Hostname: page-component-8448b6f56d-c47g7 Total loading time: 0 Render date: 2024-04-16T14:27:22.765Z Has data issue: false hasContentIssue false

Conformational dynamics of the molecular chaperone Hsp90

Published online by Cambridge University Press:  18 March 2011

Kristin A. Krukenberg
Affiliation:
Department of Biochemistry and Biophysics, University of California, San Francisco, CA, USA
Timothy O. Street
Affiliation:
Department of Biochemistry and Biophysics, University of California, San Francisco, CA, USA
Laura A. Lavery
Affiliation:
Department of Biochemistry and Biophysics, University of California, San Francisco, CA, USA
David A. Agard*
Affiliation:
Department of Biochemistry and Biophysics, University of California, San Francisco, CA, USA The Howard Hughes Medical Institute, University of California, San Francisco, CA, USA
*
*Author for correspondence: D. A. Agard, Department of Biochemistry and Biophysics, University of California, San Francisco, CA 94158, USA. Tel.: 415-476-2521; Fax: 41-476-1902; Email: agard@msg.ucsf.edu

Abstract

The ubiquitous molecular chaperone Hsp90 makes up 1–2% of cytosolic proteins and is required for viability in eukaryotes. Hsp90 affects the folding and activation of a wide variety of substrate proteins including many involved in signaling and regulatory processes. Some of these substrates are implicated in cancer and other diseases, making Hsp90 an attractive drug target. Structural analyses have shown that Hsp90 is a highly dynamic and flexible molecule that can adopt a wide variety of structurally distinct states. One driving force for these rearrangements is the intrinsic ATPase activity of Hsp90, as seen with other chaperones. However, unlike other chaperones, studies have shown that the ATPase cycle of Hsp90 is not conformationally deterministic. That is, rather than dictating the conformational state, ATP binding and hydrolysis only shift the equilibria between a pre-existing set of conformational states. For bacterial, yeast and human Hsp90, there is a conserved three-state (apo–ATP–ADP) conformational cycle; however; the equilibria between states are species specific. In eukaryotes, cytosolic co-chaperones regulate the in vivo dynamic behavior of Hsp90 by shifting conformational equilibria and affecting the kinetics of structural changes and ATP hydrolysis. In this review, we discuss the structural and biochemical studies leading to our current understanding of the conformational dynamics of Hsp90, as well as the roles that nucleotide, co-chaperones, post-translational modification and substrates play. This view of Hsp90's conformational dynamics was enabled by the use of multiple complementary structural methods including, crystallography, small-angle X-ray scattering (SAXS), electron microscopy, Förster resonance energy transfer (FRET) and NMR. Finally, we discuss the effects of Hsp90 inhibitors on conformation and the potential for developing small molecules that inhibit Hsp90 by disrupting the conformational dynamics.

Type
Review Article
Copyright
Copyright © Cambridge University Press 2011

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

11. References

Alexandrescu, A. T., Abeygunawardana, C. & Shortle, D. (1994). Structure and dynamics of a denatured 131-residue fragment of staphylococcal nuclease: a heteronuclear NMR study. Biochemistry 33, 10631072.CrossRefGoogle ScholarPubMed
Alexandrescu, A. T. & Shortle, D. (1994). Backbone dynamics of a highly disordered 131 residue fragment of staphylococcal nuclease. Journal of Molecular Biology 242, 527546.CrossRefGoogle ScholarPubMed
Ali, M. M., Roe, S. M., Vaughan, C. K., Meyer, P., Panaretou, B., Piper, P. W., Prodromou, C. & Pearl, L. H. (2006). Crystal structure of an Hsp90-nucleotide-p23/Sba1 closed chaperone complex. Nature 440, 10131017.CrossRefGoogle ScholarPubMed
