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...

The Role of mTOR Inhibitors for the Treatment of B-Cell Lymphomas

DOI: 10.1155/2012/435342

Full-Text   Cite this paper   Add to My Lib

Abstract:

Despite the fact that the majority of lymphomas initially respond to treatment, many patients relapse and die from disease that is refractory to current regimens. The need for new treatment strategies in lymphomas has led to the investigation and evaluation of novel agents that target cellular pathways. The mammalian target of rapamycin (mTOR) is a representative pathway that may be implicated in lymphomagenesis. Rapamycin and especially its derivatives (temsirolimus, everolimus, and deforolimus) represent the first described mTOR inhibitors. These agents have shown promising results in the treatment of lymphoid malignancies. On the other hand, new ATP-competitive mTOR inhibitors that provoke a broader inhibition of mTOR activity are in early stages of clinical development. The purpose of this paper is to summarize the existing knowledge about mTOR inhibitors and their use in the treatment of B-cell lymphomas. Relevant issues regarding mTOR biology in general as well as in B-cell lymphoid neoplasms are also discussed in short. 1. Introduction Current approaches in treating lymphoid malignancies have focused on the development of therapeutic regimens that selectively target dysregulated signal transduction pathways in neoplastic cells. Among aberrantly activated signaling cascades that are implicated in the pathogenesis of lymphomas is the mammalian target of rapamycin (mTOR) pathway, which is involved in many vital cellular processes [1]. Rapamycin and its analogs (rapalogs) comprise the classical mTOR inhibitors. A number of completed as well as other ongoing preclinical and clinical trials have tested these drugs in lymphomas, either as monotherapy or in combination with established chemotherapy [1]. Moreover, other anti-mTOR molecules, such as specific active-site TOR inhibitors (asTORi), with better pharmacological profiles are candidate drugs to be tested in clinical trials against lymphoid malignancies [2]. Herein we aim to review the results of trials with mTOR inhibitors in B-cell lymphomas. Firstly, the mTOR signaling network as well as possible aetiologic factors of aberrant activation of the mTORC1 signaling cascade in B-cell lymphoid malignancies are discussed in short. 2. mTOR Signaling Network Rapamycin (also known as sirolimus or Rapamune, Wyeth) is the first described mTOR inhibitor [3]. This drug, originally developed as an antifungal agent, was soon found to have immunosuppressive and antineoplasmatic actions [4]. Systemic efforts to decipher the molecular mechanisms of these actions led to the isolation of the mTOR protein and the

References

[1]  P. B. Johnston, R. Yuan, F. Cavalli, and T. E. Witzig, “Targeted therapy in lymphoma,” Journal of Hematology & Oncology, vol. 3, p. 45, 2010.
[2]  C. Vu and D. A. Fruman, “Target of rapamycin signaling in leukemia and lymphoma,” Clinical Cancer Research, vol. 16, pp. 5374–5380, 2010.
[3]  W. W. Ma and A. A. Adjei, “Novel agents on the horizon for cancer therapy,” CA: Cancer Journal for Clinicians, vol. 59, no. 2, pp. 111–137, 2009.
[4]  S. Wullschleger, R. Loewith, and M. N. Hall, “TOR signaling in growth and metabolism,” Cell, vol. 124, no. 3, pp. 471–484, 2006.
[5]  G. A. Soliman, “The mammalian target of rapamycin signalling network and gene regulation,” Current Opinion in Lipidology, vol. 16, no. 3, pp. 317–323, 2005.
[6]  K. Inoki, H. Ouyang, Y. Li, and K. L. Guan, “Signaling by target of rapamycin proteins in cell growth control,” Microbiology and Molecular Biology Reviews, vol. 69, no. 1, pp. 79–100, 2005.
[7]  K. G. Foster and D. C. Fingar, “Mammalian target of rapamycin (mTOR): conducting the cellular signaling symphony,” Journal of Biological Chemistry, vol. 285, no. 19, pp. 14071–14077, 2010.
[8]  T. R. Peterson, M. Laplante, C. C. Thoreen et al., “DEPTOR is an mTOR inhibitor frequently overexpressed in multiple myeloma cells and required for their survival,” Cell, vol. 137, no. 5, pp. 873–886, 2009.
[9]  S. Menon and B. D. Manning, “Common corruption of the mTOR signaling network in human tumors,” Oncogene, vol. 27, supplement 2, pp. S43–S51, 2008.
[10]  J. Huang and B. D. Manning, “The TSC1-TSC2 complex: a molecular switchboard controlling cell growth,” Biochemical Journal, vol. 412, no. 2, pp. 179–190, 2008.
[11]  R. M. Memmott and P. A. Dennis, “Akt-dependent and -independent mechanisms of mTOR regulation in cancer,” Cellular Signalling, vol. 21, no. 5, pp. 656–664, 2009.
