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...

Expression and Role of the Intermediate-Conductance Calcium-Activated Potassium Channel KCa3.1 in Glioblastoma

DOI: 10.1155/2012/421564

Full-Text   Cite this paper   Add to My Lib

Abstract:

Glioblastomas are characterized by altered expression of several ion channels that have important consequences in cell functions associated with their aggressiveness, such as cell survival, proliferation, and migration. Data on the altered expression and function of the intermediate-conductance calcium-activated K (KCa3.1) channels in glioblastoma cells have only recently become available. This paper aims to (i) illustrate the main structural, biophysical, pharmacological, and modulatory properties of the KCa3.1 channel, (ii) provide a detailed account of data on the expression of this channel in glioblastoma cells, as compared to normal brain tissue, and (iii) critically discuss its major functional roles. Available data suggest that KCa3.1 channels (i) are highly expressed in glioblastoma cells but only scantly in the normal brain parenchima, (ii) play an important role in the control of glioblastoma cell migration. Altogether, these data suggest KCa3.1 channels as potential candidates for a targeted therapy against this tumor. 1. Introduction Glioblastomas are the most common and aggressive among primary brain tumors. In spite of the intensive basic and clinical studies, only minor successes have been witnessed over the last decades. One-third of patients keep surviving no longer than one year from diagnosis, and average life expectancy remains dismal (12–15 months), even when radical surgical resection, chemo- and radiotherapy can be applied. The major problem with glioblastomas is their highly migratory and invasive potential into the normal brain tissue that prevents complete surgical removal of tumor cells and the extreme resistance of these cells to standard treatments [1]. To worsen the outcome of the disease is the presence in the tumor mass of a recently identified subpopulation of highly tumorigenic stem-like glioblastoma cells possessing even more invasive power, chemo- and radio-resistance than nonstem tumor cells, that are also thought to be responsible for the commonly observed tumor relapses [2–4]. Glioblastomas are characterized by a large number and variety of genetic mutations that heavily disregulate the major signaling pathways controlling cell survival, proliferation, differentiation, and invasion [5]. Among the disregulated pathways found in glioblastoma cells there are those controlling the expression of ion channels, transmembrane proteins endowed with a permeation pore that allows the passage of ions. Usually ion channels are selectively permeable to one particular ion and can open and close their permeation pore in response

References

[1]  F. B. Furnari, T. Fenton, R. M. Bachoo et al., “Malignant astrocytic glioma: genetics, biology, and paths to treatment,” Genes and Development, vol. 21, no. 21, pp. 2683–2710, 2007.
[2]  S. Bao, Q. Wu, R. E. McLendon et al., “Glioma stem cells promote radioresistance by preferential activation of the DNA damage response,” Nature, vol. 444, no. 7120, pp. 756–760, 2006.
[3]  G. Liu, X. Yuan, Z. Zeng et al., “Analysis of gene expression and chemoresistance of CD133+ cancer stem cells in glioblastoma,” Molecular Cancer, vol. 5, article 67, 2006.
[4]  A. M. Bleau, J. T. Huse, and E. C. Holland, “The ABCG2 resistance network of glioblastoma,” Cell Cycle, vol. 8, no. 18, pp. 2936–2944, 2009.
[5]  B. Purow and D. Schiff, “Advances in the genetics of glioblastoma: are we reaching critical mass?” Nature Reviews Neurology, vol. 5, no. 8, pp. 419–426, 2009.
[6]  B. Hille, Ion Channels of Excitable Membranes, Sinauer Associates, Sunderland, Mass, USA, 3rd edition, 2001.
[7]  K. Kunzelmann, “Ion channels and cancer,” Journal of Membrane Biology, vol. 205, no. 3, pp. 159–173, 2005.
[8]  A. Arcangeli, O. Crociani, E. Lastraioli, A. Masi, S. Pillozzi, and A. Becchetti, “Targeting ion channels in cancer: a novel frontier in antineoplastic therapy,” Current Medicinal Chemistry, vol. 16, no. 1, pp. 66–93, 2009.
[9]  M. L. Olsen, S. Schade, S. A. Lyons, M. D. Amaral, and H. Sontheimer, “Expression of voltage-gated chloride channels in human glioma cells,” The Journal of Neuroscience, vol. 23, no. 13, pp. 5572–5582, 2003.
[10]  C. W. Habela, M. L. Olsen, and H. Sontheimer, “ClC3 is a critical regulator of the cell cycle in normal and malignant glial cells,” The Journal of Neuroscience, vol. 28, no. 37, pp. 9205–9217, 2008.
