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...

Pathogenic Role of Store-Operated and Receptor-Operated Channels in Pulmonary Arterial Hypertension

DOI: 10.1155/2012/951497

Full-Text   Cite this paper   Add to My Lib

Abstract:

Pulmonary circulation is an important circulatory system in which the body brings in oxygen. Pulmonary arterial hypertension (PAH) is a progressive and fatal disease that predominantly affects women. Sustained pulmonary vasoconstriction, excessive pulmonary vascular remodeling, in situ thrombosis, and increased pulmonary vascular stiffness are the major causes for the elevated pulmonary vascular resistance (PVR) in patients with PAH. The elevated PVR causes an increase in afterload in the right ventricle, leading to right ventricular hypertrophy, right heart failure, and eventually death. Understanding the pathogenic mechanisms of PAH is important for developing more effective therapeutic approach for the disease. An increase in cytosolic free concentration ( ) in pulmonary arterial smooth muscle cells (PASMC) is a major trigger for pulmonary vasoconstriction and an important stimulus for PASMC migration and proliferation which lead to pulmonary vascular wall thickening and remodeling. It is thus pertinent to define the pathogenic role of signaling in pulmonary vasoconstriction and PASMC proliferation to develop new therapies for PAH. in PASMC is increased by influx through channels in the plasma membrane and by release or mobilization from the intracellular stores, such as sarcoplasmic reticulum (SR) or endoplasmic reticulum (ER). There are two entry pathways, voltage-dependent influx through voltage-dependent channels (VDCC) and voltage-independent influx through store-operated Ca2+ channels (SOC) and receptor-operated channels (ROC). This paper will focus on the potential role of VDCC, SOC, and ROC in the development and progression of sustained pulmonary vasoconstriction and excessive pulmonary vascular remodeling in PAH. 1. Introduction The only organ in the body to receive the entire cardiac output (CO) at one time is the lung. To receive a high flow of volume from the entire CO, the pulmonary circulatory system must maintain a low-resistance and low-pressure system to carry blood to the pulmonary capillaries. Deoxygenated venous blood flows through the pulmonary artery to the pulmonary capillaries where oxygen and carbon dioxide gas exchange occurs. Pulmonary hypertension (PH) is a severe chronic disorder that affects the pulmonary circulatory system. This disorder is often a deadly hemodynamic irregularity that may be idiopathic, heritable, or secondary to other diseases such as chronic obstructive pulmonary disease (COPD). Pulmonary arteries are thin and have low myogenic tone compared to systemic arteries. Therefore, pulmonary arteries rely on

References

[1]  A. L. Firth, J. Mandel, and J. X. J. Yuan, “Idiopathic pulmonary arterial hypertension,” DMM Disease Models and Mechanisms, vol. 3, no. 5-6, pp. 268–273, 2010.
[2]  D. B. Badesch, H. C. Champion, M. A. Gomez Sanchez et al., “Diagnosis and assessment of pulmonary arterial hypertension,” Journal of the American College of Cardiology, vol. 54, no. 1, pp. S55–S66, 2009.
[3]  M. Mandegar, Y. C. B. Fung, W. Huang, C. V. Remillard, L. J. Rubin, and J. X. J. Yuan, “Cellular and molecular mechanisms of pulmonary vascular remodeling: role in the development of pulmonary hypertension,” Microvascular Research, vol. 68, no. 2, pp. 75–103, 2004.
[4]  I. M. Robbins, “Epidemiolgoy and classification of pulmonary hypertension,” Advances in Pulmonary Hypertension, vol. 8, pp. 77–78, 2009.
[5]  G. Simonneau, I. M. Robbins, M. Beghetti et al., “Updated clinical classification of pulmonary hypertension,” Journal of the American College of Cardiology, vol. 54, no. 1, pp. S43–S54, 2009.
[6]  L. J. Rubin, “Diagnosis and management of pulmonary arterial hypertension: ACCP evidence-based clinical practice guidelines,” Chest, vol. 126, no. 1, pp. 4S–6S, 2004.
[7]  P. Escribano-Subias, I. Blanco, M. Lopez-Meseguer, et al., “Survival in pulmonary hypertension in spain insights from the spanish registry,” European Respiratory Journal, vol. 40, no. 3, pp. 596–603, 2012.
[8]  D. E. Bedford, W. Evans, and D. S. Short, “Solitary pulmonary hypertension,” British Heart Journal, vol. 19, no. 1, pp. 93–116, 1957.
