HSP90: References

1. Csermely,P., Schnaider,T., Soti,C., Prohaszka,Z., & Nardai,G. The 90-kDa molecular chaperone family: structure, function, and clinical applications. A comprehensive review. Pharmacol. Ther. 79, 129-168 (1998). [PubMed]

2. Felts,S.J. et al. The hsp90-related protein TRAP1 is a mitochondrial protein with distinct functional properties. J. Biol. Chem. 275, 3305-3312 (2000). [PubMed]

3. Krishna,P. & Gloor,G. The Hsp90 family of proteins in Arabidopsis thaliana. Cell Stress. Chaperones. 6, 238-246 (2001). [PubMed]

4. Pockley,A.G. & Multhoff,G. Cell stress proteins in extracellular fluids: friend or foe? Novartis. Found. Symp. 291, 86-95 (2008). [PubMed]

5. Pockley,A.G., Muthana,M., & Calderwood,S.K. The dual immunoregulatory roles of stress proteins. Trends Biochem. Sci. 33, 71-79 (2008). [PubMed]

6. Calderwood,S.K., Khaleque,M.A., Sawyer,D.B., & Ciocca,D.R. Heat shock proteins in cancer: chaperones of tumorigenesis. Trends Biochem. Sci. 31, 164-172 (2006). [PubMed]

7. Whitesell,L. & Lindquist,S.L. HSP90 and the chaperoning of cancer. Nat. Rev. Cancer 5, 761-772 (2005). [PubMed]

8. Ciocca,D.R. & Calderwood,S.K. Heat shock proteins in cancer: diagnostic, prognostic, predictive, and treatment implications. Cell Stress. Chaperones. 10, 86-103 (2005). [PubMed]

9. Lindquist,S. The heat-shock response. Annu. Rev. Biochem. 55, 1151-1191 (1986). [PubMed]

10. Lindquist,S. Regulation of protein synthesis during heat shock. Nature 293, 311-314 (1981). [PubMed]

11. Ritossa,F. A new puffing pattern induced by temperature shock and DNP in Drosophila. Experientia 18, 571-573 (1962). DOI: 10.1007/BF02172188

12. Lindquist,S. & Craig,E.A. The heat-shock proteins. Annu. Rev. Genet. 22, 631-677 (1988). [PubMed]

13. Shamovsky,I. & Nudler,E. New insights into the mechanism of heat shock response activation. Cell Mol. Life Sci. 65, 855-861 (2008). [PubMed]

14. Bardwell,J.C. & Craig,E.A. Eukaryotic Mr 83,000 heat shock protein has a homologue in Escherichia coli. Proc. Natl. Acad. Sci. U. S. A 84, 5177-5181 (1987). [PubMed]

15. Heitzer,A., Mason,C.A., Snozzi,M., & Hamer,G. Some effects of growth conditions on steady state and heat shock induced htpG gene expression in continuous cultures of Escherichia coli. Arch. Microbiol. 155, 7-12 (1990). [PubMed]

16. Conner,T.W., Lafayette,P.R., Nagao,R.T., & Key,J.L. Sequence and Expression of a HSP83 from Arabidopsis thaliana. Plant Physiol 94, 1689-1695 (1990). [PubMed]

17. Wiech,H., Buchner,J., Zimmermann,R., & Jakob,U. Hsp90 chaperones protein folding in vitro. Nature 358, 169-170 (1992). [PubMed]

18. Pratt,W.B. The hsp90-based chaperone system: involvement in signal transduction from a variety of hormone and growth factor receptors. Proc. Soc. Exp. Biol. Med. 217, 420-434 (1998). [PubMed]

19. Ullrich,S.J., Robinson,E.A., Law,L.W., Willingham,M., & Appella,E. A mouse tumor-specific transplantation antigen is a heat shock-related protein. Proc. Natl. Acad. Sci. U. S. A 83, 3121-3125 (1986). [PubMed]

20. Prodromou,C. et al. Identification and structural characterization of the ATP/ADP-binding site in the Hsp90 molecular chaperone. Cell 90, 65-75 (1997). [PubMed]

21. Shiau,A.K., Harris,S.F., Southworth,D.R., & Agard,D.A. Structural Analysis of E. coli hsp90 reveals dramatic nucleotide-dependent conformational rearrangements. Cell 127, 329-340 (2006). [PubMed]

22. Ali,M.M. et al. Crystal structure of an Hsp90-nucleotide-p23/Sba1 closed chaperone complex. Nature 440, 1013-1017 (2006). [PubMed]

23. Ratzke,C., Mickler,M., Hellenkamp,B., Buchner,J., & Hugel,T. Dynamics of heat shock protein 90 C-terminal dimerization is an important part of its conformational cycle. Proc. Natl. Acad. Sci. U. S. A 107, 16101-16106 (2010). [PubMed]

24. Didenko,T., Duarte,A.M., Karagoz,G.E., & Rudiger,S.G. Hsp90 structure and function studied by NMR spectroscopy. Biochim. Biophys. Acta 1823, 636-647 (2012). [PubMed]

25. Kampinga,H.H. et al. Guidelines for the nomenclature of the human heat shock proteins. Cell Stress. Chaperones. 14, 105-111 (2009). [PubMed]

26. Chen,B., Piel,W.H., Gui,L., Bruford,E., & Monteiro,A. The HSP90 family of genes in the human genome: insights into their divergence and evolution. Genomics 86, 627-637 (2005). [PubMed]

27. Schweinfest,C.W. et al. Cloning and sequence analysis of Hsp89alpha DeltaN, a new member of theHsp90 gene family. Biochim. Biophys. Acta 1398, 18-24 (1998). [PubMed]

28. Krone,P.H. & Sass,J.B. HSP 90 alpha and HSP 90 beta genes are present in the zebrafish and are differentially regulated in developing embryos. Biochem. Biophys. Res. Commun. 204, 746-752 (1994). [PubMed]

29. Du,S.J., Li,H., Bian,Y., & Zhong,Y. Heat-shock protein 90alpha1 is required for organized myofibril assembly in skeletal muscles of zebrafish embryos. Proc. Natl. Acad. Sci. U. S. A 105, 554-559 (2008). [PubMed]

30. Tesic,M., Marsh,J.A., Cullinan,S.B., & Gaber,R.F. Functional interactions between Hsp90 and the co-chaperones Cns1 and Cpr7 in Saccharomyces cerevisiae. J. Biol. Chem. 278, 32692-32701 (2003). [PubMed]

31. Swoboda,R.K. et al. Structure and regulation of the HSP90 gene from the pathogenic fungus Candida albicans. Infect. Immun. 63, 4506-4514 (1995). [PubMed]

32. Yamamoto,M., Takahashi,Y., Inano,K., Horigome,T., & Sugano,H. Characterization of the hydrophobic region of heat shock protein 90. J. Biochem. 110, 141-145 (1991). [PubMed]

33. Iwasaki,M. et al. Purification of heat shock protein 90 from calf uterus and rat liver and characterization of the highly hydrophobic region. Biochim. Biophys. Acta 992, 1-8 (1989). [PubMed]

34. Chang,S.C., Erwin,A.E., & Lee,A.S. Glucose-regulated protein (GRP94 and GRP78) genes share common regulatory domains and are coordinately regulated by common trans-acting factors. Mol. Cell Biol. 9, 2153-2162 (1989). [PubMed]

35. Chen,B., Zhong,D., & Monteiro,A. Comparative genomics and evolution of the HSP90 family of genes across all kingdoms of organisms. BMC. Genomics 7, 156 (2006). [PubMed]

36. Stechmann,A. & Cavalier-Smith,T. Evolutionary origins of Hsp90 chaperones and a deep paralogy in their bacterial ancestors. J. Eukaryot. Microbiol. 51, 364-373 (2004). [PubMed]

37. Han,G. et al. Large-scale phosphoproteome analysis of human liver tissue by enrichment and fractionation of phosphopeptides with strong anion exchange chromatography. Proteomics. 8, 1346-1361 (2008). [PubMed]

38. Gupta,R.S. Phylogenetic analysis of the 90 kD heat shock family of protein sequences and an examination of the relationship among animals, plants, and fungi species. Mol. Biol. Evol. 12, 1063-1073 (1995). [PubMed]

39. Nicchitta,C.V. Biochemical, cell biological and immunological issues surrounding the endoplasmic reticulum chaperone GRP94/gp96. Curr. Opin. Immunol. 10, 103-109 (1998). [PubMed]

40. Tamura,Y., Peng,P., Liu,K., Daou,M., & Srivastava,P.K. Immunotherapy of tumors with autologous tumor-derived heat shock protein preparations. Science 278, 117-120 (1997). [PubMed]

41. Young,J.C., Obermann,W.M., & Hartl,F.U. Specific binding of tetratricopeptide repeat proteins to the C-terminal 12-kDa domain of hsp90. J. Biol. Chem. 273, 18007-18010 (1998). [PubMed]

42. Chen,S., Sullivan,W.P., Toft,D.O., & Smith,D.F. Differential interactions of p23 and the TPR-containing proteins Hop, Cyp40, FKBP52 and FKBP51 with Hsp90 mutants. Cell Stress. Chaperones. 3, 118-129 (1998). [PubMed]

43. Chen,C.F. et al. A new member of the hsp90 family of molecular chaperones interacts with the retinoblastoma protein during mitosis and after heat shock. Mol. Cell Biol. 16, 4691-4699 (1996). [PubMed]

44. Bardwell,J.C. & Craig,E.A. Ancient heat shock gene is dispensable. J. Bacteriol. 170, 2977-2983 (1988). [PubMed]

45. Thomas,J.G. & Baneyx,F. ClpB and HtpG facilitate de novo protein folding in stressed Escherichia coli cells. Mol. Microbiol. 36, 1360-1370 (2000). [PubMed]

46. Jakob,U. et al. Structural organization of procaryotic and eucaryotic Hsp90. Influence of divalent cations on structure and function. J. Biol. Chem. 270, 14412-14419 (1995). [PubMed]

47. Taipale,M., Jarosz,D.F., & Lindquist,S. HSP90 at the hub of protein homeostasis: emerging mechanistic insights. Nat. Rev. Mol. Cell Biol. 11, 515-528 (2010). [PubMed]

48. Wandinger,S.K., Richter,K., & Buchner,J. The Hsp90 chaperone machinery. J. Biol. Chem. 283, 18473-18477 (2008). [PubMed]

49. Vaughan,C.K. et al. Structure of an Hsp90-Cdc37-Cdk4 complex. Mol. Cell 23, 697-707 (2006). [PubMed]

50. Dollins,D.E., Warren,J.J., Immormino,R.M., & Gewirth,D.T. Structures of GRP94-nucleotide complexes reveal mechanistic differences between the hsp90 chaperones. Mol. Cell 28, 41-56 (2007). [PubMed]

51. Cunningham,C.N., Krukenberg,K.A., & Agard,D.A. Intra- and intermonomer interactions are required to synergistically facilitate ATP hydrolysis in Hsp90. J. Biol. Chem. 283, 21170-21178 (2008). [PubMed]

52. Krukenberg,K.A., Forster,F., Rice,L.M., Sali,A., & Agard,D.A. Multiple conformations of E. coli Hsp90 in solution: insights into the conformational dynamics of Hsp90. Structure. 16, 755-765 (2008). [PubMed]

53. Southworth,D.R. & Agard,D.A. Species-dependent ensembles of conserved conformational states define the Hsp90 chaperone ATPase cycle. Mol. Cell 32, 631-640 (2008). [PubMed]

54. Vaughan,C.K., Piper,P.W., Pearl,L.H., & Prodromou,C. A common conformationally coupled ATPase mechanism for yeast and human cytoplasmic HSP90s. FEBS J. 276, 199-209 (2009). [PubMed]

55. Milioni,D. & Hatzopoulos,P. Genomic organization of hsp90 gene family in Arabidopsis. Plant Mol. Biol. 35, 955-961 (1997). [PubMed]

56. Cha,J.Y. et al. Functional characterization of orchardgrass endoplasmic reticulum-resident Hsp90 (DgHsp90) as a chaperone and an ATPase. Plant Physiol Biochem. 47, 859-866 (2009). [PubMed]

57. Krishna,P., Reddy,R.K., Sacco,M., Frappier,J.R., & Felsheim,R.F. Analysis of the native forms of the 90 kDa heat shock protein (hsp90) in plant cytosolic extracts. Plant Mol. Biol. 33, 457-466 (1997). [PubMed]

58. Liu,D., Zhang,X., Cheng,Y., Takano,T., & Liu,S. rHsp90 gene expression in response to several environmental stresses in rice (Oryza sativa L.). Plant Physiol Biochem. 44, 380-386 (2006). [PubMed]

59. Marrs,K.A. et al. Characterization of two maize HSP90 heat shock protein genes: expression during heat shock, embryogenesis, and pollen development. Dev. Genet. 14, 27-41 (1993). [PubMed]

60. Koning,A.J., Rose,R., & Comai,L. Developmental expression of tomato heat-shock cognate protein 80. Plant Physiol 100, 801-811 (1992). [PubMed]

61. Xu,Z.S. et al. Heat shock protein 90 in plants: molecular mechanisms and roles in stress responses. Int. J. Mol. Sci. 13, 15706-15723 (2012). [PubMed]

62. Sangster,T.A. & Queitsch,C. The HSP90 chaperone complex, an emerging force in plant development and phenotypic plasticity. Curr. Opin. Plant Biol. 8, 86-92 (2005). [PubMed]

63. Yabe,N., Takahashi,T., & Komeda,Y. Analysis of tissue-specific expression of Arabidopsis thaliana HSP90-family gene HSP81. Plant Cell Physiol 35, 1207-1219 (1994). [PubMed]

64. Cha,J.Y. et al. Structural and functional differences of cytosolic 90-kDa heat-shock proteins (Hsp90s) in Arabidopsis thaliana. Plant Physiol Biochem. 70, 368-373 (2013). [PubMed]

65. Erkine,A.M., Szent-Gyorgyi,C., Simmons,S.F., & Gross,D.S. The upstream sequences of the HSP82 and HSC82 genes of Saccharomyces cerevisiae: regulatory elements and nucleosome positioning motifs. Yeast 11, 573-580 (1995). [PubMed]

66. Borkovich,K.A., Farrelly,F.W., Finkelstein,D.B., Taulien,J., & Lindquist,S. hsp82 is an essential protein that is required in higher concentrations for growth of cells at higher temperatures. Mol. Cell Biol. 9, 3919-3930 (1989). [PubMed]

67. Dudgeon,D.D., Zhang,N., Ositelu,O.O., Kim,H., & Cunningham,K.W. Nonapoptotic death of Saccharomyces cerevisiae cells that is stimulated by Hsp90 and inhibited by calcineurin and Cmk2 in response to endoplasmic reticulum stresses. Eukaryot. Cell 7, 2037-2051 (2008). [PubMed]

