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...

Mitochondrial Markers for Cancer: Relevance to Diagnosis, Therapy, and Prognosis and General Understanding of Malignant Disease Mechanisms

DOI: 10.5402/2012/217162

Full-Text   Cite this paper   Add to My Lib

Abstract:

Cancer cells display changes that aid them to escape from cell death, sustain their proliferative powers, and shift their metabolism toward glycolytic energy production. Mitochondria are key organelles in many metabolic and biosynthetic pathways, and the adaptation of mitochondrial function has been recognized as crucial to the changes that occur in cancer cells. This paper zooms in on the pathologic evaluation of mitochondrial markers for diagnosing and staging of human cancer and determining the patients’ prognoses. 1. Introduction Mitochondria are membrane-enclosed cell organelles that can be found in all human cells, except for the peripheral red blood cells. An eukaryote cell contains around 1000 to 2000 mitochondria, with diameters varying between 0.5 and 1.0?μm. The mitochondrion is a key venue for cellular metabolism and the powerhouse of the cell. As a consequence, mitochondrial content is influenced by cellular energy demand. Exercise training for instance, increases the amount of mitochondria per fiber and the volume of organelles in skeletal muscle tissue [1, 2]. 1.1. Mitochondrial Metabolic Pathways To understand the crucial role played by mitochondria in cancer, it is necessary to fully grasp the extent of their metabolic activity. 1.1.1. Glycolysis Glycolysis is the sequence of cellular reactions that converts glucose into pyruvate, with the concomitant production of a relatively small amount of energy. Glycolysis occurs throughout the cell and is considered the basis of all energy processes. The 2 ATP and 2 NADH molecules it produces can enter the oxidative phosphorylation (OXPHOS) cycle, which then produces larger quantities of ATP. 1.1.2. Oxidative Phosphorylation In the Krebs or OXPHOS cycle, sequential oxidation and reduction reactions take place upon a chain of four multiprotein complexes: (1) complex I is composed of 45 protein subunits and displays NADH dehydrogenase activity. In the process, four hydrogen ions are pumped out of the mitochondrial matrix. (2) Complex II: the succinate dehydrogenase complex catalyzes the oxidation of succinate to fumarate, with concomitant reduction of ubiquinone. (3) In two cycles, complex III or coenzyme Q cytochrome c oxidoreductase reduces coenzyme Q, extracting 4 protons from the mitochondrial matrix. (4) On complex IV or cytochrome c oxidase, electrons are donated one at a time to cytochrome c and passed on to O2, producing 2 H2O molecules. In addition to the protons utilized in the reduction of O2, there is electron transfer-linked transport of 2 protons from the matrix to the mitochondrial

References

[1]  J. O. Holloszy, “Regulation by exercise of skeletal muscle content of mitochondria and GLUT4,” Journal of Physiology and Pharmacology, vol. 59, supplement 7, pp. 5–18, 2008.
[2]  A. R. Coggan, W. M. Kohrt, R. J. Spina, D. M. Bier, and J. O. Holloszy, “Endurance training decreases plasma glucose turnover and oxidation during moderate-intensity exercise in men,” Journal of Applied Physiology, vol. 68, no. 3, pp. 990–996, 1990.
[3]  R. A. Stuart, “Supercomplex organization of the oxidative phosphorylation enzymes in yeast mitochondria,” Journal of Bioenergetics and Biomembranes, vol. 40, no. 5, pp. 411–417, 2008.
[4]  MITOMAP, “A human mitochondrial genome database,” 2008, http://www.mitomap.org/MITOMAP.
[5]  M. Zeviani and S. Di Donato, “Mitochondrial disorders,” Brain, vol. 127, no. 10, pp. 2153–2172, 2004.
[6]  C. Benda, “Weitere Mitteilungen uber die Mitochondria,” Verhandlungen Physiologie Geselschaft Berlin, vol. 99, no. 4–7, pp. 376–383, 1898.
[7]  M. P. Yaffe, “The machinery of mitochondrial inheritance and behavior,” Science, vol. 283, no. 5407, pp. 1493–1497, 1999.
[8]  H. Otera, C. Wang, M. M. Cleland et al., “Mff is an essential factor for mitochondrial recruitment of Drp1 during mitochondrial fission in mammalian cells,” Journal of Cell Biology, vol. 191, no. 6, pp. 1141–1158, 2010.
[9]  D. Tondera, F. Czauderna, K. Paulick, R. Schwarzer, J. Kaufmann, and A. Santel, “The mitochondrial protein MTP18 contributes to mitochondrial fission in mammalian cells,” Journal of Cell Science, vol. 118, no. 14, pp. 3049–3059, 2005.
[10]  A. Niemann, M. Ruegg, V. La Padula, A. Schenone, and U. Suter, “Ganglioside-induced differentiation associated protein 1 is a regulator of the mitochondrial network: new implications for Charcot-Marie-Tooth disease,” Journal of Cell Biology, vol. 170, no. 7, pp. 1067–1078, 2005.
[11]  M. Zick, R. Rabl, and A. S. Reichert, “Cristae formation-linking ultrastructure and function of mitochondria,” Biochimica et Biophysica Acta, vol. 1793, no. 1, pp. 5–19, 2009.
[12]  P. Paumard, J. Vaillier, B. Coulary et al., “The ATP synthase is involved in generating mitochondrial cristae morphology,” European Molecular Biology Organization Journal, vol. 21, no. 3, pp. 221–230, 2002.
[13]  O. Warburg, “On the origin of cancer cells,” Science, vol. 123, no. 3191, pp. 309–314, 1956.
[14]  B. Altenberg and K. O. Greulich, “Genes of glycolysis are ubiquitously overexpressed in 24 cancer classes,” Genomics, vol. 84, no. 6, pp. 1014–1020, 2004.
[15]  S. Ohlmeier, A. J. Kastaniotis, J. K. Hiltunen, and U. Bergmann, “The yeast mitochondrial proteome, a study of fermentive and respiratory growth,” The Journal of Biological Chemistry, vol. 279, no. 6, pp. 3956–3979, 2004.
[16]  N. Entelis, I. Brandina, P. Kamenski, I. A. Krasheninnikov, R. P. Martin, and I. Tarassov, “A glycolytic enzyme, enolase, is recruited as a cofactor of tRNA targeting toward mitochondria in Saccharomyces cerevisiae,” Genes and Development, vol. 20, no. 12, pp. 1609–1620, 2006.
