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...
PPAR Research  2013 

Therapeutic Implications of Targeting Energy Metabolism in Breast Cancer

DOI: 10.1155/2013/109285

Full-Text   Cite this paper   Add to My Lib

Abstract:

PPARs are ligand activated transcription factors. PPARγ agonists have been reported as a new and potentially efficacious treatment of inflammation, diabetes, obesity, cancer, AD, and schizophrenia. Since cancer cells show dysregulation of glycolysis they are potentially manageable through changes in metabolic environment. Interestingly, several of the genes involved in maintaining the metabolic environment and the central energy generation pathway are regulated or predicted to be regulated by PPARγ. The use of synthetic PPARγ ligands as drugs and their recent withdrawal/restricted usage highlight the lack of understanding of the molecular basis of these drugs, their off-target effects, and their network. These data further underscores the complexity of nuclear receptor signalling mechanisms. This paper will discuss the function and role of PPARγ in energy metabolism and cancer biology in general and its emergence as a promising therapeutic target in breast cancer. 1. Introduction The peroxisome proliferator-activated receptors (PPARs) are ligand activated transcription factors, belonging to the nuclear receptor superfamily, that control the expression of genes involved in organogenesis, inflammation, cell differentiation, proliferation, lipid, and carbohydrate metabolism [1, 2]. PPARs activated by their selected ligands, heterodimerizes and its receptor with the 9-cis-retinoic acid receptor, they then bind to peroxisome proliferator response elements (PPREs), specific sequences in their target genes. The consensus PPRE site consists of a direct repeat of the sequence AGGTCA separated by a single/double nucleotide, which is designated as DR-1 site/DR-2 site [3] (Figure 1). Each major isoforms of PPAR (PPARα, PPARβ/δ, and PPARγ), encoded by a different gene, performs different functions and exhibit different tissue localizations in many parts of the human body [4]. The peroxisome proliferator-activated receptor γ (PPARγ) is the most extensively studied subtype of the PPARs [5]. PPARγ is expressed in adipose tissue, colon, immune system, hematopoietic cells, and retina involved in lipid anabolism, adipocyte differentiation, control of inflammation, macrophage maturation, embryo implantation, and molecular targets of antidiabetic thiazolidinediones [6]. Its role in cancer development and potential as a target for cancer prevention and treatment strategies has been noted in recent years. Activation of PPARγ could possibly be an approach to induce differentiation in cells thereby inhibiting proliferation of a variety of cancers. This antiproliferative effect

References

[1]  B. P. Kota, T. H. Huang, and B. D. Roufogalis, “An overview on biological mechanisms of PPARs,” Pharmacological Research, vol. 51, no. 2, pp. 85–94, 2005.
[2]  M. Tous, N. Ferré, A. Rull et al., “Dietary cholesterol and differential monocyte chemoattractant protein-1 gene expression in aorta and liver of apo E-deficient mice,” Biochemical and Biophysical Research Communications, vol. 340, no. 4, pp. 1078–1084, 2006.
[3]  H. Castelein, T. Gulick, P. E. Declercq, G. P. Mannaerts, D. D. Moore, and M. I. Baes, “The peroxisome proliferator activated receptor regulates malic enzyme gene expression,” The Journal of Biological Chemistry, vol. 269, no. 43, pp. 26754–26758, 1994.
[4]  M. van Bilsen, G. J. van der Vusse, A. J. Gilde, M. Lindhout, and K. A. J. M. van der Lee, “Peroxisome proliferator-activated receptors: lipid binding proteins controling gene expression,” Molecular and Cellular Biochemistry, vol. 239, no. 1-2, pp. 131–138, 2002.
[5]  J. J. Mansure, R. Nassim, and W. Kassouf, “Peroxisome proliferator-activated receptor gamma in bladder cancer: a promising therapeutic target in cancer,” in Cellular and Genetic Practices for Translational Medicine, vol. 8, no. 7, pp. 169–195, Research Signpost, 2011.
[6]  J. I. Park, “The role of 15d-PGJ2, a natural ligand for peroxisome proliferator-activated receptor γ (PPARγ), in cancer,” Pharmacological Research, vol. 51, no. 2, pp. 85–94, 2005.
