用户名: 密码: 验证码:
分泌型热休克蛋白90α在肿瘤发生和转移中的作用机理
详细信息    本馆镜像全文|  推荐本文 |  |   获取CNKI官网全文
摘要
热休克蛋白90α(Hsp90α)是一个伴侣蛋白,能够辅助蛋白折叠和维持细胞内多种信号传导蛋白的稳定,从而促进细胞存活和生长。在肿瘤细胞中,Hsp90α能够使过度激活或突变的信号传导蛋白保持活性,加速了肿瘤细胞的恶性转变。目前,Hsp90α已成为重要的抗肿瘤治疗靶点。
     除胞内形式以外,Hsp90α还能被分泌到胞外。分泌型Hsp90α在肿瘤细胞、神经细胞、皮肤上皮细胞和免疫细胞中均被发现,但其功能和作用机理仍不清楚。本工作首先检测了分泌型Hsp90α在恶性程度不同的乳腺癌细胞系及正常乳腺上皮细胞系培养基中的含量,发现Hsp90α的分泌水平随肿瘤细胞侵袭程度的增加而升高,表明分泌型Hsp90α与肿瘤的侵袭密切相关。接下来,本工作克隆了人Hsp90α的基因,制备了人Hsp90α的重组蛋白及抗体;体外实验证明,分泌型Hsp90α能够促进肿瘤细胞侵袭,而对增殖没有明显作用。进一步的,在黑色素瘤转移模型中,Hsp90α小鼠单抗能够明显抑制肿瘤的肺转移,更为显著的是,肿瘤的淋巴结转移被完全抑制。更进一步的,本工作证明Hsp90α能够在血浆中被检测到,肿瘤病人血浆中Hsp90α的含量与肿瘤的恶性程度,尤其是转移正相关,有希望成为肿瘤诊断和预后的标志物。
     同时,本工作探讨了分泌型Hsp90α促进肿瘤侵袭的机理。已有报道称,分泌型Hsp90α与基质金属蛋白酶-2(MMP-2)相互作用,Hsp90α的抑制剂能够降低MMP-2的活性,进而抑制肿瘤细胞侵袭;但对分泌型Hsp90α调节MMP-2活性的分子机制还没有解释。本研究证明分泌型Hsp90α能够通过与MMP-2的Hemopexin结构域相互作用,阻止MMP-2在Glu~(443)-Leu~(444)位点被酶切,进而阻止MMP-2降解失活。进一步的,我们还发现以上机制也存在于激活的内皮细胞中,分泌型Hsp90α能够促进肿瘤的新生血管生成。以上发现对分泌型Hsp90α的功能首次给予了分子机制上的解释,为分泌型Hsp90α新的相互作用蛋白的发现和机理研究提供了思路。
Heat shock protein 90-α(Hsp90α) is an intracellular molecular chaperone. It functions to maintain the normal protein folding environment and promote the cell survival under stress conditions. In tumor cells, the chaperoning activity of Hsp90αis subverted to retain the functions of mutant proteins and thus facilitate the malignant transformation. Meanwhile, Hsp90αcan also be secreted by tumor cells, but the functions of this extracellular form remains far from clear. Therefore, my study focuses on the functions of secreted Hsp90αon tumor cells and the underlying mechanisms. We demonstrate that secreted Hsp90αcan promote tumor invasiveness and is highly correlated with tumor malignancy. The blockage of secreted Hsp90αresults in significant inhibition on tumor invasion and complete prevention of lymph node metastasis. In addition, the level of plasma Hsp90αis positively correlated with tumor malignancy in cancer patients, indicating it is a potential diagnostic marker for tumor malignancy in clinical application.
     Since we found that the secreted Hsp90αis highly correlated with tumor malignancy, we next investigated the mechanism behind this phenomenon. It is reported that secreted Hsp90αinteracts with MMP-2 and the inhibition of extracellular Hsp90αreduces the activity of MMP-2. In the present study, by the means of over-expressing, knocking down and also antibody blockage, we demonstrate that the pro-invasiveness activity of extracellular Hsp90αis dependent on MMP-2. But the regulatory mechanism of Hsp90αon the activity of MMP-2 is still unknown. We find that Hsp90αcan stabilize MMP-2 and prevent the proteolytic processing of MMP-2 via directly binding to the hemopexin domain. We further demonstrated that Hsp90αcan be secreted by endothelial cells and is a positive regulator of angiogenesis. Moreover, the antibody of Hsp90αpromotes MMP-2 processing and thus retards the angiogenesis and tumor invasiveness in vitro and in vivo.
     In sum, here we reveal that secreted Hsp90αfunctions as a pro-invasiveness factor extracellularly and its level in plasma is positively correlated with the tumor malignancy, especially tumor metastasis. Meanwhile, we demonstrate that extracellular Hsp90αstabilizes MMP-2 and prevent its proteolytic processing via binding to the hemopexin domain, providing novel mechanistic explanations for both the function of extracellular Hsp90αand the regulatory mechanism of MMP-2 activity.
引文
[1] Vogelstein B, Kinzler KW. The genetic basis of human cancer. New York: McGraw-Hill, Medical Pub. Division, 2002.
    [2] Weinstein IB. Cancer. Addiction to oncogenes--the Achilles heal of cancer. Science 2002;297:63-4.
    [3] Croce CM. Oncogenes and cancer. N Engl J Med 2008;358:502-11.
    [4] Luo J, Solimini NL, Elledge SJ. Principles of cancer therapy: oncogene and non-oncogene addiction. Cell 2009;136:823-37.
    [5] Felsher DW, Bishop JM. Reversible tumorigenesis by MYC in hematopoietic lineages. Mol Cell 1999;4:199-207.
    [6] Jain M, Arvanitis C, Chu K, et al. Sustained loss of a neoplastic phenotype by brief inactivation of MYC. Science 2002;297:102-4.
    [7] Pelengaris S, Littlewood T, Khan M, et al. Reversible activation of c-Myc in skin: induction of a complex neoplastic phenotype by a single oncogenic lesion. Mol Cell 1999;3:565-77.
    [8] Chin L, Tam A, Pomerantz J, et al. Essential role for oncogenic Ras in tumour maintenance. Nature 1999;400:468-72.
    [9] Huettner CS, Zhang P, Van Etten RA, et al. Reversibility of acute B-cell leukaemia induced by BCR-ABL1. Nat Genet 2000;24:57-60.
    [10] Shirasawa S, Furuse M, Yokoyama N, et al. Altered growth of human colon cancer cell lines disrupted at activated Ki-ras. Science 1993;260:85-8.
    [11] Martins CP, Brown-Swigart L, Evan GI. Modeling the therapeutic efficacy of p53 restoration in tumors. Cell 2006;127:1323-34.
    [12] Ventura A, Kirsch DG, McLaughlin ME, et al. Restoration of p53 function leads to tumour regression in vivo. Nature 2007;445:661-5.
    [13] Xue W, Zender L, Miething C, et al. Senescence and tumour clearance is triggered by p53 restoration in murine liver carcinomas. Nature 2007;445:656-60.