Allan, R. K., Mok, D., Ward, B. K. & Ratajczak, T. (2006). Modulation of chaperone function and cochaperone interaction by novobiocin in the C-terminal domain of Hsp90: evidence that coumarin antibiotics disrupt Hsp90 dimerization. Journal of Biological Chemistry 281, 71617171.CrossRefGoogle ScholarPubMed
An, W. G., Schulte, T. W. & Neckers, L. M. (2000). The heat shock protein 90 antagonist geldanamycin alters chaperone association with p210bcr-abl and v-src proteins before their degradation by the proteasome. Cell Growth and Differentiation 11, 355360.Google ScholarPubMed
Arlander, S. J., Felts, S. J., Wagner, J. M., Stensgard, B., Toft, D. O. & Karnitz, L. M. (2006). Chaperoning checkpoint kinase 1 (Chk1), an Hsp90 client, with purified chaperones. Journal of Biological Chemistry 281, 29892998.CrossRefGoogle ScholarPubMed
Bergerat, A., de Massy, B., Gadelle, D., Varoutas, P. C., Nicolas, A. & Forterre, P. (1997). An atypical topoisomerase II from Archaea with implications for meiotic recombination. Nature 386, 414417.CrossRefGoogle ScholarPubMed
Blank, M., Mandel, M., Keisari, Y., Meruelo, D. & Lavie, G (2003). Enhanced ubiquitinylation of heat shock protein 90 as a potential mechanism for mitotic cell death in cancer cells induced with hypericin. Cancer Res 63, 82418247.Google ScholarPubMed
Chadli, A., Bruinsma, E. S., Stensgard, B. & Toft, D. (2008). Analysis of Hsp90 cochaperone interactions reveals a novel mechanism for TPR protein recognition. Biochemistry 47, 28502857.CrossRefGoogle ScholarPubMed
Chadli, A., Graham, J. D., Abel, M. G., Jackson, T. A., Gordon, D. F., Wood, W. M., Felts, S. J., Horwitz, K. B. & Toft, D. (2006). GCUNC-45 is a novel regulator for the progesterone receptor/hsp90 chaperoning pathway. Molecular and Cellular Biology 26, 17221730.CrossRefGoogle ScholarPubMed
Chiosis, G., Lucas, B., Huezo, H., Solit, D., Basso, A. & Rosen, N. (2003). Development of purine-scaffold small molecule inhibitors of Hsp90. Current Cancer Drug Targets 3, 371376.CrossRefGoogle ScholarPubMed
Chu, F., Maynard, J. C., Chiosis, G., Nicchitta, C. V. & Burlingame, A. L. (2006). Identification of novel quaternary domain interactions in the Hsp90 chaperone, GRP94. Protein Science 15, 12601269.CrossRefGoogle ScholarPubMed
Cunningham, C. N., Krukenberg, K. A. & Agard, D. A. (2008). Intra- and intermonomer interactions are required to synergistically facilitate ATP hydrolysis in Hsp90. Journal of Biological Chemistry 283, 2117021178.CrossRefGoogle ScholarPubMed
Dickey, C. A., Koren, J., Zhang, Y. J., Xu, Y. F., Jinwal, U. K., Birnbaum, M. J., Monks, B., Sun, M., Cheng, J. Q., Patterson, C., Bailey, R. M., Dunmore, J., Soresh, S., Leon, C., Morgan, D. & Petrucelli, L. (2008). Akt and CHIP coregulate tau degradation through coordinated interactions. Proceedings of the National Academy of Sciences, USA 105, 36223627.CrossRefGoogle ScholarPubMed
Dittmar, K. D. & Pratt, W. B. (1997). Folding of the glucocorticoid receptor by the reconstituted Hsp90-based chaperone machinery. The initial hsp90.p60.hsp70-dependent step is sufficient for creating the steroid binding conformation. Journal of Biological Chemistry 272, 1304713054.CrossRefGoogle ScholarPubMed
Dollins, D. E., Warren, J. J., Immormino, R. M. & Gewirth, D. T. (2007). Structures of GRP94-nucleotide complexes reveal mechanistic differences between the hsp90 chaperones. Molecular Cell 28, 4156.CrossRefGoogle ScholarPubMed
Dunbrack, R. L. Jr, Gerloff, D. L., Bower, M., Chen, X., Lichtarge, O. & Cohen, F. E. (1997). Meeting review: the second meeting on the critical assessment of techniques for protein structure prediction (CASP2), Asilomar, California, December 13–16, 1996. Folding and Design 2, R27R42.CrossRefGoogle ScholarPubMed
Duval, M., Le Boeuf, F., Huot, J. & Gratton, J. P. (2007). Src-mediated phosphorylation of Hsp90 in response to vascular endothelial growth factor (VEGF) is required for VEGF receptor-2 signaling to endothelial NO synthase. Mol Biol Cell 18, 46594668.CrossRefGoogle ScholarPubMed