68. Eisenberg,T., Carmona-Gutierrez,D., Buttner,S., Tavernarakis,N., & Madeo,F. Necrosis in yeast. Apoptosis.(2010). [PubMed]

69. Donze,O. & Picard,D. Hsp90 binds and regulates Gcn2, the ligand-inducible kinase of the alpha subunit of eukaryotic translation initiation factor 2 [corrected]. Mol. Cell Biol. 19, 8422-8432 (1999). [PubMed]

70. Franzosa,E.A., Albanese,V., Frydman,J., Xia,Y., & McClellan,A.J. Heterozygous yeast deletion collection screens reveal essential targets of Hsp90. PLoS. ONE. 6, e28211 (2011). [PubMed]

71. Silva,A. et al. Involvement of yeast HSP90 isoforms in response to stress and cell death induced by acetic acid. PLoS. ONE. 8, e71294 (2013). [PubMed]

72. Singh,S.D. et al. Hsp90 governs echinocandin resistance in the pathogenic yeast Candida albicans via calcineurin. PLoS. Pathog. 5, e1000532 (2009). [PubMed]

73. Cowen,L.E. Hsp90 orchestrates stress response signaling governing fungal drug resistance. PLoS. Pathog. 5, e1000471 (2009). [PubMed]

74. LaFayette,S.L. et al. PKC signaling regulates drug resistance of the fungal pathogen Candida albicans via circuitry comprised of Mkc1, calcineurin, and Hsp90. PLoS. Pathog. 6, e1001069 (2010). [PubMed]

75. Shapiro,R.S. et al. Hsp90 orchestrates temperature-dependent Candida albicans morphogenesis via Ras1-PKA signaling. Curr. Biol. 19, 621-629 (2009). [PubMed]

76. Robbins,N. et al. Hsp90 governs dispersion and drug resistance of fungal biofilms. PLoS. Pathog. 7, e1002257 (2011). [PubMed]

77. Cowen,L.E. et al. Harnessing Hsp90 function as a powerful, broadly effective therapeutic strategy for fungal infectious disease. Proc. Natl. Acad. Sci. U. S. A 106, 2818-2823 (2009). [PubMed]

78. Sass,J.B., Martin,C.C., & Krone,P.H. Restricted expression of the zebrafish hsp90alpha gene in slow and fast muscle fiber lineages. Int. J. Dev. Biol. 43, 835-838 (1999). [PubMed]

79. Etheridge,L., Diiorio,P., & Sagerstrom,C.G. A zebrafish unc-45-related gene expressed during muscle development. Dev. Dyn. 224, 457-460 (2002). [PubMed]

80. Sass,J.B., Weinberg,E.S., & Krone,P.H. Specific localization of zebrafish hsp90 alpha mRNA to myoD-expressing cells suggests a role for hsp90 alpha during normal muscle development. Mech. Dev. 54, 195-204 (1996). [PubMed]

81. Etard,C. et al. The UCS factor Steif/Unc-45b interacts with the heat shock protein Hsp90a during myofibrillogenesis. Dev. Biol. 308, 133-143 (2007). [PubMed]

82. Morcillo,G., Diez,J.L., Carbajal,M.E., & Tanguay,R.M. HSP90 associates with specific heat shock puffs (hsr omega) in polytene chromosomes of Drosophila and Chironomus. Chromosoma 102, 648-659 (1993). [PubMed]

83. Tissieres,A., Mitchell,H.K., & Tracy,U.M. Protein synthesis in salivary glands of Drosophila melanogaster: relation to chromosome puffs. J. Mol. Biol. 84, 389-398 (1974). [PubMed]

84. Ashburner,M. & Bonner,J.J. The induction of gene activity in drosophilia by heat shock. Cell 17, 241-254 (1979). [PubMed]

85. Olivieri,D., Senti,K.A., Subramanian,S., Sachidanandam,R., & Brennecke,J. The cochaperone shutdown defines a group of biogenesis factors essential for all piRNA populations in Drosophila. Mol. Cell 47, 954-969 (2012). [PubMed]

86. Gangaraju,V.K. et al. Drosophila Piwi functions in Hsp90-mediated suppression of phenotypic variation. Nat. Genet. 43, 153-158 (2011). [PubMed]

87. Bharadwaj,S., Ali,A., & Ovsenek,N. Multiple components of the HSP90 chaperone complex function in regulation of heat shock factor 1 In vivo. Mol. Cell Biol. 19, 8033-8041 (1999). [PubMed]

88. Duina,A.A., Kalton,H.M., & Gaber,R.F. Requirement for Hsp90 and a CyP-40-type cyclophilin in negative regulation of the heat shock response. J. Biol. Chem. 273, 18974-18978 (1998). [PubMed]

89. Zou,J., Guo,Y., Guettouche,T., Smith,D.F., & Voellmy,R. Repression of heat shock transcription factor HSF1 activation by HSP90 (HSP90 complex) that forms a stress-sensitive complex with HSF1. Cell 94, 471-480 (1998). [PubMed]

90. Marchler,G. & Wu,C. Modulation of Drosophila heat shock transcription factor activity by the molecular chaperone DROJ1. EMBO J. 20, 499-509 (2001). [PubMed]

91. Schroda,M. The Chlamydomonas genome reveals its secrets: chaperone genes and the potential roles of their gene products in the chloroplast. Photosynth. Res. 82, 221-240 (2004). [PubMed]

92. Willmund,F. & Schroda,M. HEAT SHOCK PROTEIN 90C is a bona fide Hsp90 that interacts with plastidic HSP70B in Chlamydomonas reinhardtii. Plant Physiol 138, 2310-2322 (2005). [PubMed]

93. Emelyanov,V.V. Phylogenetic relationships of organellar Hsp90 homologs reveal fundamental differences to organellar Hsp70 and Hsp60 evolution. Gene 299, 125-133 (2002). [PubMed]

94. Morita,T., Yamaguchi,H., Amagai,A., & Maeda,Y. Involvement of the TRAP-1 homologue, Dd-TRAP1, in spore differentiation during Dictyostelium development. Exp. Cell Res. 303, 425-431 (2005). [PubMed]

95. Minami,Y., Kimura,Y., Kawasaki,H., Suzuki,K., & Yahara,I. The carboxy-terminal region of mammalian HSP90 is required for its dimerization and function in vivo. Mol. Cell Biol. 14, 1459-1464 (1994). [PubMed]

96. Nemoto,T. & Sato,N. Oligomeric forms of the 90-kDa heat shock protein. Biochem. J. 330 ( Pt 2), 989-995 (1998). [PubMed]

97. Lanks,K.W. Temperature-dependent oligomerization of hsp85 in vitro. J. Cell Physiol 140, 601-607 (1989). [PubMed]

98. Minami,Y., Kawasaki,H., Miyata,Y., Suzuki,K., & Yahara,I. Analysis of native forms and isoform compositions of the mouse 90-kDa heat shock protein, HSP90. J. Biol. Chem. 266, 10099-10103 (1991). [PubMed]

99. Jakob,U., Lilie,H., Meyer,I., & Buchner,J. Transient interaction of Hsp90 with early unfolding intermediates of citrate synthase. Implications for heat shock in vivo. J. Biol. Chem. 270, 7288-7294 (1995). [PubMed]

100. Nemoto,T. et al. Dimerization characteristics of the 94-kDa glucose-regulated protein. J. Biochem. 120, 249-256 (1996). [PubMed]

101. Minami,Y., Kawasaki,H., Suzuki,K., & Yahara,I. The calmodulin-binding domain of the mouse 90-kDa heat shock protein. J. Biol. Chem. 268, 9604-9610 (1993). [PubMed]

102. Yonehara,M., Minami,Y., Kawata,Y., Nagai,J., & Yahara,I. Heat-induced chaperone activity of HSP90. J. Biol. Chem. 271, 2641-2645 (1996). [PubMed]

103. Nemoto,T.K., Ono,T., & Tanaka,K. Substrate-binding characteristics of proteins in the 90 kDa heat shock protein family. Biochem. J. 354, 663-670 (2001). [PubMed]

104. Whitesell,L., Shifrin,S.D., Schwab,G., & Neckers,L.M. Benzoquinonoid ansamycins possess selective tumoricidal activity unrelated to src kinase inhibition. Cancer Res. 52, 1721-1728 (1992). [PubMed]

105. Roe,S.M. et al. Structural basis for inhibition of the Hsp90 molecular chaperone by the antitumor antibiotics radicicol and geldanamycin. J. Med. Chem. 42, 260-266 (1999). [PubMed]

106. Pearl,L.H. & Prodromou,C. Structure and in vivo function of Hsp90. Curr. Opin. Struct. Biol. 10, 46-51 (2000). [PubMed]

107. Dixit,A. & Verkhivker,G.M. Probing molecular mechanisms of the Hsp90 chaperone: biophysical modeling identifies key regulators of functional dynamics. PLoS. ONE. 7, e37605 (2012). [PubMed]

108. Meyer,P. et al. Structural and functional analysis of the middle segment of hsp90: implications for ATP hydrolysis and client protein and cochaperone interactions. Mol. Cell 11, 647-658 (2003). [PubMed]

109. Sato,S., Fujita,N., & Tsuruo,T. Modulation of Akt kinase activity by binding to Hsp90. Proc. Natl. Acad. Sci. U. S. A 97, 10832-10837 (2000). [PubMed]

110. Muller,L., Schaupp,A., Walerych,D., Wegele,H., & Buchner,J. Hsp90 regulates the activity of wild type p53 under physiological and elevated temperatures. J. Biol. Chem. 279, 48846-48854 (2004). [PubMed]

111. Fontana,J. et al. Domain mapping studies reveal that the M domain of hsp90 serves as a molecular scaffold to regulate Akt-dependent phosphorylation of endothelial nitric oxide synthase and NO release. Circ. Res. 90, 866-873 (2002). [PubMed]

112. Richter,K., Muschler,P., Hainzl,O., Reinstein,J., & Buchner,J. Sti1 is a non-competitive inhibitor of the Hsp90 ATPase. Binding prevents the N-terminal dimerization reaction during the atpase cycle. J. Biol. Chem. 278, 10328-10333 (2003). [PubMed]

113. Terasawa,K., Minami,M., & Minami,Y. Constantly updated knowledge of Hsp90. J. Biochem. 137, 443-447 (2005). [PubMed]

114. Jackson,S.E., Queitsch,C., & Toft,D. Hsp90: from structure to phenotype. Nat. Struct. Mol. Biol. 11, 1152-1155 (2004). [PubMed]

115. Munro,S. & Pelham,H.R. A C-terminal signal prevents secretion of luminal ER proteins. Cell 48, 899-907 (1987). [PubMed]

116. Krukenberg,K.A., Böttcher,U.M., Southworth,D.R., & Agard,D.A. Grp94, the endoplasmic reticulum Hsp90, has a similar solution conformation to cytosolic Hsp90 in the absence of nucleotide. Protein Sci. 18, 1815-1827 (2009). [PubMed]

117. Krukenberg,K.A., Southworth,D.R., Street,T.O., & Agard,D.A. pH-dependent conformational changes in bacterial Hsp90 reveal a Grp94-like conformation at pH 6 that is highly active in suppression of citrate synthase aggregation. J. Mol. Biol. 390, 278-291 (2009). [PubMed]

118. Zuehlke,A. & Johnson,J.L. Hsp90 and co-chaperones twist the functions of diverse client proteins. Biopolymers 93, 211-217 (2010). PubMed]

119. Prodromou,C. et al. The ATPase cycle of Hsp90 drives a molecular ‘clamp’ via transient dimerization of the N-terminal domains. EMBO J. 19, 4383-4392 (2000). [PubMed]

120. Mickler,M., Hessling,M., Ratzke,C., Buchner,J., & Hugel,T. The large conformational changes of Hsp90 are only weakly coupled to ATP hydrolysis. Nat. Struct. Mol. Biol. 16, 281-286 (2009). [PubMed]

121. Blacklock,K. & Verkhivker,G.M. Allosteric regulation of the Hsp90 dynamics and stability by client recruiter cochaperones: protein structure network modeling. PLoS. ONE. 9, e86547 (2014). [PubMed]

122. Li,W., Sahu,D., & Tsen,F. Secreted heat shock protein-90 (Hsp90) in wound healing and cancer. Biochim. Biophys. Acta 1823, 730-741 (2012). [PubMed]

123. Tsutsumi,S. & Neckers,L. Extracellular heat shock protein 90: a role for a molecular chaperone in cell motility and cancer metastasis. Cancer Sci. 98, 1536-1539 (2007). [PubMed]

124. Saito,K., Dai,Y., & Ohtsuka,K. Enhanced expression of heat shock proteins in gradually dying cells and their release from necrotically dead cells. Exp. Cell Res. 310, 229-236 (2005). [PubMed]

125. Berwin,B., Reed,R.C., & Nicchitta,C.V. Virally induced lytic cell death elicits the release of immunogenic GRP94/gp96. J. Biol. Chem. 276, 21083-21088 (2001). [PubMed]

126. Basu,S., Binder,R.J., Suto,R., Anderson,K.M., & Srivastava,P.K. Necrotic but not apoptotic cell death releases heat shock proteins, which deliver a partial maturation signal to dendritic cells and activate the NF-kappa B pathway. Int. Immunol. 12, 1539-1546 (2000). [PubMed]

127. Hightower,L.E. & Guidon,P.T., Jr. Selective release from cultured mammalian cells of heat-shock (stress) proteins that resemble glia-axon transfer proteins. J. Cell Physiol 138, 257-266 (1989). [PubMed]

128. Clayton,A., Turkes,A., Navabi,H., Mason,M.D., & Tabi,Z. Induction of heat shock proteins in B-cell exosomes. J. Cell Sci. 118, 3631-3638 (2005). [PubMed]

129. Liao,D.F. et al. Purification and identification of secreted oxidative stress-induced factors from vascular smooth muscle cells. J. Biol. Chem. 275, 189-196 (2000). [PubMed]

130. Li,W. et al. Extracellular heat shock protein-90alpha: linking hypoxia to skin cell motility and wound healing. EMBO J. 26, 1221-1233 (2007). [PubMed]

131. Tsutsumi,S. et al. A small molecule cell-impermeant Hsp90 antagonist inhibits tumor cell motility and invasion. Oncogene 27, 2478-2487 (2008). [PubMed]

132. Sahu,D. et al. A potentially common peptide target in secreted heat shock protein-90alpha for hypoxia-inducible factor-1alpha-positive tumors. Mol. Biol. Cell 23, 602-613 (2012). [PubMed]

133. Chen,J.S. et al. Secreted heat shock protein 90alpha induces colorectal cancer cell invasion through CD91/LRP-1 and NF-kappaB-mediated integrin alphaV expression. J. Biol. Chem. 285, 25458-25466 (2010). [PubMed]

