全部 标题 作者
关键词 摘要

OALib Journal期刊
ISSN: 2333-9721
费用:99美元

查看量下载量

相关文章

更多...

Interaction of Local Anesthetics with Biomembranes Consisting of Phospholipids and Cholesterol: Mechanistic and Clinical Implications for Anesthetic and Cardiotoxic Effects

DOI: 10.1155/2013/297141

Full-Text   Cite this paper   Add to My Lib

Abstract:

Despite a long history in medical and dental application, the molecular mechanism and precise site of action are still arguable for local anesthetics. Their effects are considered to be induced by acting on functional proteins, on membrane lipids, or on both. Local anesthetics primarily interact with sodium channels embedded in cell membranes to reduce the excitability of nerve cells and cardiomyocytes or produce a malfunction of the cardiovascular system. However, the membrane protein-interacting theory cannot explain all of the pharmacological and toxicological features of local anesthetics. The administered drug molecules must diffuse through the lipid barriers of nerve sheaths and penetrate into or across the lipid bilayers of cell membranes to reach the acting site on transmembrane proteins. Amphiphilic local anesthetics interact hydrophobically and electrostatically with lipid bilayers and modify their physicochemical property, with the direct inhibition of membrane functions, and with the resultant alteration of the membrane lipid environments surrounding transmembrane proteins and the subsequent protein conformational change, leading to the inhibition of channel functions. We review recent studies on the interaction of local anesthetics with biomembranes consisting of phospholipids and cholesterol. Understanding the membrane interactivity of local anesthetics would provide novel insights into their anesthetic and cardiotoxic effects. 1. Introduction Local anesthetics clinically used so far have the common chemical structure that is composed of three portions: the hydrophobic moiety consisting of an aromatic ring, the intermediate chain, and the hydrophilic moiety consisting of an amino terminus. The aromatic residue confers lipid solubility on a drug molecule, whereas the ionizable amino group confers, water solubility. The intermediate portion provides the spatial separation between hydrophobic and hydrophilic end and structurally classifies local anesthetics into amide type and ester type (Figure 1). Figure 1: Representative amide and ester local anesthetics. Because of the presence of substituted amino groups, local anesthetics are referred to as the bases with pKa values ranging from 7.7 to 8.1 at 37°C for the amide type and from 8.4 to 8.9 at 37°C for the ester type [1], so they exist in uncharged and positively charged form. After injected, local anesthetics show an in vivo equilibrium between the uncharged and the charged fraction of molecules. According to the Henderson-Hasselbalch equation, the percentage of uncharged molecules depends