[12]  J. A. McCubrey, L. S. Steelman, W. H. Chappell et al., “Roles of the Raf/MEK/ERK pathway in cell growth, malignant transformation and drug resistance,” Biochimica et Biophysica Acta, vol. 1773, no. 8, pp. 1263–1284, 2007.
[13]  M. Cully, A. Genevet, P. Warne et al., “A role for p38 stress-activated protein kinase in regulation of cell growth via TORC1,” Molecular and Cellular Biology, vol. 30, no. 2, pp. 481–495, 2010.
[14]  L. M. Ballou and R. Z. Lin, “Rapamycin and mTOR kinase inhibitors,” Journal of Chemical Biology, vol. 1, pp. 27–36, 2008.
[15]  N. Cybulski and M. N. Hall, “TOR complex 2: a signaling pathway of its own,” Trends in Biochemical Sciences, vol. 34, no. 12, pp. 620–627, 2009.
[16]  I. Tato, R. Bartrons, F. Ventura, and J. L. Rosa, “Amino acids activate mTOR complex 2 via PI3K/Akt signalling,” Journal of Biological Chemistry, vol. 286, no. 8, pp. 6128–6142, 2011.
[17]  A. Y. Choo and J. Blenis, “Not all substrates are treated equally Implications for mTOR, rapamycin-resistance and cancer therapy,” Cell Cycle, vol. 8, no. 4, pp. 567–572, 2009.
[18]  Q. Sun, X. Chen, J. Ma, et al., “Mammalian target of rapamycin up-regulation of pyruvate kinase isoenzyme type M2 is critical for aerobic glycolysis and tumor growth,” Proceedings of the National Academy of Sciences, vol. 108, pp. 4129–4134, 2011.
[19]  D. R. Alessi, L. R. Pearce, and J. M. García-Martínez, “New insights into mTOR signaling: mTORC2 and beyond,” Science Signaling, vol. 2, no. 67, p. pe27, 2009.
[20]  W. Li, M. Petrimpol, K. D. Molle, M. N. Hall, E. J. Battegay, and R. Humar, “Hypoxia-induced endothelial proliferation requires both mTORC1 and mTORC2,” Circulation Research, vol. 100, no. 1, pp. 79–87, 2007.
[21]  F. Zhang, X. Zhang, M. Li, et al., “mTOR complex component Rictor interacts with PKCzeta and regulates cancer cell metastasis,” Cancer Research, vol. 70, pp. 9360–9370, 2010.
[22]  A. Dutton, G. M. Reynolds, C. W. Dawson, L. S. Young, and P. G. Murray, “Constitutive activation of phosphatidyl-inositide 3 kinase contributes to the survival of Hodgkin's lymphoma cells through a mechanism involving Akt kinase and mTOR,” Journal of Pathology, vol. 205, no. 4, pp. 498–506, 2005.
[23]  R. E. Brown and N. R. Kamal, “p-p70S6K (Thr 389) expression in nodular sclerosing Hodgkin's disease as evidence for receptor tyrosine kinase signaling,” Annals of Clinical and Laboratory Science, vol. 35, no. 4, pp. 413–414, 2005.
[24]  M. El-Salem, P. N. Raghunath, M. Marzec et al., “Activation of mTORC1 signaling pathway in AIDS-related lymphomas,” American Journal of Pathology, vol. 175, no. 2, pp. 817–824, 2009.
[25]  J. De and R. E. Brown, “Tissue-microarray based immunohistochemical analysis of survival pathways in nodular sclerosing classical Hodgkin lymphoma as compared with Non-Hodgkin's lymphoma,” International Journal of Clinical and Experimental Medicine, vol. 3, no. 1, pp. 55–68, 2010.
[26]  B. Zheng, P. Flumara, Y. V. Li et al., “MEK/ERK pathway is aberrantly active in Hodgkin disease: a signaling pathway shared by CD30, CD40, and RANK that regulates cell proliferation and survival,” Blood, vol. 102, no. 3, pp. 1019–1027, 2003.
[27]  E. Peponi, E. Drakos, G. Reyes, V. Leventaki, G. Z. Rassidakis, and L. J. Medeiros, “Activation of mammalian target of rapamycin signaling promotes cell cycle progression and protects cells from apoptosis in mantle cell lymphoma,” American Journal of Pathology, vol. 169, no. 6, pp. 2171–2180, 2006.
[28]  M. Rudelius, S. Pittaluga, S. Nishizuka et al., “Constitutive activation of Akt contributes to the pathogenesis and survival of mantle cell lymphoma,” Blood, vol. 108, no. 5, pp. 1668–1676, 2006.