[7]  C. C. Woo, S. Y. Loo, V. Gee et al., “Anticancer activity of thymoquinone in breast cancer cells: possible involvement of PPAR-γ pathway,” Biochemical Pharmacology, vol. 82, no. 5, pp. 464–475, 2011.
[8]  Y. L. Lu, G. L. Li, H. L. Huang, J. Zhong, and L. C. Dai, “Peroxisome proliferator-activated receptor-γ 34C > G polymorphism and colorectal cancer risk: a meta-analysis,” World Journal of Gastroenterology, vol. 16, no. 17, pp. 2170–2175, 2010.
[9]  G. Venkatachalam, A. P. Kumar, L. S. Yue, S. Pervaiz, M. V. Clement, and M. K. Sakharkar, “Computational identification and experimental validation of PPRE motifs in NHE1 and MnSOD genes of human,” BMC Genomics, vol. 10, supplement 3, article S5, 2009.
[10]  Y. Jeong, Y. Xie, W. Lee et al., “Research resource: diagnostic and therapeutic potential of nuclear receptor expression in lung cancer,” Molecular Endocrinology, vol. 26, no. 8, pp. 1443–1454, 2012.
[11]  W. Motomura, T. Okumura, N. Takahashi, T. Obara, and Y. Kohgo, “Activation of peroxisome proliferator-activated receptor γ by troglitazone inhibits cell growth through the increase of p27Kip1 in human pancreatic carcinoma cells,” Cancer Research, vol. 60, no. 19, pp. 5558–5564, 2000.
[12]  R. Govindarajan, L. Ratnasinghe, D. L. Simmons et al., “Thiazolidinediones and the risk of lung, prostate, and colon cancer in patients with diabetes,” Journal of Clinical Oncology, vol. 25, no. 12, pp. 1476–1481, 2007.
[13]  H. N. Yu, Y. R. Lee, E. M. Noh et al., “Induction of G1 phase arrest and apoptosis in MDA-MB-231 breast cancer cells by troglitazone, a synthetic peroxisome proliferator-activated receptor γ (PPARγ) ligand,” Cell Biology International, vol. 32, no. 8, pp. 906–912, 2008.
[14]  D. Bonofiglio, S. Aquila, S. Catalano et al., “Peroxisome proliferator-activated receptor-γ activates p53 gene promoter binding to the nuclear factor-κB sequence in human MCF7 breast cancer cells,” Molecular Endocrinology, vol. 20, no. 12, pp. 3083–3092, 2006.
[15]  M. Pignatelli, C. Cocca, A. Santos, and A. Perez-Castillo, “Enhancement of BRCA1 gene expression by the peroxisome proliferator-activated receptor γ in the MCF-7 breast cancer cell line,” Oncogene, vol. 22, no. 35, pp. 5446–5450, 2003.
[16]  G. L. Rubin, Y. Zhao, A. M. Kalus, and E. R. Simpson, “Peroxisome proliferator-activated receptor γ ligands inhibit estrogen biosynthesis in human breast adipose tissue: possible implications for breast cancer therapy,” Cancer Research, vol. 60, no. 6, pp. 1604–1608, 2000.
[17]  F. Turturro, E. Friday, R. Fowler, D. Surie, and T. Welbourne, “Troglitazone acts on cellular pH and DNA synthesis through a peroxisome proliferator-activated receptor γ-independent mechanism in breast cancer-derived cell lines,” Clinical Cancer Research, vol. 10, no. 20, pp. 7022–7030, 2004.
[18]  A. G. Smith and G. E. Muscat, “Orphan nuclear receptors: therapeutic opportunities in skeletal muscle,” American Journal of Physiology, vol. 291, no. 2, pp. C203–C217, 2006.
[19]  S. Mukhopadhyay, S. K. Das, and S. Mukherjee, “Expression of Mn-superoxide dismutase gene in nontumorigenic and tumorigenic human mammary epithelial cells,” Journal of Biomedicine and Biotechnology, vol. 2004, no. 4, pp. 195–202, 2004.
[20]  P. Fedele, N. Calvani, A. Marino et al., “Targeted agents to reverse resistance to endocrine therapy in metastatic breast cancer: where are we now and where are we going?” Critical Reviews in Oncology/Hematology, vol. 51, no. 2, pp. 85–94, 2012.