Falsone, S. F., Kungl, A. J., Rek, A., Cappai, R. & Zangger, K. (2009). The molecular chaperone Hsp90 modulates intermediate steps of amyloid assembly of the Parkinson-related protein alpha-synuclein. Journal of Biological Chemistry 284, 3119031199.CrossRefGoogle ScholarPubMed
Felts, S. J., Karnitz, L. M. & Toft, D. O. (2007). Functioning of the Hsp90 machine in chaperoning checkpoint kinase I (Chk1) and the progesterone receptor (PR). Cell Stress and Chaperones 12, 353363.CrossRefGoogle ScholarPubMed
Forsythe, H. L., Jarvis, J. L., Turner, J. W., Elmore, L. W. & Holt, S. E. (2001). Stable association of hsp90 and p23, but Not hsp70, with active human telomerase. Journal of Biological Chemistry 276, 1557115574.CrossRefGoogle ScholarPubMed
Frey, S., Leskovar, A., Reinstein, J. & Buchner, J. (2007). The ATPase cycle of the endoplasmic chaperone Grp94. J Biol Chem 282, 3561235620.CrossRefGoogle ScholarPubMed
Geller, R., Vignuzzi, M., Andino, R. & Frydman, J. (2007). Evolutionary constraints on chaperone-mediated folding provide an antiviral approach refractory to development of drug resistance. Genes Development 21, 195205.CrossRefGoogle ScholarPubMed
Gomez-Puertas, P., Martin-Benito, J., Carrascosa, J. L., Willison, K. R. & Valpuesta, J. M. (2004). The substrate recognition mechanisms in chaperonins. Journal of Molecular Recognition 17, 8594.CrossRefGoogle ScholarPubMed
Graf, C., Stankiewicz, M., Kramer, G. & Mayer, M. P. (2009). Spatially and kinetically resolved changes in the conformational dynamics of the Hsp90 chaperone machine. EMBO Journal 28, 602613.CrossRefGoogle ScholarPubMed
Grammatikakis, N., Lin, J. H., Grammatikakis, A., Tsichlis, P. N. & Cochran, B. H. (1999). p50(cdc37) acting in concert with Hsp90 is required for Raf-1 function. Molecular and Cellular Biology 19, 16611672.CrossRefGoogle ScholarPubMed
Grenert, J. P., Johnson, B. D. & Toft, D. O. (1999). The importance of ATP binding and hydrolysis by hsp90 in formation and function of protein heterocomplexes. Journal of Biological Chemistry 274, 1752517533.CrossRefGoogle ScholarPubMed
Harst, A., Lin, H. & Obermann, W. M. (2005). Aha1 competes with Hop, p50 and p23 for binding to the molecular chaperone Hsp90 and contributes to kinase and hormone receptor activation. Biochemical Journal 387, 789796.CrossRefGoogle Scholar
Hernandez, M. P., Chadli, A. & Toft, D. O. (2002). HSP40 binding is the first step in the HSP90 chaperoning pathway for the progesterone receptor. Journal of Biological Chemistry 277, 1187311881.CrossRefGoogle ScholarPubMed
Hessling, M., Richter, K. & Buchner, J. (2009). Dissection of the ATP-induced conformational cycle of the molecular chaperone Hsp90. Nature Structural and Molecular Biology 16, 287293.CrossRefGoogle ScholarPubMed
Immormino, R. M., Dollins, D. E., Shaffer, P. L., Soldano, K. L., Walker, M. A. & Gewirth, D. T. (2004). Ligand-induced conformational shift in the N-terminal domain of GRP94, an Hsp90 chaperone. Journal of Biological Chemistry 279, 4616246171.CrossRefGoogle ScholarPubMed
Jakob, U., Lilie, H., Meyer, I. & Buchner, J. (1995). Transient interaction of Hsp90 with early unfolding intermediates of citrate synthase. Implications for heat shock in vivo. Journal of Biological Chemistry 270, 72887294.CrossRefGoogle ScholarPubMed
Johnson, J. L. & Brown, C. (2009). Plasticity of the Hsp90 chaperone machine in divergent eukaryotic organisms. Cell Stress and Chaperones 14, 8394.CrossRefGoogle ScholarPubMed
Johnson, J. L., Halas, A. & Flom, G. (2007). Nucleotide-dependent interaction of saccharomyces cerevisiae Hsp90 with the cochaperone proteins Sti1, Cpr6, and Sba1. Molecular and Cellular Biology 27, 768776.CrossRefGoogle ScholarPubMed
Johnson, J. L. & Toft, D. O. (1995). Binding of p23 and hsp90 during assembly with the progesterone receptor. Molecular Endocrinology 9, 670678.Google ScholarPubMed