[29]  A. Psyrri, S. Papageorgiou, E. Liakata et al., “Phosphatidylinositol 3′-kinase catalytic subunit α gene amplification contributes to the pathogenesis of mantle cell lymphoma,” Clinical Cancer Research, vol. 15, no. 18, pp. 5724–5732, 2009.
[30]  P. Argyriou, S. G. Papageorgiou, V. Panteleon, et al., “Hypoxia-inducible factors in mantle cell lymphoma: implication for an activated mTORC1→HIF-1 pathway,” Annals of Hematology, vol. 90, pp. 315–322, 2011.
[31]  J. B. Dennison, M. Shanmugam, M. L. Ayres, et al., “8-Aminoadenosine inhibits Akt/mTOR and Erk signaling in mantle cell lymphoma,” Blood, vol. 116, pp. 5622–5630, 2010.
[32]  B. H. Yu, X. Y. Zhou, X. Y. Xiao, S. Y. Yan, T. Qin, and D. R. Shi, “Activation and clinicopathologic significance of AKT/mTOR signaling pathway in diffuse large B-cell lymphoma,” Chinese Journal of Pathology, vol. 38, no. 1, pp. 35–41, 2009.
[33]  M. Y. Zhao, A. Auerbach, A. M. D'Costa et al., “Phospho-p70S6K/p85S6K and cdc2/cdk1 are novel targets for diffuse large B-cell lymphoma combination therapy,” Clinical Cancer Research, vol. 15, no. 5, pp. 1708–1720, 2009.
[34]  Y. Baohua, Z. Xiaoyan, Z. Tiecheng, Q. Tao, and S. Daren, “Mutations of the PIK3CA gene in diffuse large B cell lymphoma,” Diagnostic Molecular Pathology, vol. 17, no. 3, pp. 159–165, 2008.
[35]  J. Abubaker, P. P. Bavi, S. Al-Harbi et al., “PIK3CA mutations are mutually exclusive with PTEN loss in diffuse large B-cell lymphoma,” Leukemia, vol. 21, no. 11, pp. 2368–2370, 2007.
[36]  S. Uddin, R. Bu, M. Ahmed et al., “Leptin receptor expression and its association with PI3K/AKT signaling pathway in diffuse large B-cell lymphoma,” Leukemia and Lymphoma, vol. 51, no. 7, pp. 1305–1314, 2010.
[37]  S. H. Kuo, P. Y. Yeh, L. T. Chen et al., “Overexpression of B cell activating factor of TNF family (BAFF) is associated with Helicobacter pylori independent growth of gastric diffuse large B-cell lymphoma with histologic evidence of MALT lymphoma,” Blood, vol. 112, no. 7, pp. 2927–2934, 2008.
[38]  E. P. M. Tjin, R. W. J. Groen, I. Vogelzang et al., “Functional analysis of HGF/MET signaling and aberrant HGF-activator expression in diffuse large B-cell lymphoma,” Blood, vol. 107, no. 2, pp. 760–768, 2006.
[39]  R. E. Davis, V. N. Ngo, G. Lenz et al., “Chronic active B-cell-receptor signalling in diffuse large B-cell lymphoma,” Nature, vol. 463, no. 7277, pp. 88–92, 2010.
[40]  L. Leseux, S. M. Hamdi, T. Al Saati et al., “Syk-dependent mTOR activation in follicular lymphoma cells,” Blood, vol. 108, no. 13, pp. 4156–4162, 2006.
[41]  M. Gupta, S. R. Dillon, S. C. Ziesmer et al., “A proliferation-inducing ligand mediates follicular lymphoma B-cell proliferation and cyclin D1 expression through phosphatidylinositol 3-kinase-regulated mammalian target of rapamycin activation,” Blood, vol. 113, no. 21, pp. 5206–5216, 2009.
[42]  L. Leseux, G. Laurent, C. Laurent et al., “PKC ζ-mTOR pathway: a new target for rituximab therapy in follicular lymphoma,” Blood, vol. 111, no. 1, pp. 285–291, 2008.
[43]  H. Zha, M. Raffeld, L. Charboneau et al., “Similarities of prosurvival signals in Bcl-2-positive and Bcl-2-negative follicular lymphomas identified by reverse phase protein microarray,” Laboratory Investigation, vol. 84, no. 2, pp. 235–244, 2004.
[44]  C. Gulmann, V. Espina, E. Petricoin III et al., “Proteomic analysis of apoptotic pathways reveals prognostic factors in follicular lymphoma,” Clinical Cancer Research, vol. 11, no. 16, pp. 5847–5855, 2005.
[45]  K. R. Calvo, B. Dabir, A. Kovach et al., “IL-4 protein expression and basal activation of Erk in vivo in follicular lymphoma,” Blood, vol. 112, no. 9, pp. 3818–3826, 2008.