References

[1]  G. R. Strichartz, V. Sanchez, G. R. Arthur, R. Chafetz, and D. Martin, “Fundamental properties of local anesthetics. II. Measured octanol: buffer partition coefficients and pKa values of clinically used drugs,” Anesthesia and Analgesia, vol. 71, no. 2, pp. 158–170, 1990.
[2]  A. B. López and M. P. Diago, “Failure of locoregional anesthesia in dental practice. Review of the literature,” Medicina Oral, Patología Oral y Cirugía Bucal, vol. 11, no. 6, pp. E510–E513, 2006.
[3]  D. E. Becker and K. L. Reed, “Essentials of local anesthetic pharmacology,” Anesthesia Progress, vol. 53, no. 3, pp. 98–109, 2006.
[4]  J. H. C. Tan and D. A. Saint, “Interaction of lidocaine with the cardiac sodium channel. Effects of low extracellular pH are consistent with an external blocking site,” Life Sciences, vol. 67, no. 22, pp. 2759–2766, 2000.
[5]  F. Yanagidate and G. R. Strichartz, “Local anesthetics,” in Handbook of Experimental Pharmacology, vol. 177, pp. 95–127, Springer, Berlin, Germany, 2007.
[6]  G. McCleane, “Intravenous lidocaine: an outdated or underutilized treatment for pain?” Journal of Palliative Medicine, vol. 10, no. 3, pp. 798–805, 2007.
[7]  W. Kope?, J. Telenius, and H. Khandelia, “Molecular dynamics simulations of the interactions of medicinal plant extracts and drugs with lipid bilayer membranes,” FEBS Journal, vol. 280, no. 12, pp. 2785–2805, 2013.
[8]  J. G. Paiva, P. Paradiso, A. P. Serro, A. Fernandes, and B. Saramago, “Interaction of local and general anaesthetics with liposomal membrane models: a QCM-D and DSC study,” Colloids and Surfaces B, vol. 95, pp. 65–74, 2012.
[9]  C. H?gberg and A. P. Lyubartsev, “Effect of local anesthetic lidocaine on electrostatic properties of a lipid bilayer,” Biophysical Journal, vol. 94, no. 2, pp. 525–531, 2008.
[10]  M. Pasenkiewicz-Gierula, T. Róg, J. Grochowski et al., “Effects of a carane derivative local anesthetic on a phospholipid bilayer studied by molecular dynamics simulation,” Biophysical Journal, vol. 85, no. 2, pp. 1248–1258, 2003.
[11]  A. Alberola-Die, J. Martinez-Pinna, J. M. González-Ros, I. Ivorra, and A. Morales, “Multiple inhibitory actions of lidocaine on Torpedo nicotinic acetylcholine receptors transplanted to Xenopus oocytes,” Journal of Neurochemistry, vol. 117, no. 6, pp. 1009–1019, 2011.
[12]  K. Hara and T. Sata, “The effects of the local anesthetics lidocaine and procaine on glycine and γ-aminobutyric acid receptors expressed in Xenopus oocytes,” Anesthesia and Analgesia, vol. 104, no. 6, pp. 1434–1439, 2007.
[13]  T. W. R. Lee, H. P. Grocott, D. Schwinn, and E. Jacobsohn, “High spinal anesthesia for cardiac surgery: effects on β-adrenergic receptor function, stress response, and hemodynamics,” Anesthesiology, vol. 98, no. 2, pp. 499–510, 2003.
[14]  J. T. Jastak, J. A. Yagiela, and D. Donaldson, Local Anesthesia of the Oral Cavity, WB Saunders, Philadelphia, Pa, USA, 1995.
[15]  C. Lynch III, “Meyer and Overton revisited,” Anesthesia and Analgesia, vol. 107, no. 3, pp. 864–867, 2008.
[16]  M. Eeman and M. Deleu, “From biological membranes to biomimetic model membranes,” Biotechnology, Agronomy and Society and Environment, vol. 14, no. 4, pp. 719–736, 2010.
[17]  I. Yun, E. Cho, H. Jang et al., “Amphiphilic effects of local anesthetics on rotational mobility in neuronal and model membranes,” Biochimica et Biophysica Acta, vol. 1564, no. 1, pp. 123–132, 2002.
[18]  H. Tsuchiya, M. Mizogami, and K. Takakura, “Reversed-phase liquid chromatographic retention and membrane activity relationships of local anesthetics,” Journal of Chromatography A, vol. 1073, no. 1-2, pp. 303–308, 2005.
[19]  M. Mizogami, H. Tsuchiya, and J. Harada, “Membrane effects of ropivacaine compared with those of bupivacaine and mepivacaine,” Fundamental and Clinical Pharmacology, vol. 16, no. 4, pp. 325–330, 2002.