Kekatpure, V. D., Dannenberg, A. J. & Subbaramaiah, K. (2009). HDAC6 modulates Hsp90 chaperone activity and regulates activation of ary1 hydrocarbon receptor signaling. J Biol Chem 284, 74367445.CrossRefGoogle Scholar
Kosano, H., Stensgard, B., Charlesworth, M. C., McMahon, N. & Toft, D. (1998). The assembly of progesterone receptor-hsp90 complexes using purified proteins. Journal of Biological Chemistry 273, 3297332979.CrossRefGoogle ScholarPubMed
Kovacs, J. J., Murphy, P. J., Gaillard, S., Zhao, X., Wu, J. T., Nicchitta, C. V., Yoshida, M., Toft, D. O., Pratt, W. B. & Yao, T. P. (2005). HDAC6 regulates Hsp90 acetylation and chaperone-dependent activation of glucocorticoid receptor. Mol Cell 18, 601607.CrossRefGoogle ScholarPubMed
Krukenberg, K. A., Bottcher, U. M., Southworth, D. R. & Agard, D. A. (2009a). Grp94, the endoplasmic reticulum Hsp90, has a similar solution conformation to cytosolic Hsp90 in the absence of nucleotide. Protein Science 18, 18151827.CrossRefGoogle Scholar
Krukenberg, K. A., Forster, F., Rice, L. M., Sali, A. & Agard, D. A. (2008). Multiple conformations of E. coli Hsp90 in solution: insights into the conformational dynamics of Hsp90. Structure 16, 755765.CrossRefGoogle ScholarPubMed
Krukenberg, K. A., Southworth, D. R., Street, T. O. & Agard, D. A. (2009b). 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, 278291.CrossRefGoogle ScholarPubMed
Kurt, N., Rajagopalan, S. & Cavagnero, S. (2006). Effect of hsp70 chaperone on the folding and misfolding of polypeptides modeling an elongating protein chain. Journal of Molecular Biology 355, 809820.CrossRefGoogle ScholarPubMed
Lee, P., Rao, J., Fliss, A., Yang, E., Garrett, S. & Caplan, A. J. (2002). The Cdc37 protein kinase-binding domain is sufficient for protein kinase activity and cell viability. Journal of Cell Biology 159, 10511059.CrossRefGoogle ScholarPubMed
Leskovar, A., Wegele, H., Werbeck, N. D., Buchner, J. & Reinstein, J. (2008). The ATPase cycle of the mitochondrial Hsp90 analog Trap 1. J Biol Chem 283, 1167711688.CrossRefGoogle Scholar
Liu, X. D., Morano, K. A. & Thiele, D. J. (1999). The yeast Hsp110 family member, Sse1, is an Hsp90 cochaperone. Journal of Biological Chemistry 274, 2665426660.CrossRefGoogle ScholarPubMed
Lotz, G. P., Lin, H., Harst, A. & Obermann, W. M. (2003). Aha1 binds to the middle domain of Hsp90, contributes to client protein activation, and stimulates the ATPase activity of the molecular chaperone. Journal of Biological Chemistry 278, 1722817235.CrossRefGoogle Scholar
Marcu, M. G., Chadli, A., Bouhouche, I., Catelli, M. & Neckers, L. M. (2000a). The heat shock protein 90 antagonist novobiocin interacts with a previously unrecognized ATP-binding domain in the carboxyl terminus of the chaperone. Journal of Biological Chemistry 275, 3718137186.CrossRefGoogle ScholarPubMed
Marcu, M. G., Schulte, T. W. & Neckers, L. (2000b). Novobiocin and related coumarins and depletion of heat shock protein 90-dependent signaling proteins. Journal of National Cancer Institute 92, 242248.CrossRefGoogle ScholarPubMed
Martinez-Ruiz, A., Villanueva, L., Gonzalez de Orduna, C., Lopez-Ferrer, D., Higueras, M. A., Tarin, C., Rodriguez-Crespo, I., Vazquez, J. & Lamas, S. (2005). 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, 85258530.CrossRefGoogle ScholarPubMed
Martinez-Yamout, M. A., Venkitakrishnan, R. P., Preece, N. E., Kroon, G., Wright, P. E. & Dyson, H. J. (2006). Localization of sites of interaction between p23 and Hsp90 in solution. Journal of Biological Chemistry 281, 1445714464.CrossRefGoogle ScholarPubMed
McLaughlin, S. H., Smith, H. W. & Jackson, S. E. (2002). Stimulation of the weak ATPase activity of human hsp90 by a client protein. Journal of Molecular Biology 315, 787798.CrossRefGoogle ScholarPubMed
McLaughlin, S. H., Ventouras, L. A., Lobbezoo, B. & Jackson, S. E. (2004). Independent ATPase activity of hsp90 subunits creates a flexible assembly platform. J Mol Biol 344, 813826.CrossRefGoogle ScholarPubMed
