Europe PMC

This website requires cookies, and the limited processing of your personal data in order to function. By using the site you are agreeing to this as outlined in our privacy notice and cookie policy.

Abstract 


No abstract provided.

Free full text 


Logo of nihpaAbout Author manuscriptsSubmit a manuscriptHHS Public Access; Author Manuscript; Accepted for publication in peer reviewed journal;
Gene. Author manuscript; available in PMC 2017 Nov 7.
Published in final edited form as:
PMCID: PMC5675009
NIHMSID: NIHMS909442
PMID: 26358502

HSP90AB1: helping the good and the bad

Abstract

HSP90AB1 (heat shock protein 90 kDA alpha, class B, member 1), also known as HSP90beta, is a member of the large family of HSPs which function as molecular chaperones. Chaperones, by binding to client proteins, support proper protein folding and maintain protein stability, especially after exposure to various kinds of cellular stress. Client proteins belong to various protein families including kinases, ubiquitin ligases and transcription factors. HSP90 proteins act as dimers and bind clients with the help of co-chaperones. The cochaperones influence many functions including client binding, ATPase activity or ATP binding of HSP90. HSPs are necessary for a large scale of cellular processes and therefore essential for cell survival. Since client proteins can be mutant proteins that would be degraded without the help of chaperones, HSPs also promote tumor formation and cancer cell proliferation. As such, they are also targets for new therapeutic approaches in cancer treatment. This review focuses on recent studies on HSP90AB1, if possible in comparison with its close homologue HSP90AA1.

Introduction

Heat shock proteins (HSP) are a large group of chaperones which are proteins that assist in protein folding, stabilize proteins and help to refold denatured proteins, processes that are dependent on hydrolysis of ATP (1). If proper folding is not possible, they also aid in protein degradation. The major groups of HSPs are shown in table 1. Chaperonins form a sub-class of HSPs and are characterized by a stacked double-ring structure forming barrels (Xu et al. 1997). Inside the barrel structures, they contain hydrophobic residues for client binding (Lindquist 1986). The prototypes of chaperonins are GroEL/GroES (large and small proteins of the GroE operon in E. coli, mutations of which affect the growth of lambda phage by interfering with assembly of its head protein E) in bacteria (Georgopoulos et al. 1973; Sternberg 1973b, a; Hendrix 1979; Yamamori & Yura 1980; Fayet et al. 1989) and Hsp60/Hsp10 proteins in eukaryotic cells (Johnson & Craig 1997).

Table 1

Chaperone families according to HNGC (Human Genome Organization Gene Nomenclature Committee, http://www.genenames.org/genefamilies/HSP) proposed by Kampinga et al. (Kampinga et al. 2009) and representative members of each family.

Chaperone family: short nameRepresentative membersReferences
HSP70HSPA1A (Hsp70–1), HSPA8 (HSC70/71, heat shock cognate 70/71 kDa)(Liu et al. 2012; Stricher et al. 2013)
DNAJ (HSP40)DNAJB1 (Hsp40), HSCB (Hsc20)(Ohtsuka & Hata 2000; Fan et al. 2003; Cyr & Ramos 2015)
HSPB (small heat shock proteins)HSPB1 (Hsp27)(Garrido et al. 2003; Acunzo et al. 2012)
HSPC (HSP90)HSP90AA1 (Hsp90alpha), HSP90AB1 (HSP90beta)(Pearl & Prodromou 2000; Taipale et al. 2010)
chaperoninsHSPD1 (GroEL, Hsp60), HSPE1 (GroES, Hsp10), TriC(Horwich et al. 2006; Krishna et al. 2007)

Originally, HSPs were described as proteins that were up-regulated after elevated temperatures (Lindquist 1986). Meanwhile it is recognized that HSPs are involved in the response to all kinds of stress reactions that disturb proper protein conformation such as reactions to chemicals like ethanol, arsenite, cadmium, zinc, copper, mercury, sulfhydryl reagents, calcium ionophores, steroid hormones, chelating agents, viruses and many more (Lindquist 1986). Of course, HSPs are strongly induced by DNA damage since this type of stress leads to mutations that often interfere with proper protein folding (Fornace et al. 1988). Since HSPs stabilize DNA binding proteins it is not surprising to detect genomic instability in HSP70 deficient mice (Hunt et al. 2004). Because also many undamaged proteins need assistance in folding, nearly all physiological processes require HSPs. Indeed, with the help of protein-protein interaction (PPI) studies, proteins could be identified as interaction partners that contribute to the following processes: transcription, mRNA splicing, translation, cell cycle control, DNA repair, apoptosis, intracellular transport, development, immune response, lipid and carbohydrate metabolism, cellular signaling, protein modification and many more (Gong et al. 2009; Tsaytler et al. 2009; Gano & Simon 2010; Echeverria et al. 2011; Hartson & Matts 2012; Taipale et al. 2014). HSPs are also involved in protein transport across membranes (e.g. of mitochondria or endoplasmic reticulum).

Several types of chaperones may act together, dependent on the type of protein or type of damage. One model suggests that client proteins (e.g. proteins in translation dependent on the stage of maturation) are first bound to Hsp70, then to chaperonins (Johnson & Craig 1997), then to more specialized proteins of the Hsp90 group (Hartl 1996; Johnson & Craig 1997). An alternative pathway, independent of HSP70 or HSP90, involves binding to CCT/TriC (chaperonin containing T-complex polypeptide/ TCP-1 ring complex) proteins. This pathway is used by filamentous proteins like actin and tubulin (Johnson & Craig 1997) but also other types of proteins. There are also multichaperone complexes like HSP70/HSP90. Complex formation in this case is mediated by the adapter protein and co-chaperone Hop (HSP organizing protein) which binds to the peptide sequence EEVD at the C-terminus of both proteins (Chen & Smith 1998; Scheufler et al. 2000; Brinker et al. 2002). Often, the client proteins remain bound to a HSP, but they may be released once they are stable on their own. Co-chaperones mediate substrate specificity, regulate activity of client proteins or recruit chaperones to specific locations in order to perform special functions like their roles in clathrin uncoating, synaptic vesicle fusion, or regulating cytoskeleton functions (Young et al. 2003b).

There are a lot of excellent reviews on the family of HSP proteins in general (Hartl 1996; Wegele et al. 2004; Calderwood et al. 2006), but also on special subfamilies like HSP70 (Bukau & Horwich 1998) or HSP90 (Csermely et al. 1998; Pearl & Prodromou 2000; Taipale et al. 2010; Erlejman et al. 2014a), their role in signaling protein movement (Pratt et al. 2004) and on co-chaperones (Pratt et al. 2004; Davies & Sanchez 2005; Cioffi et al. 2011; Sivils et al. 2011; Storer et al. 2011).

This review is focused on HSP90AB1, primarily in humans. Because of overlapping functions and lack of discrimination in the past, the term HSP90 is used when it is not clear which member of the HSP90 protein family was studied.

Classification

Based on several suggestions (Chen et al. 2005; Chen et al. 2006a), lastly by Kampinga et al. (Kampinga et al. 2009), the human genome organization (HUGO) gene nomenclature committee (HGNC, http://www.genenames.org/genefamilies/HSP) recognizes 5 human families of HSPs (Table 1). At the moment, all these families encompass 97 genes.

With 5 genes listed, the HSP90/HSPC family is the smallest among all human HSPs. One of them is a pseudogene (HSP90AA3P), therefore four real HSP’s of this family remain (Table 2) which have characteristic localizations in the cell. However, these localizations are only a general rule. Under certain circumstances, these proteins may be also found in other compartments of the cell and even extracellularly (Suzuki & Kulkarni 2010).

Table 2

HSP90/HSPC family according to HNGC (HUGO gene nomenclature committee, http://www.genenames.org/genefamilies/HSP) proposed by Kampinga et al. (Kampinga et al. 2009).

AbbreviationFull nameOther common names: abbreviationsOther common names: full namesLocalization
HSP90AA1heat shock protein 90 kDa alpha, class A, member 1HSP90alpha, HSP86heat shock protein 90 kDa alpha, heat shock protein 86 kDacytoplasm
HSP90AB1heat shock protein 90 kDa alpha, class B, member 1HSP90beta, Hsp84heat shock protein 90 kDa beta, heat shock protein 84 kDacytoplasm
HSP90B1heat shock protein 90 kDA beta, member 1Grp94/96glucose-regulated protein 94/96 kDaendoplasmic reticulum
Trap1TNF receptor-associated protein 1HSP90Lheat shock protein 90 likemitochondria

Cloning of HSP90AB1

Human HSP90AB1 (Rebbe et al. 1987) was cloned based on homology to HSP90AA1 which was the first HSP90 to be purified, as reported in 1982 (Welch & Feramisco 1982). Both proteins share 60 % overall homology and several regions of 50 amino acids (aa) or more share greater than 90 % homology (Rebbe et al. 1987). Mouse HSP90AB1 was cloned using the corresponding Drosophila cDNA as a hybridization probe (Moore et al. 1987; Hoffmann & Hovemann 1988). It codes for a protein containing 724 aa. The majority of the HSP90A proteins occur as homodimers (Minami et al. 1991). Human HSP90AB1 has 724 aa versus 732 aa in HSP90AA1. With the N-terminal methionins removed, which is usually the case, these numbers are 723 and 731 respectively (Lees-Miller & Anderson 1989b). In HSP90AB1, the amino acids ESEDK is removed between the phosphorylation sites (Lees-Miller & Anderson 1989b). In addition, the amino acids TQTQDQPME at the N-terminal end of HSP90AA1 are replaced by VHHG in HSP90AB1 (Lees-Miller & Anderson 1989b).

HSP90AB1 has the unique signature sequence LKID (residues 71–74) that is not present in other HSPs (Chen et al. 2006a).

Protein domains of HSP90

HSP90 is composed of five domains (Obermann et al. 1998; Chen et al. 2006a):

  • -

    N-terminal domain (NTD)

  • -

    Linker region/Charged domain 1 (CD1)

  • -

    middle domain (MD)

  • -

    charged domain 2 (CD2)

  • -

    C-terminal domain (CTD)

These main domains as well as their amino acid limits are shown in figure 1 for HSP90AB1 and the 732aa translation product of HSP90AA1. Chen et al. proposed further subdomains and shifts of domain borders based on sequence conservation or variability after analysis of a large number of HSP90 proteins of different species (Chen et al. 2006a).

An external file that holds a picture, illustration, etc.
Object name is nihms909442f1.jpg

Domain structure of human HSP90AA1 (732aa translation product, Ensembl transcript ID ENST00000216281) and HSP90AB1 (ENSEMBL transcript ID ENST00000371554). Amino acid borders of the main domains are given above each molecule. Co-chaperones that bind to the respective regions are depicted below. NTD, N-terminal domain; CTD, C-terminal domain; MD, middle domain.

The NTD contains the ATP-binding site which is essential for HSP function (Prodromou et al. 1997a; Stebbins et al. 1997; Obermann et al. 1998) and binds several co-chaperones (see below). The ATP binding site shares structural homology with other ATPases in the GHKL (gyrase, HSP90, histidine kinase and MutL) superfamilly (Pearl & Prodromou 2006).

The CD1 influences binding of other proteins to the NTD (Scheibel et al. 1999). Deletion of the CD1 domain is lethal in yeast (Hainzl et al. 2009). In addition, the sensitivity for Aha1-mediated ATPase acceleration declines, binding of p23/PTGES3 is lost and the ability to facilitate client activation is reduced (Hainzl et al. 2009). This flexible domain facilitates intermolecular rearrangements on the milliseconds timescale between a state in which the NTD is docked to the middle domain and a state in which the N-domain is more flexible (Jahn et al. 2014). The MD is the main region for client binding, but also the co-chaperone Aha1/AHSA1 (see below) binds to this domain (Table 3).