[17]  M. Capello, S. Ferri-Borgogno, P. Cappello, and F. Novelli, “α-enolase: a promising therapeutic and diagnostic tumor target,” The Federation of the Societies of Biochemistry and Molecular Biology Journal, vol. 278, no. 7, pp. 1064–1074, 2011.
[18]  S. H. Tu, C. C. Chang, C. S. Chen et al., “Increased expression of enolase α in human breast cancer confers tamoxifen resistance in human breast cancer cells,” Breast Cancer Research and Treatment, vol. 121, no. 3, pp. 539–553, 2010.
[19]  S. T. Tsai, I. H. Chien, W. H. Shen et al., “ENO1, a potential prognostic head and neck cancer marker, promotes transformation partly via chemokine CCL20 induction,” European Journal of Cancer, vol. 46, no. 9, pp. 1712–1723, 2010.
[20]  G. C. Chang, K. J. Liu, C. L. Hsieh et al., “Identification of α-enolase as an autoantigen in lung cancer: its overexpression is associated with clinical outcomes,” Clinical Cancer Research, vol. 12, no. 19, pp. 5746–5754, 2006.
[21]  S. Mazurek, C. B. Boschek, F. Hugo, and E. Eigenbrodt, “Pyruvate kinase type M2 and its role in tumor growth and spreading,” Seminars in Cancer Biology, vol. 15, no. 4, pp. 300–308, 2005.
[22]  Y. Kumar, N. Tapuria, N. Kirmani, and B. R. Davidson, “Tumour M2-pyruvate kinase: a gastrointestinal cancer marker,” European Journal of Gastroenterology and Hepatology, vol. 19, no. 3, pp. 265–276, 2007.
[23]  D. Lüftner, J. Mesterharm, C. Akrivakis et al., “Tumor type M2 pyruvate kinase expression in advanced breast cancer,” Anticancer Research, vol. 20, no. 6, pp. 5077–5082, 2000.
[24]  B. Chiavarina, D. Whitacker-Menezes, U. E. Martinez-Outschoom, et al., “Pyruvate kinase expression (PKM1 and PKM2) in cancer-associated fibroblasts drives stromal nutrient production and tumor growth,” Cancer Biology and Therapy, vol. 12, no. 12, pp. 1101–1113, 2012.
[25]  R. A. Harris, M. M. Bowker-Kinley, B. Huang, and P. Wu, “Regulation of the activity of the pyruvate dehydrogenase complex,” Advances in Enzyme Regulation, vol. 42, no. 1, pp. 249–259, 2002.
[26]  M. L. Eboli, “Pyruvate dehydrogenase levels in Morris hepatomas with different growth rate,” Cancer Letters, vol. 26, no. 2, pp. 185–190, 1985.
[27]  M. L. Eboli and A. Pasquini, “Transformation linked decrease of pyruvate dehydrogenase complex in human epidermis,” Cancer Letters, vol. 85, no. 2, pp. 239–243, 1994.
[28]  M. I. Koukourakis, A. Giatromanolaki, E. Sivridis, K. C. Gatter, and A. L. Harris, “Pyruvate dehydrogenase and pyruvate dehydrogenase kinase expression in non small cell lung cancer and tumor-associated stroma,” Neoplasia, vol. 7, no. 1, pp. 1–6, 2005.
[29]  K. M. Owens, M. Kulawiec, M. M. Desouki, A. Vanniarajan, and K. K. Singh, “Impaired OXPHOS complex III in breast cancer,” PlosOne, vol. 6, no. 8, Article ID e23846, 2011.
[30]  H. Simonnet, N. Alazard, K. Pfeiffer et al., “Low mitochondrial respiratory chain content correlates with tumor aggressiveness in renal cell carcinoma,” Carcinogenesis, vol. 23, no. 5, pp. 759–768, 2002.
[31]  H. Simonnet, J. Demont, K. Pfeiffer et al., “Mitochondrial complex I is deficient in renal oncocytomas,” Carcinogenesis, vol. 24, no. 9, pp. 1461–1466, 2003.
[32]  J. A. Mayr, D. Meierhofer, F. Zimmermann et al., “Loss of complex I due to mitochondrial DNA mutations in renal oncocytoma,” Clinical Cancer Research, vol. 14, no. 8, pp. 2270–2275, 2008.
[33]  F. A. Zimmermann, J. A. Mayr, D. Neureiter et al., “Lack of complex i is associated with oncocytic thyroid tumours,” British Journal of Cancer, vol. 100, no. 9, pp. 1434–1437, 2009.
[34]  D. Meierhofer, J. A. Mayr, U. Foetschl et al., “Decrease of mitochondrial DNA content and energy metabolism in renal cell carcinoma,” Carcinogenesis, vol. 25, no. 6, pp. 1005–1010, 2004.
[35]  Y. Ohashi, S. J. Kaneko, T. E. Cupples, and S. R. Young, “Ubiquinol cytochrome c reductase (UQCRFS1) gene amplification in primary breast cancer core biopsy samples,” Gynecologic Oncology, vol. 93, no. 1, pp. 54–58, 2004.
[36]  K. Luciakova and S. Kuzela, “Increased steady-state levels of several mitochondrial and nuclear gene transcripts in rat hepatoma with a low content of mitochondria,” European Journal of Biochemistry, vol. 205, no. 3, pp. 1187–1193, 1992.
[37]  K. D. Yu, A. X. Chen, C. Yang et al., “Current evidence on the relationship between polymorphisms in the COX-2 gene and breast cancer risk: a meta-analysis,” Breast Cancer Research and Treatment, vol. 122, no. 1, pp. 251–257, 2010.
[38]  B. G. Heerdt and L. H. Augenlicht, “Effects of fatty acids on expression of genes encoding subunits of cytochrome c oxidase and cytochrome c oxidase activity in HT29 human colonic adenocarcinoma cells,” The Journal of Biological Chemistry, vol. 266, no. 28, pp. 19120–19126, 1991.
[39]  M. Denis-Gay, J. M. Petit, J. P. Mazat, and M. H. Ratinaud, “Modifications of oxido-reductase activities in adriamycin-resistant leukaemia K562 cells,” Biochemical Pharmacology, vol. 56, no. 4, pp. 451–457, 1998.
