gudmundsen, Naja Støckel Jessen, Christina Baun, Sandra N Poetzsch, Vladimir Shalgunov, Matthias M Herth, Boris B Quednow, Chantal Martin-Soelch, Dario Dornbierer, Milan Scheidegger, Paul Cumming, Mikael Palner, Klemens Egger, Frederik Gudmundsen, Naja Støckel Jessen, Christina Baun, Sandra N Poetzsch, Vladimir Shalgunov, Matthias M Herth, Boris B Quednow, Chantal Martin-Soelch, Dario Dornbierer, Milan Scheidegger, Paul Cumming, Mikael Palner

Abstract

Rationale: The psychedelic effects of the traditional Amazonian botanical decoction known as ayahuasca are often attributed to agonism at brain serotonin 5-HT2A receptors by N,N-dimethyltryptamine (DMT). To reduce first pass metabolism of oral DMT, ayahuasca preparations additionally contain reversible monoamine oxidase A (MAO-A) inhibitors, namely β-carboline alkaloids such as harmine. However, there is lacking biochemical evidence to substantiate this pharmacokinetic potentiation of DMT in brain via systemic MAO-A inhibition. Objectives: We measured the pharmacokinetic profile of harmine and/or DMT in rat brain, and tested for pharmacodynamic effects on brain glucose metabolism and DMT occupancy at brain serotonin 5-HT2A receptors. Methods: We first measured brain concentrations of harmine and DMT after treatment with harmine and/or DMT at low sub-cutaneous doses (1 mg/kg each) or harmine plus DMT at moderate doses (3 mg/kg each). In the same groups of rats, we also measured ex vivo the effects of these treatments on the availability of serotonin 5-HT2A receptors in frontal cortex. Finally, we explored effects of DMT and/or harmine (1 mg/kg each) on brain glucose metabolism with [18F]FDG-PET. Results: Results confirmed that co-administration of harmine inhibited the formation of the DMT metabolite indole-3-acetic acid (3-IAA) in brain, while correspondingly increasing the cerebral availability of DMT. However, we were unable to detect any significant occupancy by DMT at 5-HT2A receptors measured ex vivo, despite brain DMT concentrations as high as 11.3 µM. We did not observe significant effects of low dose DMT and/or harmine on cerebral [18F]FDG-PET uptake. Conclusion: These preliminary results call for further experiments to establish the dose-dependent effects of harmine/DMT on serotonin receptor occupancy and cerebral metabolism.

Keywords: DMT; PKPD; [18 F]FDG-PET; ayahuasca; harmine; pharmahuasca; psychedelics; serotonin receptor.

Conflict of interest statement

KE, FG, NJ, CB, SP, VS, MH, BQ, and PC have nothing to declare. DD and MS co-founded Reconnect Labs, an academic spin-off at the University of Zurich. MP has an ongoing research collaboration with Compass Pathways Ltd. unrelated to the present study.

Copyright © 2023 Egger, Gudmundsen, Jessen, Baun, Poetzsch, Shalgunov, Herth, Quednow, Martin-Soelch, Dornbierer, Scheidegger, Cumming and Palner.