[11]  C. W. Habela, N. J. Ernest, A. F. Swindall, and H. Sontheimer, “Chloride accumulation drives volume dynamics underlying cell proliferation and migration,” Journal of Neurophysiology, vol. 101, no. 2, pp. 750–757, 2009.
[12]  V. A. Cuddapah, C. W. Habela, S. Watkins, et al., “Kinase activation of ClC-3 accelerates cytoplasmic condensation during mitotic cell rounding,” American Journal of Physiology, vol. 302, no. 3, pp. C527–C538, 2012.
[13]  N. J. Ernest, C. W. Habela, and H. Sontheimer, “Cytoplasmic condensation is both necessary and sufficient to induce apoptotic cell death,” Journal of Cell Science, vol. 121, part 3, pp. 290–297, 2008.
[14]  X. Liu, Y. Chang, P. H. Reinhart, and H. Sontheimer, “Cloning and characterization of glioma BK, a novel BK channel isoform highly expressed in human glioma cells,” The Journal of Neuroscience, vol. 22, no. 5, pp. 1840–1849, 2002.
[15]  V. A. Cuddapah and H. Sontheimer, “Ion channels and transporters in cancer. 2. Ion channels and the control of cancer cell migration,” American Journal of Physiology, vol. 301, no. 3, pp. C541–C549, 2011.
[16]  H. Sontheimer, “Ion channels and amino acid transporters support the growth and invasion of primary brain tumors,” Molecular Neurobiology, vol. 29, no. 1, pp. 61–71, 2004.
[17]  M. F. Ritchie, Y. Zhou, and J. Soboloff, “WT1/EGR1-mediated control of STIM1 expression and function in cancer cells,” Frontiers in Bioscience, vol. 16, pp. 2402–2415, 2011.
[18]  X. Ding, Z. He, K. Zhou et al., “Essential role of TRPC6 channels in G2/M phase transition and development of human glioma,” Journal of the National Cancer Institute, vol. 102, no. 14, pp. 1052–1068, 2010.
[19]  S. Chigurupati, R. Venkataraman, D. Barrera et al., “Receptor channel TRPC6 is a key mediator of Notch-driven glioblastoma growth and invasiveness,” Cancer Research, vol. 70, no. 1, pp. 418–427, 2010.
[20]  V. C. Bomben and H. Sontheimer, “Disruption of transient receptor potential canonical channel 1 causes incomplete cytokinesis and slows the growth of human malignant gliomas,” GLIA, vol. 58, no. 10, pp. 1145–1156, 2010.
[21]  G. Giannone, P. Rondé, M. Gaire, J. Haiech, and K. Takeda, “Calcium oscillations trigger focal adhesion disassembly in human U87 astrocytoma cells,” Journal of Biological Chemistry, vol. 277, no. 29, pp. 26364–26371, 2002.
[22]  P. Rondé, G. Giannone, I. Gerasymova, H. Stoeckel, K. Takeda, and J. Haiech, “Mechanism of calcium oscillations in migrating human astrocytoma cells,” Biochimica et Biophysica Acta, vol. 1498, no. 2-3, pp. 273–280, 2000.
[23]  G. Reetz and G. Reiser, “[Ca2+]i oscillations induced by bradykinin in rat glioma cells associated with Ca2+ store-dependent Ca2+ influx are controlled by cell volume and by membrane potential,” Cell Calcium, vol. 19, no. 2, pp. 143–156, 1996.
[24]  H. Berkefeld, B. Fakler, and U. Schulte, “Ca2+-activated K+ channels: from protein complexes to function,” Physiological Reviews, vol. 90, no. 4, pp. 1437–1459, 2010.
[25]  C. C. Chou, C. A. Lunn, and N. J. Murgolo, “KCa3.1: target and marker for cancer, autoimmune disorder and vascular inflammation?” Expert Review of Molecular Diagnostics, vol. 8, no. 2, pp. 179–187, 2008.
[26]  H. Wulff and N. A. Castle, “Therapeutic potential of KCa3.1 blockers: recent advances and promising trends,” Expert Review of Clinical Pharmacology, vol. 3, no. 3, pp. 385–396, 2010.
[27]  A. Schwab, V. Nechyporuk-Zloy, A. Fabian, and C. Stock, “Cells move when ions and water flow,” Pflügers Archiv European Journal of Physiology, vol. 453, no. 4, pp. 421–432, 2007.
[28]  X. M. Xia, B. Fakler, A. Rivard et al., “Mechanism of calcium gating in small-conductance calcium-activated potassium channels,” Nature, vol. 395, no. 6701, pp. 503–507, 1998.