    [14] Albertson DG, Pinkel D. Genomic microarrays in human genetic disease and cancer. Hum Mol Genet 2003;12 Spec No 2:R145-52.
    [15] Solimini NL, Luo J, Elledge SJ. Non-oncogene addiction and the stress phenotype of cancer cells. Cell 2007;130:986-8.
    [16] Hartwell LH, Kastan MB. Cell cycle control and cancer. Science 1994;266:1821-8.
    [17] Maser RS, DePinho RA. Connecting chromosomes, crisis, and cancer. Science 2002;297:565-9.
    [18] Windle B, Draper BW, Yin YX, et al. A central role for chromosome breakage in gene amplification, deletion formation, and amplicon integration. Genes Dev 1991;5:160-74.
    [19] Halazonetis TD, Gorgoulis VG, Bartek J. An oncogene-induced DNA damage model for cancer development. Science 2008;319:1352-5.
    [20] Harper JW, Elledge SJ. The DNA damage response: ten years after. Mol Cell 2007;28:739-45.
    [21] Bartkova J, Horejsi Z, Koed K, et al. DNA damage response as a candidate anti-cancer barrier in early human tumorigenesis. Nature 2005;434:864-70.
    [22] Gorgoulis VG, Vassiliou LV, Karakaidos P, et al. Activation of the DNA damage checkpoint and genomic instability in human precancerous lesions. Nature 2005;434:907-13.
    [23] Chen Z, Xiao Z, Gu WZ, et al. Selective Chk1 inhibitors differentially sensitize p53-deficient cancer cells to cancer therapeutics. Int J Cancer 2006;119:2784-94.
    [24] Kennedy RD, Chen CC, Stuckert P, et al. Fanconi anemia pathway-deficient tumor cells are hypersensitive to inhibition of ataxia telangiectasia mutated. J Clin Invest 2007;117:1440-9.
    [25] Komarova NL, Lengauer C, Vogelstein B, et al. Dynamics of genetic instability in sporadic and familial colorectal cancer. Cancer Biol Ther 2002;1:685-92.
    [26] Cahill DP, Lengauer C, Yu J, Riggins GJ, et al. Mutations of mitotic checkpoint genes in human cancers. Nature 1998;392:300-3.
    [27] Ganem NJ, Storchova Z, Pellman D. Tetraploidy, aneuploidy and cancer. Curr Opin Genet Dev 2007;17:157-62.
    [28] Whitesell L, Lindquist SL. HSP90 and the chaperoning of cancer. Nat Rev Cancer 2005;5:761-72.
    [29] Torres EM, Sokolsky T, Tucker CM, et al. Effects of aneuploidy on cellular physiology and cell division in haploid yeast. Science 2007;317:916-24.
    [30] Dai C, Whitesell L, Rogers AB, et al. Heat shock factor 1 is a powerful multifaceted modifier of carcinogenesis. Cell 2007;130:1005-18.
    [31] Richardson PG, Mitsiades C, Hideshima T, et al. Bortezomib: proteasome inhibition as an effective anticancer therapy. Annu Rev Med 2006;57:33-47.
    [32] Warburg O. On the origin of cancer cells. Science 1956;123:309-14.
    [33] DeBerardinis RJ, Lum JJ, Hatzivassiliou G, et al. The biology of cancer: metabolic reprogramming fuels cell growth and proliferation. Cell Metab 2008;7:11-20.
    [34] Kroemer G, Pouyssegur J. Tumor cell metabolism: cancer's Achilles' heel. Cancer Cell 2008;13:472-82.
    [35] Vander Heiden MG, Plas DR, Rathmell JC, et al. Growth factors can influence cell growth and survival through effects on glucose metabolism. Mol Cell Biol 2001;21:5899-912.
    [36] Szatrowski TP, Nathan CF. Production of large amounts of hydrogen peroxide by human tumor cells. Cancer Res 1991;51:794-8.
    [37] Lee AC, Fenster BE, Ito H, et al. Ras proteins induce senescence by altering the intracellular levels of reactive oxygen species. J Biol Chem 1999;274:7936-40.
    [38] Gogvadze V, Orrenius S, Zhivotovsky B. Mitochondria in cancer cells: what is so special about them? Trends Cell Biol 2008;18:165-73.
    [39] Dewhirst MW, Cao Y, Moeller B. Cycling hypoxia and free radicals regulate angiogenesis and radiotherapy response. Nat Rev Cancer 2008;8:425-37.
    [40] Bonnet S, Archer SL, Allalunis-Turner J, et al. A mitochondria-K+ channel axis is suppressed in cancer and its normalization promotes apoptosis and inhibits cancer growth. Cancer Cell 2007;11:37-51.
    [41] Pouyssegur J, Dayan F, Mazure NM. Hypoxia signalling in cancer and approaches to enforce tumour regression. Nature 2006;441:437-43.
    [42] Jin S, White E. Role of autophagy in cancer: management of metabolic stress. Autophagy 2007;3:28-31.
    [43] Muller AJ, Scherle PA. Targeting the mechanisms of tumoral immune tolerance with small-molecule inhibitors. Nat Rev Cancer 2006;6:613-25.
    [44] Zou W. Immunosuppressive networks in the tumour environment and their therapeutic relevance. Nat Rev Cancer 2005;5:263-74.
    [45] Garrido F, Algarra I. MHC antigens and tumor escape from immune surveillance. Adv Cancer Res 2001;83:117-58.
    [46] Sharma S, Yang SC, Zhu L, et al. Tumor cyclooxygenase-2/prostaglandin E2-dependent promotion of FOXP3 expression and CD4+ CD25+ T regulatory cell activities in lung cancer. Cancer Res 2005;65:5211-20.
    [47] Stolina M, Sharma S, Lin Y, et al. Specific inhibition of cyclooxygenase 2 restores antitumor reactivity by altering the balance of IL-10 and IL-12 synthesis. J Immunol 2000;164:361-70.
    [48] Peggs KS, Quezada SA, Allison JP. Cell intrinsic mechanisms of T-cell inhibition and application to cancer therapy. Immunol Rev 2008;224:141-65.
    [49] Lindquist S, Craig EA. The heat-shock proteins. Annu Rev Genet 1988;22:631-77.
    [50] Bukau B, Horwich AL. The Hsp70 and Hsp60 chaperone machines. Cell 1998;92:351-66.
    [51] Easton DP, Kaneko Y, Subjeck JR. The hsp110 and Grp1 70 stress proteins: newly recognized relatives of the Hsp70s. Cell Stress Chap 2000;5:276-90.
    [52] Buchner J. Hsp90 & Co. - a holding for folding. Trends Biochem Sci 1999;24:136-41.
    [53] Wegele H, Muller L, Buchner J. Hsp70 and Hsp90--a relay team for protein folding. Rev Physiol Biochem Pharmacol 2004;151:1-44.
    [54] Mayer MP, Bukau B. Hsp70 chaperones: cellular functions and molecular mechanism. Cell Mol Life Sci 2005;62:670-84.