Table 3

Important co-chaperones, modified after Taipale et al. (Taipale et al. 2010). CS, CHORD and Sgt1 domain; CT, C-terminal; CTD, C-terminal domain; MD, middle domain; MEEVD, motif named after its amino acids; NTD, N-terminal domain; TPR, tetratricopeptide peptide domain; +, acceleration; −, inhibition; 0, no influence.

AbbreviationFull Name/other names/gene nameBinding domain in HSP90Binding domain in cochaperoneATPase activityReferences
Cdc37cell division cycle 37NTDCTD(Stepanova et al. 1996; Lamphere et al. 1997)
p23Protein with ~23 kDa; PTGES3, prostaglandin E synthase 3; cPGES (cytoplasmic PGES)NTDCS(Oxelmark et al. 2006; Lovgren et al. 2007; Karagoz et al. 2011)
SGT1Small glutamine-rich TPR-containing protein 1NTDCS(Yin et al. 2006)
AHA1AHSA1, activator of heat shock 90kDa protein ATPase homolog 1MDNTD+(Wang et al. 2006; Echeverria et al. 2011)
HOPHsp-organizing protein; Sti1/STIP1, stress induced phosphoprotein 1CTD
MEEVD
TPR(Young et al. 1998; Carrello et al. 1999; Scheufler et al. 2000; King et al. 2001; Brinker et al. 2002; Millson et al. 2008)
CHIPC-terminus of HSC70-interacting protein; STUB1, STIP1 homology and U-box containing protein 1CTD
MEEVD
TPR(Ballinger et al. 1999; Shang et al. 2014)
FKBPLFK506 binding protein-like; WISp39, WAF1/CIP1 stabilizing protein 39 kDaCTD
MEEVD
TPR(McKeen et al. 2008; Sunnotel et al. 2010; Valentine et al. 2011; Yakkundi et al. 2013; Donley et al. 2014)
FKBP52FKBP52/54 (FK506-binding protein, 52/54 kDa; FKBP4, F506 binding protein 4CTD
MEEVD
TPR(Miyata et al. 1997; Carrello et al. 1999; Pirkl & Buchner 2001; Scammell et al. 2003; Davies & Sanchez 2005; Cioffi et al. 2011; Storer et al. 2011)
FKBP51FK506 binding protein 51 kDa; FKBP5, FK506 binding protein 5CTD
MEEVD
TPR(Pirkl & Buchner 2001; Hubler et al. 2003; Cioffi et al. 2011; Storer et al. 2011)
CYP40cyclophilin 40CTD
MEEVD
TPR0(Duina et al. 1996a; Pirkl & Buchner 2001; Carrello et al. 2004; Mok et al. 2006; Park et al. 2011)
PP5protein phosphatase 5CTD
MEEVD
TPR(Silverstein et al. 1997; Hahn 2005; Hinds & Sanchez 2008)
TOM70translocase of outer membrane 70 kDaCTD
MEEVD
TPR(Young et al. 2003a)
TPR2tetratricopeptide repeat containing protein 2CTD
MEEVD
TPR(Xiang et al. 2001; Kubo et al. 2013)
Cpr6 and Cpr76th and 7th cyclophilin homolog, interacting with the yeast global transcriptional regulator Rpd3CTD
MEEVD
TPR+(Duina et al. 1996a; Tenge et al. 2015)

The CD2 is only characterized by DNA analysis but not functionally defined (Chen et al. 2006a).

The CTD is the dimerization domain of HSP90. In addition, five C-terminal residues (Met-Glu-Glu-Val-Asp, MEEVD motif) make up a highly conserved tetratricopeptide repeat (TPR) domain binding site which mediates interaction with many co-chaperones (Bose et al. 1996; Owens-Grillo et al. 1996; Young et al. 1998).

Co-chaperones

As already mentioned, many HSP functions are influenced by co-chaperones. They have four major functions (Taipale et al. 2010):

  • -

    They coordinate the interplay between HSP90 and other chaperone systems, such as HSP70

  • -

    stimulate or inhibit ATPase activity of HSP90

  • -

    recruit specific classes of clients and other co-chaperones and

  • -

    have an enzymatic activity.

Some important co-chaperones are listed in table 3 and shortly characterized in the text below. The co-chaperone binding regions on HSP90 are depicted in figure 1.

The co-chaperone Cdc37 (cell division cycle 37) acts as an active scaffold molecule that allows binding of HSP90 with its C-terminal region and client proteins like kinases with its N-terminal domain. Cdc37 arrests the ATPase cycle of HSP90 by inserting an arginine side chain into the ATP binding pocket, thereby facilitating protein loading (Roe et al. 2004).The activation of several client protein kinases is regulated by chaperone-targeted dephosphorylation of Cdc37 (Vaughan et al. 2008).

Cdc37 interacts with Cdk4 (cyclin dependent kinase 4) and Cdk6, but not with Cdc2 (cell division cycle 2), Cdk2, Cdk3 or Cdk5 (Lamphere et al. 1997; Vaughan et al. 2006). The Cdc37/HSP90 complex stabilizes Cdk4 (Stepanova et al. 1996). In addition, Cdc37 facilitates complex formation between Cdk4 and Cyclin D1 in vitro (Lamphere et al. 1997). Since p16 competes with Cdc37 for binding to Cdk4 it is likely that part of its inhibitory effect is the destabilization of Cdk4 by removal of Cdc37 (Lamphere et al. 1997). However, it has been found that Cdk9 (formerly called PITALRE due to its characteristic amino acid motif) interacts both with Hsp70 and with a HSP90/Cdc37 complex during Tat activation of HIV-1 transcription (O’Keeffe et al. 2000).

The p23/PTGES3 (prostaglandin E synthase 3) protein known as co-chaperone for HSP90 proteins is also a cytoplasmic terminal PGE2 (prostaglandin E2) synthase (cPGES). P23 binds to the N-terminal domain of HSP90 (Karagoz et al. 2011). P23 deficient mice display a phenotype characteristic of mice with GR (glucocorticoid receptor) deficiency, suggesting that its main function is that of the co-chaperone and not its role in prostaglandin synthesis (Lovgren et al. 2007). Expression of the co-chaperone p23 enhances estrogen receptor (ER)-mediated transcription of genes dependent on the presence of estrogen response elements (ERE) (Oxelmark et al. 2006). P23 also has an interesting role in RNA interference (RNAi). RNAi is a mechanism to control gene expression by regulating mRNA levels. The specificity of this mechanism is determined by small double-stranded RNA molecules, the siRNA (small interfering RNA molecules). The guide strand of this siRNA, which is complementary to a section of the target mRNA, is incorporated into the RNA-induced silencing complex (RISC). The catalytic component of the RISC is the endoculease Argonaute (Ago) (Agrawal et al. 2003). It has been shown in vitro with plant proteins that HSP90 binds to Ago1 (Iki et al. 2010). RISC assembly is driven by the ATPase activity of HSP90 in a Hsc70 (heat shock cognate 70)/HSP90 complex (Iwasaki et al. 2010). The HSP90 co-chaperones p23 and FKBP52 interact with human Ago2 (hAgo2) and HSP90 before small RNA and RISC loading takes place (Pare et al. 2013). In addition, the co-chaperones Aha1 and Cdc37 are necessary for the mechanism of RNAi (Pare et al. 2013). HSP90 modifies Ago2 conformation in order to bind siRNA (Miyoshi et al. 2010).

SGT1 (Small glutamine-rich TPR-containing protein 1) interacts with both HSP90AA1 and HSP90AB1. Nuclear SGT1 is increased upon HSP90 inhibition with geldanamycin and is associated with apoptosis in HeLa cells (Yin et al. 2006).

AHA1 (activator of heat shock 90kDa protein ATPase homolog 1) interacts with both HSP90AB1 and HSP90AA1 (Echeverria et al. 2011). However, in a model of 293T cells, it hast been shown that it preferentially interacts with (the middle domain of) HSP90AA1 (Synoradzki & Bieganowski 2015). It is involved in a broad range of cellular processes such as regulation of signal transduction, RNA metabolism, cell development, regulation of ubiquitin-protein ligase actvity, glycolysis, induction of apoptosis and nucleo-cytoplasmic transport (Echeverria et al. 2011). AHA1 interacts with Importin-4 (IPO4) and karyopherin alpha 5 (Kpna5) whereas an interaction with exportin1 (Xpo1) could not be demonstrated by immunoprecipitation (Echeverria et al. 2011). AHA1 binds to HSP90 with its N-terminal domain, thereby stimulating ATP binding of HSP90 (Meyer et al. 2004). Silencing of p53 leads to elevated AHA1 levels. In addition, increased AHA1 levels, both by p53 silencing and by AHA1 overexpression, lead to increased HSP90 ATPase activity and enhanced cytochrome P450 levels both under basal conditions and after treatment with the carcinogen benzo[a]pyrene (Kochhar et al. 2014). Furthermore, after AHA1 and HSP90 activation, also Akt and GSK3beta are activated and Wnt target genes are expressed. Together with p53 rescue intervention, these experiments demonstrate a connection between the p53 and Wnt pathways (Okayama et al. 2014).

HOP (Hsp-organizing protein) mediates interaction between HSP70 and HSP90 proteins (Chen & Smith 1998). The TPR1 and TPR2A domains of Hop bind to the C-terminal (M)EEVD motifs of Hsp70 and HSP90 respectively (Scheufler et al. 2000). As explained below, this complex plays a role in stabilizing mutant p53 (King et al. 2001). Hop is also the first protein recruited to HSP90 in the chaperone cycle (see below).

CHIP/STUB1 (C-terminus of HSc70-interacting protein /STIP1 homology and U-box containing protein 1) is an E3 ubiquitin ligase that interacts and destabilizes fibroblast growth factor receptor 3 (FGFR3) (Laederich et al. 2011). FGFR3 stability is strongly dependent on HSP90 and Cdc37 (Laederich et al. 2011). Another client degraded by CHIP is PTK6 (protein tyrosine kinase 6), if not protected by HSP90 (Kang et al. 2012). PTK6 plays a role in tumor biology as described below. Mutations in KCNQ4 (potassium channel, voltage gated KQT-like subfamily Q, member 4) lead to a subtype of autosomal-dominant deafness. Mutant KCNQ4 is degraded by CHIP and protected by the HSP40-HSP70-HOP-HSP90(AB1 and AA1) pathway (Gao et al. 2013). Inducible HSP90AA1 and the constitutive HSP90AB1, had opposite effects on the cellular level of the KCNQ4 channel: HSP90AB1 over-expression partially restored KCNQ4 surface expression in cells mimicking heterozygous conditions in deafness patients (Gao et al. 2013). In TGF-beta signaling HSP70 and HSP90 differentially regulate complex formation of SMAD3 and CHIP. Increased levels of HSP70 and a decrease of HSP90 increased CHIP-mediated degradation of SMAD3 whereas the opposite changes in HSP levels stabilized SMAD3. HSP90AB1 is somewhat more effective in stabilizing SMAD3 than HSP90AA1. Consequently, HSP90AB1 overexpression leads to a higher activity of the TGFbeta pathway as measured with a luciferase reporter assay (Shang et al. 2014). In contrast, extracellular HSP90AB1 inhibits TGFbeta1 by binding to LAP (latency-associated peptide), thereby preventing the release of active TGFbeta1 from its complex with LAP and decreasing cell proliferation in a cell culture model with MG63 osteosarcoma cells (Suzuki & Kulkarni 2010).

