Actions of Analgesics on Nerve Conduction: Comparison
Please note this is a comparison between Version 3 by Eiichi Kumamoto and Version 2 by Peter Tang.

Abstract

Action potential (AP) conduction in nerve fibers plays an  crucimportantal role in transmitting nociceptive information from the periphery to cerebral cortex.  It is possible that analgesics depress nernerve AP conduction, inhibition resultings in antinociception. algesia. Many of analgesics are well-known to suppress nerve AP conduction and also voltage-sensitivegated Na+ and K+ channels involved in AP conduction. Compound action potential (CAP) hrecorded from as been used bundle of nerve fibers is a measure to know whether nerve AP conduction is affected by analgesics. This review article mentions the inhibitory effects of clinically-used analgesics, analgesic adjuvants and plant-derived analgesics on fast-conducting CAPs and voltage-gated Na+ and K+ channels. Their effects were compared in efficacy among the compounds and it was revealed that some of them have similar efficacies in suppressing CAPs. Nerve AP conduction inhibition produced by the analgesics is suggested to contribute to at least a part of their antinociceptive effects.

  • analgesic
  • antinociception
  • nerve conduction
  • sciatic nerve
  • compound action potential
  • Na+ channel
  • K+ channel
Please wait, diff process is still running!

References

  1. Fields, H.L. Pain; McGraw-Hill: New York, NY, USA, 1987. Fields, H.L. Pain; McGraw-Hill: New York, U.S.A., 1987.
  2. Willis, W.D., Jr.; Coggeshall, R.E. Sensory Mechanisms of the Spinal Cord, 2nd ed.; Plenum: New York, NY, USA, 1991. Willis, W.D. Jr.; Coggeshall, R.E. Sensory Mechanisms of the Spinal Cord, 2nd ed.; Plenum: New York, U.S.A., 1991.
  3. Merskey, H. Clarifying definition of neuropathic pain. Pain 2002, 96, 408–409. Todd, A.J. Neuronal circuitry for pain processing in the dorsal horn. Nat. Rev. Neurosci. 2010, 11, 823-836.
  4. Amir, R.; Argoff, C.E.; Bennett, G.J.; Cummins, T.R.; Durieux, M.E.; Gerner, P.; Gold, M.S.; Porreca, F.; Strichartz, G.R. The role of sodium channels in chronic inflammatory and neuropathic pain. J. Pain 2006, 7, S1–S29. Merighi, A. The histology, physiology, neurochemistry and circuitry of the substantia gelatinosa Rolandi (lamina II) in mammalian spinal cord. Prog. Neurobiol. 2018, 169, 91–134.
  5. Finnerup, N.B.; Sindrup, S.H.; Jensen, T.S. The evidence for pharmacological treatment of neuropathic pain. Pain 2010, 150, 573–581. Merskey, H. Clarifying definition of neuropathic pain. Pain 2002, 96, 408-409.
  6. Jensen, T.S. Anticonvulsants in neuropathic pain: Rationale and clinical evidence. Eur. J. Pain 2002, 6 (Suppl. A), 61–68. Amir, R.; Argoff, C.E.; Bennett, G.J.; Cummins, T.R.; Durieux, M.E.; Gerner, P.; Gold, M.S.; Porreca, F.; Strichartz, G.R. The role of sodium channels in chronic inflammatory and neuropathic pain. J. Pain 2006, 7, S1-S29.
  7. Kamibayashi, T.; Maze, M. Clinical uses of α2-adrenergic agonists. Anesthesiology 2000, 93, 1345–1349. Finnerup, N.B.; Sindrup, S.H.; Jensen, T.S. The evidence for pharmacological treatment of neuropathic pain. Pain 2010, 150, 573-581.
  8. Lynch, M.E. Antidepressants as analgesics: A review of randomized controlled trials. J. Psychiatry Neurosci. 2001, 26, 30–36. Jensen, T.S. Anticonvulsants in neuropathic pain: rationale and clinical evidence. Eur. J. Pain 2002, 6 (Suppl. A), 61-68.
  9. Sindrup, S.H.; Otto, M.; Finnerup, N.B.; Jensen, T.S. Antidepressants in the treatment of neuropathic pain. Basic Clin. Pharmacol. Toxicol. 2005, 96, 399–409. Kamibayashi, T., Maze, M. Clinical uses of α2-adrenergic agonists. Anesthesiology 2000, 93, 1345-1349.
  10. Theile, J.W.; Cummins, T.R. Recent developments regarding voltage-gated sodium channel blockers for the treatment of inherited and acquired neuropathic pain syndromes. Front. Pharmacol. 2011, 2, 54. Lynch, M.E. Antidepressants as analgesics: a review of randomized controlled trials. J. Psychiatry Neurosci. 2001, 26, 30-36.
  11. Waszkielewicz, A.M.; Gunia, A.; Słoczyńska, K.; Marona, H. Evaluation of anticonvulsants for possible use in neuropathic pain. Curr. Med. Chem. 2011, 18, 4344–4358. Sindrup, S.H.; Otto, M.; Finnerup, N.B.; Jensen, T.S. Antidepressants in the treatment of neuropathic pain. Basic Clin. Pharmacol. Toxicol. 2005, 96, 399-409.
  12. Fürst, S. Transmitters involved in antinociception in the spinal cord. Brain Res. Bull. 1999, 48, 129–141. Theile, J.W.; Cummins, T.R. Recent developments regarding voltage-gated sodium channel blockers for the treatment of inherited and acquired neuropathic pain syndromes. Front. Pharmacol. 2011, 2, 54.
  13. Kumamoto, E. Cellular mechanisms for antinociception produced by oxytocin and orexins in the rat spinal lamina II—Comparison with those of other endogenous pain modulators. Pharmaceuticals 2019, 12, 136. Waszkielewicz, A.M.; Gunia, A.; Słoczyńska, K.; Marona, H. Evaluation of anticonvulsants for possible use in neuropathic pain. Curr. Med. Chem. 2011, 18, 4344-4358.
  14. Zeilhofer, H.U.; Wildner, H.; Yévenes, G.E. Fast synaptic inhibition in spinal sensory processing and pain control. Physiol. Rev. 2012, 92, 193–235. Fakhri, S.; Abbaszadeh, F.; Jorjani, M. On the therapeutic targets and pharmacological treatments for pain relief following spinal cord injury: a mechanistic review. Biomed. Pharmacother. 2021, 139, 111563.
  15. Kiernan, M.C.; Bostock, H.; Park, S.B.; Kaji, R.; Krarup, C.; Krishnan, A.V.; Kuwabara, S.; Lin, C.S.; Misawa, S.; Moldovan, M.; et al. Measurement of axonal excitability: Consensus guidelines. Clin. Neurophysiol. 2020, 131, 308–323. Kocot-Kępska, M.; Zajączkowska, R.; Mika, J.; Kopsky, D.J.; Wordliczek, J.; Dobrogowski, J.; Przeklasa-Muszyńska, A. Topical treatments and their molecular/cellular mechanisms in patients with peripheral neuropathic pain - narrative review. Pharmaceutics 2021, 13, 450.
  16. Levitan, I.B.; Karczmarek, L.K. The Neuron, 3rd ed.; Oxford University Press: New York, NY, USA, 2002. Fürst, S. Transmitters involved in antinociception in the spinal cord. Brain Res. Bull. 1999, 48, 129-141.
  17. Kumamoto, E.; Mizuta, K.; Fujita, T. Peripheral nervous system in the frog as a tool to examine the regulation of the transmission of neuronal information. In Frogs: Biology, Ecology and Uses; Murray, J.L., Ed.; Nova Science Publishers, Inc.: New York, NY, USA, 2012; pp. 89–106. Kumamoto, E. Cellular mechanisms for antinociception produced by oxytocin and orexins in the rat spinal lamina II – comparison with those of other endogenous pain modulators. Pharmaceuticals 2019, 12, 136.
  18. Kobayashi, J.; Ohta, M.; Terada, Y. C fiber generates a slow Na+ spike in the frog sciatic nerve. Neurosci. Lett. 1993, 162, 93–96. Zeilhofer, H.U.; Wildner, H.; Yévenes, G.E. Fast synaptic inhibition in spinal sensory processing and pain control. Physiol. Rev. 2012, 92, 193-235.
  19. Fujita, T.; Kumamoto, E. Inhibition by endomorphin-1 and endomorphin-2 of excitatory transmission in adult rat substantia gelatinosa neurons. Neuroscience 2006, 139, 1095–1105. Gouveia, D.N.; Pina, L.T.S.; Rabelo, T.K.; da Rocha Santos, W.B.; Quintans, J.S.S.; Guimarães, A.G. Monoterpenes as perspective to chronic pain management: a systematic review. Curr. Drug Targets 2018, 19, 960-972.
  20. Kohno, T.; Kumamoto, E.; Higashi, H.; Shimoji, K.; Yoshimura, M. Actions of opioids on excitatory and inhibitory transmission in substantia gelatinosa of adult rat spinal cord. J. Physiol. 1999, 518, 803–813. Wang, Z.-J.; Heinbockel, T. Essential oils and their constituents targeting the GABAergic system and sodium channels as treatment of neurological diseases. Molecules 2018, 23, 1061.
  21. Yoshimura, M.; North, R.A. Substantia gelatinosa neurones hyperpolarized in vitro by enkephalin. Nature 1983, 305, 529–530. Gouveia, D.N.; Guimarães, A.G.; da Rocha Santos, W.B.; Quintans-Júnior, L.J. Natural products as a perspective for cancer pain management: a systematic review. Phytomedicine 2019, 58, 152766.
  22. North, R.A. Opioid actions on membrane ion channels. In Handbook of Experimental Pharmacology; Herz, A., Ed.; Springer: Berlin, Germany, 1993; Volume 104, pp. 773–797. Kiernan, M.C.; Bostock, H.; Park, S.B.; Kaji, R.; Krarup, C.; Krishnan, A.V.; Kuwabara, S.; Lin, C.S.; Misawa, S.; Moldovan, M.; Sung, J.; Vucic, S.; Wainger, B.J.; Waxman, S.; Burke, D. Measurement of axonal excitability: consensus guidelines. Clin. Neurophysiol. 2020, 131, 308-323.
  23. Yaksh, T.L. Pharmacology and mechanisms of opioid analgesic activity. Acta Anaesthesiol. Scand. 1997, 41, 94–111. Levitan, I.B.; Karczmarek, L.K. The Neuron, 3rd ed.; Oxford University Press: New York, U.S.A., 2002.
  24. Labuz, D.; Mousa, S.A.; Schäfer, M.; Stein, C.; Machelska, H. Relative contribution of peripheral versus central opioid receptors to antinociception. Brain Res. 2007, 1160, 30–38. Kumamoto, E.; Mizuta, K.; Fujita, T. Peripheral nervous system in the frog as a tool to examine the regulation of the transmission of neuronal information. In Frogs: Biology, Ecology and Uses; Murray, J.L., Ed.; Nova Science Publishers, Inc.: New York, U.S.A., 2012, pp. 89-106.
  25. Shannon, H.E.; Lutz, E.A. Comparison of the peripheral and central effects of the opioid agonists loperamide and morphine in the formalin test in rats. Neuropharmacology 2002, 42, 253–261. Kobayashi, J.; Ohta, M.; Terada, Y. C fiber generates a slow Na+ spike in the frog sciatic nerve. Neurosci. Lett. 1993, 162, 93-96.
  26. Smith, T.W.; Buchan, P.; Parsons, D.N.; Wilkinson, S. Peripheral antinociceptive effects of N-methyl morphine. Life Sci. 1982, 31, 1205–1208. Suzuki, R.; Fujita, T.; Mizuta, K.; Kumamoto, E. Inhibition by non-steroidal anti-inflammatory drugs of compound action potentials in frog sciatic nerve fibers. Biomed. Pharmacother. 2018, 103, 326-335.
  27. Wenk, H.N.; Brederson, J.-D.; Honda, C.N. Morphine directly inhibits nociceptors in inflamed skin. J. Neurophysiol. 2006, 95, 2083–2097. Katsuki, R.; Fujita, T.; Koga, A.; Liu, T.; Nakatsuka, T.; Nakashima, M.; Kumamoto, E. Tramadol, but not its major metabolite (mono-O-demethyl tramadol) depresses compound action potentials in frog sciatic nerves. Br. J. Pharmacol. 2006, 149, 319-327.
  28. Stein, C.; Schäfer, M.; Machelska, H. Attacking pain at its source: New perspectives on opioids. Nature Med. 2003, 9, 1003–1008. Mizuta, K.; Fujita, T.; Nakatsuka, T.; Kumamoto, E. Inhibitory effects of opioids on compound action potentials in frog sciatic nerves and their chemical structures. Life Sci. 2008, 83, 198-207.
  29. Yuge, O.; Matsumoto, M.; Kitahata, L.M.; Collins, J.G.; Senami, M. Direct opioid application to peripheral nerves does not alter compound action potentials. Anesth. Analg. 1985, 64, 667–671. Uemura, Y.; Fujita, T.; Ohtsubo, S.; Hirakawa, N.; Sakaguchi, Y.; Kumamoto, E. Effects of various antiepileptics used to alleviate neuropathic pain on compound action potential in frog sciatic nerves: comparison with those of local anesthetics. Biomed. Res. Int. 2014, 2014, 540238.
  30. Gissen, A.J.; Gugino, L.D.; Datta, S.; Miller, J.; Covino, B.G. Effects of fentanyl and sufentanil on peripheral mammalian nerves. Anesth. Analg. 1987, 66, 1272–1276. Magori, N.; Fujita, T.; Mizuta, K.; Kumamoto, E. Inhibition by general anesthetic propofol of compound action potentials in the frog sciatic nerve and its chemical structure. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2019, 392, 359-369.
  31. Jaffe, R.A.; Rowe, M.A. A comparison of the local anesthetic effects of meperidine, fentanyl, and sufentanil on dorsal root axons. Anesth. Analg. 1996, 83, 776–781. Mizuta, K.; Fujita, T.; Yamagata, H.; Kumamoto, E. Bisphenol A inhibits compound action potentials in the frog sciatic nerve in a manner independent of estrogen receptors. Biochem. Biophys. Rep. 2017, 10, 145-151.