    [55] Pratt WB, Toft DO. Regulation of signaling protein function and trafficking by the hsp90/hsp70-based chaperone machinery. Exp Biol Med (Maywood) 2003;228:111-33.
    [56] Spiess C, Meyer AS, Reissmann S, et al. Mechanism of the eukaryotic chaperonin: protein folding in the chamber of secrets. Trends Cell Biol 2004;14:598-604.
    [57] Young JC, Agashe VR, Siegers K, et al. Pathways of chaperone-mediated protein folding in the cytosol. Nat Rev Mol Cell Biol 2004;5:781-91.
    [58] Arrigo AP. Heat shock proteins as molecular chaperones. M S-Med Sci 2005;21:619-25.
    [59] Schlesinger MJ. How the Cell Copes with Stress and the Function of Heat-Shock Proteins. Pediatr Res 1994;36:1-6.
    [60] Hut HM, Kampinga HH, Sibon OC. Hsp70 protects mitotic cells against heat-induced centrosome damage and division abnormalities. Mol Biol Cell 2005;16:3776-85.
    [61] Beere HM. "The stress of dying": the role of heat shock proteins in the regulation of apoptosis. J Cell Sci 2004;117:2641-51.
    [62] Nylandsted J, Gyrd-Hansen M, Danielewicz A, et al. Heat shock protein 70 promotes cell survival by inhibiting lysosomal membrane permeabilization. J Exp Med 2004;200:425-35.
    [63] Calderwood SK. Regulatory interfaces between the stress protein response and other gene expression programs in the cell. Methods 2005;35:139-48.
    [64] Mosser DD, Morimoto RI. Molecular chaperones and the stress of oncogenesis. Oncogene 2004;23:2907-18.
    [65] Neckers L, Ivy SP. Heat shock protein 90. Curr Opin Oncol 2003;15:419-24.
    [66] Pratt WB, Galigniana MD, Harrell JM, et al. Role of hsp90 and the hsp90-binding immunophilins in signalling protein movement. Cell Signal 2004;16:857-72.
    [67] Neckers L. Hsp90 inhibitors as novel cancer chemotherapeutic agents. Trends Mol Med 2002;8:S55-S61.
    [68] Nimmanapalli R, O'Bryan E, Bhalla K. Geldanamycin and its analogue 17-allylamino-17-demethoxygeldanamycin lowers Bcr-Abl levels and induces apoptosis and differentiation of Bcr-Abl-positive human leukemic blasts. Cancer Res 2001;61:1799-804.
    [69] Rahmani M, Reese E, Dai Y, et al. Cotreatment with suberanoylanilide hydroxamic acid and 17-allylamino 17-demethoxygeldanamycin synergistically induces apoptosis in Bcr-Abl+ Cells sensitive and resistant to STI571 (imatinib mesylate) in association with down-regulation of Bcr-Abl, abrogation of signal transducer and activator of transcription 5 activity, and Bax conformational change. Mol Pharmacol 2005;67:1166-76.
    [70] Calderwood SK, Khaleque MA, Sawyer DB, et al. Heat shock proteins in cancer: chaperones of tumorigenesis. Trends Biochem Sci 2006;31:164-72.
    [71] Gerner EW, Schneider MJ. Induced thermal resistance in HeLa cells. Nature 1975;256:500-2.
    [72] Beere HM. Stressed to death: regulation of apoptotic signaling pathways by the heat shock proteins. Sci STKE 2001;2001:re1.
    [73] Hsu AL, Murphy CT, Kenyon C. Regulation of aging and age-related disease by DAF-16 and heat-shock factor. Science 2003;300:1142-5.
    [74] Tatar M, Khazaeli AA, Curtsinger JW. Chaperoning extended life. Nature 1997;390:30-.
    [75] Nelson DA, White E. Exploiting different ways to die. Gene Dev 2004;18:1223-6.
    [76] Campisi J. Senescent cells, tumor suppression, and organismal aging: Good citizens, bad neighbors. Cell 2005;120:513-22.
    [77] Workman P. Altered states: selectively drugging the Hsp90 cancer chaperone. Trends Mol Med 2004;10:47-51.
    [78] Rohde M, Daugaard M, Jensen MH, et al. Members of the heat-shock protein 70 family promote cancer cell growth by distinct mechanisms. Gene Dev 2005;19:570-82.
    [79] Wadhwa R, Taira K, Kaul SC. An Hsp70 family chaperone, mortalin/mthsp70/PBP74/Grp75: what, when, and where? Cell Stress Chap 2002;7:309-16.
    [80] Folkman J. Role of angiogenesis in tumor growth and metastasis. Semin Oncol 2002;29:15-8.
    [81] Zhou J, Schmid T, Frank R, et al. PI3K/Akt is required for heat shock proteins to protect hypoxia-inducible factor 1alpha from pVHL-independent degradation. J Biol Chem 2004;279:13506-13.
    [82] Sun J, Liao JK. Induction of angiogenesis by heat shock protein 90 mediated by protein kinase Akt and endothelial nitric oxide synthase. Arterioscler Thromb Vasc Biol 2004;24:2238-44.
    [83] Pfosser A, Thalgott M, Buttner K, et al. Liposomal Hsp90 cDNA induces neovascularization via nitric oxide in chronic ischemia. Cardiovasc Res 2005;65:728-36.
    [84] Keezer SM, Ivie SE, Krutzsch HC, et al. Angiogenesis inhibitors target the endothelial cell cytoskeleton through altered regulation of heat shock protein 27 and cofilin. Cancer Res 2003;63:6405-12.
    [85] Rousseau S, Houle F, Landry J, et al. p38 MAP kinase activation by vascular endothelial growth factor mediates actin reorganization and cell migration in human endothelial cells. Oncogene 1997;15:2169-77.
    [86] Hoang AT, Huang JP, Rudra-Ganguly N, et al. A novel association between the human heat shock transcription factor 1 (HSF1) and prostate adenocarcinoma. Am J Pathol 2000;156:857-64.
    [87] Ciocca DR, Calderwood SK. Heat shock proteins in cancer: diagnostic, prognostic, predictive, and treatment implications. Cell Stress Chap 2005;10:86-103.
    [88] Eustace BK, Sakurai T, Stewart JK, et al. Functional proteomic screens reveal an essential extracellular role for hsp90 alpha in cancer cell invasiveness. Nat Cell Biol 2004;6:507-14.
    [89] Eustace BK, Jay DG. Extracellular roles for the molecular chaperone, hsp90. Cell Cycle 2004;3:1098-100.
    [90] Price JT, Quinn JMW, Sims NA, et al. The heat shock protein 90 inhibitor, 17-allylamino-17-demethoxygeldanamycin, enhances osteoclast formation and potentiates bone metastasis of a human breast cancer cell line. Cancer Res 2005;65:4929-38.
    [91] Calderwood SK, Theriault JR, Gong JL. Message in a bottle: Role of the 70-kDa heat shock protein family in anti-tumor immunity. Eur J Immunol 2005;35:2518-27.