FKBP (FK506 binding proteins) are members of the immunophilin family of proteins which are proteins that bind to immunosuppressants like FK506 or rapamycin. They encode proteins with peptidyl-prolyl cis-trans isomerase (PPIase) activity. FKBPL (FKBP-like) interacts with the HSP90-steroid hormone receptor complex (McKeen et al. 2008). Mutations of FKBPL are associated with azoospermia most likely due to a decrease in androgen receptor (AR) DNA binding caused by the destabilization of this protein (Sunnotel et al. 2010).

FKBPL influences the cytoskeleton through a PPIase domain that interacts with the dynamin motor protein dynamitin (McKeen et al. 2008). In addition, FKBPL binds to CD44 and inhibits cell migration, as shown in endothelial cells and breast cancer cells (Yakkundi et al. 2013). A high level of FKBPL and the FKBPL-interacting protein RBCK1 (RanBP-type and C3HC4-type zinc finger-containing protein 1) correlates with increased patient survival in breast cancer (Donley et al. 2014), suggesting decreased migration and consequential inhibition of metastasis as an underlying mechanism. The role of FKBP52 in RNAi has been mentioned above (Pare et al. 2013). Like FKBPL, FKBP52 interacts with steroid hormone receptors. It does not only stabilize them, but is also necessary for their nuclear translocation (Storer et al. 2011). The interaction between FKBP52 (FKBP4) and HSP90 is regulated by phosphorylation of FKBP52 by casein kinase II (CKII) (Miyata et al. 1997). GR and mineralocorticoid receptor (MR) hormone binding activity and their activity upon gene transcription is higher in association with FKBP52 compared to PP5, whereas FKBP51 has an inhibitory effect (Davies et al. 2005; Gallo et al. 2007). As already mentioned for FKBPL, loss of FKBP52 in mice also leads to azoospermia (Storer et al. 2011). FKBP52 influences microtubule dynamics by isomerization of the microtubule binding protein Tau to its cis configuration, which enhances its dephosphorylation by PP5 (protein phosphatase 5) (Cioffi et al. 2011). Ret51 (a variant of the Ret transmembrane tyrosine kinase) activation by both glial cell line-derived neurotrophic factor (GDNF) and nerve growth factor (NGF) triggers the formation of RET51/FKBP52 complex (Fusco et al. 2010). Mutations in Ret51 and FKBP52 that disrupt this complex have been found in rare cases of parkinson’s disease (PD) (Fusco et al. 2010).

Overexpression of FKBP52 increases aggregation of alpha-synuclein, commonly observed in patients with PD (Gerard et al. 2010). However, FKBP52 seems to play an opposite role in Alzheimers disease (AD). Overexpressioin of FKBP52 in Drosophila suppressed the toxicity of amyloid beta, a protein which forms aggregates in AD (Sanokawa-Akakura et al. 2010). During early neural differentiation, FKBP52 enhances and FKBP51 inhibits neurite outgrowth (Quinta et al. 2010). Whereas in undifferentiated cells, FKBP52, HSP90 and p23 are located in the perinuclear region, they are rapidly dispersed in the cytoplasm upon induction of differentiation (Quinta & Galigniana 2012). Neuronal damage induces redistribution of these three proteins to the perinuclear region. During adipocyte differentiation, FKPB51 rapidly increases whereas FKBP52 decreases. In addition, FKPB51 translocates from the mitochondria to the nucleus, a process which is dependent on PKA (protein kinase A) (Toneatto et al. 2013). Also in other cell types, like 3T3 fibroblasts, FKBP51 has been found in the mitochondria, which is also true for GR, but this subcellular localization and the trafficking of both proteins is independent from each other (Gallo et al. 2011). FKPB51 protects cells from oxidative stress and translocates to the nucleus upon exposure to oxidative stress (Gallo et al. 2011). FKBP38 (not mentioned in the table) has an anti-apoptotic function by anchoring the pro-apoptotic proteins Bcl-2 and Bcl-x(L) to mitochondria (Shirane & Nakayama 2003).

Cyp40 (cyclophilin 40), also an immunophilin, is involved in steroid hormone receptor assembly. It prefers HSP90 over HSP70 binding (Carrello et al. 2004). Cyp40 binds to the DNA repair protein Ku70 and to Rack1 (receptor of activated protein kinase C 1) (Park et al. 2011). Rack1 promotes stabilization of HIF-1alpha (hypoxia inducible factor 1 alpha) in prostate cancer cells (Yu et al. 2014). However, RACK1 can also down-regulate HIF-1alpha after cobalt chloride treatment of MCF-7 breast cancer cells (Park et al. 2011).

PP5 (protein phosphatase 5, also abbreviated PPP5C) is a serine/threonine protein phosphatase that functions in signaling pathways that control cellular responses to stress, steroid hormones and DNA damage (Chinkers 2001; Yang et al. 2005; Hinds & Sanchez 2008). PP5 phosphatase activity is auto-inhibited in a state unbound to HSP90 or fatty acids by a special conformation of the TPR domain and a C-terminal subdomain. Binding to HSP90 or fatty acids (e.g. arachidonic acid) to the TPR domain activates phosphatase activity (Yang et al. 2005). PP5 is up-regulated in breast cancer, as shown by immunohistochemical staining of a tissue microarray (TMA) (Golden et al. 2008).

The translocation machinery (or “translocase”) of the outer mitochondrial membrane (TOM) is a multiprotein complex containing, among other proteins, TOM20 and TOM70 (named after their approximate molecular weight). Tom70 preferentially binds pre-proteins with internal targeting sequences (Brix et al. 1997). TOM70 binds to a multichaperone complex including HSP70 and HSP90 via specialized TPR domains in which three TPR domains are organized into a superhelical structure (Scheufler et al. 2000; Young et al. 2003a). Proteins are translocated through the import pore to the TIM (translocase of the inner mitochondrial membrane) machinery in an ATP-dependent manner (Young et al. 2003a). Pre-proteins with N-terminal pre-sequences are translocated with the help of Tom20, a process which is less dependent on chaperones (Brix et al. 1997).

TPR2 binds to Rad9 (Xiang et al. 2001), which is part of the Rad1-Hus1-Rad9 complex that is important for DNA-damage induced cell-cycle checkpoint activation (Parrilla-Castellar et al. 2004). The J-domain (DNAJ homology domain) of TPR2 regulates TPR2-Rad9 interaction (Xiang et al. 2001) and thereby potentially the stability of the checkpoint complex.

Deletion of CPR7 (7th cyclophilin homolog, interacting with the yeast global transcriptional regulator Rpd3) caused severe growth defects when combined with mutations that decrease the amount of Hsp90 or Hop. Cpr7 null mutations caused defects in GR and pp60/v-src kinase activity (Duina et al. 1996b). In addition, Cpr7 deletion resulted in a significant impairment of cell division (Duina et al. 1996a). Both Cpr6 and Cpr7 interact with the ribosome and may be involved in the regulation of protein synthesis (Tenge et al. 2015).

The HSP90 homodimer can bind several co-chaperones at once, even different co-chaperones that bind to the TPR domain. For example, a HSP90(2)-FKBP52(1)-p23(2)-HOP(2)-complex could be demonstrated by immunoprecipitation, chromatographic methods and dynamic light scattering (Hildenbrand et al. 2011).

The chaperone cycle

After binding of the client protein to the HSP90 homodimer (open configuration), ATP binds to the NTD and induces transition of the complex to a closed ATP bound form. During proper folding, the conformation closes further and the NTDs dimerize transiently, coupled to ATP hydrolysis (Prodromou et al. 2000). Finally, the configuration opens up, the folded client protein is released and ADP dissociates (figure 2).

An external file that holds a picture, illustration, etc.
Object name is nihms909442f2.jpg

Simplified chaperone cycle. ATP and ADP bound states during protein folding as well as open configuration of HSP90 homodimer are shown. NTD, N-terminal domain; CTD, C-terminal domain; MD, middle domain.

For proper functioning of this simplified chaperone cycle, several co-chaperones are necessary. The interaction with co-chaperones occurs in a ordered, sequential manner (Li et al. 2011). The first co-chaperone entering the cycle is the HSP90 ATPase inhibitor Hop (Chen & Smith 1998), that binds to the Hsp90 homodimer. Together with Hop or after that, Hsc70 binds. After that, a PPIase, such as Fkbp51, Fkbp52 or Fkbp37 binds to the opposite side of the HSP90 dimer. HSP90 converts to a closed conformation after binding of ATP and binding of p23 (Johnson et al. 1994; Johnson & Toft 1995; Freeman et al. 2000; McLaughlin et al. 2006; Forafonov et al. 2008), releasing Hop and HSC70 from the complex, leaving space for another PPIase molecule. After ATP hydrolysis, all remaining co-chaperones are released from the HSP90 dimer which converts to an open conformation. This allows the start of another HSP90 cycle (Li et al. 2011). In a cell-free system with progesterone receptor (PR) as a client protein it could be shown that such an assembly/disassembly cycle has a t1/2 of about 5 min (Smith 1993).

Mechanism of HSP90 induction

In general, heat shock generates a higher induction of HSP90AA1 compared to HSPAB1 (Ullrich et al. 1989). The first level of regulation is during gene transcription. HSF1 (heat shock factor 1), which binds to HSE (heat shock elements), is thought to be the major transcription factor for HSPs (Ciocca et al. 2013). However, the basal level of HSP90AB1 transcripts does not depend on HSF1 since HSF1 knockout leads to no transcript reduction in a model of oocytes (Metchat et al. 2009). In contrast, HSP90AA1 transcripts are reduced by ~50 % in HSF1 knockout oocytes (Metchat et al. 2009). The functional importance of the two HSE in the HSP90AA1 promoter has been demonstrated with reporter constructs in Jurkat cells (Zhang et al. 1999). The HSP90AB1 promoter contains a TATA box and a CAAT box. In addition, there is an Sp1 binding site and typical as well as some atypical HSEs are located up- and downstream the transcription initiation site (Rebbe et al. 1989). The two typical HSEs in the first intron are functionally important for inducibility of HSP90AB1 by heat shock (Shen et al. 1997). Binding of KLF4 (Krüppel-Like-Factor 4) to the promoters of HSP90AB1 and HSP90AA1 leads to a higher expression of these HSPs (Liu et al. 2010). During induction of HSP90AB1 after heat shock, Stat1 phosphorylation is indispensable. The dominant kinases for phosphorylation of Stat1 (Y701 and S727) are Jak2 and PKCepsilon (Cheng et al. 2010). However, the activation of these kinases requires the association with an HSP90 protein, leading to a positive auto-regulatory loop. For full activation, Brg1 (brm (Brahma)/SWI2-related gene), an ATPase subunit of the SWI/SNF chromatin remodeling complex, is recruited to the transcription factor complex (Cheng et al. 2010). The mTORC1 (mTOR complex 1) regulates mRNA translation of HSP90AB1 as detected in a high-resolution transcriptome-scale ribosome profiling analysis. It has been found that a TOP (5′ terminal oligopyrimidine) motif in HSP90AB1 is necessary for this translational regulation, although the mechanism is not known (Thoreen et al. 2012). As a general negative feedback mechanism, the HSP90 complex represses HSF1 activation, thereby inhibiting an over-activation of the HSF1 response (Zou et al. 1998).

Posttranslational modifications

The tyrosine (Y) phosphorylation of Hsp90AA1 and Hsp90AB1 induced by LPS in endothelial cells is mediated by pp60src (Barabutis et al. 2013). Mass spectrometry identified Y309 as a major site of Y phosphorylation on Hsp90AA1 (and Y300 on Hsp90AB1) (Barabutis et al. 2013). Phosphorylation of Ser225 and Ser254 (counted after removal of the first methionine) of HSP90AB1 inhibited binding of AHR (arylhydrocarbon receptor) (Ogiso et al. 2004). The phosphorylation of HSP90 is inhibited by 17-allyl-amino-demethoxy-geldanamycin (17-AAG) (Barabutis et al. 2013). DNA-activated protein kinase (DNA-PK), a protein that is important for DNA repair, phosphorylates two threonine residues in HSP90AA1. Since HSP90AB1 lacks these threonine residues, no DNA-PK dependent phosphorylation occurs in HSP90AB1 (Lees-Miller & Anderson 1989a).