[21]  T. N. Seyfried and L. M. Shelton, “Cancer as a metabolic disease,” Nutrition & Metabolism, vol. 7, article 7, 2010.
[22]  R. J. Klement and U. K?mmerer, “Is there a role for carbohydrate restriction in the treatment and prevention of cancer?” Nutrition & Metabolism, vol. 8, article 75, 2011.
[23]  X. Bi, Q. Lin, T. W. Foo, S. Joshi, T. You, H. M. Shen, et al., “Proteomic analysis of colorectal cancer reveals alterations in metabolic pathways: mechanism of tumorigenesis,” Molecular & Cellular Proteomics, vol. 5, pp. 1119–1130, 2006.
[24]  R. D. Unwin, R. A. Craven, P. Harnden et al., “Proteomic changes in renal cancer and co-ordinate demonstration of both the glycolytic and mitochondrial aspects of the Warburg effect,” Proteomics, vol. 3, no. 8, pp. 1620–1632, 2003.
[25]  B. Perroud, J. Lee, N. Valkova et al., “Pathway analysis of kidney cancer using proteomics and metabolic profiling,” Molecular Cancer, vol. 5, article 64, 2006.
[26]  A. Isidoro, E. Casado, A. Redondo et al., “Breast carcinomas fulfill the Warburg hypothesis and provide metabolic markers of cancer prognosis,” Carcinogenesis, vol. 26, no. 12, pp. 2095–2104, 2005.
[27]  L. M. Amon, S. J. Pitteri, C. I. Li et al., “Concordant release of glycolysis proteins into the plasma preceding a diagnosis of ER+ breast cancer,” Cancer Research, vol. 72, no. 8, pp. 1935–1942, 2012.
[28]  M. F. de Oliveira, N. D. Amoêdo, and F. D. Rumjanek, “Energy and redox homeostasis in tumor cells,” International Journal of Cell Biology, vol. 2012, Article ID 593838, 15 pages, 2012.
[29]  P. S. Ward and C. B. Thompson, “Metabolic reprogramming: a cancer hallmark even Warburg did not anticipate,” Cancer Cell, vol. 21, no. 3, pp. 297–308, 2012.
[30]  A. F. A. Mentis and E. Kararizou, “Metabolism and cancer: an up-to-date review of a mutual connection,” Asian Pacific Journal of Cancer Prevention, vol. 11, no. 6, pp. 1437–1444, 2010.
[31]  X. Zha, Q. Sun, and H. Zhang, “mTOR upregulation of glycolytic enzymes promotes tumor development,” Cell Cycle, vol. 10, no. 7, pp. 1015–1016, 2011.
[32]  J. A. Menendez, L. Vellon, C. Oliveras-Ferraros, S. Cufí, and A. Vazquez-Martin, “mTOR-regulated senescence and autophagy during reprogramming of somatic cells to pluripotency: a roadmap from energy metabolism to stem cell renewal and aging,” Cell Cycle, vol. 10, no. 21, pp. 3658–3677, 2011.
[33]  A. C. Williams, T. J. Collard, and C. Paraskeva, “An acidic environment leads to p53 dependent induction of apoptosis in human adenoma and carcinoma cell lines: implications for clonal selection during colorectal carcinogenesis,” Oncogene, vol. 18, no. 21, pp. 3199–3204, 1999.
[34]  S. Pavlides, D. Whitaker-Menezes, R. Castello-Cros et al., “The reverse Warburg effect: aerobic glycolysis in cancer associated fibroblasts and the tumor stroma,” Cell Cycle, vol. 8, no. 23, pp. 3984–4001, 2009.
[35]  M. Guppy, P. Leedman, X. Zu, and V. Russell, “Contribution by different fuels and metabolic pathways to the total ATP turnover of proliferating MCF-7 breast cancer cells,” Biochemical Journal, vol. 364, part 1, pp. 309–315, 2002.
[36]  R. A. Gatenby, E. T. Gawlinski, A. F. Gmitro, B. Kaylor, and R. J. Gillies, “Acid-mediated tumor invasion: a multidisciplinary study,” Cancer Research, vol. 66, no. 10, pp. 5216–5223, 2006.