[9]  W. Huang, R. T. Yen, M. McLaurine, and G. Bledsoe, “Morphometry of the human pulmonary vasculature,” Journal of Applied Physiology, vol. 81, no. 5, pp. 2123–2133, 1996.
[10]  J. K. Belknap, E. C. Orton, B. Ensley, A. Tucker, and K. R. Stenmark, “Hypoxia increases bromodeoxyuridine labeling indices in bovine neonatal pulmonary arteries,” American Journal of Respiratory Cell and Molecular Biology, vol. 16, no. 4, pp. 366–371, 1997.
[11]  N. W. Morrell, S. Adnot, S. L. Archer et al., “Cellular and molecular basis of pulmonary arterial hypertension,” Journal of the American College of Cardiology, vol. 54, no. 1, pp. S20–S31, 2009.
[12]  J. X. J. Yuan and L. J. Rubin, “Pathogenesis of pulmonary arterial hypertension: the need for multiple hits,” Circulation, vol. 111, no. 5, pp. 534–538, 2005.
[13]  L. Reid, “Vascular remodeling,” in The Pulmonary Circulation: Normal and Abnormal Mechansimc, Management, and the National Registry, A. Fishman, Ed., p. 264, University of Pennsylvania Press, Philadelphia, Pa, USA, 1990.
[14]  S. Zhang, I. Fantozzi, D. D. Tigno et al., “Bone morphogenetic proteins induce apoptosis in human pulmonary vascular smooth muscle cells,” American Journal of Physiology, vol. 285, no. 3, pp. L740–L754, 2003.
[15]  B. Meyrick and L. Reid, “Hypoxia and incorporation of 3H-thymidine by cells of the rat pulmonary arteries and alveolar wall,” American Journal of Pathology, vol. 96, no. 1, pp. 51–70, 1979.
[16]  J. Chamley-Campbell, G. R. Campbell, and R. Ross, “The smooth muscle cell in culture,” Physiological Reviews, vol. 59, no. 1, pp. 1–61, 1979.
[17]  G. K. Owens, M. S. Kumar, and B. R. Wamhoff, “Molecular regulation of vascular smooth muscle cell differentiation in development and disease,” Physiological Reviews, vol. 84, no. 3, pp. 767–801, 2004.
[18]  K. Sobue, K. Hayashi, and W. Nishida, “Expressional regulation of smooth muscle cell-specific genes in association with phenotypic modulation,” Molecular and Cellular Biochemistry, vol. 190, no. 1-2, pp. 105–118, 1999.
[19]  M. G. Frid, E. P. Moiseeva, and K. R. Stenmark, “Multiple phenotypically distinct smooth muscle cell populations exist in the adult and developing bovine pulmonary arterial media in vivo,” Circulation Research, vol. 75, no. 4, pp. 669–681, 1994.
[20]  E. D. Michelakis, V. Hampl, A. Nsair et al., “Diversity in mitochondrial function explains differences in vascular oxygen sensing,” Circulation Research, vol. 90, no. 12, pp. 1307–1315, 2002.
[21]  L. J. Rubin, “Primary pulmonary hypertension,” New England Journal of Medicine, vol. 336, no. 2, pp. 111–117, 1997.
[22]  K. R. Stenmark, N. Davie, M. Frid, E. Gerasimovskaya, and M. Das, “Role of the adventitia in pulmonary vascular remodeling,” Physiology, vol. 21, no. 2, pp. 134–145, 2006.
[23]  G. E. Hardingham, S. Chawla, C. M. Johnson, and H. Bading, “Distinct functions of nuclear and cytoplasmic calcium in the control of gene expression,” Nature, vol. 385, no. 6613, pp. 260–265, 1997.
[24]  A. P. Somlyo and A. V. Somlyo, “Signal transduction and regulation in smooth muscle,” Nature, vol. 372, no. 6503, pp. 231–236, 1994.
[25]  S. S. Mcdaniel, O. Platoshyn, J. Wang et al., “Capacitative entry in agonist-induced pulmonary vasoconstriction,” American Journal of Physiology, vol. 280, no. 5, pp. L870–L880, 2001.
[26]  A. R. Means, “Calcium, calmodulin and cell cycle regulation,” FEBS Letters, vol. 347, no. 1, pp. 1–4, 1994.
[27]  D. D. Ginty, “Calcium regulation of gene expression: isn't that spatial?” Neuron, vol. 18, no. 2, pp. 183–186, 1997.
[28]  M. J. Berridge, “Calcium signalling and cell proliferation,” BioEssays, vol. 17, no. 6, pp. 491–500, 1995.