Figures

FIGURE 1
FIGURE 1
Experimental workflow in Experiments 1 and 2. (A) Experiment 1: Rats received either 2 × vehicle, harmine and vehicle, vehicle and DMT, or harmine and DMT (s.c.), then received successive injections of [3H]ketanserin (i.v.) and [18F]FDG (i.p). They were then allowed to roam freely for 45 min whereupon they were rapidly anesthetized for a PET/CT recording lasting 30 min. Next, rats were euthanized and brain samples (cerebellum and frontal cortex) were taken to measure [3H]ketanserin content by liquid scintillation counting, and the cerebral concentrations of harmine, DMT, and the metabolite 3-IAA by HPLC-MS. (B) Experiment 2: Rats received either 2 × vehicle, or harmine and DMT (s.c.), followed by [18F]MH.MZ injection (i.p). Rats were then allowed to roam freely for 45 min and then euthanized. Samples from their brain (cerebellum and frontal cortex) were taken to measure serotonin receptor occupancy/competition with a gamma counter and DMT, harmine and their metabolite concentrations with UHPLC-MS. For details, please refer to the text. DMT = N,N-dimethyltryptamine, [18F]FDG = [18F]fluorodeoxyglucose, PET = positron emission tomography. CT = computed tomography, UHPLC-MS = ultra-high performance liquid chromatography-mass spectroscopy. Figure created with BioRender.com.
FIGURE 2
FIGURE 2
Mean globally normalized SUV (SUVglob) results from experiment 1. [18F]FDG-PET data in groups of (n = 5) control and (n = 6) rats treated with harmine, DMT, or harmine + DMT (1 mg/kg s.c. each). The first two columns display a structural MRI template that was used for parcelation of brain PET data. The second column depicts the regions of interest, namely (from rostral to caudal): orbitofrontal cortex (OFC, blue), medial prefrontal cortex (mPFC, purple), nucleus accumbens (NAc, dark blue), striatum (red), anterodorsal hippocampus (dark green), posterior hippocampus (light green), thalamus (neon green), visual cortex (blue), cerebellum (yellow). The SUVglob images represent the mean of six frames of 5 minutes duration, starting at 60 min after [18F]FDG administration.

