Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Review Article
  • Published:

Heat shock transcription factor 1 as a therapeutic target in neurodegenerative diseases

Key Points

  • For neurodegenerative diseases such as Huntington's disease, spinocerebellar muscular atrophy, amyotrophic lateral sclerosis, Parkinson's disease and Alzheimer's disease there is a lack of effective treatments that directly address the underlying biochemical aetiology of neuronal dysfunction and cell death.

  • Protein misfolding, cellular stress and neuronal cell death are common features of neurodegenerative diseases.

  • A diverse set of chaperone proteins act in concert to fold misfolded proteins, disaggregate damaged proteins and prevent programmed cell death.

  • Heat shock transcription factor 1 (HSF1) coordinately activates the expression of chaperone protein gene expression.

  • Genetic and pharmacological experiments in cell culture, fruitfly and mouse models of neurodegenerative disease suggest that enhancing the cellular protein folding and anti-apoptotic machinery by elevating levels of chaperone proteins could have potential therapeutic efficacy in neurodegenerative diseases.

  • Current small-molecule HSF1 activators have undesirable properties — including direct proteotoxicity, inhibition of the central cellular chaperone heat shock protein 90 and other characteristics — that limit their development for clinical use.

  • As the master activator of chaperone protein expression, HSF1 is an attractive pharmacological target for the development of optimized small-molecule activators for therapeutic intervention in neurodegenerative diseases.

Abstract

Neurodegenerative diseases such as Alzheimer's disease, Parkinson's disease, Huntington's disease, amyotrophic lateral sclerosis and prion-based neurodegeneration are associated with the accumulation of misfolded proteins, resulting in neuronal dysfunction and cell death. However, current treatments for these diseases predominantly address disease symptoms, rather than the underlying protein misfolding and cell death, and are not able to halt or reverse the degenerative process. Studies in cell culture, fruitfly, worm and mouse models of protein misfolding-based neurodegenerative diseases indicate that enhancing the protein-folding capacity of cells, via elevated expression of chaperone proteins, has therapeutic potential. Here, we review advances in strategies to harness the power of the natural cellular protein-folding machinery through pharmacological activation of heat shock transcription factor 1 — the master activator of chaperone protein gene expression — to treat neurodegenerative diseases.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Chaperone proteins and maintenance of protein homeostasis.
Figure 2: HSF1 activation and attenuation cycle.
Figure 3: HSF1 regulation by post-translational modifications.
Figure 4: Proposed mechanisms to promote the direct activation of HSF1.
Figure 5: Importance of early intervention in neurodegenerative diseases.

Similar content being viewed by others

References

  1. Ross, C. A. & Poirier, M. A. Protein aggregation and neurodegenerative disease. Nature Med. 10, S10–S17 (2004).

    Article  CAS  PubMed  Google Scholar 

  2. Zhang, Q. C. et al. A compact β model of huntingtin toxicity. J. Biol. Chem. 286, 8188–8196 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Haass, C. & Selkoe, D. J. Soluble protein oligomers in neurodegeneration: lessons from the Alzheimer's amyloid β-peptide. Nature Rev. Mol. Cell Biol. 8, 101–112 (2007).

    Article  CAS  Google Scholar 

  4. Muchowski, P. J. Protein misfolding, amyloid formation, and neurodegeneration: a critical role for molecular chaperones? Neuron 35, 9–12 (2002).

    Article  CAS  PubMed  Google Scholar 

  5. Arrasate, M., Mitra, S., Schweitzer, E. S., Segal, M. R. & Finkbeiner, S. Inclusion body formation reduces levels of mutant huntingtin and the risk of neuronal death. Nature 431, 805–810 (2004).

    Article  CAS  PubMed  Google Scholar 

  6. Fiumara, F., Fioriti, L., Kandel, E. R. & Hendrickson, W. A. Essential role of coiled coils for aggregation and activity of Q/N-rich prions and polyQ proteins. Cell 143, 1121–1135 (2010). This study demonstrated that coiled-coil motifs in polyQ proteins contribute to the aggregation and cytotoxicity of these proteins.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Cookson, M. R. The role of leucine-rich repeat kinase 2 (LRRK2) in Parkinson's disease. Nature Rev. Neurosci. 11, 791–797 (2010).

    Article  CAS  Google Scholar 

  8. Jankovic, J. Parkinson's disease: clinical features and diagnosis. J. Neurol. Neurosurg. Psychiatr. 79, 368–376 (2008).

    Article  CAS  Google Scholar 

  9. Buschert, V., Bokde, A. L. W. & Hampel, H. Cognitive intervention in Alzheimer disease. Nature Rev. Neurol. 6, 508–517 (2010).

    Article  CAS  Google Scholar 

  10. Carter, M. D., Simms, G. A. & Weaver, D. F. The development of new therapeutics for Alzheimer's disease. Clin. Pharmacol. Ther. 88, 475–486 (2010).

    Article  CAS  PubMed  Google Scholar 

  11. Boillée, S., Vande Velde, C. & Cleveland, D. W. ALS: a disease of motor neurons and their nonneuronal neighbors. Neuron 52, 39–59 (2006).

    Article  CAS  PubMed  Google Scholar 

  12. Verity, N. C. & Mallucci, G. R. Rescuing neurons in prion disease. Biochem. J. 433, 19–29 (2010).

    Article  CAS  Google Scholar 

  13. Walker, F. O. Huntington's disease. Lancet 369, 218–228 (2007).

    Article  CAS  PubMed  Google Scholar 

  14. Hartl, F. U. & Hayer-Hartl, M. Molecular chaperones in the cytosol: from nascent chain to folded protein. Science 295, 1852–1858 (2002).

    Article  CAS  PubMed  Google Scholar 

  15. Chai, Y., Koppenhafer, S. L., Bonini, N. M. & Paulson, H. L. Analysis of the role of heat shock protein (Hsp) molecular chaperones in polyglutamine disease. J. Neurosci. 19, 10338–10347 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Warrick, J. M. et al. Suppression of polyglutamine-mediated neurodegeneration in Drosophila by the molecular chaperone HSP70. Nature Genet. 23, 425–428 (1999).

    Article  CAS  PubMed  Google Scholar 

  17. Chan, H. Y., Warrick, J. M., Gray-Board, G. L., Paulson, H. L. & Bonini, N. M. Mechanisms of chaperone suppression of polyglutamine disease: selectivity, synergy and modulation of protein solubility in Drosophila. Hum. Mol. Genet. 9, 2811–2820 (2000). This study showed that HSP40 and HSP70 synergize to ameliorate the cytotoxicity of polyQ proteins in fruitfly disease models by modulating the solubility of these proteins.

