Phytocannabinoids, Endocannabinoids and Synthetic Cannabinoids: Comparison
Please note this is a comparison between Version 2 by Camila Xu and Version 1 by Silvana Alfei.

Generally, the term ‘cannabinoids’ refers to a heterogeneous family of compounds that exhibit activity upon particular human cannabinoid receptors, namely CB1 and CB2. They encompass the natural compounds present in the Cannabis plants, lipid mediators called ECs naturally produced by human cells, as well as by all vertebrates on planet Earth, and the synthetic analogs of both groups designed by scientist, called SCs.

  • bacterial resistance
  • methicillin-resistant S. aureus (MRSA)
  • multi drug resistant (MDR) bacteria
  • Cannabis sativa
  • phytocannabinoids (PCs)

1. Introduction

Given the rapid emergence of multi drug resistant (MDR), extensively drug-resistant (XDR) and pandrug-resistant (PDR) pathogens, against which current antibiotics are no longer functioning, we are rapidly moving into a post-antibiotic era where infections will be practically untreatable [1]. According to the definition of the World Health Organization (WHO), antimicrobial resistance is a natural event that occurs when microbes become tolerant to drugs originally active, thus rendering several infections more difficult or impossible to treat [2,3][2][3]. Particularly, WHO has identified twelve families of bacteria to be considered as the most dangerous to human health. These families have been assigned to three priority groups, comprising critical pathogens (Acinetobacter, Pseudomonas, and Enterobacteriaceae), high priority pathogens (Enterococcus faecium, Staphylococcus aureus, Helicobacter pylori, Campylobacter, Salmonella spp., and Neisseria gonorrhoeae), and medium priority pathogens (Streptococcus pneumoniae, and Shigella spp.) [3,4][3][4]. Resistance in bacteria can be acquired or natural, but several mechanisms exist by which pathogens can become resistant to antibiotics (Figure 1).
Figure 1.
Mechanisms by which pathogens can become resistant.
As shown in Figure 1, antimicrobial resistance mechanisms include drug inactivation, decreased intracellular drug concentration, and altered drug targets [5].

1.1. Mechanism of Antimicrobial Resistance vs. Strategies to Develop Novel Antibiotics

Drug inactivation can occur either by enzymatic or chemical degradation, while decreased intracellular drug concentration can occur because of increasing drugs efflux and decreasing drugs influx [5]. In this regard, porin mutations in resistant strains alter the permeability of bacterial membranes, thus reducing the uptake of antibiotics into the bacterial cell. On the contrary, the hyperexpression of efflux pumps, which pump antibiotics out of the cell, dramatically reduces their concentration inside the cell [6]. Also, by the action of enzymes that chemically modify components of the bacterial outer membrane essential for antibiotic binding, some Gram-negative bacteria such as P. aeruginosa, Acinetobacter baumannii and others develop resistance to glycopeptide and polymyxin antibiotics. Furthermore, methyltransferases are a class of enzymes capable to modify the target thus promoting the resistance to antibiotics including aminoglycoside, lincosamide, macrolide, streptogramin, and oxazolidinone [7]. Another phenomenon known as “target protection” occurs when antibiotic target’s resistance proteins, such as the tetracycline ribosomal protection proteins (TRPPs), protect bacteria from the antibiotic-induced inhibition [8]. Additionally, the antibiotic resistance could be caused by the use of antibiotics in feed diet for animal production. The overuse, abuse, and misuse of β-lactams, aminoglycosides, tetracyclines, macrolides, and other antibiotics, with the purpose of promoting the development of animals, can cause the presence of residual antibiotics in the products intended for human consumption obtained from those animals, and can determine antibiotics pollution into the environment [9,10,11][9][10][11]. It was reported that some bacterial infections in humans are sustained by animal pathogens, namely zoonotic pathogens, thus proving that antibiotic resistance can be directly or indirectly transmitted from animal to humans [9]. A few practices, including the improvement of animal feed, waste management, and animal natural immunity, as well as the use of antibiotic alternatives such as prebiotics, probiotic vaccines, and bacteriophages can regulate and limit the antibiotic resistance, thus maintaining the potency of the available drugs [12]. However, more strategies to counteract antibiotic resistance are necessary, and currently they include the use of nanotechnology, computational methods, the use of antibiotic alternatives, drug repurposing, the synthesis of novel antibacterial agents, prodrugs, the development of efficient diagnostic agents also named rapid diagnostic tests (RDTs), the use of combination therapy, as well as the awareness, and knowledge of antibiotic prescribing (Table 1).