    [92] Asea A, Kraeft SK, Kurt-Jones EA, et al. HSP70 stimulates cytokine production through a CD14-dependant pathway, demonstrating its dual role as a chaperone and cytokine. Nat Med 2000;6:435-42.
    [93] Viatour P, Merville MP, Bours V, et al. Phosphorylation of NF-kappaB and IkappaB proteins: implications in cancer and inflammation. Trends Biochem Sci 2005;30:43-52.
    [94] Daniels GA, Sanchez-Perez L, Diaz RM, et al. A simple method to cure established tumors by inflammatory killing of normal cells. Nat Biotechnol 2004;22:1125-32.
    [95] Bausero MA, Page DT, Osinaga E, et al. Surface expression of Hsp25 and Hsp72 differentially regulates tumor growth and metastasis. Tumor Biol 2004;25:243-51.
    [96] Farkas B, Hantschel M, Magyarlaki M, et al. Heat shock protein 70 membrane expression and melanoma-associated marker phenotype in primary and metastatic melanoma. Melanoma Res 2003;13:147-52.
    [97] Kaur J, Das SN, Srivastava A, et al. Cell surface expression of 70 kDa heat shock protein in human oral dysplasia and squamous cell carcinoma: correlation with clinicopathological features. Oral Oncol 1998;34:93-8.
    [98] Tsutsumi S, Neckers L. Extracellular heat shock protein 90: A role for a molecular chaperone in cell motility and cancer metastasis. Cancer Sci 2007;98:1536-9.
    [99] Altmeyer A, Maki RG, Feldweg AM, et al. Tumor-specific cell surface expression of the -KDEL containing, endoplasmic reticular heat shock protein gp96. Int J Cancer 1996;69:340-9.
    [100] Wang X, Song X, Zhuo W, et al. The regulatory mechanism of Hsp90alpha secretion and its function in tumor malignancy. Proc Natl Acad Sci U S A 2009;106:21288-93.
    [101] Multhoff G, Hightower LE. Cell surface expression of heat shock proteins and the immune response. Cell Stress Chap 1996;1:167-76.
    [102] Young RA, Elliott TJ. Stress Proteins, Infection, and Immune Surveillance. Cell 1989;59:5-8.
    [103] Young D, Roman E, Moreno C, et al. Molecular Chaperones and the Immune-Response. Philos T Roy Soc B 1993;339:363-8.
    [104] Srivastava PK, Maki RG. Stress-Induced Proteins in Immune-Response to Cancer. Curr Top Microbiol 1991;167:109-23.
    [105] Srivastava PK, Udono H. Heat-Shock Protein-Peptide Complexes in Cancer-Immunotherapy. Curr Opin Immunol 1994;6:728-32.
    [106] Srivastava PK. Heat-Shock Proteins in Immune-Response to Cancer - the 4th Paradigm. Experientia 1994;50:1054-60.
    [107] Udono H, Srivastava PK. Heat-Shock Protein-70 Associated Peptides Elicit Specific Cancer Immunity. J Exp Med 1993;178:1391-6.
    [108] Udono H, Srivastava PK. Comparison of Tumor-Specific Immunogenicities of Stress-Induced Proteins Gp96, Hsp90, and Hsp70. J Immunol 1994;152:5398-403.
    [109] Suto R, Srivastava PK. A Mechanism for the Specific Immunogenicity of Heat-Shock Protein-Chaperoned Peptides. Science 1995;269:1585-8.
    [110] Tamura Y. Immunotherapy of tumors with autologous tumor-derived heat shock protein preparations (vol 278, pg 117-20, 1997). Science 1999;283:1118-1119.
    [111] Li LH, Clark TD, Cowie CH, et al. Effects of Geldanamycin and Its Derivatives on Rna-Directed DNA-Polymerase and Infectivity of Rauscher Leukemia-Virus. Cancer Treat Rep 1977;61:815-24.
    [112] Price PJ, Suk WA, Skeen PC, et al. Geldanamycin Inhibition of 3-Methylcholanthrene-Induced Rat Embryo Cell Transformation. P Soc Exp Biol Med 1977;155:461-3.
    [113] Whitesell L, Shifrin SD, Schwab G, et al. Benzoquinonoid Ansamycins Possess Selective Tumoricidal Activity Unrelated to Src Kinase Inhibition. Cancer Res 1992;52:1721-8.
    [114] Uehara Y, Hori M, Takeuchi T, et al. Phenotypic Change from Transformed to Normal Induced by Benzoquinonoid Ansamycins Accompanies Inactivation of P60src in Rat-Kidney Cells Infected with Rous-Sarcoma Virus. Mol Cell Biol 1986;6:2198-206.
    [115] Stebbins CE, Russo AA, Schneider C, et al. Crystal structure of an Hsp90-geldanamycin complex: Targeting of a protein chaperone by an antitumor agent. Cell 1997;89:239-50.
    [116] Prodromou C, Roe SM, O'Brien R, et al. Identification and structural characterization of the ATP/ADP-binding site in the Hsp90 molecular chaperone. Cell 1997;90:65-75.
    [117] Whitesell L, Mimnaugh EG, De Costa B, et al. Inhibition of heat shock protein HSP90-pp60v-src heteroprotein complex formation by benzoquinone ansamycins: essential role for stress proteins in oncogenic transformation. Proc Natl Acad Sci U S A 1994;91:8324-8.
    [118] Scheibel T, Buchner J. The Hsp90 complex - A super-chaperone machine as a novel drug target. Biochem Pharmacol 1998;56:675-82.
    [119] Sharma SV, Agatsuma T, Nakano H. Targeting of the protein chaperone, HSP90, by the transformation suppressing agent, radicicol. Oncogene 1998;16:2639-45.
    [120] Kim JH, Hahn EW, Benjamin FJ. Treatment of superficial cancers by combination hyperthermia and radiation therapy. Clin Bull 1979;9:13-6.
    [121] Issels RD. Hyperthermia and thermochemotherapy. Cancer Treat Res 1993;67:143-60.
    [122] Dewhirst MW, Prosnitz L, Thrall D, et al. Hyperthermic treatment of malignant diseases: current status and a view toward the future. Semin Oncol 1997;24:616-25.
    [123] Crile G, Jr. The effects of heat and radiation on cancers implanted on the feet of mice. Cancer Res 1963;23:372-80.
    [124] Fiorentini G, Szasz A. Hyperthermia today: electric energy, a new opportunity in cancer treatment. J Cancer Res Ther 2006;2:41-6.
    [125] Pearl LH, Prodromou C. Structure and mechanism of the Hsp90 molecular chaperone machinery. Annu Rev Biochem 2006;75:271-94.
    [126] Wright L, Barril X, Dymock B, et al. Structure-activity relationships in purine-based inhibitor binding to HSP90 isoforms. Chem & Biol 2004;11:775-85.
    [127] Millson SH, Truman AW, Racz A, et al. Expressed as the sole Hsp90 of yeast, the alpha and beta isoforms of human Hsp90 differ with regard to their capacities for activation of certain client proteins, whereas only Hsp90 beta generates sensitivity to the Hsp90 inhibitor radicicol. Febs J 2007;274:4453-63.