    Article  CAS  PubMed  Google Scholar 

  18. Auluck, P. K. & Bonini, N. M. Pharmacological prevention of Parkinson disease in Drosophila. Nature Med. 8, 1185–1186 (2002). This paper showed that pharmacological activation of HSF1 via the HSP90 inhibitor geldanamycin can ameliorate disease phenotypes in a fruitfly model of Parkinson's disease.

    Article  CAS  PubMed  Google Scholar 

  19. Auluck, P., Meulener, M. & Bonini, N. Mechanisms of suppression of α-synuclein neurotoxicity by geldanamycin in Drosophila. J. Biol. Chem. 280, 2873–2878 (2005).

    Article  CAS  PubMed  Google Scholar 

  20. Alavez, S., Vantipalli, M. C., Zucker, D. J., Klang, I. M. & Lithgow, G. J. Amyloid-binding compounds maintain protein homeostasis during ageing and extend lifespan. Nature 472, 226–229 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Ben-Zvi, A., Miller, E. A. & Morimoto, R. I. Collapse of proteostasis represents an early molecular event in Caenorhabditis elegans aging. Proc. Natl Acad. Sci. USA 106, 14914–14919 (2009). This study describes a widespread failure in protein folding that occurs in early adulthood and coincides with reduced activation of HSF1 and chaperone protein expression in C. elegans.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Fonte, V. et al. Interaction of intracellular β amyloid peptide with chaperone proteins. Proc. Natl Acad. Sci. USA 99, 9439–9444 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Satyal, S. H. et al. Polyglutamine aggregates alter protein folding homeostasis in Caenorhabditis elegans. Proc. Natl Acad. Sci. USA 97, 5750–5755 (2000). This study shows that the expression of polyQ proteins in C. elegans disrupts general protein folding, causes aggregation of otherwise soluble proteins and constitutively promotes the activation of HSF1 and chaperone proteins.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Teixeira-Castro, A. et al. Neuron-specific proteotoxicity of mutant ataxin-3 in C. elegans: rescue by the DAF-16 and HSF-1 pathways. Hum. Mol. Genet. 20, 2996–3009 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Wang, J. et al. An ALS-linked mutant SOD1 produces a locomotor defect associated with aggregation and synaptic dysfunction when expressed in neurons of Caenorhabditis elegans. PLoS Genet. 5, e1000350 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Lanneau, D., de Thonel, A., Maurel, S., Didelot, C. & Garrido, C. Apoptosis versus cell differentiation: role of heat shock proteins HSP90, HSP70 and HSP27. Prion 1, 53–60 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  27. Batulan, Z. et al. High threshold for induction of the stress response in motor neurons is associated with failure to activate HSF1. J. Neurosci. 23, 5789–5798 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Bonelli, M. A. et al. Attenuated expression of 70-kDa heat shock protein in WI-38 human fibroblasts during aging in vitro. Exp. Cell Res. 252, 20–32 (1999).

    Article  CAS  PubMed  Google Scholar 

  29. Gutsmann-Conrad, A., Heydari, A. R., You, S. & Richardson, A. The expression of heat shock protein 70 decreases with cellular senescence in vitro and in cells derived from young and old human subjects. Exp. Cell Res. 241, 404–413 (1998).

    Article  CAS  PubMed  Google Scholar 

  30. Gutsmann-Conrad, A., Pahlavani, M. A., Heydari, A. R. & Richardson, A. Expression of heat shock protein 70 decreases with age in hepatocytes and splenocytes from female rats. Mech. Ageing Dev. 107, 255–270 (1999).

    Article  CAS  PubMed  Google Scholar 

  31. Fargnoli, J., Kunisada, T., Fornace, A. J., Schneider, E. L. & Holbrook, N. J. Decreased expression of heat shock protein 70 mRNA and protein after heat treatment in cells of aged rats. Proc. Natl Acad. Sci. USA 87, 846–850 (1990).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Fawcett, T. W., Sylvester, S. L., Sarge, K. D., Morimoto, R. I. & Holbrook, N. J. Effects of neurohormonal stress and aging on the activation of mammalian heat shock factor 1. J. Biol. Chem. 269, 32272–32278 (1994).

    CAS  PubMed  Google Scholar 

  33. Pahlavani, M. A., Harris, M. D., Moore, S. A., Weindruch, R. & Richardson, A. The expression of heat shock protein 70 decreases with age in lymphocytes from rats and rhesus monkeys. Exp. Cell Res. 218, 310–318 (1995).

    Article  CAS  PubMed  Google Scholar 

  34. Bailey, C. K., Andriola, I. F. M., Kampinga, H. H. & Merry, D. E. Molecular chaperones enhance the degradation of expanded polyglutamine repeat androgen receptor in a cellular model of spinal and bulbar muscular atrophy. Hum. Mol. Genet. 11, 515–523 (2002).

    Article  CAS  PubMed  Google Scholar 

  35. Fujimoto, M. et al. Active HSF1 significantly suppresses polyglutamine aggregate formation in cellular and mouse models. J. Biol. Chem. 280, 34908–34916 (2005). This study demonstrates that the expression of a constitutively active HSF1 allele ameliorates pathogenic phenotypes in a mouse model of Huntington's disease.

    Article  CAS  PubMed  Google Scholar 

  36. Muchowski, P. J. et al. Hsp70 and Hsp40 chaperones can inhibit self-assembly of polyglutamine proteins into amyloid-like fibrils. Proc. Natl Acad. Sci. USA 97, 7841–7846 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Wacker, J. L., Zareie, M. H., Fong, H., Sarikaya, M. & Muchowski, P. J. Hsp70 and Hsp40 attenuate formation of spherical and annular polyglutamine oligomers by partitioning monomer. Nature Struct. Mol. Biol. 11, 1215–1222 (2004).

    Article  CAS  Google Scholar 

  38. Wyttenbach, A. et al. Effects of heat shock, heat shock protein 40 (HDJ-2), and proteasome inhibition on protein aggregation in cellular models of Huntington's disease. Proc. Natl Acad. Sci. USA 97, 2898–2903 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Feder, J. H., Rossi, J. M., Solomon, J., Solomon, N. & Lindquist, S. The consequences of expressing hsp70 in Drosophila cells at normal temperatures. Genes Dev. 6, 1402–1413 (1992).