Table 1.
Strategies for combating antibiotic resistance.

1.2. Cannabinoids as Strategic Compounds to Develop New Antibiotics

Omitting to comment on each strategy reported in Table 1, and instead focusing on the development of alternative antibiotics, it can be observed that cannabinoids, better known for many other pharmacological and psychotropic effects are included in this category. Particularly, cannabinoids are prenylated polyketides produced in Cannabis plants and particularly in Cannabis sativa, which is an herbaceous plant that has been used for millennia for both medicinal and recreational purposes. C. sativa possesses a plethora of pharmacological properties and mind-altering effects, largely due to its content in cannabinoids, more precisely phytocannabinois (PCs), given their vegetable origin [28]. Collectively, more than 1600 chemical compounds have been isolated from C. sativa, of which over 500 are phytochemicals including cannabinoids, flavonoids terpenoids and sterols [28]. Among phytochemicals, more than 180 are cannabinoids, about 125 have been isolated, that can be classified into 11 structural families [28,29][28][29]. The most abundant representatives of these families are Δ9-tetrahydrocannabinol (Δ9-THC, also the main psychoactive cannabinoid), cannabidiol (CBD), and cannabichromene (CBC). Additionally, other classes whose prototypes are Δ8-E-tetrahydrocannabinol (Δ8-THC), cannabigerol (CBG), cannabinodiol (CBND), cannabielsoin (CBE), cannabicyclol (CBL), cannabinol (CBN), cannabitriol (CBT), and a miscellaneous group have been identified [28,29][28][29]. Currently, despite its psychotropic effects, Δ9- THC is used as therapeutic agent in the treatment of chemotherapy-associated nausea and vomiting, AIDS related loss of appetite, as well as pain and muscle spasms in multiple sclerosis [30]. Also, its carboxylic acid precursor, THCA, not exerting psycho-active effects in humans, is currently examined for its immunomodulatory, anti-inflammatory, neuroprotective and anti-neoplastic effects as well for its effectiveness in reducing adiposity and preventing metabolic disease caused by diet-induced obesity [31]. CBD, non-psychotropic as well, is currently investigated for application in the treatment of Alzheimer’s disease, Parkinson’s disease, epilepsy, cancer and for its neuroprotective efficacy [32]. Although the most studied cannabinoids for medicinal purposes are CBD and Δ9-THC, nowadays the research focus moves increasingly towards other PCs, such as the not psychoactive CBC, currently investigated for its anti-inflammatory, anti-fungal, antibiotic and analgesic effects [30], CBG and cannabigerolic acid (CBGA), which is the precursor of the decarboxylated CBG and could be considered as the “mother of all cannabinoids” (see later). Particularly, CBG has many putative benefits ranging from anti-inflammatory action to pain reliever [33]. Among other more investigated therapeutic properties, PCs including Δ9-THC, Δ8-THC, CBD, CBN, CBG, and CBC and some their correspondent carboxylic acids have shown from moderate to potent antimicrobial properties mainly against Gram-positive bacteria (MICs 0.5–8 µg/mL), and especially against strains of S. aureus, including MRSA, EMRSA, as well as fluoroquinolone and tetracycline-resistant strains, [34]. Particularly, even if the precise mechanisms used by PCs remains unknown so far, recent investigations have revealed that PCs inhibits bacteria by injuring their cytoplasmic membrane [35,36][35][36]. Recently, Luz-Veiga et al. have reported the antibacterial activity of both CBD and CBG, being CBG the most potent compound, and their capability to inhibit Staphylococci adherence to keratinocytes without compromising skin microbiota, thus being very promising as antibacterial agents to treat skin infection by topical administration [37]. Blaskovich et al., in addition to confirm the antibacterial activity of CBD on Gram-positive pathogens, including highly resistant S. aureus, S. pneumoniae, and Clostridioides difficile, demonstrated that CBD has excellent activity against biofilms, little propensity to induce resistance, and topical in vivo efficacy [38]. Moreover, the authors reported that CBD can selectively kill a subset of Gram-negative bacteria that includes the ‘urgent threat’ pathogen Neisseria gonorrhoeae [38]. Additionally, the interaction of CBD with broad-spectrum antibiotics such as ampicillin, kanamycin, and polymyxin B was studied by Gildea et al. [39]. By disrupting membrane integrity at extremely low dosages, CBD-antibiotic co-therapy showed synergistic activity against Salmonella typhimurium, offering an intriguing alternative in the treatment of this clinically relevant bacterium. The impressively strong antibacterial activity against MRSA of CBG has been reported by Farha et al. in the year 2020 [33]. Even in comparison with standard therapy with vancomycin, CBG outcompetes classical approaches against MRSA. Additionally, CBG demonstrated to inhibit the capability of MRSA to generate de novo biofilm, showed to succeed in disaggregating the pre-formed biofilm, to kill rapidly stationary phase cells (persisters), and to effectively inhibit MRSA also in vivo, in a murine model. The authors speculated that C. sativa may produce PCs as a natural defense mechanism against pathogens and suggested PCs as a new compound class serving as novel antibiotic drug [33].