[29]  J. E. Keen, R. Khawaled, D. L. Farrens et al., “Domains responsible for constitutive and Ca2+-dependent interactions between calmodulin and small conductance Ca2+-activated potassium channels,” The Journal of Neuroscience, vol. 19, no. 20, pp. 8830–8838, 1999.
[30]  W. Li, D. B. Halling, A. W. Hall, and R. W. Aldrich, “EF hands at the N-lobe of calmodulin are required for both SK channel gating and stable SK-calmodulin interaction,” Journal of General Physiology, vol. 134, no. 4, pp. 281–293, 2009.
[31]  P. Pedarzani and M. Stocker, “Molecular and cellular basis of small- and intermediate-conductance, calcium-activated potassium channel function in the brain,” Cellular and Molecular Life Sciences, vol. 65, no. 20, pp. 3196–3217, 2008.
[32]  B. Fioretti, E. Castigli, M. R. Micheli et al., “Expression and modulation of the intermediate- conductance Ca2+-activated K+ channel in glioblastoma GL-15 cells,” Cellular Physiology and Biochemistry, vol. 18, no. 1–3, pp. 47–56, 2006.
[33]  W. J. Joiner, L. U. Y. Wang, M. D. Tang, and L. K. Kaczmarek, “hSK4, a member of a novel subfamily of calcium-activated potassium channels,” Proceedings of the National Academy of Sciences of the United States of America, vol. 94, no. 20, pp. 11013–11018, 1997.
[34]  N. J. Logsdon, J. Kang, J. A. Togo, E. P. Christian, and J. Aiyar, “A novel gene, hKCa4, encodes the calcium-activated potassium channel in human T lymphocytes,” Journal of Biological Chemistry, vol. 272, no. 52, pp. 32723–32726, 1997.
[35]  T. M. Ishii, C. Silvia, B. Hirschberg, C. T. Bond, J. P. Adelman, and J. Maylie, “A human intermediate conductance calcium-activated potassium channel,” Proceedings of the National Academy of Sciences of the United States of America, vol. 94, no. 21, pp. 11651–11656, 1997.
[36]  M. A. Bailey, M. Grabe, and D. C. Devor, “Characterization of the PCMBS-dependent modification of KCa3.1 channel gating,” Journal of General Physiology, vol. 136, no. 4, pp. 367–387, 2010.
[37]  B. Hirschberg, J. Maylie, J. P. Adelman, and N. V. Marrion, “Gating of recombinant small-conductance Ca-activated K+ channels by calcium,” Journal of General Physiology, vol. 111, no. 4, pp. 565–581, 1998.
[38]  J. Ledoux, A. D. Bonev, and M. T. Nelson, “Ca2+-activated K+ channels in murine endothelial cells: block by intracellular calcium and magnesium,” Journal of General Physiology, vol. 131, no. 2, pp. 125–135, 2008.
[39]  C. Barmeyer, C. Rahner, Y. Yang, F. J. Sigworth, H. J. Binder, and V. M. Rajendran, “Cloning and identification of tissue-specific expression of KCNN4 splice variants in rat colon,” American Journal of Physiology, vol. 299, no. 2, pp. C251–C263, 2010.
[40]  H. Wulff, A. Kolski-Andreaco, A. Sankaranarayanan, J. M. Sabatier, and V. Shakkottai, “Modulators of small- and intermediate-conductance calcium-activated potassium channels and their therapeutic indications,” Current Medicinal Chemistry, vol. 14, no. 13, pp. 1437–1457, 2007.
[41]  N. A. Castle, D. O. London, C. Creech, Z. Fajloun, J. W. Stocker, and J. M. Sabatier, “Maurotoxin: a potent inhibitor of intermediate conductance Ca2+-activated potassium channels,” Molecular Pharmacology, vol. 63, no. 2, pp. 409–418, 2003.
[42]  V. Visan, Z. Fajloun, J. M. Sabatier, and S. Grissmer, “Mapping of maurotoxin binding sites on hKv1.2, hKv1.3, and hlKCa1 channels,” Molecular Pharmacology, vol. 66, no. 5, pp. 1103–1112, 2004.
[43]  H. Rauer, M. D. Lanigan, M. W. Pennington et al., “Structure-guided transformation of charybdotoxin yields an analog that selectively targets Ca2+-activated over voltage-gated K+ Channels,” The Journal of Biological Chemistry, vol. 275, no. 2, pp. 1201–1208, 2000.
[44]  L. R. Berkowitz and E. P. Orringer, “Effect of cetiecil, an in vitro antisickling agent, on erythrocyte membrane cation permeability,” Journal of Clinical Investigation, vol. 68, no. 5, pp. 1215–1220, 1981.