[20]  H. Tsuchiya and M. Mizogami, “Membrane interactivity of charged local anesthetic derivative and stereoselectivity in membrane interaction of local anesthetic enantiomers,” Local and Regional Anesthesia, vol. 1, pp. 1–9, 2008.
[21]  K. J. McClellan and D. Faulds, “Ropivacaine: an update of its use in regional anaesthesia,” Drugs, vol. 60, no. 5, pp. 1065–1093, 2000.
[22]  H. ?nyüksel, V. Sethi, G. L. Weinberg, P. K. Dudeja, and I. Rubinstein, “Bupivacaine, but not lidocaine, disrupts cardiolipin-containing small biomimetic unilamellar liposomes,” Chemico-Biological Interactions, vol. 169, no. 3, pp. 154–159, 2007.
[23]  H. Tsuchiya, T. Ueno, M. Mizogami, and K. Takakura, “Local anesthetics structure-dependently interact with anionic phospholipid membranes to modify the fluidity,” Chemico-Biological Interactions, vol. 183, no. 1, pp. 19–24, 2010.
[24]  L. Groban, D. D. Deal, J. C. Vernon, R. L. James, and J. Butterworth, “Cardiac resuscitation after incremental overdosage with lidocaine, bupivacaine, levobupivacaine, and ropivacaine in anesthetized dogs,” Anesthesia and Analgesia, vol. 92, no. 1, pp. 37–43, 2001.
[25]  L. Groban, D. D. Deal, J. C. Vernon, R. L. James, and J. Butterworth, “Ventricular arrhythmias with or without programmed electrical stimulation after incremental overdosage with lidocaine, bupivacaine, levobupivacaine, and ropivacaine,” Anesthesia and Analgesia, vol. 91, no. 5, pp. 1103–1111, 2000.
[26]  J. Wang and G. Zhang, “Influence of membrane physical state on lysosomal potassium ion permeability,” Cell Biology International, vol. 29, no. 6, pp. 393–401, 2005.
[27]  G. Balogh, I. Horváth, E. Nagy et al., “The hyperfluidization of mammalian cell membranes acts as a signal to initiate the heat shock protein response,” FEBS Journal, vol. 272, no. 23, pp. 6077–6086, 2005.
[28]  S. Schreier, S. V. P. Malheiros, and E. de Paula, “Surface active drugs: self-association and interaction with membranes and surfactants. Physicochemical and biological aspects,” Biochimica et Biophysica Acta, vol. 1508, no. 1-2, pp. 210–234, 2000.
[29]  A. Pohorille, M. A. Wilson, M. H. New, and C. Chipot, “Concentrations of anesthetics across the water-membrane interface; the Meyer-Overton hypothesis revisited,” Toxicology Letters, vol. 100-101, pp. 421–430, 1998.
[30]  A. Sunami, I. W. Glaaser, and H. A. Fozzard, “A critical residue for isoform difference in tetrodotoxin affinity is a molecular determinant of the external access path for local anesthetics in the cardiac sodium channel,” Proceedings of the National Academy of Sciences of the United States of America, vol. 97, no. 5, pp. 2326–2331, 2000.
[31]  T. K. Y. Lim, B. A. MacLeod, C. R. Ries, and S. K. W. Schwarz, “The quaternary lidocaine derivative, QX-314, produces long-lasting local anesthesia in animal models in vivo,” Anesthesiology, vol. 107, no. 2, pp. 305–311, 2007.
[32]  G. K. Wang, “Local anesthetics,” in Neural Mechanisms of Anesthesia, J. F. Antognini, E. E. Carstens, and D. E. Raines, Eds., Humana Press, Totowa, NJ, USA, 2003.
[33]  K. A. Slaviero, S. J. Clarke, and L. P. Rivory, “Inflammatory response: an unrecognised source of variability in the pharmacokinetics and pharmacodynamics of cancer chemotherapy,” The Lancet Oncology, vol. 4, no. 4, pp. 224–232, 2003.
[34]  S. Sattari, W. F. Dryden, L. A. Eliot, and F. Jamali, “Despite increased plasma concentration, inflammation reduces potency of calcium channel antagonists due to lower binding to the rat heart,” British Journal of Pharmacology, vol. 139, no. 5, pp. 945–954, 2003.
[35]  E. Claffey, A. Reader, J. Nusstein, M. Beck, and J. Weaver, “Anesthetic efficacy of articaine for inferior alveolar nerve blocks in patients with irreversible pulpitis,” Journal of Endodontics, vol. 30, no. 8, pp. 568–571, 2004.
[36]  N. Srinivasan, M. Kavitha, C. S. Loganathan, and G. Padmini, “Comparison of anesthetic efficacy of 4% articaine and 2% lidocaine for maxillary buccal infiltration in patients with irreversible pulpitis,” Oral Surgery, Oral Medicine, Oral Pathology, Oral Radiology and Endodontology, vol. 107, no. 1, pp. 133–136, 2009.