[37]  P. I. Homem de Bittencourt Jr., C. M. Peres, M. M. Yano, M. H. Hirata, and R. Curi, “Pyruvate is a lipid precursor for rat lymphocytes in culture: evidence for a lipid exporting capacity,” Biochemistry and Molecular Biology International, vol. 30, no. 4, pp. 631–641, 1993.
[38]  H. Kondoh, M. E. Lleonart, D. Bernard, and J. Gil, “Protection from oxidative stress by enhanced glycolysis; a possible mechanism of cellular immortalization,” Histology and Histopathology, vol. 22, no. 1, pp. 85–90, 2007.
[39]  R. A. Gatenby and R. J. Gillies, “Why do cancers have high aerobic glycolysis?” Nature Reviews Cancer, vol. 4, no. 11, pp. 891–899, 2004.
[40]  R. J. Gillies and R. A. Gatenby, “Adaptive landscapes and emergent phenotypes: why do cancers have high glycolysis?” Journal of Bioenergetics and Biomembranes, vol. 39, no. 3, pp. 251–257, 2007.
[41]  G. Biamonti and J. F. Caceres, “Cellular stress and RNA splicing,” Trends in Biochemical Sciences, vol. 34, no. 3, pp. 146–153, 2009.
[42]  A. E. Greijer, P. van der Groep, D. Kemming et al., “Up-regualtion of gene expression by hypoxia is mediated predominantly by hypoxia-inducible factor I (HIF-I),” Journal of Pathology, vol. 206, no. 3, pp. 291–304, 2005.
[43]  N. Serkova and L. G. Boros, “Detection of resistance to imatinib by metabolic profiling: clinical and drug development implications,” American Journal of PharmacoGenomics, vol. 5, no. 5, pp. 293–302, 2005.
[44]  P. Hsu and D. Sabatini, “Cancer cell metabolism: Warburg and beyond,” Cell, vol. 134, no. 5, pp. 703–707, 2008.
[45]  J. Dhahbi, H. Kim, P. Mote, R. Beaver, and S. Spindler, “Temporal linkage between the phenotypic and genomic responses to caloric restriction,” Proceedings of the National Academy of Sciences of the United States of America, vol. 101, no. 15, pp. 5524–5529, 2004.
[46]  S. Y. Lunt and M. G. Vander Heiden, “Aerobic glycolysis: meeting the metabolic requirements of cell proliferation,” Annual Review of Cell and Developmental Biology, vol. 27, pp. 441–464, 2011.
[47]  S. John, J. N. Weiss, and B. Ribalet, “Subcellular localization of hexokinases I and II directs the metabolic fate of glucose,” PLoS ONE, vol. 6, no. 3, Article ID e17674, 2011.
[48]  S. P. Mathupala, Y. H. Ko, and P. L. Pedersen, “Hexokinase-2 bound to mitochondria: cancer's stygian link to the “Warburg effect” and a pivotal target for effective therapy,” Seminars in Cancer Biology, vol. 19, no. 1, pp. 17–24, 2009.
[49]  S. P. Mathupala, Y. H. Ko, and P. L. Pedersen, “Hexokinase II: cancer's double-edged sword acting as both facilitator and gatekeeper of malignancy when bound to mitochondria,” Oncogene, vol. 25, no. 34, pp. 4777–4786, 2006.
[50]  E. E. Mendoza, M. G. Pocceschi, X. Kong et al., “Control of glycolytic flux by AMP-activated protein kinase in tumor cells adapted to low pH,” Translational Oncology, vol. 5, no. 3, pp. 208–216, 2012.
[51]  E. C. Ferguson and J. C. Rathmell, “New roles for pyruvate kinase M2: working out the Warburg effect,” Trends in Biochemical Sciences, vol. 33, no. 8, pp. 359–362, 2008.
[52]  C. J. David, M. Chen, M. Assanah, P. Canoll, and J. L. Manley, “HnRNP proteins controlled by c-Myc deregulate pyruvate kinase mRNA splicing in cancer,” Nature, vol. 463, no. 7279, pp. 364–368, 2010.