Unfortunately, since in C. sativa, CBGA is promptly and directly converted to CBDA and THCA, leaving no CBGA pool available to form CBG, the CBG levels in plants are exceptionally low. In this context, it has been suggested that a possible strategy to increase the CBG yield from hemp biomass could consist in harvesting much earlier in the ripening phase of the plants before the other cannabinoids are formed and detract the CBGA from the cannabinoid pool [40]. On the other hand, having available reliable synthetic procedures to prepare natural PCs would consent the accessibility to considerable quantities of CBG, as well as of other microbiologically promising minor cannabinoids, unlikely provided naturally by Cannabis plants, thus allowing further studies finalized to the development of novel PCs-based antibiotics.

2. Phytocannabinoids (PCs), Endocannabinoids (ECs) and Synthetic Cannabinoids (SCs)

2.1. Phytocannabinoids (PCs) and Endocannabinoids (ECs)

Generally, the term ‘cannabinoids’ refers to a heterogeneous family of compounds that exhibit activity upon particular human cannabinoid receptors, namely CB1 and CB2 [41,42][41][42]. They encompass the natural compounds present in the Cannabis plants, lipid mediators called ECs naturally produced by human cells, as well as by all vertebrates on planet Earth, and the synthetic analogs of both groups designed by scientist, called SCs [42]. Natural cannabinoids from Cannabis are more specifically called PCs referring to their original plant source, differently from ECs which are produced from human cells [43,44][43][44]. PCs and ECs could include compounds structurally very different both between the two families and inside the same class, as shown in Figure 2 and Figure 3, which report the structure of the most relevant PCs and ECs, respectively.
Figure 2.
Chemical structure of the main PCs found in
C. sativa
acting on CB1 and/or CB2 receptors.
Figure 3.
Chemical structure of the main ECs found in humans acting on CB1 and/or CB2 receptors.
Both PCs and ECs exert their effects by interacting with CB1 and CB2 receptors, found throughout the human body, and whose locations have been listed in Table 2. Table 2 has constructed using the valuable information contained in the relevant work by Fraguas-Sánchez et al. [45].
Table 2.
Locations of CB1 and CB2 receptors in the human body.

References

  1. Vivas, R.; Barbosa, A.A.T.; Dolabela, S.S.; Jain, S. Multidrug-Resistant Bacteria and Alternative Methods to Control Them: An Overview. Microb. Drug Resist. 2019, 25, 890–908.
  2. Mancuso, G.; Midiri, A.; Gerace, E.; Biondo, C. Bacterial Antibiotic Resistance: The Most Critical Pathogens. Pathogens 2021, 10, 1310.
  3. WHO. Antimicrobial Resistance. Available online: https://www.who.int/news-room/fact-sheets/detail/antimicrobial-resistance (accessed on 3 May 2023).
  4. Stojković, D.; Petrović, J.; Carević, T.; Soković, M.; Liaras, K. Synthetic and Semisynthetic Compounds as Antibacterials Targeting Virulence Traits in Resistant Strains: A Narrative Updated Review. Antibiotics 2023, 12, 963.