[29]  I. A. Graef, F. Chen, L. Chen, A. Kuo, and G. R. Crabtree, “Signals transduced by /calcineurin and NFATc3/c4 pattern the developing vasculature,” Cell, vol. 105, no. 7, pp. 863–875, 2001.
[30]  M. Sheng, G. McFadden, and M. E. Greenberg, “Membrane depolarization and calcium induce c-fos transcription via phosphorylation of transcription factor CREB,” Neuron, vol. 4, no. 4, pp. 571–582, 1990.
[31]  O. Platoshyn, V. A. Golovina, C. L. Bailey et al., “Sustained membrane depolarization and pulmonary artery smooth muscle cell proliferation,” American Journal of Physiology, vol. 279, no. 5, pp. C1540–C1549, 2000.
[32]  A. D. Short, J. Bian, T. K. Ghosh, R. T. Waldron, S. L. Rybak, and D. L. Gill, “Intracellular pool content is linked to control of cell growth,” Proceedings of the National Academy of Sciences of the United States of America, vol. 90, no. 11, pp. 4986–4990, 1993.
[33]  N. Takuwa, W. Zhou, and Y. Takuwa, “Calcium, calmodulin and cell cycle progression,” Cellular Signalling, vol. 7, no. 2, pp. 93–104, 1995.
[34]  C. V. Remillard and J. X. J. Yuan, “TRP channels, CCE, and the pulmonary vascular smooth muscle,” Microcirculation, vol. 13, no. 8, pp. 671–692, 2006.
[35]  A. Makino, A. L. Firth, and J. X. J. Yuan, “Endothelial and smooth muscle cell ion channels in pulmonary vasoconstriction and vascular remodeling,” in Comprehensive Physiology, John Wiley & Sons, 2011.
[36]  M. D. Cahalan, “STIMulating store-operated entry,” Nature Cell Biology, vol. 11, no. 6, pp. 669–677, 2009.
[37]  T. T. Chen, K. D. Luykenaar, E. J. Walsh, M. P. Walsh, and W. C. Cole, “Key role of channels in vasoregulation,” Circulation Research, vol. 99, no. 1, pp. 53–60, 2006.
[38]  F. Plane, R. Johnson, P. Kerr et al., “Heteromultimeric channels contribute to myogenic control of arterial diameter,” Circulation Research, vol. 96, no. 2, pp. 216–224, 2005.
[39]  M. D. Cahalan, “How to STIMulate calcium channels,” Science, vol. 330, no. 6000, pp. 43–44, 2010.
[40]  C. Y. Park, A. Shcheglovitov, and R. Dolmetsch, “The CRAC channel activator STIM1 binds and inhibits L-type voltage-gated calcium channels,” Science, vol. 330, no. 6000, pp. 101–105, 2010.
[41]  Y. Wang, X. Deng, S. Mancarella et al., “The calcium store sensor, STIM1, reciprocally controls Orai and Ca V1.2 channels,” Science, vol. 330, no. 6000, pp. 105–109, 2010.
[42]  J. W. Putney Jr., “A model for receptor-regulated calcium entry,” Cell Calcium, vol. 7, no. 1, pp. 1–12, 1986.
[43]  J. Roos, P. J. DiGregorio, A. V. Yeromin et al., “STIM1, an essential and conserved component of store-operated Ca 2+ channel function,” Journal of Cell Biology, vol. 169, no. 3, pp. 435–445, 2005.
[44]  C. Wang, J. F. Li, L. Zhao et al., “Inhibition of SOC/ /NFAT pathway is involved in the anti-proliferative effect of sildenafil on pulmonary artery smooth muscle cells,” Respiratory Research, vol. 10, article 123, 2009.
[45]  J. Wang, L. Weigand, W. Lu, J. T. Sylvester, G. L. Semenza, and L. A. Shimoda, “Hypoxia inducible factor 1 mediates hypoxia-induced TRPC expression and elevated intracellular in pulmonary arterial smooth muscle cells,” Circulation Research, vol. 98, no. 12, pp. 1528–1537, 2006.
[46]  R. T. Williams, S. S. M. Manji, N. J. Parker et al., “Identification and characterization of the STIM (stromal interaction molecule) gene family: coding for a novel class of transmembrane proteins,” Biochemical Journal, vol. 357, no. 3, pp. 673–685, 2001.
[47]  L. Zheng, P. B. Stathopulos, R. Schindl, G. Y. Li, C. Romanin, and M. Ikura, “Auto-inhibitory role of the EF-SAM domain of STIM proteins in store-operated calcium entry,” Proceedings of the National Academy of Sciences of the United States of America, vol. 108, no. 4, pp. 1337–1342, 2011.