  32. Jurna, I.; Grossmann, W. The effect of morphine on mammalian nerve fibres. Eur. J. Pharmacol. 1977, 44, 339–348. Tomohiro, D.; Mizuta, K.; Fujita, T.; Nishikubo, Y.; Kumamoto, E. Inhibition by capsaicin and its related vanilloids of compound action potentials in frog sciatic nerves. Life Sci. 2013, 92, 368-378.
  33. Coggeshall, R.E.; Zhou, S.; Carlton, S.M. Opioid receptors on peripheral sensory axons. Brain Res. 1997, 764, 126–132. Hirao, R.; Fujita, T.; Sakai, A.; Kumamoto, E. Compound action potential inhibition produced by various antidepressants in the frog sciatic nerve. Eur. J. Pharmacol. 2018, 819, 122-128.
  34. Fields, H.L.; Emson, P.C.; Leigh, B.K.; Gilbert, R.F.T.; Iversen, L.L. Multiple opiate receptor sites on primary afferent fibres. Nature 1980, 284, 351–353. Kosugi, T.; Mizuta, K.; Fujita, T.; Nakashima, M.; Kumamoto, E. High concentrations of dexmedetomidine inhibit compound action potentials in frog sciatic nerves without α2 adrenoceptor activation. Br. J. Pharmacol. 2010, 160, 1662-1676.
  35. Wenk, H.N.; Honda, C.N. Immunohistochemical localization of delta opioid receptors in peripheral tissues. J. Comp. Neurol. 1999, 408, 567–579. Kumamoto, E. Effects of plant-derived compounds on excitatory synaptic transmission and nerve conduction in the nervous system – involvement in pain modulation. Curr. Top. Phytochemistry 2018, 14, 45-70.
  36. Hunter, E.G.; Frank, G.B. An opiate receptor on frog sciatic nerve axons. Can. J. Physiol. Pharmacol. 1979, 57, 1171–1174. Ferreira, S.H. Prostaglandins, aspirin-like drugs and analgesia. Nat. New Biol. 1972, 240, 200-203.
  37. Klotz, U. Tramadol—The impact of its pharmacokinetic and pharmacodynamic properties on the clinical management of pain. Arzneimittelforschung 2003, 53, 681–687. Takayama, K.; Hirose, A.; Suda, I.; Miyazaki, A.; Oguchi, M.; Onotogi, M.; Fotopoulos, G. Comparison of the anti-inflammatory and analgesic effects in rats of diclofenac-sodium, felbinac and indomethacin patches. Int. J. Biomed. Sci. 2011, 7, 222-229.
  38. Lintz, W.; Erlacin, S.; Frankus, E.; Uragg, H. Metabolismus von Tramadol bei Mensch und Tier. Arzneimittelforschung 1981, 31, 1932–1943. Vane, J.R. Introduction: mechanism of action of NSAIDs. Br. J. Rheumatol. 1996, 35 (Suppl. 1), 1-3.
  39. Hennies, H.-H.; Friderichs, E.; Schneider, J. Receptor binding, analgesic and antitussive potency of tramadol and other selected opioids. Arzneimittelforschung 1988, 38, 877–880. Simmons, D.L.; Botting, R.M.; Hla, T. Cyclooxygenase isozymes: the biology of prostaglandin synthesis and inhibition. Pharmacol. Rev. 2004, 56, 387-437.
  40. Raffa, R.B.; Friderichs, E.; Reimann, W.; Shank, R.P.; Codd, E.E.; Vaught, J.L. Opioid and nonopioid components independently contribute to the mechanism of action of tramadol, an ’atypical’ opioid analgesic. J. Pharmacol. Exp. Ther. 1992, 260, 275–285. Grosser, T.; Smyth, E.; FitzGerald, G.A. Anti-inflammatory, antipyretic, and analgesic agents; pharmacotherapy of gout. In: Goodman & Gilman’s The Pharmacological Basis of Therapeutics, 12th ed.; Brunton, L.L.; Chabner, B.A.; Knollmann, B.C., Eds.; McGraw-Hill, Medical Publishing Division: New York, U.S.A., 2011, pp. 959-1004.
  41. Koga, A.; Fujita, T.; Totoki, T.; Kumamoto, E. Tramadol produces outward currents by activating μ-opioid receptors in adult rat substantia gelatinosa neurones. Br. J. Pharmacol. 2005, 145, 602–607. Kaduševičius E. Novel applications of NSAIDs: insight and future perspectives in cardiovascular, neurodegenerative, diabetes and cancer disease therapy. Int. J. Mol. Sci. 2021, 22, 6637.
  42. Koga, A.; Fujita, T.; Piao, L.-H.; Nakatsuka, T.; Kumamoto, E. Inhibition by O-desmethyltramadol of glutamatergic excitatory transmission in adult rat spinal substantia gelatinosa neurons. Mol. Pain 2019, 15, 1744806918824243. Garg, P.; Sanguinetti, M.C. Structure-activity relationship of fenamates as Slo2.1 channel activators. Mol. Pharmacol. 2012, 82, 795-802.
  43. Yamasaki, H.; Funai, Y.; Funao, T.; Mori, T.; Nishikawa, K. Effects of tramadol on substantia gelatinosa neurons in the rat spinal cord: An in vivo patch-clamp analysis. PLoS ONE 2015, 10, e0125147. Ortiz, M.I.; Torres-López, J.E.; Castañeda-Hernández, G.; Rosas, R.; Vidal-Cantú, G.C.; Granados-Soto, V. Pharmacological evidence for the activation of K+ channels by diclofenac. Eur. J. Pharmacol. 2002, 438, 85-91.
  44. Altunkaya, H.; Ozer, Y.; Kargi, E.; Babuccu, O. Comparison of local anaesthetic effects of tramadol with prilocaine for minor surgical procedures. Br. J. Anaesth. 2003, 90, 320–322. Ortiz, M.I.; Castañeda-Hernández, G.; Granados-Soto, V. Pharmacological evidence for the activation of Ca2+-activated K+ channels by meloxicam in the formalin test. Pharmacol. Biochem. Behav. 2005, 81, 725-731.
  45. Altunkaya, H.; Ozer, Y.; Kargi, E.; Ozkocak, I.; Hosnuter, M.; Demirel, C.B.; Babuccu, O. The postoperative analgesic effect of tramadol when used as subcutaneous local anesthetic. Anesth. Analg. 2004, 99, 1461–1464. Ortiz, M.I.; Granados-Soto, V.; Castañeda-Hernández, G. The NO-cGMP-K+ channel pathway participates in the antinociceptive effect of diclofenac, but not of indomethacin. Pharmacol. Biochem. Behav. 2003, 76, 187-195.
  46. Pang, W.-W.; Mok, M.S.; Chang, D.-P.; Huang, M.-H. Local anesthetic effect of tramadol, metoclopramide, and lidocaine following intradermal injection. Reg. Anesth. Pain Med. 1998, 23, 580–583. Peretz, A.; Degani, N.; Nachman, R.; Uziyel, Y.; Gibor, G.; Shabat, D.; Attali, B. Meclofenamic acid and diclofenac, novel templates of KCNQ2/Q3 potassium channel openers, depress cortical neuron activity and exhibit anticonvulsant properties. Mol. Pharmacol. 2005, 67, 1053-1066.
  47. Le Roux, P.J.; Coetzee, J.F. Tramadol today. Curr. Opin. Anaesth. 2000, 13, 457–461. Gwanyanya, A.; Macianskiene, R.; Mubagwa, K. Insights into the effects of diclofenac and other non-steroidal anti-inflammatory agents on ion channels. J. Pharm. Pharmacol. 2012, 64, 1359-1375.
  48. Tsai, Y.-C.; Chang, P.-J.; Jou, I.-M. Direct tramadol application on sciatic nerve inhibits spinal somatosensory evoked potentials in rats. Anesth. Analg. 2001, 92, 1547–1551. Guinamard, R.; Simard, C.; Del Negro, C. Flufenamic acid as an ion channel modulator. Pharmacol. Ther. 2013, 138, 272-284.
  49. Katsuki, R.; Fujita, T.; Koga, A.; Liu, T.; Nakatsuka, T.; Nakashima, M.; Kumamoto, E. Tramadol, but not its major metabolite (mono-O-demethyl tramadol) depresses compound action potentials in frog sciatic nerves. Br. J. Pharmacol. 2006, 149, 319–327. Voilley, N.; de Weille, J.; Mamet, J.; Lazdunski, M. Nonsteroid anti-inflammatory drugs inhibit both the activity and the inflammation-induced expression of acid-sensing ion channels in nociceptors. J. Neurosci. 2001, 21, 8026-8033.
  50. Mert, T.; Gunes, Y.; Guven, M.; Gunay, I.; Ozcengiz, D. Comparison of nerve conduction blocks by an opioid and a local anesthetic. Eur. J. Pharmacol. 2002, 439, 77–81. Inoue, N.; Ito, S.; Nogawa, M.; Tajima, K.; Kyoi, T. Etodolac blocks the allyl isothiocyanate-induced response in mouse sensory neurons by selective TRPA1 activation. Pharmacology 2012, 90, 47-54.
  51. Güven, M.; Mert, T.; Günay, I. Effects of tramadol on nerve action potentials in rat: Comparisons with benzocaine and lidocaine. Int. J. Neurosci. 2005, 115, 339–349. Suzuki, H.; Sasaki, E.; Nakagawa, A.; Muraki, Y.; Hatano, N.; Muraki, K. Diclofenac, a nonsteroidal anti-inflammatory drug, is an antagonist of human TRPM3 isoforms. Pharmacol. Res. Perspect. 2016, 4, e00232.
  52. Mert, T.; Gunes, Y.; Guven, M.; Gunay, I.; Gocmen, C. Differential effects of lidocaine and tramadol on modified nerve impulse by 4-aminopyridine in rats. Pharmacology 2003, 69, 68–73. Papworth, J.; Colville-Nash, P.; Alam, C.; Seed, M.; Willoughby, D. The depletion of substance P by diclofenac in the mouse. Eur. J. Pharmacol. 1997, 325, R1-R2.
  53. Gillen, C.; Haurand, M.; Kobelt, D.J.; Wnendt, S. Affinity, potency and efficacy of tramadol and its metabolites at the cloned human μ-opioid receptor. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2000, 362, 116–121. Silva, L.C.R.; Castor, M.G.Me., Souza, T.C.; Duarte, I.D.G.; Romero, T.R.L. NSAIDs induce peripheral antinociception by interaction with the adrenergic system. Life Sci. 2015, 130, 7-11.
  54. Driessen, B.; Reimann, W. Interaction of the central analgesic, tramadol, with the uptake and release of 5-hydroxytryptamine in the rat brain in vitro. Br. J. Pharmacol. 1992, 105, 147–151. Vazquez, E.; Hernandez, N.; Escobar, W.; Vanegas, H. Antinociception induced by intravenous dipyrone (metamizol) upon dorsal horn neurons: involvement of endogenous opioids at the periaqueductal gray matter, the nucleus raphe magnus, and the spinal cord in rats. Brain Res. 2005, 1048, 211-217.
  55. Driessen, B.; Reimann, W.; Giertz, H. Effects of the central analgesic tramadol on the uptake and release of noradrenaline and dopamine in vitro. Br. J. Pharmacol. 1993, 108, 806–811. Silva, L.C.R.; Castor, M.G.Me.; Navarro, L.C.; Romero, T.R.L.; Duarte, I.D.G. κ-Opioid receptor participates of NSAIDs peripheral antinociception. Neurosci. Lett. 2016, 622, 6-9.
  56. Leffler, A.; Frank, G.; Kistner, K.; Niedermirtl, F.; Koppert, W.; Reeh, P.W.; Nau, C. Local anesthetic-like inhibition of voltage-gated Na+ channels by the partial μ-opioid receptor agonist buprenorphine. Anesthesiology 2012, 116, 1335–1346. Fowler, C.J. NSAIDs: eNdocannabinoid stimulating anti-inflammatory drugs? Trends Pharmacol. Sci. 2012, 33, 468-473.
  57. Haeseler, G.; Foadi, N.; Ahrens, J.; Dengler, R.; Hecker, H.; Leuwer, M. Tramadol, fentanyl and sufentanil but not morphine block voltage-operated sodium channels. Pain 2006, 126, 234–244. McCormack, K.; Brune, K. Dissociation between the antinociceptive and anti-inflammatory effects of the nonsteroidal anti-inflammatory drugs. A survey of their analgesic efficacy. Drugs 1991, 41, 533-547.
  58. Tsai, T.-Y.; Tsai, Y.-C.; Wu, S.-N.; Liu, Y.-C. Tramadol-induced blockade of delayed rectifier potassium current in NG108-15 neuronal cells. Eur. J. Pain 2006, 10, 597–601. Lee, H.M.; Kim, H.I.; Shin, Y.K.; Lee, C.S.; Park, M.; Song, J.-H. Diclofenac inhibition of sodium currents in rat dorsal root ganglion neurons. Brain Res. 2003, 992, 120-127.
  59. Grond, S.; Meuser, T.; Uragg, H.; Stahlberg, H.J.; Lehmann, K.A. Serum concentrations of tramadol enantiomers during patient-controlled analgesia. Br. J. Clin. Pharmacol. 1999, 48, 254–257. Acosta, M.C.; Luna, C.; Graff, G.; Meseguer, V.M.; Viana, F.; Gallar, J.; Belmonte, C. Comparative effects of the nonsteroidal anti-inflammatory drug nepafenac on corneal sensory nerve fibers responding to chemical irritation. Invest. Ophthalmol. Vis. Sci. 2007, 48, 182-188.
  60. Mizuta, K.; Fujita, T.; Nakatsuka, T.; Kumamoto, E. Inhibitory effects of opioids on compound action potentials in frog sciatic nerves and their chemical structures. Life Sci. 2008, 83, 198–207. Fei, X.-W.; Liu, L.-Y.; Xu, J.-G.; Zhang, Z.-H.; Mei, Y.-A. The non-steroidal anti-inflammatory drug, diclofenac, inhibits Na+ current in rat myoblasts. Biochem. Biophys. Res. Commun. 2006, 346, 1275-1283.
  61. Brodin, P.; Skoglund, L.A. Dose-response inhibition of rat compound nerve action potential by dextropropoxyphene and codeine compared to morphine and cocaine in vitro. Gen. Pharmacol. 1990, 21, 551–553. Yarishkin, O.V.; Hwang, E.M.; Kim, D.; Yoo, J.C.; Kang, S.S.; Kim, D.R.; Shin, J.-H.-J.; Chung, H.-J.; Jeong, H.-S.; Kang, D.; Han, J.; Park, J.-Y.; Hong, S.-G. Diclofenac, a non-steroidal anti-inflammatory drug, inhibits L-type Ca2+ channels in neonatal rat ventricular cardiomyocytes. Korean J. Physiol. Pharmacol. 2009, 13, 437-442.