134. Suzuki,S. & Kulkarni,A.B. Extracellular heat shock protein HSP90beta secreted by MG63 osteosarcoma cells inhibits activation of latent TGF-beta1. Biochem. Biophys. Res. Commun. 398, 525-531 (2010). [PubMed]

135. Hegmans,J.P. et al. Proteomic analysis of exosomes secreted by human mesothelioma cells. Am. J. Pathol. 164, 1807-1815 (2004). [PubMed]

136. Mignot,G., Roux,S., Thery,C., Segura,E., & Zitvogel,L. Prospects for exosomes in immunotherapy of cancer. J. Cell Mol. Med. 10, 376-388 (2006). [PubMed]

137. Savina,A., Furlan,M., Vidal,M., & Colombo,M.I. Exosome release is regulated by a calcium-dependent mechanism in K562 cells. J. Biol. Chem. 278, 20083-20090 (2003). [PubMed]

138. Lancaster,G.I. & Febbraio,M.A. Exosome-dependent trafficking of HSP70: a novel secretory pathway for cellular stress proteins. J. Biol. Chem. 280, 23349-23355 (2005). [PubMed]

139. Nickel,W. The mystery of nonclassical protein secretion. A current view on cargo proteins and potential export routes. Eur. J. Biochem. 270, 2109-2119 (2003). [PubMed]

140. Cheng,C.F. et al. Transforming growth factor alpha (TGFalpha)-stimulated secretion of HSP90alpha: using the receptor LRP-1/CD91 to promote human skin cell migration against a TGFbeta-rich environment during wound healing. Mol. Cell Biol. 28, 3344-3358 (2008). [PubMed]

141. Wayne,N., Mishra,P., & Bolon,D.N. Hsp90 and client protein maturation. Methods Mol. Biol. 787, 33-44 (2011). [PubMed]

142. Taipale,M. et al. Quantitative analysis of HSP90-client interactions reveals principles of substrate recognition. Cell 150, 987-1001 (2012). [PubMed]

143. Young,J.C., Agashe,V.R., Siegers,K., & Hartl,F.U. Pathways of chaperone-mediated protein folding in the cytosol. Nat. Rev. Mol. Cell Biol. 5, 781-791 (2004). [PubMed]

144. Zhao,R. et al. Navigating the chaperone network: an integrative map of physical and genetic interactions mediated by the hsp90 chaperone. Cell 120, 715-727 (2005). [PubMed]

145. Zhao,R. & Houry,W.A. Hsp90: a chaperone for protein folding and gene regulation. Biochem. Cell Biol. 83, 703-710 (2005). [PubMed]

146. Pearl,L.H. & Prodromou,C. Structure, function, and mechanism of the Hsp90 molecular chaperone. Adv. Protein Chem. 59, 157-186 (2001). [PubMed]

147. Picard,D. Heat-shock protein 90, a chaperone for folding and regulation. Cell Mol. Life Sci. 59, 1640-1648 (2002). [PubMed]

148. Cheng,C.F., Fan,J., Zhao,Z., Woodley,D.T., & Li,W. Secreted heat shock protein-90 α : a more effective and safer target for anti-cancer drugs? Curr. Signal Transduct. Ther. 5, 121-127 (2010). [DOI: 10.2174/157436210791112208]

149. Caplan,A.J., Jackson,S., & Smith,D. Hsp90 reaches new heights. Conference on the Hsp90 chaperone machine. EMBO Rep. 4, 126-130 (2003). [PubMed]

150. Wegele,H., Müller,L., & Buchner,J. Hsp70 and Hsp90–a relay team for protein folding. Rev. Physiol Biochem. Pharmacol. 151, 1-44 (2004). [PubMed]

151. Zhao,R. & Houry,W.A. Molecular interaction network of the Hsp90 chaperone system. Adv. Exp. Med. Biol. 594, 27-36 (2007). [PubMed]

152. Workman,P. Altered states: selectively drugging the Hsp90 cancer chaperone. Trends Mol. Med. 10, 47-51 (2004). [PubMed]

153. Xu,W. & Neckers,L. Targeting the molecular chaperone heat shock protein 90 provides a multifaceted effect on diverse cell signaling pathways of cancer cells. Clin. Cancer Res. 13, 1625-1629 (2007). [PubMed]

154. Kamal,A. et al. A high-affinity conformation of Hsp90 confers tumour selectivity on Hsp90 inhibitors. Nature 425, 407-410 (2003). [PubMed]

155. Chen,Z. et al. Inhibition of ALK, PI3K/MEK, and HSP90 in murine lung adenocarcinoma induced by EML4-ALK fusion oncogene. Cancer Res. 70, 9827-9836 (2010). [PubMed]

156. Banilas,G., Korkas,E., Englezos,V., Nisiotou,A.A., & Hatzopoulos,P. Genome-wide analysis of the heat shock protein 90 gene family in grapevine (Vitis vinifera L.). Aust. J. Grape Wine Res. 18, 29-38 (2012). [DOI: 10.1111/j.1755-0238.2011.00166.x]

157. Zhang,L., Fan,Y., Shi,F., Qin,S., & Liu,B. Molecular cloning, characterization, and expression analysis of a cytosolic HSP90 gene from Haematococcus pluvialis. J. Appl. Phycol. 24, 1601-1612 (2012). [DOI: 10.1007/s10811-012-9821-5]

158. Wójcik,M. & Tukiendorf,A. Glutathione in adaptation of Arabidopsis thaliana to cadmium stress. Biol. Plant. 55, 125-132 (2011). [DOI: 10.1007/s10535-011-0017-7]

159. Maksymiec,W. Effects of jasmonate and some other signalling factors on bean and onion growth during the initial phase of cadmium action. Biol. Plant. 55, 112-118 (2011). [DOI: 10.1007/s10535-011-0015-9]

160. Grigorova,B., Vaseva,I., Demirevska,K., & Feller,U. Combined drought and heat stress in wheat: changes in some heat shock proteins. Biol. Plant. 55, 105-111 (2011). [DOI: 10.1007/s10535-011-0014-x]

161. Jarosz,D.F. & Lindquist,S. Hsp90 and environmental stress transform the adaptive value of natural genetic variation. Science 330, 1820-1824 (2010). [PubMed]

162. Rizhsky,L., Liang,H., & Mittler,R. The combined effect of drought stress and heat shock on gene expression in tobacco. Plant Physiol 130, 1143-1151 (2002). [PubMed]

163. Cano,L.Q., Lavery,D.N., & Bevan,C.L. Mini-review: Foldosome regulation of androgen receptor action in prostate cancer. Mol. Cell Endocrinol. 369, 52-62 (2013). [PubMed]

164. Gelmann,E.P. Molecular biology of the androgen receptor. J. Clin. Oncol. 20, 3001-3015 (2002). [PubMed]

165. Lavery,D.N. & McEwan,I.J. Structure and function of steroid receptor AF1 transactivation domains: induction of active conformations. Biochem. J. 391, 449-464 (2005). [PubMed]

166. Pratt,W.B. & Toft,D.O. Regulation of signaling protein function and trafficking by the hsp90/hsp70-based chaperone machinery. Exp. Biol. Med. (Maywood. ) 228, 111-133 (2003). [PubMed]

167. Hernandez,M.P., Chadli,A., & Toft,D.O. HSP40 binding is the first step in the HSP90 chaperoning pathway for the progesterone receptor. J. Biol. Chem. 277, 11873-11881 (2002). [PubMed]

168. Hernandez,M.P., Sullivan,W.P., & Toft,D.O. The assembly and intermolecular properties of the hsp70-Hop-hsp90 molecular chaperone complex. J. Biol. Chem. 277, 38294-38304 (2002). [PubMed]

169. Schmid,A.B. et al. The architecture of functional modules in the Hsp90 co-chaperone Sti1/Hop. EMBO J. 31, 1506-1517 (2012). [PubMed]

170. Gaiser,A.M., Brandt,F., & Richter,K. The non-canonical Hop protein from Caenorhabditis elegans exerts essential functions and forms binary complexes with either Hsc70 or Hsp90. J. Mol. Biol. 391, 621-634 (2009). [PubMed]

171. Ebong,I.O. et al. Heterogeneity and dynamics in the assembly of the heat shock protein 90 chaperone complexes. Proc. Natl. Acad. Sci. U. S. A 108, 17939-17944 (2011). [PubMed]

172. Li,J., Richter,K., & Buchner,J. Mixed Hsp90-cochaperone complexes are important for the progression of the reaction cycle. Nat. Struct. Mol. Biol. 18, 61-66 (2011). [PubMed]

173. Morishima,Y. et al. The hsp90 cochaperone p23 is the limiting component of the multiprotein hsp90/hsp70-based chaperone system in vivo where it acts to stabilize the client protein: hsp90 complex. J. Biol. Chem. 278, 48754-48763 (2003). [PubMed]

174. Harst,A., Lin,H., & Obermann,W.M. Aha1 competes with Hop, p50 and p23 for binding to the molecular chaperone Hsp90 and contributes to kinase and hormone receptor activation. Biochem. J. 387, 789-796 (2005). [PubMed]

175. Pratt,W.B. & Toft,D.O. Steroid receptor interactions with heat shock protein and immunophilin chaperones. Endocr. Rev. 18, 306-360 (1997). [PubMed]

176. Simons,S.S., Jr., Sistare,F.D., & Chakraborti,P.K. Steroid binding activity is retained in a 16-kDa fragment of the steroid binding domain of rat glucocorticoid receptors. J. Biol. Chem. 264, 14493-14497 (1989). [PubMed]

177. Bledsoe,R.K. et al. Crystal structure of the glucocorticoid receptor ligand binding domain reveals a novel mode of receptor dimerization and coactivator recognition. Cell 110, 93-105 (2002). [PubMed]

178. Vandevyver,S., Dejager,L., & Libert,C. On the trail of the glucocorticoid receptor: into the nucleus and back. Traffic. 13, 364-374 (2012). [PubMed]

179. Galigniana,M.D., Echeverria,P.C., Erlejman,A.G., & Piwien-Pilipuk,G. Role of molecular chaperones and TPR-domain proteins in the cytoplasmic transport of steroid receptors and their passage through the nuclear pore. Nucleus. 1, 299-308 (2010). [PubMed]

180. Echeverria,P.C. & Picard,D. Molecular chaperones, essential partners of steroid hormone receptors for activity and mobility. Biochim. Biophys. Acta 1803, 641-649 (2010). [PubMed]

181. Smith,D.F. Chaperones in progesterone receptor complexes. Semin. Cell Dev. Biol. 11, 45-52 (2000). [PubMed]

182. Heinlein,C.A. & Chang,C. Role of chaperones in nuclear translocation and transactivation of steroid receptors. Endocrine. 14, 143-149 (2001). [PubMed]

183. Davies,T.H., Ning,Y.M., & Sanchez,E.R. A new first step in activation of steroid receptors: hormone-induced switching of FKBP51 and FKBP52 immunophilins. J. Biol. Chem. 277, 4597-4600 (2002). [PubMed]

184. Makhnevych,T. & Houry,W.A. The role of Hsp90 in protein complex assembly. Biochim. Biophys. Acta 1823, 674-682 (2012). [PubMed]

185. Yamano,T. et al. Hsp90-mediated assembly of the 26 S proteasome is involved in major histocompatibility complex class I antigen processing. J. Biol. Chem. 283, 28060-28065 (2008). [PubMed]

186. Imai,J., Maruya,M., Yashiroda,H., Yahara,I., & Tanaka,K. The molecular chaperone Hsp90 plays a role in the assembly and maintenance of the 26S proteasome. EMBO J. 22, 3557-3567 (2003). [PubMed]

187. Srivastava,P.K., Menoret,A., Basu,S., Binder,R.J., & McQuade,K.L. Heat shock proteins come of age: primitive functions acquire new roles in an adaptive world. Immunity. 8, 657-665 (1998). [PubMed]

188. Henderson,B. & Kaiser,F. Do reciprocal interactions between cell stress proteins and cytokines create a new intra-/extra-cellular signalling nexus? Cell Stress. Chaperones. 18, 685-701 (2013). [PubMed]

189. Stephanou,A. et al. Cardiotrophin-1 induces heat shock protein accumulation in cultured cardiac cells and protects them from stressful stimuli. J. Mol. Cell Cardiol. 30, 849-855 (1998). [PubMed]

190. Wax,S., Piecyk,M., Maritim,B., & Anderson,P. Geldanamycin inhibits the production of inflammatory cytokines in activated macrophages by reducing the stability and translation of cytokine transcripts. Arthritis Rheum. 48, 541-550 (2003). [PubMed]

191. Broemer,M., Krappmann,D., & Scheidereit,C. Requirement of Hsp90 activity for IkappaB kinase (IKK) biosynthesis and for constitutive and inducible IKK and NF-kappaB activation. Oncogene 23, 5378-5386 (2004). [PubMed]

192. Salminen,A., Paimela,T., Suuronen,T., & Kaarniranta,K. Innate immunity meets with cellular stress at the IKK complex: regulation of the IKK complex by HSP70 and HSP90. Immunol. Lett. 117, 9-15 (2008). [PubMed]

193. Shimp,S.K., III et al. HSP90 inhibition by 17-DMAG reduces inflammation in J774 macrophages through suppression of Akt and nuclear factor-kappaB pathways. Inflamm. Res. 61, 521-533 (2012). [PubMed]

194. Zeng,Y. et al. Natural killer cells play a key role in the antitumor immunity generated by chaperone-rich cell lysate vaccination. Int. J. Cancer 119, 2624-2631 (2006). [PubMed]

195. Dowling,P., Walsh,N., & Clynes,M. Membrane and membrane-associated proteins involved in the aggressive phenotype displayed by highly invasive cancer cells. Proteomics. 8, 4054-4065 (2008). [PubMed]

196. Taiyab,A. & Rao,C. HSP90 modulates actin dynamics: inhibition of HSP90 leads to decreased cell motility and impairs invasion. Biochim. Biophys. Acta 1813, 213-221 (2011). [PubMed]

197. Amano,M. et al. Formation of actin stress fibers and focal adhesions enhanced by Rho-kinase. Science 275, 1308-1311 (1997). [PubMed]

198. Watanabe,N. et al. p140mDia, a mammalian homolog of Drosophila diaphanous, is a target protein for Rho small GTPase and is a ligand for profilin. EMBO J. 16, 3044-3056 (1997). [PubMed]

199. Cheng,C.F. et al. A fragment of secreted Hsp90alpha carries properties that enable it to accelerate effectively both acute and diabetic wound healing in mice. J. Clin. Invest 121, 4348-4361 (2011). [PubMed]