Methylated HSP90AB1 increases the proliferation of cancer cells (Hamamoto et al. 2014). Methylation of lysines 531 and 574 of HSP90AB1 occurs by SMYD2 (SET/MYND domain containing 2) histone H3 lysine (K) 36-specific methyltransferase (KMT) (Donlin et al. 2012; Hamamoto et al. 2014). Methylated HSP90AB1 is associated with increased proliferation of cancer cells (Hamamoto et al. 2014). Acetylation seems to inhibit HSP90 function (Yang et al. 2013) as shown in a model of Gaucher disease. This is a lysosomal storage disorder caused by mutations in the enzyme glucocerebrosidase (GCase), leading to a misfolded enzyme. HSP90AB1 is essential for the degradation of misfolded GCase by guiding it to a valosin-containing protein (VCP)-associated proteasomal degradation pathway (Dai & Li 2001; Zhong et al. 2004; Yang et al. 2013). Inhibition of HDAC (histone deacetylases) keeps HSP90 in an acetylated state and inhibits GCase degradation, thereby alleviating the disease phenotype (Yang et al. 2013). Conversely, HSP90 also influences HDAC levels. Hsp90 interacts with HDAC6 and is a key effector of HDAC6 levels (New et al. 2013). Nitric oxide (NO) induces S-nitrosylation of HSP90AB1 at its C-terminus and thereby reduces its activity (Retzlaff et al. 2009). Obviously, also citrullination occurs on HSP90 proteins since autoantibodies against citrullinated HSP90AA1 and HSP90AB1 are detected in patients with rheumatoid arthritis-associated interstitial lung disease (RA-ILD), but not in patients with RA without ILD, MCTD (mixed connective tissue disease) and IPF (interstitial pulmonary fibrosis) (Harlow et al. 2013).

Inhibition of HSP90

Derivatives of geldanamycin, like 17-dimethylaminoethylamino-17-demethoxy-geldanamycin (17-DMAG) or 17-AAG bind to the ATP pocket in the NTD of HSP90 (Grenert et al. 1997; Prodromou et al. 1997b; Stebbins et al. 1997; Roe et al. 1999). The HSP90 inhibitor radicicol, a macrocyclic antifungal antibiotic, also binds to the ATP binding pocket of HSP90 (Schulte et al. 1998; Roe et al. 1999). Celastrol, a triterpenoid compound that inhibits HSP90AB1 activity, alters the three-dimensional structure of the co-chaperone p23 resulting in a more selective destabilization of steroid receptors (Chadli et al. 2010) or acts by disrupting the HSP90-Cdc37 complex (Zhang et al. 2009). Celastrol also inhibits the production of matrix metalloproteinases (MMPs), inducible nitric oxide synthase (iNOS) and cyclooxygenase-2 (COX-2) from chondrocytes derived from patients with osteoarthritis (Ding et al. 2013). This effect may be mediated by direct interaction of HSP90AB1 with these proteins or by its interaction with NF-kappaB (nuclear factor kappaB) (Ding et al. 2013). The natural product drug BTIMNP_D004 inhibits, dependent on cell type, HSP90AB1 and/or HSP90AA1 and a number of proliferation-associated signaling proteins (Samadi et al. 2009). A kind of non-specific HSP inhibition is exerted by free radicals formed during oxidative stress. They can lead to a cleavage of HSP90AB1 at two positions near the N-terminus of the protein, which occurs through a Fenton-Reaction (Beck et al. 2012).

Inductors of HSP90(AB1)

As already described, there is a great number of stress types that are able to induce HSP protein levels. In this paragraph, some studies are summarized that show a specific induction of HSP90AB1. Infection is a basic type of stress and therefore a common reason for HSP induction. Increased HSP90AB1 has been detected after influenza infection in infected cells (Wahl et al. 2010). Bacterial challenge in the abalone Haliotis tuberculata leads to an up-regulation of HSP90AB1 (Travers et al. 2010). LPS treatment or challenge with src-adenovirus lead to an increased phosphorylation of HSP90AB1 (Y300) in endothelial cells (Barabutis et al. 2013).

Necrosis causes an inflammatory reaction in order to clean up the damaged region. Extracellular HSP90AB1 can elicit such a response independently from IL-8 and MCP-1 in cultured retinal pigment epithelial (RPE) cells. A cell-impermeable geldanamycin derivative can inhibit the production of inflammatory cytokines in this model (Qin et al. 2011). H. pylori induces the translocation of HSP90AB1 from the cytosol to the membrane causing interaction of HSP90AB1 and Rac1, which leads to the activation of NADPH oxidase and production of ROS in gastric epithelial cells (Cha et al. 2010).

Heat shock or increased expression of HSP90AB1 restore reduced infectivity of HIV with incompletely processed capsid-spacer protein 1 (CA-SP1) (Joshi & Stoddart 2011; Joshi et al. 2013). Increased thermotolerance of hepatoma cells (Reuber H35 cells) in late G1/early S-phase as compared to G0 phase correlated with enhanced basal synthesis of HSP65, HSP68 and HSP90AB1 as well as with increased phosphorylation of non-histone chromosomal proteins (van Dongen et al. 1986). Not only heat stress but also cold stress can induce HSPs. HSP90AB1 is up-regulated during hibernation in the hearts of ground squirrels and protects them from ischemia-reperfusion injury and apoptosis (Grabek et al. 2011).

Benzo[a]pyrene (B[a]P), a cancer promoting agent, induces HSP90AB1 mRNA (Sadeu & Foster 2013). Chronic opioid treatment leads to increased HSP90AB1 transcript and protein levels. This contributed to an enhanced responsiveness of adenylate cyclase to forskolin treatment. Thus, HSP90AB1 contributes to opioid induced adenylate cyclase sensitization (Koshimizu et al. 2010). Overexpression of TRIM14 (T-cell receptor interacting molecule 14) leads to increased levels of HSP90AB1 (Nenasheva et al. 2014). Trim14 increases mesodermal differentiation and inhibits ectodermal differentiation of mouse ES cells (Novosadova et al. 2009).

Also normal developmental processes as shown in Drosophila can induce HSP (Lindquist 1986). Induction of HSP90AB1 occurs during adipogenesis in a cell culture model (Fromm-Dornieden et al. 2012).

HSP90AB1 and protein transport

As already briefly mentioned, HSP90AB1 is necessary for the shuttle of client proteins between cytoplasm and nucleus (Galigniana et al. 2010a; Galigniana 2012). The interaction of steroid hormone receptors with Hsp90 is required for their translocation to the nucleus (Czar et al. 1997). In addition to Hsp90 homodimers, transport protein complexes consist of Hsp70, Hop, Hsp40, p23 and PP5 (Chen et al. 1996; Silverstein et al. 1997; Pratt et al. 1999; Murphy et al. 2001). The protein complexes are transported along microtubular tracks with the help of the motor protein dynein (Harrell et al. 2004; Pratt et al. 2004), as shown for the tumor suppressor protein p53 (Galigniana et al. 2004). Directionality is mediated by recognition sequences in the client proteins and/or by interaction with additional proteins. For example, immunophilins target the complexes in the nuclear direction (Pratt et al. 1999), as shown for FKBP52 (Czar et al. 1995). For this transport, a nuclear localization sequence (NLS) is necessary. However, it has been shown, that in addition to the classical NLS (NS1), also other recognition sequences contribute to the guiding of proteins in the direction of the nucleus (Piwien Pilipuk et al. 2007). For transport in the opposite direction (nucleus-to-membrane transport) it has been suggested that the transport is mediated by interaction of the transport complex with Cdc37 instead of FKBP52 (Pratt et al. 1999).

Nuclear transport of GR occurs after hormone binding and exchange of FKBP51 for FKBP52, generating a GR-HSP90-FKBP52-dynein complex that travels to the nucleus (Davies et al. 2002). FKBP52 facilitates hormone binding (Riggs et al. 2003). Interaction with dynein is mediated by the PPI domain of immunophilins (Galigniana et al. 2002) via the dynamitin component of the dynactin complex (Galigniana et al. 2004). The requirement of dynein for steroid receptor movement along microtubules has been demonstrated in a cell-free system (Harrell et al. 2004). At the nuclear pore, the complex interacts with importin-beta and the nucleopore glycopotein Nup62 and passes undissociated (Echeverria et al. 2009). Dissociation occurs within the nucleus, releasing GR for binding to DNA. A similar mechanism applies for the MR (Galigniana et al. 2010b), and in general also for plant cells (Harrell et al. 2002). With the MR model it has been shown that the whole undissociated complex is detectable in the nucleus for several minutes after translocation (Galigniana et al. 2010b; Grossmann et al. 2012). One other protein that can be associated with the nuclear translocation complex is RAC3 (receptor-associated co-activator 3), a protein that exerts an anti-apoptotic activity by activating NF-kappaB, AKT and p38 MAPK and by inhibition of caspase 9, AIF (apoptosis-inducing factor) and ERK2 (extracellular signal regulated kinase 3/MAPK6) (Colo et al. 2008; Alvarado et al. 2014).

HSP90 and protein degradation

A few examples of HSP90-associated protein degradation have been described. Apo(lipo)proteinB (apoB) is essential for the assembly and secretion of lipoproteins. In a cell-free system it has been shown that apoB degradation is dependent on HSP90 (Gusarova et al. 2001). The degradation of misfolded proteins can be mediated by HSP90. The VHL (von-Hippel-Lindau) tumor suppressor protein possesses ubiquitin ligase activity and is involved in the degradation of hypoxia inducible factor (HIF) (Maxwell et al. 1999). Folding of VHL and degradation of mutated VHL is dependent on HSP70 whereas the chaperonin TriC is necessary for its folding but not for the degradation (McClellan et al. 2005). In contrast, HSP90 is not responsible for VHL folding or maintenance of misfolded protein solubility but for its degradation (McClellan et al. 2005). The role of HSP90AB1 in degradation of misfolded GCase in Gaucher disease has been described above and its role in cystic fibrosis is described in the following paragraph.

Diseases aggravated by HSP90AB1 or its co-chaperones

Cystic fibrosis (CF, mucoviscidosis) is a genetic disorder with increased viscosity of various secretions leading to damage in the affected organs, most importantly lung and pancreas. The genetic basis is a mutation in the cAMP-regulated chloride ion channel CFTR (cystic fibrosis transmembrane conductance regulator), in most cases a deletion of Phe508. This causes a maturation defect of the protein, leaving the nucleotide binding domain 2 (NBD2) permanently in a protease-sensitive state (Riordan 2005) which leads to increased processing in the ERAD (endoplasmic reticulum associated degradation) pathway. Linked by Hop, CFTR binds to both an Hsc-Hsp40/70 complex and also to Hsp90. Whether proteins are properly folded for export from ER by ERAF (ER-associated folding) machinery or degraded by ERAD seems to be dynamically controlled by co-chaperones. Deletion of AHA1 leads to stabilization of mutated CFTR, a finding that may lead to a potential therapy for CF (Wang et al. 2006).

Diseases alleviated by HSP90AB1

Bronchopulmonary dysplasia (BPD), a disease caused by prolonged high oxygen delivery to immature lungs involves cellular damage by oxygen free radicals. Thioredoxin-1 (Trx) is a radical scavenger stabilized by HSP90ab1 and HSP90AB2. Consequently HSP90 proteins prevent hyperoxic cell death (Floen et al. 2014).