[48]  S. L. Zhang, Y. Yu, J. Roos et al., “STIM1 is a sensor that activates CRAC channels and migrates from the store to the plasma membrane,” Nature, vol. 437, no. 7060, pp. 902–905, 2005.
[49]  X. Deng, Y. Wang, Y. Zhou, J. Soboloff, and D. L. Gill, “STIM and orai: dynamic intermembrane coupling to control cellular calcium signals,” Journal of Biological Chemistry, vol. 284, no. 34, pp. 22501–22505, 2009.
[50]  J. Liou, M. L. Kim, D. H. Won et al., “STIM is a sensor essential for -store-depletion-triggered influx,” Current Biology, vol. 15, no. 13, pp. 1235–1241, 2005.
[51]  R. M. Luik, B. Wang, M. Prakriya, M. M. Wu, and R. S. Lewis, “Oligomerization of STIM1 couples ER calcium depletion to CRAC channel activation,” Nature, vol. 454, no. 7203, pp. 538–542, 2008.
[52]  J. Y. Kim and S. Muallem, “Unlocking SOAR releases STIM,” EMBO Journal, vol. 30, no. 9, pp. 1673–1675, 2011.
[53]  O. Brandman, J. Liou, W. S. Park, and T. Meyer, “STIM2 is a feedback regulator that stabilizes basal cytosolic and endoplasmic reticulum levels,” Cell, vol. 131, no. 7, pp. 1327–1339, 2007.
[54]  L. Zheng, P. B. Stathopulos, G. Y. Li, and M. Ikura, “Biophysical characterization of the EF-hand and SAM domain containing sensory region of STIM1 and STIM2,” Biochemical and Biophysical Research Communications, vol. 369, no. 1, pp. 240–246, 2008.
[55]  S. Parvez, A. Beck, C. Peinelt et al., “STIM2 protein mediates distinct store-dependent and store-independent modes of CRAC channel activation,” FASEB Journal, vol. 22, no. 3, pp. 752–761, 2008.
[56]  M. Y. Song, A. Makino, and J. X. Yuan, “STIM2 contributes to enhanced store-operated Ca entry in pulmonary artery smooth muscle cells from patients with idiopathic pulmonary arterial hypertension,” Pulmonary Circulation, vol. 1, pp. 84–94, 2011.
[57]  V. A. Golovina, O. Platoshyn, C. L. Bailey et al., “Upregulated TRP and enhanced capacitative entry in human pulmonary artery myocytes during proliferation,” American Journal of Physiology, vol. 280, no. 2, pp. H746–H755, 2001.
[58]  H. L. Sweeney, Z. Yang, G. Zhi, J. T. Stull, and K. M. Trybus, “Charge replacement near the phosphorylatable serine of the myosin regulatory light chain mimics aspects of phosphorylation,” Proceedings of the National Academy of Sciences of the United States of America, vol. 91, no. 4, pp. 1490–1494, 1994.
[59]  A. Penna, A. Demuro, A. V. Yeromin et al., “The CRAC channel consists of a tetramer formed by Stim-induced dimerization of Orai dimers,” Nature, vol. 456, no. 7218, pp. 116–120, 2008.
[60]  Y. Wang, X. Deng, T. Hewavitharana, J. Soboloff, and D. L. Gill, “STIM, ORAI and TRPC channels in the control of calcium entry signals in smooth muscle,” Clinical and Experimental Pharmacology and Physiology, vol. 35, no. 9, pp. 1127–1133, 2008.
[61]  D. L. Cioffi, C. Barry, and T. Stevens, “Store-operated calcium entry channels in pulmonary endotheliumml: the emerging story of TRPCS and Orai1,” Advances in Experimental Medicine and Biology, vol. 661, pp. 137–154, 2010.
[62]  M. Prakriya, S. Feske, Y. Gwack, S. Srikanth, A. Rao, and P. G. Hogan, “Orai1 is an essential pore subunit of the CRAC channel,” Nature, vol. 443, no. 7108, pp. 230–233, 2006.
[63]  P. G. Hogan, R. S. Lewis, and A. Rao, “Molecular basis of calcium signaling in lymphocytes: STIM and ORAI,” Annual Review of Immunology, vol. 28, pp. 491–533, 2010.