[37]  I. Poto?nik and F. Bajrovi?, “Failure of inferior alveolar nerve block in endodontics,” Dental Traumatology, vol. 15, no. 6, pp. 247–251, 1999.
[38]  A. Punnia-Moorthy, “Evaluation of pH changes in inflammation of the subcutaneous air pouch lining in the rat, induced by carrageenan, dextran and Staphylococcus aureus,” Journal of Oral Pathology, vol. 16, no. 1, pp. 36–44, 1987.
[39]  D. de Backer, “Lactic acidosis,” Minerva Anestesiologica, vol. 69, no. 4, pp. 281–284, 2003.
[40]  H. Tsuchiya, M. Mizogami, T. Ueno, and K. Takakura, “Interaction of local anaesthetics with lipid membranes under inflammatory acidic conditions,” Inflammopharmacology, vol. 15, no. 4, pp. 164–170, 2007.
[41]  T. Ueno, H. Tsuchiya, M. Mizogami, and K. Takakura, “Local anesthetic failure associated with inflammation: verification of the acidosis mechanism and the hypothetic participation of inflammatory peroxynitrite,” Journal of Inflammation Research, vol. 1, pp. 41–48, 2008.
[42]  A. Punnia-Moorthy, “Buffering capacity of normal and inflamed tissues following the injection of local anaesthetic solutions,” British Journal of Anaesthesia, vol. 61, no. 2, pp. 154–159, 1988.
[43]  B. Gunaydin and A. T. Demiryurek, “Interaction of lidocaine with reactive oxygen and nitrogen species,” European Journal of Anaesthesiology, vol. 18, no. 12, pp. 816–822, 2001.
[44]  H. Tsuchiya, T. Ueno, M. Mizogami, and K. Takakura, “Antioxidant activity analysis by liposomal membrane system and application to anesthetics,” Analytical Sciences, vol. 24, no. 12, pp. 1557–1562, 2008.
[45]  T. Ueno, M. Mizogami, K. Takakura, and H. Tsuchiya, “Membrane effect of lidocaine is inhibited by interaction with peroxynitrite,” Journal of Anesthesia, vol. 22, no. 1, pp. 96–99, 2008.
[46]  T. Ueno, M. Mizogami, K. Takakura, and H. Tsuchiya, “Peroxynitrite affects lidocaine by acting on membrane-constituting lipids,” Journal of Anesthesia, vol. 22, no. 4, pp. 475–478, 2008.
[47]  N. M. Flake and M. S. Gold, “Inflammation alters sodium currents and excitability of temporomandibular joint afferents,” Neuroscience Letters, vol. 384, no. 3, pp. 294–299, 2005.
[48]  J. Wang, J. A. Strong, W. Xie, and J. Zhang, “Local inflammation in rat dorsal root ganglion alters excitability and ion currents in small-diameter sensory neurons,” Anesthesiology, vol. 107, no. 2, pp. 322–332, 2007.
[49]  J. M. Dippenaar, “Local anaesthetic toxicity,” Southern African Journal of Anaesthesia and Analgesia, vol. 13, no. 3, pp. 23–28, 2007.
[50]  C. Gevirtz, “Local anesthetic systemic toxicity: prevention and treatment,” Topics in Pain Management, vol. 27, no. 12, pp. 1–6, 2012.
[51]  J. E. Heavner, “Cardiac toxicity of local anesthetics in the intact isolated heart model: a review,” Regional Anesthesia and Pain Medicine, vol. 27, no. 6, pp. 545–555, 2002.
[52]  L. Pardo, T. J. J. Blanck, and E. Recio-Pinto, “The neuronal lipid membrane permeability was markedly increased by bupivacaine and mildly affected by lidocaine and ropivacaine,” European Journal of Pharmacology, vol. 455, no. 2-3, pp. 81–90, 2002.
[53]  P. Friederich, “Basic concepts of ion channel physiology and anaesthetic drug effects,” European Journal of Anaesthesiology, vol. 20, no. 5, pp. 343–353, 2003.
[54]  R. H. Houtkooper and F. M. Vaz, “Cardiolipin, the heart of mitochondrial metabolism,” Cellular and Molecular Life Sciences, vol. 65, no. 16, pp. 2493–2506, 2008.
[55]  K. Felice and H. Schumann, “Intravenous lipid emulsion for local anesthetic toxicity: a review of the literature,” Journal of Medical Toxicology, vol. 4, no. 3, pp. 184–191, 2008.
[56]  J. Muhonen, J. M. Holopainen, and S. K. Wiedmer, “Interactions between local anesthetics and lipid dispersions studied with liposome electrokinetic capillary chromatography,” Journal of Chromatography A, vol. 1216, no. 15, pp. 3392–3397, 2009.