    [128] Chadli A, Felts SJ, Toft DO. GCUNC45 is the first Hsp90 co-chaperone to show alpha/beta isoform specificity. J Biol Chem 2008;283:9509-12.
    [129] Grammatikakis N, Vultur A, Ramana CV, et al. The role of Hsp90N, a new member of the Hsp90 family, in signal transduction and neoplastic transformation. J Biol Chem 2002;277:8312-20.
    [130] Versteeg S, Mogk A, Schumann W. The Bacillus subtilis htpG gene is not involved in thermal stress management. Mol Gen Genet 1999;261:582-8.
    [131] Borkovich KA, Farrelly FW, Finkelstein DB, et al. Hsp82 Is an Essential Protein That Is Required in Higher Concentrations for Growth of Cells at Higher Temperatures. Mol Cell Biol 1989;9:3919-30.
    [132] Meyer P, Prodromou C, Hu B, et al. Structural and functional analysis of the middle segment of hsp90: implications for ATP hydrolysis and client protein and cochaperone interactions. Mol Cell 2003;11:647-58.
    [133] Huai Q, Wang HC, Liu YD, et al. Structures of the N-terminal and middle domains of E-coli Hsp90 and conformation changes upon ADP binding. Structure 2005;13:579-90.
    [134] Harris SF, Shiau AK, Agard DA. The crystal structure of the carboxy-terminal dimerization domain of htpG, the Escherichia coli Hsp90, reveals a potential substrate binding site. Structure 2004;12:1087-97.
    [135] Dutta R, Inouye M. GHKL, an emergent ATPase/kinase superfamily. Trends Biochem Sci 2000;25:24-8.
    [136] McLaughlin SH, Ventouras LA, Lobbezoo B, et al. Independent ATPase activity of Hsp90 subunits creates a flexible assembly platform. J Mol Biol 2004;344:813-26.
    [137] Hawle P, Siepmann M, Harst A, et al. The middle domain of Hsp90 acts as a discriminator between different types of client proteins. Mol Cell Biol 2006;26:8385-95.
    [138] Minami Y, Kimura Y, Kawasaki H, et al. The carboxy-terminal region of mammalian HSP90 is required for its dimerization and function in vivo. Mol Cell Biol 1994;14:1459-64.
    [139] Soti C, Racz A, Csermely P. A Nucleotide-dependent molecular switch controls ATP binding at the C-terminal domain of Hsp90. N-terminal nucleotide binding unmasks a C-terminal binding pocket. J Biol Chem 2002;277:7066-75.
    [140] Garnier C, Lafitte D, Tsvetkov PO, et al. Binding of ATP to heat shock protein 90: evidence for an ATP-binding site in the C-terminal domain. J Biol Chem 2002;277:12208-14.
    [141] Scheufler C, Brinker A, Bourenkov G, et al. Structure of TPR domain-peptide complexes: critical elements in the assembly of the Hsp70-Hsp90 multichaperone machine. Cell 2000;101:199-210.
    [142] Prodromou C, Siligardi G, O'Brien R, et al. Regulation of Hsp90 ATPase activity by tetratricopeptide repeat (TPR)-domain co-chaperones. EMBO J 1999;18:754-62.
    [143] Zuehlke A, Johnson JL. Hsp90 and co-chaperones twist the functions of diverse client proteins. Biopolymers 2010;93:211-7.
    [144] Smith DF, Whitesell L, Nair SC, et al. Progesterone receptor structure and function altered by geldanamycin, an hsp90-binding agent. Mol Cell Biol 1995;15:6804-12.
    [145] Pratt WB. The hsp90-based chaperone system: involvement in signal transduction from a variety of hormone and growth factor receptors. Proc Soc Exp Biol Med 1998;217:420-34.
    [146] Holt SE, Aisner DL, Baur J, et al. Functional requirement of p23 and Hsp90 in telomerase complexes. Genes Dev 1999;13:817-26.
    [147] Park SJ, Suetsugu S, Takenawa T. Interaction of HSP90 to N-WASP leads to activation and protection from proteasome-dependent degradation. EMBO J 2005;24:1557-70.
    [148] Garcia-Cardena G, Fan R, Shah V, et al. Dynamic activation of endothelial nitric oxide synthase by Hsp90. Nature 1998;392:821-4.
    [149] Pearl LH. Hsp90 and Cdc37 -- a chaperone cancer conspiracy. Curr Opin Genet Dev 2005;15:55-61.
    [150] Zhao RM, Davey M, Hsu YC, et al. Navigating the chaperone network: An integrative map of physical and genetic interactions mediated by the Hsp90 chaperone. Cell 2005;120:715-27.
    [151] Jolly C, Morimoto RI. Role of the heat shock response and molecular chaperones in oncogenesis and cell death. J Natl Cancer Inst 2000;92:1564-72.
    [152] Muller L, Schaupp A, Walerych D, et al. Hsp90 regulates the activity of wild type p53 under physiological and elevated temperatures. J Biol Chem 2004;279:48846-54.
    [153] Walerych D, Kudla G, Gutkowska M, et al. Hsp90 chaperones wild-type p53 tumor suppressor protein. J Biol Chem 2004;279:48836-45.
    [154] Whitesell L, Sutphin PD, Pulcini EJ, et al. The physical association of multiple molecular chaperone proteins with mutant p53 is altered by Geldanamycin, an hsp90-binding agent. Mol Cell Biol 1998;18:1517-24.
    [155] Blagosklonny MV, Toretsky J, Bohen S, et al. Mutant conformation of p53 translated in vitro or in vivo requires functional HSP90. Proc Natl Acad Sci USA 1996;93:8379-83.
    [156] Sepehrnia B, Paz IB, Dasgupta G, et al. Heat shock protein 84 forms a complex with mutant p53 protein predominantly within a cytoplasmic compartment of the cell. J Biol Chem 1996;271:15084-90.
    [157] Falsone SF, Leptihn S, Osterauer A, et al. Oncogenic mutations reduce the stability of Src kinase. J Mol Biol 2004;344:281-91.
    [158] Oppermann H, Levinson W, Bishop JM. A Cellular Protein That Associates with the Transforming Protein of Rous-Sarcoma Virus Is Also a Heat-Shock Protein. Proc Natl Acad Sci-Biol 1981;78:1067-71.
    [159] Brugge J, Yonemoto W, Darrow D. Interaction between the Rous-Sarcoma Virus Transforming Protein and 2 Cellular Phosphoproteins - Analysis of the Turnover and Distribution of This Complex. Mol Cell Biol 1983;3:9-19.
    [160] Xu Y, Lindquist S. Heat-shock protein hsp90 governs the activity of pp60v-src kinase. Proc Natl Acad Sci U S A 1993;90:7074-8.
    [161] Fujimoto J, Shiota M, Iwahara T, et al. Characterization of the transforming activity of p80, a hyperphosphorylated protein in a Ki-1 lymphoma cell line with chromosomal translocation t(2;5). Proc Natl Acad Sci U S A 1996;93:4181-6.