[45]  L. R. Berkowitz and E. P. Orringer, “Effects of cetiedil on monovalent cation permeability in the erythrocyte: an explanation for the efficacy of cetiedil in the treatment of sickle cell anemia,” Blood Cells, vol. 8, no. 2, pp. 283–288, 1982.
[46]  J. Alvarez, M. Montero, and J. Garcia-Sancho, “High affinity inhibition of Ca2+-dependent K+ channels by cytochrome P-450 inhibitors,” The Journal of Biological Chemistry, vol. 267, no. 17, pp. 11789–11793, 1992.
[47]  J. C. Ellory, S. J. Culliford, P. A. Smith, M. W. Wolowyk, and E. E. Knaus, “Specific inhibition of Ca-activated K channels in red cells by selected dihydropyridine derivatives,” British Journal of Pharmacology, vol. 111, no. 3, pp. 903–905, 1994.
[48]  C. J. Roxburgh, C. R. Ganellin, S. Athmani et al., “Synthesis and structure-activity relationships of cetiedil analogues as blockers of the Ca2+-activated K+ permeability of erythrocytes,” Journal of Medicinal Chemistry, vol. 44, no. 20, pp. 3244–3253, 2001.
[49]  H. Wulff, M. J. Miller, W. H?nsel, S. Grissmer, M. D. Cahalan, and K. G. Chandy, “Design of a potent and selective inhibitor of the intermediate-conductance Ca2+-activated K+ channel, IKCa1: a potential immunosuppressant,” Proceedings of the National Academy of Sciences of the United States of America, vol. 97, no. 14, pp. 8151–8156, 2000.
[50]  J. W. Stocker, L. De Franceschi, G. A. McNaughton-Smith, R. Corrocher, Y. Beuzard, and C. Brugnara, “ICA-17043, a novel Gardos channel blocker, prevents sickled red blood cell dehydration in vitro and in vivo in SAD mice,” Blood, vol. 101, no. 6, pp. 2412–2418, 2003.
[51]  H. Wulff, G. A. Gutman, M. D. Cahalan, and K. G. Chandy, “Delineation of the clotrimazole/TRAM-34 binding site on the intermediate conductance calcium-activated potassium channel, IKCa1,” The Journal of Biological Chemistry, vol. 276, no. 34, pp. 32040–32045, 2001.
[52]  K. Urbahns, E. Horváth, J. P. Stasch, and F. Mauler, “4-Phenyl-4H-pyrans as channel blockers,” Bioorganic & Medicinal Chemistry Letters, vol. 13, no. 16, pp. 2637–2639, 2003.
[53]  K. Urbahns, S. Goldmann, J. Krüger et al., “ -channel blockers. Part 2: discovery of cyclohexadienes,” Bioorganic & Medicinal Chemistry Letters, vol. 15, no. 2, pp. 401–404, 2005.
[54]  D. C. Devor, A. K. Singh, R. A. Frizzell, and R. J. Bridges, “Modulation of Cl- secretion by benzimidazolones. I. Direct activation of a Ca2+-dependent K+ channel,” American Journal of Physiology, vol. 271, no. 5, part 1, pp. L775–L784, 1996.
[55]  S. Singh, C. A. Syme, A. K. Singh, D. C. Devor, and R. J. Bridges, “Benzimidazolone activators of chloride secretion: potential therapeutics for cystic fibrosis and chronic obstructive pulmonary disease,” The Journal of Pharmacology and Experimental Therapeutics, vol. 296, no. 2, pp. 600–611, 2001.
[56]  D. Str?baek, L. Teuber, T. D. J?rgensen et al., “Activation of human IK and SK Ca2+-activated K+ channels by NS309 (6,7-dichloro-1H-indole-2,3-dione 3-oxime),” Biochimica et Biophysica Acta, vol. 1665, no. 1-2, pp. 1–5, 2004.
[57]  C. A. Syme, A. C. Gerlach, A. K. Singh, and D. C. Devor, “Pharmacological activation of cloned intermediate- and small-conductance Ca2+-activated K+ channels,” American Journal of Physiology, vol. 278, no. 3, pp. C570–C581, 2000.
[58]  K. A. Pedersen, R. L. Schro?der, B. Skaaning-Jensen, D. Str?b?k, S. P. Olesen, and P. Christophersen, “Activation of the human intermediate-conductance Ca2+-activated K+ channel by 1-ethyl-2-benzimidazolinone is strongly Ca2+-dependent,” Biochimica et Biophysica Acta, vol. 1420, no. 1-2, pp. 231–240, 1999.