[40]  N. A. Alam, A. J. Rowan, N. C. Wortham et al., “Genetic and functional analyses of FH mutations in multiple cutaneous and uterine leiomyomatosis, hereditary leiomyomatosis and renal cancer, and fumarate hydratase deficiency,” Human Molecular Genetics, vol. 12, no. 11, pp. 1241–1252, 2003.
[41]  P. J. Ratcliffe, “Fumarate hydratase deficiency and cancer: activation of hypoxia signaling?” Cancer Cell, vol. 11, no. 4, pp. 303–305, 2007.
[42]  K. K. Hyoung, S. P. Won, H. K. Sung et al., “Mitochondrial alterations in human gastric carcinoma cell line,” American Journal of Physiology, vol. 293, no. 2, pp. C761–C771, 2007.
[43]  L. Wang, Q. Ke, W. Chen et al., “Polymorphisms of MTHFD, plasma homocysteine levels, and risk of gastric cancer in a high-risk Chinese population,” Clinical Cancer Research, vol. 13, no. 8, pp. 2526–2532, 2007.
[44]  A. S. Andrew, J. Gui, A. C. Sanderson et al., “Bladder cancer SNP panel predicts susceptibility and survival,” Human Genetics, vol. 125, no. 5-6, pp. 527–539, 2009.
[45]  T. Geiger, S. F. Madden, W. M. Gallagher, J. Cox, and M. Mann, “Proteomic portrait of human breast cancer progression identifies novel prognostic markers,” Cancer Research, vol. 72, no. 9, pp. 2428–2439, 2012.
[46]  R. Strong, T. Nakanishi, D. Ross, and C. Fenselau, “Alterations in the mitochondrial proteome of adriamycin resistant MCF-7 breast cancer cells,” Journal of Proteome Research, vol. 5, no. 9, pp. 2389–2395, 2006.
[47]  D. L. Croteau and V. A. Bohr, “Repair of oxidative damage to nuclear and mitochondrial DNA in mammalian cells,” The Journal of Biological Chemistry, vol. 272, no. 41, pp. 25409–25412, 1997.
[48]  M. Verma, R. K. Naviaux, M. Tanaka, D. Kumar, C. Franceschi, and K. K. Singh, “Mitochondrial DNA and cancer epidemiology,” Cancer Research, vol. 67, no. 2, pp. 437–439, 2007.
[49]  G. Gasparre, A. M. Porcelli, E. Bonora et al., “Disruptive mitochondrial DNA mutations in complex I subunits are markers of oncocytic phenotype in thyroid tumors,” Proceedings of the National Academy of Sciences of the United States of America, vol. 104, no. 21, pp. 9001–9006, 2007.
[50]  K. Polyak, Y. B. Li, H. Zhu et al., “Somatic mutations of the mitochondrial genome in human colorectal tumours,” Nature Genetics, vol. 20, no. 3, pp. 291–293, 1998.
[51]  W. Habano, T. Sugai, T. Yoshida, and S. Nakamura, “Mitochondrial gene mutation, but not large-scale deletion, is a feature of colorectal carcinomas with mitochondrial microsatellite instability,” International Journal of Cancer, vol. 83, no. 5, pp. 625–629, 1999.
[52]  D. J. Tan, J. Chang, W. L. Chen et al., “Somatic mitochondrial DNA mutations in oral cancer of betel quid chewers,” Annals of the New York Academy of Sciences, vol. 1011, pp. 310–316, 2004.
[53]  S. Dasgupta, C. B. Shao, T. E. Keane, et al., “Detection of mitochondrial deoxyribonucleic acid alterations in urine from urothelial cells carcinoma patients,” International Journal of Cancer, vol. 131, no. 1, pp. 158–164, 2012.
[54]  A. I. Guney, D. S. Ergec, H. H. Tavukcu, et al., “Detection of mitochondrial DNA mutations in nonmuscle invasive bladder cancer,” Genetic Testing and Molecular Biomarkers, vol. 16, no. 7, pp. 672–678, 2012.
[55]  S. Dasgupta, M. O. Hoque, S. Upadhyay, and D. Sidransky, “Mitochondrial cytochrome B gene mutation promotes tumor growth in bladder cancer,” Cancer Research, vol. 68, no. 3, pp. 700–706, 2008.
[56]  J. A. Petros, A. K. Baumann, E. Ruiz-Pesini et al., “MtDNA mutations increase tumorigenicity in prostate cancer,” Proceedings of the National Academy of Sciences of the United States of America, vol. 102, no. 3, pp. 719–724, 2005.
[57]  J. A. Smeitink, M. Zeviani, D. M. Turnbull, and H. T. Jacobs, “Mitochondrial medicine: a metabolic perspective on the pathology of oxidative phosphorylation disorders,” Cell Metabolism, vol. 3, no. 1, pp. 9–13, 2006.
[58]  X. G. Guo, C. T. Liu, H. Dai, and Q. N. Guo, “Mutations in the mitochondrial ATase6 gene are frequent in Chinese patients with osteosarcoma,” Experimental and Molecular Pathology. In press.
[59]  F. O. Aikhionbare, S. Mehrabi, K. Kumaresan et al., “Mitochondrial DNA sequence variants in epithelial ovarian tumor subtypes and stages,” Journal of Carcinogenesis, vol. 6, article e1, 2007.
[60]  M. I. Ekstrand, M. Falkenberg, A. Rantanen et al., “Mitochondrial transcription factor A regulates mtDNA copy number in mammals,” Human Molecular Genetics, vol. 13, no. 9, pp. 935–944, 2004.
[61]  S. Horai and K. Hayasaka, “Intraspecific nucleotide sequence differences in the major noncoding region of human mitochondrial DNA,” American Journal of Human Genetics, vol. 46, no. 4, pp. 828–842, 1990.
[62]  S. A. Liu, R. S. Jiang, F. J. Chen, W. Y. Wang, and J. C. Lin, “Somatic mutations in the D-loop of mitochondrial DNA in oral squamous cell carcinoma,” European Archives of Oto-Rhino-Laryngology, vol. 269, no. 6, pp. 1665–1670, 2012.
[63]  A. Nagy, M. Wilhelm, and G. Kovacs, “Mutations of mtDNA in renal cell tumours arising in end-stage renal disease,” Journal of Pathology, vol. 199, no. 2, pp. 237–242, 2003.