200. Eustace,B.K. et al. Functional proteomic screens reveal an essential extracellular role for hsp90 alpha in cancer cell invasiveness. Nat. Cell Biol. 6, 507-514 (2004). [PubMed]

201. Stellas,D., El,H.A., & Patsavoudi,E. Monoclonal antibody 4C5 prevents activation of MMP2 and MMP9 by disrupting their interaction with extracellular HSP90 and inhibits formation of metastatic breast cancer cell deposits. BMC. Cell Biol. 11, 51 (2010). [PubMed]

202. Sidera,K., Gaitanou,M., Stellas,D., Matsas,R., & Patsavoudi,E. A critical role for HSP90 in cancer cell invasion involves interaction with the extracellular domain of HER-2. J. Biol. Chem. 283, 2031-2041 (2008). [PubMed]

203. Schiller,P. et al. Cis-acting elements involved in the regulated expression of a human HSP70 gene. J. Mol. Biol. 203, 97-105 (1988). [PubMed]

204. Akerfelt,M., Morimoto,R.I., & Sistonen,L. Heat shock factors: integrators of cell stress, development and lifespan. Nat. Rev. Mol. Cell Biol. 11, 545-555 (2010). [PubMed]

205. Pirkkala,L., Nykanen,P., & Sistonen,L. Roles of the heat shock transcription factors in regulation of the heat shock response and beyond. FASEB J. 15, 1118-1131 (2001). [PubMed]

206. Harrison,C.J., Bohm,A.A., & Nelson,H.C. Crystal structure of the DNA binding domain of the heat shock transcription factor. Science 263, 224-227 (1994). [PubMed]

207. Zorzi,E. & Bonvini,P. Inducible hsp70 in the regulation of cancer cell survival: analysis of chaperone induction, expression and activity. Cancers. (Basel) 3, 3921-3956 (2011). [PubMed]

208. Ali,A., Bharadwaj,S., O’Carroll,R., & Ovsenek,N. HSP90 interacts with and regulates the activity of heat shock factor 1 in Xenopus oocytes. Mol. Cell Biol. 18, 4949-4960 (1998). [PubMed]

209. Calderwood,S.K. et al. Signal Transduction Pathways Leading to Heat Shock Transcription. Sign. Transduct. Insights. 2, 13-24 (2010). [PubMed]

210. Leach,M.D., Tyc,K.M., Brown,A.J., & Klipp,E. Modelling the regulation of thermal adaptation in Candida albicans, a major fungal pathogen of humans. PLoS. ONE. 7, e32467 (2012). [PubMed]

211. Guo,Y. et al. Evidence for a mechanism of repression of heat shock factor 1 transcriptional activity by a multichaperone complex. J. Biol. Chem. 276, 45791-45799 (2001). [PubMed]

212. Baler,R., Welch,W.J., & Voellmy,R. Heat shock gene regulation by nascent polypeptides and denatured proteins: hsp70 as a potential autoregulatory factor. J. Cell Biol. 117, 1151-1159 (1992). [PubMed]

213. Abravaya,K., Myers,M.P., Murphy,S.P., & Morimoto,R.I. The human heat shock protein hsp70 interacts with HSF, the transcription factor that regulates heat shock gene expression. Genes Dev. 6, 1153-1164 (1992). [PubMed]

214. Shi,Y., Mosser,D.D., & Morimoto,R.I. Molecular chaperones as HSF1-specific transcriptional repressors. Genes Dev. 12, 654-666 (1998). [PubMed]

215. Westerheide,S.D., Anckar,J., Stevens,S.M., Jr., Sistonen,L., & Morimoto,R.I. Stress-inducible regulation of heat shock factor 1 by the deacetylase SIRT1. Science 323, 1063-1066 (2009). [PubMed]

216. Hietakangas,V. et al. Phosphorylation of serine 303 is a prerequisite for the stress-inducible SUMO modification of heat shock factor 1. Mol. Cell Biol. 23, 2953-2968 (2003). [PubMed]

217. Tang,D. et al. Expression of heat shock proteins and heat shock protein messenger ribonucleic acid in human prostate carcinoma in vitro and in tumors in vivo. Cell Stress. Chaperones. 10, 46-58 (2005). [PubMed]

218. Radons,J. Inflammatory stress and sarcomagenesis: a vicious interplay. Cell Stress. Chaperones. 19, 1-13 (2014). [PubMed]

219. Visone,R. & Croce,C.M. MiRNAs and cancer. Am. J. Pathol. 174, 1131-1138 (2009). [PubMed]

220. Spizzo,R., Nicoloso,M.S., Croce,C.M., & Calin,G.A. SnapShot: MicroRNAs in Cancer. Cell 137, 586 (2009). [PubMed]

221. Li,G. et al. Heat shock protein 90B1 plays an oncogenic role and is a target of microRNA-223 in human osteosarcoma. Cell Physiol Biochem. 30, 1481-1490 (2012). [PubMed]

222. Kariya,A., Furusawa,Y., Yunoki,T., Kondo,T., & Tabuchi,Y. A microRNA-27a mimic sensitizes human oral squamous cell carcinoma HSC-4 cells to hyperthermia through downregulation of Hsp110 and Hsp90. Int. J. Mol. Med. 34, 334-340 (2014). [PubMed]

223. Matkovich,S.J., Hu,Y., Eschenbacher,W.H., Dorn,L.E., & Dorn,G.W. Direct and indirect involvement of microRNA-499 in clinical and experimental cardiomyopathy. Circ. Res. 111, 521-531 (2012). [PubMed]

224. Akira,S. et al. Molecular cloning of APRF, a novel IFN-stimulated gene factor 3 p91-related transcription factor involved in the gp130-mediated signaling pathway. Cell 77, 63-71 (1994). [PubMed]

225. Stephanou,A., Isenberg,D.A., Akira,S., Kishimoto,T., & Latchman,D.S. The nuclear factor interleukin-6 (NF-IL6) and signal transducer and activator of transcription-3 (STAT-3) signalling pathways co-operate to mediate the activation of the hsp90beta gene by interleukin-6 but have opposite effects on its inducibility by heat shock. Biochem. J. 330 ( Pt 1), 189-195 (1998). [PubMed]

226. Stephanou,A., Isenberg,D.A., Nakajima,K., & Latchman,D.S. Signal transducer and activator of transcription-1 and heat shock factor-1 interact and activate the transcription of the Hsp-70 and Hsp-90beta gene promoters. J. Biol. Chem. 274, 1723-1728 (1999). [PubMed]

227. Wang,Y., Chen,L., Hagiwara,N., & Knowlton,A.A. Regulation of heat shock protein 60 and 72 expression in the failing heart. J. Mol. Cell Cardiol. 48, 360-366 (2010). [PubMed]

228. Cappello,F. et al. Convergent sets of data from in vivo and in vitro methods point to an active role of Hsp60 in chronic obstructive pulmonary disease pathogenesis. PLoS. ONE. 6, e28200 (2011). [PubMed]

229. Guzhova,I.V., Darieva,Z.A., Melo,A.R., & Margulis,B.A. Major stress protein Hsp70 interacts with NF-kB regulatory complex in human T-lymphoma cells. Cell Stress. Chaperones. 2, 132-139 (1997). [PubMed]

230. Ammirante,M. et al. The activity of hsp90 alpha promoter is regulated by NF-kappa B transcription factors. Oncogene 27, 1175-1178 (2008). [PubMed]

231. Rappa,F. et al. HSP-molecular chaperones in cancer biogenesis and tumor therapy: an overview. Anticancer Res. 32, 5139-5150 (2012). [PubMed]

232. Mollapour,M. & Neckers,L. Post-translational modifications of Hsp90 and their contributions to chaperone regulation. Biochim. Biophys. Acta 1823, 648-655 (2012). [PubMed]

233. Mollapour,M. et al. Asymmetric Hsp90 N domain SUMOylation recruits Aha1 and ATP-competitive inhibitors. Mol. Cell 53, 317-329 (2014). [PubMed]

234. Overath,T. et al. Mapping of O-GlcNAc sites of 20 S proteasome subunits and Hsp90 by a novel biotin-cystamine tag. Mol. Cell Proteomics. 11, 467-477 (2012). [PubMed]

235. Dougherty,J.J., Rabideau,D.A., Iannotti,A.M., Sullivan,W.P., & Toft,D.O. Identification of the 90 kDa substrate of rat liver type II casein kinase with the heat shock protein which binds steroid receptors. Biochim. Biophys. Acta 927, 74-80 (1987). [PubMed]

236. Dougherty,J.J., Puri,R.K., & Toft,D.O. Phosphorylation in vivo of chicken oviduct progesterone receptor. J. Biol. Chem. 257, 14226-14230 (1982). [PubMed]

237. Mollapour,M. et al. Threonine 22 phosphorylation attenuates Hsp90 interaction with cochaperones and affects its chaperone activity. Mol. Cell 41, 672-681 (2011). [PubMed]

238. Mollapour,M., Tsutsumi,S., Kim,Y.S., Trepel,J., & Neckers,L. Casein kinase 2 phosphorylation of Hsp90 threonine 22 modulates chaperone function and drug sensitivity. Oncotarget. 2, 407-417 (2011). [PubMed]

239. Mimnaugh,E.G., Worland,P.J., Whitesell,L., & Neckers,L.M. Possible role for serine/threonine phosphorylation in the regulation of the heteroprotein complex between the hsp90 stress protein and the pp60v-src tyrosine kinase. J. Biol. Chem. 270, 28654-28659 (1995). [PubMed]

240. Wandinger,S.K., Suhre,M.H., Wegele,H., & Buchner,J. The phosphatase Ppt1 is a dedicated regulator of the molecular chaperone Hsp90. EMBO J. 25, 367-376 (2006). [PubMed]

241. Zhang,Y., Leung,D.Y., Nordeen,S.K., & Goleva,E. Estrogen inhibits glucocorticoid action via protein phosphatase 5 (PP5)-mediated glucocorticoid receptor dephosphorylation. J. Biol. Chem. 284, 24542-24552 (2009). [PubMed]

242. Kurokawa,M., Zhao,C., Reya,T., & Kornbluth,S. Inhibition of apoptosome formation by suppression of Hsp90beta phosphorylation in tyrosine kinase-induced leukemias. Mol. Cell Biol. 28, 5494-5506 (2008). [PubMed]

243. Lei,H., Venkatakrishnan,A., Yu,S., & Kazlauskas,A. Protein kinase A-dependent translocation of Hsp90 alpha impairs endothelial nitric-oxide synthase activity in high glucose and diabetes. J. Biol. Chem. 282, 9364-9371 (2007). [PubMed]

244. Lees-Miller,S.P. & Anderson,C.W. Two human 90-kDa heat shock proteins are phosphorylated in vivo at conserved serines that are phosphorylated in vitro by casein kinase II. J. Biol. Chem. 264, 2431-2437 (1989). [PubMed]

245. Barati,M.T., Rane,M.J., Klein,J.B., & McLeish,K.R. A proteomic screen identified stress-induced chaperone proteins as targets of Akt phosphorylation in mesangial cells. J. Proteome. Res. 5, 1636-1646 (2006). [PubMed]

246. Old,W.M. et al. Functional proteomics identifies targets of phosphorylation by B-Raf signaling in melanoma. Mol. Cell 34, 115-131 (2009). [PubMed]

247. Lees-Miller,S.P. & Anderson,C.W. The human double-stranded DNA-activated protein kinase phosphorylates the 90-kDa heat-shock protein, hsp90 alpha at two NH2-terminal threonine residues. J. Biol. Chem. 264, 17275-17280 (1989). [PubMed]

248. Mollapour,M. et al. Swe1Wee1-dependent tyrosine phosphorylation of Hsp90 regulates distinct facets of chaperone function. Mol. Cell 37, 333-343 (2010). [PubMed]

249. Mollapour,M., Tsutsumi,S., & Neckers,L. Hsp90 phosphorylation, Wee1 and the cell cycle. Cell Cycle 9, 2310-2316 (2010). [PubMed]

250. Zhang,Y. et al. Mice lacking histone deacetylase 6 have hyperacetylated tubulin but are viable and develop normally. Mol. Cell Biol. 28, 1688-1701 (2008). [PubMed]

251. Yang,Y. et al. Role of acetylation and extracellular location of heat shock protein 90alpha in tumor cell invasion. Cancer Res. 68, 4833-4842 (2008). [PubMed]

252. Bali,P. et al. Inhibition of histone deacetylase 6 acetylates and disrupts the chaperone function of heat shock protein 90: a novel basis for antileukemia activity of histone deacetylase inhibitors. J. Biol. Chem. 280, 26729-26734 (2005). [PubMed]

253. Kovacs,J.J. et al. HDAC6 regulates Hsp90 acetylation and chaperone-dependent activation of glucocorticoid receptor. Mol. Cell 18, 601-607 (2005). [PubMed]

254. Murphy,P.J., Morishima,Y., Kovacs,J.J., Yao,T.P., & Pratt,W.B. Regulation of the dynamics of hsp90 action on the glucocorticoid receptor by acetylation/deacetylation of the chaperone. J. Biol. Chem. 280, 33792-33799 (2005). [PubMed]

255. Scroggins,B.T. et al. An acetylation site in the middle domain of Hsp90 regulates chaperone function. Mol. Cell 25, 151-159 (2007). [PubMed]

256. Aoyagi,S. & Archer,T.K. Modulating molecular chaperone Hsp90 functions through reversible acetylation. Trends Cell Biol. 15, 565-567 (2005). [PubMed]

257. Lee,S.M. et al. Bcr-Abl-independent imatinib-resistant K562 cells show aberrant protein acetylation and increased sensitivity to histone deacetylase inhibitors. J. Pharmacol. Exp. Ther. 322, 1084-1092 (2007). [PubMed]

258. Park,J.H. et al. Class II histone deacetylases play pivotal roles in heat shock protein 90-mediated proteasomal degradation of vascular endothelial growth factor receptors. Biochem. Biophys. Res. Commun. 368, 318-322 (2008). [PubMed]

259. Nishioka,C. et al. MS-275, a novel histone deacetylase inhibitor with selectivity against HDAC1, induces degradation of FLT3 via inhibition of chaperone function of heat shock protein 90 in AML cells. Leuk. Res. 32, 1382-1392 (2008). [PubMed]

260. Zhou,Q., Agoston,A.T., Atadja,P., Nelson,W.G., & Davidson,N.E. Inhibition of histone deacetylases promotes ubiquitin-dependent proteasomal degradation of DNA methyltransferase 1 in human breast cancer cells. Mol. Cancer Res. 6, 873-883 (2008). [PubMed]

261. Retzlaff,M. et al. Hsp90 is regulated by a switch point in the C-terminal domain. EMBO Rep. 10, 1147-1153 (2009). [PubMed]