[57]  G. Weinberg, R. Ripper, D. L. Feinstein, and W. Hoffman, “Lipid emulsion infusion rescues dogs from bupivacaine-induced cardiac toxicity,” Regional Anesthesia and Pain Medicine, vol. 28, no. 3, pp. 198–202, 2003.
[58]  R. J. Litz, M. Popp, S. N. Stehr, and T. Koch, “Successful resuscitation of a patient with ropivacaine-induced asystole after axillary plexus block using lipid infusion,” Anaesthesia, vol. 61, no. 8, pp. 800–801, 2006.
[59]  G. Foxall, R. McCahon, J. Lamb, J. G. Hardman, and N. M. Bedforth, “Levobupivacaine-induced seizures and cardiovascular collapse treated with Intralipid,” Anaesthesia, vol. 62, no. 5, pp. 516–518, 2007.
[60]  S. Ciechanowicz and V. Patil, “Lipid emulsion for local anesthetic systemic toxicity,” Anesthesiology Research and Practice, vol. 2012, Article ID 131784, 11 pages, 2012.
[61]  J. A. Warren, R. B. Thoma, A. Georgescu, and S. J. Shah, “Intravenous lipid infusion in the successful resuscitation of local anesthetic-induced cardiovascular collapse after supraclavicular brachial plexus block,” Anesthesia and Analgesia, vol. 106, no. 5, pp. 1578–1580, 2008.
[62]  E. Calenda and S. A. Dinescu, “Failure of lipid emulsion to reverse neurotoxicity after an ultrasound-guided axillary block with ropivacaine and mepivacaine,” Journal of Anesthesia, vol. 23, no. 3, pp. 472–473, 2009.
[63]  E. Bourne, C. Wright, and C. Royse, “A review of local anesthetic cardiotoxicity and treatment with lipid emulsion,” Local and Regional Anesthesia, vol. 3, no. 1, pp. 11–19, 2010.
[64]  J. Mazoit, R. Le Guen, H. Beloeil, and D. Benhamou, “Binding of long-lasting local anesthetics to lipid emulsions,” Anesthesiology, vol. 110, no. 2, pp. 380–386, 2009.
[65]  AAGBI, Management of Severe Local Anaesthetic Toxicity 2 (2010), AAGBI Guidelines, 2010, http://www.aagbi.org/publications/publications-guidelines/M/R.
[66]  P. Tsibiribi, C. Bui-Xuan, B. Bui-Xuan et al., “The effects of ropivacaine at clinically relevant doses on myocardial ischemia in pigs,” Journal of Anesthesia, vol. 20, no. 4, pp. 341–343, 2006.
[67]  D. J. Pacini, G. Boachie-Ansah, and K. A. Kane, “Modification by hypoxia, hyperkalaemia and acidosis of the cardiac electrophysiological effects of a range of antiarrhythmic drugs,” British Journal of Pharmacology, vol. 107, no. 3, pp. 665–670, 1992.
[68]  M. Freysz, Q. Timour, L. Bertrix, J. Loufoua-Moundanga, S. Omar, and G. Faucon, “Enhancement by ischemia of the risk of cardiac disorders, especially fibrillation, in regional anesthesia with bupivacaine,” Acta Anaesthesiologica Scandinavica, vol. 37, no. 4, pp. 350–356, 1993.
[69]  R. Ferrari, A. Cargnoni, P. Bernocchi et al., “Metabolic adaptation during a sequence of no-flow and low-flow ischemia: a possible trigger for hibernation,” Circulation, vol. 94, no. 10, pp. 2587–2596, 1996.
[70]  Y. Watanabe, S. Dohi, H. Iida, and T. Ishiyama, “The effects of bupivacaine and ropivacaine on baroreflex sensitivity with or without respiratory acidosis and alkalosis in rats,” Anesthesia and Analgesia, vol. 84, no. 2, pp. 398–404, 1997.
[71]  H. Tsuchiya, M. Mizogami, T. Ueno, and K. Shigemi, “Cardiotoxic local anesthetics increasingly interact with biomimetic membranes under ischemia-like acidic conditions,” Biological and Pharmaceutical Bulletin, vol. 35, no. 6, pp. 988–992, 2012.
[72]  G. Paradies, G. Petrosillo, M. Pistolese, N. Di Venosa, A. Federici, and F. M. Ruggiero, “Decrease in mitochondrial complex I activity in ischemic/reperfused rat heart: involvement of reactive oxygen species and cardiolipin,” Circulation Research, vol. 94, no. 1, pp. 53–59, 2004.
[73]  M. M. Lalu, W. Wang, and R. Schulz, “Peroxynitrite in myocardial ischemia-reperfusion injury,” Heart Failure Reviews, vol. 7, no. 4, pp. 359–369, 2002.
[74]  H. Tsuchiya, M. Mizogami, and K. Shigemi, “Increasing membrane interactions of local anaesthetics as hypothetic mechanism for their cardiotoxicity enhanced by myocardial ischaemia,” Basic and Clinical Pharmacology and Toxicology, vol. 111, no. 5, pp. 303–308, 2012.