[64]  L. M. Tseng, P. H. Yin, C. W. Chi et al., “Mitochondrial DNA mutations and mitochondrial DNA depletion in breast cancer,” Genes Chromosomes and Cancer, vol. 45, no. 7, pp. 629–638, 2006.
[65]  H. C. Lee, S. H. Li, J. C. Lin, C. C. Wu, D. C. Yeh, and Y. H. Wei, “Somatic mutations in the D-loop and decrease in the copy number of mitochondrial DNA in human hepatocellular carcinoma,” Mutation Research, vol. 547, no. 1-2, pp. 71–78, 2004.
[66]  M. Pinheiro, I. Veiga, C. Pinto et al., “Mitochondrial genome alterations in rectal and sigmoid carcinomas,” Cancer Letters, vol. 280, no. 1, pp. 38–43, 2009.
[67]  C. Xu, D. Tran-Thanh, C. Ma, et al., “Mitochondrial D310 mutations in the early development of breast cancer,” British Journal of Cancer, vol. 106, no. 9, pp. 1506–1511, 2012.
[68]  P. Parrella, Y. Xiao, M. Fliss et al., “Detection of mitochondrial DNA mutations in primary breast cancer and fine-needle aspirates,” Cancer Research, vol. 61, no. 20, pp. 7623–7626, 2001.
[69]  M. Y. Tang, S. Baez, M. Pruyas et al., “Mitochondrial DNA mutation at the D310 (displacement loop) mononucleotide sequence in the pathogenesis of gallbladder carcinoma,” Clinical Cancer Research, vol. 10, no. 3, pp. 1041–1046, 2004.
[70]  M. S. Fliss, H. Usadel, O. L. Caballero et al., “Facile detection of mitochondrial DNA mutations in tumors and bodily fluids,” Science, vol. 287, no. 5460, pp. 2017–2019, 2000.
[71]  A. Lièvre, C. Chapusot, A. M. Bouvier et al., “Clinical value of mitochondrial mutations in colorectal cancer,” Journal of Clinical Oncology, vol. 23, no. 15, pp. 3517–3525, 2005.
[72]  J. S. Carew and P. Huang, “Mitochondrial defects in cancer,” Molecular Cancer, vol. 1, article 9, 2002.
[73]  H. C. Lee and Y. H. Wei, “Mitochondrial DNA instability and metabolic shift in human cancers,” International Journal of Molecular Sciences, vol. 10, no. 2, pp. 674–701, 2009.
[74]  W. W. Jiang, B. Masayesva, M. Zahurak et al., “Increased mitochondrial DNA content in saliva associated with head and neck cancer,” Clinical Cancer Research, vol. 11, no. 7, pp. 2486–2491, 2005.
[75]  C. S. Lin, S. C. Chang, L. S. Wang et al., “The role of mitochondrial DNA alterations in esophageal squamous cell carcinomas,” Journal of Thoracic and Cardiovascular Surgery, vol. 139, no. 1, pp. 189–197, 2010.
[76]  M. M. Kim, J. D. Clinger, B. G. Masayesva et al., “Mitochondrial DNA quantity increases with histopathologic grade in premalignant and malignant head and neck lesions,” Clinical Cancer Research, vol. 10, no. 24, pp. 8512–8515, 2004.
[77]  E. Mambo, A. Chatterjee, M. Xing et al., “Tumor-specific changes in mtDNA content in human cancer,” International Journal of Cancer, vol. 116, no. 6, pp. 920–924, 2005.
[78]  Y. Wang, V. W. S. Liu, W. C. Xue, P. C. K. Tsang, A. N. Y. Cheung, and H. Y. S. Ngan, “The increase of mitochondrial DNA content in endometrial adenocarcinoma cells: a quantitative study using laser-captured microdissected tissues,” Gynecologic Oncology, vol. 98, no. 1, pp. 104–110, 2005.
[79]  G. Tallini, M. Ladanyi, J. Rosai, and S. C. Jhanwar, “Analysis of nuclear and mitochondrial DNA alterations in thyroid and renal oncocytic tumors,” Cytogenetics and Cell Genetics, vol. 66, no. 4, pp. 253–259, 1994.
[80]  T. Mizumachi, L. Muskhelishvili, A. Naito et al., “Increased distributional variance of mitochondrial DNA content associated with prostate cancer cells as compared with normal prostate cells,” Prostate, vol. 68, no. 4, pp. 408–417, 2008.
[81]  C. W. Wu, P. H. Yin, W. Y. Hung et al., “Mitochondria DNA mutations and mitochondrial DNA depletion in gastric cancer,” Genes Chromosomes and Cancer, vol. 44, no. 1, pp. 19–28, 2005.
[82]  C. S. Lina, L. S. Wang, C. M. Tsaig, and Y. H. Wei, “Low copy number and low oxidative damage of mitochondrial DNA are associated with tumor progression in lung cancer tissues after neoadjuvant chemotherapy,” Interactive Cardiovascular and Thoracic Surgery, vol. 7, no. 6, pp. 954–958, 2008.
[83]  N. Mehra, M. Penning, J. Maas, N. Van Daal, R. H. Giles, and E. E. Voest, “Circulating mitochondrial nucleic acids have prognostic value for survival in patients with advanced prostate cancer,” Clinical Cancer Research, vol. 13, no. 2, pp. 421–426, 2007.
[84]  C. W. Hsu, P. H. Yin, H. C. Lee, C. W. Chi, and L. M. Tseng, “Mitochondrial DNA content as a potential marker to predict response to anthracycline in breast cancer patients,” The Breast Journal, vol. 16, no. 3, pp. 264–270, 2010.
[85]  X. Zhu, X. Peng, M. X. Guan, and Q. Yan, “Pathogenic mutations of nuclear genes associated with mitochondrial disorders,” Acta Biochimica et Biophysica Sinica, vol. 41, no. 3, pp. 179–187, 2009.
[86]  J. Permuth-Wey, Y. A. Chen, Y. Y. Tsai et al., “Inherited variants in mitochondrial biogenesis genes may influence epithelial ovarian cancer risk,” Cancer Epidemiology Biomarkers and Prevention, vol. 20, no. 6, pp. 1131–1145, 2011.
[87]  L. Wang, S. K. McDonnell, S. J. Hebbering, et al., “Polymorphisms in mitochondrial genes and prostate cancer risk,” Cancer Epidemiol Biomarkers, vol. 17, no. 12, pp. 3558–3666, 2008.