  62. Kumamoto, E.; Mizuta, K.; Fujita, T. Opioid actions in primary-afferent fibers—Involvement in analgesia and anesthesia. Pharmaceuticals 2011, 4, 343–365. Kuo, C.-C.; Huang, R.-C.; Lou, B.-S. Inhibition of Na+ current by diphenhydramine and other diphenyl compounds: molecular determinants of selective binding to the inactivated channels. Mol. Pharmacol. 2000, 57, 135-143.
  63. Bräu, M.E.; Nau, C.; Hempelmann, G.; Vogel, W. Local anesthetics potently block a potential insensitive potassium channel in myelinated nerve. J. Gen. Physiol. 1995, 105, 485–505. Yang, Y.-C.; Kuo, C.-C. An inactivation stabilizer of the Na+ channel acts as an opportunistic pore blocker modulated by external Na+. J. Gen. Physiol. 2005, 125, 465-481.
  64. Bräu, M.E.; Vogel, W.; Hempelmann, G. Fundamental properties of local anesthetics: Half-maximal blocking concentrations for tonic block of Na+ and K+ channels in peripheral nerve. Anesth. Analg. 1998, 87, 885–889. Yau, H.-J.; Baranauskas, G.; Martina, M. Flufenamic acid decreases neuronal excitability through modulation of voltage-gated sodium channel gating. J. Physiol. 2010, 588, 3869-3882.
  65. Tokuno, H.A.; Bradberry, C.W.; Everill, B.; Agulian, S.K.; Wilkes, S.; Baldwin, R.M.; Tamagnan, G.D.; Kocsis, J.D. Local anesthetic effects of cocaethylene and isopropylcocaine on rat peripheral nerves. Brain Res. 2004, 996, 159–167. Nakamura, M.; Jang, I.-S. pH-dependent inhibition of tetrodotoxin-resistant Na+ channels by diclofenac in rat nociceptive neurons. Prog. Neuropsychopharmacol. Biol. Psychiatry 2016, 64, 35-43.
  66. Chen, Z.R.; Irvine, R.J.; Somogyi, A.A.; Bochner, F. Mu receptor binding of some commonly used opioids and their metabolites. Life Sci. 1991, 48, 2165–2171. Sun, J.-F.; Xu, Y.-J.; Kong, X.-H.; Su, Y.; Wang, Z.-Y. Fenamates inhibit human sodium channel Nav1.7 and Nav1.8. Neurosci. Lett. 2019, 696, 67-73.
  67. Mizuta, K.; Fujita, T.; Kumamoto, E. Inhibition by morphine and its analogs of action potentials in adult rat dorsal root ganglion neurons. J. Neurosci. Res. 2012, 90, 1830–1841. Chen, X.; Gallar, J.; Belmonte, C. Reduction by antiinflammatory drugs of the response of corneal sensory nerve fibers to chemical irritation. Invest. Ophthalmol. Vis. Sci. 1997, 38, 1944-1953.
  68. Staiman, A.; Seeman, P. The impulse-blocking concentrations of anesthetics, alcohols, anticonvulsants, barbiturates, and narcotics on phrenic and sciatic nerves. Can. J. Physiol. Pharmacol. 1974, 52, 535–550. Kumamoto, E. Inhibition of fast nerve conduction produced by analgesics and analgesic adjuvants – possible involvement in pain alleviation. Pharmaceuticals 2020, 13, 62.
  69. Scholz, A. Mechanisms of (local) anaesthetics on voltage-gated sodium and other ion channels. Br. J. Anaesth. 2002, 89, 52–61. Gil-Flores, M.; Ortiz, M.I.; Castañeda-Hernández, G.; Chávez-Piña, A.E. Acemetacin antinociceptive mechanism is not related to NO or K+ channel pathways. Methods Find. Exp. Clin. Pharmacol. 2010, 32, 101-105.
  70. Hu, S.; Rubly, N. Effects of morphine on ionic currents in frog node of Ranvier. Eur. J. Pharmacol. 1983, 95, 185–192. Gögelein, H.; Dahlem, D.; Englert, H.C.; Lang, H.J. Flufenamic acid, mefenamic acid and niflumic acid inhibit single nonselective cation channels in the rat exocrine pancreas. FEBS Lett. 1990, 268, 79-82.
  71. Frazier, D.T.; Murayama, K.; Abbott, N.J.; Narahashi, T. Effects of morphine on internally perfused squid giant axons. Proc. Soc. Exp. Biol. Med. 1972, 139, 434–438. Hu, H.; Tian, J.; Zhu, Y.; Wang, C.; Xiao, R.; Herz, J.M.; Wood, J.D.; Zhu, M.X. Activation of TRPA1 channels by fenamate nonsteroidal anti-inflammatory drugs. Pflügers Arch. 2010, 459, 579-592.
  72. Wagner, L.E., II; Eaton, M.; Sabnis, S.S.; Gingrich, K.J. Meperidine and lidocaine block of recombinant voltage-dependent Na+ channels: Evidence that meperidine is a local anesthetic. Anesthesiology 1999, 91, 1481–1490. Tatematsu, Y.; Hayashi, H.; Taguchi, R.; Fujita, H.; Yamamoto, A.; Ohkura, K. Effect of N-phenylanthranilic acid scaffold nonsteroidal anti-inflammatory drugs on the mitochondrial permeability transition. Biol. Pharm. Bull. 2016, 39, 278-284.
  73. Viel, E.J.; Eledjam, J.J.; De La Coussaye, J.E.; D’Athis, F. Brachial plexus block with opioids for postoperative pain relief: Comparison between buprenorphine and morphine. Reg. Anesth. 1989, 14, 274–278. Glass, J.S.; Hardy, C.L.; Meeks, N.M.; Carroll, B.T. Acute pain management in dermatology: risk assessment and treatment. J. Am. Acad. Dermatol. 2015, 73, 543-560.
  74. Gutstein, H.B.; Akil, H. Opioid analgesics. In Goodman & Gilman’s the Pharmacological Basis of Therapeutics, 11th ed.; Brunton, L.L., Lazo, J.S., Parker, K.L., Eds.; McGraw-Hill, Medical Publishing Division: New York, NY, USA, 2006; pp. 547–590. Fujita, T.; Kumamoto, E. Inhibition by endomorphin-1 and endomorphin-2 of excitatory transmission in adult rat substantia gelatinosa neurons. Neuroscience 2006, 139, 1095-1105.
  75. King, M.; Su, W.; Chang, A.; Zuckerman, A.; Pasternak, G.W. Transport of opioids from the brain to the periphery by P-glycoprotein: Peripheral actions of central drugs. Nat. Neurosci. 2001, 4, 268–274. Kohno, T.; Kumamoto, E.; Higashi, H.; Shimoji, K.; Yoshimura, M. Actions of opioids on excitatory and inhibitory transmission in substantia gelatinosa of adult rat spinal cord. J. Physiol. 1999, 518, 803-813.
  76. Stein, C.; Comisel, K.; Haimerl, E.; Yassouridis, A.; Lehrberger, K.; Herz, A.; Peter, K. Analgesic effect of intraarticular morphine after arthroscopic knee surgery. N. Engl. J. Med. 1991, 325, 1123–1126. Yoshimura, M.; North, R.A. Substantia gelatinosa neurones hyperpolarized in vitro by enkephalin. Nature 1983, 305, 529-530.
  77. Mays, K.S.; Lipman, J.J.; Schnapp, M. Local analgesia without anesthesia using peripheral perineural morphine injections. Anesth. Analg. 1987, 66, 417–420. North, R.A. Opioid actions on membrane ion channels. In: Handbook of Experimental Pharmacology, Vol. 104; Herz, A., Ed.; Springer: Berlin, Germany, 1993, pp. 773-797.
  78. Cleary, J.; Mikus, G.; Somogyi, A.; Bochner, F. The influence of pharmacogenetics on opioid analgesia: Studies with codeine and oxycodone in the Sprague-Dawley/Dark Agouti rat model. J. Pharmacol. Exp. Ther. 1994, 271, 1528–1534. Yaksh, T.L. Pharmacology and mechanisms of opioid analgesic activity. Acta Anaesthesiol. Scand. 1997, 41, 94-111.
  79. Mikus, G.; Somogyi, A.A.; Bochner, F.; Eichelbaum, M. Codeine O-demethylation: Rat strain differences and the effects of inhibitors. Biochem. Pharmacol. 1991, 41, 757–762. Smith, T.W.; Buchan, P.; Parsons, D.N.; Wilkinson, S. Peripheral antinociceptive effects of N-methyl morphine. Life Sci. 1982, 31, 1205-1208.
  80. Ferreira, S.H. Prostaglandins, aspirin-like drugs and analgesia. Nat. New Biol. 1972, 240, 200–203. Stein, C.; Comisel, K.; Haimerl, E.; Yassouridis, A.; Lehrberger, K.; Herz, A.; Peter, K. Analgesic effect of intraarticular morphine after arthroscopic knee surgery. N. Engl. J. Med. 1991, 325, 1123-1126.
  81. Takayama, K.; Hirose, A.; Suda, I.; Miyazaki, A.; Oguchi, M.; Onotogi, M.; Fotopoulos, G. Comparison of the anti-inflammatory and analgesic effects in rats of diclofenac-sodium, felbinac and indomethacin patches. Int. J. Biomed. Sci. 2011, 7, 222–229. Shannon, H.E.; Lutz, E.A. Comparison of the peripheral and central effects of the opioid agonists loperamide and morphine in the formalin test in rats. Neuropharmacology 2002, 42, 253-261.
  82. Grosser, T.; Smyth, E.; FitzGerald, G.A. Anti-inflammatory, antipyretic, and analgesic agents; pharmacotherapy of gout. In Goodman & Gilman’s the Pharmacological Basis of Therapeutics, 12th ed.; Brunton, L.L., Chabner, B.A., Knollmann, B.C., Eds.; McGraw-Hill, Medical Publishing Division: New York, NY, USA, 2011; pp. 959–1004. Wenk, H.N.; Brederson, J.-D.; Honda, C.N. Morphine directly inhibits nociceptors in inflamed skin. J. Neurophysiol. 2006, 95, 2083-2097.
  83. Simmons, D.L.; Botting, R.M.; Hla, T. Cyclooxygenase isozymes: The biology of prostaglandin synthesis and inhibition. Pharmacol. Rev. 1996, 56 (Suppl. 1), 387–437. Labuz, D.; Mousa, S.A.; Schäfer, M.; Stein, C.; Machelska, H. Relative contribution of peripheral versus central opioid receptors to antinociception. Brain Res. 2007, 1160, 30-38.
  84. Vane, J.R. Introduction: Mechanism of action of NSAIDs. Br. J. Rheumatol. 1996, 35 (Suppl. 1), 1–3. Stein, C.; Schäfer, M.; Machelska, H. Attacking pain at its source: new perspectives on opioids. Nature Med. 2003, 9, 1003-1008.
  85. Voilley, N.; de Weille, J.; Mamet, J.; Lazdunski, M. Nonsteroid anti-inflammatory drugs inhibit both the activity and the inflammation-induced expression of acid-sensing ion channels in nociceptors. J. Neurosci. 2001, 21, 8026–8033. Yuge, O.; Matsumoto, M.; Kitahata, L.M.; Collins, J.G.; Senami, M. Direct opioid application to peripheral nerves does not alter compound action potentials. Anesth. Analg. 1985, 64, 667-671.
  86. Inoue, N.; Ito, S.; Nogawa, M.; Tajima, K.; Kyoi, T. Etodolac blocks the allyl isothiocyanate-induced response in mouse sensory neurons by selective TRPA1 activation. Pharmacology 2012, 90, 47–54. Gissen, A.J.; Gugino, L.D.; Datta, S.; Miller, J.; Covino, B.G. Effects of fentanyl and sufentanil on peripheral mammalian nerves. Anesth. Analg. 1987, 66, 1272-1276.
  87. Suzuki, H.; Sasaki, E.; Nakagawa, A.; Muraki, Y.; Hatano, N.; Muraki, K. Diclofenac, a nonsteroidal anti-inflammatory drug, is an antagonist of human TRPM3 isoforms. Pharmacol. Res. Perspect. 2016, 4, e00232. Jaffe, R.A.; Rowe, M.A. A comparison of the local anesthetic effects of meperidine, fentanyl, and sufentanil on dorsal root axons. Anesth. Analg. 1996, 83, 776-781.
  88. Garg, P.; Sanguinetti, M.C. Structure-activity relationship of fenamates as Slo2.1 channel activators. Mol. Pharmacol. 2012, 82, 795–802. Jurna, I.; Grossmann, W. The effect of morphine on mammalian nerve fibres. Eur. J. Pharmacol. 1977, 44, 339-348.
  89. Ortiz, M.I.; Torres-López, J.E.; Castañeda-Hernández, G.; Rosas, R.; Vidal-Cantú, G.C.; Granados-Soto, V. Pharmacological evidence for the activation of K+ channels by diclofenac. Eur. J. Pharmacol. 2002, 438, 85–91. Coggeshall, R.E.; Zhou, S.; Carlton, S.M. Opioid receptors on peripheral sensory axons. Brain Res. 1997, 764, 126-132.
  90. Ortiz, M.I.; Castañeda-Hernández, G.; Granados-Soto, V. Pharmacological evidence for the activation of Ca2+-activated K+ channels by meloxicam in the formalin test. Pharmacol. Biochem. Behav. 2005, 81, 725–731. Fields, H.L.; Emson, P.C.; Leigh, B.K.; Gilbert, R.F.T.; Iversen, L.L. Multiple opiate receptor sites on primary afferent fibres. Nature 1980, 284, 351-353.
  91. Ortiz, M.I.; Granados-Soto, V.; Castañeda-Hernández, G. The NO-cGMP-K+ channel pathway participates in the antinociceptive effect of diclofenac, but not of indomethacin. Pharmacol. Biochem. Behav. 2003, 76, 187–195. Wenk, H.N.; Honda, C.N. Immunohistochemical localization of delta opioid receptors in peripheral tissues. J. Comp. Neurol. 1999, 408, 567-579.