[46]  P. Wlodarski, M. Kasprzycka, X. Liu et al., “Activation of mammalian target of rapamycin in transformed B lymphocytes is nutrient dependent but independent of Akt, mitogen-activated protein kinase/extracellular signal-regulated kinase kinase, insulin growth factor-1, and serum,” Cancer Research, vol. 65, no. 17, pp. 7800–7808, 2005.
[47]  S. H. Sin, D. Roy, L. Wang et al., “Rapamycin is efficacious against primary effusion lymphoma (PEL) cell lines in vivo by inhibiting autocrine signaling,” Blood, vol. 109, no. 5, pp. 2165–2173, 2007.
[48]  A. P. Bhatt, P. M. Bhende, S. H. Sin, D. Roy, D. P. Dittmer, and B. Damania, “Dual inhibition of PI3K and mTOR inhibits autocrine and paracrine proliferative loops in PI3K/Akt/mTOR-addicted lymphomas,” Blood, vol. 115, no. 22, pp. 4455–4463, 2010.
[49]  J. D. Col, P. Zancai, L. Terrin et al., “Distinct functional significance of Akt and mTOR constitutive activation in mantle cell lymphoma,” Blood, vol. 111, no. 10, pp. 5142–5151, 2008.
[50]  X. Lu, H. Nechushtan, F. Ding et al., “Distinct IL-4-induced gene expression, proliferation, and intracellular signaling in germinal center B-cell-like and activated B-cell-like diffuse large-cell lymphomas,” Blood, vol. 105, no. 7, pp. 2924–2932, 2005.
[51]  C. Renné, K. Willenbrock, J. I. Martin-Subero et al., “High expression of several tyrosine kinases and activation of the PI3K/AKT pathway in mediastinal large B cell lymphoma reveals further similarities to Hodgkin lymphoma,” Leukemia, vol. 21, no. 4, pp. 780–787, 2007.
[52]  S. Tauzin, H. Ding, D. Burdevet, B. Borisch, and D. C. Hoessli, “Membrane-associated signaling in human B-lymphoma lines,” Experimental Cell Research, vol. 317, pp. 151–162, 2011.
[53]  T. Portis and R. Longnecker, “Epstein-Barr virus (EBV) LMP2A mediates B-lymphocyte survival through constitutive activation of the Ras/PI3K/Akt pathway,” Oncogene, vol. 23, no. 53, pp. 8619–8628, 2004.
[54]  R. Swart, I. K. Ruf, J. Sample, and R. Longnecker, “Latent membrane protein 2A-mediated effects on the phosphatidylinositol 3-kinase/Akt pathway,” Journal of Virology, vol. 74, no. 22, pp. 10838–10845, 2000.
[55]  C. C. Tomlinson and B. Damania, “The K1 protein of Kaposi's sarcoma-associated herpesvirus activates the akt signaling pathway,” Journal of Virology, vol. 78, no. 4, pp. 1918–1927, 2004.
[56]  P. G. Smith, F. Wang, K. N. Wilkinson et al., “The phosphodiesterase PDE4B limits cAMP-associated PI3K/AKT-dependent apoptosis in diffuse large B-cell lymphoma,” Blood, vol. 105, no. 1, pp. 308–316, 2005.
[57]  M. A. Shipp, K. N. Ross, P. Tamayo et al., “Diffuse large B-cell lymphoma outcome prediction by gene-expression profiling and supervised machine learning,” Nature Medicine, vol. 8, no. 1, pp. 68–74, 2002.
[58]  K. K. Hoyer, S. W. French, D. E. Turner et al., “Dysregulated TCL1 promotes multiple classes of mature B cell lymphoma,” Proceedings of the National Academy of Sciences of the United States of America, vol. 99, no. 22, pp. 14392–14397, 2002.
[59]  G. Lenz, G. W. Wright, N. C. T. Emre et al., “Molecular subtypes of diffuse large B-cell lymphoma arise by distinct genetic pathways,” Proceedings of the National Academy of Sciences of the United States of America, vol. 105, no. 36, pp. 13520–13525, 2008.
[60]  A. Navarro, S. Beà, V. Fernández et al., “MicroRNA expression, chromosomal alterations, and immunoglobulin variable heavy chain hypermutations in mantle cell lymphomas,” Cancer Research, vol. 69, no. 17, pp. 7071–7078, 2009.
[61]  A. V. Miletic, A. N. Anzelon-Mills, D. M. Mills, et al., “Coordinate suppression of B cell lymphoma by PTEN and SHIP phosphatases,” The Journal of Experimental Medicine, vol. 207, pp. 2407–2420, 2010.
[62]  D. Iwakiri and K. Takada, “Phosphatidylinositol 3-kinase is a determinant of responsiveness to B cell antigen receptor-mediated epstein-barr virus activation,” Journal of Immunology, vol. 172, no. 3, pp. 1561–1566, 2004.