[75]  E. J. Lesnefsky and C. L. Hoppel, “Cardiolipin as an oxidative target in cardiac mitochondria in the aged rat,” Biochimica et Biophysica Acta, vol. 1777, no. 7-8, pp. 1020–1027, 2008.
[76]  G. Paradies, G. Petrosillo, V. Paradies, and F. M. Ruggiero, “Role of cardiolipin peroxidation and Ca2+ in mitochondrial dysfunction and disease,” Cell Calcium, vol. 45, no. 6, pp. 643–650, 2009.
[77]  H. Tsuchiya, T. Ueno, and M. Mizogami, “Stereostructure-based differences in the interactions of cardiotoxic local anesthetics with cholesterol-containing biomimetic membranes,” Bioorganic and Medicinal Chemistry, vol. 19, no. 11, pp. 3410–3415, 2011.
[78]  D. J. Kaczmarek, C. Herzog, J. Larmann et al., “Lidocaine protects from myocardial damage due to ischemia and reperfusion in mice by its antiapoptotic effects,” Anesthesiology, vol. 110, no. 5, pp. 1041–1049, 2009.
[79]  F. Lenfant, J. Lahet, C. Courderot-Masuyer, M. Freysz, and L. Rochette, “Lidocaine has better antioxidant potential than ropivacaine and bupivacaine: in vitro comparison in a model of human erythrocytes submitted to an oxidative stress,” Biomedicine and Pharmacotherapy, vol. 58, no. 4, pp. 248–254, 2004.
[80]  S. Pope, J. M. Land, and S. J. R. Heales, “Oxidative stress and mitochondrial dysfunction in neurodegeneration; cardiolipin a critical target?” Biochimica et Biophysica Acta, vol. 1777, no. 7-8, pp. 794–799, 2008.
[81]  E. J. Lesnefsky, T. J. Slabe, M. S. K. Stoll, P. E. Minkler, and C. L. Hoppel, “Myocardial ischemia selectively depletes cardiolipin in rabbit heart subsarcolemmal mitochondria,” American Journal of Physiology, vol. 280, no. 6, pp. H2770–H2778, 2001.
[82]  K. M. Rentsch, “The importance of stereoselective determination of drugs in the clinical laboratory,” Journal of Biochemical and Biophysical Methods, vol. 54, no. 1-3, pp. 1–9, 2002.
[83]  C. Nau and G. R. Strichartz, “Drug chirality in anesthesia,” Anesthesiology, vol. 97, no. 2, pp. 497–502, 2002.
[84]  R. H. Foster and A. Markham, “Levobupivacaine: a review of its pharmacology and use as a local anaesthetic,” Drugs, vol. 59, no. 3, pp. 551–579, 2000.
[85]  M. Vladimirov, C. Nau, W. M. Mok, and G. Strichartz, “Potency of bupivacaine stereoisomers tested in vitro and in vivo: biochemical, electrophysiological, and neurobehavioral studies,” Anesthesiology, vol. 93, no. 3, pp. 744–755, 2000.
[86]  G. Lyons, M. Columb, R. C. Wilson, and R. V. Johnson, “Epidural pain relief in labour: potencies of levobupivacaine and racemic bupivacaine,” British Journal of Anaesthesia, vol. 81, no. 6, pp. 899–901, 1998, Erratum in: British Journal of Anaesthesia, vol. 82, no. 3, p. 488, 1998.
[87]  Y. Lim, C. E. Ocampo, and A. T. Sia, “A comparison of duration of analgesia of intrathecal 2.5?mg of bupivacaine, ropivacaine, and levobupivacaine in combined spinal epidural analgesia for patients in labor,” Anesthesia and Analgesia, vol. 98, no. 1, pp. 235–239, 2004.
[88]  B. M. Graf, E. Martin, Z. J. Bosnjak, and D. F. Stowe, “Stereospecific effect of bupivacaine isomers on atrioventricular conduction in the isolated perfused guinea pig heart,” Anesthesiology, vol. 86, no. 2, pp. 410–419, 1997.
[89]  S. G. Morrison, J. J. Dominguez, P. Frascarolo, and S. Reiz, “A comparison of the electrocardiographic cardiotoxic effects of racemic bupivacaine, levobupivacaine, and ropivacaine in anesthetized swine,” Anesthesia and Analgesia, vol. 90, no. 6, pp. 1308–1314, 2000.
[90]  C. Valenzuela, D. J. Snyders, P. B. Bennett, J. Tamargo, and L. M. Hondeghem, “Stereoselective block of cardiac sodium channels by bupivacaine in guinea pig ventricular myocytes,” Circulation, vol. 92, no. 10, pp. 3014–3024, 1995.
[91]  C. Valenzuela, E. Delpón, M. M. Tamkun, J. Tamargo, and D. J. Snyders, “Stereoselective block of a human cardiac potassium channel (Kv1.5) by bupivacaine enantiomers,” Biophysical Journal, vol. 69, no. 2, pp. 418–427, 1995.