  92. Peretz, A.; Degani, N.; Nachman, R.; Uziyel, Y.; Gibor, G.; Shabat, D.; Attali, B. Meclofenamic acid and diclofenac, novel templates of KCNQ2/Q3 potassium channel openers, depress cortical neuron activity and exhibit anticonvulsant properties. Mol. Pharmacol. 2005, 67, 1053–1066. Klotz, U. Tramadol - the impact of its pharmacokinetic and pharmacodynamic properties on the clinical management of pain. Arzneimittelforschung 2003, 53, 681-687.
  93. Guinamard, R.; Simard, C.; Del Negro, C. Flufenamic acid as an ion channel modulator. Pharmacol. Ther. 2013, 138, 272–284. Lintz, W.; Erlacin, S.; Frankus, E.; Uragg, H. Metabolismus von Tramadol bei Mensch und Tier. Arzneimittelforschung 1981, 31, 1932-1943.
  94. Gwanyanya, A.; Macianskiene, R.; Mubagwa, K. Insights into the effects of diclofenac and other non-steroidal anti-inflammatory agents on ion channels. J. Pharm. Pharmacol. 2012, 64, 1359–1375. Hennies, H.-H.; Friderichs, E.; Schneider, J. Receptor binding, analgesic and antitussive potency of tramadol and other selected opioids. Arzneimittelforschung 1988, 38, 877-880.
  95. Papworth, J.; Colville-Nash, P.; Alam, C.; Seed, M.; Willoughby, D. The depletion of substance P by diclofenac in the mouse. Eur. J. Pharmacol. 1997, 325, R1–R2. Raffa, R.B.; Friderichs, E.; Reimann, W.; Shank, R.P.; Codd, E.E.; Vaught, J.L. Opioid and nonopioid components independently contribute to the mechanism of action of tramadol, an 'atypical' opioid analgesic. J. Pharmacol. Exp. Ther. 1992, 260, 275-285.
  96. Silva, L.C.R.; Castor, M.G.M.; Souza, T.C.; Duarte, I.D.G.; Romero, T.R.L. NSAIDs induce peripheral antinociception by interaction with the adrenergic system. Life Sci. 2015, 130, 7–11. Koga, A.; Fujita, T.; Totoki, T.; Kumamoto, E. Tramadol produces outward currents by activating μ-opioid receptors in adult rat substantia gelatinosa neurones. Br. J. Pharmacol. 2005, 145, 602-607.
  97. Silva, L.C.R.; Castor, M.G.Me.; Navarro, L.C.; Romero, T.R.L.; Duarte, I.D.G. κ-Opioid receptor participates of NSAIDs peripheral antinociception. Neurosci. Lett. 2016, 622, 6–9. Koga, A.; Fujita, T.; Piao, L.-H.; Nakatsuka, T.; Kumamoto, E. Inhibition by O-desmethyltramadol of glutamatergic excitatory transmission in adult rat spinal substantia gelatinosa neurons. Mol. Pain 2019, 15.
  98. Vazquez, E.; Hernandez, N.; Escobar, W.; Vanegas, H. Antinociception induced by intravenous dipyrone (metamizol) upon dorsal horn neurons: Involvement of endogenous opioids at the periaqueductal gray matter, the nucleus raphe magnus, and the spinal cord in rats. Brain Res. 2005, 1048, 211–217. Yamasaki, H.; Funai, Y.; Funao, T.; Mori, T.; Nishikawa, K. Effects of tramadol on substantia gelatinosa neurons in the rat spinal cord: an in vivo patch-clamp analysis. PLoS One 2015, 10, e0125147.
  99. Fowler, C.J. NSAIDs: eNdocannabinoid stimulating anti-inflammatory drugs? Trends Pharmacol. Sci. 2012, 33, 468–473. Altunkaya, H.; Ozer, Y.; Kargi, E.; Babuccu, O. Comparison of local anaesthetic effects of tramadol with prilocaine for minor surgical procedures. Br. J. Anaesth. 2003, 90, 320-322.
  100. McCormack, K.; Brune, K. Dissociation between the antinociceptive and anti-inflammatory effects of the nonsteroidal anti-inflammatory drugs. A survey of their analgesic efficacy. Drugs 1991, 41, 533–547. Altunkaya, H.; Ozer, Y.; Kargi, E.; Ozkocak, I.; Hosnuter, M.; Demirel, C.B.; Babuccu, O. The postoperative analgesic effect of tramadol when used as subcutaneous local anesthetic. Anesth. Analg. 2004, 99, 1461-1464.
  101. Suzuki, R.; Fujita, T.; Mizuta, K.; Kumamoto, E. Inhibition by non-steroidal anti-inflammatory drugs of compound action potentials in frog sciatic nerve fibers. Biomed. Pharmacother. 2018, 103, 326–335. Pang, W.-W.; Mok, M.S.; Chang, D.-P.; Huang, M.-H. Local anesthetic effect of tramadol, metoclopramide, and lidocaine following intradermal injection. Reg. Anesth. Pain Med. 1998, 23, 580-583.
  102. Lee, H.M.; Kim, H.I.; Shin, Y.K.; Lee, C.S.; Park, M.; Song, J.-H. Diclofenac inhibition of sodium currents in rat dorsal root ganglion neurons. Brain Res. 2003, 992, 120–127. Le Roux, P.J.; Coetzee, J.F. Tramadol today. Curr. Opin. Anaesth. 2000, 13, 457-461.
  103. Acosta, M.C.; Luna, C.; Graff, G.; Meseguer, V.M.; Viana, F.; Gallar, J.; Belmonte, C. Comparative effects of the nonsteroidal anti-inflammatory drug nepafenac on corneal sensory nerve fibers responding to chemical irritation. Investig. Ophthalmol. Vis. Sci. 2007, 48, 182–188. Tsai, Y.-C.; Chang, P.-J.; Jou, I.-M. Direct tramadol application on sciatic nerve inhibits spinal somatosensory evoked potentials in rats. Anesth. Analg. 2001, 92, 1547-1551.
  104. Fei, X.-W.; Liu, L.-Y.; Xu, J.-G.; Zhang, Z.-H.; Mei, Y.-A. The non-steroidal anti-inflammatory drug, diclofenac, inhibits Na+ current in rat myoblasts. Biochem. Biophys. Res. Commun. 2006, 346, 1275–1283. Shin, H.-W.; Ju, B.-J.; Jang, Y.-K.; You, H.-S.; Kang, H.; Park, J.-Y. Effect of tramadol as an adjuvant to local anesthetics for brachial plexus block: a systematic review and meta-analysis. PLoS One 2017, 12, e0184649.
  105. Yarishkin, O.V.; Hwang, E.M.; Kim, D.; Yoo, J.C.; Kang, S.S.; Kim, D.R.; Shin, J.-H.-J.; Chung, H.-J.; Jeong, H.-S.; Kang, D.; et al. Diclofenac, a non-steroidal anti-inflammatory drug, inhibits L-type Ca2+ channels in neonatal rat ventricular cardiomyocytes. Korean J. Physiol. Pharmacol. 2009, 13, 437–442. Mert, T.; Gunes, Y.; Guven, M.; Gunay, I.; Ozcengiz, D. Comparison of nerve conduction blocks by an opioid and a local anesthetic. Eur. J. Pharmacol. 2002, 439, 77-81.
  106. Kuo, C.-C.; Huang, R.-C.; Lou, B.-S. Inhibition of Na+ current by diphenhydramine and other diphenyl compounds: Molecular determinants of selective binding to the inactivated channels. Mol. Pharmacol. 2000, 57, 135–143. Güven, M.; Mert, T.; Günay, I. Effects of tramadol on nerve action potentials in rat: comparisons with benzocaine and lidocaine. Int. J. Neurosci. 2005, 115, 339-349.
  107. Yang, Y.-C.; Kuo, C.-C. An inactivation stabilizer of the Na+ channel acts as an opportunistic pore blocker modulated by external Na+. J. Gen. Physiol. 2005, 125, 465–481. Mert, T.; Gunes, Y.; Guven, M.; Gunay, I.; Gocmen, C. Differential effects of lidocaine and tramadol on modified nerve impulse by 4-aminopyridine in rats. Pharmacology 2003, 69, 68-73.
  108. Yau, H.-J.; Baranauskas, G.; Martina, M. Flufenamic acid decreases neuronal excitability through modulation of voltage-gated sodium channel gating. J. Physiol. 2010, 588, 3869–3882. Gillen, C.; Haurand, M.; Kobelt, D.J.; Wnendt, S. Affinity, potency and efficacy of tramadol and its metabolites at the cloned human μ-opioid receptor. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2000, 362, 116-121.
  109. Nakamura, M.; Jang, I.-S. pH-dependent inhibition of tetrodotoxin-resistant Na+ channels by diclofenac in rat nociceptive neurons. Prog. Neuropsychopharmacol. Biol. Psychiatry 2016, 64, 35–43. Driessen, B.; Reimann, W. Interaction of the central analgesic, tramadol, with the uptake and release of 5-hydroxytryptamine in the rat brain in vitro. Br. J. Pharmacol. 1992, 105, 147-151.
  110. Sun, J.-F.; Xu, Y.-J.; Kong, X.-H.; Su, Y.; Wang, Z.-Y. Fenamates inhibit human sodium channel Nav1.7 and Nav1.8. Neurosci. Lett. 2019, 696, 67–73. Driessen, B.; Reimann, W.; Giertz, H. Effects of the central analgesic tramadol on the uptake and release of noradrenaline and dopamine in vitro. Br. J. Pharmacol. 1993, 108, 806-811.
  111. Chen, X.; Gallar, J.; Belmonte, C. Reduction by antiinflammatory drugs of the response of corneal sensory nerve fibers to chemical irritation. Investig. Ophthalmol. Vis. Sci. 1997, 38, 1944–1953. Leffler, A.; Frank, G.; Kistner, K.; Niedermirtl, F.; Koppert, W.; Reeh, P.W.; Nau, C. Local anesthetic-like inhibition of voltage-gated Na+ channels by the partial μ-opioid receptor agonist buprenorphine. Anesthesiology 2012, 116, 1335-1346.
  112. Mizuta, K.; Fujita, T.; Yamagata, H.; Kumamoto, E. Bisphenol A inhibits compound action potentials in the frog sciatic nerve in a manner independent of estrogen receptors. Biochem. Biophys. Rep. 2017, 10, 145–151. Haeseler, G.; Foadi, N.; Ahrens, J.; Dengler, R.; Hecker, H.; Leuwer, M. Tramadol, fentanyl and sufentanil but not morphine block voltage-operated sodium channels. Pain 2006, 126, 234-244.
  113. Gil-Flores, M.; Ortiz, M.I.; Castañeda-Hernández, G.; Chávez-Piña, A.E. Acemetacin antinociceptive mechanism is not related to NO or K+ channel pathways. Methods Find. Exp. Clin. Pharmacol. 2010, 32, 101–105. Tsai, T.-Y.; Tsai, Y.-C.; Wu, S.-N.; Liu, Y.-C. Tramadol-induced blockade of delayed rectifier potassium current in NG108-15 neuronal cells. Eur. J. Pain 2006, 10, 597-601.
  114. Gögelein, H.; Dahlem, D.; Englert, H.C.; Lang, H.J. Flufenamic acid, mefenamic acid and niflumic acid inhibit single nonselective cation channels in the rat exocrine pancreas. FEBS Lett. 1990, 268, 79–82. Grond, S.; Meuser, T.; Uragg, H.; Stahlberg, H.J.; Lehmann, K.A. Serum concentrations of tramadol enantiomers during patient-controlled analgesia. Br. J. Clin. Pharmacol. 1999, 48, 254-257.
  115. Hu, H.; Tian, J.; Zhu, Y.; Wang, C.; Xiao, R.; Herz, J.M.; Wood, J.D.; Zhu, M.X. Activation of TRPA1 channels by fenamate nonsteroidal anti-inflammatory drugs. Pflügers Arch. 2010, 459, 579–592. Brodin, P.; Skoglund, L.A. Dose-response inhibition of rat compound nerve action potential by dextropropoxyphene and codeine compared to morphine and cocaine in vitro. Gen. Pharmacol. 1990, 21, 551-553.
  116. Tatematsu, Y.; Hayashi, H.; Taguchi, R.; Fujita, H.; Yamamoto, A.; Ohkura, K. Effect of N-phenylanthranilic acid scaffold nonsteroidal anti-inflammatory drugs on the mitochondrial permeability transition. Biol. Pharm. Bull. 2016, 39, 278–284. Hunter, E.G.; Frank, G.B. An opiate receptor on frog sciatic nerve axons. Can. J. Physiol. Pharmacol. 1979, 57, 1171-1174.
  117. Glass, J.S.; Hardy, C.L.; Meeks, N.M.; Carroll, B.T. Acute pain management in dermatology: Risk assessment and treatment. J. Am. Acad. Dermatol. 2015, 73, 543–560. Kumamoto, E.; Mizuta, K.; Fujita, T. Opioid actions in primary-afferent fibers – involvement in analgesia and anesthesia. Pharmaceuticals 2011, 4, 343-365.
  118. Bräu, M.E.; Nau, C.; Hempelmann, G.; Vogel, W. Local anesthetics potently block a potential insensitive potassium channel in myelinated nerve. J. Gen. Physiol. 1995, 105, 485-505.
  119. Bräu, M.E.; Vogel, W.; Hempelmann, G. Fundamental properties of local anesthetics: half-maximal blocking concentrations for tonic block of Na+ and K+ channels in peripheral nerve. Anesth. Analg. 1998, 87, 885-889.
  120. Tokuno, H.A.; Bradberry, C.W.; Everill, B.; Agulian, S.K.; Wilkes, S.; Baldwin, R.M.; Tamagnan, G.D.; Kocsis, J.D. Local anesthetic effects of cocaethylene and isopropylcocaine on rat peripheral nerves. Brain Res. 2004, 996, 159-167.
  121. Chen, Z.R.; Irvine, R.J.; Somogyi, A.A.; Bochner, F. Mu receptor binding of some commonly used opioids and their metabolites. Life Sci. 1991, 48, 2165-2171.
  122. Mizuta, K.; Fujita, T.; Kumamoto, E. Inhibition by morphine and its analogs of action potentials in adult rat dorsal root ganglion neurons. J. Neurosci. Res. 2012, 90, 1830-1841.