    Article  CAS  PubMed  Google Scholar 

  40. Dai, C., Whitesell, L., Rogers, A. B. & Lindquist, S. Heat shock factor 1 is a powerful multifaceted modifier of carcinogenesis. Cell 130, 1005–1018 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Akerfelt, M., Morimoto, R. I. & Sistonen, L. Heat shock factors: integrators of cell stress, development and lifespan. Nature Rev. Mol. Cell Biol. 11, 545–555 (2010).

    Article  CAS  Google Scholar 

  42. Gonsalves, S. E., Moses, A. M., Razak, Z., Robert, F. & Westwood, J. T. Whole-genome analysis reveals that active heat shock factor binding sites are mostly associated with non-heat shock genes in Drosophila melanogaster. PLoS ONE 6, e15934 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Hahn, J.-S., Hu, Z., Thiele, D. J. & Iyer, V. R. Genome-wide analysis of the biology of stress responses through heat shock transcription factor. Mol. Cell Biol. 24, 5249–5256 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Trinklein, N. D., Murray, J. I., Hartman, S. J., Botstein, D. & Myers, R. M. The role of heat shock transcription factor 1 in the genome-wide regulation of the mammalian heat shock response. Mol. Biol. Cell 15, 1254–1261 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Ostling, P., Björk, J. K., Roos-Mattjus, P., Mezger, V. & Sistonen, L. Heat shock factor 2 (HSF2) contributes to inducible expression of hsp genes through interplay with HSF1. J. Biol. Chem. 282, 7077–7086 (2007).

    Article  CAS  PubMed  Google Scholar 

  46. Sandqvist, A. et al. Heterotrimerization of heat-shock factors 1 and 2 provides a transcriptional switch in response to distinct stimuli. Mol. Biol. Cell 20, 1340–1347 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Shinkawa, T. et al. Heat shock factor 2 is required for maintaining proteostasis against febrile range thermal stress and polyglutamine aggregation. Mol. Biol. Cell 22, 3571–3583 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. 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).

    Article  CAS  PubMed  Google Scholar 

  49. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Conde, R., Xavier, J., McLoughlin, C., Chinkers, M. & Ovsenek, N. Protein phosphatase 5 is a negative modulator of heat shock factor 1. J. Biol. Chem. 280, 28989–28996 (2005).

    Article  CAS  PubMed  Google Scholar 

  52. 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).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. 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).

    Article  CAS  PubMed  Google Scholar 

  55. Arlander, S. J. H. et al. Chaperoning checkpoint kinase 1 (Chk1), an Hsp90 client, with purified chaperones. J. Biol. Chem. 281, 2989–2998 (2006).

    Article  CAS  PubMed  Google Scholar 

  56. Hernández, 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).

    Article  CAS  PubMed  Google Scholar 

  57. King, F. W., Wawrzynow, A., Höhfeld, J. & Zylicz, M. Co-chaperones Bag-1, Hop and Hsp40 regulate Hsc70 and Hsp90 interactions with wild-type or mutant p53. EMBO J. 20, 6297–6305 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Carmichael, J., Sugars, K. L., Bao, Y. P. & Rubinsztein, D. C. Glycogen synthase kinase-3β inhibitors prevent cellular polyglutamine toxicity caused by the Huntington's disease mutation. J. Biol. Chem. 277, 33791–33798 (2002).

    Article  CAS  PubMed  Google Scholar 

  59. Chu, B., Soncin, F., Price, B. D., Stevenson, M. A. & Calderwood, S. K. Sequential phosphorylation by mitogen-activated protein kinase and glycogen synthase kinase 3 represses transcriptional activation by heat shock factor-1. J. Biol. Chem. 271, 30847–30857 (1996).

    Article  CAS  PubMed  Google Scholar 

  60. Chu, B., Zhong, R., Soncin, F., Stevenson, M. A. & Calderwood, S. K. Transcriptional activity of heat shock factor 1 at 37 °C is repressed through phosphorylation on two distinct serine residues by glycogen synthase kinase 3 and protein kinases Cα and Cζ. J. Biol. Chem. 273, 18640–18646 (1998).

    Article  CAS  PubMed  Google Scholar 

  61. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Kline, M. P. & Morimoto, R. I. Repression of the heat shock factor 1 transcriptional activation domain is modulated by constitutive phosphorylation. Mol. Cell Biol. 17, 2107–2115 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Knauf, U., Newton, E. M., Kyriakis, J. & Kingston, R. E. Repression of human heat shock factor 1 activity at control temperature by phosphorylation. Genes Dev. 10, 2782–2793 (1996).

    Article  CAS  PubMed  Google Scholar 

  64. Murshid, A. et al. Protein kinase A binds and activates heat shock factor 1. PLoS ONE 5, e13830 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Wang, X. et al. Phosphorylation of HSF1 by MAPK-activated protein kinase 2 on serine 121, inhibits transcriptional activity and promotes HSP90 binding. J. Biol. Chem. 281, 782–791 (2006).

    Article  CAS  PubMed  Google Scholar 

  66. Pelham, H. R. A regulatory upstream promoter element in the Drosophila Hsp 70 heat-shock gene. Cell 30, 517–528 (1982).

    Article  CAS  PubMed  Google Scholar 

  67. Pelham, H. R. & Bienz, M. A synthetic heat-shock promoter element confers heat-inducibility on the herpes simplex virus thymidine kinase gene. EMBO J. 1, 1473–1477 (1982).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Perisic, O., Xiao, H. & Lis, J. T. Stable binding of Drosophila heat shock factor to head-to-head and tail-to-tail repeats of a conserved 5 bp recognition unit. Cell 59, 797–806 (1989).

    Article  CAS  PubMed  Google Scholar 

  69. Clos, J. et al. Molecular cloning and expression of a hexameric Drosophila heat shock factor subject to negative regulation. Cell 63, 1085–1097 (1990).

    Article  CAS  PubMed  Google Scholar 

  70. Sorger, P. K. & Nelson, H. C. Trimerization of a yeast transcriptional activator via a coiled-coil motif. Cell 59, 807–813 (1989).