[64]  R. Berra-Romani, A. Mazzocco-Spezzia, M. V. Pulina, and V. A. Golovina, “ handling is altered when arterial myocytes progress from a contractile to a proliferative phenotype in culture,” American Journal of Physiology, vol. 295, no. 3, pp. C779–C790, 2008.
[65]  S. G. Baryshnikov, M. V. Pulina, A. Zulian, C. I. Linde, and V. A. Golovina, “Orai1, a critical component of store-operated entry, is functionally associated with Na+/ exchanger and plasma membrane pump in proliferating human arterial myocytes,” American Journal of Physiology, vol. 297, no. 5, pp. C1103–C1112, 2009.
[66]  J. M. Bisaillon, R. K. Motiani, J. C. Gonzalez-Cobos et al., “Essential role for STIM1/Orai1-mediated calcium influx in PDGF-induced smooth muscle migration,” American Journal of Physiology, vol. 298, no. 5, pp. C993–C1005, 2010.
[67]  W. Zhang, K. E. Halligan, X. Zhang et al., “Orai1-mediated ICRAC Is essential for neointima formation after vascular injury,” Circulation Research, vol. 109, pp. 534–542, 2011.
[68]  J. M. Edwards, Z. P. Neeb, M. A. Alloosh et al., “Exercise training decreases store-operated entry associated with metabolic syndrome and coronary atherosclerosis,” Cardiovascular Research, vol. 85, no. 3, pp. 631–640, 2010.
[69]  A. Ogawa, A. L. Firth, K. A. Smith, M. V. Maliakal, and J. X. Yuan, “PDGF enhances store-operated entry by upregulating STIM1/Orai1 via activation of Akt/mTOR in human pulmonary arterial smooth muscle cells,” American Journal of Physiology Cell Physiology, vol. 302, pp. C405–C411, 2012.
[70]  T. Y. Chuang, L. C. Au, L. C. Wang, L. T. Ho, D. M. Yang, and C. C. Juan, “Potential effect of resistin on the ET-1-increased reactions of blood pressure in rats and signaling in vascular smooth muscle cells,” Journal of Cell Physiology, vol. 227, pp. 1610–1618, 2012.
[71]  A. Lis, C. Peinelt, A. Beck et al., “CRACM1, CRACM2, and CRACM3 are store-operated channels with distinct functional properties,” Current Biology, vol. 17, no. 9, pp. 794–800, 2007.
[72]  A. V. Yeromin, S. L. Zhang, W. Jiang, Y. Yu, O. Safrina, and M. D. Cahalan, “Molecular identification of the CRAC channel by altered ion selectivity in a mutant of Orai,” Nature, vol. 443, no. 7108, pp. 226–229, 2006.
[73]  W. I. DeHaven, J. T. Smyth, R. R. Boyles, and J. W. Putney, “Calcium inhibition and calcium potentiation of Orai1, Orai2, and Orai3 calcium release-activated calcium channels,” Journal of Biological Chemistry, vol. 282, no. 24, pp. 17548–17556, 2007.
[74]  P. C. Sundivakkam, M. Freichel, V. Singh, et al., “The sensor stromal interaction molecule 1 (STIM1) is necessary and sufficient for the store-operated entry function of transient receptor potential canonical (TRPC) 1 and 4 channels in endothelial cells,” Molecular Pharmacology, vol. 81, pp. 510–526, 2012.
[75]  G. M. Salido, I. Jardín, and J. A. Rosado, “The TRPC ion channels: association with Orai1 and STIM1 proteins and participation in capacitative and non-capacitative calcium entry,” Advances in Experimental Medicine and Biology, vol. 704, pp. 413–433, 2011.
[76]  J. W. Putney Jr., M. Trebak, G. Vazquez, B. Wedel, and G. S. J. Bird, “Signalling mechanisms for TRPC3 channels,” Novartis Foundation Symposium, vol. 258, pp. 123–139, 2004.
[77]  G. Vazquez, B. J. Wedel, M. Trebak, G. S. J. Bird, and J. W. Putney Jr., “Expression level of the canonical transient receptor potential 3 (TRPC3) channel determines its mechanism of activation,” Journal of Biological Chemistry, vol. 278, no. 24, pp. 21649–21654, 2003.
[78]  A. P. Albert and W. A. Large, “Store-operated -permeable non-selective cation channels in smooth muscle cells,” Cell Calcium, vol. 33, no. 5-6, pp. 345–356, 2003.
[79]  P. C. Sundivakkam, A. M. Kwiatek, T. T. Sharma, R. D. Minshall, A. B. Malik, and C. Tiruppathi, “Caveolin-1 scaffold domain interacts with TRPC1 and IP3R3 to regulate store release-induced entry in endothelial cells,” American Journal of Physiology, vol. 296, no. 3, pp. C403–C413, 2009.