[53]  B. Shashni, K. R. Sakharkar, Y. Nagasaki, and M. K. Sakharkar, “Glycolytic enzymes PGK1 and PKM2 as novel transcriptional targets of PPAR gamma in breast cancer pathophysiology,” Journal of Drug Targeting. In press.
[54]  G. Venkatachalam, A. P. Kumar, K. R. Sakharkar, S. Thangavel, M. V. Clement, and M. K. Sakharkar, “PPARγ disease gene network and identification of therapeutic targets for prostate cancer,” Journal of Drug Targeting, vol. 19, no. 9, pp. 781–796, 2011.
[55]  A. Lagana, J. Vadnais, P. U. Le et al., “Regulation of the formation of tumor cell pseudopodia by the Na(+)/H(+) exchanger NHE1,” Journal of Cell Science, vol. 113, part 20, pp. 3649–3662, 2000.
[56]  L. K. Putney, S. P. Denker, and D. L. Barber, “The changing face of the Na+/H+ exchanger, NHE1: structure, regulation, and cellular actions,” Annual Review of Pharmacology and Toxicology, vol. 42, pp. 527–552, 2002.
[57]  S. J. Reshkin, A. Bellizzi, S. Caldeira et al., “Na+/H+ exchanger-dependent intracellular alkalinization is an early event in malignant transformation and plays an essential role in the development of subsequent transformation-associated phenotypes,” The FASEB Journal, vol. 14, no. 14, pp. 2185–2197, 2000.
[58]  S. Grinstein and S. J. Dixon, “Ion transport, membrane potential, and cytoplasmic pH in lymphocytes: changes during activation,” Physiological Reviews, vol. 69, no. 2, pp. 417–481, 1989.
[59]  S. M. Bell, S. M. Schreiner, P. J. Schultheis, et al., “Targeted disruption of the murine NHE1 locus induces ataxia, growth retardation, and seizures,” American Journal of Physiology, vol. 276, pp. C788–C795, 1999.
[60]  J. Pouyssegur, A. Franchi, and G. Pages, “pHi, aerobic glycolysis and vascular endothelial growth factor in tumour growth,” Novartis Foundation Symposium, vol. 240, pp. 186–196, 2001.
[61]  F. Turturro, E. Friday, R. Fowler, D. Surie, and T. Welbourne, “Troglitazone acts on cellular pH and DNA synthesis through a peroxisome proliferator-activated receptor γ-independent mechanism in breast cancer-derived cell lines,” Clinical Cancer Research, vol. 10, no. 20, pp. 7022–7030, 2004.
[62]  S. Akram, H. F. C. Teong, L. Fliegel, S. Pervaiz, and M. V. Clément, “Reactive oxygen species-mediated regulation of the Na+-H+ exchanger 1 gene expression connects intracellular redox status with cells' sensitivity to death triggers,” Cell Death and Differentiation, vol. 13, no. 4, pp. 628–641, 2006.
[63]  S. A. Kliewer, J. M. Lenhard, T. M. Willson, I. Patel, D. C. Morris, and J. M. Lehmann, “A prostaglandin J2 metabolite binds peroxisome proliferator-activated receptor γ and promotes adipocyte differentiation,” Cell, vol. 83, no. 5, pp. 813–819, 1995.
[64]  H. Pelicano, D. Carney, and P. Huang, “ROS stress in cancer cells and therapeutic implications,” Drug Resistance Updates, vol. 7, no. 2, pp. 97–110, 2004.
[65]  L. Gibellini, M. Pinti, M. Nasi et al., “Interfering with ROS metabolism in cancer cells: the potential role of quercetin,” Cancers, vol. 2, no. 2, pp. 1288–1311, 2010.
[66]  K. B. Storey, “Oxidative stress: animal adaptations in nature,” Brazilian Journal of Medical and Biological Research, vol. 29, no. 12, pp. 1715–1733, 1996.
[67]  A. M. L. Janssen, C. B. Bosman, W. van Duijn et al., “Superoxide dismutases in gastric and esophageal cancer and the prognostic impact in gastric cancer,” Clinical Cancer Research, vol. 6, no. 8, pp. 3183–3192, 2000.