    Article  CAS  PubMed  Google Scholar 

  71. Ahn, S.-G. & Thiele, D. J. Redox regulation of mammalian heat shock factor 1 is essential for Hsp gene activation and protection from stress. Genes Dev. 17, 516–528 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Neef, D. W., Turski, M. L. & Thiele, D. J. Modulation of heat shock transcription factor 1 as a therapeutic target for small molecule intervention in neurodegenerative disease. PLoS Biol. 8, e1000291 (2010). In this study the authors generated a humanized HSF1-based yeast screen to identify HSF1A, a novel pharmacological activator of HSF1 that is efficacious in ameliorating polyQ protein-associated protein aggregation and cytotoxicity in cell culture and fruitfly disease models.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Trott, A. et al. Activation of heat shock and antioxidant responses by the natural product celastrol: transcriptional signatures of a thiol-targeted molecule. Mol. Biol. Cell 19, 1104–1112 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Rabindran, S. K., Haroun, R. I., Clos, J., Wisniewski, J. & Wu, C. Regulation of heat shock factor trimer formation: role of a conserved leucine zipper. Science 259, 230–234 (1993).

    Article  CAS  PubMed  Google Scholar 

  75. Guettouche, T., Boellmann, F., Lane, W. S. & Voellmy, R. Analysis of phosphorylation of human heat shock factor 1 in cells experiencing a stress. BMC Biochem. 6, 4 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Holmberg, C. I. et al. Phosphorylation of serine 230 promotes inducible transcriptional activity of heat shock factor 1. EMBO J. 20, 3800–3810 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Kim, S.-A., Yoon, J.-H., Lee, S.-H. & Ahn, S.-G. Polo-like kinase 1 phosphorylates heat shock transcription factor 1 and mediates its nuclear translocation during heat stress. J. Biol. Chem. 280, 12653–12657 (2005).

    Article  CAS  PubMed  Google Scholar 

  78. Westerheide, S. D., Anckar, J., Stevens, S. M., Sistonen, L. & Morimoto, R. I. Stress-inducible regulation of heat shock factor 1 by the deacetylase SIRT1. Science 323, 1063–1066 (2009). This study demonstrated that the DNA binding activity of HSF1 is inhibited by acetylation within the DNA binding domain, and HSF1 is maintained in a deacetylated state via SIRT1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Yang, J., Bridges, K., Chen, K. Y. & Liu, A. Y.-C. Riluzole increases the amount of latent HSF1 for an amplified heat shock response and cytoprotection. PLoS ONE 3, e2864 (2008). This work reported that riluzole, which is a treatment for ALS, promotes an increase in steady-state HSF1 levels potentially via the inhibition of chaperone-mediated autophagy.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Trepel, J., Mollapour, M., Giaccone, G. & Neckers, L. Targeting the dynamic HSP90 complex in cancer. Nature Rev. Cancer 10, 537–549 (2010).

    Article  CAS  Google Scholar 

  81. Dickey, C. A. et al. HSP induction mediates selective clearance of tau phosphorylated at proline-directed Ser/Thr sites but not KXGS (MARK) sites. FASEB J. 20, 753–755 (2006).

    Article  CAS  PubMed  Google Scholar 

  82. Dickey, C. A. et al. Development of a high throughput drug screening assay for the detection of changes in tau levels — proof of concept with HSP90 inhibitors. Curr. Alzheimer Res. 2, 231–238 (2005).

    Article  CAS  PubMed  Google Scholar 

  83. Dou, F. et al. Chaperones increase association of tau protein with microtubules. Proc. Natl Acad. Sci. USA 100, 721–726 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Petrucelli, L. et al. CHIP and Hsp70 regulate tau ubiquitination, degradation and aggregation. Hum. Mol. Genet. 13, 703–714 (2004).

    Article  CAS  PubMed  Google Scholar 

  85. Flower, T. R., Chesnokova, L. S., Froelich, C. A., Dixon, C. & Witt, S. N. Heat shock prevents α-synuclein-induced apoptosis in a yeast model of Parkinson's disease. J. Mol. Biol. 351, 1081–1100 (2005).

    Article  CAS  PubMed  Google Scholar 

  86. Shen, H.-Y., He, J.-C., Wang, Y., Huang, Q.-Y. & Chen, J.-F. Geldanamycin induces heat shock protein 70 and protects against MPTP-induced dopaminergic neurotoxicity in mice. J. Biol. Chem. 280, 39962–39969 (2005).

    Article  CAS  PubMed  Google Scholar 

  87. Agrawal, N. et al. Identification of combinatorial drug regimens for treatment of Huntington's disease using Drosophila. Proc. Natl Acad. Sci. USA 102, 3777–3781 (2005). This study demonstrated that the HSP90 inhibitor geldanamycin and the histone deacetylase inhibitor suberoylanilide hydroxamic acid have combinatorial efficacy in ameliorating cytotoxicity in a fruitfly model of neurodegenerative disease.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Fujikake, N. et al. Heat shock transcription factor 1-activating compounds suppress polyglutamine-induced neurodegeneration through induction of multiple molecular chaperones. J. Biol. Chem. 283, 26188–26197 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Hay, D. G. et al. Progressive decrease in chaperone protein levels in a mouse model of Huntington's disease and induction of stress proteins as a therapeutic approach. Hum. Mol. Genet. 13, 1389–1405 (2004).

    Article  CAS  PubMed  Google Scholar 

  90. Sittler, A. et al. Geldanamycin activates a heat shock response and inhibits huntingtin aggregation in a cell culture model of Huntington's disease. Hum. Mol. Genet. 10, 1307–1315 (2001).

    Article  CAS  PubMed  Google Scholar 

  91. 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).

    Article  CAS  PubMed  Google Scholar 

  92. Yu, X. M. et al. Hsp90 inhibitors identified from a library of novobiocin analogues. J. Am. Chem. Soc. 127, 12778–12779 (2005).

    Article  CAS  PubMed  Google Scholar 

  93. Ansar, S. et al. A non-toxic Hsp90 inhibitor protects neurons from Aβ-induced toxicity. Bioorg. Med. Chem. Lett. 17, 1984–1990 (2007).

    Article  CAS  PubMed  Google Scholar 

  94. Kimura, H. et al. ITZ-1, a client-selective Hsp90 inhibitor, efficiently induces heat shock factor 1 activation. Chem. Biol. 17, 18–27 (2010).

    Article  CAS  PubMed  Google Scholar 

  95. Salehi, A. H. et al. AEG3482 is an antiapoptotic compound that inhibits Jun kinase activity and cell death through induced expression of heat shock protein 70. Chem. Biol. 13, 213–223 (2006).