[88]  N. Burnichon, J. J. Brière, R. Libé et al., “SDHA is a tumor suppressor gene causing paraganglioma,” Human Molecular Genetics, vol. 19, no. 15, pp. 3011–3020, 2010.
[89]  B. E. Baysal, J. E. Willett-Brozick, E. C. Lawrence et al., “Prevalence of SDHB, SDHC, and SDHD germline mutations in clinic patients with head and neck paragangliomas,” Journal of Medical Genetics, vol. 39, no. 3, pp. 178–183, 2002.
[90]  S. Niemann and U. Muller, “Mutations in SDHC cause autosomal dominant paraganglioma, type 3,” Nature Genetics, vol. 26, no. 3, pp. 268–270, 2000.
[91]  B. E. Baysal, R. E. Ferrell, J. E. Willett-Brozick et al., “Mutations in SDHD, a mitochondrial complex II gene, in hereditary paraganglioma,” Science, vol. 287, no. 5454, pp. 848–851, 2000.
[92]  D. Astuti, F. Latif, A. Dallol et al., “Gene mutations in the succinate dehydrogenase subunit SDHB cause susceptibility to familial pheochromocytoma and to familial paraganglioma,” American Journal of Human Genetics, vol. 69, no. 1, pp. 49–54, 2001.
[93]  C. Ricketts, E. R. Woodward, P. Killick et al., “Germline SDHB mutations and familial renal cell carcinoma,” Journal of the National Cancer Institute, vol. 100, no. 17, pp. 1260–1262, 2008.
[94]  H. X. Hao, O. Khalimonchuk, M. Schraders et al., “SDH5, a gene required for flavination of succinate dehydrogenase, is mutated in paraganglioma,” Science, vol. 325, no. 5944, pp. 1139–1142, 2009.
[95]  S. N. J. Sait, M. U. Qadir, J. M. Conroy, S. I. Matsui, N. J. Nowak, and M. R. Baer, “Double minute chromosomes in acute myeloid leukemia and myelodysplastic syndrome: identification of new amplification regions by fluorescence in situ hybridization and spectral karyotyping,” Genes Chromosomes and Cancer, vol. 34, no. 1, pp. 42–47, 2002.
[96]  S. J. Kaneko, T. Gerasimova, S. T. Smith, K. O. Lloyd, K. Suzumori, and S. R. Young, “CA125 and UQCRFS1 FISH studies of ovarian carcinoma,” Gynecologic Oncology, vol. 90, no. 1, pp. 29–36, 2003.
[97]  L. Wang, S. K. McDonnell, S. J. Hebbring et al., “Polymorphisms in mitochondrial genes and prostate cancer risk,” Cancer Epidemiology Biomarkers and Prevention, vol. 17, no. 12, pp. 3558–3566, 2008.
[98]  K. K. Singh, V. Ayyasamy, K. M. Owens, M. S. Koul, and M. Vujcic, “Mutations in mitochondrial DNA polymerase-γ promote breast tumorigenesis,” Journal of Human Genetics, vol. 54, no. 9, pp. 516–524, 2009.
[99]  M. Blomberg Jensen, H. Leffers, J. H. Petersen, G. Daugaard, N. E. Skakkebaek, and E. Rajpert-De Meyts, “Association of the polymorphism of the CAG repeat in the mitochondrial DNA polymerase gamma gene (POLG) with testicular germ-cell cancer,” Annals of Oncology, vol. 19, no. 11, pp. 1910–1914, 2008.
[100]  M. M. Desouki, M. Kulawiec, S. Bansal, G. Das, and K. K. Singh, “Cross talk between mitochondria and superoxide generating NADPH oxidase in breast and ovarian tumors,” Cancer Biology and Therapy, vol. 4, no. 12, pp. 1367–1373, 2005.
[101]  K. Kahlos, S. Anttila, T. Asikainen et al., “Manganese superoxide dismutase in healthy human pleural mesothelium and in malignant pleural mesothelioma,” American Journal of Respiratory Cell and Molecular Biology, vol. 18, no. 4, pp. 570–580, 1998.
[102]  Y. Hu, D. G. Rosen, Y. Zhou et al., “Mitochondrial manganese-superoxide dismutase expression in ovarian cancer: role in cell proliferation and response to oxidative stress,” The Journal of Biological Chemistry, vol. 280, no. 47, pp. 39485–39492, 2005.
[103]  E. Tsanou, E. Ioachim, E. Briasoulis et al., “Immunohistochemical expression of superoxide dismutase (MnSOD) anti-oxidant enzyme in invasive breast carcinoma,” Histology and Histopathology, vol. 19, no. 3, pp. 807–813, 2004.
[104]  S. M. Tsai, M. F. Hou, S. H. Wu et al., “Expression of manganese superoxide dismutase in patients with breast cancer,” Kaohsiung Journal of Medical Sciences, vol. 27, no. 5, pp. 167–172, 2011.
[105]  F. Ria, M. Landriscina, F. Remiddi et al., “The level of manganese superoxide dismutase content is an independent prognostic factor for glioblastoma. Biological mechanisms and clinical implications,” British Journal of Cancer, vol. 84, no. 4, pp. 529–534, 2001.
[106]  C. K. Park, J. H. Jung, M. J. Moon et al., “Tissue expression of manganese superoxide dismutase is a candidate prognostic marker for glioblastoma,” Oncology, vol. 77, no. 3-4, pp. 178–181, 2009.
[107]  D. Ratnasinghe, J. A. Tangrea, M. R. Andersen et al., “Glutathione peroxidase codon 198 polymorphism variant increases lung cancer risk,” Cancer Research, vol. 60, no. 22, pp. 6381–6383, 2000.
[108]  D. Angnani, O. Camacho-Vanegas, C. Camacho, et al., “Decreased levels of serum glutathione peroxidase 3 are associated with papillary serous ovarian cancer and disease progression,” Journal of Ovarian Cancer, vol. 4, article 18, 2011.
[109]  E. Falck, S. Karlsson, J. Carlsson, G. Helenius, M. Karlsson, and K. Klinga-Levan, “Loss of glutathione peroxidase 3 expression is correlated with epigenetic mechanisms in endometrial adenocarcinoma,” Cancer Cell International, vol. 10, article 46, 2010.
[110]  C. Glorieux, N. Dejeans, B. Sid, R. Beck, P. B. Calderon, and J. Verrax, “Catalase overexpression in mammary cancer cells leads to a less aggressive phenotype and an altered response to chemotherapy,” Biochemical Pharmacology, vol. 82, no. 10, pp. 1384–1390, 2011.