  5. Chancey, S.T.; Zahner, D.; Stephens, D.S. Acquired inducible antimicrobial resistance in Gram-positive bacteria. Future Microbiol. 2012, 7, 959–978.
  6. Spengler, G.; Kincses, A.; Gajdacs, M.; Amaral, L. New Roads Leading to Old Destinations: Efflux Pumps as Targets to Reverse Multidrug Resistance in Bacteria. Molecules 2017, 22, 468.
  7. Schaenzer, A.J.; Wright, G.D. Antibiotic Resistance by Enzymatic Modification of Antibiotic Targets. Trends Mol. Med. 2020, 26, 768–782.
  8. Wilson, D.N.; Hauryliuk, V.; Atkinson, G.C.; O’Neill, A.J. Target protection as a key antibiotic resistance mechanism. Nat. Rev. Microbiol. 2020, 18, 637–648.
  9. Larsson, D.G.J.; Flach, C.F. Antibiotic resistance in the environment. Nat. Rev. Microbiol. 2022, 20, 257–269.
  10. Guetiya Wadoum, R.E.; Zambou, N.F.; Anyangwe, F.F.; Njimou, J.R.; Coman, M.M.; Verdenelli, M.C.; Cecchini, C.; Silvi, S.; Orpianesi, C.; Cresci, A.; et al. Abusive use of antibiotics in poultry farming in Cameroon and the public health implications. Br. Poult. Sci. 2016, 57, 483–493.
  11. Baynes, R.E.; Dedonder, K.; Kissell, L.; Mzyk, D.; Marmulak, T.; Smith, G.; Tell, L.; Gehring, R.; Davis, J.; Riviere, J.E. Health concerns and management of select veterinary drug residues. Food Chem. Toxicol. Int. J. Publ. Br. Ind. Biol. Res. Assoc. 2016, 88, 112–122.
  12. Ghosh, C.; Sarkar, P.; Issa, R.; Haldar, J. Alternatives to Conventional Antibiotics in the Era of Antimicrobial Resistance. Trend. Microbiol. 2019, 27, 323–338.
  13. Gupta, A.; Mumtaz, S.; Li, C.H.; Hussain, I.; Rotello, V.M. Combatting antibiotic-resistant bacteria using nanomaterials. Chem. Soc. Rev. 2019, 48, 415–427.
  14. Sarkar, D.J.; Mohanty, D.; Raut, S.S.; Das, B.K. Antibacterial properties and in silico odelling perspective of nano ZnO transported oxytetracycline-Zn2+ complex + against oxytetracycline-resistant Aeromonas hydrophila. J. Antibiot. 2022, 75, 635–649.
  15. Li, Q. Application of Fragment-Based Drug Discovery to Versatile Targets. Front. Mol. Biosci. 2020, 7, 180.
  16. Boyd, N.K.; Teng, C.; Frei, C.R. Brief Overview of Approaches and Challenges in New Antibiotic Development: A Focus On Drug Repurposing. Front. Cell. Infect. Microbiol. 2021, 11, 684515.
  17. Mazur, M.; Masłowiec, D. Antimicrobial Activity of Lactones. Antibiotics 2022, 11, 1327.
  18. de Ruyck, J.; Dupont, C.; Lamy, E.; Le Moigne, V.; Biot, C.; Guérardel, Y.; Herrmann, J.L.; Blaise, M.; Grassin-Delyle, S.; Kremer, L.; et al. Structure-Based Design and Synthesis of Piperidinol-Containing Molecules as New Mycobacterium abscessus Inhibitors. Chem. Open 2020, 9, 351–365.
  19. Dias, C.; Pais, J.P.; Nunes, R.; Blázquez-Sánchez, M.-T.; Marquês, J.T.; Almeida, A.F.; Serra, P.; Xavier, N.M.; Vila-Viçosa, D.; Machuqueiro, M.; et al. Sugar-based bactericides targeting phosphatidylethanolamine-enriched membranes. Nat. Commun. 2018, 9, 4857.
  20. Thakur, A.; Verma, M.; Setia, P.; Bharti, R.; Sharma, R.; Sharma, A.; Negi, N.P.; Anand, V.; Bansal, R. DFT analysis and in vitro studies of isoxazole derivatives as potent antioxidant and antibacterial agents synthesized via one-pot methodology. Res. Chem. Intermed. 2023, 49, 859–883.