[92]  M. Mizogami, H. Tsuchiya, T. Ueno, M. Kashimata, and K. Takakura, “Stereospecific interaction of bupivacaine enantiomers with lipid membranes,” Regional Anesthesia and Pain Medicine, vol. 33, no. 4, pp. 304–311, 2008.
[93]  H. Tsuchiya and M. Mizogami, “R(+)-, Rac-, and S(?)-bupivacaine stereostructure-specifically interact with membrane lipids at cardiotoxically relevant concentrations,” Anesthesia and Analgesia, vol. 114, no. 2, pp. 310–312, 2012.
[94]  T. González, C. Arias, R. Caballero et al., “Effects of levobupivacaine, ropivacaine and bupivacaine on HERG channels: stereoselective bupivacaine block,” British Journal of Pharmacology, vol. 137, no. 8, pp. 1269–1279, 2002.
[95]  P. Friederich, A. Solth, S. Schillemeit, and D. Isbrandt, “Local anaesthetic sensitivities of cloned HERG channels from human heart: comparison with HERG/MiRP1 and HERG/MiRP1T8A,” British Journal of Anaesthesia, vol. 92, no. 1, pp. 93–101, 2004.
[96]  H. Tsuchiya and M. Mizogami, “The membrane interaction of drugs as one of mechanisms for their enantioselective effects,” Medical Hypotheses, vol. 79, no. 1, pp. 65–67, 2012.
[97]  N. Nandi and D. Vollhardt, “Chiral discrimination and recognition in Langmuir monolayers,” Current Opinion in Colloid and Interface Science, vol. 13, no. 1-2, pp. 40–46, 2008.
[98]  S. Lalitha, A. S. Kumar, K. J. Stine, and D. F. Covey, “Chirality in membranes: first evidence that enantioselective interactions between cholesterol and cell membrane lipids can be a determinant of membrane physical properties,” Journal of Supramolecular Chemistry, vol. 1, no. 2, pp. 53–61, 2001.
[99]  A. Kessel, N. Ben-Tal, and S. May, “Interactions of cholesterol with lipid bilayers: the preferred configuration and fluctuations,” Biophysical Journal, vol. 81, no. 2, pp. 643–658, 2001.
[100]  K. Simons and D. Toomre, “Lipid rafts and signal transduction,” Nature Reviews Molecular Cell Biology, vol. 1, no. 1, pp. 31–39, 2000.
[101]  M.-O. Parat, “Could endothelial caveolae be the target of general anaesthetics?” British Journal of Anaesthesia, vol. 96, no. 5, pp. 547–550, 2006.
[102]  K. M. S. O'Connell, J. R. Martens, and M. M. Tamkun, “Localization of ion channels to lipid raft domains within the cardiovascular system,” Trends in Cardiovascular Medicine, vol. 14, no. 2, pp. 37–42, 2004.
[103]  Y. Xiang, V. O. Rybin, S. F. Steinberg, and B. Kobilka, “Caveolar localization dictates physiologic signaling of β2-adrenoceptors in neonatal cardiac myocytes,” The Journal of Biological Chemistry, vol. 277, no. 37, pp. 34280–34286, 2002.
[104]  Y. Fujimura, H. Tachibana, and K. Yamada, “Lipid raft-associated catechin suppresses the FcεRI expression by inhibiting phosphorylation of the extracellular signal-regulated kinase1/2,” FEBS Letters, vol. 556, no. 1-3, pp. 204–210, 2004.
[105]  A. Ausuili, A. Torrecillas, F. J. Aranda et al., “Edelfosine is incorporated into rafts and alters their organization,” Journal of Physical Chemistry B, vol. 112, no. 37, pp. 11643–11654, 2008.
[106]  K. Kamata, S. Manno, M. Ozaki, and Y. Takakuwa, “Functional evidence for presence of lipid rafts in erythrocyte membranes: Gsα in rafts is essential for signal transduction,” American Journal of Hematology, vol. 83, no. 5, pp. 371–375, 2008.
[107]  K. Kamata, S. Manno, Y. Takakuwa, and M. Ozaki, “Reversible effect of lidocaine on raft formation,” International Congress Series, vol. 1283, pp. 281–282, 2005.
[108]  H. Tsuchiya, T. Ueno, M. Mizogami, and K. Takakura, “Do local anesthetics interact preferentially with membrane lipid rafts? Comparative interactivities with raft-like membranes,” Journal of Anesthesia, vol. 24, no. 4, pp. 639–642, 2010.
[109]  C. Bandeiras, A. P. Serro, K. Luzyanin, A. Fernandes, and B. Saramago, “Anesthetics interacting with lipid rafts,” European Journal of Pharmaceutical Sciences, vol. 48, no. 1-2, pp. 153–165, 2013.