[63]  L. C. Platanias, “Map kinase signaling pathways and hematologic malignancies,” Blood, vol. 101, no. 12, pp. 4667–4679, 2003.
[64]  K. S. J. Elenitoba-Johnson, S. D. Jenson, R. T. Abbott et al., “Involvement of multiple signaling pathways in follicular lymphoma transformation: p38-mitogen-activated protein kinase as a target for therapy,” Proceedings of the National Academy of Sciences of the United States of America, vol. 100, no. 12, pp. 7259–7264, 2003.
[65]  A. Rinaldi, I. Kwee, M. Taborelli et al., “Genomic and expression profiling identifies the B-cell associated tyrosine kinase Syk as a possible therapeutic target in mantle cell lymphoma,” British Journal of Haematology, vol. 132, no. 3, pp. 303–316, 2006.
[66]  X. Huang, S. Wullschleger, M. Shpiro et al., “Important role of the LKB1-AMPK pathway in suppressing tumorigenesis in PTEN-deficient mice,” Biochemical Journal, vol. 412, no. 2, pp. 211–221, 2008.
[67]  K. J. Mavrakis, H. Zhu, R. L. A. Silva et al., “Tumorigenic activity and therapeutic inhibition of Rheb GTPase,” Genes and Development, vol. 22, no. 16, pp. 2178–2188, 2008.
[68]  D. D. Sarbassov, S. M. Ali, S. Sengupta et al., “Prolonged rapamycin treatment inhibits mTORC2 assembly and Akt/PKB,” Molecular Cell, vol. 22, no. 2, pp. 159–168, 2006.
[69]  D. A. Foster and A. Toschi, “Targeting mTOR with rapamycin: one dose does not fit all,” Cell Cycle, vol. 8, no. 7, pp. 1026–1029, 2009.
[70]  M. Gupta, S. M. Ansell, A. J. Novak, S. Kumar, S. H. Kaufmann, and T. E. Witzig, “Inhibition of histone deacetylase overcomes rapamycin-mediated resistance in diffuse large B-cell lymphoma by inhibiting Akt signaling through mTORC2,” Blood, vol. 114, no. 14, pp. 2926–2935, 2009.
[71]  K. Wanner, S. Hipp, M. Oelsner et al., “Mammalian target of rapamycin inhibition induces cell cycle arrest in diffuse large B cell lymphoma (DLBCL) cells and sensitises DLBCL cells to rituximab,” British Journal of Haematology, vol. 134, no. 5, pp. 475–484, 2006.
[72]  S. Hipp, I. Ringshausen, M. Oelsner, C. Bogner, C. Peschel, and T. Decker, “Inhibition of the mammalian target of rapamycin and the induction of cell cycle arrest in mantle cell lymphoma cells,” Haematologica, vol. 90, no. 10, pp. 1433–1434, 2005.
[73]  T. Peng, T. R. Golub, and D. M. Sabatini, “The immunosuppressant rapamycin mimics a starvation-like signal distinct from amino acid and glucose deprivation,” Molecular and Cellular Biology, vol. 22, no. 15, pp. 5575–5584, 2002.
[74]  J. LoPiccolo, G. M. Blumenthal, W. B. Bernstein, and P. A. Dennis, “Targeting the PI3K/Akt/mTOR pathway: effective combinations and clinical considerations,” Drug Resistance Updates, vol. 11, no. 1-2, pp. 32–50, 2008.
[75]  L. Bonapace, B. C. Bornhauser, M. Schmitz et al., “Induction of autophagy-dependent necroptosis is required for childhood acute lymphoblastic leukemia cells to overcome glucocorticoid resistance,” Journal of Clinical Investigation, vol. 120, no. 4, pp. 1310–1323, 2010.
[76]  H. Takeuchi, Y. Kondo, K. Fujiwara et al., “Synergistic augmentation of rapamycin-induced autophagy in malignant glioma cells by phosphatidylinositol 3-kinase/protein kinase B inhibitors,” Cancer Research, vol. 65, no. 8, pp. 3336–3346, 2005.
[77]  D. T. Teachey, S. A. Grupp, and V. I. Brown, “Mammalian target of rapamycin inhibitors and their potential role in therapy in leukaemia and other haematological malignancies,” British Journal of Haematology, vol. 145, no. 5, pp. 569–580, 2009.
[78]  R. Yuan, A. Kay, W. J. Berg, and D. Lebwohl, “Targeting tumorigenesis: development and use of mTOR inhibitors in cancer therapy,” Journal of Hematology and Oncology, vol. 2, article 45, 2009.