    [162] Bonvini P, Gastaldi T, Falini B, et al. Nucleophosmin-anaplastic lymphoma kinase (NPM-ALK), a novel Hsp90-client tyrosine kinase: down-regulation of NPM-ALK expression and tyrosine phosphorylation in ALK(+) CD30(+) lymphoma cells by the Hsp90 antagonist 17-allylamino,17-demethoxygeldanamycin. Cancer Res 2002;62:1559-66.
    [163] Naoe T, Kiyoi H, Yamamoto Y, et al. FLT3 tyrosine kinase as a target molecule for selective antileukemia therapy. Cancer Chemoth Pharm 2001;48:S27-S30.
    [164] Minami Y, Kiyoi H, Yamamoto Y, et al. Selective apoptosis of tandemly duplicated FLT3-transformed leukemia cells by Hsp90 inhibitors. Leukemia 2002;16:1535-40.
    [165] Druker BJ, Tamura S, Buchdunger E, et al. Effects of a selective inhibitor of the Abl tyrosine kinase on the growth of Bcr-Abl positive cells. Nat Med 1996;2:561-6.
    [166] An WG, Schulte TW, Neckers LM. The heat shock protein 90 antagonist geldanamycin alters chaperone association with p210bcr-abl and v-src proteins before their degradation by the proteasome. Cell Growth Differ 2000;11:355-60.
    [167] Shiotsu Y, Neckers LM, Wortman I, et al. Novel oxime derivatives of radicicol induce erythroid differentiation associated with preferential G(1) phase accumulation against chronic myelogenous leukemia cells through destabilization of Bcr-Abl with Hsp90 complex. Blood 2000;96:2284-91.
    [168] Shah NP, Nicoll JM, Nagar B, et al. Multiple BCR-ABL kinase domain mutations confer polyclonal resistance to the tyrosine kinase inhibitor imatinib (STI571) in chronic phase and blast crisis chronic myeloid leukemia. Cancer Cell 2002;2:117-25.
    [169] Gorre ME, Ellwood-Yen K, Chiosis G, et al. BCR-ABL point mutants isolated from patients with imatinib mesylate-resistant chronic myeloid leukemia remain sensitive to inhibitors of the BCR-ABL chaperone heat shock protein 90. Blood 2002;100:3041-4.
    [170] Sidera K, Samiotaki M, Yfanti E, et al. Involvement of cell surface HSP90 in cell migration reveals a novel role in the developing nervous system. J Biol Chem 2004;279:45379-88.
    [171] Stellas D, Karameris A, Patsavoudi E. Monoclonal antibody 4C5 immunostains human melanomas and inhibits melanoma cell invasion and metastasis. Clin Cancer Res 2007;13:1831-8.
    [172] Li W, Li Y, Guan S, et al. Extracellular heat shock protein-90alpha: linking hypoxia to skin cell motility and wound healing. EMBO J 2007;26:1221-33.
    [173] Sidera K, Gaitanou M, Stellas D, et al. A critical role for surface HSP90 in cancer cell invasion involves extracellular interaction with HER-2. J Biol Chem 2008;283:2031-2041
    [174] Xu W, Yuan X, Xiang Z, et al. Surface charge and hydrophobicity determine ErbB2 binding to the Hsp90 chaperone complex. Nat Struct Mol Biol 2005;12:120-6.
    [175] Lei HT, Romeo G, Kazlauskas A. Heat shock protein 90 alpha-dependent translocation of annexin II to the surface of endothelial cells modulates plasmin activity in the diabetic rat aorta. Circ Res 2004;94:902-9.
    [176] Basu S, Binder RJ, Ramalingam T, et al. CD91 is a common receptor for heat shock proteins gp96, hsp90, hsp70, and calreticulin. Immunity 2001;14:303-13.
    [177] Triantafilou M, Triantafilou K. Heat-shock protein 70 and heat-shock protein 90 associate with Toll-like receptor 4 in response to bacterial lipopolysaccharide. Biochemical Society Transactions 2004;32:636-9.
    [178] Jin S, Song YC, Emili A, et al. JlpA of Campylobacter jejuni interacts with surface-exposed heat shock protein 90alpha and triggers signalling pathways leading to the activation of NF-kappaB and p38 MAP kinase in epithelial cells. Cell Microbiol 2003;5:165-74.
    [179] Reyes-Del Valle J, Chavez-Salinas S, Medina F, et al. Heat shock protein 90 and heat shock protein 70 are components of dengue virus receptor complex in human cells. J Virol 2005;79:4557-67.
    [180] Ullrich SJ, Robinson EA, Law LW, et al. A mouse tumor-specific transplantation antigen is a heat shock-related protein. Proc Natl Acad Sci U S A 1986;83:3121-5.
    [181] Becker B, Multhoff G, Farkas B, et al. Induction of Hsp90 protein expression in malignant melanomas and melanoma metastases. Exp Dermatol 2004;13:27-32.
    [182] Ferrarini M, Heltai S, Zocchi MR, et al. Unusual expression and localization of heat-shock proteins in human tumor cells. Int J Cancer 1992;51:613-9.
    [183] Sapozhnikov AM, Ponomarev ED, Tarasenko TN, et al. Spontaneous apoptosis and expression of cell surface heat-shock proteins in cultured EL-4 lymphoma cells. Cell Prolif 1999;32:363-78.
    [184] Binder RJ, Vatner R, Srivastava P. The heat-shock protein receptors: some answers and more questions. Tissue Antigens 2004;64:442-51.
    [185] Tsutsumi S, Scroggins B, Koga F, et al. A small molecule cell-impermeant Hsp90 antagonist inhibits tumor cell motility and invasion. Oncogene 2008;27:2478-87.
    [186] Picard D. Heat-shock protein 90, a chaperone for folding and regulation. Cell Mol Life Sci 2002;59:1640-8.
    [187] Caplan AJ, Mandal AK, Theodoraki MA. Molecular chaperones and protein kinase quality control. Trends Cell Biol 2007;17:87-92.
    [188] Kim YS, Alarcon SV, Lee S, et al. Update on Hsp90 inhibitors in clinical trial. Curr Top Med Chem 2009;9:1479-92.
    [189] Chen B, Piel WH, Gui L, et al. The HSP90 family of genes in the human genome: insights into their divergence and evolution. Genomics 2005;86:627-37.
    [190] Nemoto TK, Ono T, Tanaka K. Substrate-binding characteristics of proteins in the 90 kDa heat shock protein family. Biochem J 2001;354:663-70.
    [191] Minn AJ, Gupta GP, Siegel PM, et al. Genes that mediate breast cancer metastasis to lung. Nature 2005;436:518-24.
    [192] Buccione R, Orth JD, McNiven MA. Foot and mouth: podosomes, invadopodia and circular dorsal ruffles. Nat Rev Mol Cell Biol 2004;5:647-57.
    [193] Coopman PJ, Do MTH, Thompson EW, et al. Phagocytosis of cross-linked gelatin matrix by human breast carcinoma cells correlates with their invasive capacity. Clin Cancer Res 1998;4:507-15.