McLaughlin, S. H., Sobott, F., Yao, Z. P., Zhang, W., Nielsen, P. R., Grossmann, J. G., Laue, E. D., Robinson, C. V. & Jackson, S. E. (2006). The co-chaperone p23 arrests the Hsp90 ATPase cycle to trap client proteins. Journal of Molecular Biology 356, 746758.CrossRefGoogle ScholarPubMed
Meyer, P., Prodromou, C., Hu, B., Vaughan, C., Roe, S. M., Panaretou, B., Piper, P. W. & Pearl, L. H. (2003). Structural and functional analysis of the middle segment of hsp90: implications for ATP hydrolysis and client protein and cochaperone interactions. Molecular Cell 11, 647658.CrossRefGoogle ScholarPubMed
Meyer, P., Prodromou, C., Liao, C., Hu, B., Mark Roe, S., Vaughan, C. K., Vlasic, I., Panaretou, B., Piper, P. W. & Pearl, L. H. (2004). Structural basis for recruitment of the ATPase activator Aha1 to the Hsp90 chaperone machinery. EMBO Journal 23, 511519.CrossRefGoogle Scholar
Millson, S. H., Truman, A. W., Wolfram, F., King, V., Panaretou, B., Prodromou, C., Pearl, L. H. & Piper, P. W. (2004). Investigating the protein-protein interactions of the yeast Hsp90 chaperone system by two-hybrid analysis: potential uses and limitations of this approach. Cell Stress and Chaperones 9, 359368.CrossRefGoogle ScholarPubMed
Mimnaugh, E. G., Chavany, C. & Neckers, L. (1996). Polyubiquitination and proteasomal degradation of the p185c-erbB-2 receptor protein-tyrosine kinase induced by geldanamycin. Journal of Biological Chemistry 271, 2279622801.CrossRefGoogle ScholarPubMed
Mollapour, M., Tsutsumi, S., Donnelly, A. C., Beebe, K., Tokita, M. J., Lee, M. J., Lee, S., Morra, G., Bourboulia, D., Scroggins, B. T., Colombo, G., Blagg, B. S., Panaretou, B., Stetler-Stevenson, W. G., Trepel, J. B., Piper, P. W., Prodromou, C., Pearl, L. H. & Neckers, L. (2010). Swe1Wee1-dependent tyrosine phosphorylation of Hsp90 regulates distinct facets of chaperone function. Mol Cell 37, 333343.CrossRefGoogle ScholarPubMed
Morra, G., Verkhivker, G. & Colombo, G. (2009). Modeling signal propagation mechanisms and ligand-based conformational dynamics of the Hsp90 molecular chaperone full-length dimer. PLoS Computational Biology 5, e1000323.CrossRefGoogle ScholarPubMed
Nathan, D. F. & Lindquist, S. (1995). Mutational analysis of Hsp90 function: interactions with a steroid receptor and a protein kinase. Molecular and Cellular Biology 15, 39173925.CrossRefGoogle ScholarPubMed
Neckers, L. (2007). Heat shock protein 90: the cancer chaperone. Journal of Bioscience 32, 517530.CrossRefGoogle ScholarPubMed
Neckers, L. & Ivy, S. P. (2003). Heat shock protein 90. Current Opinion in Oncology 15, 419524.CrossRefGoogle ScholarPubMed
Nishiya, Y., Shibata, K., Saito, S., Yano, K., Oneyama, C., Nakano, H. & Sharma, S. V. (2009). Drug-target identification from total cellular lysate by drug-induced conformational changes. Analytical Biochemistry 385, 314320.CrossRefGoogle ScholarPubMed
Ogiso, H., Kagi, N., Matsumoto, E., Nishimoto, M., Arai, R., Shirouzu, M., Mimura, J., Fujii-Kuriyama, Y. & Yokoyama, S.. (2004). Phosphorylation analysis of 90 kDa heat shock protein within the cytosolic arylhydrocarbon receptor complex. Biochemistry 43, 1551015519.CrossRefGoogle ScholarPubMed
Onuoha, S. C., Coulstock, E. T., Grossmann, J. G. & Jackson, S. E. (2008). Structural studies on the co-chaperone Hop and its complexes with Hsp90. Journal of Molecular Biology 379, 732744.CrossRefGoogle ScholarPubMed
Palermo, C. M., Westlake, C. A. & Gasiewicz, T. A. (2005). Epigallocatechin gallate inhibits aryl hydrocarbon receptor gene transcription through an indirect mechanism involving binding to a 90 kDa heat shock protein. Biochemistry 44, 50415052.CrossRefGoogle ScholarPubMed
Panaretou, B., Siligardi, G., Meyer, P., Maloney, A., Sullivan, J. K., Singh, S., Millson, S. H., Clarke, P. A., Naaby-Hansen, S., Stein, R., Cramer, R., Mollapour, M., Workman, P., Piper, P. W., Pearl, L. H. & Prodromou, C. (2002). Activation of the ATPase activity of hsp90 by the stress-regulated cochaperone aha1. Molecular Cell 10, 13071318.CrossRefGoogle ScholarPubMed