[111]  J. Goh, L. Enns, S. Fatemie et al., “Mitochondrial targeted catalase suppresses invasive breast cancer in mice,” BMC Cancer, vol. 11, article 191, 2011.
[112]  M. Bakhanashvili, S. Grinberg, E. Bonda, A. J. Simon, S. Moshitch-Moshkovitz, and G. Rahav, “P53 in mitochondria enhances the accuracy of DNA synthesis,” Cell Death and Differentiation, vol. 15, no. 12, pp. 1865–1874, 2008.
[113]  J. Qian, K. Hirasawa, D. G. Bostwick et al., “Loss of p53 and c-myc overrepresentation in stage T2-3 N1-3M0 prostate cancer are potential markers for cancer progression,” Modern Pathology, vol. 15, no. 1, pp. 35–44, 2002.
[114]  I. Kim, S. Rodriguez-Enriquez, and J. J. Lemasters, “Minireview: selective degradation of mitochondria by mitophagy,” Archives of Biochemistry and Biophysics, vol. 462, no. 2, pp. 245–253, 2007.
[115]  B. Antonsson, “Mitochondria and the Bcl-2 proteins in apoptosis signaling pathways,” Molecular and Cellular Biochemistry, vol. 256-257, no. 1-2, pp. 141–155, 2004.
[116]  N. Rampino, H. Yamamoto, Y. Ionov et al., “Somatic frameshift mutations in the BAX gene in colon cancers of the microsatellite mutator phenotype,” Science, vol. 275, no. 5302, pp. 967–969, 1997.
[117]  P. Costantini, E. Jacotot, D. Decaudin, and G. Kroemer, “Mitochondrion as a novel target of anticancer chemotherapy,” Journal of the National Cancer Institute, vol. 92, no. 13, pp. 1042–1053, 2000.
[118]  D. Hanahan and R. A. Weinberg, “The hallmarks of cancer,” Cell, vol. 100, no. 1, pp. 57–70, 2000.
[119]  A. Schubert and S. Grimm, “Cyclophilin D, a component of the permeability transition-pore, is an apoptosis repressor,” Cancer Research, vol. 64, no. 1, pp. 85–93, 2004.
[120]  D. Whitacker-Menezes, U. E. Martinez-Outschoorn, N. Flomenberg, et al., “Hyperactivation of oxidative mitochondrial metabolism in epithelial cancer cells in situ: visualizing the therapeutic effect of metformin in tumor tissue,” Cell Cycle, vol. 10, no. 23, pp. 4047–4064, 2011.
[121]  C. Wang, G. L. Zhou, S. Vedantam, P. Li, and J. Field, “Mitochondrial shuttling of CAP1 promotes actin- and cofilin-dependent apoptosis,” Journal of Cell Science, vol. 121, no. 17, pp. 2913–2920, 2008.
[122]  I. M. Fearnley, J. Carroll, R. J. Shannon, M. J. Runswick, J. E. Walker, and J. Hirst, “GRIM-19, a cell death regulatory gene product, is a subunit of bovine mitochondrial NADH: ubiquinone oxidoreductase (complex I),” The Journal of Biological Chemistry, vol. 276, no. 42, pp. 38345–38348, 2001.
[123]  G. Huang, H. Lu, A. Hao et al., “GRIM-19, a cell death regulatory protein, is essential for assembly and function of mitochondrial complex I,” Molecular and Cellular Biology, vol. 24, no. 19, pp. 8447–8456, 2004.
[124]  I. Alchanati, S. C. Nallar, P. Sun et al., “A proteomic analysis reveals the loss of expression of the cell death regulatory gene GRIM-19 in human renal cell carcinomas,” Oncogene, vol. 25, no. 54, pp. 7138–7147, 2006.
[125]  L. B. Gong, X. L. Luo, S. Y. Liu, D. D. Tao, J. P. Gong, and J. B. Hu, “Correlations of GRIM-19 and its target gene product STAT3 to malignancy of human colorectal carcinoma,” Ai Zheng, vol. 26, no. 7, pp. 683–687, 2007.
[126]  Y. Zhou, M. Li, Y. Wei et al., “Down-regulation of GRIM-19 expression is associated with hyperactivation of STAT3-induced gene expression and tumor growth in human cervical cancers,” Journal of Interferon and Cytokine Research, vol. 29, no. 10, pp. 695–703, 2009.
[127]  X. Y. Fan, Z. F. Jiang, L. Cai, and R. Y. Liu, “Expression and clinical significance of GRIM-19 in lung cancer,” Medical Oncology. In press.
[128]  S. Moreira, M. Correia, P. Soares, and V. Máximo, “GRIM-19 function in cancer development,” Mitochondrion, vol. 11, no. 5, pp. 693–699, 2011.
[129]  B. H. Kang and D. C. Altieri, “Compartmentalized cancer drug discovery targeting mitochondrial Hsp90 chaperones,” Oncogene, vol. 28, no. 42, pp. 3681–3688, 2009.
[130]  T. Lebret, R. W. G. Watson, V. Molinié et al., “Heat shock proteins HSP27, HSP60, HSP70, and HSP90: expression in bladder carcinoma,” Cancer, vol. 98, no. 5, pp. 970–977, 2003.
[131]  L. Xu, L. A. Voloboueva, Y. Ouyang, J. F. Emery, and R. G. Giffard, “Overexpression of mitochondrial Hsp70/Hsp75 in rat brain protects mitochondria, reduces oxidative stress, and protects from focal ischemia,” Journal of Cerebral Blood Flow and Metabolism, vol. 29, no. 2, pp. 365–374, 2009.
[132]  B. H. Kang, J. Plescia, T. Dohi, J. Rosa, S. J. Doxsey, and D. C. Altieri, “Regulation of tumor cell mitochondrial homeostasis by an organelle-specific HSP90 chaperone network,” Cell, vol. 131, no. 2, pp. 257–270, 2007.
[133]  J. C. Ghosh, M. D. Siegelin, T. Dohi, and D. C. Altieri, “Heat shock protein 60 regulation of the mitochondrial permeability transition pore in tumor cells,” Cancer Research, vol. 70, no. 22, pp. 8988–8993, 2010.