    Article  CAS  PubMed  Google Scholar 

  96. Schnaider, T., Somogyi, J., Csermely, P. & Szamel, M. The Hsp90-specific inhibitor, geldanamycin, blocks CD28-mediated activation of human T lymphocytes. Life Sci. 63, 949–954 (1998).

    Article  CAS  PubMed  Google Scholar 

  97. Westerheide, S. et al. Celastrols as inducers of the heat shock response and cytoprotection. J. Biol. Chem. 279, 56053–56060 (2004).

    Article  CAS  PubMed  Google Scholar 

  98. Hieronymus, H. et al. Gene expression signature-based chemical genomic prediction identifies a novel class of HSP90 pathway modulators. Cancer Cell 10, 321–330 (2006).

    Article  CAS  PubMed  Google Scholar 

  99. Zhang, T. et al. A novel Hsp90 inhibitor to disrupt Hsp90/Cdc37 complex against pancreatic cancer cells. Mol. Cancer Ther. 7, 162–170 (2008).

    Article  CAS  PubMed  Google Scholar 

  100. Yang, H., Chen, D., Cui, Q. C., Yuan, X. & Dou, Q. P. Celastrol, a triterpene extracted from the Chinese “Thunder of God Vine,” is a potent proteasome inhibitor and suppresses human prostate cancer growth in nude mice. Cancer Res. 66, 4758–4765 (2006).

    Article  CAS  PubMed  Google Scholar 

  101. Kiaei, M. et al. Celastrol blocks neuronal cell death and extends life in transgenic mouse model of amyotrophic lateral sclerosis. Neurodegener. Dis. 2, 246–254 (2005).

    Article  CAS  PubMed  Google Scholar 

  102. Allison, A. C., Cacabelos, R., Lombardi, V. R., Alvarez, X. A. & Vigo, C. Celastrol, a potent antioxidant and anti-inflammatory drug, as a possible treatment for Alzheimer's disease. Prog. Neuropsychopharmacol. Biol. Psychiatry 25, 1341–1357 (2001).

    Article  CAS  PubMed  Google Scholar 

  103. Wang, J., Gines, S., MacDonald, M. E. & Gusella, J. F. Reversal of a full-length mutant huntingtin neuronal cell phenotype by chemical inhibitors of polyglutamine-mediated aggregation. BMC Neurosci. 6, 1 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Zhang, Y.-Q. & Sarge, K. D. Celastrol inhibits polyglutamine aggregation and toxicity though induction of the heat shock response. J. Mol. Med. 85, 1421–1428 (2007).

    Article  CAS  PubMed  Google Scholar 

  105. Cleren, C., Calingasan, N. Y., Chen, J. & Beal, M. F. Celastrol protects against MPTP- and 3-nitropropionic acid-induced neurotoxicity. J. Neurochem. 94, 995–1004 (2005).

    Article  CAS  PubMed  Google Scholar 

  106. Faust, K. et al. Neuroprotective effects of compounds with antioxidant and anti-inflammatory properties in a Drosophila model of Parkinson's disease. BMC Neurosci. 10, 109 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Hansen, J. & Bross, P. A cellular viability assay to monitor drug toxicity. Methods Mol. Biol. 648, 303–311 (2010).

    Article  CAS  PubMed  Google Scholar 

  108. Kalmar, B. & Greensmith, L. Activation of the heat shock response in a primary cellular model of motoneuron neurodegeneration — evidence for neuroprotective and neurotoxic effects. Cell. Mol. Biol. Lett. 14, 319–335 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Wang, S., Liu, K., Wang, X., He, Q. & Chen, X. Toxic effects of celastrol on embryonic development of zebrafish (Danio rerio). Drug Chem. Toxicol. 34, 61–65 (2011).

    Article  CAS  PubMed  Google Scholar 

  110. Ohtsuka, K., Kawashima, D., Gu, Y. & Saito, K. Inducers and co-inducers of molecular chaperones. Int. J. Hyperthermia 21, 703–711 (2005).

    Article  CAS  PubMed  Google Scholar 

  111. Hirakawa, T., Rokutan, K., Nikawa, T. & Kishi, K. Geranylgeranylacetone induces heat shock proteins in cultured guinea pig gastric mucosal cells and rat gastric mucosa. Gastroenterology 111, 345–357 (1996).

    Article  CAS  PubMed  Google Scholar 

  112. Katsuno, M. et al. Pharmacological induction of heat-shock proteins alleviates polyglutamine-mediated motor neuron disease. Proc. Natl Acad. Sci. USA 102, 16801–16806 (2005). This work demonstrated that pharmacological activation of HSF1 via geranylgeranylacetone promotes the activation of chaperone protein expression and ameliorates cytotoxicity in a mouse model of spinal and bulbar muscular atrophy.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Otaka, M. et al. The induction mechanism of the molecular chaperone HSP70 in the gastric mucosa by geranylgeranylacetone (HSP-inducer). Biochem. Biophys. Res. Commun. 353, 399–404 (2007).

    Article  CAS  PubMed  Google Scholar 

  114. Patury, S., Miyata, Y. & Gestwicki, J. E. Pharmacological targeting of the Hsp70 chaperone. Curr. Top. Med. Chem. 9, 1337–1351 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Hirota, K. et al. Geranylgeranylacetone enhances expression of thioredoxin and suppresses ethanol-induced cytotoxicity in cultured hepatocytes. Biochem. Biophys. Res. Commun. 275, 825–830 (2000).

    Article  CAS  PubMed  Google Scholar 

  116. Okada, S. et al. Geranylgeranylacetone induces apoptosis in HL-60 cells. Cell Struct. Funct. 24, 161–168 (1999).

    Article  CAS  PubMed  Google Scholar 

  117. Endo, S. et al. Geranylgeranylacetone, an inducer of the 70-kDa heat shock protein (HSP70), elicits unfolded protein response and coordinates cellular fate independently of HSP70. Mol. Pharmacol. 72, 1337–1348 (2007).

    Article  CAS  PubMed  Google Scholar 

  118. Tam, S., Geller, R., Spiess, C. & Frydman, J. The chaperonin TRiC controls polyglutamine aggregation and toxicity through subunit-specific interactions. Nature Cell Biol. 8, 1155–1162 (2006). This study shows that the TRIC cytosolic chaperone complex binds to the pathogenic huntingtin protein and reduces huntingtin-mediated cytotoxicity.

    Article  CAS  PubMed  Google Scholar 

  119. Tam, S. et al. The chaperonin TRiC blocks a huntingtin sequence element that promotes the conformational switch to aggregation. Nature Struct. Mol. Biol. 16, 1279–1285 (2009).