  123. Staiman, A.; Seeman, P. The impulse-blocking concentrations of anesthetics, alcohols, anticonvulsants, barbiturates, and narcotics on phrenic and sciatic nerves. Can. J. Physiol. Pharmacol. 1974, 52, 535-550.
  124. Scholz, A. Mechanisms of (local) anaesthetics on voltage-gated sodium and other ion channels. Br. J. Anaesth. 2002, 89, 52-61.
  125. Hu, S.; Rubly, N. Effects of morphine on ionic currents in frog node of Ranvier. Eur. J. Pharmacol. 1983, 95, 185-192.
  126. Frazier, D.T.; Murayama, K.; Abbott, N.J.; Narahashi, T. Effects of morphine on internally perfused squid giant axons. Proc. Soc. Exp. Biol. Med. 1972, 139, 434-438.
  127. Wagner, L.E. II; Eaton, M.; Sabnis, S.S.; Gingrich, K.J. Meperidine and lidocaine block of recombinant voltage-dependent Na+ channels: evidence that meperidine is a local anesthetic. Anesthesiology 1999, 91, 1481-1490.
  128. Gutstein, H.B.; Akil, H. Opioid analgesics. In: Goodman & Gilman’s The Pharmacological Basis of Therapeutics, 11th ed.; Brunton, L.L.; Lazo, J.S.; Parker, K.L., Eds.; McGraw-Hill, Medical Publishing Division: New York, U.S.A., 2006, pp. 547-590.
  129. Viel, E.J.; Eledjam, J.J.; De La Coussaye, J.E.; D'Athis, F. Brachial plexus block with opioids for postoperative pain relief: comparison between buprenorphine and morphine. Reg. Anesth. 1989, 14, 274-278.
  130. King, M.; Su, W.; Chang, A.; Zuckerman, A.; Pasternak, G.W. Transport of opioids from the brain to the periphery by P-glycoprotein: peripheral actions of central drugs. Nature Neurosci. 2001, 4, 268-274.
  131. Mays, K.S.; Lipman, J.J.; Schnapp, M. Local analgesia without anesthesia using peripheral perineural morphine injections. Anesth. Anal. 1987, 66, 417-420.
  132. Cleary, J.; Mikus, G.; Somogyi, A.; Bochner, F. The influence of pharmacogenetics on opioid analgesia: studies with codeine and oxycodone in the Sprague-Dawley/Dark Agouti rat model. J. Pharmacol. Exp. Therap. 1994, 271, 1528-1534.
  133. Mikus, G.; Somogyi, A.A.; Bochner, F.; Eichelbaum, M. Codeine O-demethylation: rat strain differences and the effects of inhibitors. Biochem. Pharmacol. 1991, 41, 757-762.
  134. Hermanns, H.; Hollmann, M.W.; Stevens, M.F.; Lirk, P.; Brandenburger, T.; Piegeler, T.; Werdehausen, R. Molecular mechanisms of action of systemic lidocaine in acute and chronic pain: a narrative review. Br. J. Anaesth. 2019, 123, 335-349.
  135. Hille, B. Ionic Channels of Excitable Membranes; Sinauer Associates Inc.: Massachusetts, 1984.
  136. Kirillova, I.; Teliban, A.; Gorodetskaya, N.; Grossmann, L.; Bartsch, F.; Rausch, V.H.; Struck, M.; Tode, J.; Baron, R.; Jänig, W. Effect of local and intravenous lidocaine on ongoing activity in injured afferent nerve fibers. Pain 2011, 152, 1562-1571.
  137. Shin, J.W.; Pancaro, C.; Wang, C.F.; Gerner, P. Low-dose systemic bupivacaine prevents the development of allodynia after thoracotomy in rats. Anesth. Analg. 2008, 107, 1587-1591.
  138. Delorme, C.; Navez, M.L.; Legout, V.; Deleens, R.; Moyse, D. Treatment of neuropathic pain with 5% lidocaine-medicated plaster: five years of clinical experience. Pain Res. Manag. 2011, 16, 259-263.
  139. Kalso, E.; Tramèr, M.R.; McQuay, H.J.; Moore, R.A. Systemic local-anaesthetic-type drugs in chronic pain: a systematic review. Eur. J. Pain 1998, 2, 3-14.
  140. Tremont-Lukats, I.W.; Challapalli, V.; McNicol, E.D.; Lau, J.; Carr, D.B. Systemic administration of local anesthetics to relieve neuropathic pain: a systematic review and meta-analysis. Anesth. Analg. 2005, 101, 1738-1749.
  141. Zhu, B.; Zhou, X.; Zhou, Q.; Wang, H.; Wang, S.; Luo, K. Intra-venous lidocaine to relieve neuropathic pain: a systematic review and meta-analysis. Front. Neurol. 2019, 10, 954.
  142. Piao, L.-H.; Fujita, T.; Jiang, C.-Y.; Liu, T.; Yue, H.-Y.; Nakatsuka, T.; Kumamoto, E. TRPA1 activation by lidocaine in nerve terminals results in glutamate release increase. Biochem. Biophys. Res. Commun. 2009, 379, 980-984.
  143. Piao, L.-H.; Fujita, T.; Yu, T.; Kumamoto, E. Presynaptic facilitation by tetracaine of glutamatergic spontaneous excitatory transmission in the rat spinal substantia gelatinosa – involvement of TRPA1 channels. Brain Res. 2017, 1657, 245-252.
  144. Leffler, A.; Fischer, M.J.; Rehner, D.; Kienel, S.; Kistner, K.; Sauer, S.K.; Gavva, N.R.; Reeh, P.W.; Nau, C. The vanilloid receptor TRPV1 is activated and sensitized by local anesthetics in rodent sensory neurons. J. Clin. Invest. 2008, 118, 763-776.
  145. Leffler, A.; Lattrell, A.; Kronewald, S.; Niedermirtl, F.; Nau, C. Activation of TRPA1 by membrane permeable local anesthetics. Mol. Pain 2011, 7, 62.
  146. Oda, A.; Iida, H.; Tanahashi, S.; Osawa, Y.; Yamaguchi, S.; Dohi, S. Effects of α2-adrenoceptor agonists on tetrodotoxin-resistant Na+ channels in rat dorsal root ganglion neurons. Eur. J. Anaesthesiol. 2007, 24, 934-941.
  147. Zhao, K.; Dong, Y.; Su, G.; Wang, T.; Ji, T.; Wu, N.; Cui, X.; Li, W.; Yang, Y.; Chen, X. Effect of systemic lidocaine on postoperative early recovery quality in patients undergoing supratentorial tumor resection. Drug Des. Devel. Ther. 2022, 16, 1171–1181.
  148. Chan, V.W.S.; Weisbrod, M.J.; Kaszas, Z.; Dragomir, C. Comparison of ropivacaine and lidocaine for intravenous regional anesthesia in volunteers: a preliminary study on anesthetic efficacy and blood level. Anesthesiology 1999, 90, 1602-1608.
  149. McClellan, K.J.; Faulds, D. Ropivacaine: an update of its use in regional anaesthesia. Drugs 2000, 60, 1065-1093.
  150. Bader, A.M.; Datta, S.; Flanagan, H.; Covino, B.G. Comparison of bupivacaine- and ropivacaine-induced conduction blockade in the isolated rabbit vagus nerve. Anesth. Analg. 1989, 68, 724-727.
  151. Yilmaz-Rastoder, E.; Gold, M.S.; Hough, K.A.; Gebhart, G.F.; Williams, B.A. Effect of adjuvant drugs on the action of local anesthetics in isolated rat sciatic nerves. Reg. Anesth. Pain Med. 2012, 37, 403-409.
  152. Lee-Son, S.; Wang, G.K.; Concus, A.; Crill, E.; Strichartz, G. Stereoselective inhibition of neuronal sodium channels by local anesthetics. Evidence for two sites of action? Anesthesiology 1992, 77, 324-335.
  153. Foster, R.H.; Markham, A. Levobupivacaine: a review of its pharmacology and use as a local anaesthetic. Drugs 2000, 59, 551-579.
  154. Vladimirov, M.; Nau, C.; Mok, W.M.; Strichartz, G. Potency of bupivacaine stereoisomers tested in vitro and in vivo: biochemical, electrophysiological, and neurobehavioral studies. Anesthesiology 2000, 93, 744-755.
  155. Stoetzer, C.; Martell, C.; de la Roche, J.; Leffler, A. Inhibition of voltage-gated Na+ channels by bupivacaine is enhanced by the adjuvants buprenorphine, ketamine, and clonidine. Reg. Anesth. Pain Med. 2017, 42, 462-468.
  156. Gerner, P.; Mujtaba, M.; Sinnott, C.J.; Wang, G.K. Amitriptyline versus bupivacaine in rat sciatic nerve blockade. Anesthesiology 2001, 94, 661-667.
  157. Bedford, J.A.; Turner, C.E.; Elsohly, H.N. Local anesthetic effects of cocaine and several extracts of the coca leaf (E. coca). Pharmacol. Biochem. Behav. 1984, 20, 819-821.
  158. Pagala, M.K.D.; Venkatachari, S.A.T.; Herzlich, B.; Ravindran, K.; Namba, T.; Grob, D. Effect of cocaine on responses of mouse phrenic nerve-diaphragm preparation. Life Sci. 1991, 48, 795-802.
  159. Carney, T.P. Alkaloids as local anesthetics. In: The Alkaloids, Vol. 5; Manske, R.H.F., Ed.; Academic Press: New York, U.S.A. 1955, pp. 211-227.
  160. Matthews, J.C.; Collins, A. Interactions of cocaine and cocaine congeners with sodium channels. Biochem. Pharmacol. 1983, 32, 455-460.
  161. O'Leary, M.E., Chahine, M. Cocaine binds to a common site on open and inactivated human heart (Nav1.5) sodium channels. J. Physiol. 2002, 541, 701-716.
  162. Liu, D.; Hariman, R.J.; Bauman, J.L. Cocaine concentration-effect relationship in the presence and absence of lidocaine: evidence of competitive binding between cocaine and lidocaine. J. Pharmacol. Exp. Therap. 1996, 276, 568-577.
  163. Chen, Y.-H.; Lin, C.-H.; Lin, P.-L.; Tsai, M.-C. Cocaine elicits action potential bursts in a central snail neuron: the role of delayed rectifying K+ current. Neuroscience 2006, 138, 257-280.
  164. Štolc, S.; Mai, P.-M. Comparison of local anesthetic activity of pentacaine (trapencaine) and some of its derivatives by three different techniques. Pharmazie 1993, 48, 210-212.
  165. Butterworth, J.F. IV; Lief, P.A.; Strichartz, G.R. The pH-dependent local anesthetic activity of diethylaminoethanol, a procaine metabolite. Anesthesiology 1988, 68, 501-506.
  166. Ribeiro, J.A.; Sebastião, A.M. Antagonism of tetrodotoxin- and procaine-induced axonal blockade by adenine nucleotides in the frog sciatic nerve. Br. J. Pharmacol. 1984, 81, 277-282.
  167. Kalichman, M.W.; Moorhouse, D.F.; Powell, H.C.; Myers, R.R. Relative neural toxicity of local anesthetics. J. Neuropathol. Exp. Neurol. 1993, 52, 234-240.
  168. Lee, H.-S. Recent advances in topical anesthesia. J. Dent. Anesth. Pain Med. 2016, 16, 237-244.
  169. Thygesen, M.M.; Rasmussen, M.M.; Madsen, J.G.; Pedersen, M.; Lauridsen, H. Propofol (2, 6‐diisopropylphenol) is an applicable immersion anesthetic in the axolotl with potential uses in hemodynamic and neurophysiological experiments. Regeneration 2017, 4, 124-131.
  170. Guénette, S.A.; Giroux, M.-C.; Vachon, P. Pain perception and anaesthesia in research frogs. Exp. Anim. 2013, 62, 87-92.
  171. Vanable, J.W. Benzocaine: an excellent amphibian anesthetic. Axolotl Newsletter 1985, 14, 19-21.
  172. Starke, K.; Wagner, J.; Schümann, H.J. Adrenergic neuron blockade by clonidine: comparison with guanethidine and local anesthetics. Arch. Int. Pharmacodyn. 1972, 195, 291-308.
  173. Gissen, A.J.; Covino, B.G.; Gregus, J. Differential sensitivities of mammalian nerve fibers to local anesthetic agents. Anesthesiology 1980, 53, 467-474.
  174. Macdonald, R.L. Cellular effects of antiepileptic drugs. In Epilepsy: A Comprehensive Textbook; Engel, J. Jr.; Pedley, T.A., Eds.; Lippincon-Raven Publishers; Philadelphia, U.S.A., 1997, pp. 1383-1391.
  175. Kubacka, M.; Rapacz, A.; Sałat, K.; Filipek, B.; Cios, A.; Pociecha, K.; Wyska, E.; Hubicka, U.; Żuromska-Witek, B.; Kwiecień, A.; Marona, H.; Waszkielewicz, A.M. KM-416, a novel phenoxyalkylaminoalkanol derivative with anticonvulsant properties exerts analgesic, local anesthetic, and antidepressant-like activities. Pharmacodynamic, pharmacokinetic, and forced degradation studies. Eur. J. Pharmacol. 2020, 886, 173540.
  176. Xie, X.; Dale, T.J.; John, V.H.; Cater, H.L.; Peakman, T.C.; Clare, J.J. Electrophysiological and pharmacological properties of the human brain type IIA Na+ channel expressed in a stable mammalian cell line. Pflügers Arch. 2001, 441, 425-433.
  177. McLean, M.J.; Macdonald, R.L. Carbamazepine and 10,11-epoxycarbamazepine produce use- and voltage-dependent limitation of rapidly firing action potentials of mouse central neurons in cell culture. J. Pharmacol. Exp. Ther. 1986, 238, 727-738.
  178. Cruccu, G.; Gronseth, G.; Alksne, J.; Argoff, C.; Brainin, M.; Burchiel, K.; Nurmikko, T.; Zakrzewska, J.M. AAN-EFNS guidelines on trigeminal neuralgia management. Eur. J. Neurol. 2008, 15, 1013-1028.
  179. Vargas-Espinosa, M.-L., Sanmartí-García, G.; Vázquez-Delgado, E.; Gay-Escoda, C. Antiepileptic drugs for the treatment of neuropathic pain: a systematic review. Med. Oral Patol. Oral Cir. Bucal 2012, 17, e786-e793.