262. Zhang,H.H., Wang,Y.P., & Chen,D.B. Analysis of nitroso-proteomes in normotensive and severe preeclamptic human placentas. Biol. Reprod. 84, 966-975 (2011). [PubMed]

263. Martinez-Ruiz,A. et al. S-nitrosylation of Hsp90 promotes the inhibition of its ATPase and endothelial nitric oxide synthase regulatory activities. Proc. Natl. Acad. Sci. U. S. A 102, 8525-8530 (2005). [PubMed]

264. Piwien-Pilipuk,G., Ayala,A., Machado,A., & Galigniana,M.D. Impairment of mineralocorticoid receptor (MR)-dependent biological response by oxidative stress and aging: correlation with post-translational modification of MR and decreased ADP-ribosylatable level of elongating factor 2 in kidney cells. J. Biol. Chem. 277, 11896-11903 (2002). [PubMed]

265. Piwien-Pilipuk,G. & Galigniana,M.D. Oxidative stress induced by L-buthionine-(S,R)-sulfoximine, a selective inhibitor of glutathione metabolism, abrogates mouse kidney mineralocorticoid receptor function. Biochim. Biophys. Acta 1495, 263-280 (2000). [PubMed]

266. Stancato,L.F., Hutchison,K.A., Chakraborti,P.K., Simons,S.S., Jr., & Pratt,W.B. Differential effects of the reversible thiol-reactive agents arsenite and methyl methanethiosulfonate on steroid binding by the glucocorticoid receptor. Biochemistry 32, 3729-3736 (1993). [PubMed]

267. Okamoto,K. et al. Redox-dependent regulation of nuclear import of the glucocorticoid receptor. J. Biol. Chem. 274, 10363-10371 (1999). [PubMed]

268. Chen,W.Y. et al. Tubocapsenolide A, a novel withanolide, inhibits proliferation and induces apoptosis in MDA-MB-231 cells by thiol oxidation of heat shock proteins. J. Biol. Chem. 283, 17184-17193 (2008). [PubMed]

269. Abu-Farha,M. et al. Proteomic analyses of the SMYD family interactomes identify HSP90 as a novel target for SMYD2. J. Mol. Cell Biol. 3, 301-308 (2011). [PubMed]

270. Donlin,L.T. et al. Smyd2 controls cytoplasmic lysine methylation of Hsp90 and myofilament organization. Genes Dev. 26, 114-119 (2012). [PubMed]

271. Franco,M.C. et al. Nitration of Hsp90 induces cell death. Proc. Natl. Acad. Sci. U. S. A 110, E1102-E1111 (2013). [PubMed]

272. Sarkar,A.A. & Zohn,I.E. Hectd1 regulates intracellular localization and secretion of Hsp90 to control cellular behavior of the cranial mesenchyme. J. Cell Biol. 196, 789-800 (2012). [PubMed]

273. Kundrat,L. & Regan,L. Identification of residues on Hsp70 and Hsp90 ubiquitinated by the cochaperone CHIP. J. Mol. Biol. 395, 587-594 (2010). [PubMed]

274. Echeverria,P.C., Bernthaler,A., Dupuis,P., Mayer,B., & Picard,D. An interaction network predicted from public data as a discovery tool: application to the Hsp90 molecular chaperone machine. PLoS. ONE. 6, e26044 (2011). [PubMed]

275. Samant,R.S., Clarke,P.A., & Workman,P. The expanding proteome of the molecular chaperone HSP90. Cell Cycle 11, 1301-1308 (2012). [PubMed]

276. Erlejman,A.G., Lagadari,M., Toneatto,J., Piwien-Pilipuk,G., & Galigniana,M.D. Regulatory role of the 90-kDa-heat-shock protein (Hsp90) and associated factors on gene expression. Biochim. Biophys. Acta 1839, 71-87 (2014). [PubMed]

277. Pearl,L.H., Prodromou,C., & Workman,P. The Hsp90 molecular chaperone: an open and shut case for treatment. Biochem. J. 410, 439-453 (2008). [PubMed]

278. McClellan,A.J. et al. Diverse cellular functions of the Hsp90 molecular chaperone uncovered using systems approaches. Cell 131, 121-135 (2007). [PubMed]

279. Prodromou,C. et al. Regulation of Hsp90 ATPase activity by tetratricopeptide repeat (TPR)-domain co-chaperones. EMBO J. 18, 754-762 (1999). [PubMed]

280. Siligardi,G. et al. Regulation of Hsp90 ATPase activity by the co-chaperone Cdc37p/p50cdc37. J. Biol. Chem. 277, 20151-20159 (2002). [PubMed]

281. Eckl,J.M. & Richter,K. Functions of the Hsp90 chaperone system: lifting client proteins to new heights. Int. J. Biochem. Mol. Biol. 4, 157-165 (2013). [PubMed]

282. Young,J.C. & Hartl,F.U. Polypeptide release by Hsp90 involves ATP hydrolysis and is enhanced by the co-chaperone p23. EMBO J. 19, 5930-5940 (2000). [PubMed]

283. Trepel,J., Mollapour,M., Giaccone,G., & Neckers,L. Targeting the dynamic HSP90 complex in cancer. Nat. Rev. Cancer 10, 537-549 (2010). [PubMed]

284. Zhang,H. & Burrows,F. Targeting multiple signal transduction pathways through inhibition of Hsp90. J. Mol. Med. (Berl) 82, 488-499 (2004). [PubMed]

285. Maloney,A. & Workman,P. HSP90 as a new therapeutic target for cancer therapy: the story unfolds. Expert. Opin. Biol. Ther. 2, 3-24 (2002). [PubMed]

286. Sumimoto,H., Imabayashi,F., Iwata,T., & Kawakami,Y. The BRAF-MAPK signaling pathway is essential for cancer-immune evasion in human melanoma cells. J. Exp. Med. 203, 1651-1656 (2006). [PubMed]

287. Banerji,U., Affolter,A., Judson,I., Marais,R., & Workman,P. BRAF and NRAS mutations in melanoma: potential relationships to clinical response to HSP90 inhibitors. Mol. Cancer Ther. 7, 737-739 (2008). [PubMed]

288. Calapre,L., Gray,E.S., & Ziman,M. Heat stress: a risk factor for skin carcinogenesis. Cancer Lett. 337, 35-40 (2013). [PubMed]

289. Marquette,A., Bagot,M., Bensussan,A., & Dumaz,N. Recent discoveries in the genetics of melanoma and their therapeutic implications. Arch. Immunol. Ther. Exp. (Warsz. ) 55, 363-372 (2007). [PubMed]

290. Benjamin,C.L., Ullrich,S.E., Kripke,M.L., & Ananthaswamy,H.N. p53 tumor suppressor gene: a critical molecular target for UV induction and prevention of skin cancer. Photochem. Photobiol. 84, 55-62 (2008). [PubMed]

291. Green,D.R. At the gates of death. Cancer Cell 9, 328-330 (2006). [PubMed]

292. Park,S.J., Borin,B.N., Martinez-Yamout,M.A., & Dyson,H.J. The client protein p53 adopts a molten globule-like state in the presence of Hsp90. Nat. Struct. Mol. Biol. 18, 537-541 (2011). [PubMed]

293. Hagn,F. et al. Structural analysis of the interaction between Hsp90 and the tumor suppressor protein p53. Nat. Struct. Mol. Biol. 18, 1086-1093 (2011). [PubMed]

294. Walerych,D. et al. ATP binding to Hsp90 is sufficient for effective chaperoning of p53 protein. J. Biol. Chem. 285, 32020-32028 (2010). [PubMed]

295. Padmini,E. & Usha,R.M. Heat-shock protein 90 alpha (HSP90alpha) modulates signaling pathways towards tolerance of oxidative stress and enhanced survival of hepatocytes of Mugil cephalus. Cell Stress. Chaperones. 16, 411-425 (2011). [PubMed]

296. Millson,S.H. et al. Investigating the protein-protein interactions of the yeast Hsp90 chaperone system by two-hybrid analysis: potential uses and limitations of this approach. Cell Stress. Chaperones. 9, 359-368 (2004). [PubMed]

297. Eckert,K. et al. The Pih1-Tah1 cochaperone complex inhibits Hsp90 molecular chaperone ATPase activity. J. Biol. Chem. 285, 31304-31312 (2010). [PubMed]

298. Boulon,S. et al. The Hsp90 chaperone controls the biogenesis of L7Ae RNPs through conserved machinery. J. Cell Biol. 180, 579-595 (2008). [PubMed]

299. Jha,S. & Dutta,A. RVB1/RVB2: running rings around molecular biology. Mol. Cell 34, 521-533 (2009). [PubMed]

300. Sasaki,T. et al. HLA-B-associated transcript 3 (Bat3)/Scythe is essential for p300-mediated acetylation of p53. Genes Dev. 21, 848-861 (2007). [PubMed]

301. Pandey,P. et al. Negative regulation of cytochrome c-mediated oligomerization of Apaf-1 and activation of procaspase-9 by heat shock protein 90. EMBO J. 19, 4310-4322 (2000). [PubMed]

302. Zhang,R. et al. Hsp90-Akt phosphorylates ASK1 and inhibits ASK1-mediated apoptosis. Oncogene 24, 3954-3963 (2005). [PubMed]

303. Fulda,S., Galluzzi,L., & Kroemer,G. Targeting mitochondria for cancer therapy. Nat. Rev. Drug Discov. 9, 447-464 (2010). [PubMed]

304. Shang,Y. et al. Hsp70 and Hsp90 oppositely regulate TGF-beta signaling through CHIP/Stub1. Biochem. Biophys. Res. Commun. 446, 387-392 (2014). [PubMed]

305. Wrighton,K.H., Lin,X., & Feng,X.H. Critical regulation of TGFbeta signaling by Hsp90. Proc. Natl. Acad. Sci. U. S. A 105, 9244-9249 (2008). [PubMed]

306. Karkoulis,P.K., Stravopodis,D.J., Konstantakou,E.G., & Voutsinas,G.E. Targeted inhibition of heat shock protein 90 disrupts multiple oncogenic signaling pathways, thus inducing cell cycle arrest and programmed cell death in human urinary bladder cancer cell lines. Cancer Cell Int. 13, 11 (2013). [PubMed]

307. Mehta,P.P. et al. A novel class of specific Hsp90 small molecule inhibitors demonstrate in vitro and in vivo anti-tumor activity in human melanoma cells. Cancer Lett. 300, 30-39 (2011). [PubMed]

308. Okayama,S. et al. p53 Protein Regulates Hsp90 ATPase Activity and Thereby Wnt Signaling by Modulating Aha1 Expression. J. Biol. Chem. 289, 6513-6525 (2014). [PubMed]

309. Schwock,J. et al. Targeting focal adhesion kinase with dominant-negative FRNK or Hsp90 inhibitor 17-DMAG suppresses tumor growth and metastasis of SiHa cervical xenografts. Cancer Res. 69, 4750-4759 (2009). [PubMed]

310. Correia,A.L., Mori,H., Chen,E.I., Schmitt,F.C., & Bissell,M.J. The hemopexin domain of MMP3 is responsible for mammary epithelial invasion and morphogenesis through extracellular interaction with HSP90beta. Genes Dev. 27, 805-817 (2013). [PubMed]

311. Sims,J.D., McCready,J., & Jay,D.G. Extracellular heat shock protein (Hsp)70 and Hsp90alpha assist in matrix metalloproteinase-2 activation and breast cancer cell migration and invasion. PLoS. ONE. 6, e18848 (2011). [PubMed]

312. Song,X. et al. The regulatory mechanism of extracellular Hsp90{alpha} on matrix metalloproteinase-2 processing and tumor angiogenesis. J. Biol. Chem. 285, 40039-40049 (2010). [PubMed]

313. Nelson,A.R., Fingleton,B., Rothenberg,M.L., & Matrisian,L.M. Matrix metalloproteinases: biologic activity and clinical implications. J. Clin. Oncol. 18, 1135-1149 (2000). [PubMed]

314. Lagarrigue,F. et al. Matrix metalloproteinase-9 is upregulated in nucleophosmin-anaplastic lymphoma kinase-positive anaplastic lymphomas and activated at the cell surface by the chaperone heat shock protein 90 to promote cell invasion. Cancer Res. 70, 6978-6987 (2010). [PubMed]

315. Triantafilou,M. & Triantafilou,K. Heat-shock protein 70 and heat-shock protein 90 associate with Toll-like receptor 4 in response to bacterial lipopolysaccharide. Biochem. Soc. Trans. 32, 636-639 (2004). [PubMed]

316. Wheeler,D.S. et al. Extracellular Hsp72, an endogenous DAMP, is released by virally infected airway epithelial cells and activates neutrophils via Toll-like receptor (TLR)-4. Respir. Res. 10, 31 (2009). [PubMed]

317. Wallin,R.P. et al. Heat-shock proteins as activators of the innate immune system. Trends Immunol. 23, 130-135 (2002). [PubMed]

318. Bianchi,M.E. DAMPs, PAMPs and alarmins: all we need to know about danger. J. Leukoc. Biol. 81, 1-5 (2007). [PubMed]

319. Bohonowych,J.E. et al. Extracellular Hsp90 mediates an NF-kappaB dependent inflammatory stromal program: implications for the prostate tumor microenvironment. Prostate 74, 395-407 (2014). [PubMed]

320. Chung,S.W. et al. Extracellular heat shock protein 90 induces interleukin-8 in vascular smooth muscle cells. Biochem. Biophys. Res. Commun. 378, 444-449 (2009). [PubMed]

321. Ancrile,B., Lim,K.H., & Counter,C.M. Oncogenic Ras-induced secretion of IL6 is required for tumorigenesis. Genes Dev. 21, 1714-1719 (2007). [PubMed]

322. Sparmann,A. & Bar-Sagi,D. Ras-induced interleukin-8 expression plays a critical role in tumor growth and angiogenesis. Cancer Cell 6, 447-458 (2004). [PubMed]

323. Binder,R.J., Vatner,R., & Srivastava,P. The heat-shock protein receptors: some answers and more questions. Tissue Antigens 64, 442-451 (2004). [PubMed]

324. Murshid,A., Gong,J., & Calderwood,S.K. Heat shock protein 90 mediates efficient antigen cross presentation through the scavenger receptor expressed by endothelial cells-I. J. Immunol. 185, 2903-2917 (2010). [PubMed]

325. Panjwani,N.N., Popova,L., & Srivastava,P.K. Heat shock proteins gp96 and hsp70 activate the release of nitric oxide by APCs. J. Immunol. 15, 2997 (2002). [PubMed]

326. Berwin,B. et al. Scavenger receptor-A mediates gp96/GRP94 and calreticulin internalization by antigen-presenting cells. EMBO J. 22, 6127-6136 (2003). [PubMed]