[68]  O. Kanbagli, G. Ozdemirler, T. Bulut, S. Yamaner, G. Ayka?-Toker, and M. Uysal, “Mitochondrial lipid peroxides and antioxidant enzymes in colorectal adenocarcinoma tissues,” Japanese Journal of Cancer Research, vol. 91, no. 12, pp. 1258–1263, 2000.
[69]  K. Punnonen, M. Ahotupa, K. Asaishi, M. Hy?ty, R. Kudo, and R. Punnonen, “Antioxidant enzyme activities and oxidative stress in human breast cancer,” Journal of Cancer Research and Clinical Oncology, vol. 120, no. 6, pp. 374–377, 1994.
[70]  C. S. Cobbs, D. S. Levi, K. Aldape, and M. A. Israel, “Manganese superoxide dismutase expression in human central nervous system tumors,” Cancer Research, vol. 56, no. 14, pp. 3192–3195, 1996.
[71]  E. A. Hileman, G. Achanta, and P. Huang, “Superoxide dismutase: an emerging target for cancer therapeutics,” Expert Opinion on Therapeutic Targets, vol. 5, no. 6, pp. 697–710, 2001.
[72]  T. Nishiura, K. Suzuki, T. Kawaguchi et al., “Elevated serum manganese superoxide dismutase in acute leukemias,” Cancer Letters, vol. 62, no. 3, pp. 211–215, 1992.
[73]  S. Senthil, R. M. Veerappan, M. Ramakrishna Rao, and K. V. Pugalendi, “Oxidative stress and antioxidants in patients with cardiogenic shock complicating acute myocardial infarction,” Clinica Chimica Acta, vol. 348, no. 1-2, pp. 131–137, 2004.
[74]  S. P. Weisberg, D. McCann, M. Desai, M. Rosenbaum, R. L. Leibel, and A. W. Ferrante Jr., “Obesity is associated with macrophage accumulation in adipose tissue,” The Journal of Clinical Investigation, vol. 112, no. 12, pp. 1796–1808, 2003.
[75]  G. R. Hajer, T. W. van Haeften, and F. L. Visseren, “Adipose tissue dysfunction in obesity, diabetes, and vascular diseases,” European Heart Journal, vol. 29, no. 24, pp. 2959–2971, 2008.
[76]  C. Kumar-Sinha, K. W. Ignatoski, M. E. Lippman, S. P. Ethier, and A. M. Chinnaiyan, “Transcriptome analysis of HER2 reveals a molecular connection to fatty acid synthesis,” Cancer Research, vol. 63, no. 1, pp. 132–139, 2003.
[77]  J. A. Menendez, I. Mehmi, V. A. Verma, P. K. Teng, and R. Lupu, “Pharmacological inhibition of fatty acid synthase (FAS): a novel therapeutic approach for breast cancer chemoprevention through its ability to suppress Her-2/neu (erbB-2) oncogene-induced malignant transformation,” Molecular Carcinogenesis, vol. 41, no. 3, pp. 164–178, 2004.
[78]  J. A. Menendez, B. P. Oza, E. Atlas, V. A. Verma, I. Mehmi, and R. Lupu, “Inhibition of tumor-associated fatty acid synthase activity antagonizes estradiol- and tamoxifen-induced agonist transactivation of estrogen receptor (ER) in human endometrial adenocarcinoma cells,” Oncogene, vol. 23, no. 28, pp. 4945–4958, 2004.
[79]  S. Hardy, Y. Langelier, and M. Prentki, “Oleate activates phosphatidylinositol 3-kinase and promotes proliferation and reduces apoptosis of MDA-MB-231 breast cancer cells, whereas palmitate has opposite effects,” Cancer Research, vol. 60, no. 22, pp. 6353–6358, 2000.
[80]  S. Hardy, W. El-Assaad, E. Przybytkowski, E. Joly, M. Prentki, and Y. Langelier, “Saturated fatty acid-induced apoptosis in MDA-MB-231 breast cancer cells. A role for cardiolipin,” The Journal of Biological Chemistry, vol. 278, no. 34, pp. 31861–31870, 2003.
[81]  B. Desvergne and W. Wahli, “Peroxisome proliferators-activated receptors: nuclear control of metabolism,” Endocrine Reviews, vol. 20, pp. 649–688, 1999.