Pearl, L. H. & Prodromou, C. (2006). Structure and mechanism of the Hsp90 molecular chaperone machinery. Annual Review of Biochemistry 75, 271294.CrossRefGoogle ScholarPubMed
Phillips, J. J., Yao, Z. P., Zhang, W., McLaughlin, S., Laue, E. D., Robinson, C. V. & Jackson, S. E. (2007). Conformational dynamics of the molecular chaperone Hsp90 in complexes with a co-chaperone and anticancer drugs. Journal of Molecular Biology 372, 11891203.CrossRefGoogle ScholarPubMed
Prodromou, C., Panaretou, B., Chohan, S., Siligardi, G., O'Brien, R., Ladbury, J. E., Roe, S. M., Piper, P. W. & Pearl, L. H. (2000). The ATPase cycle of Hsp90 drives a molecular ‘clamp’ via transient dimerization of the N-terminal domains. EMBO Journal 19, 43834392.CrossRefGoogle ScholarPubMed
Prodromou, C., Roe, S. M., O'Brien, R., Ladbury, J. E., Piper, P. W. & Pearl, L. H. (1997a). Identification and structural characterization of the ATP/ADP-binding site in the Hsp90 molecular chaperone. Cell 90, 6575.CrossRefGoogle ScholarPubMed
Prodromou, C., Roe, S. M., Piper, P. W. & Pearl, L. H. (1997b). A molecular clamp in the crystal structure of the N-terminal domain of the yeast Hsp90 chaperone. Nature Structural Biology 4, 477482.CrossRefGoogle ScholarPubMed
Prodromou, C., Siligardi, G., O'Brien, R., Woolfson, D. N., Regan, L., Panaretou, B., Ladbury, J. E., Piper, P. W. & Pearl, L. H. (1999). Regulation of Hsp90 ATPase activity by tetratricopeptide repeat (TPR)-domain co-chaperones. EMBO Journal 18, 754762.CrossRefGoogle ScholarPubMed
Retzlaff, M., Hagn, F., Mitschke, L., Hessling, M., Gugel, F., Kessler, H., Richter, K. & Buchner, J. (2010). Asymmetric activation of the hsp90 dimer by its cochaperone aha1. Molecular Cell 37, 344354.CrossRefGoogle ScholarPubMed
Richter, K. & Buchner, J. (2001). Hsp90: chaperoning signal transduction. Journal of Cellular Physiology 188, 281290.CrossRefGoogle ScholarPubMed
Richter, K., Moser, S., Hagn, F., Friedrich, R., Hainzl, O., Heller, M., Schlee, S., Kessler, H., Reinstein, J. & Buchner, J. (2006). Intrinsic inhibition of the Hsp90 ATPase activity. Journal of Biological Chemistry 281, 1130111311.CrossRefGoogle ScholarPubMed
Richter, K., Muschler, P., Hainzl, O., Reinstein, J. & Buchner, J. (2003). Sti1 is a non-competitive inhibitor of the Hsp90 ATPase. Binding prevents the N-terminal dimerization reaction during the atpase cycle. Journal of Biological Chemistry 278, 1032810333.CrossRefGoogle ScholarPubMed
Richter, K., Walter, S. & Buchner, J. (2004). The Co-chaperone Sba1 connects the ATPase reaction of Hsp90 to the progression of the chaperone cycle. Journal of Molecular Biology 342, 14031413.CrossRefGoogle Scholar
Riggs, D. L., Roberts, P. J., Chirillo, S. C., Cheung-Flynn, J., Prapapanich, V., Ratajczak, T., Gaber, R., Picard, D. & Smith, D. F. (2003). The Hsp90-binding peptidylprolyl isomerase FKBP52 potentiates glucocorticoid signaling in vivo. EMBO Journal 22, 11581167.CrossRefGoogle ScholarPubMed
Roe, S. M., Ali, M. M., Meyer, P., Vaughan, C. K., Panaretou, B., Piper, P. W., Prodromou, C. & Pearl, L. H. (2004). The mechanism of Hsp90 regulation by the protein kinase-specific cochaperone p50(cdc37). Cell 116, 8798.CrossRefGoogle ScholarPubMed
Rudiger, S., Freund, S. M., Veprintsev, D. B. & Fersht, A. R. (2002). CRINEPT-TROSY NMR reveals p53 core domain bound in an unfolded form to the chaperone Hsp90. Proceedings of the National Academy of Sciences, USA 99, 1108511090.CrossRefGoogle Scholar
Rudiger, S., Germeroth, L., Schneider-Mergener, J. & Bukau, B. (1997). Substrate specificity of the DnaK chaperone determined by screening cellulose-bound peptide libraries. EMBO Journal 16, 15011507.CrossRefGoogle ScholarPubMed
Scheibel, T., Siegmund, H. I., Jaenicke, R., Ganz, P., Lilie, H. & Buchner, J. (1999). The charged region of Hsp90 modulates the function of the N-terminal domain. Proceedings of the National Academy of Sciences, USA 96, 12971302.CrossRefGoogle ScholarPubMed