327. Basu,S., Binder,R.J., Ramalingam,T., & Srivastava,P.K. CD91 is a common receptor for heat shock proteins gp96, hsp90, hsp70, and calreticulin. Immunity. 14, 303-313 (2001). [PubMed]

328. Binder,R.J., Han,D.K., & Srivastava,P.K. CD91: a receptor for heat shock protein gp96. Nat. Immunol. 1, 151-155 (2000). [PubMed]

329. Vabulas,R.M. et al. The endoplasmic reticulum-resident heat shock protein Gp96 activates dendritic cells via the Toll-like receptor 2/4 pathway. J. Biol. Chem. 277, 20847-20853 (2002). [PubMed]

330. Jozefowski,S., Biedrol,R., Srottek,M., Chadzilska,M., & Marcinkiewicz,J. The class A scavenger receptor SR-A/CD204 and the class B scavenger receptor CD36 regulate immune functions of macrophages differently. Innate. Immun.(2014). [PubMed]

331. Pick,E. et al. High HSP90 expression is associated with decreased survival in breast cancer. Cancer Res. 67, 2932-2937 (2007). [PubMed]

332. Ciocca,D.R., Gago,F.E., Fanelli,M.A., & Calderwood,S.K. Co-expression of steroid receptors (estrogen receptor alpha and/or progesterone receptors) and Her-2/neu: Clinical implications. J. Steroid Biochem. Mol. Biol. 102, 32-40 (2006). [PubMed]

333. 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. 61, 1799-1804 (2001). [PubMed]

334. Neckers,L. Hsp90 inhibitors as novel cancer chemotherapeutic agents. Trends Mol. Med. 8, S55-S61 (2002). [PubMed]

335. McCarthy,M.M. et al. HSP90 as a marker of progression in melanoma. Ann. Oncol. 19, 590-594 (2008). [PubMed]

336. Shimamura,T. & Shapiro,G.I. Heat shock protein 90 inhibition in lung cancer. J. Thorac. Oncol. 3, S152-S159 (2008). [PubMed]

337. Lim,S.O. et al. Expression of heat shock proteins (HSP27, HSP60, HSP70, HSP90, GRP78, GRP94) in hepatitis B virus-related hepatocellular carcinomas and dysplastic nodules. World J. Gastroenterol. 11, 2072-2079 (2005). [PubMed]

338. Lebret,T. et al. Heat shock proteins HSP27, HSP60, HSP70, and HSP90: expression in bladder carcinoma. Cancer 98, 970-977 (2003). [PubMed]

339. Elpek,G.O., Karaveli,S., Simsek,T., Keles,N., & Aksoy,N.H. Expression of heat-shock proteins hsp27, hsp70 and hsp90 in malignant epithelial tumour of the ovaries. APMIS 111, 523-530 (2003). [PubMed]

340. Chiu,C.C. et al. Molecular chaperones as a common set of proteins that regulate the invasion phenotype of head and neck cancer. Clin. Cancer Res. 17, 4629-4641 (2011). [PubMed]

341. Cardillo,M.R. & Ippoliti,F. IL-6, IL-10 and HSP-90 expression in tissue microarrays from human prostate cancer assessed by computer-assisted image analysis. Anticancer Res. 26, 3409-3416 (2006). [PubMed]

342. Cornford,P.A. et al. Heat shock protein expression independently predicts clinical outcome in prostate cancer. Cancer Res. 60, 7099-7105 (2000). [PubMed]

343. Miyake,H., Muramaki,M., Kurahashi,T., Takenaka,A., & Fujisawa,M. Expression of potential molecular markers in prostate cancer: correlation with clinicopathological outcomes in patients undergoing radical prostatectomy. Urol. Oncol. 28, 145-151 (2010). [PubMed]

344. Wang,X. et al. The regulatory mechanism of Hsp90alpha secretion and its function in tumor malignancy. Proc. Natl. Acad. Sci. U. S. A 106, 21288-21293 (2009). [PubMed]

345. Kakeda,M., Arock,M., Schlapbach,C., & Yawalkar,N. Increased expression of heat shock protein 90 in keratinocytes and mast cells in patients with psoriasis. J. Am. Acad. Dermatol. 70, 683-690 (2014). [PubMed]

346. Kasperkiewicz,M. et al. Evidence for a role of autoantibodies to heat shock protein 60, 70, and 90 in patients with dermatitis herpetiformis. Cell Stress Chaperones in press (2014). [PubMed]

347. Pires,E.S. & Khole,V.V. A block in the road to fertility: autoantibodies to heat-shock protein 90-beta in human ovarian autoimmunity. Fertil. Steril. 92, 1395-1409 (2009). [PubMed]

348. Yonekura,K. et al. Prevalence of anti-heat shock protein antibodies in cerebrospinal fluids of patients with Guillain-Barre syndrome. J. Neuroimmunol. 156, 204-209 (2004). [PubMed]

349. Carman,A., Kishinevsky,S., Koren,J., III, Luo,W., & Chiosis,G. Regulatory chaperone complexes in neurodegenerative diseases: a perspective on therapeutic intervention. Curr. Alzheimer Res. 11, 59-68 (2014). [PubMed]

350. Geller,R., Taguwa,S., & Frydman,J. Broad action of Hsp90 as a host chaperone required for viral replication. Biochim. Biophys. Acta 1823, 698-706 (2012). [PubMed]

351. Geller,R., Vignuzzi,M., Andino,R., & Frydman,J. Evolutionary constraints on chaperone-mediated folding provide an antiviral approach refractory to development of drug resistance. Genes Dev. 21, 195-205 (2007). [PubMed]

352. Geller,R., Andino,R., & Frydman,J. Hsp90 inhibitors exhibit resistance-free antiviral activity against respiratory syncytial virus. PLoS. ONE. 8, e56762 (2013). [PubMed]

353. Uryu,K. et al. Convergence of heat shock protein 90 with ubiquitin in filamentous alpha-synuclein inclusions of alpha-synucleinopathies. Am. J. Pathol. 168, 947-961 (2006). [PubMed]

354. Karagöz,G.E. et al. Hsp90-Tau complex reveals molecular basis for specificity in chaperone action. Cell 156, 963-974 (2014). [PubMed]

355. Thompson,A.D. et al. Analysis of the tau-associated proteome reveals that exchange of Hsp70 for Hsp90 is involved in tau degradation. ACS Chem. Biol. 7, 1677-1686 (2012). [PubMed]

356. Dou,F. et al. Chaperones increase association of tau protein with microtubules. Proc. Natl. Acad. Sci. U. S. A 100, 721-726 (2003). [PubMed]

357. George,J.M. The synucleins. Genome Biol. 3, reviews3002.1-reviews3002.6 (2002). [PubMed]

358. Wright,J.A., Wang,X., & Brown,D.R. Unique copper-induced oligomers mediate alpha-synuclein toxicity. FASEB J. 23, 2384-2393 (2009). [PubMed]

359. Falsone,S.F., Kungl,A.J., Rek,A., Cappai,R., & Zangger,K. The molecular chaperone Hsp90 modulates intermediate steps of amyloid assembly of the Parkinson-related protein alpha-synuclein. J. Biol. Chem. 284, 31190-31199 (2009). [PubMed]

360. Luk,K.C., Mills,I.P., Trojanowski,J.Q., & Lee,V.M. Interactions between Hsp70 and the hydrophobic core of alpha-synuclein inhibit fibril assembly. Biochemistry 47, 12614-12625 (2008). [PubMed]

361. Dickey,C.A. et al. The high-affinity HSP90-CHIP complex recognizes and selectively degrades phosphorylated tau client proteins. J. Clin. Invest 117, 648-658 (2007). [PubMed]

362. Luo,W. et al. Roles of heat-shock protein 90 in maintaining and facilitating the neurodegenerative phenotype in tauopathies. Proc. Natl. Acad. Sci. U. S. A 104, 9511-9516 (2007). [PubMed]

363. Mackenzie,I.R., Rademakers,R., & Neumann,M. TDP-43 and FUS in amyotrophic lateral sclerosis and frontotemporal dementia. Lancet Neurol. 9, 995-1007 (2010). [PubMed]

364. Sehati,S. et al. Metabolic alterations in yeast lacking copper-zinc superoxide dismutase. Free Radic. Biol. Med. 50, 1591-1598 (2011). [PubMed]

365. Choi,J.S., Cho,S., Park,S.G., Park,B.C., & Lee,D.H. Co-chaperone CHIP associates with mutant Cu/Zn-superoxide dismutase proteins linked to familial amyotrophic lateral sclerosis and promotes their degradation by proteasomes. Biochem. Biophys. Res. Commun. 321, 574-583 (2004). [PubMed]

366. Batulan,Z. et al. Induction of multiple heat shock proteins and neuroprotection in a primary culture model of familial amyotrophic lateral sclerosis. Neurobiol. Dis. 24, 213-225 (2006). [PubMed]

367. Ling,S.C. et al. ALS-associated mutations in TDP-43 increase its stability and promote TDP-43 complexes with FUS/TLS. Proc. Natl. Acad. Sci. U. S. A 107, 13318-13323 (2010). [PubMed]

368. Gregory,J.M., Barros,T.P., Meehan,S., Dobson,C.M., & Luheshi,L.M. The aggregation and neurotoxicity of TDP-43 and its ALS-associated 25 kDa fragment are differentially affected by molecular chaperones in Drosophila. PLoS. ONE. 7, e31899 (2012). [PubMed]

369. Basso,M. et al. Characterization of detergent-insoluble proteins in ALS indicates a causal link between nitrative stress and aggregation in pathogenesis. PLoS. ONE. 4, e8130 (2009). [PubMed]

370. Zinkie,S., Gentil,B.J., Minotti,S., & Durham,H.D. Expression of the protein chaperone, clusterin, in spinal cord cells constitutively and following cellular stress, and upregulation by treatment with Hsp90 inhibitor. Cell Stress Chaperones 18, 745-758 (2013). [PubMed]

371. Rizzi,F., Coletta,M., & Bettuzzi,S. Chapter 2: Clusterin (CLU): From one gene and two transcripts to many proteins. Adv. Cancer Res. 104, 9-23 (2009). [PubMed]

372. Rizzi,F. & Bettuzzi,S. Clusterin (CLU) and prostate cancer. Adv. Cancer Res. 105, 1-19 (2009). [PubMed]

373. Panico,F., Rizzi,F., Fabbri,L.M., Bettuzzi,S., & Luppi,F. Clusterin (CLU) and lung cancer. Adv. Cancer Res. 105, 63-76 (2009). [PubMed]

374. Duval,A. et al. Expression and prognostic significance of heat-shock proteins in myelodysplastic syndromes. Haematologica 91, 713-714 (2006). [PubMed]

375. Tefferi,A. & Vardiman,J.W. Myelodysplastic syndromes. N. Engl. J. Med. 361, 1872-1885 (2009). [PubMed]

376. Heaney,M.L. & Golde,D.W. Myelodysplasia. N. Engl. J. Med. 340, 1649-1660 (1999). [PubMed]

377. Heyman,M.R. Recent advances in biology and treatment of myelodysplasia. Curr. Opin. Oncol. 3, 44-53 (1991). [PubMed]

378. Flandrin-Gresta,P. et al. Heat Shock Protein 90 is overexpressed in high-risk myelodysplastic syndromes and associated with higher expression and activation of Focal Adhesion Kinase. Oncotarget. 3, 1158-1168 (2012). [PubMed]

379. Yufu,Y., Nishimura,J., & Nawata,H. High constitutive expression of heat shock protein 90 alpha in human acute leukemia cells. Leuk. Res. 16, 597-605 (1992). [PubMed]

380. Aanei,C.M. et al. Focal adhesion protein abnormalities in myelodysplastic mesenchymal stromal cells. Exp. Cell Res. 317, 2616-2629 (2011). [PubMed]

381. Flandrin,P. et al. Significance of heat-shock protein (HSP) 90 expression in acute myeloid leukemia cells. Cell Stress Chaperones 13, 357-364 (2008). [PubMed]

382. Gu,Y., Lewis,D.F., Zhang,Y., Groome,L.J., & Wang,Y. Increased superoxide generation and decreased stress protein Hsp90 expression in human umbilical cord vein endothelial cells (HUVECs) from pregnancies complicated by preeclampsia. Hypertens. Pregnancy. 25, 169-182 (2006). [PubMed]

383. Padmini,E., Venkatraman,U., & Srinivasan,L. Mechanism of JNK signal regulation by placental HSP70 and HSP90 in endothelial cell during preeclampsia. Toxicol. Mech. Methods 22, 367-374 (2012). [PubMed]

384. Ekambaram,P., Jayachandran,T., & Dhakshinamoorthy,L. Differential expression of HSP90alpha and heme oxygenase in cord blood RBC during preeclampsia. Toxicol. Mech. Methods 23, 113-119 (2013). [PubMed]

385. Lopatin,D.E., Shelburne,C.E., Van,P.N., Kowalski,C.J., & Bagramian,R.A. Humoral immunity to stress proteins and periodontal disease. J. Periodontol. 70, 1185-1193 (1999). [PubMed]

386. Shelburne,C.E., Coopamah,M.D., Sweier,D.G., An,F.Y., & Lopatin,D.E. HtpG, the Porphyromonas gingivalis HSP-90 homologue, induces the chemokine CXCL8 in human monocytic and microvascular vein endothelial cells. Cell Microbiol. 9, 1611-1619 (2007). [PubMed]

387. Shelburne,C.E. et al. Serum antibodies to Porphyromonas gingivalis chaperone HtpG predict health in periodontitis susceptible patients. PLoS. ONE. 3, e1984 (2008). [PubMed]

388. Rutherford,S.L. & Lindquist,S. Hsp90 as a capacitor for morphological evolution. Nature 396, 336-342 (1998). [PubMed]

389. Kim,Y.S. et al. Update on Hsp90 inhibitors in clinical trial. Curr. Top. Med. Chem. 9, 1479-1492 (2009). [PubMed]

390. Fadden,P. et al. Application of chemoproteomics to drug discovery: identification of a clinical candidate targeting hsp90. Chem. Biol. 17, 686-694 (2010). [PubMed]

391. Socinski,M.A. et al. A multicenter phase II study of ganetespib monotherapy in patients with genotypically defined advanced non-small cell lung cancer. Clin. Cancer Res. 19, 3068-3077 (2013). [PubMed]

392. Modi,S. et al. HSP90 inhibition is effective in breast cancer: a phase II trial of tanespimycin (17-AAG) plus trastuzumab in patients with HER2-positive metastatic breast cancer progressing on trastuzumab. Clin. Cancer Res. 17, 5132-5139 (2011). [PubMed]

393. Garraway,L.A. & Jänne,P.A. Circumventing cancer drug resistance in the era of personalized medicine. Cancer Discov. 2, 214-226 (2012). [PubMed]