[79]  J. Tabernero, F. Rojo, E. Calvo et al., “Dose- and schedule-dependent inhibition of the mammalian target of rapamycin pathway with everolimus: a phase I tumor pharmacodynamic study in patients with advanced solid tumors,” Journal of Clinical Oncology, vol. 26, no. 10, pp. 1603–1610, 2008.
[80]  Z. Zeng, D. D. Sarbassov, I. J. Samudio et al., “Rapamycin derivatives reduce mTORC2 signaling and inhibit AKT activation in AML,” Blood, vol. 109, no. 8, pp. 3509–3512, 2007.
[81]  L. Wang, W. Y. Shi, Z. Y. Wu, et al., “Cytostatic and anti-angiogenic effects of temsirolimus in refractory mantle cell lymphoma,” Journal of Hematology & Oncology, vol. 3, p. 30, 2010.
[82]  V. Y. Yazbeck, D. Buglio, G. V. Georgakis et al., “Temsirolimus downregulates p21 without altering cyclin D1 expression and induces autophagy and synergizes with vorinostat in mantle cell lymphoma,” Experimental Hematology, vol. 36, no. 4, pp. 443–450, 2008.
[83]  M. Hidalgo, J. C. Buckner, C. Erlichman et al., “A phase I and pharmacokinetic study of temsirolimus (CCI-779) administered intravenously daily for 5 days every 2 weeks to patients with advanced cancer,” Clinical Cancer Research, vol. 12, no. 19, pp. 5755–5763, 2006.
[84]  E. Raymond, J. Alexandre, S. Faivre et al., “Safety and pharmacokinetics of escalated doses of weekly intravenous infusion of CCI-779, a novel mTOR inhibitor, in patients with cancer,” Journal of Clinical Oncology, vol. 22, no. 12, pp. 2336–2347, 2004.
[85]  A. O'Donnell, S. Faivre, H. A. Burris III et al., “Phase I pharmacokinetic and pharmacodynamic study of the oral mammalian target of rapamycin inhibitor everolimus in patients with advanced solid tumors,” Journal of Clinical Oncology, vol. 26, no. 10, pp. 1588–1595, 2008.
[86]  C. Tanaka, T. O'Reilly, J. M. Kovarik et al., “Identifying optimal biologic doses of everolimus (RAD001) in patients with cancer based on the modeling of preclinical and clinical pharmacokinetic and pharmacodynamic data,” Journal of Clinical Oncology, vol. 26, no. 10, pp. 1596–1602, 2008.
[87]  M. M. Mita, A. C. Mita, Q. S. Chu et al., “Phase I trial of the novel mammalian target of rapamycin inhibitor deforolimus (AP23573; MK-8669) administered intravenously daily for 5 days every 2 weeks to patients with advanced malignancies,” Journal of Clinical Oncology, vol. 26, no. 3, pp. 361–367, 2008.
[88]  G. J. Fetterly, M. M. Mita, C. D. Britten, et al., “Tolcher AW Pharmacokinetics of oral deforolimus (AP23573, MK-8669),” Journal of Clinical Oncology, vol. 26, Abstract 3509, 2008.
[89]  E. Drakos, G. Z. Rassidakis, and L. J. Medeiros, “Mammalian target of rapamycin (mTOR) pathway signalling in lymphomas,” Expert Reviews in Molecular Medicine, vol. 10, no. 4, article e4, pp. 1–21, 2008.
[90]  E. S. Jaffe, “The 2008 WHO classification of lymphomas: implications for clinical practice and translational research,” Hematology, pp. 523–531, 2009.
[91]  J. E. Romaguera, “Mantle cell lymphoma: frontline and salvage therapy,” Current Hematologic Malignancy Reports, vol. 3, no. 4, pp. 204–209, 2008.
[92]  T. Haritunians, A. Mori, J. O'Kelly, Q. T. Luong, F. J. Giles, and H. P. Koeffler, “Antiproliferative activity of RAD001 (everolimus) as a single agent and combined with other agents in mantle cell lymphoma,” Leukemia, vol. 21, no. 2, pp. 333–339, 2007.
[93]  K. W. L. Yee, Z. Zeng, M. Konopleva et al., “Phase I/II study of the mammalian target of rapamycin inhibitor everolimus (RAD001) in patients with relapsed or refractory hematologic malignancies,” Clinical Cancer Research, vol. 12, no. 17, pp. 5165–5173, 2006.
[94]  T. E. Witzig, S. M. Geyer, I. Ghobrial et al., “Phase II trial of single-agent temsirolimus (CCI-779) for relapsed mantle cell lymphoma,” Journal of Clinical Oncology, vol. 23, no. 23, pp. 5347–5356, 2005.
[95]  S. M. Ansell, D. J. Inwards, K. M. Rowland et al., “Low-dose, single-agent temsirolimus for relapsed mantle cell lymphoma: a phase 2 trial in the North Central Cancer Treatment Group,” Cancer, vol. 113, no. 3, pp. 508–514, 2008.