Schulte, T. W., Blagosklonny, M. V., Ingui, C. & Neckers, L. (1995). Disruption of the Raf-1-Hsp90 molecular complex results in destabilization of Raf-1 and loss of Raf-1-Ras association. Journal of Biological Chemistry 270, 2458524588.CrossRefGoogle ScholarPubMed
Scroggins, B. T., Robzyk, K., Wang, D., Marcu, M. G., Tsutsumi, S., Beebe, K., Cotter, R. J., Felts, S., Toft, D., Karnitz, L., Rosen, N. & Neckers, L. (2007). An acetylation site in the middle domain of Hsp90 regulates chaperone function. Mol Cell 25, 151159.CrossRefGoogle ScholarPubMed
Shah, V., Wiest, R., Garcia-Cardena, G., Cadelina, G., Groszmann, R. J. & Sessa, W. C. (1999). Hsp90 regulation of endothelial nitric oxide synthase contributes to vascular control in portal hypertension. American Journal of Physiology 277, G463G468.Google ScholarPubMed
Shao, J., Irwin, A., Hartson, S. D. & Matts, R. L. (2003). Functional dissection of cdc37: characterization of domain structure and amino acid residues critical for protein kinase binding. Biochemistry 42, 1257712588.CrossRefGoogle ScholarPubMed
Shiau, A. K., Harris, S. F., Southworth, D. R. & Agard, D. A. (2006). Structural Analysis of E. coli hsp90 reveals dramatic nucleotide-dependent conformational rearrangements. Cell 127, 329340.CrossRefGoogle ScholarPubMed
Shortle, D. (2002). The expanded denatured state: an ensemble of conformations trapped in a locally encoded topological space. Advances in Protein Chemistry 62, 123.CrossRefGoogle Scholar
Siligardi, G., Hu, B., Panaretou, B., Piper, P. W., Pearl, L. H. & Prodromou, C. (2004). Co-chaperone regulation of conformational switching in the Hsp90 ATPase cycle. Journal of Biological Chemistry 279, 5198951998.CrossRefGoogle ScholarPubMed
Smith, D. F. (1993). Dynamics of heat shock protein 90-progesterone receptor binding and the disactivation loop model for steroid receptor complexes. Molecular Endocrinology 7, 14181429.Google ScholarPubMed
Smith, D. F., Faber, L. E. & Toft, D. O. (1990). Purification of unactivated progesterone receptor and identification of novel receptor-associated proteins. Journal of Biological Chemistry 265, 39964003.CrossRefGoogle ScholarPubMed
Solit, D. B., Osman, I., Polsky, D., Panageas, K. S., Daud, A., Goydos, J. S., Teitcher, J., Wolchok, J. D., Germino, F. J., Krown, S. E., Coit, D., Rosen, N. & Chapman, P. B. (2008). Phase II trial of 17-allylamino-17-demethoxygeldanamycin in patients with metastatic melanoma. Clinical Cancer Research 14, 83028307.CrossRefGoogle ScholarPubMed
Soti, C., Racz, A. & Csermely, P. (2002). A nucleotide-dependent molecular switch controls ATP binding at the C-terminal domain of Hsp90. N-terminal nucleotide binding unmasks a C-terminal binding pocket. Journal of Biological Chemistry 277, 70667075.CrossRefGoogle ScholarPubMed
Southworth, D. R. & Agard, D. A. (2008). Species-dependent ensembles of conserved conformational states define the Hsp90 chaperone ATPase cycle. Molecular Cell 32, 631640.CrossRefGoogle ScholarPubMed
Stebbins, C. E., Russo, A. A., Schneider, C., Rosen, N., Hartl, F. U. & Pavletich, N. P. (1997). Crystal structure of an Hsp90-geldanamycin complex: targeting of a protein chaperone by an antitumor agent. Cell 89, 239250.CrossRefGoogle ScholarPubMed
Stepanova, L., Leng, X., Parker, S. B. & Harper, J. W. (1996). Mammalian p50Cdc37 is a protein kinase-targeting subunit of Hsp90 that binds and stabilizes Cdk4. Genes and Development 10, 14911502.CrossRefGoogle ScholarPubMed
Street, T. O., Krukenberg, K. A., Rosgen, J., Bolen, D. W. & Agard, D. A. (2010). Osmolyte-induced conformational changes in the Hsp90 molecular chaperone. Protein Science 19, 5765.CrossRefGoogle ScholarPubMed
Sullivan, W., Stensgard, B., Caucutt, G., Bartha, B., McMahon, N., Alnemri, E. S., Litwack, G. & Toft, D. (1997). Nucleotides and two functional states of hsp90. Journal of Biological Chemistry 272, 80078012.CrossRefGoogle ScholarPubMed