394. Zaarur,N., Gabai,V.L., Porco,J.A., Jr., Calderwood,S., & Sherman,M.Y. Targeting heat shock response to sensitize cancer cells to proteasome and Hsp90 inhibitors. Cancer Res. 66, 1783-1791 (2006). [PubMed]

395. Dai,C. et al. Loss of tumor suppressor NF1 activates HSF1 to promote carcinogenesis. J. Clin. Invest 122, 3742-3754 (2012). [PubMed]

396. Au,Q., Zhang,Y., Barber,J.R., Ng,S.C., & Zhang,B. Identification of inhibitors of HSF1 functional activity by high-content target-based screening. J. Biomol. Screen. 14, 1165-1175 (2009). [PubMed]

397. Gehrmann,M., Radons,J., Molls,M., & Multhoff,G. The therapeutic implications of clinically applied modifiers of heat shock protein 70 (Hsp70) expression by tumor cells. Cell Stress. Chaperones. 13, 1-10 (2008). [PubMed]

398. Syrigos,K.N. et al. Clinical significance of heat shock protein-70 expression in bladder cancer. Urology 61, 677-680 (2003). [PubMed]

399. Kaur,J., Srivastava,A., & Ralhan,R. Expression of 70-kDa heat shock protein in oral lesions: marker of biological stress or pathogenicity. Oral Oncol. 34, 496-501 (1998). [PubMed]

400. Lazaris,A.C., Theodoropoulos,G.E., Aroni,K., Saetta,A., & Davaris,P.S. Immunohistochemical expression of C-myc oncogene, heat shock protein 70 and HLA-DR molecules in malignant cutaneous melanoma. Virchows Arch. 426, 461-467 (1995). [PubMed]

401. Ralhan,R. & Kaur,J. Differential expression of Mr 70,000 heat shock protein in normal, premalignant, and malignant human uterine cervix. Clin. Cancer Res. 1, 1217-1222 (1995). [PubMed]

402. Murshid,A., Gong,J., Stevenson,M.A., & Calderwood,S.K. Heat shock proteins and cancer vaccines: developments in the past decade and chaperoning in the decade to come. Expert. Rev. Vaccines. 10, 1553-1568 (2011). [PubMed]

403. Calderwood,S.K., Stevenson,M.A., & Murshid,A. Heat shock proteins, autoimmunity, and cancer treatment. Autoimmune. Dis. 2012, 486069 (2012). [PubMed]

404. Murshid,A., Gong,J., & Calderwood,S.K. Purification, preparation, and use of chaperone-peptide complexes for tumor immunotherapy. Methods Mol. Biol. 960, 209-217 (2013). [PubMed]

405. Udono,H. & Srivastava,P.K. Heat shock protein 70-associated peptides elicit specific cancer immunity. J. Exp. Med. 178, 1391-1396 (1993). [PubMed]

406. Testori,A. et al. Phase III comparison of vitespen, an autologous tumor-derived heat shock protein gp96 peptide complex vaccine, with physician’s choice of treatment for stage IV melanoma: the C-100-21 Study Group. J. Clin. Oncol. 26, 955-962 (2008). [PubMed]

407. Wood,C. et al. An adjuvant autologous therapeutic vaccine (HSPPC-96; vitespen) versus observation alone for patients at high risk of recurrence after nephrectomy for renal cell carcinoma: a multicentre, open-label, randomised phase III trial. Lancet 372, 145-154 (2008). [PubMed]

408. Rivoltini,L. et al. Human tumor-derived heat shock protein 96 mediates in vitro activation and in vivo expansion of melanoma- and colon carcinoma-specific T cells. J. Immunol. 171, 3467-3474 (2003). [PubMed]

409. Li,Z. et al. Combination of imatinib mesylate with autologous leukocyte-derived heat shock protein and chronic myelogenous leukemia. Clin. Cancer Res. 11, 4460-4468 (2005). [PubMed]

410. Janetzki,S. et al. Immunization of cancer patients with autologous cancer-derived heat shock protein gp96 preparations: a pilot study. Int. J. Cancer 88, 232-238 (2000). [PubMed]

411. Reitsma,D.J. & Combest,A.J. Challenges in the development of an autologous heat shock protein based anti-tumor vaccine. Hum. Vaccin. Immunother. 8, 1152-1155 (2012). [PubMed]

412. Chandawarkar,R.Y., Wagh,M.S., & Srivastava,P.K. The dual nature of specific immunological activity of tumor-derived gp96 preparations. J. Exp. Med. 189, 1437-1442 (1999). [PubMed]

413. Liu,S. et al. CD8+ lymphocyte infiltration is an independent favorable prognostic indicator in basal-like breast cancer. Breast Cancer Res. 14, R48 (2012). [PubMed]

414. Shinagawa,N. et al. Immunotherapy with dendritic cells pulsed with tumor-derived gp96 against murine lung cancer is effective through immune response of CD8+ cytotoxic T lymphocytes and natural killer cells. Cancer Immunol. Immunother. 57, 165-174 (2008). [PubMed]

415. Li,G. et al. Human ovarian tumour-derived chaperone-rich cell lysate (CRCL) elicits T cell responses in vitro. Clin. Exp. Immunol. 148, 136-145 (2007). [PubMed]

416. Pilla,L. et al. A phase II trial of vaccination with autologous, tumor-derived heat-shock protein peptide complexes Gp96, in combination with GM-CSF and interferon-alpha in metastatic melanoma patients. Cancer Immunol. Immunother. 55, 958-968 (2006). [PubMed]

417. Kojima,T. et al. Granulocyte-macrophage colony-stimulating factor gene-transduced tumor cells combined with tumor-derived gp96 inhibit tumor growth in mice. Hum. Gene Ther. 14, 715-728 (2003). [PubMed]

418. Zerbini,A. et al. Increased immunostimulatory activity conferred to antigen-presenting cells by exposure to antigen extract from hepatocellular carcinoma after radiofrequency thermal ablation. J. Immunother. 31, 271-282 (2008). [PubMed]

419. Demaria,S., Bhardwaj,N., McBride,W.H., & Formenti,S.C. Combining radiotherapy and immunotherapy: a revived partnership. Int. J. Radiat. Oncol. Biol. Phys. 63, 655-666 (2005). [PubMed]

420. Shi,Y., Evans,J.E., & Rock,K.L. Molecular identification of a danger signal that alerts the immune system to dying cells. Nature 425, 516-521 (2003). [PubMed]

421. Chen,Q. et al. Fever-range thermal stress promotes lymphocyte trafficking across high endothelial venules via an interleukin 6 trans-signaling mechanism. Nat. Immunol. 7, 1299-1308 (2006). [PubMed]

422. Kottke,T. et al. Induction of hsp70-mediated Th17 autoimmunity can be exploited as immunotherapy for metastatic prostate cancer. Cancer Res. 67, 11970-11979 (2007). [PubMed]

423. Multhoff,G., Molls,M., & Radons,J. Chronic inflammation in cancer development. Front. Immunol. 2:98., doi: 10.3389/fimmu.2011.00098 (2012). [PubMed]

424. Warger,T. et al. Interaction of TLR2 and TLR4 ligands with the N-terminal domain of Gp96 amplifies innate and adaptive immune responses. J. Biol. Chem. 281, 22545-22553 (2006). [PubMed]

425. Okuya,K. et al. Spatiotemporal regulation of heat shock protein 90-chaperoned self-DNA and CpG-oligodeoxynucleotide for type I IFN induction via targeting to static early endosome. J. Immunol. 184, 7092-7099 (2010). [PubMed]

426. Rajaiah,R. & Moudgil,K.D. Heat-shock proteins can promote as well as regulate autoimmunity. Autoimmun. Rev. 8, 388-393 (2009). [PubMed]

427. Liu,Z. et al. Treg suppress CTL responses upon immunization with HSP gp96. Eur. J. Immunol. 39, 3110-3120 (2009). [PubMed]

428. Chen,L. Co-inhibitory molecules of the B7-CD28 family in the control of T-cell immunity. Nat. Rev. Immunol. 4, 336-347 (2004). [PubMed]

429. Tarhini,A.A. & Kirkwood,J.M. CTLA-4-blocking immunotherapy with ipilimumab for advanced melanoma. Oncology (Williston. Park) 24, 1302, 1304 (2010). [PubMed]

430. Dulos,J. et al. PD-1 blockade augments Th1 and Th17 and suppresses Th2 responses in peripheral blood from patients with prostate and advanced melanoma cancer. J. Immunother. 35, 169-178 (2012). [PubMed]

431. Thumar,J.R. & Kluger,H.M. Ipilimumab: a promising immunotherapy for melanoma. Oncology (Williston. Park) 24, 1280-1288 (2010). [PubMed]

432. Jaber,S.H. et al. Skin reactions in a subset of patients with stage IV melanoma treated with anti-cytotoxic T-lymphocyte antigen 4 monoclonal antibody as a single agent. Arch. Dermatol. 142, 166-172 (2006). [PubMed]

433. Callahan,M.K., Wolchok,J.D., & Allison,J.P. Anti-CTLA-4 antibody therapy: immune monitoring during clinical development of a novel immunotherapy. Semin. Oncol. 37, 473-484 (2010). [PubMed]

434. Gong,J. et al. A heat shock protein 70-based vaccine with enhanced immunogenicity for clinical use. J. Immunol. 184, 488-496 (2010). [PubMed]

435. Enomoto,Y. et al. Enhanced immunogenicity of heat shock protein 70 peptide complexes from dendritic cell-tumor fusion cells. J. Immunol. 177, 5946-5955 (2006). [PubMed]

436. Weng,D. et al. Induction of cytotoxic T lymphocytes against ovarian cancer-initiating cells. Int. J. Cancer 129, 1990-2001 (2011). [PubMed]

437. Jego,G., Hazoume,A., Seigneuric,R., & Garrido,C. Targeting heat shock proteins in cancer. Cancer Lett. 332, 275-285 (2013). [PubMed]

438. Supko,J.G., Hickman,R.L., Grever,M.R., & Malspeis,L. Preclinical pharmacologic evaluation of geldanamycin as an antitumor agent. Cancer Chemother. Pharmacol. 36, 305-315 (1995). [PubMed]

439. Eiseman,J.L. et al. Pharmacokinetics and pharmacodynamics of 17-demethoxy 17-[[(2-dimethylamino)ethyl]amino]geldanamycin (17DMAG, NSC 707545) in C.B-17 SCID mice bearing MDA-MB-231 human breast cancer xenografts. Cancer Chemother. Pharmacol. 55, 21-32 (2005). [PubMed]

440. Georgakis,G.V., Li,Y., & Younes,A. The heat shock protein 90 inhibitor 17-AAG induces cell cycle arrest and apoptosis in mantle cell lymphoma cell lines by depleting cyclin D1, Akt, Bid and activating caspase 9. Br. J. Haematol. 135, 68-71 (2006). [PubMed]

441. Pacey,S., Banerji,U., Judson,I., & Workman,P. Hsp90 inhibitors in the clinic. Handb. Exp. Pharmacol.331-358 (2006). [PubMed]

442. Kummar,S. et al. Phase I trial of 17-dimethylaminoethylamino-17-demethoxygeldanamycin (17-DMAG), a heat shock protein inhibitor, administered twice weekly in patients with advanced malignancies. Eur. J. Cancer 46, 340-347 (2010). [PubMed]

443. Lancet,J.E. et al. Phase I study of the heat shock protein 90 inhibitor alvespimycin (KOS-1022, 17-DMAG) administered intravenously twice weekly to patients with acute myeloid leukemia. Leukemia 24, 699-705 (2010). [PubMed]

444. Palacios,C., Lopez-Perez,A.I., & Lopez-Rivas,A. Down-regulation of RIP expression by 17-dimethylaminoethylamino-17-demethoxygeldanamycin promotes TRAIL-induced apoptosis in breast tumor cells. Cancer Lett. 287, 207-215 (2010). [PubMed]

445. Ayrault,O. et al. Inhibition of Hsp90 via 17-DMAG induces apoptosis in a p53-dependent manner to prevent medulloblastoma. Proc. Natl. Acad. Sci. U. S. A 106, 17037-17042 (2009). [PubMed]

446. Chatterjee,M. et al. STAT3 and MAPK signaling maintain overexpression of heat shock proteins 90alpha and beta in multiple myeloma cells, which critically contribute to tumor-cell survival. Blood 109, 720-728 (2007). [PubMed]

447. Pillai,R.N. & Ramalingam,S.S. Heat shock protein 90 inhibitors in non-small-cell lung cancer. Curr. Opin. Oncol. 26, 159-164 (2014). [PubMed]

448. Wagner,A.J. et al. A phase I study of the HSP90 inhibitor retaspimycin hydrochloride (IPI-504) in patients with gastrointestinal stromal tumors or soft-tissue sarcomas. Clin. Cancer Res. 19, 6020-6029 (2013). [PubMed]

449. Modi,S. et al. A multicenter trial evaluating retaspimycin HCL (IPI-504) plus trastuzumab in patients with advanced or metastatic HER2-positive breast cancer. Breast Cancer Res. Treat. 139, 107-113 (2013). [PubMed]

450. Hanson,B.E. & Vesole,D.H. Retaspimycin hydrochloride (IPI-504): a novel heat shock protein inhibitor as an anticancer agent. Expert. Opin. Investig. Drugs 18, 1375-1383 (2009). [PubMed]

451. Siegel,D. et al. A phase 1 study of IPI-504 (retaspimycin hydrochloride) in patients with relapsed or relapsed and refractory multiple myeloma. Leuk. Lymphoma 52, 2308-2315 (2011). [PubMed]

452. Floris,G. et al. The Novel HSP90 inhibitor, IPI-493, is highly effective in human gastrostrointestinal stromal tumor xenografts carrying heterogeneous KIT mutations. Clin. Cancer Res. 17, 5604-5614 (2011). [PubMed]

453. Bedi,A. et al. BCR-ABL-mediated inhibition of apoptosis with delay of G2/M transition after DNA damage: a mechanism of resistance to multiple anticancer agents. Blood 86, 1148-1158 (1995). [PubMed]

454. Radujkovic,A. et al. Synergistic activity of imatinib and 17-AAG in imatinib-resistant CML cells overexpressing BCR-ABL–Inhibition of P-glycoprotein function by 17-AAG. Leukemia 19, 1198-1206 (2005). [PubMed]

455. Erlichman,C. Tanespimycin: the opportunities and challenges of targeting heat shock protein 90. Expert. Opin. Investig. Drugs 18, 861-868 (2009). [PubMed]

456. Wu,Y.C., Yen,W.Y., Lee,T.C., & Yih,L.H. Heat shock protein inhibitors, 17-DMAG and KNK437, enhance arsenic trioxide-induced mitotic apoptosis. Toxicol. Appl. Pharmacol. 236, 231-238 (2009). [PubMed]