Genetic variations

Genetic variations in HSPAB1 have been described that are likely to have an influence on stress-induced mortality, GR level, and GRE (glucocorticoid-response element) binding actvity in C57BL/6 mice compared to BALB/c mice (Shen et al. 2010). Some SNPs in HSP90AB1 in laying hens are associated with longer life and higher productivity (Sun et al. 2013), but a very limited number of SNPs in human HSP90AB1 are likely to have a functional consequence (Urban et al. 2012). SNPs in HSP90AB1 in Thai native cattle are associated with increased heat tolerance (Charoensook et al. 2012). A study on a turkish population found significantly more polymorphisms in HSP90AA1, HSP90AB1 and HSP90B1 in patients with non-small cell lung cancer (NSCLC) (Coskunpinar et al. 2014).

Role in molecular evolution

Mutations can cause defects in protein folding. Their stability can often only be maintained by binding to HSPs. In addition, molecular evolution by gene duplication and subsequent progressive mutation is only conceivable with the help of a strong and versatile chaperone machinery (Rutherford & Lindquist 1998; Yahara 1999; Jarosz & Lindquist 2010; Jarosz et al. 2010).

HSP90AB1 client proteins

Large scale protein-protein interaction studies revealed many new client proteins (Tsaytler et al. 2009; Taipale et al. 2014). Although the spectrum of interaction partners can be well described, they cannot be easily predicted based on protein structures. Therefore, it remains largely enigmatic why some proteins are clients of HSPs and others not. However, interaction of HSP90 chaperone and its Cdc37 co-chaperone with kinases depends on the thermal stability of the kinase domain (Taipale et al. 2012) which can be exploited for high-throughput in-vivo screens for kinase inhibitors (Lambert et al. 2013; Taipale et al. 2013). In addition to the well-known interaction with steroid hormone receptors (Catelli et al. 1985; Mendel & Orti 1988; Rexin et al. 1991), HSP90 interacts with 7 % of transcription factors, 30 % of ubiquitin ligases and 60 % of kinases. A study of the MAPK interactome revealed the interaction of HSP90AB1 with a large number of MAPK, and its functional relevance in examples such as MAPK6 and p38MAPK/MAPK14 has been demonstrated (Bandyopadhyay et al. 2010).

PTK6 (protein tyrosine kinase 6) is a client protein of HSP90 as described below in the cancer paragraph (Kang et al. 2012). ERK5 (extracellular regulated kinase 5) is retained in the cytoplasm by binding to a HSP90AB1-Cdc37 complex. Activated ERK5 dissociates from HSP90AB1-Cdc37 and can subsequently be transported to the nucleus (Erazo et al. 2013). Nanog is an important transcription factor in ES (embryonic stem) cells. TRIM8, a member of the tripartite motif (TRIM) family, binds to HSP90AB1 and inhibits translocation of SMAD3 into the nucleus thereby reducing the transcription of Nanog in ES cells (Okumura et al. 2011). A novel ADP-dependent HSP90 interaction with the cysteine- and histidine-rich domain (CHORD)-containing protein CHORDC1 (Gano & Simon 2010), a protein associated with inclusion conjunctivitis.

The serine/threonine kinase Wee1 is necessary G2/M checkpoint in yeast and is a client of HSP90AB1. USP50 (ubiquitin-specific protease 50), a deubiqitinating enzyme (DUB), prevents Wee1 degradation, thereby repressing entry into mitosis following activation of the G2/M DNA damage checkpoint (Aressy et al. 2010). Conversely, HSP90AB1 interacts with cell cycle regulator Cdc25A, which promotes cell cycle progression (Giessrigl et al. 2012). HSP90 stabilizes tau protein which forms precipitates in Alzheimer disease and other tauopathies (Karagoz et al. 2014). FKBP51 recruits HSP90 to tau and helps to stabilize tau and to maintain hyperphosphorylated Tau levels, a status that perpetuates Alzheimer’s disease (Jinwal et al. 2010). Thus, a possible therapeutic approach for tauopathies is the inhibition of HSPs. In a mouse model, it has been demonstrated that inhibition of HSP90 promotes Tau degradation (Dickey et al. 2007). One possible mediator is the co-chaperone CHIP, a tau ubiquitin ligase. It has also been shown that the inhibition of another co-chaperone, Cdc37, leads to tau degradation (Jinwal et al. 2011). This is also true for other co-chaperones like p23, PP5, FKBP51 and S100A1 (Jinwal et al. 2013). All these might be valuable therapeutic targets. Methylene blue (MB) is a small molecule that leads to tau degradation. Treatment with MB leads to dissociation of HSP70 from tau and increased association of HSP90 with tau (Thompson et al. 2012). In this model, HSP90 is important in tau clearance from cells, as shown by siRNA mediated HSP90 knockdown.

The transcription factor NF-kappaB is involved in cell cycle control, cellular differentiation, and inflammation. It is activated by degradation of its inhibitor IkappaB (inhibitor of NF-kappaB), mainly IkappaBalpha. This ubiquitin-dependent degradation is initiated by phosphorylation of an IKK (IkappaB kinase or IKBK) complex that usually consists of three subunits, IKK alpha, beta and gamma (Karin & Lin 2002; Senftleben & Karin 2002). HSP90 is required for IKK biosynthesis and function (Broemer et al. 2004). A heterocomplex of Cdc37 and HSP90 is required for TNFalpha dependent activation of IKK (and NF-kappaB) (Chen et al. 2002; Hinz et al. 2007). In contrast, HSP70 binds to TRAF6 (TNF receptor associated factor 6), thereby inhibiting LPS (lipopolysaccharide)-triggered NF-kappaB activation (Chen et al. 2006b). Immunophilins have opposite effects on NF-kappaB. Whereas FKBP52 enhances NF-kappaB activity, FKBP51 plays an inhibitory role (Erlejman et al. 2014b).

HSPs interact with the cytoskeleton, on one side leading to modification and stabilization of cytoskeletal structures and on the other side using the cytoskeleton for transport purposes (Quinta et al. 2011). HSP90 binds to actin and crosslinks actin filaments (Koyasu et al. 1986). HSP90 also binds to N-WASP (neuronal Wiskott-Aldrich Syndrome Protein, now called WASPL, WASP-like), a protein which activates the Arp2/3 (actin related/released protein 2/3) complex. The Arp2/3 complexes serve as nucleation sites for actin polymerization. HSP90 bundles branched actin filaments and promotes the formation of unbranched actin structures (Park et al. 2007). HSP90 increases N-WASP phosphorylation by v-Src, thereby enhancing actin polymerization (Park et al. 2005). In addition, HSP90 protects N-WASP from proteasomal degradation (Park et al. 2005). Upon HSP90 inhibition, association of HSP90 with G-actin (globular actin) is increased, associated with decreased actin polymerization and concomitantly decreased cell motility and invasion (Taiyab & Rao Ch 2011). HSP90 also binds to several other actin-binding proteins like calponin (Ma et al. 2000) and LIMK (LIM kinase) (Li et al. 2006). LIMK contains a LIM domain, named after the three proteins LIN-11, ISL1 and MEC-3 that contain this domain. HSP90 activates LIMK by promoting its homodimerization (Li et al. 2006). LIMK phosphorylates and thereby activates cofilin, a protein that promotes disassembly of actin filaments. Thus, HSP90 is able to promote antagonizing functions with regard to actin polymerization. ILK (integrin linked kinase) is located at the interface between actin filaments and cell membrane. ILK is stabilized by HSP90 and HSP90 inhibition leads to decreased cell migration (Radovanac et al. 2013). Both HSP90AA1 and HSP90AB1 bind to microtubules, dependent on the level of tubulin acetylation (Giustiniani et al. 2009). This tubulin acetylation also influences binding and activity of the HSP90 client proteins Akt and p53 (Giustiniani et al. 2009). HSP90 protects tubulin, the major component of microtubules, from thermal injury (Weis et al. 2010). Some effects of HSP90 on the cytoskeleton may be mediated by other proteins like the lysine demethylase Kdm3a. Mutant Kdm3a mice display male infertility and their cells show an altered fractionation of actin and tubulin (Kasioulis et al. 2014).

Diseases with down-regulated HSP90

Hsp90ab1, expressed in oligodendrocyte precursor cells, has been shown to be down-regulated in a rat model of acute spinal cord injury (ASCI) (Zhou et al. 2014). HIV virus causes a progressive decline in CD4+ T-cells. This involves an aging-like phenotype with decline of telomerase activity. This is, at least partly, caused by a reduction of HSP90AB1 which is mediated by the HIV-protein Tat (Comandini et al. 2013). In a rat model of radiation-induced fibrosing alveolitis, it has been shown that HSP90AB1 was progressively down-regulated in radiation-damaged lung cells other than mast cells (Haase et al. 2014).

Diseases with up-regulated HSP90

HSP90 is increased in systemic sclerosis (SSc) of the skin. Inhibition of Hsp90 by 17-DMAG inhibited canonical TGF-beta signaling and completely prevented the stimulatory effects of TGF-beta on collagen synthesis and myofibroblast differentiation (Tomcik et al. 2014). Thus, HSP is pro-fibrotic in this model. Conversely, HSP90 inhibitors repress TGFbeta1 signaling (Noh et al. 2012) by a mechanism dependent on Smurf2-mediated degradation of TGFbetaRII. This lead to a decreased expression of alpha-smooth muscle actin, fibronectin, and collagen I and a decrease of renal fibrosis in an animal model (Noh et al. 2012).

Miscellaneous roles of HSP90 in diseases

Huntington’s disease, an autosomal dominant disease, is caused by increased amounts of repetitive trinucleotide repeats (CAG) in the vicinity of promoters which leads to precipitation of polyglutamin protein. This leads to massive loss of interneurons in the caudate nucleus, especially affecting GABA- and cholinergic neurons. The trinucleotide expansion comprises up to 100 repeats whereas 11–35 repeats are normal. Inhibition of cytoplasmic HSP90 supports clearance of Huntingtin protein and is a promising approach for the treatment of Huntington’s disease (Ernst et al. 2014).

In trauma patients, a higher HSP90ab1 promoter activity (−144AA genotype) was associated with low expression of the inflammatory cytokine TNFalpha (tumor necrosis factor alpha) and with lower organ dysfunction scores (Zhao et al. 2013). High levels of HSP90ab1 resulted in lower hypoxia-induced apoptosis in a model of intestinal epithelial cells (IEC) (Zhang et al. 2013).

DNA damage recognition/DNA repair/Free radicals

The tumor suppressor protein p53 is normally degraded after binding to the E3 ubiquitin ligase MDM2 (mouse double minute 2 homolog). Mutant p53 is commonly overexpressed in cancer cells. HSP90AB1 binds to mutant p53 in cancer cells and stabilizes it (Sepehrnia et al. 1996). P53 binds with its DNA binding domain to the middle and CTD of HSP90AB1 (Muller et al. 2004). It has been demonstrated that the complex is only stable in the presence of HSC70, HSP40, HOP and ATP (King et al. 2001). Degradation of wild type p53, which is initiated by MDM2, is inhibited by the co-chaperone TPR2 that binds p53 and thereby blocks complex formation with MDM2 (Kubo et al. 2013). In the latter complex, no HSP90 has been described (Kubo et al. 2013). After exposure to ionizing radiation, inhibition of HSP90 by a geldanamycin analogue inhibits p53 activation and subsequent apoptosis in human peripheral blood mononuclear cells (Fukumoto & Kiang 2011). In a cell-free system, HSP90AA1 has been shown to stabilize wild-type p53 and to maintain binding of p53 to the p21 promoter sequence (Walerych et al. 2004).