    [194] Thompson EW, Paik S, Brunner N, et al. Association of increased basement membrane invasiveness with absence of estrogen receptor and expression of vimentin in human breast cancer cell lines. J Cell Physiol 1992;150:534-44.
    [195] Bowden ET, Barth M, Thomas D, et al. An invasion-related complex of cortactin, paxillin and PKC mu associates with invadopodia at sites of extracellular matrix degradation. Oncogene 1999;18:4440-9.
    [196] Seals DF, Azucena EF, Pass I, et al. The adaptor protein Tks5/Fish is required for podosome formation and function, and for the protease-driven invasion of cancer cells. Cancer Cell 2005;7:155-65.
    [197] Zhang LH, Tian B, Diao LR, et al. Dominant expression of 85-kDa form of cortactin in colorectal cancer. J Cancer Res Clin Oncol 2006;132:113-20.
    [198] Yamaguchi H, Wyckoff J, Condeelis J. Cell migration in tumors. Curr Opin Cell Biol 2005;17:559-64.
    [199] Weaver AM. Invadopodia: specialized cell structures for cancer invasion. Clin Exp Metastasis 2006;23:97-105.
    [200] Clark ES, Whigham AS, Yarbrough WG, et al. Cortactin is an essential regulator of matrix metalloproteinase secretion and extracellular matrix degradation in invadopodia. Cancer Res 2007;67:4227-35.
    [201] Baldassarre M, Pompeo A, Beznoussenko G, et al. Dynamin participates in focal extracellular matrix degradation by invasive cells. Mol Biol Cell 2003;14:1074-84.
    [202] Chuang YY, Tran NL, Rusk N, et al. Role of synaptojanin 2 in glioma cell migration and invasion. Cancer Res 2004;64:8271-5.
    [203] Hashimoto S, Onodera Y, Hashimoto A, et al. Requirement for Arf6 in breast cancer invasive activities. Proc Natl Acad Sci U S A 2004;101:6647-52.
    [204] McHugh B, Krause SA, Yu B, et al. Invadolysin: a novel, conserved metalloprotease links mitotic structural rearrangements with cell migration. J Cell Biol 2004;167:673-86.
    [205] Onodera Y, Hashimoto S, Hashimoto A, et al. Expression of AMAP1, an ArfGAP, provides novel targets to inhibit breast cancer invasive activities. EMBO J 2005;24:963-73.
    [206] Tague SE, Muralidharan V, D'Souza-Schorey C. ADP-ribosylation factor 6 regulates tumor cell invasion through the activation of the MEK/ERK signaling pathway. Proc Natl Acad Sci USA 2004;101:9671-6.
    [207] Yamaguchi H, Lorenz M, Kempiak S, et al. Molecular mechanisms of invadopodium formation: the role of the N-WASP-Arp2/3 complex pathway and cofilin. J Cell Biol 2005;168:441-52.
    [208] Miki H, Takenawa T. Regulation of actin dynamics by WASP family proteins. J Biochem 2003;134:309-13.
    [209] Millard TH, Sharp SJ, Machesky LM. Signalling to actin assembly via the WASP (Wiskott-Aldrich syndrome protein)-family proteins and the Arp2/3 complex. Biochem J 2004;380:1-17.
    [210] Stradal TE, Rottner K, Disanza A, et al. Regulation of actin dynamics by WASP and WAVE family proteins. Trends Cell Biol 2004;14:303-11.
    [211] Nakahara H, Otani T, Sasaki T, et al. Involvement of Cdc42 and Rac small G proteins in invadopodia formation of RPMI7951 cells. Genes Cells 2003;8:1019-27.
    [212] Stylli SS, Kaye AH, Lock P. Invadopodia: at the cutting edge of tumour invasion. J Clin Neurosci 2008;15:725-37.
    [213] Ammer AG, Weed SA. Cortactin branches out: roles in regulating protrusive actin dynamics. Cell Motil Cytoskeleton 2008;65:687-707.
    [214] Robinson RC, Turbedsky K, Kaiser DA, et al. Crystal structure of Arp2/3 complex. Science 2001;294:1679-84.
    [215] Clark ES, Weaver AM. A new role for cortactin in invadopodia: regulation of protease secretion. Eur J Cell Biol 2008;87:581-90.
    [216] Gonzalez L, Agullo-Ortuno MT, Garcia-Martinez JM, et al. Role of c-Src in human MCF7 breast cancer cell tumorigenesis. J Biol Chem 2006;281:20851-64.
    [217] Kempiak SJ, Yamaguchi H, Sarmiento C, et al. A neural Wiskott-Aldrich syndrome protein-mediated pathway for localized activation of actin polymerization that is regulated by cortactin. J Biol Chem 2005;280:5836-42.
    [218] Nagase H, Visse R, Murphy G. Structure and function of matrix metalloproteinases and TIMPs. Cardiovasc Res 2006;69:562-73.
    [219] Page-McCaw A, Ewald AJ, Werb Z. Matrix metalloproteinases and the regulation of tissue remodelling. Nat Rev Mol Cell Biol 2007;8:221-33.
    [220] Harris ED, Jr., Krane SM. An endopeptidase from rheumatoid synovial tissue culture. Biochim Biophys Acta 1972;258:566-76.
    [221] Parks WC, Wilson CL, Lopez-Boado YS. Matrix metalloproteinases as modulators of inflammation and innate immunity. Nat Rev Immunol 2004;4:617-29.
    [222] Deryugina EI, Quigley JP. Matrix metalloproteinases and tumor metastasis. Cancer Metastasis Rev 2006;25:9-34.
    [223] Liotta LA, Tryggvason K, Garbisa S, et al. Metastatic potential correlates with enzymatic degradation of basement membrane collagen. Nature 1980;284:67-8.
    [224] Visse R, Nagase H. Matrix metalloproteinases and tissue inhibitors of metalloproteinases: structure, function, and biochemistry. Circ Res 2003;92:827-39.
    [225] Ra HJ, Parks WC. Control of matrix metalloproteinase catalytic activity. Matrix Biol 2007;26:587-96.
    [226] Kinoshita T, Sato H, Takino T, et al. Processing of a precursor of 72-kilodalton type IV collagenase/gelatinase A by a recombinant membrane-type 1 matrix metalloproteinase. Cancer Res 1996;56:2535-8.
    [227] Wang Z, Juttermann R, Soloway PD. TIMP-2 is required for efficient activation of proMMP-2 in vivo. J Biol Chem 2000;275:26411-5.
    [228] Zucker S, Drews M, Conner C, et al. Tissue inhibitor of metalloproteinase-2 (TIMP-2) binds to the catalytic domain of the cell surface receptor, membrane type 1-matrix metalloproteinase 1 (MT1-MMP). J Biol Chem 1998;273:1216-22.
    [229] Strongin AY, Collier I, Bannikov G, et al. Mechanism of cell surface activation of 72-kDa type IV collagenase. Isolation of the activated form of the membrane metalloprotease. J Biol Chem 1995;270:5331-8.