[134]  F. Xiang, Y. S. Huang, X. H. Shi, and Q. Zhang, “Mitochondrial chaperone tumour necrosis factor receptor-associated protein 1 protects cardiomyocytes from hypoxic injury by regulating mitochondrial permeability transition pore opening,” The Federation of the Societies of Biochemistry and Molecular Biology Journal, vol. 277, no. 8, pp. 1929–1938, 2010.
[135]  S. Moalic-Juge, B. Liagre, R. Duval et al., “The anti-apoptotic property of NS-398 at high dose can be mediated in part through NF-kappaB activation, hsp70 induction and a decrease in caspase-3 activity in human osteosarcoma cells,” International Journal of Oncology, vol. 20, no. 6, pp. 1255–1262, 2002.
[136]  C. Gaudin, F. Kremer, E. Angevin, V. Scott, and F. Triebel, “A hsp70-2 mutation recognized by CTL on a human renal cell carcinoma,” Journal of Immunology, vol. 162, no. 3, pp. 1730–1738, 1999.
[137]  F. Capello, S. David, N. Ardizzone, et al., “Expression of heat shock proteins HSP10, HSP27, HSP60, HSP70, and HSP90 in urothelial carcinoma of urinary bladder,” Journal of Cancer Molecules, vol. 2, no. 2, pp. 73–77, 2006.
[138]  J. J. Hansen, P. Bross, M. Westergaard et al., “Genomic structure of the human mitochondrial chaperonin genes: HSP60 and HSP10 are localised head to head on chromosome 2 separated by a bidirectional promoter,” Human Genetics, vol. 112, no. 1, pp. 71–77, 2003.
[139]  F. Cappello, S. David, F. Rappa et al., “The expression of Hsp60 and Hsp10 in large bowel carcinomas with lymph node metastases,” BioMedCentral Cancer, vol. 5, article 139, 2005.
[140]  P. E. Castle, R. Ashfaq, F. Ansari, and C. Y. Muller, “Immunohistochemical evaluation of heat shock proteins in normal and preinvasive lesions of the cervix,” Cancer Letters, vol. 229, no. 2, pp. 245–252, 2005.
[141]  Y. J. Hwang, S. P. Lee, S. Y. Kim et al., “Expression of heat shock protein 60 kDa is upregulated in cervical cancer,” Yonsei Medical Journal, vol. 50, no. 3, pp. 399–406, 2009.
[142]  C. Castilla, B. Congregado, J. M. Conde et al., “Immunohistochemical expression of Hsp60 correlates with tumor progression and hormone resistance in prostate cancer,” Journal of Urology, vol. 76, no. 4, pp. 1017.e1–1017.e6, 2010.
[143]  A. Faried, M. Sohda, M. Nakajima, T. Miyazaki, H. Kato, and H. Kuwano, “Expression of heat-shock protein Hsp60 correlated with the apoptotic index and patient prognosis in human oesophageal squamous cell carcinoma,” European Journal of Cancer, vol. 40, no. 18, pp. 2804–2811, 2004.
[144]  J. Schneider, E. Jiménez, K. Marenbach, H. Romero, D. Marx, and H. Meden, “Immunohistochemical detection of HSP60-expression in human ovarian cancer. Correlation with survival in a series of 247 patients,” Anticancer Research, vol. 19, no. 3, pp. 2141–2146, 1999.
[145]  Z. Rui, J. Jian-Guo, T. Yuan-Peng, P. Hai, and R. Bing-Gen, “Use of serological proteomic methods to find biomarkers associated with breast cancer,” Proteomics, vol. 3, no. 4, pp. 433–439, 2003.
[146]  S. P. Langdon, G. J. Rabiasz, G. L. Hirst et al., “Expression of the heat shock protein HSP27 in human ovarian cancer,” Clinical Cancer Research, vol. 1, no. 12, pp. 1603–1609, 1995.
[147]  M. Zhao, F. Shen, Y. X. Yin, Y. Y. Yang, D. J. Xiang, and Q. Chen, “Increased expression of heat shock protein 27 correlates with peritoneal metastasis in epithelial ovarian cancer,” Reproductive Science, vol. 19, no. 7, pp. 748–753, 2012.
[148]  O. Straume, T. Shimamura, M. J. G. Lampa, et al., “Suppression of heat shock protein 27 induces long-term dormancy in human breast cancer,” Proceedings of the National Academy of Sciences, vol. 109, no. 22, pp. 8699–8704, 2012.
[149]  N. Pfanner and N. Wiedemann, “Mitochondrial protein import: two membranes, three translocases,” Current Opinion in Cell Biology, vol. 14, no. 4, pp. 400–411, 2002.
[150]  F. Sotgia, D. Whitaker-Menezes, U. E. Martinez-Outschoon, et al., “Mitochondrial metabolism in cancer metastasis,” Cell Cycle, vol. 11, no. 7, pp. 1445–1454, 2012.
[151]  X. Xu, M. Qiao, Y. Zhang et al., “Quantitative proteomics study of breast cancer cell lines isolated from a single patient: discovery of TIMM17A as a marker for breast cancer,” Proteomics, vol. 10, no. 7, pp. 1374–1390, 2010.
[152]  M. Salhab, N. Patani, W. Jiang, and K. Mokbel, “High TIMM17A expression is associated with adverse pathological and clinical outcomes in human breast cancer,” Breast Cancer, vol. 19, no. 2, pp. 153–160, 2012.
[153]  A. P. Sutter, K. Maaser, P. Grabowski et al., “Peripheral benzodiazepine receptor ligands induce apoptosis and cell cycle arrest in human hepatocellular carcinoma cells and enhance chemosensitivity to paclitaxel, docetaxel, doxorubicin and the Bcl-2 inhibitor HA14-1,” Journal of Hepatology, vol. 41, no. 5, pp. 799–807, 2004.
[154]  Z. Han, R. S. Slack, W. Li, and V. Papadopoulos, “Expression of peripheral benzodiazepine receptor (PBR) in human tumors: relationship to breast, colorectal, and prostate tumor progression,” Journal of Receptors and Signal Transduction, vol. 23, no. 2-3, pp. 225–238, 2003.
[155]  M. Volante, S. La Rosa, I. Castellano, G. Finzi, C. Capella, and G. Bussolati, “Clinico-pathological features of a series of 11 oncocytic endocrine tumours of the pancreas,” Virchows Archiv, vol. 448, no. 5, pp. 545–551, 2006.