Te, J., Jia, L., Rogers, J., Miller, A. & Hartson, S. D. (2007). Novel subunits of the mammalian Hsp90 signal transduction chaperone. Journal of Proteome Research 6, 19631973.CrossRefGoogle ScholarPubMed
Vasko, R. C., Rodriguez, R. A., Cunningham, C. N., Ardi, V. C., Agard, D. A. & McAlpine, S. R. (2010). Mechanistic studies of sansalvamide a-amide: an allosteric modulator of Hsp90. ACS Medicinal Chemistry Letters 1, 48.CrossRefGoogle ScholarPubMed
Vaughan, C. K., Gohlke, U., Sobott, F., Good, V. M., Ali, M. M., Prodromou, C., Robinson, C. V., Saibil, H. R. & Pearl, L. H. (2006). Structure of an Hsp90-Cdc37-Cdk4 complex. Molecular Cell 23, 697707.CrossRefGoogle ScholarPubMed
Walter, S. & Buchner, J. (2002). Molecular chaperones – cellular machines for protein folding. Angewandte Chem (International Edition in English) 41, 10981113.3.0.CO;2-9>CrossRefGoogle ScholarPubMed
Wandinger, S. K., Richster, K. & Buchner, J. (2008). The Hsp90 chaperone machinery. J Biol Chem 283, 1847318477.CrossRefGoogle ScholarPubMed
Wegele, H., Muschler, P., Bunck, M., Reinstein, J. & Buchner, J. (2003). Dissection of the contribution of individual domains to the ATPase mechanism of Hsp90. Journal of Biological Chemistry 278, 3930339310.CrossRefGoogle Scholar
Wegele, H., Wandinger, S. K., Schmid, A. B., Reinstein, J. & Buchner, J. (2006). Substrate transfer from the chaperone Hsp70 to Hsp90. Journal of Molecular Biology 356, 802811.CrossRefGoogle ScholarPubMed
Weissman, J. S., Hohl, C. M., Kovalenko, O., Kashi, Y., Chen, S., Braig, K., Saibil, H. R., Fenton, W. A. & Horwich, A. L. (1995). Mechanism of GroEL action: productive release of polypeptide from a sequestered position under GroES. Cell 83, 577587.CrossRefGoogle ScholarPubMed
Whitesell, L., Mimnaugh, E. G., De Costa, B., Myers, C. E. & Neckers, L. M. (1994). Inhibition of heat shock protein HSP90-pp60v-src heteroprotein complex formation by benzoquinone ansamycins: essential role for stress proteins in oncogenic transformation. Proceedings of the National Academy of Sciences, USA 91, 83248328.CrossRefGoogle ScholarPubMed
Weikl, T., Muschler, P., Richter, K., Veit, T., Reinstein, J. & Buchner, J. (2000). C-terminal regions of Hsp90 are important for trapping the nucleotide during the ATPase cycle. J Mol Biol 303, 583592.CrossRefGoogle ScholarPubMed
Wiech, H., Buchner, J., Zimmermann, R. & Jakob, U. (1992). Hsp90 chaperones protein folding in vitro. Nature 358, 169170.CrossRefGoogle ScholarPubMed
Workman, P. (2004). Combinatorial attack on multistep oncogenesis by inhibiting the Hsp90 molecular chaperone. Cancer Letters 206, 149157.CrossRefGoogle ScholarPubMed
Young, J. C., Moarefi, I. & Hartl, F. U. (2001). Hsp90: a specialized but essential protein-folding tool. Journal of Cell Biology 154, 267273.CrossRefGoogle ScholarPubMed
Yu, X., Guo, Z. S., Marcu, M. G., Neckers, L., Nguyen, D. M., Chen, G. A. & Schrump, D. S. (2002). Modulation of p53, ErbB1, ErbB2, and Raf-1 expression in lung cancer cells by depsipeptide FR901228. J Natl Cancer Inst 94, 504513.CrossRefGoogle ScholarPubMed
Yun, B. G., Huang, W., Leach, N., Hartson, S. D. & Matts, R. L. (2004). Novobiocin induces a distinct conformation of hsp90 and alters hsp90-cochaperone-client interactions. Biochemistry 43, 82178229.CrossRefGoogle ScholarPubMed
Zhang, W., Hirshberg, M., McLaughlin, S. H., Lazar, G. A., Grossmann, J. G., Nielsen, P. R., Sobott, F., Robinson, C. V., Jackson, S. E. & Laue, E. D. (2004). Biochemical and structural studies of the interaction of Cdc37 with Hsp90. Journal of Molecular Biology 340, 891907.CrossRefGoogle ScholarPubMed
Zhao, R., Davey, M., Hsu, Y. C., Kaplanek, P., Tong, A., Parsons, A. B., Krogan, N., Cagney, G., Mai, D., Greenblatt, J., Boone, C., Emili, A. & Houry, W. A. (2005). Navigating the chaperone network: an integrative map of physical and genetic interactions mediated by the hsp90 chaperone. Cell 120, 715727.CrossRefGoogle ScholarPubMed