[110]  C. Ioannides, “Drug-phytochemical interactions,” Inflammopharmacology, vol. 11, no. 1, pp. 7–42, 2003.
[111]  S. Zhou, Z. Zhou, C. Li et al., “Identification of drugs that interact with herbs in drug development,” Drug Discovery Today, vol. 12, no. 15-16, pp. 664–673, 2007.
[112]  Y. Vorobeychik, V. Gordin, J. Mao, and L. Chen, “Combination therapy for neuropathic pain: a review of current evidence,” CNS Drugs, vol. 25, no. 12, pp. 1023–1034, 2011.
[113]  D. ?picarová and J. Pale?ek, “The role of spinal cord vanilloid (TRPV1) receptors in pain modulation,” Physiological Research, vol. 57, supplement 3, pp. S69–S77, 2008.
[114]  M. J. Caterina, M. A. Schumacher, M. Tominaga, T. A. Rosen, J. D. Levine, and D. Julius, “The capsaicin receptor: a heat-activated ion channel in the pain pathway,” Nature, vol. 389, no. 6653, pp. 816–824, 1997.
[115]  K. Venkatachalam and C. Montell, “TRP channels,” Annual Review of Biochemistry, vol. 76, pp. 387–417, 2007.
[116]  A. M. Binshtok, B. P. Bean, and C. J. Woolf, “Inhibition of nociceptors by TRPV1-mediated entry of impermeant sodium channel blockers,” Nature, vol. 449, no. 7162, pp. 607–610, 2007.
[117]  C. R. Ries, R. Pillai, C. C. Chung, J. T. Wang, B. A. MacLeod, and S. K. Schwarz, “QX-314 produces long-lasting local anesthesia modulated by transient receptor potential vanilloid receptors in mice,” Anesthesiology, vol. 111, no. 1, pp. 122–126, 2009.
[118]  J. Shen, L. E. Fox, and J. Cheng, “Differential effects of peripheral versus central coadministration of QX-314 and capsaicin on neuropathic pain in rats,” Anesthesiology, vol. 117, no. 2, pp. 365–380, 2012.
[119]  P. Gerner, A. M. Binshtok, C. Wang et al., “Capsaicin combined with local anesthetics preferentially prolongs sensory/nociceptive block in rat sciatic nerve,” Anesthesiology, vol. 109, no. 5, pp. 872–878, 2008.
[120]  E. W. McCleskey, “Neuroscience: a local route to pain relief,” Nature, vol. 449, no. 7162, pp. 545–546, 2007.
[121]  H. Shin, H. J. Park, S. A. Raymond, and G. R. Strichartz, “Potentiation by capsaicin of lidocaine's tonic impulse block in isolated rat sciatic nerve,” Neuroscience Letters, vol. 174, no. 1, pp. 14–16, 1994.
[122]  H. Shin, H. J. Park, and S. A. Raymond, “Potentiation by capsaicin of lidocaine's phasic impulse block in isolated rat sciatic nerve,” Pharmacological Research, vol. 30, no. 1, pp. 73–79, 1994.
[123]  H. Tsuchiya, “Biphasic membrane effects of capsaicin, an active component in Capsicum species,” Journal of Ethnopharmacology, vol. 75, no. 2-3, pp. 295–299, 2001, Erratum in: Journal of Ethnopharmacology, vol. 76, no. 3, p. 313, 2001.
[124]  H. Tsuchiya and M. Mizogami, “Plant components exhibit pharmacological activities and drug interactions by acting on lipid membranes,” Pharmacognosy Communications, vol. 2, no. 4, pp. 58–71, 2012.
[125]  K. Yang, C. Tsai, Y. Wang et al., “Apple polyphenol phloretin potentiates the anticancer actions of paclitaxel through induction of apoptosis in human Hep G2 cells,” Molecular Carcinogenesis, vol. 48, no. 5, pp. 420–431, 2009.
[126]  R. Cseh and R. Benz, “Interaction of phloretin with lipid monolayers: relationship between structural changes and dipole potential change,” Biophysical Journal, vol. 77, no. 3, pp. 1477–1488, 1999.
[127]  B. G. Auner and C. Valenta, “Influence of phloretin on the skin permeation of lidocaine from semisolid preparations,” European Journal of Pharmaceutics and Biopharmaceutics, vol. 57, no. 2, pp. 307–312, 2004.
[128]  G. ?berg and G. Sydnes, “An inhibitory effect of polyphloretin phosphate (PPP) on local anaesthesia,” Acta Anaesthesiologica Scandinavica, vol. 19, no. 5, pp. 377–383, 1975.
[129]  A. Nordenram and K. Danielsson, “The effect of polyphloretin phosphate (PPP) on oral infiltration anaesthesia. An experimental investigation,” Swedish Dental Journal, vol. 12, no. 6, pp. 217–220, 1988.

Full-Text

comments powered by Disqus

Contact Us

service@oalib.com

QQ:3279437679

WhatsApp +8615387084133