  21. Patil, S.A.; Patil, S.A.; Ble-González, E.A.; Isbel, S.R.; Hampton, S.M.; Bugarin, A. Carbazole Derivatives as Potential Antimicrobial Agents. Molecules 2022, 27, 6575.
  22. Jubeh, B.; Breijyeh, Z.; Karaman, R. Antibacterial Prodrugs to Overcome Bacterial Resistance. Molecules 2020, 25, 1543.
  23. Bassetti, M.; Kanj, S.S.; Kiratisin, P.; Rodrigues, C.; Van Duin, D.; Villegas, M.V.; Yu, Y. Early appropriate diagnostics and treatment of MDR Gram-negative infections. JAC-Antimicrob. Resist. 2022, 4, dlac089.
  24. Tyers, M.; Wright, G.D. Drug combinations: A strategy to extend the life of antibiotics in the 21st century. Nat. Rev. Microbiol. 2019, 17, 141–155.
  25. Alfei, S.; Schito, A.M. β-Lactam Antibiotics and β-Lactamase Enzymes Inhibitors, Part 2: Our Limited Resources. Pharmaceuticals 2022, 15, 476.
  26. Alfei, S.; Zuccari, G. Recommendations to Synthetize Old and New β-Lactamases Inhibitors: A Review to Encourage Further Production. Pharmaceuticals 2022, 15, 384.
  27. Karasneh, R.A.; Al-Azzam, S.I.; Ababneh, M.; Al-Azzeh, O.; Al-Batayneh, O.B.; Muflih, S.M.; Khasawneh, M.; Khassawneh, A.M.; Khader, Y.S.; Conway, B.R.; et al. Prescribers’ Knowledge, Attitudes and Behaviors on Antibiotics, Antibiotic Use and Antibiotic Resistance in Jordan. Antibiotics 2021, 10, 858.
  28. Radwan, M.M.; Chandra, S.; Gul, S.; ElSohly, M.A. Cannabinoids, Phenolics, Terpenes and Alkaloids of Cannabis. Molecules 2021, 26, 2774.
  29. Tahir, M.N.; Shahbazi, F.; Rondeau-Gagné, S.; Trant, J.F. The biosynthesis of the cannabinoids. J. Cannabis. Res. 2021, 3, 7.
  30. Pagano, C.; Navarra, G.; Coppola, L.; Avilia, G.; Bifulco, M.; Laezza, C. Cannabinoids: Therapeutic Use in Clinical Practice. Int. J. Mol. Sci. 2022, 23, 3344.
  31. Palomares, B.; Ruiz-Pino, F.; Garrido-Rodriguez, M.; Eugenia Prados, M.; Sánchez-Garrido, M.A.; Velasco, I.; Vazquez, M.J.; Nadal, X.; Ferreiro-Vera, C.; Morrugares, R.; et al. Tetrahydrocannabinolic Acid A (THCA-A) Reduces Adiposity and Prevents Metabolic Disease Caused by Diet-Induced Obesity. Biochem. Pharmacol. 2020, 171, 113693.
  32. Pisanti, S.; Malfitano, A.M.; Ciaglia, E.; Lamberti, A.; Ranieri, R.; Cuomo, G.; Abate, M.; Faggiana, G.; Proto, M.C.; Fiore, D.; et al. Cannabidiol: State of the Art and New Challenges for Therapeutic Applications. Pharmacol. Ther. 2017, 175, 133–150.
  33. Farha, M.A.; El-Halfawy, O.M.; Gale, R.T.; MacNair, C.R.; Carfrae, L.A.; Zhang, X.; Jentsch, N.G.; Magolan, J.; Brown, E.D. Uncovering the Hidden Antibiotic Potential of Cannabis. ACS Infect. Dis. 2020, 6, 338–346.
  34. Breijyeh, Z.; Karaman, R. Design and Synthesis of Novel Antimicrobial Agents. Antibiotics 2023, 12, 628.
  35. Saleemi, M.A.; Yahaya, N.; Zain, N.N.M.; Raoov, M.; Yong, Y.K.; Noor, N.S.; Lim, V. Antimicrobial and Cytotoxic Effects of Cannabinoids: An Updated Review with Future Perspectives and Current Challenges. Pharmaceuticals 2022, 15, 1228.