[96]  G. Hess, R. Herbrecht, J. Romaguera et al., “Phase III study to evaluate temsirolimus compared with investigator's choice therapy for the treatment of relapsed or refractory mantle cell lymphoma,” Journal of Clinical Oncology, vol. 27, no. 23, pp. 3822–3829, 2009.
[97]  R. Yuan, A. Kay, W. J. Berg, and D. Lebwohl, “Targeting tumorigenesis: development and use of mTOR inhibitors in cancer therapy,” Journal of Hematology and Oncology, vol. 2, article 45, 2009.
[98]  S. M. Ansell, H. Tang, P. J. Kurtin, et al., “Temsirolimus and rituximab in patients with relapsed or refractory mantle cell lymphoma: a phase 2 study,” The Lancet Oncology, vol. 12, pp. 361–368, 2011.
[99]  T. E. Witzig, C. B. Reeder, B. R. Laplant, et al., “A phase II trial of the oral mTOR inhibitor everolimus in relapsed aggressive lymphoma,” Leukemia, vol. 25, pp. 341–347, 2011.
[100]  D. A. Rizzieri, E. Feldman, J. F. Dipersio et al., “A phase 2 clinical trial of deforolimus (AP23573, MK-8669), a novel mammalian target of rapamycin inhibitor, in patients with relapsed or refractory hematologic malignancies,” Clinical Cancer Research, vol. 14, no. 9, pp. 2756–2762, 2008.
[101]  J. O. Armitage and D. D. Weisenburger, “New approach to classifying non-Hodgkin's lymphomas: clinical features of the major histologic subtypes. Non-Hodgkin's Lymphoma Classification Project,” Journal of Clinical Oncology, vol. 16, pp. 2780–2795, 1998.
[102]  S. Uddin, A. R. Hussain, A. K. Siraj et al., “Role of phosphatidylinositol 3′-kinase/AKT pathway in diffuse large B-cell lymphoma survival,” Blood, vol. 108, no. 13, pp. 4178–4186, 2006.
[103]  S. M. Smith, K. van Besien, T. Karrison, et al., “Temsirolimus has activity in non-mantle cell non-Hodgkin's lymphoma subtypes: the University of Chicago phase II consortium,” Journal of Clinical Oncology, vol. 28, pp. 4740–4746, 2010.
[104]  D. C. Fingar, S. Salama, C. Tsou, ED. Harlow, and J. Blenis, “Mammalian cell size is controlled by mTOR and its downstream targets S6K1 and 4EBP1/eIF4E,” Genes and Development, vol. 16, no. 12, pp. 1472–1487, 2002.
[105]  J. Montagne, M. J. Stewart, H. Stocker, E. Hafen, S. C. Kozma, and G. Thomas, “Drosophila S6 kinase: a regulator of cell size,” Science, vol. 285, no. 5436, pp. 2126–2129, 1999.
[106]  H. Stocker and E. Hafen, “Genetic control of cell size,” Current Opinion in Genetics and Development, vol. 10, no. 5, pp. 529–535, 2000.
[107]  T. V. Isaacson, “Expression of mTOR pathway proteins in malignant lymphoma,” Laboratory Investigation, vol. 246A, 2007.
[108]  G. V. Georgakis, Y. Li, G. Z. Rassidakis, L. J. Medeiros, G. B. Mills, and A. Younes, “Inhibition of the phosphatidylinositol-3 kinase/Akt promotes G1 cell cycle arrest and apoptosis in Hodgkin lymphoma,” British Journal of Haematology, vol. 132, no. 4, pp. 503–511, 2006.
[109]  F. Jundt, N. Raetzel, C. Muller, et al., “A rapamycin derivative (everolimus) controls proliferation through down-regulation of truncated CCAAT enhancer binding protein {beta} and NF-{kappa}B activity in Hodgkin and anaplastic large cell lymphomas,” Blood, vol. 106, pp. 1801–1817, 2005.
[110]  P. B. Johnston, D. J. Inwards, J. P. Colgan et al., “A Phase II trial of the oral mTOR inhibitor everolimus in relapsed Hodgkin lymphoma,” American Journal of Hematology, vol. 85, no. 5, pp. 320–324, 2010.
[111]  M. R. Janes and D. A. Fruman, “Targeting TOR dependence in cancer,” Oncotarget, vol. 1, pp. 69–76, 2010.
[112]  S. M. Maira, F. Stauffer, J. Brueggen et al., “Identification and characterization of NVP-BEZ235, a new orally available dual phosphatidylinositol 3-kinase/mammalian target of rapamycin inhibitor with potent in vivo antitumor activity,” Molecular Cancer Therapeutics, vol. 7, no. 7, pp. 1851–1863, 2008.