    Article  CAS  Google Scholar 

  120. Hargitai, J. et al. Bimoclomol, a heat shock protein co-inducer, acts by the prolonged activation of heat shock factor-1. Biochem. Biophys. Res. Commun. 307, 689–695 (2003).

    Article  CAS  PubMed  Google Scholar 

  121. Vígh, L. et al. Bimoclomol: a nontoxic, hydroxylamine derivative with stress protein-inducing activity and cytoprotective effects. Nature Med. 3, 1150–1154 (1997).

    Article  PubMed  Google Scholar 

  122. Török, Z. et al. Heat shock protein coinducers with no effect on protein denaturation specifically modulate the membrane lipid phase. Proc. Natl Acad. Sci. USA 100, 3131–3136 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Nánási, P. P. & Jednákovits, A. Multilateral in vivo and in vitro protective effects of the novel heat shock protein coinducer, bimoclomol: results of preclinical studies. Cardiovasc. Drug Rev. 19, 133–151 (2001).

    Article  PubMed  Google Scholar 

  124. Kalmar, B. et al. Late stage treatment with arimoclomol delays disease progression and prevents protein aggregation in the SOD1G93A mouse model of ALS. J. Neurochem. 107, 339–350 (2008).

    Article  CAS  PubMed  Google Scholar 

  125. Kieran, D. et al. Treatment with arimoclomol, a coinducer of heat shock proteins, delays disease progression in ALS mice. Nature Med. 10, 402–405 (2004). This study demonstrates that pharmacological activation of HSF1 via arimoclomol ameliorates pathogenic phenotypes and extends lifespan in a mouse model of ALS.

    Article  CAS  PubMed  Google Scholar 

  126. Lanka, V., Wieland, S., Barber, J. & Cudkowicz, M. Arimoclomol: a potential therapy under development for ALS. Expert Opin. Investig. Drugs 18, 1907–1918 (2009).

    Article  CAS  PubMed  Google Scholar 

  127. Liu, A. Y. C. et al. Neuroprotective drug riluzole amplifies the heat shock factor 1 (HSF1)- and glutamate transporter 1 (GLT1)-dependent cytoprotective mechanisms for neuronal survival. J. Biol. Chem. 286, 2785–2794 (2011).

    Article  CAS  PubMed  Google Scholar 

  128. Jurivich, D. A., Sistonen, L., Kroes, R. A. & Morimoto, R. I. Effect of sodium salicylate on the human heat shock response. Science 255, 1243–1245 (1992).

    Article  CAS  PubMed  Google Scholar 

  129. Lee, B. S., Chen, J., Angelidis, C., Jurivich, D. A. & Morimoto, R. I. Pharmacological modulation of heat shock factor 1 by antiinflammatory drugs results in protection against stress-induced cellular damage. Proc. Natl Acad. Sci. USA 92, 7207–7211 (1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Winegarden, N. A., Wong, K. S., Sopta, M. & Westwood, J. T. Sodium salicylate decreases intracellular ATP, induces both heat shock factor binding and chromosomal puffing, but does not induce hsp 70 gene transcription in Drosophila. J. Biol. Chem. 271, 26971–26980 (1996).

    Article  CAS  PubMed  Google Scholar 

  131. Housby, J. N. et al. Non-steroidal anti-inflammatory drugs inhibit the expression of cytokines and induce HSP70 in human monocytes. Cytokine 11, 347–358 (1999).

    Article  CAS  PubMed  Google Scholar 

  132. Palayoor, S. T., Youmell, M. Y., Calderwood, S. K., Coleman, C. N. & Price, B. D. Constitutive activation of IκB kinase α and NF-κB in prostate cancer cells is inhibited by ibuprofen. Oncogene 18, 7389–7394 (1999).

    Article  CAS  PubMed  Google Scholar 

  133. Stevenson, M. A., Zhao, M. J., Asea, A., Coleman, C. N. & Calderwood, S. K. Salicylic acid and aspirin inhibit the activity of RSK2 kinase and repress RSK2-dependent transcription of cyclic AMP response element binding protein- and NF-κ B-responsive genes. J. Immunol. 163, 5608–5616 (1999).

    CAS  PubMed  Google Scholar 

  134. Ishihara, K., Yamagishi, N. & Hatayama, T. Suppression of heat- and polyglutamine-induced cytotoxicity by nonsteroidal anti-inflammatory drugs. Eur. J. Biochem. 271, 4552–4558 (2004).

    Article  CAS  PubMed  Google Scholar 

  135. Ianaro, A. et al. Anti-inflammatory activity of 15-deoxy-δ12,14-PGJ2 and 2-cyclopenten-1-one: role of the heat shock response. Mol. Pharmacol. 64, 85–93 (2003).

    Article  CAS  PubMed  Google Scholar 

  136. Rossi, A., Elia, G. & Santoro, M. G. 2-cyclopenten-1-one, a new inducer of heat shock protein 70 with antiviral activity. J. Biol. Chem. 271, 32192–32196 (1996).

    Article  CAS  PubMed  Google Scholar 

  137. Zhou, Y. et al. Chloro-oxime derivatives as novel small molecule chaperone amplifiers. Bioorg. Med. Chem. Lett. 19, 3128–3135 (2009).

    Article  CAS  PubMed  Google Scholar 

  138. Zhou, Y. et al. Pyrimido[5,4-e][1,2,4]triazine-5,7(1H,6H)-dione derivatives as novel small molecule chaperone amplifiers. Bioorg. Med. Chem. Lett. 19, 4303–4307 (2009).

    Article  CAS  PubMed  Google Scholar 

  139. Zhang, B. et al. Identification of small-molecule HSF1 amplifiers by high content screening in protection of cells from stress induced injury. Biochem. Biophys. Res. Commun. 390, 925–930 (2009).

    Article  CAS  PubMed  Google Scholar 

  140. Hayashida, N. et al. Heat shock factor 1 ameliorates proteotoxicity in cooperation with the transcription factor NFAT. EMBO J. 29, 3459–3469 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Batista-Nascimento, L., Neef, D. W., Liu, P. C. C., Rodrigues-Pousada, C. & Thiele, D. J. Deciphering human heat shock transcription factor 1 regulation via post-translational modification in yeast. PLoS ONE 6, e15976 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Rimoldi, M., Servadio, A. & Zimarino, V. Analysis of heat shock transcription factor for suppression of polyglutamine toxicity. Brain Res. Bull. 56, 353–362 (2001). This study shows that constitutively active HSF1, via loss of repressive phosphorylation events, prevents protein aggregation in cell culture models of polyglutamine disease.