[59]  A. C. Gerlach, N. N. Gangopadhyay, and D. C. Devor, “Kinase-dependent regulation of the intermediate conductance, calcium- dependent potassium channel, hIK1,” The Journal of Biological Chemistry, vol. 275, no. 1, pp. 585–598, 2000.
[60]  S. Srivastava, Z. Li, L. Lin et al., “The phosphatidylinositol 3-phosphate phosphatase myotubularin- related protein 6 (MTMR6) is a negative regulator of the Ca2+-activated K+ channel 3.1,” Molecular and Cellular Biology, vol. 25, no. 9, pp. 3630–3638, 2005.
[61]  A. Wulf and A. Schwab, “Regulation of a calcium-sensitive K+ channel (cIK1) by protein kinase C,” Journal of Membrane Biology, vol. 187, no. 1, pp. 71–79, 2002.
[62]  S. Srivastava, P. Choudhury, Z. Li et al., “Phosphatidylinositol 3-phosphate indirectly activates KCa3.1 via 14 amino acids in the carboxy terminus of KCa3.1,” Molecular Biology of the Cell, vol. 17, no. 1, pp. 146–154, 2006.
[63]  S. Srivastava, Z. Li, K. Ko et al., “Histidine phosphorylation of the potassium channel KCa3.1 by nucleoside diphosphate kinase B is required for activation of KCa3.1 and CD4 T cells,” Molecular Cell, vol. 24, no. 5, pp. 665–675, 2006.
[64]  H. Klein, L. Garneau, N. T. N. Trinh et al., “Inhibition of the KCa3.1 channels by AMP-activated protein kinase in human airway epithelial cells,” American Journal of Physiology, vol. 296, no. 2, pp. C285–C295, 2009.
[65]  S. Srivastava, L. Di, O. Zhdanova et al., “The class II phosphatidylinositol 3 kinase C2β is required for the activation of the K+ channel KCa3.1 and CD4 T-cells,” Molecular Biology of the Cell, vol. 20, no. 17, pp. 3783–3791, 2009.
[66]  C. M. Balut, Y. Gao, C. Luke, and D. C. Devor, “Immunofluorescence-based assay to identify modulators of the number of plasma membrane KCa3.1 channels,” Future Medicinal Chemistry, vol. 2, no. 5, pp. 707–713, 2010.
[67]  Y. Gao, C. M. Balut, M. A. Bailey, G. Patino-Lopez, S. Shaw, and D. C. Devor, “Recycling of the Ca2+-activated K+ channel, KCa2.3, is dependent upon RME-1, Rab35/EPI64C, and an N-terminal domain,” The Journal of Biological Chemistry, vol. 285, no. 23, pp. 17938–17953, 2010.
[68]  C. M. Balut, Y. Gao, S. A. Murray, P. H. Thibodeau, and D. C. Devor, “ESCRT-dependent targeting of plasma membrane localized KCa3.1 to the lysosomes,” American Journal of Physiology, vol. 299, no. 5, pp. C1015–C1027, 2010.
[69]  C. M. Balut, C. M. Loch, and D. C. Devor, “Role of ubiquitylation and USP8-dependent deubiquitylation in the endocytosis and lysosomal targeting of plasma membrane KCa3.1,” The FASEB Journal, vol. 25, no. 11, pp. 3938–3948, 2011.
[70]  M. Sciaccaluga, B. Fioretti, L. Catacuzzeno et al., “CXCL12-induced glioblastoma cell migration requires intermediate conductance Ca2+-activated K+ channel activity,” American Journal of Physiology, vol. 299, no. 1, pp. C175–C184, 2010.
[71]  S. Ghanshani, H. Wulff, M. J. Miller et al., “Up-regulation of the IKCa1 potassium channel during T-cell activation: molecular mechanism and functional consequences,” The Journal of Biological Chemistry, vol. 275, no. 47, pp. 37137–37149, 2000.
[72]  E. C. Holland, “Gliomagenesis: genetic alterations and mouse models,” Nature Reviews Genetics, vol. 2, no. 2, pp. 120–129, 2001.
[73]  A. Cheong, A. J. Bingham, J. Li et al., “Downregulated REST transcription factor is a switch enabling critical potassium channel expression and cell proliferation,” Molecular Cell, vol. 20, no. 1, pp. 45–52, 2005.
[74]  Z. Gao, K. Ure, P. Ding et al., “The master negative regulator REST/NRSF controls adult neurogenesis by restraining the neurogenic program in quiescent stem cells,” The Journal of Neuroscience, vol. 31, no. 26, pp. 9772–9786, 2011.
[75]  N. Ballas, C. Grunseich, D. D. Lu, J. C. Speh, and G. Mandel, “REST and its corepressors mediate plasticity of neuronal gene chromatin throughout neurogenesis,” Cell, vol. 121, no. 4, pp. 645–657, 2005.