References

    1. Aghajanian G. K., Marek G. J. (1999). Serotonin and hallucinogens. Neuropsychopharmacology 21, 16S–23S. 10.1016/S0893-133X(98)00135-3
    1. Bærentzen S., Casado-Sainz A., Lange D., Shalgunov V., Tejada I. M., Xiong M., et al. (2019). The chemogenetic receptor ligand clozapine N-oxide induces in vivo neuroreceptor occupancy and reduces striatal glutamate levels. Front. Neurosci. 13, 187. 10.3389/fnins.2019.00187
    1. Barker S. A., Monti J. A., Christian S. T. (1980). Metabolism of the hallucinogen N,N-dimethyltryptamine in rat brain homogenates. Biochem. Pharmacol. 29, 1049–1057. 10.1016/0006-2952(80)90169-0
    1. Barker S. A. (2018). N, N-dimethyltryptamine (DMT), an endogenous hallucinogen: Past, present, and future research to determine its role and function. Front. Neurosci. 12, 536. 10.3389/fnins.2018.00536
    1. Bergstrom M., Westerberg G., Langstrom B. (1997). 11C-harmine as a tracer for monoamine oxidase A (MAO-A): in vitro and in vivo studies. Nucl. Med. Biol. 24, 287–293. 10.1016/s0969-8051(97)00013-9
    1. Berlowitz I., Ghasarian C., Walt H., Mendive F., Alvarado V., Martin-Soelch C. (2018). Conceptions and practices of an integrative treatment for substance use disorders involving amazonian medicine: Traditional healers’ perspectives. Braz J. Psychiatry 40, 200–209. 10.1590/1516-4446-2016-2117
    1. Brito-da-Costa A. M., Dias-da-Silva D., Gomes N. G. M., Dinis-Oliveira R. J., Madureira-Carvalho Á. (2020). Toxicokinetics and toxicodynamics of ayahuasca alkaloids N,N-dimethyltryptamine (DMT), harmine, harmaline and tetrahydroharmine: Clinical and forensic impact. Pharm. (Basel) 13, 334. 10.3390/ph13110334
    1. Buckholtz N. S., Boggan W. O. (1977). Inhibition by β-carbolines of monoamine uptake into a synaptosomal preparation: Structure-activity relationships. Life Sci. 20, 2093–2099. 10.1016/0024-3205(77)90190-4
    1. Callaway J. C., McKenna D. J., Grob C. S., Brito G. S., Raymon L. P., Poland R. E., et al. (1999). Pharmacokinetics of Hoasca alkaloids in healthy humans. J. Ethnopharmacol. 65, 243–256. 10.1016/s0378-8741(98)00168-8
    1. Canal C. E. (2018). Serotonergic psychedelics: Experimental approaches for assessing mechanisms of action. Handb. Exp. Pharmacol. 252, 227–260. 10.1007/164_2018_107
    1. Carbonaro T. M., Gatch M. B. (2016). Neuropharmacology of N,N-dimethyltryptamine. Brain Res. Bull. 126, 74–88. 10.1016/j.brainresbull.2016.04.016
    1. Carhart-Harris R. L., Nutt D. J. (2017). Serotonin and brain function: a tale of two receptors. J. Psychopharmacol. 31, 1091–1120. 10.1177/0269881117725915
    1. Cohen I., Vogel W. H. (1972). Determination and physiological disposition of dimethyltryptamine and diethyltryptamine in rat brain, liver and plasma. Biochem. Pharmacol. 21, 1214–1216. 10.1016/0006-2952(72)90119-0
    1. Cozzi N. V., Gopalakrishnan A., Anderson L. L., Feih J. T., Shulgin A. T., Daley P. F., et al. (2009). Dimethyltryptamine and other hallucinogenic tryptamines exhibit substrate behavior at the serotonin uptake transporter and the vesicle monoamine transporter. J. Neural Transm. 116, 1591–1599. 10.1007/s00702-009-0308-8
    1. Cumming P. (2011). Absolute abundances and affinity states of dopamine receptors in mammalian brain: A review. Synapse 65, 892–909. 10.1002/syn.20916
    1. Cumming P., Scheidegger M., Dornbierer D., Palner M., Quednow B. B., Martin-Soelch C. (2021). Molecular and functional imaging studies of psychedelic drug action in animals and humans. Molecules 26, 2451. 10.3390/molecules26092451
    1. Custodio R. J., Ortiz D. M., Lee H. J., Sayson L. V., Buctot D., Kim M., et al. (2022). 5-HT2CR is as important as 5-ht2ar in inducing hallucinogenic effects in serotonergic compounds. SSRN Electron. J. 10.2139/ssrn.4121838