[80]  W. Lu, P. Ran, D. Zhang et al., “Sildenafil inhibits chronically hypoxic upregulation of canonical transient receptor potential expression in rat pulmonary arterial smooth muscle,” American Journal of Physiology, vol. 298, no. 1, pp. C114–C123, 2010.
[81]  G. Peng, W. Lu, X. Li et al., “Expression of store-operated entry and transient receptor potential canonical and vanilloid-related proteins in rat distal pulmonary venous smooth muscle,” American Journal of Physiology, vol. 299, no. 5, pp. L621–L630, 2010.
[82]  J. Wang, L. A. Shimoda, and J. T. Sylvester, “Capacitative calcium entry and TRPC channel proteins are expressed in rat distal pulmonary arterial smooth muscle,” American Journal of Physiology, vol. 286, no. 4, pp. L848–L858, 2004.
[83]  M. J. Lin, G. P. H. Leung, W. M. Zhang et al., “Chronic hypoxia-induced upregulation of store-operated and receptor-operated channels in pulmonary arterial smooth muscle cells: a novel mechanism of hypoxic pulmonary hypertension,” Circulation Research, vol. 95, no. 5, pp. 496–505, 2004.
[84]  B. Kumar, K. Dreja, S. S. Shah et al., “Upregulated TRPC1 channel in vascular injury in vivo and its role in human neointimal hyperplasia,” Circulation Research, vol. 98, no. 4, pp. 557–563, 2006.
[85]  X. R. Liu, M. F. Zhang, N. Yang, et al., “Enhanced store-operated entry and TRPC channel expression in pulmonary arteries of monocrotaline-induced pulmonary hypertensive rats,” American Journal of Physiology, vol. 302, pp. C77–C87, 2012.
[86]  J. P. T. Ward, G. A. Knock, V. A. Snetkov, and P. I. Aaronson, “Protein kinases in vascular smooth muscle tone-role in the pulmonary vasculature and hypoxic pulmonary vasoconstriction,” Pharmacology and Therapeutics, vol. 104, no. 3, pp. 207–231, 2004.
[87]  S. N. Saleh, A. P. Albert, and W. A. Large, “Activation of native TRPC1/C5/C6 channels by endothelin-1 is mediated by both PIP3 and PIP2 in rabbit coronary artery myocytes,” Journal of Physiology, vol. 587, no. 22, pp. 5361–5375, 2009.
[88]  J. W. Landsberg and J. X. J. Yuan, “Calcium and TRP channels in pulmonary vascular smooth muscle cell proliferation,” News in Physiological Sciences, vol. 19, no. 2, pp. 44–50, 2004.
[89]  A. B. Parekh and J. W. Putney, “Store-operated calcium channels,” Physiological Reviews, vol. 85, no. 2, pp. 757–810, 2005.
[90]  J. P. Yuan, W. Zeng, G. N. Huang, P. F. Worley, and S. Muallem, “STIM1 heteromultimerizes TRPC channels to determine their function as store-operated channels,” Nature Cell Biology, vol. 9, no. 6, pp. 636–645, 2007.
[91]  J. Li, P. Sukumar, C. J. Milligan et al., “Interactions, functions, and independence of plasma membrane STIM1 and TRPC1 in vascular smooth muscle cells,” Circulation Research, vol. 103, no. 8, pp. e97–e104, 2008.
[92]  W. Zeng, J. P. Yuan, M. S. Kim et al., “STIM1 gates TRPC channels, but not Orai1, by electrostatic interaction,” Molecular Cell, vol. 32, no. 3, pp. 439–448, 2008.
[93]  M. S. Kim, W. Zeng, J. P. Yuan, D. M. Shin, P. F. Worley, and S. Muallem, “Native store-operated influx requires the channel function of Orai 1 and TRPC1,” Journal of Biological Chemistry, vol. 284, no. 15, pp. 9733–9741, 2009.
[94]  Y. Liao, C. Erxleben, J. Abramowitz et al., “Functional interactions among Orai1, TRPCs, and STIM1 suggest a STIM-regulated heteromeric Orai/TRPC model for SOCE/Icrac channels,” Proceedings of the National Academy of Sciences of the United States of America, vol. 105, no. 8, pp. 2895–2900, 2008.