[82]  G. Medina-Gomez, S. Gray, and A. Vidal-Puig, “Adipogenesis and lipotoxicity: role of peroxisome proliferator-activated receptor γ (PPARγ) and PPARγcoactivator-1 (PGC1),” Public Health Nutrition, vol. 10, no. 10, pp. 1132–1137, 2007.
[83]  L. Han, R. Zhou, J. Niu, M. A. McNutt, P. Wang, and T. Tong, “SIRT1 is regulated by a PPARγ-SIRT1 negative feedback loop associated with senescence,” Nucleic Acids Research, vol. 38, no. 21, pp. 7458–7471, 2010.
[84]  E. Giovannucci, D. M. Harlan, M. C. Archer et al., “Diabetes and cancer: a consensus report,” CA Cancer Journal for Clinicians, vol. 60, no. 4, pp. 207–221, 2010.
[85]  S. Liao, J. Li, L. Wang, Y. Zhang, and C. Wang, “Type 2 diabetes mellitus and characteristics of breast cancer in China,” The Asian Pacific Journal of Cancer Prevention, vol. 11, pp. 933–937, 2010.
[86]  P. Boyle, M. Boniol, A. Koechlin et al., “Diabetes and breast cancer risk: a meta-analysis,” British Journal of Cancer, vol. 51, no. 2, pp. 85–94, 2012.
[87]  D. O. Carpenter, “Environmental contaminants as risk factors for developing diabetes,” Reviews on Environmental Health, vol. 23, no. 1, pp. 59–74, 2008.
[88]  J. T. Huang, J. S. Welch, M. Ricote et al., “Interleukin-4-dependent production of PPAR-γ ligands in macrophages by 12/15-lipoxygenase,” Nature, vol. 400, no. 6742, pp. 378–382, 1999.
[89]  L. Nagy, P. Tontonoz, J. G. A. Alvarez, H. Chen, and R. M. Evans, “Oxidized LDL regulates macrophage gene expression through ligand activation of PPARγ,” Cell, vol. 93, no. 2, pp. 229–240, 1998.
[90]  J. Berger, P. Bailey, C. Biswas et al., “Thiazolidinediones produce a conformational change in peroxisomal proliferator-activated receptor-γ: binding and activation correlate with antidiabetic actions in db/db mice,” Endocrinology, vol. 137, no. 10, pp. 4189–4195, 1996.
[91]  J. M. Lehmann, L. B. Moore, T. A. Smith-Oliver, W. O. Wilkison, T. M. Willson, and S. A. Kliewer, “An antidiabetic thiazolidinedione is a high affinity ligand for peroxisome proliferator-activated receptor γ (PPARγ),” The Journal of Biological Chemistry, vol. 270, no. 22, pp. 12953–12956, 1995.
[92]  K. G. Lambe and J. D. Tugwood, “A human peroxisome-proliferator-activated receptor-γ is activated by inducers of adipogenesis, including thiazalidinedione drugs,” European Journal of Biochemistry, vol. 239, no. 1, pp. 1–7, 1996.
[93]  J. J. Acton III, R. M. Black, A. B. Jones, et al., “Benzoyl 2-methyl indoles as selective PPARgamma modulators,” Bioorganic & Medicinal Chemistry Letters, vol. 15, pp. 357–362, 2005.
[94]  A. M. Fair, Q. Dai, X. O. Shu et al., “Energy balance, insulin resistance biomarkers, and breast cancer risk,” Cancer Detection and Prevention, vol. 31, no. 3, pp. 214–219, 2007.
[95]  P. Pisani, “Hyper-insulinaemia and cancer, meta-analyses of epidemiological studies,” Archives of Physiology and Biochemistry, vol. 114, no. 1, pp. 63–70, 2008.
[96]  A. Belfiore, M. Genua, and R. Malaguarnera, “PPAR-γ agonists and their effects on IGF-I receptor signaling: implications for cancer,” PPAR Research, vol. 2009, Article ID 830501, 18 pages, 2009.
[97]  H. Y. J?rvinen, “Thiazolidinediones,” The New England Journal of Medicine, vol. 351, no. 11, pp. 1106–1118, 2004.