    Article  CAS  PubMed  Google Scholar 

  143. Banerjee Mustafi, S., Chakraborty, P. K. & Raha, S. Modulation of Akt and ERK1/2 pathways by resveratrol in chronic myelogenous leukemia (CML) cells results in the downregulation of Hsp70. PLoS ONE 5, e8719 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Khaleque, M. A. et al. Induction of heat shock proteins by heregulin β1 leads to protection from apoptosis and anchorage-independent growth. Oncogene 24, 6564–6573 (2005).

    Article  CAS  PubMed  Google Scholar 

  145. Xavier, I. et al. Glycogen synthase kinase 3β negatively regulates both DNA-binding and transcriptional activities of heat shock factor 1. J. Biol. Chem. 275, 29147–29152 (2000).

    Article  CAS  PubMed  Google Scholar 

  146. Anckar, J. et al. Inhibition of DNA binding by differential sumoylation of heat shock factors. Mol. Cell Biol. 26, 955–964 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Bernier-Villamor, V., Sampson, D. A., Matunis, M. J. & Lima, C. D. Structural basis for E2-mediated SUMO conjugation revealed by a complex between ubiquitin-conjugating enzyme Ubc9 and RanGAP1. Cell 108, 345–356 (2002).

    Article  CAS  PubMed  Google Scholar 

  148. Brunet Simioni, M. et al. Heat shock protein 27 is involved in SUMO-2/3 modification of heat shock factor 1 and thereby modulates the transcription factor activity. Oncogene 28, 3332–3344 (2009).

    Article  CAS  PubMed  Google Scholar 

  149. Fukuda, I. et al. Ginkgolic acid inhibits protein SUMOylation by blocking formation of the E1-SUMO intermediate. Chem. Biol. 16, 133–140 (2009).

    Article  CAS  PubMed  Google Scholar 

  150. Parker, J. A. et al. Resveratrol rescues mutant polyglutamine cytotoxicity in nematode and mammalian neurons. Nature Genet. 37, 349–350 (2005). This study shows that activation of SIR-2 (the C. elegans homolog of SIRT1) via resveratrol rescues neuronal dysfunction in C. elegans and mouse models of polyQ disease.

    Article  CAS  PubMed  Google Scholar 

  151. Kim, D. et al. SIRT1 deacetylase protects against neurodegeneration in models for Alzheimer's disease and amyotrophic lateral sclerosis. EMBO J. 26, 3169–3179 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  152. Ladiwala, A. R. A. et al. Resveratrol selectively remodels soluble oligomers and fibrils of amyloid Aβ into off-pathway conformers. J. Biol. Chem. 285, 24228–24237 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Marambaud, P., Zhao, H. & Davies, P. Resveratrol promotes clearance of Alzheimer's disease amyloid-β peptides. J. Biol. Chem. 280, 37377–37382 (2005).

    Article  CAS  PubMed  Google Scholar 

  154. Lu, K.-T. et al. Neuroprotective effects of resveratrol on MPTP-induced neuron loss mediated by free radical scavenging. J. Agric. Food Chem. 56, 6910–6913 (2008).

    Article  CAS  PubMed  Google Scholar 

  155. Zhang, F. et al. Resveratrol protects dopamine neurons against lipopolysaccharide-induced neurotoxicity through its anti-inflammatory actions. Mol. Pharmacol. 78, 466–477 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Salamanca, H. H., Fuda, N., Shi, H. & Lis, J. T. An RNA aptamer perturbs heat shock transcription factor activity in Drosophila melanogaster. Nucleic Acids Res. 39, 6729–6740 (2011). This work describes an RNA aptamer that interacts with the DNA binding domain of HSF1 and inhibits its binding to promoter heat shock elements.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. Liu, P. C. & Thiele, D. J. Modulation of human heat shock factor trimerization by the linker domain. J. Biol. Chem. 274, 17219–17225 (1999).

    Article  CAS  PubMed  Google Scholar 

  158. Finkbeiner, S. Bridging the Valley of Death of therapeutics for neurodegeneration. Nature Med. 16, 1227–1232 (2010).

    Article  CAS  PubMed  Google Scholar 

  159. Aguzzi, A. & O'Connor, T. Protein aggregation diseases: pathogenicity and therapeutic perspectives. Nature Rev. Drug Discov. 9, 237–248 (2010).

    Article  CAS  Google Scholar 

  160. Hampel, H. et al. Biomarkers for Alzheimer's disease: academic, industry and regulatory perspectives. Nature Rev. Drug Discov. 9, 560–574 (2010).

    Article  CAS  Google Scholar 

  161. Schapira, A. H. V. Challenges to the development of disease-modifying therapies in Parkinson's disease. Eur. J. Neurol. 18 (Suppl. 1), 16–21 (2011).

    Article  PubMed  Google Scholar 

  162. Murray, A. N., Solomon, J. P., Wang, Y. J., Balch, W. E. & Kelly, J. W. Discovery and characterization of a mammalian amyloid disaggregation activity. Protein Sci. 19, 836–846 (2010). This work describes the discovery of a mammalian disaggregase with the ability to disaggregate β-amyloid aggregates.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Opar, A. Hope builds for earlier detection of Alzheimer's disease. Nature Rev. Drug Discov. 9, 579–581 (2010).

    Article  CAS  Google Scholar 

  164. Nielsen, P. A., Andersson, O., Hansen, S. H., Simonsen, K. B. & Andersson, G. Models for predicting blood–brain barrier permeation. Drug Discov. Today 16, 472–475 (2011).

    Article  CAS  PubMed  Google Scholar 

  165. Pardridge, W. M. Alzheimer's disease drug development and the problem of the blood-brain barrier. Alzheimers Dement. 5, 427–432 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  166. Lipinski, C. A., Lombardo, F., Dominy, B. W. & Feeney, P. J. Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings. Adv. Drug Deliv. Rev. 46, 3–26 (2001).

    Article  CAS  PubMed  Google Scholar 

  167. Cudkowicz, M. E. et al. Arimoclomol at dosages up to 300 mg/day is well tolerated and safe in amyotrophic lateral sclerosis. Muscle Nerve 38, 837–844 (2008).