[76]  P. W. L. Tas, H. G. Kress, and K. Koschel, “Presence of a charybdotoxin sensitive Ca2+-activated K+ channel in rat glioma C6 cells,” Neuroscience Letters, vol. 94, no. 3, pp. 279–284, 1988.
[77]  F. A. De-Allie, S. R. Bolsover, A. V. Nowicky, and P. N. Strong, “Characterization of Ca2+-activated 86Rb+ fluxes in rat C6 glioma cells: a system for identifying novel IKCa-channel toxins,” British Journal of Pharmacology, vol. 117, no. 3, pp. 479–487, 1996.
[78]  D. Manor and N. Moran, “Modulation of small conductance calcium-activated potassium channels in C6 glioma cells,” Journal of Membrane Biology, vol. 140, no. 1, pp. 69–79, 1994.
[79]  S. Supattapone and C. C. Ashley, “Endothelin opens potassium channels in glial cells,” European Journal of Neuroscience, vol. 3, no. 4, pp. 349–355, 1991.
[80]  M. M. Gleason, E. C. Griffin, P. Nambi, and N. Aiyar, “Endothelin stimulates 86Rb efflux in rat glioma C6-Bu-1 cells,” Neuropeptides, vol. 20, no. 1, pp. 17–23, 1991.
[81]  G. Reiser, F. Donié, and F. J. Binm?ller, “Serotonin regulates cytosolic Ca2+ activity and membrane potential in a neuronal and in a glial cell line via 5-HT3 and 5-HT2 receptors by different mechanisms,” Journal of Cell Science, vol. 93, part 3, pp. 545–555, 1989.
[82]  G. Reiser, F. J. Binm?ller, P. N. Strong, and B. Hamprecht, “Activation of a K+ conductance by bradykinin and by inositol-1,4,5-trisphosphate in rat glioma cells: involvement of intracellular and extracellular Ca2+,” Brain Research, vol. 506, no. 2, pp. 205–214, 1990.
[83]  A. Ogura and T. Amano, “Serotonin-receptor coupled with membrane electrogenesis in a rat glioma clone,” Brain Research, vol. 297, no. 2, pp. 387–391, 1984.
[84]  T. Weiger, D. R. Stevens, L. Wunder, and H. L. Haas, “Histamine H1 receptors in C6 glial cells are coupled to calcium-dependent potassium channels via release of calcium from internal stores,” Naunyn-Schmiedeberg's Archives of Pharmacology, vol. 355, no. 5, pp. 559–565, 1997.
[85]  B. Fioretti, E. Castigli, I. Calzuola, A. A. Harper, F. Franciolini, and L. Catacuzzeno, “NPPB block of the intermediate-conductance Ca2+-activated K+ channel,” European Journal of Pharmacology, vol. 497, no. 1, pp. 1–6, 2004.
[86]  A. K. Weaver, V. C. Bomben, and H. Sontheimer, “Expression and function of calcium-activated potassium channels in human glioma cells,” GLIA, vol. 54, no. 3, pp. 223–233, 2006.
[87]  L. Nowak, P. Ascher, and Y. Berwald-Netter, “Ionic channels in mouse astrocytes in culture,” The Journal of Neuroscience, vol. 7, no. 1, pp. 101–109, 1987.
[88]  I. F. Abdullaev, A. Rudkouskaya, A. A. Mongin, and Y. H. Kuo, “Calcium-activated potassium channels BK and IK1 are functionally expressed in human gliomas but do not regulate cell proliferation,” Plos ONE, vol. 5, no. 8, Article ID e12304, 2010.
[89]  H. Sontheimer, M. Perouansky, D. Hoppe, H. D. Lux, R. Grantyn, and H. Kettenmann, “Glial cells of the oligodendrocyte lineage express proton-activated Na+ channels,” Journal of Neuroscience Research, vol. 24, no. 4, pp. 496–500, 1989.
[90]  A. Bringmann, B. Biedermann, and A. Reichenbach, “Expression of potassium channels during postnatal differentiation of rabbit Müller glial cells,” European Journal of Neuroscience, vol. 11, no. 8, pp. 2883–2896, 1999.
[91]  R. Khanna, L. Roy, X. Zhu, and L. C. Schlichter, “K+ channels and the microglial respiratory burst,” American Journal of Physiology, vol. 280, no. 4, pp. C796–C806, 2001.