[98]  H. J. Burstein, G. D. Demetri, E. Mueller, P. Sarraf, B. M. Spiegelman, and E. P. Winer, “Use of the peroxisome proliferator-activated receptor (PPAR) γ ligand troglitazone as treatment for refractory breast cancer: a phase II study,” Breast Cancer Research and Treatment, vol. 79, no. 3, pp. 391–397, 2003.
[99]  S. Kawa, T. Nikaido, H. Unno, N. Usuda, K. Nakayama, and K. Kiyosawa, “Growth inhibition and differentiation of pancreatic cancer cell lines by PPARγ ligand troglitazone,” Pancreas, vol. 24, no. 1, pp. 1–7, 2002.
[100]  E. Mueller, M. Smith, P. Sarraf, et al., “Effects of ligand activation of peroxisome proliferator-activated receptor γ in human prostate cancer,” Proceedings of the National Academy of Sciences of the United States of America, vol. 97, no. 20, pp. 10990–10995, 2000.
[101]  T. Shimada, K. Kojima, K. Yoshiura, H. Hiraishi, and A. Terano, “Characteristics of the peroxisome proliferator activated receptor γ (PPARγ) ligand induced apoptosis in colon cancer cells,” Gut, vol. 50, no. 5, pp. 658–664, 2002.
[102]  M. Ricote, A. C. Li, T. M. Willson, C. J. Kelly, and C. K. Glass, “The peroxisome proliferator-activated receptor-γ is a negative regulator of macrophage activation,” Nature, vol. 391, no. 6662, pp. 79–82, 1998.
[103]  U. Kintscher, C. J. Lyon, and R. E. Law, “Angiotensin II, PPAR-gamma and atherosclerosis,” Frontiers in Bioscience, vol. 1, pp. 359–369, 2004.
[104]  G. Helbig, K. W. Christopherson, P. Bhat-Nakshatri et al., “NF-κB promotes breast cancer cell migration and metastasis by inducing the expression of the chemokine receptor CXCR4,” The Journal of Biological Chemistry, vol. 278, no. 24, pp. 21631–21638, 2003.
[105]  B. Sung, S. Park, B. P. Yu, and H. Y. Chung, “Amelioration of age-related inflammation and oxidative stress by PPARγ activator: suppression of NF-κB by 2, 4-thiazolidinedione,” Experimental Gerontology, vol. 41, pp. 590–599, 2006.
[106]  B. Sung, S. Park, B. P. Yu, and H. Y. Chung, “Modulation of PPAR in aging, inflammation, and calorie restriction,” The Journals of Gerontology A, vol. 59, no. 10, pp. B997–B1006, 2004.
[107]  K. L. Houseknecht, B. M. Cole, and P. J. Steele, “Peroxisome proliferator-activated receptor gamma (PPARγ) and its ligands: a review,” Domestic Animal Endocrinology, vol. 22, no. 1, pp. 1–23, 2002.
[108]  M. Ricote, A. C. Li, T. M. Willson, J. Kelly, and C. K. Glass, “The peroxisome proliferators activated receptor-gamma is a negative regulator of macrophage activation,” Nature, vol. 391, pp. 79–82, 1998.
[109]  T. K. Kerppola, D. Luk, and T. Curran, “Fos is a preferential target of glucocorticoid receptor inhibition of AP-1 activity in vitro,” Molecular and Cellular Biology, vol. 13, pp. 3782–3791, 1993.
[110]  Y. Kamei, L. Xu, T. Heinzel et al., “A CBP integrator complex mediates transcriptional activation and AP-1 inhibition by nuclear receptors,” Cell, vol. 85, no. 3, pp. 403–414, 1996.
[111]  C. K. Glass and M. G. Rosenfeld, “The coregulator exchange in transcriptional functions of nuclear receptors,” Genes & Development, vol. 14, pp. 121–141, 2000.
[112]  S. Tyagi, P. Gupta, A. S. Saini, C. Kaushal, and S. Sharma, “The peroxisome proliferator-activated receptor: a family of nuclear receptors role in various diseases,” Journal of Advanced Pharmaceutical Technology & Research, vol. 2, no. 4, pp. 236–240, 2011.

Full-Text

comments powered by Disqus

Contact Us

service@oalib.com

QQ:3279437679

WhatsApp +8615387084133