    [230] Strongin AY, Marmer BL, Grant GA, et al. Plasma membrane-dependent activation of the 72-kDa type IV collagenase is prevented by complex formation with TIMP-2. J Biol Chem 1993;268:14033-9.
    [231] Xu X, Wang Y, Lauer-Fields JL, et al. Contributions of the MMP-2 collagen binding domain to gelatin cleavage. Substrate binding via the collagen binding domain is required for hydrolysis of gelatin but not short peptides. Matrix Biol 2004;23:171-81.
    [232] Overall CM, King AE, Bigg HF, et al. Identification of the TIMP-2 binding site on the gelatinase A hemopexin C-domain by site-directed mutagenesis and the yeast two-hybrid system. Ann N Y Acad Sci 1999;878:747-53.
    [233] Wallon UM, Overall CM. The hemopexin-like domain (C domain) of human gelatinase A (matrix metalloproteinase-2) requires Ca2+ for fibronectin and heparin binding. Binding properties of recombinant gelatinase A C domain to extracellular matrix and basement membrane components. J Biol Chem 1997;272:7473-81.
    [234] Bergmann U, Tuuttila A, Stetler-Stevenson WG, et al. Autolytic activation of recombinant human 72 kilodalton type IV collagenase. Biochemistry 1995;34:2819-25.
    [235] Howard EW, Bullen EC, Banda MJ. Regulation of the autoactivation of human 72-kDa progelatinase by tissue inhibitor of metalloproteinases-2. J Biol Chem 1991;266:13064-9.
    [236] Grenert JP, Johnson BD, Toft DO. The importance of ATP binding and hydrolysis by hsp90 in formation and function of protein heterocomplexes. J Biol Chem 1999;274:17525-33.
    [237] Brooks PC, Silletti S, von Schalscha TL, et al. Disruption of angiogenesis by PEX, a noncatalytic metalloproteinase fragment with integrin binding activity. Cell 1998;92:391-400.
    [238] Piccard H, Van den Steen PE, Opdenakker G. Hemopexin domains as multifunctional liganding modules in matrix metal loproteinases and other proteins. J Leuk Biol 2007;81:870-92.
    [239] Ra HJ, Parks WC. Control of matrix metalloproteinase catalytic activity. Matrix Biology 2007;26:587-96.
    [240] Stetler-Stevenson WG, Seo DW. TIMP-2: an endogenous inhibitor of angiogenesis. Trends Mol Med 2005;11:97-103.
    [241] Bertini I, Calderone V, Fragai M, et al. Evidence of reciprocal reorientation of the catalytic and hemopexin-like domains of full-length MMP-12. J Am Chem Soc 2008;130:7011-21.
    [242] Rosenblum G, Van den Steen PE, Cohen SR, et al. Insights into the structure and domain flexibility of full-length pro-matrix metalloproteinase-9/gelatinase B. Structure 2007;15:1227-36.
    [243] Bertini I, Fragai M, Luchinat C, et al. Interdomain flexibility in full-length matrix metalloproteinase-1 (MMP-1). J Biol Chem 2009; 284:12821-8.
    [244] Diaz N, Suarez D, Valdes H. From the X-ray compact structure to the elongated form of the full-length MMP-2 enzyme in solution: a molecular dynamics study. J Am Chem Soc 2008;130:14070-1.
    [245] Osenkowski P, Meroueh SO, et al. Mutational and structural analyses of the hinge region of membrane type 1-matrix metalloproteinase and enzyme processing. J Biol Chem 2005;280:26160-8.
    [246] Minami Y, Kawasaki H, Minami M, et al. A critical role for the proteasome activator PA28 in the Hsp90-dependent protein refolding. J Biol Chem 2000;275:9055-61.
    [247] Yonehara M, Minami Y, Kawata Y, et al. Heat-induced chaperone activity of HSP90. J Biol Chem 1996;271:2641-5.
    [248] Yang Y, Rao R, Shen J, et al. Role of acetylation and extracellular location of heat shock protein 90alpha in tumor cell invasion. Cancer Res 2008;68:4833-42.
    [249] Wang X, Song X, Zhuo W, et al. The regulatory mechanism of Hsp90{alpha} secretion and its function in tumor malignancy. Proc Natl Acad Sci U S A 2009;106:21288-93.
    [250] Dean RA, Overall CM. Proteomics discovery of metalloproteinase substrates in the cellular context by iTRAQ labeling reveals a diverse MMP-2 substrate degradome. Mol Cell Proteomics 2007;6:611-23.
    [251] Csermely P, Schnaider T, Soti C, et al. The 90-kDa molecular chaperone family: structure, function, and clinical applications. A comprehensive review. Pharmacol Ther 1998;79:129-68.
    [252] Overall CM. Molecular determinants of metalloproteinase substrate specificity - Matrix metalloproteinase substrate binding domains, modules, and exosites. Mol Biotech 2002;22:51-86.
    [253] Butler GS, Butler MJ, Atkinson SJ, et al. The TIMP2 membrane type 1 metalloproteinase "receptor" regulates the concentration and efficient activation of progelatinase A. A kinetic study. J Biol Chem 1998;273:871-80.
    [254] Rundhaug JE. Matrix metalloproteinases, angiogenesis, and cancer: commentary re: A. C. Lockhart et al., Reduction of wound angiogenesis in patients treated with BMS-275291, a broad spectrum matrix metalloproteinase inhibitor. Clin. Cancer Res., 9: 00-00, 2003. Clin Cancer Res 2003;9:551-4.
    [255] Carmeliet P. Angiogenesis in life, disease and medicine. Nature 2005;438:932-6.
    [256] Gariano RF, Gardner TW. Retinal angiogenesis in development and disease. Nature 2005;438:960-6.
    [257] Carmeliet P, Jain RK. Angiogenesis in cancer and other diseases. Nature 2000;407:249-57.
    [258] Folkman J. Tumor angiogenesis: therapeutic implications. N Engl J Med 1971;285:1182-6.
    [259] Ferrara N, Kerbel RS. Angiogenesis as a therapeutic target. Nature 2005;438:967-74.
    [260] Folkman J. Angiogenesis: an organizing principle for drug discovery? Nat Rev Drug Discov 2007;6:273-86.
    [261] Cao Y. Endogenous angiogenesis inhibitors and their therapeutic implications. Int J Biochem Cell Biol 2001;33:357-69.
    [262] Cao Y. Tumor angiogenesis and molecular targets for therapy. Front Biosci 2009;14:3962-73.
    [263] Ades EW, Candal FJ, Swerlick RA, et al. HMEC-1: establishment of an immortalized human microvascular endothelial cell line. J Invest Dermatol 1992;99:683-90.
    [264] Jaffe EA, Nachman RL, Becker CG, et al. Culture of human endothelial cells derived from umbilical veins. Identification by morphologic and immunologic criteria. J Clin Invest 1973;52:2745-56.

© 2004-2018 中国地质图书馆版权所有 京ICP备05064691号 京公网安备11010802017129号

地址:北京市海淀区学院路29号 邮编:100083

电话:办公室:(+86 10)66554848;文献借阅、咨询服务、科技查新:66554700