The stability of DNA-PKcs is dependent on the activity of HSP90, since inhibition of the HSP90 with radicicol decreased the expression of DNA-PKcs in HeLa cells (Falsone et al. 2005). HSP90(AA1 and AB1) specifically binds to REV1 (reversionless 1), a Y-family polymerase that causes lack of UV-induced reversion of the arg4–17 ochre allele in yeast and plays a central role in mutagenic translesion DNA synthesis (TLS) (Lawrence & Christensen 1976; Ohmori et al. 2001; Pozo et al. 2011). HSP90 inhibition disrupts the interaction between REV1 and monoubiqitinated PCNA (proliferating cell nuclear antigen), a molecule that recruits REV1 to damaged DNA sites. Thus, HSP90 inhibition suppresses TLS-mediated mutagenesis (Pozo et al. 2011). HSP90AB1 inhibition decreased IL-1beta induced NO (nitric oxide) production in chrondrocytes (Calamia et al. 2011).

Cancer

HSPs are frequently up-regulated in cancer. One way of promoting cancer cell survival and proliferation is by stabilizing proteins with activating mutations, and other cancer promoting proteins that would otherwise be degraded. A systematic analysis of the expression of HSP90AA1 and HSP90AB1 proteins, their co-chaperones (Aha1, Cdc37, p23, Tpr2) and the Hsp90 dependent transcription factor HSF1 in 17 cancer types suggested that the overexpression is tumor-specific and obviously random. However, the study demonstrated that in more than half of the tumor tissues at least one of the factors mentioned is overexpressed (McDowell et al. 2009). In salivary gland tumors, the expression of HSP90AA1 and HSP90AB1 correlated with malignancy, proliferation, neural invasion and metastasis (Wang et al. 2013). HSP90AB1 is also up-regulated in lung cancer cells and this up-regulation correlates with pathological grade, lymphatic invasion and poor survival (Biaoxue et al. 2012). In lung cancer cell lines, it has been shown that resistance to ionizing radiation is partially caused by stabilization of a HSP90-HIF1alpha (hypoxia inducible factor 1 alpha) complex and that inhibition of HSP90 increased radiosensitivity of the tumor cells (Kim et al. 2009). HSP90AB1 is overexpressed in GISTs (gastrointestinal stromal tumors) (Tsumuraya et al. 2010), tumors that originate from gastrointestinal pacemaker cells (Cajal cells). In melanoma tumor cells it was shown that Bcl-2 protein stabilizes HIF-1 (hypoxia-Inducible Factor 1), a critical mediator of the cellular response to hypoxia. Increased HIF-1 leads to increased tumor angiogenesis, metastasis, therapeutic resistance and poor prognosis. The stabilization of HIF-1 is due to the formation of a ternary complex between HIF-1, Bcl-2 and HSP90AB1 (Trisciuoglio et al. 2010). The proliferation of Multiple Myeloma (MM) cells is critically dependent upon a number of protein kinases. One of them is GRK6 (G-protein coupled receptor protein kinase 6) which is required for STAT3 phosphorylation and subsequent activation of MCL1 (myeloid cell leukemia 1) protein (Tiedemann et al. 2010). GRK6 expression is regulated by HSP90AA1 and HSP90AB1 (Tiedemann et al. 2010). Tumor cell growth to a visible size is critically dependent on angiogenesis. HSP90AB1 interacts with and stabilizes the mRNA of the VEGF-A-induced pro-angiogenic protein BAZF (BCL-6 associated zinc finger protein) and thus may positively regulate angiogenesis (Miwa et al. 2013). A fibronectin (FN) matrix is important for cell migration and invasion, processes that are features of cancer cells (Lowrie et al. 2004; Lobert et al. 2010). HSP90AB1 stabilizes extracellular FN in a study with the breast cancer cell line Hs578T (Hunter et al. 2014). Reduction of HSP90AB1 leads to a reduction in extracellular fibronectin (Hunter et al. 2014). The tumor suppressor kinases of the hippo pathway, LATS1 (large tumor suppressor kinase 1) and LATS2, are clients of HSP90AB1. In this case, inhibition of HSP90AB1 leads to a disruption of the LATS tumor suppressor pathway, i.e. to a promotion of cancer growth (Huntoon et al. 2010). In a cell culture model with the hepatocellular carcinoma cell line HepG2 it was shown that HSP90AB1 expression was up-regulated by Hepatitis B virus encoded X protein (HBx) (Li et al. 2009). The oncogenes c-Myc and MYCN determine malignancy in neuroblastomas. Inhibition of HSP90 leads to inhibition of neuroblastoma cell proliferation and to accumulation of p53 (Regan et al. 2011).

Up-regulation of HSP90 is one reason for increased telomerase activity in tumors, as shown in a model of prostate cancer (Akalin et al. 2001). HSP70, p23 and HSP90 are associated with hTERT (human telomerase reverse transcriptase subunit), but upon recruitment of hTR (human telomerase RNA template), HSP70 dissociates from this complex (Forsythe et al. 2001). Overexpression of hTERT together with HSP90 in pheochromocytomas has been shown to be associated with more aggressive tumors (Boltze et al. 2003). In transformed natural killer (NK) cells it has been shown that, after IL-2 stimulation, hTERT forms a complex with HSP90, Akt, mTOR (mammalian target of rapamycin) and p70S6 kinase (70 kDa ribosomal subunit S6 kinase) (Kawauchi et al. 2005). In contrast, disruption of HSP90 function by geldanamycin leads to proteasome-mediated degradation of hTERT. In this model, the RING finger gene MKRN1 (makorin ring finger protein 1) is the E3 ligase that mediated ubiquitination of hTERT (Kim et al. 2005). In addition to the mechanisms described above, HSP90 can interact with the hTERT promoter and thereby increase promoter activity. This interaction is independent of the transcription factors Sp1 and c-Myc, that also bind to this promoter (Kim et al. 2008).

Caspase-induced cleavage of the co-chaperone p23 leads to a truncated p23 molecule, deltap23. Overexpression of a truncated p23 corresponding to deltap23 leads to a decrease in hTERT levels and to down-regulation of telomerase activity (Woo et al. 2009).

Protein tyrosine kinase 6 (PTK6, also known as breast tumor kinase, BRK) is expressed in most breast cancer tissues and cell lines as well as in other tumors. In normal breast epithelial cells, it is expressed but obviously not active (Peng et al. 2014). In breast cancer cells, PTK6 is activated by Y342 phosphorylation and recruited to the cell membrane (Peng et al. 2014). HSP90 stabilizes PTK6 and may therefore be a therapeutic target for cancers with overexpression of PTK6 (Kang et al. 2012). High mRNA expression driven by chromosome coding region amplifications of HPS90AB1 correlated with poor prognosis in Her2-negative/ER-positive breast cancer (Cheng et al. 2012).

Treatment of AML (acute myeloic leukemia) patients with the HDACi (histone deacetylase inhibitor) valproic acid (VPA) increases survival in more than 20 % of patients with advanced disease due to increased apoptosis of AML cells. HSP90AB1 is among the genes up-regulated in response to VPA and might be a resistance mechanism. Indeed, the rate of apoptosis in AML cells could be increased by parallel administration of geldanamycin (Forthun et al. 2012).

The consumption of cruciferous vegetables is associated with a lower incidence of cancer. They contain isothiocyanates, membrane-permeable electrophilic compounds that form adducts with thiols. One major target of such a thiocarbamoylation is Cys-521 of HSP90AB1. This thiocarbamoylation leads to a dissociation of HSF-1 (heat shock factor 1) from HSP90AB1 and translocation of active HSP-1 to the nucleus, provoking a heat shock response (Shibata et al. 2011). However, the mechanistic link between thiocarbamoylation and suppression of cancer cell growth remains to be established.

Sometimes, mutations in the HSP90 gene might cause drug resistance. Selection of inhibitor-resistant mutants in yeast yielded resistant cell lines. For example, a HSP90AB1 I123T mutation was sufficient to confer inhibitor resistance and a corresponding HSP90AA1 I128T mutation yielded cell lines resistant to inhibitors of the Hsp90 ATPase (Zurawska et al. 2010).

Since tumor growth is supported by HSP proteins, inhibition of HSP is an obvious strategy for tumor treatment. Inhibition of HSP90 (with small-molecule NXD30001) suppressed proliferation of NF2 (neurofibromatosis type 2 gene)-deficient cells (Tanaka et al. 2013). Cancers with mutated KRAS cannot be treated with EGFR inhibitors, but they require the serine/threonine kinase STK33 for their viability and proliferation. Since STK33 is a target for the HSP90/CDC37 complex, HSP inhibitors might be a new treatment option for those tumors (Azoitei et al. 2012). Multiple ongoing clinical trials test the utility of HSP inhibition in a great variety of tumors (Den & Lu 2012; Jhaveri et al. 2012; Hong et al. 2013; Ramalingam et al. 2015).

Miscellaneous functions

In a study of 55 women with primary ovarian failure and 65 women with infertility, about 30 % had anti-ovarian antibodies (AOA) (Pires et al. 2007). None of the 60 control women had AOA. Using chromatography and mass spectrometry, the predominant antigen could be identified as HSP90AB1 (Pires & Khole 2009).

Concluding remarks

Since HSP90AB1 is involved in stabilization and transport of all types of proteins, it belongs to the proteins with most universal impact on biological functions. Because it is an ubiquitous protein, its importance for these various processes is often not enough estimated. Its overexpression in cancers together with its stabilizing function of proteins that promote essential functions of tumor cells make HSP inhibition a promising strategy for tumor treatment. Many clinical trials are under way. In addition, modulation of chaperone or co-chaperone activity might benefit inflammatory diseases, like rheumatoid arthritis, diseases caused by protein misfolding, like cystic fibrosis, or caused by increased protein aggregation like Alzheimer’s disease.

Highlights

  • This is a review on HSP90AB1 (heat shock protein 90 kDA alpha, class B, member 1), also known as HSP90beta.

  • HSP90AB1 binds to all kinds of proteins including steroid hormone receptors, transcription factors, ubiqitin ligases and kinases.

  • HSP90AB1 is a chaperone that stabilizes proteins in a properly folded structure, but is also involved in protein transport and protein degradation.

  • Inhibition of HSP90AB1 function is a promising therapeutic approach for a variety of diseases including cancer, inflammatory diseases and inherited diseases.

Acknowledgments

This review and the corresponding Gene Wiki article are written as part of the Gene Wiki Review series–a series resulting from a collaboration between the journal GENE and the Gene Wiki Initiative. The Gene Wiki Initiative is supported by National Institutes of Health (GM089820). Additional support for Gene Wiki Reviews is provided by Elsevier, the publisher of GENE. The corresponding Gene Wiki entry for this review can be found here: https://en.wikipedia.org/wiki/HSP90AB1.