[95]  Y. Liao, C. Erxleben, E. Yildirim, J. Abramowitz, D. L. Armstrong, and L. Birnbaumer, “Orai proteins interact with TRPC channels and confer responsiveness to store depletion,” Proceedings of the National Academy of Sciences of the United States of America, vol. 104, no. 11, pp. 4682–4687, 2007.
[96]  T. C. Kwong, X. Liu, L. O. Hwei, and I. S. Ambudkar, “Functional requirement for Orai1 in store-operated TRPC1-STIM1 channels,” Journal of Biological Chemistry, vol. 283, no. 19, pp. 12935–12940, 2008.
[97]  D. L. Cioffi, S. Wu, and H. Chen, “Orai1 determines calcium selectivity of an endogenous TRPC heterotetramer channel,” Circulation Research, vol. 110, pp. 1435–1444, 2012.
[98]  N. Kunichika, J. W. Landsberg, Y. Yu et al., “Bosentan inhibits transient receptor potential channel expression in pulmonary vascular myocytes,” American Journal of Respiratory and Critical Care Medicine, vol. 170, no. 10, pp. 1101–1107, 2004.
[99]  N. Kunichika, Y. Yu, C. V. Remillard, O. Platoshyn, S. Zhang, and J. X. J. Yuan, “Overexpression of TRPC1 enhances pulmonary vasoconstriction induced by capacitative entry,” American Journal of Physiology, vol. 287, no. 5, pp. L962–L969, 2004.
[100]  Y. Yu, I. Fantozzi, C. V. Remillard et al., “Enhanced expression of transient receptor potential channels in idiopathic pulmonary arterial hypertension,” Proceedings of the National Academy of Sciences of the United States of America, vol. 101, no. 38, pp. 13861–13866, 2004.
[101]  A. Dietrich, Y. S. M. Mederos, M. Gollasch et al., “Increased vascular smooth muscle contractility in TRPC6 -/- mice,” Molecular and Cellular Biology, vol. 25, pp. 6980–6989, 2005.
[102]  Y. Yu, S. H. Keller, C. V. Remillard et al., “A functional single-nucleotide polymorphism in the TRPC6 gene promoter associated with idiopathic pulmonary arterial hypertension,” Circulation, vol. 119, no. 17, pp. 2313–2322, 2009.
[103]  M. Trebak, G. Vazquez, G. S. J. Bird, and J. W. Putney, “The TRPC3/6/7 subfamily of cation channels,” Cell Calcium, vol. 33, no. 5-6, pp. 451–461, 2003.
[104]  N. Weissmann, A. Dietrich, B. Fuchs et al., “Classical transient receptor potential channel 6 (TRPC6) is essential for hypoxic pulmonary vasoconstriction and alveolar gas exchange,” Proceedings of the National Academy of Sciences of the United States of America, vol. 103, no. 50, pp. 19093–19098, 2006.
[105]  A. Dietrich, H. Kalwa, B. Fuchs, F. Grimminger, N. Weissmann, and T. Gudermann, “In vivo TRPC functions in the cardiopulmonary vasculature,” Cell Calcium, vol. 42, no. 2, pp. 233–244, 2007.
[106]  R. Inoue, T. Okada, H. Onoue et al., “The transient receptor potential protein homologue TRP6 is the essential component of vascular alpha;1-adrenoceptor-activated -permeable cation channel,” Circulation Research, vol. 88, no. 3, pp. 325–332, 2001.
[107]  S. Jung, R. Strotmann, G. Schultz, and T. D. Plant, “TRCP6 is a candidate channel involved in receptor stimulated cation currents in A7r5 smooth muscle cells,” American Journal of Physiology, vol. 282, no. 2, pp. C347–C359, 2002.
[108]  B. Fuchs, M. Rupp, H. A. Ghofrani et al., “Diacylglycerol regulates acute hypoxic pulmonary vasoconstriction via TRPC6,” Respiratory Research, vol. 12, article 20, 2011.
[109]  H. Sawada, Y. Mitani, J. Maruyama et al., “A nuclear factor-κB inhibitor pyrrolidine dithiocarbamate ameliorates pulmonary hypertension in rats,” Chest, vol. 132, no. 4, pp. 1265–1274, 2007.
[110]  Y. Jia and L. Y. Lee, “Role of TRPV receptors in respiratory diseases,” Biochimica et Biophysica Acta, vol. 1772, pp. 915–927, 2007.
[111]  Y. X. Wang, J. Wang, C. Wang et al., “Functional expression of transient receptor potential vanilloid-related channels in chronically hypoxic human pulmonary arterial smooth muscle cells,” Journal of Membrane Biology, vol. 223, no. 3, pp. 151–159, 2008.