  180. Lang, D.G.; Wang, C.M.; Cooper, B.R. Lamotrigine, phenytoin and carbamazepine interactions on the sodium current present in N4TG1 mouse neuroblastoma cells. J. Pharmacol. Exp. Ther. 1993, 266, 829-835.
  181. Benes, J.; Parada A.; Figueiredo, A.A.; Alves, P.C.; Freitas, A.P.; Learmonth, D.A.; Cunha, R.A.; Garrett, J.; Soares-da-Silva, P. Anticonvulsant and sodium channel-blocking properties of novel 10,11-dihydro-5H-dibenz[b,f]azepine-5- carboxamide derivatives. J. Med. Chem. 1999, 42, 2582-2587.
  182. Huang, C.-W.; Huang, C.-C.; Lin, M.-W.; Tsai, J.-J.; Wu, S.-N. The synergistic inhibitory actions of oxcarbazepine on voltage-gated sodium and potassium currents in differentiated NG108-15 neuronal cells and model neurons. Int. J. Neuropsychopharmacol. 2008, 11, 597-610.
  183. Kuo, C.-C. A common anticonvulsant binding site for phenytoin, carbamazepine, and lamotrigine in neuronal Na+ channels. Mol. Pharmacol. 1998, 54, 712-721.
  184. Molnár, P.; Erdö, S.L. Vinpocetine is as potent as phenytoin to block voltage-gated Na+ channels in rat cortical neurons. Eur. J. Pharmacol. 1995, 273, 303-306.
  185. Qiao, X.; Sun, G.; Clare, J.J.; Werkman, T.R.; Wadman, W.J. Properties of human brain sodium channel α-subunits expressed in HEK293 cells and their modulation by carbamazepine, phenytoin and lamotrigine. Br. J. Pharmacol. 2014, 171, 1054-1067.
  186. Neumcke, B.; Schwarz, J.R.; Stämpfli, R. A comparison of sodium currents in rat and frog myelinated nerve: normal and modified sodium inactivation. J. Physiol. 1987, 382, 175-191.
  187. Maneuf, Y.P.; Gonzalez, M.I.; Sutton, K.S.; Chung, F.-Z.; Pinnock, R.D.; Lee, K. Cellular and molecular action of the putative GABA-mimetic, gabapentin. Cell. Mol. Life Sci. 2003, 60, 742-750.
  188. Chen, J.; Li, L.; Chen, S.-R.; Chen, H.; Xie, J.-D.; Sirrieh, R.E.; MacLean, D.M.; Zhang, Y.; Zhou, M.-H.; Jayaraman, V.; Pan, H.-L. The α2δ-1-NMDA Receptor complex Is critically involved in neuropathic pain development and gabapentin therapeutic actions. Cell Rep. 2018, 22, 2307-2321.
  189. Zona, C.; Ciotti, M.T.; Avoli, M. Topiramate attenuates voltage-gated sodium currents in rat cerebellar granule cells. Neurosci. Lett. 1997, 231, 123-126.
  190. Curia, G.; Aracri, P.; Colombo, E.; Scalmani, P.; Mantegazza, M.; Avanzini, G.; Franceschetti, S. Phosphorylation of sodium channels mediated by protein kinase-C modulates inhibition by topiramate of tetrodotoxin-sensitive transient sodium current. Br. J. Pharmacol. 2007, 150, 792-797.
  191. Chapman, A.; Keane, P.E.; Meldrum, B.S.; Simiand, J.; Vernieres, J.C. Mechanism of anticonvulsant action of valproate. Prog. Neurobiol. 1982, 19, 315-359.
  192. Perucca, E. A pharmacological and clinical review on topiramate, a new antiepileptic drug. Pharmacol. Res. 1997, 35, 241-256.
  193. Braga, M.F.M.; Aroniadou-Anderjaska, V.; Li, H.; Rogawski, M.A. Topiramate reduces excitability in the basolateral amygdala by selectively inhibiting GluK1 (GluR5) kainate receptors on interneurons and positively modulating GABAA receptors on principal neurons. J. Pharmacol. Exp. Ther. 2009, 330, 558-566.
  194. Lee, C.-Y.; Fu, W.-M.; Chen, C.-C.; Su, M.-J.; Liou, H.-H. Lamotrigine inhibits postsynaptic AMPA receptor and glutamate release in the dentate gyrus. Epilepsia 2008, 49, 888-897.
  195. Blackburn-Munro, G.; Ibsen, N.; Erichsen, H.K. A comparison of the anti-nociceptive effects of voltage-activated Na+ channel blockers in the formalin test. Eur. J. Pharmacol. 2002, 445, 231-238.
  196. Shannon, H.E.; Eberle, E.L.; Peters, S.C. Comparison of the effects of anticonvulsant drugs with diverse mechanisms of action in the formalin test in rats. Neuropharmacology 2005, 48, 1012-1020.
  197. Douglas-Hall, P.; Dzahini, O.; Gaughran, F.; Bile, A.; Taylor, D. Variation in dose and plasma level of lamotrigine in patients discharged from a mental health trust. Ther. Adv. Psychopharmacol. 2017, 7, 17-24.
  198. Morselli, P.L. Carbamazepine: absorption, distribution, and excretion. In: Antiepileptic Drugs, 4th ed.; Levy, R.H.; Mattson, R.H.; Meldrum, B.S., Eds.; Raven Press: New York, U.S.A., 1995, pp. 515-528.
  199. Ardid, D.; Jourdan, D.; Mestre, C.; Villanueva, L.; Le Bars, D.; Eschalier, A. Involvement of bulbospinal pathways in the antinociceptive effect of clomipramine in the rat. Brain Res. 1995, 695, 253-256.
  200. Max, M.B.; Lynch, S.A.; Muir, J.; Shoaf, S.E.; Smoller, B.; Dubner, R. Effects of desipramine, amitriptyline, and fluoxetine on pain in diabetic neuropathy. N. Engl. J. Med. 1992, 326, 1250-1256.
  201. Anjaneyulu, M.; Chopra, K. Possible involvement of cholinergic and opioid receptor mechanisms in fluoxetine mediated antinociception response in streptozotocin-induced diabetic mice. Eur. J. Pharmacol. 2006, 538, 80-84.
  202. Cervantes-Durán, C.; Rocha-González, H.I.; Granados-Soto, V. Peripheral and spinal 5-HT receptors participate in the pronociceptive and antinociceptive effects of fluoxetine in rats. Neuroscience 2013, 252, 396-409.
  203. Ghelardini, C.; Galeotti, N.; Bartolini, A. Antinociception induced by amitriptyline and imipramine is mediated by α2A-adrenoceptors. Jpn. J. Pharmacol. 2000, 82, 130-137.
  204. Hall, H.; Ögren, S.-O. Effects of antidepressant drugs on different receptors in the brain. Eur. J. Pharmacol. 1981, 70, 393-407.
  205. O’Donnell, J.M.; Shelton, R.C. Drug therapy of depression and anxiety disorders. In Goodman & Gilman’s The Pharmacological Basis of Therapeutics, 12th ed.; Brunton, L.L.; Chabner, B.A.; Knollmann, B.C., Eds.; McGraw-Hill, Medical Publishing Division: New York, U.S.A., 2011, pp. 397-415.
  206. Wong, D.T.; Bymaster, F.P. Dual serotonin and noradrenaline uptake inhibitor class of antidepressants - potential for greater efficacy or just hype? Prog. Drug Res. 2002, 58, 169-222.
  207. Traboulsie, A.; Chemin, J.; Kupfer, E.; Nargeot, J., Lory, P. T-type calcium channels are inhibited by fluoxetine and its metabolite norfluoxetine. Mol. Pharmacol. 2006, 69, 1963-1968.
  208. Wu, W.; Ye, Q.; Wang, W.; Yan, L.; Wang, Q.; Xiao, H.; Wan, Q. Amitriptyline modulates calcium currents and intracellular calcium concentration in mouse trigeminal ganglion neurons. Neurosci. Lett. 2012, 506, 307-311.
  209. Reynolds, I.J.; Miller, R.J. Tricyclic antidepressants block N-methyl-D-aspartate receptors: similarities to the action of zinc. Br. J. Pharmacol. 1988, 95, 95-102.
  210. Sernagor, E.; Kuhn, D.; Vyklicky, L. Jr.; Mayer, M.L. Open channel block of NMDA receptor responses evoked by tricyclic antidepressants. Neuron 1989, 2, 1221-1227.
  211. Watanabe, Y.; Saito, H.; Abe, K. Tricyclic antidepressants block NMDA receptor-mediated synaptic responses and induction of long-term potentiation in rat hippocampal slices. Neuropharmacology 1993, 32, 479-486.
  212. Barygin, O.I.; Nagaeva, E.I.; Tikhonov, D.B.; Belinskaya, D.A.; Vanchakova, N.P.; Shestakova, N.N. Inhibition of the NMDA and AMPA receptor channels by antidepressants and antipsychotics. Brain Res. 2017, 1660, 58-66.
  213. Nagata, K.; Imai, T.; Yamashita, T.; Tsuda, M.; Tozaki-Saitoh, H.; Inoue, K. Antidepressants inhibit P2X4 receptor function: a possible involvement in neuropathic pain relief. Mol. Pain 2009, 5, 20.
  214. Kremer, M.; Yalcin, I.; Goumon, Y.; Wurtz, X.; Nexon, L.; Daniel, D.; Megat, S.; Ceredig, R.A.; Ernst, C.; Turecki, G.; Chavant, V.; Théroux, J.-F.; Lacaud, A.; Joganah, L.-E.; Lelievre, V.; Massotte, D.; Lutz, P.-E.; Gilsbach, R.; Salvat, E.; Barrot, M. A dual noradrenergic mechanism for the relief of neuropathic allodynia by the antidepressant drugs duloxetine and amitriptyline. J. Neurosci. 2018, 38, 9934-9954.
  215. Le Cudennec, C.; Castagné, V. Face-to-face comparison of the predictive validity of two models of neuropathic pain in the rat: analgesic activity of pregabalin, tramadol and duloxetine. Eur. J. Pharmacol. 2014, 735, 17-25.
  216. Wong, D.T.; Bymaster, F.P.; Mayle, D.A.; Reid, L.R.; Krushinski, J.H.; Robertson, D.W. LY248686, a new inhibitor of serotonin and norepinephrine uptake. Neuropsychopharmacology 1993, 8, 23-33.
  217. Russell, I.J.; Mease, P.J.; Smith, T.R.; Kajdasz, D.K.; Wohlreich, M.M.; Detke, M.J.; Walker, D.J.; Chappell, A.S.; Arnold, L.M. Efficacy and safety of duloxetine for treatment of fibromyalgia in patients with or without major depressive disorder: results from a 6-month, randomized, double-blind, placebo-controlled, fixed-dose trial. Pain 2008, 136, 432-444.
  218. Müller, N.; Schennach, R.; Riedel, M.; Möller, H.-J. Duloxetine in the treatment of major psychiatric and neuropathic disorders. Expert Rev. Neurother. 2008, 8, 527-536.
  219. Stark, P.; Fuller, R.W.; Wong, D.T. The pharmacologic profile of fluoxetine. J. Clin. Psychiatry 1985, 46, 7-13.
  220. Korzeniewska-Rybicka, I.; Płaźnik, A. Analgesic effect of antidepressant drugs. Pharmacol. Biochem. Behav. 1998, 59, 331-338.
  221. Richeimer, S.H.; Bajwa, Z.H.; Kahraman, S.S.; Ransil, B.J.; Warfield, C.A. Utilization patterns of tricyclic antidepressants in a multidisciplinary pain clinic: a survey. Clin. J. Pain 1997, 13, 324-329.
  222. Okuda, K.; Takanishi, T.; Yoshimoto, K.; Ueda, S. Trazodone hydrochloride attenuates thermal hyperalgesia in a chronic constriction injury rat model. Eur. J. Anaesthesiol. 2003, 20, 409-415.
  223. Richelson, E.; Pfenning, M. Blockade by antidepressants and related compounds of biogenic amine uptake into rat brain synaptosomes: most antidepressants selectively block norepinephrine uptake. Eur. J. Pharmacol. 1984, 104, 277-286.
  224. Schreiber, S.; Backer, M.M.; Herman, I.; Shamir, D.; Boniel, T.; Pick, C.G. The antinociceptive effect of trazodone in mice is mediated through both μ-opioid and serotonergic mechanisms. Behav. Brain Res. 2000, 114, 51-56.
  225. Davidoff, G.; Guarracini, M.; Roth, E.; Sliwa, J.; Yarkony, G. Trazodone hydrochloride in the treatment of dysesthetic pain in traumatic myelopathy: a randomized, double-blind, placebo-controlled study. Pain 1987, 29, 151-161.
  226. Baastrup, C.; Finnerup, N.B. Pharmacological management of neuropathic pain following spinal cord injury. CNS Drugs 2008, 22, 455-475.
  227. Stoetzer, C.; Papenberg, B.; Doll, T.; Völker, M.; Heineke, J.; Stoetzer, M.; Wegner, F.; Leffler, A. Differential inhibition of cardiac and neuronal Na+ channels by the selective serotonin-norepinephrine reuptake inhibitors duloxetine and venlafaxine. Eur. J. Pharmacol. 2016, 783, 1-10.
  228. Wang, S.-Y.; Calderon, J.; Wang, G.K. Block of neuronal Na+ channels by antidepressant duloxetine in a state-dependent manner. Anesthesiology 2010, 113, 655-665.
  229. Pancrazio, J.J.; Kamatchi, G.L.; Roscoe, A.K.; Lynch, C. 3rd. Inhibition of neuronal Na+ channels by antidepressant drugs. J. Pharmacol. Exp. Ther. 1998, 284, 208-214.
  230. Ishii, Y.; Sumi, T. Amitriptyline inhibits striatal efflux of neurotransmitters via blockade of voltage-dependent Na+ channels. Eur. J. Pharmacol. 1992, 221, 377-380.
  231. Leffler, A.; Reiprich, A.; Mohapatra, D.P.; Nau, C. Use-dependent block by lidocaine but not amitriptyline is more pronounced in tetrodotoxin (TTX)-resistant Nav1.8 than in TTX-sensitive Na+ channels. J. Pharmacol. Exp. Ther. 2007, 320, 354-364.