Abbreviation

17-DMAG17-dimethylaminoethylamino-17-demethoxy-geldanamycin
ADAlzheimers disease
Agoargonaute protein
AHAactivator of heat shock 90kDa protein ATPase homolog
AHRarylhydrocarbon receptor
AIFapoptosis-inducing factor
AOAanti-ovarian antibodies
ARandrogen receptor
Arp2/3actin related/released protein 2/3
ASCIacute spinal cord injury
B[a]PBenzo[a]pyrene
BAZFBCL-6 associated zinc finger protein
BclB-cell lymphoma protein (influences apoptosis)
BPDbronchopulmonary dysplasia
Brg1brm (Brahma)/SWI2-related gene
BTKbreast tumor kinase
CA-SPcapsid-spacer protein
CCTchaperonin containing T-complex polypeptide
CDcharged domain
Cdc37cell division cycle 37
CDKcyclin dependent kinase
CFcystic fibrosis
CHIPC-terminus of HSc70-interacting protein (same as STUB1)
CHORDcysteine- and histidine-rich domain
CHORDC1cysteine- and histidine-rich domain containing protein 1
CKIIcasein kinase II
cPGEScytoplasmic terminal prostaglandin E2 synthase
CPR77th cyclophilin homolog, interacting with the yeast global transcriptional regulator Rpd3
CTDC-terminal domain
Cyp-40cyclophilin 40
DNA-PKDNA damage-activated protein kinase
DUBdeubiqitinating enzyme
EGFRepidermal growth factor receptor
ERestrogen receptor
ERADendoplasmic reticulum associated degradation
ERAFER-associated folding
EREestrogen response elements
ERKextracellular signal regulated kinase
ES cellsembryonic stem cells
FGFRfibroblast growth factor receptor
FKBPFK506-binding protein
G-actinglobular actin
GCaseglucocerebrosidase
GDNFglial cell line-derived neurotrophic factor
GISTgastrointestinal stromal tumor
GREglucocorticoid-response element
GRKG-protein coupled receptor protein kinase
GroELlarge protein of the GroE operon in E. coli, mutations of which affect the growth of lambda phage by interfering with assembly of its head protein E
GroESsmall protein of the GroE operon in E. coli, mutations of which affect the growth of lambda phage by interfering with assembly of its head protein E
HBxHepatitis B virus encoded X protein
HDAChistone deacetylase
HDACihistone deacetylase inhibitor (HDI)
HDIhistone deacetylase inhibitor
HGNChuman genome organization gene nomenclature committee
HIFhypoxia inducible factor
HOPHSP organizing protein
HSEheat shock element
HSF1heat shock factor 1
HSPheat shock protein
HSP90AB1heat shock protein 90 kDA alpha, class B, member 1 (also known as HSP90beta)
hTERThuman telomerase reverse transcriptase subunit
hTRhuman telomerase RNA template
HUGOhuman genome organization
IECintestinal epithelial cells
IkappaBinhibitor of NF-kappaB
IKBKIkappaB kinase
IKKIkappaB kinase
ILKintegrin linked kinase
IPFinterstitial pulmonary fibrosis
IPO4Importin-4
J-domainDNAJ homology domain
KCNQ4potassium channel, voltage gated KQT-like subfamily Q, member 4
KLF4Krüppel-Like-Factor 4
KMThistone H3 lysine (K) 36-specific methyltransferase
Kpna5karyopherin alpha 5
LAPlatency-associated peptide
LIMKLIM kinase
MAPKmitogen activated protein kinase
MBmethylene blue
MCL1myeloid cell leukemia protein 1
MCTDmixed connective tissue disease
MDmiddle domain
MDM2mouse double minute 2 homolog
MEEVDMet-Glu-Glu-Val-Asp motif
MKRN1makorin ring finger protein 1
MMPmatrix metalloproteinases
mTORmammalian target of rapamycin
MTORC1mammalian TOR complex 1
NF2neurofibromatosis type 2
NF-kappaBnuclear factor kappaB
NGFnerve growth factor
NLSnuclear localization sequence
NOnitric oxide
NSCLCnon-small cell lung cancer
NTDN-terminal domain
N-WASPneuronal Wiskott-Aldrich Syndrome Protein, now called WASPL
p23protein with ~23 kDa (same as PTGES3)
p70S670 kDa ribosomal subunit S6 kinase kinase
PDParkinson’s disease
PGE2prostaglandin E2
PKAprotein kinase A
PP5protein phosphatase 5, also abbreviated PPP5C
PPIasepeptidyl-prolyl cis-trans isomerase
PRprogesterone receptor
PTGES3prostaglandin E synthase 3 (same as p23
PTKprotein tyrosine kinase
PTK6protein tyrosine kinase6 (also called BTK)
RACreceptor-associated co-activator
Rackreceptor of activated protein kinase C
RA-ILDrheumatoid arthritis-associated interstitial lung disease
RBCK1RanBP-type and C3HC4-type zinc finger-containing protein 1
REV-1reversionless 1
RISCRNA-induced silencing complex
RNAiRNA interference
RPEretinal pigment epithelium
SGTsmall glutamine-rich TPR-containing protein
siRNAsmall interfering RNA
SScsystemic sclerosis
STUB1STIP1 homology and U-box containing protein 1 (same as CHIP)
TIMtranslocation machinery (translocase) of the inner mitochondrial membrane
TLStranslesion DNA synthesis
TMAtissue microarray
TNFalphatumor necrosis factor alpha
TOMtranslocation machinery (translocase) of the outer mitochondrial membrane
TOP5′ terminal oligopyrimidine
TPRtetratricopeptide repeat
TriCTCP-1 ring complex protein
TRIMT-cell receptor interacting molecule
Trxthioredoxin-1
VPAvalproic acid
WASPLWiskott-Aldrich Syndrome Protein-like