    Article  CAS  PubMed  Google Scholar 

  168. Milane, A. et al. Brain and plasma riluzole pharmacokinetics: effect of minocycline combination. J. Pharm. Pharm. Sci. 12, 209–217 (2009).

    Article  CAS  PubMed  Google Scholar 

  169. Kumar, S. et al. Extracellular phosphorylation of the amyloid β-peptide promotes formation of toxic aggregates during the pathogenesis of Alzheimer's disease. EMBO J. 30, 2255–2265 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  170. Williams, T. L. & Serpell, L. C. Membrane and surface interactions of the Alzheimer's Aβ peptide: insights into the mechanism of cytotoxicity. FEBS J. 278, 3905–3917 (2011).

    Article  CAS  PubMed  Google Scholar 

  171. Cohen, F. E. & Kelly, J. W. Therapeutic approaches to protein-misfolding diseases. Nature 426, 905–909 (2003).

    Article  CAS  PubMed  Google Scholar 

  172. Labbadia, J. et al. Altered chromatin architecture underlies progressive impairment of the heat shock response in mouse models of Huntington disease. J. Clin. Invest. 121, 3306–3319 (2011). This study demonstrates that activation of HSF1-dependent chaperone protein expression via an HSP90 inhibitor transiently ameliorates disease phenotypes in a mouse model of polyQ-based disease as a result of decreased promoter acetylation.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  173. Balch, W. E., Morimoto, R. I., Dillin, A. & Kelly, J. W. Adapting proteostasis for disease intervention. Science 319, 916–919 (2008).

    Article  CAS  PubMed  Google Scholar 

  174. Biamonte, M. A. et al. Heat shock protein 90: inhibitors in clinical trials. J. Med. Chem. 53, 3–17 (2010).

    Article  CAS  PubMed  Google Scholar 

  175. 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).

    Article  CAS  PubMed  Google Scholar 

  176. Nowakowski, G. S. et al. A phase I trial of twice-weekly 17-allylamino-demethoxy-geldanamycin in patients with advanced cancer. Clin. Cancer Res. 12, 6087–6093 (2006).

    Article  CAS  PubMed  Google Scholar 

  177. Brandt, G. E. L., Schmidt, M. D., Prisinzano, T. E. & Blagg, B. S. J. Gedunin, a novel Hsp90 inhibitor: semisynthesis of derivatives and preliminary structure–activity relationships. J. Med. Chem. 51, 6495–6502 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Kikuchi, T. et al. Cytotoxic and apoptosis-inducing activities of limonoids from the seeds of Azadirachta indica (neem). J. Nat. Prod. 74, 866–870 (2011).

    Article  CAS  PubMed  Google Scholar 

  179. Traynor, B. J. et al. Neuroprotective agents for clinical trials in ALS: a systematic assessment. Neurology 67, 20–27 (2006).

    Article  CAS  PubMed  Google Scholar 

  180. Bensimon, G. et al. Riluzole treatment, survival and diagnostic criteria in Parkinson plus disorders: the NNIPPS study. Brain 132, 156–171 (2009).

    Article  PubMed  Google Scholar 

  181. Nanke, Y. et al. Geranylgeranylacetone, a non-toxic inducer of heat shock protein, induces cell death in fibroblast-like synoviocytes from patients with rheumatoid arthritis. Mod. Rheumatol. 19, 379–383 (2009).

    Article  CAS  PubMed  Google Scholar 

  182. Nishida, T. et al. Geranylgeranylacetone protects against acetaminophen-induced hepatotoxicity by inducing heat shock protein 70. Toxicology 219, 187–196 (2006).

    Article  CAS  PubMed  Google Scholar 

  183. Shirakabe, H. et al. Clinical evaluation of teprenone, a mucosal protective agent, in the treatment of patients with gastric ulcers: a nationwide, multicenter clinical study. Clin. Ther. 17, 924–935 (1995).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank T. Nevitt for her critical comments on the manuscript. This work was supported in part by the US National Institutes of Health (NIH) National Research Service Award Postdoctoral Fellowship GM076954 (to D.W.N.) and the NIH grant R01-GM059911 (to D.J.T.). A.M.J. is a trainee of the Duke University Pharmacological Sciences Training Program.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Dennis J. Thiele.

Ethics declarations

Competing interests

Dennis J. Thiele and Daniel W. Neef are inventors on patent applications describing small-molecule activators of human heat shock transcription factor 1. Dennis J. Thiele is a co-founder and a shareholder of Chaperone Therapeutic, Inc.

Related links

Related links

FURTHER INFORMATION

Dennis J. Thiele's homepage

ClinicalTrials.gov website

Glossary

Dyskinesia

A condition in which voluntary movement is lost and an increase in chorea-like involuntary movement is observed.

Leucine zipper

A structural motif that stabilizes inter- or intramolecular protein–protein interactions via hydrophobic and charged interactions across coiled-coils and is commonly found in oligomerization domains.

Sumoylation

A post-translational modification that is indicated by the addition of a small ubiquitin-like modifier (SUMO) moiety that can affect protein stability, localization and activity.

Residence time

The duration of time that heat shock transcription factor 1 is bound to heat shock elements in the promoter region of target genes such as those encoding chaperone proteins.

Chaperone-mediated autophagy

A process by which cytosolic proteins are selectively degraded through interaction with heat shock cognate protein 70, which facilitates direct translocation into lysosomes for proteolysis.

Unfolded protein response

A conserved physiological response involving endoplasmic reticulum (ER)-initiated signal-transduction events, induced by accumulation of unfolded proteins in the lumen of the ER.

SOD1G93A mice

Transgenic mice expressing the G93A mutant form of human superoxide dismutase 1 (SOD1) that causes familial amyotrophic lateral sclerosis (ALS), which are commonly used as a model for ALS.

RNA aptamer

A specifically designed oligonucleotide with a secondary structure that elicits high affinity for a desired target.

p53R172H mouse model

A mouse model expressing a mutated form of the tumour suppressor protein p53, R172H, which results in increased oncogenesis.

R6/2 mouse model

A widely used transgenic mouse model — expressing exon 1 of the human huntingtin gene containing 150 CAG repeats — that rapidly develops Huntington's disease-like symptoms.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Neef, D., Jaeger, A. & Thiele, D. Heat shock transcription factor 1 as a therapeutic target in neurodegenerative diseases. Nat Rev Drug Discov 10, 930–944 (2011). https://doi.org/10.1038/nrd3453

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrd3453

This article is cited by

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research