  232. Nicholson, G.M.; Blanche, T.; Mansfield, K.; Tran, Y. Differential blockade of neuronal voltage-gated Na+ and K+ channels by antidepressant drugs. Eur. J. Pharmacol. 2002, 452, 35-48.
  233. Song, J.-H.; Ham, S.-S.; Shin, Y.-K.; Lee, C.-S. Amitriptyline modulation of Na+ channels in rat dorsal root ganglion neurons. Eur. J. Pharmacol. 2000, 401, 297-305.
  234. Wang, G.K.; Russell, C.; Wang, S.-Y. State-dependent block of voltage-gated Na+ channels by amitriptyline via the local anesthetic receptor and its implication for neuropathic pain. Pain 2004, 110, 166-174.
  235. Yan, L.; Wang, Q.; Fu, Q.; Ye, Q.; Xiao, H.; Wan, Q. Amitriptyline inhibits currents and decreases the mRNA expression of voltage-gated sodium channels in cultured rat cortical neurons. Brain Res. 2010, 1336, 1-9.
  236. Dick, I.E.; Brochu, R.M.; Purohit, Y.; Kaczorowski, G.J.; Martin, W.J.; Priest, B.T. Sodium channel blockade may contribute to the analgesic efficacy of antidepressants. J. Pain 2007, 8, 315-324.
  237. Liang, J.; Liu, X.; Pan, M.; Dai, W.; Dong, Z.; Wang, X.; Liu, R.; Zheng, J.; Yu, S. Blockade of Nav1.8 currents in nociceptive trigeminal neurons contributes to anti-trigeminovascular nociceptive effect of amitriptyline. Neuromolecular Med. 2014, 16, 308-321.
  238. Catterall, W.A.; Mackie, K. Local anesthetics. In: Goodman & Gilman’s The Pharmacological Basis of Therapeutics, 12th ed.; Brunton, L.L.; Chabner, B.A.; Knollmann, B.C., Eds.; McGraw-Hill, Medical Publishing Division: New York, U.S.A., 2011, pp. 565-582.
  239. Caruso, R.; Ostuzzi, G.; Turrini, G.; Ballette, F.; Recla, E.; DallʼOlio, R.; Croce, E.; Casoni, B.; Grassi, L.; Barbui, C. Beyond pain: can antidepressants improve depressive symptoms and quality of life in patients with neuropathic pain? A systematic review and meta-analysis. Pain 2019, 160, 2186-2198.
  240. Mika, J.; Zychowska, M.; Makuch, W.; Rojewska, E.; Przewlocka, B. Neuronal and immunological basis of action of antidepressants in chronic pain - clinical and experimental studies. Pharmacol. Rep. 2013, 65, 1611-1621.
  241. Banafshe, H.R.; Hajhashemi, V.; Minaiyan, M.; Mesdaghinia, A.; Abed, A. Antinociceptive effects of maprotiline in a rat model of peripheral neuropathic pain: possible involvement of opioid system. Iran J. Basic Med. Sci. 2015, 18, 752-757.
  242. Bhana, N.; Goa, K.L.; McClellan, K.J. Dexmedetomidine. Drugs 2000, 59, 263-268.
  243. Costa-Pereira, J.T.; Ribeiro, J.; Martins, I.; Tavares, I. Role of spinal cord α2-adrenoreceptors in noradrenergic inhibition of nociceptive transmission during chemotherapy-induced peripheral neuropathy. Front. Neurosci. 2020, 13, 1413.
  244. Fisher, B.; Zornow, M.H.; Yaksh, T.L.; Peterson, B.M. Antinociceptive properties of intrathecal dexmedetomidine in rats. Eur. J. Pharmacol. 1991, 192, 221-225.
  245. Takano, Y.; Yaksh, T.L. Relative efficacy of spinal alpha-2 agonists, dexmedetomidine, clonidine and ST-91, determined in vivo by using N-ethoxycarbonyl-2-ethoxy-1,2-dihydroquinoline, an irreversible antagonist. J. Pharmacol. Exp. Ther. 1991, 258, 438-446.
  246. Filos, K.S.; Goudas, L.C.; Patroni, O.; Polyzou, V. Hemodynamic and analgesic profile after intrathecal clonidine in humans. A dose-response study. Anesthesiology 1994, 81, 591-601.
  247. Sullivan, A.F.; Kalso, E.A.; McQuay, H.J.; Dickenson, A.H. The antinociceptive actions of dexmedetomidine on dorsal horn neuronal responses in the anaesthetized rat. Eur. J. Pharmacol. 1992, 215, 127-133.
  248. Brummett, C.M.; Norat, M.A.; Palmisano, J.M.; Lydic, R. Perineural administration of dexmedetomidine in combination with bupivacaine enhances sensory and motor blockade in sciatic nerve block without inducing neurotoxicity in rat. Anesthesiology 2008, 109, 502-511.
  249. Calasans-Maia, J.A.; Zapata-Sudo, G.; Sudo, R.T. Dexmedetomidine prolongs spinal anaesthesia induced by levobupivacaine 0.5% in guinea-pigs. J. Pharm. Pharmacol. 2005, 57, 1415-1420.
  250. Tsutsui, Y.; Sunada, K. Adding dexmedetomidine to articaine increases the latency of thermal antinociception in rats. Anesth. Prog. 2020, 67, 72–78.
  251. Marolf, V.; Ida, K.K.; Siluk, D.; Struck-Lewicka, W.; Markuszewski, M.J.; Sandersen, C. Effects of perineural administration of ropivacaine combined with perineural or intravenous administration of dexmedetomidine for sciatic and saphenous nerve blocks in dogs. Am. J. Vet. Res. 2021, 82, 449-458.
  252. Kanazi, G.E.; Aouad, M.T.; Jabbour-Khoury, S.I.; Al Jazzar, M.D.; Alameddine, M.M.; Al-Yaman, R.; Bulbul, M.; Baraka, A.S. Effect of low-dose dexmedetomidine or clonidine on the characteristics of bupivacaine spinal block. Acta Anaesthesiol. Scand. 2006, 50, 222-227.
  253. Madan, R.; Bharti, N.; Shende, D.; Khokhar, S.K.; Kaul, H.L. A dose response study of clonidine with local anesthetic mixture for peribulbar block: a comparison of three doses. Anesth. Analg. 2001, 93, 1593-1597.
  254. Memiş, D.; Turan, A.; Karamanlioĝlu, B.; Pamukçu, Z.; Kurt, I. Adding dexmedetomidine to lidocaine for intravenous regional anesthesia. Anesth. Analg. 2004, 98, 835-840.
  255. Singelyn, F.J.; Gouverneur, J.-M.; Robert, A. A minimum dose of clonidine added to mepivacaine prolongs the duration of anesthesia and analgesia after axillary brachial plexus block. Anesth. Analg. 1996, 83, 1046-1050.
  256. Tschernko, E.M.; Klepetko, H.; Gruber, E.; Kritzinger, M.; Klimscha, W.; Jandrasits, O.; Haider, W. Clonidine added to the anesthetic solution enhances analgesia and improves oxygenation after intercostal nerve block for thoracotomy. Anesth. Analg. 1998, 87, 107-111.
  257. Mohyiedin, H.; Kamelia, A.A.; Ekram, F.S.; Al Shaimaa, A.K. The effect of various additives to local anesthetics on the duration of analgesia of supraclavicular brachial plexus block. J. Anest. & Inten. Care Med. 2019, 9, 555756.
  258. Ouchi, K. Dexmedetomidine 2 ppm is appropriate for the enhancement effect of local anesthetic action of lidocaine in inferior alveolar nerve block: a preliminary, randomized cross-over study. Clin. J. Pain 2020, 36, 618-625.
  259. Sane, S.; Shokouhi, S.; Golabi, P.; Rezaeian, M.; Kazemi Haki, B. The effect of dexmedetomidine in combination with bupivacaine on sensory and motor block time and pain score in supraclavicular block. Pain Res. Manag. 2021, 2021, 8858312.
  260. Eisenach, J.C.; De Kock, M.; Klimscha, W. α2-Adrenergic agonists for regional anesthesia. A clinical review of clonidine (1984-1995). Anesthesiology 1996, 85, 655-674.
  261. Concepcion, M.; Maddi, R.; Francis, D.; Rocco, A.G.; Murray, E.; Covino, B.G. Vasoconstrictors in spinal anesthesia with tetracaine – a comparison of epinephrine and phenylephrine. Anesth. Analg. 1984, 63, 134-138.
  262. Vaida, G.T.; Moss, P.; Capan, L.M.; Turndorf, H. Prolongation of lidocaine spinal anesthesia with phenylephrine. Anesth. Analg. 1986, 65, 781-785.
  263. Gaumann, D.M.; Brunet, P.C.; Jirounek, P. Clonidine enhances the effects of lidocaine on C-fiber action potential. Anesth. Analg. 1992, 74, 719-725.
  264. Kawasaki, Y.; Kumamoto, E.; Furue, H.; Yoshimura, M. α2 Adrenoceptor-mediated presynaptic inhibition of primary afferent glutamatergic transmission in rat substantia gelatinosa neurons. Anesthesiology 2003, 98, 682-689.
  265. Pan, Y.-Z.; Li, D.-P.; Pan, H.-L. Inhibition of glutamatergic synaptic input to spinal lamina IIo neurons by presynaptic α2-adrenergic receptors. J. Neurophysiol. 2002, 87, 1938-1947.
  266. Butterworth, J.F. IV; Strichartz, G.R. The α2-adrenergic agonists clonidine and guanfacine produce tonic and phasic block of conduction in rat sciatic nerve fibers. Anesth. Analg. 1993, 76, 295-301.
  267. Ishii, H.; Kohno, T.; Yamakura, T.; Ikoma, M.; Baba, H. Action of dexmedetomidine on the substantia gelatinosa neurons of the rat spinal cord. Eur. J. Neurosci. 2008, 27, 3182-3190.
  268. Yoshitomi, T.; Kohjitani, A.; Maeda, S.; Higuchi, H.; Shimada, M.; Miyawaki, T. Dexmedetomidine enhances the local anesthetic action of lidocaine via an α-2A adrenoceptor. Anesth. Analg. 2008, 107, 96-101.
  269. Mohamed, S.A.; Sayed, D.M.; El Sherif, F.A.; Abd El-Rahman, A.M. Effect of local wound infiltration with ketamine versus dexmedetomidine on postoperative pain and stress after abdominal hysterectomy, a randomized trial. Eur. J. Pain 2018, 22, 951-960.
  270. Coughlan, M.G.; Lee, J.G.; Bosnjak, Z.J.; Schmeling, W.T.; Kampine, J.P.; Warltier, D.C. Direct coronary and cerebral vascular responses to dexmedetomidine. Significance of endogenous nitric oxide synthesis. Anesthesiology 1992, 77, 998-1006.
  271. Correa-Sales, C.; Rabin, B.C.; Maze, M. A hypnotic response to dexmedetomidine, an α2 agonist, is mediated in the locus coeruleus in rats. Anesthesiology 1992, 76, 948-952.
  272. Sjöholm, B.; Voutilainen, R.; Luomala, K.; Savola, J.-M.; Scheinin, M. Characterization of [3H]atipamezole as a radioligand for α2-adrenoceptors. Eur. J. Pharmacol. 1992, 215, 109-117.
  273. MacDonald, E.; Kobilka, B.K.; Scheinin, M. Gene targeting-homing in on α2-adrenoceptor-subtype function. Trends Pharmacol. Sci. 1997, 18, 211-219.
  274. Virtanen, R. Pharmacological profiles of medetomidine and its antagonist, atipamezole. Acta Vet. Scand. 1989, 85, 29-37.
  275. Bylund, D.B.; Eikenberg, D.C.; Hieble, J.P.; Langer, S.Z.; Lefkowitz, R.J.; Minneman, K.P.; Molinoff, P.B.; Ruffolo, R.R.; Trendelenburg, U. International union of pharmacology nomenclature of adrenoceptors. Pharmacol. Rev. 1994, 46, 121-136.
  276. Chen, B.-S.; Peng, H.; Wu, S.-N. Dexmedetomidine, an α2-adrenergic agonist, inhibits neuronal delayed-rectifier potassium current and sodium current. Br. J. Anaesth. 2009, 103, 244-254.
  277. Ebert, T.J.; Hall, J.E.; Barney, J.A.; Uhrich, T.D.; Colinco, M.D. The effects of increasing plasma concentrations of dexmedetomidine in humans. Anesthesiology 2000, 93, 382-394.
  278. Slingsby, L.S.; Taylor, P.M. Thermal antinociception after dexmedetomidine administration in cats: a dose-finding study. J. Vet. Pharmacol. Therap. 2008, 31, 135-142.
  279. Bharti, N.; Sardana, D.K.; Bala, I. The analgesic efficacy of dexmedetomidine as an adjunct to local anesthetics in supraclavicular brachial plexus block: a randomized controlled trial. Anesth. Analg. 2015, 121, 1655-1660.
  280. Nassar, M.A.; Stirling, L.C.; Forlani, G.; Baker, M.D.; Matthews, E.A.; Dickenson, A.H.; Wood, J.N. Nociceptor-specific gene deletion reveals a major role for Nav1.7 (PN1) in acute and inflammatory pain. Proc. Natl. Acad. Sci. USA 2004, 101, 12706-12711.
  281. Patapoutian, A.; Tate, S.; Woolf, C.J. Transient receptor potential channels: targeting pain at the source. Nat. Rev. Drug Discov. 2009, 8, 55-68.
  282. Julius, D. TRP channels and pain. Annu. Rev. Cell Dev. Biol. 2013, 29, 355-384.
  283. Yang, K; Kumamoto, E.; Furue, H.; Yoshimura, M. Capsaicin facilitates excitatory but not inhibitory synaptic transmission in substantia gelatinosa of the rat spinal cord. Neurosci. Lett. 1998, 255, 135-138.
  284. Kosugi, M.; Nakatsuka, T.; Fujita, T.; Kuroda, Y.; Kumamoto, E. Activation of TRPA1 channel facilitates excitatory synaptic transmission in substantia gelatinosa neurons of the adult rat spinal cord. J. Neurosci. 2007, 27, 4443-4451.
  285. Inoue, M.; Fujita, T.; Goto, M.; Kumamoto, E. Presynaptic enhancement by eugenol of spontaneous excitatory transmission in rat spinal substantia gelatinosa neurons is mediated by transient receptor potential A1 channels. Neuroscience 2012, 210, 403-415.