    1. Erspamer V. (1955). Observations on the fate of indolalkylamines in the organism. J. Physiol. 127, 118–133. 10.1113/jphysiol.1955.sp005242
    1. Geyer M., Vollenweider F. (2008). Serotonin research: contributions to understanding psychoses. Trends Pharmacol. Sci. 29, 445–453. 10.1016/j.tips.2008.06.006
    1. Gouzoulis-Mayfrank E., Schreckenberger M., Sabri O., Arning C., Thelen B., Spitzer M., et al. (1999). Neurometabolic effects of psilocybin, 3,4-methylenedioxyethylamphetamine (MDE) and d-methamphetamine in healthy volunteers. A double-blind, placebo-controlled PET study with [18F]FDG. Neuropsychopharmacology 20, 565–581. 10.1016/s0893-133x(98)00089-x
    1. Guan Y., Louis E. D., Zheng W. (2001). Toxicokinetics of tremorogenic natural products, harmane and harmine, in male Sprague-Dawley rats. J. Toxicol. Environ. Health A 64, 645–660. 10.1080/152873901753246241
    1. Halberstadt A. L., Buell M. R., Masten V. L., Risbrough V. B., Geyer M. A. (2008). Modification of the effects of 5-methoxy-N,N-dimethyltryptamine on exploratory behavior in rats by monoamine oxidase inhibitors. Psychopharmacol. Berl. 201, 55–66. 10.1007/s00213-008-1247-z
    1. Halberstadt A. L., Chatha M., Klein A. K., Wallach J., Brandt S. D. (2020). Correlation between the potency of hallucinogens in the mouse head-twitch response assay and their behavioral and subjective effects in other species. Neuropharmacology 167, 107933. 10.1016/J.NEUROPHARM.2019.107933
    1. Hansen H. D., Ettrup A., Herth M. M., Dyssegaard A., Ratner C., Gillings N., et al. (2013). Direct comparison of [(18) F] and [(18) F] altanserin for 5-HT2A receptor imaging with PET. Synapse 67, 328–337. 10.1002/syn.21643
    1. Hasler F., Quednow B. B., Treyer V., Schubiger P. A., Buck A., Vollenweider F. X. (2009a). “Role of prefrontal serotonin-2A receptors in self-experience during psilocybin induced altered states,” in Annual Conference of the Swiss Society of Sleep Research, Sleep Medicine and Chronobiology (SSSSC) and the Swiss Society of Biological Psychiatry (SSBP), Bern, Switzerland, March 25 and 26, 2009, 59–60.
    1. Hasler F., Quednow B., Treyer V., Schubiger P., Buck A., Vollenweider F. (2009b). Role of prefrontal serotonin-2A receptors in self-experience during psilocybin induced altered states. Neuropsychobiology 59, 60. 10.1159/000209314
    1. Herth M. M., Debus F., Piel M., Palner M., Knudsen G. M., Lüddens H., et al. (2008). Total synthesis and evaluation of [18F]MHMZ. Bioorg Med. Chem. Lett. 18, 1515–1519. 10.1016/j.bmcl.2007.12.054
    1. Herth M. M., Kramer V., Piel M., Palner M., Riss P. J., Knudsen G. M., et al. (2009). Synthesis and in vitro affinities of various MDL 100907 derivatives as potential 18F-radioligands for 5-HT2A receptor imaging with PET. Bioorg Med. Chem. 17, 2989–3002. 10.1016/j.bmc.2009.03.021
    1. Jenner P., Marsden C. D., Thanki C. M. (1980). Behavioural changes induced by N,N-dimethyl-tryptamine in rodents. Br. J. Pharmacol. 69, 69–80. 10.1111/j.1476-5381.1980.tb10884.x
    1. Keiser M. J., Setola V., Irwin J. J., Laggner C., Abbas A. I., Hufeisen S. J., et al. (2009). Predicting new molecular targets for known drugs. Nature 462, 175–181. 10.1038/nature08506
    1. Koeck P. (2015). “Predicting the hallucinogenic action of N,N-dimethyltryptamine and harmine using multireceptor positron emission tomography – A mathematical model based on retrospective data analysis,”. Thesis for: Dr. med. univ. 10.13140/RG.2.2.33559.42401
    1. Kometer M., Schmidt A., Jäncke L., Vollenweider F. X. (2013). Activation of serotonin 2A receptors underlies the psilocybin-induced effects on α oscillations, N170 visual-evoked potentials, and visual hallucinations. J. Neurosci. 33, 10544–10551. 10.1523/jneurosci.3007-12.2013