  36. Chen, J.; Zhang, H.; Wang, S.; Du, Y.; Wei, B.; Wu, Q.; Wang, H. Inhibitors of Bacterial Extracellular Vesicles. Front. Microbiol. 2022, 13, 835058.
  37. Luz-Veiga, M.; Amorim, M.; Pinto-Ribeiro, I.; Oliveira, A.L.S.; Silva, S.; Pimentel, L.L.; Rodríguez-Alcalá, L.M.; Madureira, R.; Pintado, M.; Azevedo-Silva, J.; et al. Cannabidiol and Cannabigerol Exert Antimicrobial Activity without Compromising Skin Microbiota. Int. J. Mol. Sci. 2023, 24, 2389.
  38. Blaskovich, M.A.T.; Kavanagh, A.M.; Elliott, A.G.; Zhang, B.; Ramu, S.; Amado, M.; Lowe, G.J.; Hinton, A.O.; Pham, D.M.T.; Zuegg, J.; et al. The antimicrobial potential of cannabidiol. Commun. Biol. 2021, 4, 7.
  39. Gildea, L.; Ayariga, J.A.; Xu, J.; Villafane, R.; Robertson, B.K.; Samuel-Foo, M.; Ajayi, O.S. Cannabis sativa CBD Extract Exhibits Synergy with Broad-Spectrum Antibiotics against Salmonella enterica subsp. Enterica serovar typhimurium. Microorganisms 2022, 10, 2360.
  40. Calapai, F.; Cardia, L.; Esposito, E.; Ammendolia, I.; Mondello, C.; Lo Giudice, R.; Gangemi, S.; Calapai, G.; Mannucci, C. Pharmacological Aspects and Biological Effects of Cannabigerol and Its Synthetic Derivatives. Evid.-Based Complement. Altern. Med. 2022, 2022, 3336516.
  41. Whiting, P.F.; Wolff, R.F.; Deshpande, S.; Di Nisio, M.; Duffy, S.; Hernandez, A.V.; Keurentjes, J.C.; Lang, S.; Misso, K.; Ryder, S.; et al. Cannabinoids for Medical Use: A Systematic Review and Meta-analysis. JAMA 2015, 313, 2456–2473.
  42. Vučković, S.; Srebro, D.; Vujović, K.S.; Vučetić, Č.; Prostran, M. Cannabinoids and Pain: New Insights From Old Molecules. Front. Pharmacol. 2018, 9, 1259.
  43. Lafaye, G.; Karila, L.; Blecha, L.; Benyamina, A. Cannabis, Cannabinoids, and Health. DCNS 2017, 19, 309–316.
  44. Berman, P.; Futoran, K.; Lewitus, G.M.; Mukha, D.; Benami, M.; Shlomi, T.; Meiri, D. A New ESI-LC/MS Approach for Comprehensive Metabolic Profiling of Phytocannabinoids in Cannabis. Sci. Rep. 2018, 8, 14280.
  45. Fraguas-Sánchez, A.I.; Fernández-Carballido, A.; Torres-Suárez, A.I. Phyto-, Endo- and Synthetic Cannabinoids: Promising Chemotherapeutic Agents in the Treatment of Breast and Prostate Carcinomas. Expert. Opin. Investig. Drugs. 2016, 25, 1311–1323.
  46. Mackie, K. Cannabinoid Receptors: Where They are and What They do. J. Neuroendocr. 2008, 20, 10–14.
  47. Brennecke, B.; Gazzi, T.; Atz, K.; Fingerle, J.; Kuner, P.; Schindler, T.; Weck, G.; Nazaré, M.; Grether, U. Cannabinoid receptor type 2 ligands: An analysis of granted patents since 2010. Pharm. Patent Anal. 2021, 10, 111–163.
  48. Gertsch, J.; Raduner, S.; Altmann, K.-H. New Natural Noncannabinoid Ligands for Cannabinoid Type-2 (CB2) Receptors. J. Recept. Signal Transduct. 2006, 26, 709–730.