  286. Luo, Q.-T.; Fujita, T.; Jiang, C.-Y.; Kumamoto, E. Carvacrol presynaptically enhances spontaneous excitatory transmission and produces outward current in adult rat spinal substantia gelatinosa neurons. Brain Res. 2014, 1592, 44-54.
  287. Xu, Z.-H.; Wang, C.; Fujita, T.; Jiang, C.-Y.; Kumamoto, E. Action of thymol on spontaneous excitatory transmission in adult rat spinal substantia gelatinosa neurons. Neurosci. Lett. 2015, 606, 94-99.
  288. Kang, Q.; Jiang, C.-Y.; Fujita, T.; Kumamoto, E. Spontaneous L-glutamate release enhancement in rat substantia gelatinosa neurons by (-)-carvone and (+)-carvone which activate different types of TRP channel. Biochem. Biophys. Res. Commun. 2015, 459, 498-503.
  289. Jiang, C.-Y.; Wang, C.; Xu, N.-X.; Fujita, T.; Murata, Y.; Kumamoto, E. 1,8- and 1,4-cineole enhance spontaneous excitatory transmission by activating different types of transient receptor potential channels in the rat spinal substantia gelatinosa. J. Neurochem. 2016, 136, 764-777.
  290. Wang, C.; Fujita, T.; Yasuda, H.; Kumamoto, E. Spontaneous excitatory transmission enhancement produced by linalool and its isomer geraniol in rat spinal substantia gelatinosa neurons - involvement of transient receptor potential channels. Phytomedicine Plus 2022, 2, 100155.
  291. Kumamoto, E.; Fujita, T.; Jiang, C.-Y. TRP channels involved in spontaneous L-glutamate release enhancement in the adult rat spinal substantia gelatinosa. Cells 2014, 3, 331-362.
  292. Kumamoto, E.; Fujita, T. Differential activation of TRP channels in the adult rat spinal substantia gelatinosa by stereoisomers of plant-derived chemicals. Pharmaceuticals 2016, 9, 46.
  293. Guimarães, A.G.; Quintans, J.S.S.; Quintans-Júnior, L.J. Monoterpenes with analgesic activity - a systematic review. Phytother. Res. 2013, 27, 1-15.
  294. Tsuchiya, H. Anesthetic agents of plant origin: a review of phytochemicals with anesthetic activity. Molecules 2017, 22, 1369.
  295. Nolano, M.; Simone, D.A.; Wendelschafer-Crabb, G.; Johnson, T.; Hazen, E.; Kennedy, W.R. Topical capsaicin in humans: parallel loss of epidermal nerve fibers and pain sensation. Pain 1999, 81, 135-145.
  296. Malmberg, A.B.; Mizisin, A.P.; Calcutt, N.A.; von Stein, T.; Robbins, W.R.; Bley, K.R. Reduced heat sensitivity and epidermal nerve fiber immunostaining following single applications of a high-concentration capsaicin patch. Pain 2004, 111, 360-367.
  297. Kawasaki, H.; Mizuta, K.; Fujita, T.; Kumamoto, E. Inhibition by menthol and its related chemicals of compound action potentials in frog sciatic nerves. Life Sci. 2013, 92, 359-367.
  298. Matsushita, A.; Ohtsubo, S.; Fujita, T.; Kumamoto, E. Inhibition by TRPA1 agonists of compound action potentials in the frog sciatic nerve. Biochem. Biophys. Res. Commun. 2013, 434, 179-184.
  299. Ohtsubo, S.; Fujita, T.; Matsushita, A.; Kumamoto, E. Inhibition of the compound action potentials of frog sciatic nerves by aroma oil compounds having various chemical structures. Pharmacol. Res. Perspect. 2015, 3, e00127.
  300. Lundbæk, J.A.; Birn, P.; Tape, S.E.; Toombes, G.E.S.; Søgaard, R.; Koeppe II, R.E.; Gruner, S.M.; Hansen, A.J.; Andersen, O.S. Capsaicin regulates voltage-dependent sodium channels by altering lipid bilayer elasticity. Mol. Pharmacol. 2005, 68, 680-689.
  301. Cao, X.; Cao, X.; Xie, H.; Yang, R.; Lei, G.; Li, F.; Li, A.; Liu, C.; Liu, L. Effects of capsaicin on VGSCs in TRPV1-/- mice. Brain Res. 2007, 1163, 33-43.
  302. Wang, S.-Y.; Mitchell, J.; Wang, G.K. Preferential block of inactivation-deficient Na+ currents by capsaicin reveals a non-TRPV1 receptor within the Na+ channel. Pain 2007, 127, 73-83.
  303. Cho, J.S.; Kim, T.H.; Lim, J.-M.; Song, J.-H. Effects of eugenol on Na+ currents in rat dorsal root ganglion neurons. Brain Res. 2008, 1243, 53-62.
  304. Haeseler, G.; Maue, D.; Grosskreutz, J.; Bufler, J.; Nentwig, B.; Piepenbrock, S.; Dengler, R.; Leuwer, M. Voltage-dependent block of neuronal and skeletal muscle sodium channels by thymol and menthol. Eur. J. Anaesthesiol. 2002, 19, 571-579.
  305. Joca, H.C.; Cruz-Mendes, Y.; Oliveira-Abreu, K.; Maia-Joca, R.P.M.; Barbosa, R.; Lemos, T.L.; Lacerda Beirão, P.S.; Leal-Cardoso, J.H. Carvacrol decreases neuronal excitability by inhibition of voltage-gated sodium channels. J. Nat. Prod. 2012, 75, 1511-1517.
  306. Leal-Cardoso, J.H.; da Silva-Alves, K.S.; Ferreira-da-Silva, F.W.; dos Santos-Nascimento, T.; Joca, H.C.; de Macedo, F.H.P.; de Albuquerque-Neto, P.M.; Magalhães, P.J.C.; Lahlou, S.; Cruz, J.S.; Barbosa, R. Linalool blocks excitability in peripheral nerves and voltage-dependent Na+ current in dissociated dorsal root ganglia neurons. Eur. J. Pharmacol. 2010, 645, 86-93.
  307. de Araújo, D.A.M.; Freitas, C.; Cruz, J.S. Essential oils components as a new path to understand ion channel molecular pharmacology. Life Sci. 2011, 89, 540-544.
  308. Joca, H.C.; Vieira, D.C.O.; Vasconcelos, A.P.; Araújo, D.A.M,; Cruz, J.S. Carvacrol modulates voltage-gated sodium channels kinetics in dorsal root ganglia. Eur. J. Pharmacol. 2015, 756, 22-29.
  309. Nozoe, T. Über die farbstoffe im holzteile des “hinoki” baumes. I. Hinokitin und hinokitiol. Bull. Chem. Soc. Jpn. 1936, 11, 295-298.
  310. Magori, N.; Fujita, T.; Kumamoto, E. Hinokitiol inhibits compound action potentials in the frog sciatic nerve. Eur. J. Pharmacol. 2018, 819, 254-260.
  311. Baba, T.; Nakano, H.; Tamai, K.; Sawamura, D.; Hanada, K.; Hashimoto, I.; Arima, Y. Inhibitory effect of β-thujaplicin on ultraviolet B-induced apoptosis in mouse keratinocytes. J. Invest. Dermatol. 1998, 110, 24-28.
  312. Shih, Y.-H.; Lin, D.-J.; Chang, K.-W.; Hsia, S.-M.; Ko, S.-Y.; Lee, S.-Y.; Hsue, S.-S.; Wang, T.-H.; Chen, Y.-L.; Shieh, T.-M. Evaluation physical characteristics and comparison antimicrobial and anti-inflammation potentials of dental root canal sealers containing hinokitiol in vitro. PLoS One 2014, 9, e94941.
  313. Morita, Y.; Matsumura, E.; Okabe, T.; Shibata, M.; Sugiura, M.; Ohe, T.; Tsujibo, H.; Ishida, N.; Inamori, Y. Biological activity of tropolone. Biol. Pharm. Bull. 2003, 26, 1487-1490.
  314. Morita, Y.; Matsumura, E.; Okabe, T.; Fukui, T.; Ohe, T.; Ishida, N.; Inamori, Y. Biological activity of β-dolabrin, γ-thujaplicin, and 4-acetyltropolone, hinokitiol-related compounds. Biol. Pharm. Bull. 2004, 27, 1666-1669.
  315. Yamato, M.; Ando, J.; Sakaki, K.; Hashigaki, K.; Wataya, Y.; Tsukagoshi, S.; Tashiro, T.; Tsuruo, T. Synthesis and antitumor activity of tropolone derivatives. 7. Bistropolones containing connecting methylene chains. J. Med. Chem. 1992, 35, 267-273.
  316. Inamori, Y.; Tsujibo, H.; Ohishi, H.; Ishii, F.; Mizugaki, M.; Aso, H.; Ishida, N. Cytotoxic effect of hinokitiol and tropolone on the growth of mammalian cells and on blastogenesis of mouse splenic T cells. Biol. Pharm. Bull. 1993, 16, 521-523.
  317. Yasumoto, E.; Nakano, K.; Nakayachi, T.; Morshed, S.R.; Hashimoto, K.; Kikuchi, H.; Nishikawa, H.; Kawase, M.; Sakagami, H. Cytotoxic activity of deferiprone, maltol and related hydroxyketones against human tumor cell lines. Anticancer Res. 2004, 24, 755-762.
  318. Nagao, Y.; Sata, M. Effect of oral care gel on the quality of life for oral lichen planus in patients with chronic HCV infection. Virol. J. 2011, 8, 348.
  319. James, R.; Glen, J.B. Synthesis, biological evaluation, and preliminary structure-activity considerations of a series of alkylphenols as intravenous anesthetic agents. J. Med. Chem. 1980, 23, 1350-1357.
  320. Doze, V.A.; Westphal, L.M.; White, P.F. Comparison of propofol with methohexital for outpatient anesthesia. Anesth. Analg. 1986, 65, 1189-1195.
  321. Shafer, A.; Doze, V.A.; Shafer, S.L.; White, P.F. Pharmacokinetics and pharmacodynamics of propofol infusions during general anesthesia. Anesthesiology 1988, 69, 348-356.
  322. Antkowiak, B.; Rammes, G. GABA(A) receptor-targeted drug development - new perspectives in perioperative anesthesia. Expert Opin. Drug Discov. 2019, 14, 683-699.
  323. Vasileiou, I.; Xanthos, T.; Koudouna, E.; Perrea, D.; Klonaris, C.; Katsargyris, A.; Papadimitriou, L. Propofol: a review of its non-anaesthetic effects. Eur. J. Pharmacol. 2009, 605, 1-8.
  324. Hanrahan, S.J.; Greger, B.; Parker, R.A.; Ogura, T.; Obara, S.; Egan, T.D.; House, P.A. The effects of propofol on local field potential spectra, action potential firing rate, and their temporal relationship in humans and felines. Front. Hum. Neurosci. 2013, 7, 136.
  325. Shi, Q.-Q.; Sun, X.; Fang, H. A mechanism study on propofol's action on middle latency auditory evoked potential by neurons in ventral partition of medial geniculate body in rats. Eur. Rev. Med. Pharmacol. Sci. 2014, 18, 1859-1868.
  326. Takazawa, T.; Furue, H.; Nishikawa, K.; Uta, D.; Takeshima, K.; Goto, F.; Yoshimura, M. Actions of propofol on substantia gelatinosa neurones in rat spinal cord revealed by in vitro and in vivo patch-clamp recordings. Eur. J. Neurosci. 2009, 29, 518-528.
  327. Hijikata, Y. Analgesic treatment with Kampo prescription. Expert Rev. Neurother. 2006, 6, 795-802.
  328. Kono, T.; Kanematsu, T.; Kitajima, M. Exodus of Kampo, traditional Japanese medicine, from the complementary and alternative medicines: is it time yet? Surgery 2009, 146, 837-840.
  329. Motoo, Y.; Arai, I.; Hyodo, I.; Tsutani, K. Current status of Kampo (Japanese herbal) medicines in Japanese clinical practice guidelines. Complement. Ther. Med. 2009, 17, 147-154.
  330. Mochiki, E.; Yanai, M.; Ohno, T.; Kuwano, H. The effect of traditional Japanese medicine (Kampo) on gastrointestinal function. Surg. Today 2010, 40, 1105-1111.
  331. Wachtel-Galor, S.; Benzie, I.F.F. Herbal medicine: an introduction to its history, usage, regulation, current trends, and research needs. In: Herbal medicine: biomolecular and clinical aspects, 2nd ed.; Benzie, I.F.F.; Wachtel-Galor, S., Eds.; CRC Press, Boca Raton, FL, U.S.A., 2011, Chapter 1.
  332. Sunagawa, M.; Takayama, Y.; Kato, M.; Tanaka, M.; Fukuoka, S.; Okumo, T.; Tsukada, M.; Yamaguchi, K. Kampo formulae for the treatment of neuropathic pain - especially the mechanism of action of Yokukansan - . Front. Mol. Neurosci. 2021, 14, 705023.
  333. Matsushita, A.; Fujita, T.; Ohtsubo, S.; Kumamoto, E. Traditional Japanese medicines inhibit compound action potentials in the frog sciatic nerve. J. Ethnopharmacol. 2016, 178, 272-280.
  334. Fink, B.R.; Cairns, A.M. Differential slowing and block of conduction by lidocaine in individual afferent myelinated and unmyelinated axons. Anesthesiology 1984, 60, 111-120.
  335. Brodin, P. Differential inhibition of A, B and C fibres in the rat vagus nerve by lidocaine, eugenol and formaldehyde. Arch. Oral Biol. 1985, 30, 477-480.
  336. Nakamura, M.; Jang, I.-S. Indomethacin inhibits tetrodotoxin-resistant Na+ channels at acidic pH in rat nociceptive neurons. Neuropharmacology 2016, 105, 454-462.
  337. Joshi, S.K.; Mikusa, J.P.; Hernandez, G.; Baker, S.; Shieh, C.C.; Neelands, T.; Zhang, X.F.; Niforatos, W.; Kage, K.; Han, P.; Krafte, D.; Faltynek, C.; Sullivan, J.P.; Jarvis, M.F.; Honore, P. Involvement of the TTX-resistant sodium channel Nav 1.8 in inflammatory and neuropathic, but not post-operative, pain states. Pain 2006, 123, 75-82.
More
Video Production Service