[112]  B. Minke and M. Parnas, “Insights on TRP channels from in vivo studies in Drosophila,” Annual Review of Physiology, vol. 68, pp. 649–684, 2006.
[113]  X. R. Yang, M. J. Lin, L. S. McIntosh, and J. S. K. Sham, “Functional expression of transient receptor potential melastatin- and vanilloid-related channels in pulmonary arterial and aortic smooth muscle,” American Journal of Physiology, vol. 290, no. 6, pp. L1267–L1276, 2006.
[114]  S. Earley, T. J. Heppner, M. T. Nelson, and J. E. Brayden, “TRPV4 forms a novel signaling complex with ryanodine receptors and BKCa channels,” Circulation Research, vol. 97, no. 12, pp. 1270–1279, 2005.
[115]  M. T. Nelson, H. Cheng, M. Rubart et al., “Relaxation of arterial smooth muscle by calcium sparks,” Science, vol. 270, no. 5236, pp. 633–637, 1995.
[116]  T. Ducret, C. Guibert, R. Marthan, and J. P. Savineau, “Serotonin-induced activation of TRPV4-like current in rat intrapulmonary arterial smooth muscle cells,” Cell Calcium, vol. 43, no. 4, pp. 315–323, 2008.
[117]  T. Kark, Z. Bagi, E. Lizanecz et al., “Tissue-specific regulation of microvascular diameter: opposite functional roles of neuronal and smooth muscle located vanilloid receptor-1,” Molecular Pharmacology, vol. 73, no. 5, pp. 1405–1412, 2008.
[118]  K. Muraki, Y. Iwata, Y. Katanosaka et al., “TRPV2 is a component of osmotically sensitive cation channels in murine aortic myocytes,” Circulation Research, vol. 93, no. 9, pp. 829–838, 2003.
[119]  H. Y. Yoo, S. J. Park, E. Y. Seo, et al., “Role of thromboxane A(2)-activated nonselective cation channels in hypoxic pulmonary vasoconstriction of rat,” American Journal of Physiology, vol. 302, pp. C307–C317, 2012.
[120]  R. Inoue, L. J. Jensen, J. Shi et al., “Transient receptor potential channels in cardiovascular function and disease,” Circulation Research, vol. 99, no. 2, pp. 119–131, 2006.
[121]  B. Nilius, J. Prenen, G. Droogmans et al., “Voltage dependence of the -activated cation channel TRPM4,” Journal of Biological Chemistry, vol. 278, no. 33, pp. 30813–30820, 2003.
[122]  B. Nilius, J. Prenen, A. Janssens et al., “The selectivity filter of the cation channel TRPM4,” Journal of Biological Chemistry, vol. 280, no. 24, pp. 22899–22906, 2005.
[123]  A. L. Gonzales, G. C. Amberg, and S. Earley, “ release from the sarcoplasmic reticulum is required for sustained TRPM4 activity in cerebral artery smooth muscle cells,” American Journal of Physiology, vol. 299, no. 2, pp. C279–C288, 2010.
[124]  S. A. Reading and J. E. Brayden, “Central role of TRPM4 channels in cerebral blood flow regulation,” Stroke, vol. 38, no. 8, pp. 2322–2328, 2007.
[125]  Y. He, G. Yao, C. Savoia, and R. M. Touyz, “Transient receptor potential melastatin 7 ion channels regulate magnesium homeostasis in vascular smooth muscle cells: role of angiotensin II,” Circulation Research, vol. 96, no. 2, pp. 207–215, 2005.
[126]  C. Nasuhoglu, S. Feng, Y. Mao et al., “Modulation of cardiac PIP2 by cardioactive hormones and other physiologically relevant interventions,” American Journal of Physiology, vol. 283, no. 1, pp. C223–C234, 2002.
[127]  C. D. Johnson, D. Melanaphy, A. Purse, S. A. Stokesberry, P. Dickson, and A. V. Zholos, “Transient receptor potential melastatin 8 channel involvement in the regulation of vascular tone,” American Journal of Physiology, vol. 296, no. 6, pp. H1868–H1877, 2009.
[128]  F. Mahieu, G. Owsianik, L. Verbert et al., “TRPM8-independent menthol-induced release from endoplasmic reticulum and Golgi,” Journal of Biological Chemistry, vol. 282, no. 5, pp. 3325–3336, 2007.

Full-Text

comments powered by Disqus

Contact Us

service@oalib.com

QQ:3279437679

WhatsApp +8615387084133