457. Robles,A.I. et al. Schedule-dependent synergy between the heat shock protein 90 inhibitor 17-(dimethylaminoethylamino)-17-demethoxygeldanamycin and doxorubicin restores apoptosis to p53-mutant lymphoma cell lines. Clin. Cancer Res. 12, 6547-6556 (2006). [PubMed]

458. Marcu,M.G., Chadli,A., Bouhouche,I., Catelli,M., & Neckers,L.M. The heat shock protein 90 antagonist novobiocin interacts with a previously unrecognized ATP-binding domain in the carboxyl terminus of the chaperone. J. Biol. Chem. 275, 37181-37186 (2000). [PubMed]

459. Heddle,J.G., Barnard,F.M., Wentzell,L.M., & Maxwell,A. The interaction of drugs with DNA gyrase: a model for the molecular basis of quinolone action. Nucleosides Nucleotides Nucleic Acids 19, 1249-1264 (2000). [PubMed]

460. Heide,L. New aminocoumarin antibiotics as gyrase inhibitors. Int. J. Med. Microbiol. 304, 31-36 (2014). [PubMed]

461. Allan,R.K., Mok,D., Ward,B.K., & Ratajczak,T. Modulation of chaperone function and cochaperone interaction by novobiocin in the C-terminal domain of Hsp90: evidence that coumarin antibiotics disrupt Hsp90 dimerization. J. Biol. Chem. 281, 7161-7171 (2006). [PubMed]

462. Samuels,D.S. & Garon,C.F. Coumermycin A1 inhibits growth and induces relaxation of supercoiled plasmids in Borrelia burgdorferi, the Lyme disease agent. Antimicrob. Agents Chemother. 37, 46-50 (1993). [PubMed]

463. Wu,L.X. et al. Disruption of the Bcr-Abl/Hsp90 protein complex: a possible mechanism to inhibit Bcr-Abl-positive human leukemic blasts by novobiocin. Leukemia 22, 1402-1409 (2008). [PubMed]

464. Cohen,S.M. et al. Novel C-terminal Hsp90 inhibitor for head and neck squamous cell cancer (HNSCC) with in vivo efficacy and improved toxicity profiles compared with standard agents. Ann. Surg. Oncol. 19 Suppl 3, S483-S490 (2012). [PubMed]

465. Samadi,A.K. et al. A novel C-terminal HSP90 inhibitor KU135 induces apoptosis and cell cycle arrest in melanoma cells. Cancer Lett. 312, 158-167 (2011). [PubMed]

466. Matthews,S.B. et al. Characterization of a novel novobiocin analogue as a putative C-terminal inhibitor of heat shock protein 90 in prostate cancer cells. Prostate 70, 27-36 (2010). [PubMed]

467. Shelton,S.N. et al. KU135, a novel novobiocin-derived C-terminal inhibitor of the 90-kDa heat shock protein, exerts potent antiproliferative effects in human leukemic cells. Mol. Pharmacol. 76, 1314-1322 (2009). [PubMed]

468. Schulte,T.W. et al. Antibiotic radicicol binds to the N-terminal domain of Hsp90 and shares important biologic activities with geldanamycin. Cell Stress Chaperones 3, 100-108 (1998). [PubMed]

469. Soga,S., Shiotsu,Y., Akinaga,S., & Sharma,S.V. Development of radicicol analogues. Curr. Cancer Drug Targets. 3, 359-369 (2003). [PubMed]

470. Huntoon,C.J. et al. Heat shock protein 90 inhibition depletes LATS1 and LATS2, two regulators of the mammalian hippo tumor suppressor pathway. Cancer Res. 70, 8642-8650 (2010). [PubMed]

471. Higashi,C. et al. The effects of heat shock protein 90 inhibitors on apoptosis and viral replication in primary effusion lymphoma cells. Biol. Pharm. Bull. 35, 725-730 (2012). [PubMed]

472. Geng,X., Yang,Z.-Q., & Danishefsky,S.J. Synthetic development of radicicol and cycloproparadicicol: highly promising anticancer agents targeting Hsp90. Synlett 8, 1325-1333 (2004). [DOI: 10.1055/s-2004-829052]

473. Shiotsu,Y. 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 96, 2284-2291 (2000). [PubMed]

474. Garon,E.B. et al. The HSP90 inhibitor NVP-AUY922 potently inhibits non-small cell lung cancer growth. Mol. Cancer Ther. 12, 890-900 (2013). [PubMed]

475. Eccles,S.A. et al. NVP-AUY922: a novel heat shock protein 90 inhibitor active against xenograft tumor growth, angiogenesis, and metastasis. Cancer Res. 68, 2850-2860 (2008). [PubMed]

476. Huo,C. et al. The challenge of developing green tea polyphenols as therapeutic agents. Inflammopharmacology. 16, 248-252 (2008). [PubMed]

477. Hugel,H.M. & Jackson,N. Redox chemistry of green tea polyphenols: therapeutic benefits in neurodegenerative diseases. Mini. Rev. Med. Chem. 12, 380-387 (2012). [PubMed]

478. Steinmann,J., Buer,J., Pietschmann,T., & Steinmann,E. Anti-infective properties of epigallocatechin-3-gallate (EGCG), a component of green tea. Br. J. Pharmacol. 168, 1059-1073 (2013). [PubMed]

479. Yin,Z., Henry,E.C., & Gasiewicz,T.A. (-)-Epigallocatechin-3-gallate is a novel Hsp90 inhibitor. Biochemistry 48, 336-345 (2009). [PubMed]

480. Bhat,R. et al. Towards the discovery of drug-like epigallocatechin gallate analogs as Hsp90 inhibitors. Bioorg. Med. Chem. Lett. 24, 2263-2266 (2014). [PubMed]

481. Sharp,S.Y. et al. In vitro biological characterization of a novel, synthetic diaryl pyrazole resorcinol class of heat shock protein 90 inhibitors. Cancer Res. 67, 2206-2216 (2007). [PubMed]

482. Sharp,S.Y. et al. Inhibition of the heat shock protein 90 molecular chaperone in vitro and in vivo by novel, synthetic, potent resorcinylic pyrazole/isoxazole amide analogues. Mol. Cancer Ther. 6, 1198-1211 (2007). [PubMed]

483. Hong,D.S. et al. Targeting the molecular chaperone heat shock protein 90 (HSP90): lessons learned and future directions. Cancer Treat. Rev. 39, 375-387 (2013). [PubMed]

484. Sessa,C. et al. First-in-human phase I dose-escalation study of the HSP90 inhibitor AUY922 in patients with advanced solid tumors. Clin. Cancer Res. 19, 3671-3680 (2013). [PubMed]

485. Samuel,T.A. et al. AUY922, a novel HSP90 inhibitor: Final results of a first-in-human study in patients with advanced solid malignancies. J. Clin. Oncol. 28 (Suppl.), 2528 (2010). [Abstract]

486. Schroder,C.P. et al. Use of biomarkers and imaging to evaluate the treatment effect of AUY922, an HSP90 inhibitor, in patients with HER2+ or ER+ metastatic breast cancer. J. Clin. Oncol. 29 (Suppl.), e11024 (2011). [Abstract]

487. Kong,A. et al. Phase IB/II study of the HSP90 inhibitor AUY922, in combination with trastuzumab, in patients with HER2+ advanced breast cancer. J. Clin. Oncol. 30 (Suppl.), 530 (2012). [Abstract]

488. Johnson,M.L. et al. A phase II study of HSP90 inhibitor AUY922 and erlotinib (E) for patients (pts) with EGFR-mutant lung cancer and acquired resistance (AR) to EGFR tyrosine kinase inhibitors (EGFR TKIs). J. Clin. Oncol. 31 (Suppl.), 8036 (2013). [Abstract]

489. Shapiro,G. et al. Phase I pharmacokinetic and pharmacodynamic study of the heat shock protein 90 inhibitor AT13387 in patients with refractory solid tumors. J. Clin. Oncol. 28 (Suppl.), 3069 (2010). [Abstract]

490. Goldman,J.W. et al. A first in human, safety, pharmacokinetics, and clinical activity phase I study of once weekly administration of the Hsp90 inhibitor ganetespib (STA-9090) in patients with solid malignancies. BMC. Cancer 13, 152 (2013). [PubMed]

491. Goldman,J.W. et al. A phase I dose-escalation study of the Hsp90 inhibitor STA-9090 administered once weekly in patients with solid tumors. J. Clin. Oncol. 28 (Suppl.), 2529 (2010). [Abstract]

492. Cleary,J.M. et al. A phase I dose-escalation study of the Hsp90 inhibitor STA-9090 administered twice weekly in patients with solid tumors. J. Clin. Oncol. 28 (Suppl.), 3083 (2010). [Abstract]

493. Nagaraju,G.P. et al. Antiangiogenic effects of ganetespib in colorectal cancer mediated through inhibition of HIF-1alpha and STAT-3. Angiogenesis. 16, 903-917 (2013). [PubMed]

494. Kim,S.H. et al. Discovery of (2S)-1-[4-(2-{6-amino-8-[(6-bromo-1,3-benzodioxol-5-yl)sulfanyl]-9H-purin-9-yl}et hyl)piperidin-1-yl]-2-hydroxypropan-1-one (MPC-3100), a purine-based Hsp90 inhibitor. J. Med. Chem. 55, 7480-7501 (2012). [PubMed]

495. Yu,M.K. et al. MPC-3100, a fully synthetic, orally bioavailable Hsp90 inhibitor, in cancer patients. J. Clin. Oncol. 28 (Suppl.), e13112 (2010). [Abstract]

496. Caldas-Lopes,E. et al. Hsp90 inhibitor PU-H71, a multimodal inhibitor of malignancy, induces complete responses in triple-negative breast cancer models. Proc. Natl. Acad. Sci. U. S. A 106, 8368-8373 (2009). [PubMed]

497. Cerchietti,L.C. et al. A purine scaffold Hsp90 inhibitor destabilizes BCL-6 and has specific antitumor activity in BCL-6-dependent B cell lymphomas. Nat. Med. 15, 1369-1376 (2009). [PubMed]

498. Lundgren,K. et al. BIIB021, an orally available, fully synthetic small-molecule inhibitor of the heat shock protein Hsp90. Mol. Cancer Ther. 8, 921-929 (2009). [PubMed]

499. Boll,B. et al. Heat shock protein 90 inhibitor BIIB021 (CNF2024) depletes NF-kappaB and sensitizes Hodgkin’s lymphoma cells for natural killer cell-mediated cytotoxicity. Clin. Cancer Res. 15, 5108-5116 (2009). [PubMed]

500. Yin,X. et al. BIIB021, a novel Hsp90 inhibitor, sensitizes head and neck squamous cell carcinoma to radiotherapy. Int. J. Cancer 126, 1216-1225 (2010). [PubMed]

501. Zhang,H. et al. BIIB021, a synthetic Hsp90 inhibitor, has broad application against tumors with acquired multidrug resistance. Int. J. Cancer 126, 1226-1234 (2010). [PubMed]

502. Castro,J.E. et al. ZAP-70 is a novel conditional heat shock protein 90 (Hsp90) client: inhibition of Hsp90 leads to ZAP-70 degradation, apoptosis, and impaired signaling in chronic lymphocytic leukemia. Blood 106, 2506-2512 (2005). [PubMed]

503. Okawa,Y. et al. SNX-2112, a selective Hsp90 inhibitor, potently inhibits tumor cell growth, angiogenesis, and osteoclastogenesis in multiple myeloma and other hematologic tumors by abrogating signaling via Akt and ERK. Blood 113, 846-855 (2009). [PubMed]

504. Rajan,A. et al. A phase I study of PF-04929113 (SNX-5422), an orally bioavailable heat shock protein 90 inhibitor, in patients with refractory solid tumor malignancies and lymphomas. Clin. Cancer Res. 17, 6831-6839 (2011). [PubMed]

505. Reddy,N. et al. Phase I trial of the HSP90 inhibitor PF-04929113 (SNX5422) in adult patients with recurrent, refractory hematologic malignancies. Clin. Lymphoma Myeloma. Leuk. 13, 385-391 (2013). [PubMed]

506. Huang,K.H. et al. Discovery of novel 2-aminobenzamide inhibitors of heat shock protein 90 as potent, selective and orally active antitumor agents. J. Med. Chem. 52, 4288-4305 (2009). [PubMed]

507. Wang,S. et al. SNX-25a, a novel Hsp90 inhibitor, inhibited human cancer growth more potently than 17-AAG. Biochem. Biophys. Res. Commun. in press (2014). [PubMed]

508. Plescia,J. et al. Rational design of shepherdin, a novel anticancer agent. Cancer Cell 7, 457-468 (2005). [PubMed]

509. Yi,F. & Regan,L. A novel class of small molecule inhibitors of Hsp90. ACS Chem. Biol. 3, 645-654 (2008). [PubMed]

510. Soldano,K.L., Jivan,A., Nicchitta,C.V., & Gewirth,D.T. Structure of the N-terminal domain of GRP94. Basis for ligand specificity and regulation. J. Biol. Chem. 278, 48330-48338 (2003). [PubMed]

511. Cheung,K.M. et al. The identification, synthesis, protein crystal structure and in vitro biochemical evaluation of a new 3,4-diarylpyrazole class of Hsp90 inhibitors. Bioorg. Med. Chem. Lett. 15, 3338-3343 (2005). [PubMed]

512. Rowlands,M.G. et al. High-throughput screening assay for inhibitors of heat-shock protein 90 ATPase activity. Anal. Biochem. 327, 176-183 (2004). [PubMed]

513. Pachl,J. et al. A randomized, blinded, multicenter trial of lipid-associated amphotericin B alone versus in combination with an antibody-based inhibitor of heat shock protein 90 in patients with invasive candidiasis. Clin. Infect. Dis. 42, 1404-1413 (2006). [PubMed]

514. Prodromou,C., Roe,S.M., Piper,P.W., & Pearl,L.H. A molecular clamp in the crystal structure of the N-terminal domain of the yeast Hsp90 chaperone. Nat. Struct. Biol. 4, 477-482 (1997). [PubMed]

515. Clare,D.K. & Saibil,H.R. ATP-driven molecular chaperone machines. Biopolymers 99, 846-859 (2013). [PubMed]

516. Lele,Z. et al. Disruption of zebrafish somite development by pharmacologic inhibition of Hsp90. Dev. Biol. 210, 56-70 (1999). [PubMed]

517. Whitesell,L. et al. Inhibition of heat shock protein HSP90-pp6Ov-src heteroprotein complex formation by benzoquinone ansamycins: Essential role for stress proteins in oncogenic transformation.  Proc. Natl. Acad. Sci. USA 91, 8324-8328 (1994). [PubMed]