    1. Li S., Zhang Y., Deng G., Wang Y., Qi S., Cheng X., et al. (2017). Exposure characteristics of the analogous β-carboline alkaloids harmaline and harmine based on the efflux transporter of multidrug resistance protein 2. Front. Pharmacol. 8, 541. 10.3389/fphar.2017.00541
    1. López-Giménez J. F., Mengod G., Palacios J. M., Vilaró M. T. (1997). Selective visualization of rat brain 5-HT2A receptors by autoradiography with [3H]MDL 100,907. Naunyn Schmiedeb. Arch. Pharmacol. 356, 446–454. 10.1007/pl00005075
    1. Lyon R. A., Titeler M., Seggel M. R., Glennon R. A. (1988). Indolealkylamine analogs share 5-HT2 binding characteristics with phenylalkylamine hallucinogens. Eur. J. Pharmacol. 145, 291–297. 10.1016/0014-2999(88)90432-3
    1. Madsen M. K., Fisher P. M., Burmester D., Dyssegaard A., Stenbæk D. S., Kristiansen S., et al. (2019). Psychedelic effects of psilocybin correlate with serotonin 2A receptor occupancy and plasma psilocin levels. Neuropsychopharmacology 44, 1328–1334. 10.1038/s41386-019-0324-9
    1. Maschauer S., Haller A., Riss P. J., Kuwert T., Prante O., Cumming P. (2015). Specific binding of [18F]fluoroethyl-harmol to monoamine oxidase A in rat brain cryostat sections, and compartmental analysis of binding in living brain. J. Neurochem. 135, 908–917. 10.1111/jnc.13370
    1. McAlonan K., Cavanaugh J., Wurtz R. H. (2008). Guarding the gateway to cortex with attention in visual thalamus. Nature 456, 391–394. 10.1038/nature07382
    1. McCulloch D. E., Madsen M. K., Stenbaek D. S., Kristiansen S., Ozenne B., Jensen P. S., et al. (2022). Lasting effects of a single psilocybin dose on resting-state functional connectivity in healthy individuals. J. Psychopharmacol. 36, 74–84. 10.1177/02698811211026454
    1. McEwen C. M., Jr., Cullen K. T., Sober A. J. (1966). Rabbit serum monoamine oxidase: Purification and characterization. J. Biol. Chem. 241, 4544–4556. 10.1016/S0021-9258(18)99753-X
    1. McKenna D. J., Repke D. B., Lo L., Peroutka S. J. (1990). Differential interactions of indolealkylamines with 5-hydroxytryptamine receptor subtypes. Neuropharmacology 29, 193–198. 10.1016/0028-3908(90)90001-8
    1. McKenna D. J., Towers G. H., Abbott F. (1984). Monoamine oxidase inhibitors in South American hallucinogenic plants: Tryptamine and beta-carboline constituents of ayahuasca. J. Ethnopharmacol. 10, 195–223. 10.1016/0378-8741(84)90003-5
    1. Nichols D. E. (2004). Hallucinogens. Pharmacol. Ther. 101, 131–181. 10.1016/j.pharmthera.2003.11.002
    1. Nichols D. E. (2016). Psychedelics. Pharmacol. Rev. 68, 264–355. 10.1124/pr.115.011478
    1. Odland A. U., Kristensen J. L., Andreasen J. T. (2022). Animal behavior in psychedelic research. Pharmacol. Rev. 74, 1176–1205. 10.1124/pharmrev.122.000590
    1. Ott J. (1999). Pharmahuasca: human pharmacology of oral DMT plus harmine. J. Psychoact. Drugs 31, 171–177. 10.1080/02791072.1999.10471741
    1. Pierce P. A., Peroutka S. J. (1989). Hallucinogenic drug interactions with neurotransmitter receptor binding sites in human cortex. Psychopharmacol. Berl. 97, 118–122. 10.1007/BF00443425
    1. Preller K. H., Burt J. B., Ji J. L., Schleifer C. H., Adkinson B. D., Stampfli P., et al. (2018). Changes in global and thalamic brain connectivity in LSD-induced altered states of consciousness are attributable to the 5-HT2A receptor. Elife 7, e35082. 10.7554/eLife.35082
    1. Preller K. H., Razi A., Zeidman P., Stampfli P., Friston K. J., Vollenweider F. X. (2019). Effective connectivity changes in LSD-induced altered states of consciousness in humans. Proc. Natl. Acad. Sci. U. S. A. 116, 2743–2748. 10.1073/pnas.1815129116
    1. Quednow B. B., Kometer M., Geyer M. A., Vollenweider F. X. (2012). Psilocybin-induced deficits in automatic and controlled inhibition are attenuated by ketanserin in healthy human volunteers. Neuropsychopharmacology 37, 630–640. 10.1038/npp.2011.228