[156]  J. M. Cuezva, M. Krajewska, M. L. de Heredia et al., “The bioenergetic signature of cancer: a marker of tumor progression,” Cancer Research, vol. 62, no. 22, pp. 6674–6681, 2002.
[157]  T. J. Kwon, J. Y. Ro, and B. Mackay, “Clear-cell carcinoma: an ultrastructural study of 57 tumors from various sites,” Ultrastructural Pathology, vol. 20, no. 6, pp. 519–527, 1996.
[158]  I. Bosun, “Ultramicroscopic studies in retinoblastoma,” Oftalmologia, vol. 41, no. 3, pp. 231–233, 1997.
[159]  S. R. Bornstein, J. W. Brown, A. Carballeira, J. Goodman, W. A. Scherbaum, and L. M. Fishman, “Ultrastructural dynamics of mitochondrial morphology in varying functional forms of human adrenal cortical adenoma,” Hormone and Metabolic Research, vol. 28, no. 4, pp. 177–182, 1996.
[160]  R. W. Gilkerson, D. H. Margineantu, R. A. Capaldi, and J. M. L. Selker, “Mitochondrial DNA depletion causes morphological changes in the mitochondrial reticulum of cultured human cells,” The Federation of the Societies of Biochemistry and Molecular Biology Letters, vol. 474, no. 1, pp. 1–4, 2000.
[161]  B. Tandler and F. H. Shipkey, “Ultrastructure of Warthin's tumor. I. Mitochondria,” Journal of Ultrasructure Research, vol. 11, no. 3-4, pp. 292–305, 1964.
[162]  H. K. Kim, W. S. Park, S. H. Kang, et al., “Mitochondrial alterations in human gastric carcinoma cell line,” American Journal of Physiology Cell Physiology, vol. 293, no. 2, pp. C761–C771, 2007.
[163]  S. K. Tickoo, M. W. Lee, J. N. Eble et al., “Ultrastructural observations on mitochondria and microvesicles in renal oncocytoma, chromophobe renal cell carcinoma, and eosinophilic variant of conventional (clear cell) renal cell carcinoma,” American Journal of Surgical Pathology, vol. 24, no. 9, pp. 1247–1256, 2000.
[164]  L. Putignani, S. Raffa, R. Pescosolido, et al., “Preliminary evidences on mitochondrial injury and impaired oxidative metabolism in breast cancer,” Mitochondrion, vol. 12, no. 3, pp. 363–396, 2012.
[165]  G. B. John, Y. Shang, L. Li et al., “The mitochondrial inner membrane protein mitofilin controls cristae morphology,” Molecular Biology of the Cell, vol. 16, no. 3, pp. 1543–1554, 2005.
[166]  L. Scorrano, M. Ashiya, K. Buttle et al., “A distinct pathway remodels mitochondrial cristae and mobilizes cytochrome c during apoptosis,” Developmental Cell, vol. 2, no. 1, pp. 55–67, 2002.
[167]  M. Lutter, G. A. Perkins, and X. Wang, “The pro-apoptotic Bcl-2 family member tBid localizes to mitochondrial contact sites,” BioMedCentral Cell Biology, vol. 2, article 22, 2001.
[168]  F. Legros, A. Lombès, P. Frachon, and M. Rojo, “Mitochondrial fusion in human cells is efficient, requires the inner membrane potential, and is mediated by mitofusins,” Molecular Biology of the Cell, vol. 13, no. 12, pp. 4343–4354, 2002.
[169]  J. Rehman, H. J. Zhang, P. T. Toth, et al., “Inhibition of mitochondrial fission prevents cell cycle progression in lung cancer,” The Federation of the Societies of Biochemistry and Molecular Biology Journal, vol. 26, no. 5, pp. 2175–2186, 2012.
[170]  Y. Y. Chiang, S. L. Chen, Y. T. Hsiao et al., “Nuclear expression of dynamin-related protein 1 in lung adenocarcinomas,” Modern Pathology, vol. 22, no. 9, pp. 1139–1150, 2009.
[171]  D. M. Su, Q. Zhang, X. Wang et al., “Two types of human malignant melanoma cell lines revealed by expression patterns of mitochondrial and survival-apoptosis genes: implications for malignant melanoma therapy,” Molecular Cancer Therapeutics, vol. 8, no. 5, pp. 1292–1304, 2009.
[172]  B. De Paepe, J. Smet, M. Lammens et al., “Immunohistochemical analysis of the oxidative phosphorylation complexes in skeletal muscle from patients with mitochondrial DNA encoded tRNA gene defects,” Journal of Clinical Pathology, vol. 62, no. 2, pp. 172–176, 2009.
[173]  B. De Paepe, J. Smet, J. G. Leroy et al., “Diagnostic value of immunostaining in cultured skin fibroblasts from patients with oxidative phosphorylation defects,” Pediatric Research, vol. 59, no. 1, pp. 2–6, 2006.
[174]  B. De Paepe, J. Smet, A. Vanlander, et al., “Fluorescence imaging of mitochondria in cultured skin fibroblasts: a useful method for the detection of oxidative phosphorylation defects,” Pediatric Research, vol. 72, no. 3, pp. 232–240, 2012.
[175]  J. P. Jakupciak, A. Maggrah, S. Maragh et al., “Facile whole mitochondrial genome resequencing from nipple aspirate fluid using MitoChip v2.0,” BioMedCentral Cancer, vol. 8, article 95, 2008.
[176]  L. F. Dong, P. Low, J. C. Dyason et al., “α-Tocopheryl succinate induces apoptosis by targeting ubiquinone-binding sites in mitochondrial respiratory complex II,” Oncogene, vol. 27, no. 31, pp. 4324–4335, 2008.
[177]  S. J. Ralph and J. Neuzil, “Mitochondria as targets for cancer therapy,” Molecular Nutrition & Food Research, vol. 53, no. 1, pp. 9–28, 2009.
[178]  J. Rohlena, L. F. Dong, K. Kluckova, et al., “Mitochondrially targeted alpha-Tocopheryl succinate is antiangiogenic: potential benefit against tumor angiogenesis but caution against wound healing,” Antioxidants and Redox Signaling, vol. 15, no. 12, pp. 2923–2935, 2011.

Full-Text

comments powered by Disqus

Contact Us

service@oalib.com

QQ:3279437679

WhatsApp +8615387084133

WeChat 1538708413