[92]  V. Kaushal, P. D. Koeberle, Y. Wang, and L. C. Schlichter, “The Ca2+-activated K+ channel KCNN4/KCa3.1 contributes to microglia activation and nitric oxide-dependent neurodegeneration,” The Journal of Neuroscience, vol. 27, no. 1, pp. 234–244, 2007.
[93]  D. Bouhy, N. Ghasemlou, S. Lively et al., “Inhibition of the Ca2+-dependent K+ channel, KCNN4/KCa3.1, improves tissue protection and locomotor recovery after spinal cord injury,” The Journal of Neuroscience, vol. 31, no. 45, pp. 16298–16308, 2011.
[94]  F. Vogalis, J. R. Harvey, C. B. Neylon, and J. B. Furness, “Regulation of K+ channels underlying the slow afterhyperpolarization in enteric afterhyperpolarization-generating myenteric neurons: role of calcium and phosphorylation,” Clinical and Experimental Pharmacology and Physiology, vol. 29, no. 10, pp. 935–943, 2002.
[95]  J. B. Furness, K. Kearney, H. L. Robbins et al., “Intermediate conductance potassium (IK) channels occur in human enteric neurons,” Autonomic Neuroscience, vol. 112, no. 1-2, pp. 93–97, 2004.
[96]  C. B. Neylon, C. J. Fowler, and J. B. Furness, “Regulation of the slow afterhyperpolarization in enteric neurons by protein kinase A,” Autonomic Neuroscience, vol. 126-127, pp. 258–263, 2006.
[97]  J. D. Engbers, D. Anderson, H. Asmara, et al., “Intermediate conductance calcium-activated potassium channels modulate summation of parallel fiber input in cerebellar Purkinje cells,” Proceedings of the National Academy of Sciences ot the United States of America, vol. 109, no. 7, pp. 2601–2606, 2012.
[98]  P. K. Bahia, R. Suzuki, D. C. H. Benton et al., “A functional role for small-conductance calcium-activated potassium channels in sensory pathways including nociceptive processes,” The Journal of Neuroscience, vol. 25, no. 14, pp. 3489–3498, 2005.
[99]  T. A. Longden, K. M. Dunn, H. J. Draheim, M. T. Nelson, A. H. Weston, and G. Edwards, “Intermediate-conductance calcium-activated potassium channels participate in neurovascular coupling,” British Journal of Pharmacology, vol. 164, no. 3, pp. 922–933, 2011.
[100]  H. Liu, Y. Li, and K. P. Raisch, “Clotrimazole induces a late G1 cell cycle arrest and sensitizes glioblastoma cells to radiation in vitro,” Anti-Cancer Drugs, vol. 21, no. 9, pp. 841–849, 2010.
[101]  M. Humayun Khalid, Y. Tokunaga, A. J. Caputy, and E. Walters, “Inhibition of tumor growth and prolonged survival of rats with intracranial gliomas following administration of clotrimazole,” Journal of Neurosurgery, vol. 103, no. 1, pp. 79–86, 2005.
[102]  M. H. Khalid, S. Shibata, and T. Hiura, “Effects of clotrimazole on the growth, morphological characteristics, and cisplatin sensitivity of human glioblastoma cells in vitro,” Journal of Neurosurgery, vol. 90, no. 5, pp. 918–927, 1999.
[103]  W. F. Wonderlin and J. S. Strobl, “Potassium channels, proliferation and G1 progression,” Journal of Membrane Biology, vol. 154, no. 2, pp. 91–107, 1996.
[104]  A. Bajetto, F. Barbieri, A. Dorcaratto et al., “Expression of CXC chemokine receptors 1-5 and their ligands in human glioma tissues: role of CXCR4 and SDF1 in glioma cell proliferation and migration,” Neurochemistry International, vol. 49, no. 5, pp. 423–432, 2006.
[105]  S. Barbero, A. Bajetto, R. Bonavia et al., “Expression of the chemokine receptor CXCR4 and its ligand stromal cell-derived factor 1 in human brain tumors and their involvement in glial proliferation in vitro,” Annals of the New York Academy of Sciences, vol. 973, pp. 60–69, 2002.
[106]  A. Salmaggi, M. Gelati, B. Pollo et al., “CXCL12 expression is predictive of a shorter time to tumor progression in low-grade glioma: a single-institution study in 50 patients,” Journal of Neuro-Oncology, vol. 74, no. 3, pp. 287–293, 2005.
[107]  J. Zhang, S. Sarkar, and V. W. Yong, “The chemokine stromal cell derived factor-1 (CXCL12) promotes glioma invasiveness through MT2-matrix metalloproteinase,” Carcinogenesis, vol. 26, no. 12, pp. 2069–2077, 2005.