  49. Li, X.; Chang, H.; Bouma, J.; de Paus, L.V.; Mukhopadhyay, P.; Paloczi, J.; Mustafa, M.; van der Horst, C.; Kumar, S.S.; Wu, L.; et al. Structural Basis of Selective Cannabinoid CB2 Receptor Activation. Nat. Commun. 2023, 14, 1447.
  50. Lambert, D.M. Pharmacologic Targeting of the CB2 Cannabinoid Receptor for Application in Centrally-Mediated Chronic Pain. Ph.D. Thesis, University of British Columbia, Vancouver, BC, Canada, 2019. Available online: https://open.library.ubc.ca/collections/ubctheses/24/items/1.0376050 (accessed on 27 June 2023).
  51. Fezza, F.; Bari, M.; Florio, R.; Talamonti, E.; Feole, M.; Maccarrone, M. Endocannabinoids, Related Compounds and Their Metabolic Routes. Molecules 2014, 19, 17078–17106.
  52. Sharma, D.S.; Paddibhatla, I.; Raghuwanshi, S.; Malleswarapu, M.; Sangeeth, A.; Kovuru, N.; Dahariya, S.; Gautam, D.K.; Pallepati, A.; Gutti, R.K. Endocannabinoid system: Role in blood cell development, neuroimmune interactions and associated disorders. J. Neuroimmunol. 2021, 353, 577501.
  53. Formato, M.; Crescente, G.; Scognamiglio, M.; Fiorentino, A.; Pecoraro, M.T.; Piccolella, S.; Catauro, M.; Pacifico, S. (−)-Cannabidiolic Acid, a Still Overlooked Bioactive Compound: An Introductory Review and Preliminary Research. Molecules 2020, 25, 2638.
  54. Nguyen, G.N.; Jordan, E.N.; Kayser, O. Synthetic Strategies for Rare Cannabinoids Derived from Cannabis sativa. J. Nat. Prod. 2022, 85, 1555–1568.
  55. Schwilke, E.W.; Schwope, D.M.; Karschner, E.L.; Lowe, R.H.; Darwin, W.D.; Kelly, D.L.; Goodwin, R.S.; Gorelick, D.A.; Huestis, M.A. Δ9-Tetrahydrocannabinol (THC), 11-Hydroxy-THC, and 11-Nor-9-Carboxy-THC Plasma Pharmacokinetics during and after Continuous High-Dose Oral THC. Clin. Chem. 2009, 55, 2180–2189.
  56. Martin, B.R.; Jefferson, R.; Winckler, R.; Wiley, J.L.; Huffman, J.W.; Crocker, P.J.; Saha, B.; Razdan, R.K. Manipulation of the tetrahydrocannabinol side chain delineates agonists, partial agonists, and antagonists. J. Pharmacol. Exp. Ther. 1999, 290, 1065–1079.
  57. Andersson, D.A.; Gentry, C.; Alenmyr, L.; Killander, D.; Lewis, S.E.; Andersson, A.; Bucher, B.; Galzi, J.-L.; Sterner, O.; Bevan, S. TRPA1 mediates spinal antinociception induced by acetaminophen and the cannabinoid. δ 9-tetrahydrocannabiorcol. Nat. Commun. 2011, 2, 551.
  58. Bow, E.W.; Rimoldi, J.M. The structure–function relationships of classical cannabinoids: CB1/CB2 modulation. Perspect. Med. Chem. 2016, 8, 17–39.
  59. Thomas, A.; Ross, R.A.; Saha, B.; Mahadevan, A.; Razdan, R.K.; Pertwee, R.G. 6″-azidohex-2″-yne-cannabidiol: A potential neutral, competitive cannabinoid cb1 receptor antagonist. Eur. J. Pharmacol. 2004, 487, 213–221.
  60. O’Donnell, B.; Meissner, H.; Gupta, V. Dronabinol. In StatPearls; Updated 5 September 2022; StatPearls Publishing: Treasure Island, FL, USA, 2023. Available online: https://www.ncbi.nlm.nih.gov/books/NBK557531/ (accessed on 27 June 2023).
  61. (R)-(+)-Methanandamide. Available online: https://www.tocris.com/products/r-methanandamide_1121 (accessed on 3 May 2023).