    1. Ray T. S. (2010). Psychedelics and the human receptorome. PLoS One 5, e9019. 10.1371/journal.pone.0009019
    1. Riba J., Rodríguez-Fornells A., Urbano G., Morte A., Antonijoan R., Montero M., et al. (2001). Subjective effects and tolerability of the South American psychoactive beverage Ayahuasca in healthy volunteers. Psychopharmacol. Berl. 154, 85–95. 10.1007/s002130000606
    1. Riba J., Romero S., Grasa E., Mena E., Carrio I., Barbanoj M. J. (2006). Increased frontal and paralimbic activation following ayahuasca, the pan-Amazonian inebriant. Psychopharmacol. Berl. 186, 93–98. 10.1007/s00213-006-0358-7
    1. Riss P. J., Hong Y. T., Williamson D., Caprioli D., Sitnikov S., Ferrari V., et al. (2011). Validation and quantification of [18F]altanserin binding in the rat brain using blood input and reference tissue modeling. J. Cereb. Blood Flow Metabolism 31, 2334–2342. 10.1038/jcbfm.2011.94
    1. Rivier L., Lindgren J.-E. (1972). “Ayahuasca,” the South American hallucinogenic drink: An ethnobotanical and chemical investigation. Econ. Bot. 26, 101–129. 10.1007/BF02860772
    1. Sanches R. F., de Lima Osorio F., Dos Santos R. G., Macedo L. R., Maia-de-Oliveira J. P., Wichert-Ana L., et al. (2016). Antidepressant effects of a single dose of ayahuasca in patients with recurrent depression: A SPECT study. J. Clin. Psychopharmacol. 36, 77–81. 10.1097/JCP.0000000000000436
    1. Schiffer W. K., Mirrione M. M., Biegon A., Alexoff D. L., Patel V., Dewey S. L. (2006). Serial microPET measures of the metabolic reaction to a microdialysis probe implant. J. Neurosci. Methods 155, 272–284. 10.1016/j.jneumeth.2006.01.027
    1. Seabold S., Perktold J. (2010). Statsmodels: Econometric and statistical modeling with Python.
    1. Sitaram B. R., Lockett L., Talomsin R., Blackman G. L., McLeod W. R. (1987). In vivo metabolism of 5-methoxy-N, N-dimethyltryptamine and N,N-dimethyltryptamine in the rat. Biochem. Pharmacol. 36, 1509–1512. 10.1016/0006-2952(87)90118-3
    1. Smith R. L., Canton H., Barrett R. J., Sanders-Bush E. (1998). Agonist properties of N,N-dimethyltryptamine at serotonin 5-ht2a and 5-HT2C receptors. Pharmacol. Biochem. Behav. 61, 323–330. 10.1016/S0091-3057(98)00110-5
    1. Valle M., Maqueda A. E., Rabella M., Rodríguez-Pujadas A., Antonijoan R. M., Romero S., et al. (2016). Inhibition of alpha oscillations through serotonin-2A receptor activation underlies the visual effects of ayahuasca in humans. Eur. Neuropsychopharmacol. 26, 1161–1175. 10.1016/j.euroneuro.2016.03.012
    1. Vogt S. B., Ley L., Erne L., Straumann I., Becker A. M., Klaiber A., et al. (2023). Acute effects of intravenous DMT in a randomized placebo-controlled study in healthy participants. Transl. Psychiatry 13, 172. 10.1038/s41398-023-02477-4
    1. Vollenweider F. X., Leenders K. L., Scharfetter C., Maguire P., Stadelmann O., Angst J. (1997). Positron emission tomography and fluorodeoxyglucose studies of metabolic hyperfrontality and psychopathology in the psilocybin model of psychosis. Neuropsychopharmacology 16, 357–372. 10.1016/s0893-133x(96)00246-1
    1. Wang Lu L. J., Domino E. (1976). Effects of iproniazid and tranylcypromine of the half-life of N,N-dimethyltryptamine in rat brain and liver. Biochem. Pharmacol. 25, 1521–1527. 10.1016/0006-2952(76)90071-x
    1. Wang W., Jones H. E., Andolina I. M., Salt T. E., Sillito A. M. (2006). Functional alignment of feedback effects from visual cortex to thalamus. Nat. Neurosci. 9, 1330–1336. 10.1038/nn1768
    1. Wang Y., Wang H., Zhang L., Zhang Y., Sheng Y., Deng G., et al. (2019). Subchronic toxicity and concomitant toxicokinetics of long-term oral administration of total alkaloid extracts from seeds of peganum harmala linn: A 28-day study in rats. J. Ethnopharmacol. 238, 111866. 10.1016/j.jep.2019.111866

Source: PubMed

3
Se inscrever