[108]  N. De La Iglesia, G. Konopka, K. L. Lim et al., “Deregulation of a STAT3-interleukin 8 signaling pathway promotes human glioblastoma cell proliferation and invasiveness,” The Journal of Neuroscience, vol. 28, no. 23, pp. 5870–5878, 2008.
[109]  M. Ehtesham, J. A. Winston, P. Kabos, and R. C. Thompson, “CXCR4 expression mediates glioma cell invasiveness,” Oncogene, vol. 25, no. 19, pp. 2801–2806, 2006.
[110]  K. Wakabayashi, F. Kambe, X. Cao et al., “Inhibitory effects of cyclosporin A on calcium mobilization-dependent interleukin-8 expression and invasive potential of human glioblastoma U251MG cells,” Oncogene, vol. 23, no. 41, pp. 6924–6932, 2004.
[111]  M. E. Berens, M. D. Rief, J. R. Shapiro et al., “Proliferation and motility responses of primary and recurrent gliomas related to changes in epidermal growth factor receptor expression,” Journal of Neuro-Oncology, vol. 27, no. 1, pp. 11–22, 1996.
[112]  R. J. Seitz and W. Wechsler, “Immunohistochemical demonstration of serum proteins in human cerebral gliomas,” Acta Neuropathologica, vol. 73, no. 2, pp. 145–152, 1987.
[113]  C. V. Lund, M. T. N. Nguyen, G. C. Owens et al., “Reduced glioma infiltration in Src-deficient mice,” Journal of Neuro-Oncology, vol. 78, no. 1, pp. 19–29, 2006.
[114]  L. Catacuzzeno, F. Aiello, B. Fioretti et al., “Serum-activated K and Cl currents underlay U87-MG glioblastoma cell migration,” Journal of Cellular Physiology, vol. 226, no. 7, pp. 1926–1933, 2011.
[115]  L. Soroceanu, T. J. Manning Jr., and H. Sontheimer, “Modulation of glioma cell migration and invasion using Cl- and K+ ion channel blockers,” The Journal of Neuroscience, vol. 19, no. 14, pp. 5942–5954, 1999.
[116]  Y. D. Gao, P. J. Hanley, S. Rinné, M. Zuzarte, and J. Daut, “Calcium-activated K+ channel (KCa3.1) activity during Ca2+ store depletion and store-operated Ca2+ entry in human macrophages,” Cell Calcium, vol. 48, no. 1, pp. 19–27, 2010.
[117]  K. Funabashi, S. Ohya, H. Yamamura et al., “Accelerated Ca2+ entry by membrane hyperpolarization due to Ca2+-activated K+ channel activation in response to histamine in chondrocytes,” American Journal of Physiology, vol. 298, no. 4, pp. C786–C797, 2010.
[118]  E. Shumilina, R. S. Lam, F. W?lbing et al., “Blunted IgE-mediated activation of mast cells in mice lacking the Ca2+-activated K+ channel KCa3.1,” The Journal of Immunology, vol. 180, no. 12, pp. 8040–8047, 2008.
[119]  B. Fioretti, L. Catacuzzeno, L. Sforna et al., “Histamine hyperpolarizes human glioblastoma cells by activating the intermediate-conductance Ca2+-activated K+ channel,” American Journal of Physiology, vol. 297, no. 1, pp. C102–C110, 2009.
[120]  L. Catacuzzeno, B. Fioretti, and F. Franciolini, “A theoretical study on the role of Ca2+-activated K+ channels in the regulation of hormone-induced Ca2+ oscillations and their synchronization in adjacent cells,” Journal of Theoretical Biology. In press.
[121]  B. Fioretti, F. Franciolini, and L. Catacuzzeno, “A model of intracellular Ca2+ oscillations based on the activity of the intermediate-conductance Ca2+-activated K+ channels,” Biophysical Chemistry, vol. 113, no. 1, pp. 17–23, 2005.
[122]  J. Folkman, “Tumor suppression by p53 is mediated in part by the antiangiogenic activity of endostatin and tumstatin,” Science's STKE, vol. 2006, no. 354, p. pe35, 2006.
[123]  D. Hanahan and J. Folkman, “Patterns and emerging mechanisms of the angiogenic switch during tumorigenesis,” Cell, vol. 86, no. 3, pp. 353–364, 1996.
[124]  P. Kleihues, D. N. Louis, B. W. Scheithauer et al., “The WHO classification of tumors of the nervous system,” Journal of Neuropathology & Experimental Neurology, vol. 61, no. 3, pp. 215–225, 2002.

Full-Text

comments powered by Disqus

Contact Us

service@oalib.com

QQ:3279437679

WhatsApp +8615387084133