[113]  V. Serra, B. Markman, M. Scaltriti et al., “NVP-BEZ235, a dual PI3K/mTOR inhibitor, prevents PI3K signaling and inhibits the growth of cancer cells with activating PI3K mutations,” Cancer Research, vol. 68, no. 19, pp. 8022–8030, 2008.
[114]  K. Kojima, M. Shimanuki, M. Shikami et al., “The dual PI3 kinase/mTOR inhibitor PI-103 prevents p53 induction by Mdm2 inhibition but enhances p53-mediated mitochondrial apoptosis in p53 wild-type AML,” Leukemia, vol. 22, no. 9, pp. 1728–1736, 2008.
[115]  M. E. Feldman, B. Apsel, A. Uotila et al., “Active-site inhibitors of mTOR target rapamycin-resistant outputs of mTORC1 and mTORC2,” PLoS Biology, vol. 7, no. 2, article e38, pp. 0371–0383, 2009.
[116]  P. M. Bhende, S. I. Park, M. S. Lim, D. P. Dittmer, and B. Damania, “The dual PI3K/mTOR inhibitor, NVP-BEZ235, is efficacious against follicular lymphoma,” Leukemia, vol. 24, pp. 1781–1784, 2010.
[117]  J. M. García-Martínez, J. Moran, R. G. Clarke et al., “Ku-0063794 is a specific inhibitor of the mammalian target of rapamycin (mTOR),” Biochemical Journal, vol. 421, no. 1, pp. 29–42, 2009.
[118]  C. C. Thoreen, S. A. Kang, J. W. Chang et al., “An ATP-competitive mammalian target of rapamycin inhibitor reveals rapamycin-resistant functions of mTORC1,” Journal of Biological Chemistry, vol. 284, no. 12, pp. 8023–8032, 2009.
[119]  K. Yu, L. Toral-Barza, C. Shi et al., “Biochemical, cellular, and in vivo activity of novel ATP-competitive and selective inhibitors of the mammalian target of rapamycin,” Cancer Research, vol. 69, no. 15, pp. 6232–6240, 2009.
[120]  M. R. Janes, J. J. Limon, L. So et al., “Effective and selective targeting of leukemia cells using a TORC1/2 kinase inhibitor,” Nature Medicine, vol. 16, no. 2, pp. 205–213, 2010.
[121]  C. M. Chresta, B. R. Davies, I. Hickson, et al., “AZD-8055 is a potent, selective and orally bioavailable ATP-competitive mammalian target of rapamycin kinase inhibitor with in vitro and in vivo antitumor activity,” Cancer Research, vol. 70, pp. 288–298, 2009.
[122]  K. Yu, C. Shi, L. Toral-Barza et al., “Beyond rapalog therapy: preclinical pharmacology and antitumor activity of WYE-125132, an ATP-competitive and specific inhibitor of mTORC1 and mTORC2,” Cancer Research, vol. 70, no. 2, pp. 621–631, 2010.
[123]  R. J. Dowling, I. Topisirovic, T. Alain, et al., “mTORC1-mediated cell proliferation, but not cell growth, controlled by the 4E-BPs,” Science, vol. 328, pp. 1172–1176, 2010.
[124]  G. J. Brunn, J. Williams, C. Sabers, G. Wiederrecht, J. C. Lawrence Jr., and R. T. Abraham, “Direct inhibition of the signaling functions of the mammalian target of rapamycin by the phosphoinositide 3-kinase inhibitors, wortmannin and LY294002,” EMBO Journal, vol. 15, no. 19, pp. 5256–5267, 1996.
[125]  F. Chiarini, C. Grimaldi, F. Ricci, et al., “Activity of the novel dual phosphatidylinositol 3-kinase/mammalian target of rapamycin inhibitor NVP-BEZ235 against T-cell acute lymphoblastic leukemia,” Cancer Research, vol. 70, pp. 8097–8107, 2010.
[126]  T. C. Zhang, H. J. Chu, J. Q. Zhao, X. Y. Zhou, and D. R. Shi, “Study on PI3K inhibitor LY294002 for chemotherapeutic sensitization in diffuse large B cell lymphoma cell lines,” Zhonghua Xue Ye Xue Za Zhi, vol. 31, pp. 671–674, 2010.
[127]  A. C. Hsieh, M. Costa, O. Zollo et al., “Genetic dissection of the oncogenic mTOR pathway reveals druggable addiction to translational control via 4EBP-eIF4E,” Cancer Cell, vol. 17, no. 3, pp. 249–261, 2010.

Full-Text

comments powered by Disqus

Contact Us

service@oalib.com

QQ:3279437679

WhatsApp +8615387084133

WeChat 1538708413