  62. Gratzke, C.; Streng, T.; Stief, C.G.; Downs, T.R.; Alroy, I.; Rosenbaum, J.S.; Andersson, K.E.; Hedlund, P. Effects of cannabinor, a novel selective cannabinoid 2 receptor agonist, on bladder function in normal rats. Eur. Urol. 2010, 57, 1093–1100.
  63. D’Aquila, P.S. Microstructure analysis of the effects of the cannabinoid agents HU-210 and rimonabant in rats licking for sucrose. Eur. J. Pharmacol. 2020, 887, 173468.
  64. Ikeda, H.; Ikegami, M.; Kai, M.; Ohsawa, M.; Kamei, J. Activation of spinal cannabinoid CB2 receptors inhibits neuropathic pain in streptozotocin-induced diabetic mice. Neuroscience 2013, 250, 446–454.
  65. Du, J.J.; Liu, Z.Q.; Yan, Y.; Xiong, J.; Jia, X.T.; Di, Z.L.; Ren, J.J. The Cannabinoid WIN 55,212-2 Reduces Delayed Neurologic Sequelae After Carbon Monoxide Poisoning by Promoting Microglial M2 Polarization Through ST2 Signaling. J. Mol. Neurosci. MN 2020, 70, 422–432.
  66. Verty, A.N.; Stefanidis, A.; McAinch, A.J.; Hryciw, D.H.; Oldfield, B. Anti-Obesity Effect of the CB2 Receptor Agonist JWH-015 in Diet-Induced Obese Mice. PLoS ONE 2015, 10, e0140592.
  67. Howlett, A.C.; Thomas, B.F.; Huffman, J.W. The Spicy Story of Cannabimimetic Indoles. Molecules 2021, 26, 6190.
  68. Abadji, V.; Lin, S.; Taha, G.; Griffin, G.; Stevenson, L.A.; Pertwee, R.G.; Makriyannis, A. (R)-Methanandamide: A Chiral Novel Anandamide Possessing Higher Potency and Metabolic Stability. J. Med. Chem. 1994, 37, 1889–1893.
  69. WIN 55212-2. Available online: https://pubchem.ncbi.nlm.nih.gov/compound/5311501 (accessed on 3 May 2023).
  70. JWH-133. Available online: https://pubchem.ncbi.nlm.nih.gov/compound/6918505 (accessed on 3 May 2023).
  71. Hassenberg, C.; Clausen, F.; Hoffmann, G.; Studer, A.; Schürenkamp, J. Investigation of phase II metabolism of 11-hydroxy-Δ-9-tetrahydrocannabinol and metabolite verification by chemical synthesis of 11-hydroxy-Δ-9-tetrahydrocannabinol-glucuronide. Int. J. Legal Med. 2020, 134, 2105–2119.
  72. Engels, F.K.; de Jong, F.A.; Mathijssen, R.H.J.; Erkens, J.A.; Herings, R.M.; Verweij, J. Medicinal Cannabis in Oncology. Eu. J. Cancer 2007, 43, 2638–2644.
  73. Ward, S.J.; McAllister, S.D.; Kawamura, R.; Murase, R.; Neelakantan, H.; Walker, E.A. Cannabidiol Inhibits Paclitaxel-Induced Neuropathic Pain through 5-HT1A Receptors without Diminishing Nervous System Function or Chemotherapy Efficacy. Br. J. Pharmacol. 2014, 171, 636–645.
  74. Keating, G.M. Delta-9-Tetrahydrocannabinol/Cannabidiol Oromucosal Spray (Sativex®): A Review in Multiple Sclerosis-Related Spasticity. Drugs 2017, 77, 563–574.
  75. Reddy, D.S.; Golub, M.V. The Pharmacological Basis of Cannabis Therapy for Epilepsy. J. Pharmacol. Exp. Ther. 2016, 357, 45.
  76. Navarro, G.; Gonzalez, A.; Sánchez-Morales, A.; Casajuana-Martin, N.; Gómez-Ventura, M.; Cordomí, A.; Busqué, F.; Alibés, R.; Pardo, L.; Franco, R. Design of Negative and Positive Allosteric Modulators of the Cannabinoid CB2 Receptor Derived from the Natural Product Cannabidiol. J. Med. Chem. 2021, 64, 9354–9364.
  77. Luft, F.C. Rehabilitating rimonabant. J. Mol. Med. 2013, 91, 777–779.
More
Video Production Service