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • Acunzo J, Katsogiannou M, Rocchi P. Small heat shock proteins HSP27 (HspB1), alphaB-crystallin (HspB5) and HSP22 (HspB8) as regulators of cell death. Int J Biochem Cell Biol. 2012;44:1622–31. [Abstract] [Google Scholar]
  • Agrawal N, Dasaradhi PV, Mohmmed A, Malhotra P, Bhatnagar RK, Mukherjee SK. RNA interference: biology, mechanism, and applications. Microbiol Mol Biol Rev. 2003;67:657–85. [Europe PMC free article] [Abstract] [Google Scholar]
  • Akalin A, Elmore LW, Forsythe HL, Amaker BA, McCollum ED, Nelson PS, Ware JL, Holt SE. A novel mechanism for chaperone-mediated telomerase regulation during prostate cancer progression. Cancer Res. 2001;61:4791–6. [Abstract] [Google Scholar]
  • Alvarado CV, Rubio MF, Fernandez Larrosa PN, Panelo LC, Azurmendi PJ, Ruiz Grecco M, Martinez-Noel GA, Costas MA. The levels of RAC3 expression are up regulated by TNF in the inflammatory response. FEBS Open Bio. 2014;4:450–7. [Europe PMC free article] [Abstract] [Google Scholar]
  • Aressy B, Jullien D, Cazales M, Marcellin M, Bugler B, Burlet-Schiltz O, Ducommun B. A screen for deubiquitinating enzymes involved in the G(2)/M checkpoint identifies USP50 as a regulator of HSP90-dependent Wee1 stability. Cell Cycle. 2010;9:3815–22. [Abstract] [Google Scholar]
  • Azoitei N, Hoffmann CM, Ellegast JM, Ball CR, Obermayer K, Gossele U, Koch B, Faber K, Genze F, Schrader M, et al. Targeting of KRAS mutant tumors by HSP90 inhibitors involves degradation of STK33. J Exp Med. 2012;209:697–711. [Europe PMC free article] [Abstract] [Google Scholar]
  • Ballinger CA, Connell P, Wu Y, Hu Z, Thompson LJ, Yin LY, Patterson C. Identification of CHIP, a novel tetratricopeptide repeat-containing protein that interacts with heat shock proteins and negatively regulates chaperone functions. Mol Cell Biol. 1999;19:4535–45. [Europe PMC free article] [Abstract] [Google Scholar]
  • Bandyopadhyay S, Chiang CY, Srivastava J, Gersten M, White S, Bell R, Kurschner C, Martin CH, Smoot M, Sahasrabudhe S, et al. A human MAP kinase interactome. Nat Methods. 2010;7:801–5. [Europe PMC free article] [Abstract] [Google Scholar]
  • Barabutis N, Handa V, Dimitropoulou C, Rafikov R, Snead C, Kumar S, Joshi A, Thangjam G, Fulton D, Black SM, et al. LPS induces pp60c-src-mediated tyrosine phosphorylation of Hsp90 in lung vascular endothelial cells and mouse lung. Am J Physiol Lung Cell Mol Physiol. 2013;304:L883–93. [Europe PMC free article] [Abstract] [Google Scholar]
  • Beck R, Dejeans N, Glorieux C, Creton M, Delaive E, Dieu M, Raes M, Leveque P, Gallez B, Depuydt M, et al. Hsp90 is cleaved by reactive oxygen species at a highly conserved N-terminal amino acid motif. PLoS One. 2012;7:e40795. [Europe PMC free article] [Abstract] [Google Scholar]
  • Biaoxue R, Xiling J, Shuanying Y, Wei Z, Xiguang C, Jinsui W, Min Z. Upregulation of Hsp90-beta and annexin A1 correlates with poor survival and lymphatic metastasis in lung cancer patients. J Exp Clin Cancer Res. 2012;31:70. [Europe PMC free article] [Abstract] [Google Scholar]
  • Boltze C, Lehnert H, Schneider-Stock R, Peters B, Hoang-Vu C, Roessner A. HSP90 is a key for telomerase activation and malignant transition in pheochromocytoma. Endocrine. 2003;22:193–201. [Abstract] [Google Scholar]
  • Bose S, Weikl T, Bugl H, Buchner J. Chaperone function of Hsp90-associated proteins. Science. 1996;274:1715–7. [Abstract] [Google Scholar]
  • Brinker A, Scheufler C, Von Der Mulbe F, Fleckenstein B, Herrmann C, Jung G, Moarefi I, Hartl FU. Ligand discrimination by TPR domains. Relevance and selectivity of EEVD-recognition in Hsp70 × Hop × Hsp90 complexes. J Biol Chem. 2002;277:19265–75. [Abstract] [Google Scholar]
  • Brix J, Dietmeier K, Pfanner N. Differential recognition of preproteins by the purified cytosolic domains of the mitochondrial import receptors Tom20, Tom22, and Tom70. J Biol Chem. 1997;272:20730–5. [Abstract] [Google Scholar]
  • 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. 2004;23:5378–86. [Abstract] [Google Scholar]
  • Bukau B, Horwich AL. The Hsp70 and Hsp60 chaperone machines. Cell. 1998;92:351–66. [Abstract] [Google Scholar]
  • Calamia V, de Andres MC, Oreiro N, Ruiz-Romero C, Blanco FJ. Hsp90beta inhibition modulates nitric oxide production and nitric oxide-induced apoptosis in human chondrocytes. BMC Musculoskelet Disord. 2011;12:237. [Europe PMC free article] [Abstract] [Google Scholar]
  • Calderwood SK, Khaleque MA, Sawyer DB, Ciocca DR. Heat shock proteins in cancer: chaperones of tumorigenesis. Trends Biochem Sci. 2006;31:164–72. [Abstract] [Google Scholar]
  • Carrello A, Ingley E, Minchin RF, Tsai S, Ratajczak T. The common tetratricopeptide repeat acceptor site for steroid receptor-associated immunophilins and hop is located in the dimerization domain of Hsp90. J Biol Chem. 1999;274:2682–9. [Abstract] [Google Scholar]
  • Carrello A, Allan RK, Morgan SL, Owen BA, Mok D, Ward BK, Minchin RF, Toft DO, Ratajczak T. Interaction of the Hsp90 cochaperone cyclophilin 40 with Hsc70. Cell Stress Chaperones. 2004;9:167–81. [Europe PMC free article] [Abstract] [Google Scholar]
  • Catelli MG, Binart N, Jung-Testas I, Renoir JM, Baulieu EE, Feramisco JR, Welch WJ. The common 90-kd protein component of non-transformed ‘8S’ steroid receptors is a heat-shock protein. EMBO J. 1985;4:3131–5. [Europe PMC free article] [Abstract] [Google Scholar]
  • Cha B, Lim JW, Kim KH, Kim H. HSP90beta interacts with Rac1 to activate NADPH oxidase in Helicobacter pylori-infected gastric epithelial cells. Int J Biochem Cell Biol. 2010;42:1455–61. [Abstract] [Google Scholar]
  • Chadli A, Felts SJ, Wang Q, Sullivan WP, Botuyan MV, Fauq A, Ramirez-Alvarado M, Mer G. Celastrol inhibits Hsp90 chaperoning of steroid receptors by inducing fibrillization of the Co-chaperone p23. J Biol Chem. 2010;285:4224–31. [Europe PMC free article] [Abstract] [Google Scholar]
  • Charoensook R, Gatphayak K, Sharifi AR, Chaisongkram C, Brenig B, Knorr C. Polymorphisms in the bovine HSP90AB1 gene are associated with heat tolerance in Thai indigenous cattle. Trop Anim Health Prod. 2012;44:921–8. [Europe PMC free article] [Abstract] [Google Scholar]
  • Chen B, Zhong D, Monteiro A. Comparative genomics and evolution of the HSP90 family of genes across all kingdoms of organisms. BMC Genomics. 2006a;7:156. [Europe PMC free article] [Abstract] [Google Scholar]
  • Chen B, Piel WH, Gui L, Bruford E, Monteiro A. The HSP90 family of genes in the human genome: insights into their divergence and evolution. Genomics. 2005;86:627–37. [Abstract] [Google Scholar]
  • Chen G, Cao P, Goeddel DV. TNF-induced recruitment and activation of the IKK complex require Cdc37 and Hsp90. Mol Cell. 2002;9:401–10. [Abstract] [Google Scholar]
  • Chen H, Wu Y, Zhang Y, Jin L, Luo L, Xue B, Lu C, Zhang X, Yin Z. Hsp70 inhibits lipopolysaccharide-induced NF-kappaB activation by interacting with TRAF6 and inhibiting its ubiquitination. FEBS Lett. 2006b;580:3145–52. [Abstract] [Google Scholar]
  • Chen MS, Silverstein AM, Pratt WB, Chinkers M. The tetratricopeptide repeat domain of protein phosphatase 5 mediates binding to glucocorticoid receptor heterocomplexes and acts as a dominant negative mutant. J Biol Chem. 1996;271:32315–20. [Abstract] [Google Scholar]
  • Chen S, Smith DF. Hop as an adaptor in the heat shock protein 70 (Hsp70) and hsp90 chaperone machinery. J Biol Chem. 1998;273:35194–200. [Abstract] [Google Scholar]
  • Cheng MB, Zhang Y, Zhong X, Sutter B, Cao CY, Chen XS, Cheng XK, Xiao L, Shen YF. Stat1 mediates an auto-regulation of hsp90beta gene in heat shock response. Cell Signal. 2010;22:1206–13. [Abstract] [Google Scholar]
  • Cheng Q, Chang JT, Geradts J, Neckers LM, Haystead T, Spector NL, Lyerly HK. Amplification and high-level expression of heat shock protein 90 marks aggressive phenotypes of human epidermal growth factor receptor 2 negative breast cancer. Breast Cancer Res. 2012;14:R62. [Europe PMC free article] [Abstract] [Google Scholar]
  • Chinkers M. Protein phosphatase 5 in signal transduction. Trends Endocrinol Metab. 2001;12:28–32. [Abstract] [Google Scholar]
  • Ciocca DR, Arrigo AP, Calderwood SK. Heat shock proteins and heat shock factor 1 in carcinogenesis and tumor development: an update. Arch Toxicol. 2013;87:19–48. [Europe PMC free article] [Abstract] [Google Scholar]
  • Cioffi DL, Hubler TR, Scammell JG. Organization and function of the FKBP52 and FKBP51 genes. Curr Opin Pharmacol. 2011;11:308–13. [Europe PMC free article] [Abstract] [Google Scholar]
  • Colo GP, Rubio MF, Nojek IM, Werbajh SE, Echeverria PC, Alvarado CV, Nahmod VE, Galigniana MD, Costas MA. The p160 nuclear receptor co-activator RAC3 exerts an anti-apoptotic role through a cytoplasmatic action. Oncogene. 2008;27:2430–44. [Abstract] [Google Scholar]
  • Comandini A, Naro C, Adamo R, Akbar AN, Lanna A, Bonmassar E, Franzese O. Molecular mechanisms involved in HIV-1-Tat mediated inhibition of telomerase activity in human CD4(+) T lymphocytes. Mol Immunol. 2013;54:181–92. [Abstract] [Google Scholar]
  • Coskunpinar E, Akkaya N, Yildiz P, Oltulu YM, Aynaci E, Isbir T, Yaylim I. The significance of HSP90AA1, HSP90AB1 and HSP90B1 gene polymorphisms in a Turkish population with non-small cell lung cancer. Anticancer Res. 2014;34:753–7. [Abstract] [Google Scholar]
  • 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. 1998;79:129–68. [Abstract] [Google Scholar]
  • Cyr DM, Ramos CH. Specification of Hsp70 function by Type I and Type II Hsp40. Subcell Biochem. 2015;78:91–102. [Abstract] [Google Scholar]
  • Czar MJ, Galigniana MD, Silverstein AM, Pratt WB. Geldanamycin, a heat shock protein 90-binding benzoquinone ansamycin, inhibits steroid-dependent translocation of the glucocorticoid receptor from the cytoplasm to the nucleus. Biochemistry. 1997;36:7776–85. [Abstract] [Google Scholar]
  • Czar MJ, Lyons RH, Welsh MJ, Renoir JM, Pratt WB. Evidence that the FK506-binding immunophilin heat shock protein 56 is required for trafficking of the glucocorticoid receptor from the cytoplasm to the nucleus. Mol Endocrinol. 1995;9:1549–60. [Abstract] [Google Scholar]
  • Dai RM, Li CC. Valosin-containing protein is a multi-ubiquitin chain-targeting factor required in ubiquitin-proteasome degradation. Nat Cell Biol. 2001;3:740–4. [Abstract] [Google Scholar]
  • Davies TH, Sanchez ER. FKBP52. Int J Biochem Cell Biol. 2005;37:42–7. [Abstract] [Google Scholar]
  • Davies TH, Ning YM, Sanchez ER. A new first step in activation of steroid receptors: hormone-induced switching of FKBP51 and FKBP52 immunophilins. J Biol Chem. 2002;277:4597–600. [Abstract] [Google Scholar]
  • Davies TH, Ning YM, Sanchez ER. Differential control of glucocorticoid receptor hormone-binding function by tetratricopeptide repeat (TPR) proteins and the immunosuppressive ligand FK506. Biochemistry. 2005;44:2030–8. [Abstract] [Google Scholar]
  • Den RB, Lu B. Heat shock protein 90 inhibition: rationale and clinical potential. Ther Adv Med Oncol. 2012;4:211–8. [Europe PMC free article] [Abstract] [Google Scholar]
  • Dickey CA, Kamal A, Lundgren K, Klosak N, Bailey RM, Dunmore J, Ash P, Shoraka S, Zlatkovic J, Eckman CB, et al. The high-affinity HSP90-CHIP complex recognizes and selectively degrades phosphorylated tau client proteins. J Clin Invest. 2007;117:648–58. [Abstract] [Google Scholar]
  • Ding QH, Cheng Y, Chen WP, Zhong HM, Wang XH. Celastrol, an inhibitor of heat shock protein 90beta potently suppresses the expression of matrix metalloproteinases, inducible nitric oxide synthase and cyclooxygenase-2 in primary human osteoarthritic chondrocytes. Eur J Pharmacol. 2013;708:1–7. [Abstract] [Google Scholar]
  • Donley C, McClelland K, McKeen HD, Nelson L, Yakkundi A, Jithesh PV, Burrows J, McClements L, Valentine A, Prise KM, et al. Identification of RBCK1 as a novel regulator of FKBPL: implications for tumor growth and response to tamoxifen. Oncogene. 2014;33:3441–50. [Abstract] [Google Scholar]
  • Donlin LT, Andresen C, Just S, Rudensky E, Pappas CT, Kruger M, Jacobs EY, Unger A, Zieseniss A, Dobenecker MW, et al. Smyd2 controls cytoplasmic lysine methylation of Hsp90 and myofilament organization. Genes Dev. 2012;26:114–9. [Europe PMC free article] [Abstract] [Google Scholar]
  • Duina AA, Marsh JA, Gaber RF. Identification of two CyP-40-like cyclophilins in Saccharomyces cerevisiae, one of which is required for normal growth. Yeast. 1996a;12:943–52. [Abstract] [Google Scholar]
  • Duina AA, Chang HC, Marsh JA, Lindquist S, Gaber RF. A cyclophilin function in Hsp90-dependent signal transduction. Science. 1996b;274:1713–5. [Abstract] [Google Scholar]
  • Echeverria PC, 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. 2011;6:e26044. [Europe PMC free article] [Abstract] [Google Scholar]
  • Echeverria PC, Mazaira G, Erlejman A, Gomez-Sanchez C, Piwien Pilipuk G, Galigniana MD. Nuclear import of the glucocorticoid receptor-hsp90 complex through the nuclear pore complex is mediated by its interaction with Nup62 and importin beta. Mol Cell Biol. 2009;29:4788–97. [Europe PMC free article] [Abstract] [Google Scholar]
  • Erazo T, Moreno A, Ruiz-Babot G, Rodriguez-Asiain A, Morrice NA, Espadamala J, Bayascas JR, Gomez N, Lizcano JM. Canonical and kinase activity-independent mechanisms for extracellular signal-regulated kinase 5 (ERK5) nuclear translocation require dissociation of Hsp90 from the ERK5-Cdc37 complex. Mol Cell Biol. 2013;33:1671–86. [Europe PMC free article] [Abstract] [Google Scholar]
  • Erlejman AG, Lagadari M, Toneatto J, Piwien-Pilipuk G, Galigniana MD. Regulatory role of the 90-kDa-heat-shock protein (Hsp90) and associated factors on gene expression. Biochim Biophys Acta. 2014a;1839:71–87. [Abstract] [Google Scholar]
  • Erlejman AG, De Leo SA, Mazaira GI, Molinari AM, Camisay MF, Fontana V, Cox MB, Piwien-Pilipuk G, Galigniana MD. NF-kappaB transcriptional activity is modulated by FK506-binding proteins FKBP51 and FKBP52: a role for peptidyl-prolyl isomerase activity. J Biol Chem. 2014b;289:26263–76. [Europe PMC free article] [Abstract] [Google Scholar]
  • Ernst JT, Neubert T, Liu M, Sperry S, Zuccola H, Turnbull A, Fleck B, Kargo W, Woody L, Chiang P, et al. Identification of novel HSP90alpha/beta isoform selective inhibitors using structure-based drug design. demonstration of potential utility in treating CNS disorders such as Huntington’s disease. J Med Chem. 2014;57:3382–400. [Abstract] [Google Scholar]
  • Falsone SF, Gesslbauer B, Tirk F, Piccinini AM, Kungl AJ. A proteomic snapshot of the human heat shock protein 90 interactome. FEBS Lett. 2005;579:6350–4. [Abstract] [Google Scholar]
  • Fan CY, Lee S, Cyr DM. Mechanisms for regulation of Hsp70 function by Hsp40. Cell Stress Chaperones. 2003;8:309–16. [Europe PMC free article] [Abstract] [Google Scholar]
  • Fayet O, Ziegelhoffer T, Georgopoulos C. The groES and groEL heat shock gene products of Escherichia coli are essential for bacterial growth at all temperatures. J Bacteriol. 1989;171:1379–85. [Europe PMC free article] [Abstract] [Google Scholar]
  • Floen MJ, Forred BJ, Bloom EJ, Vitiello PF. Thioredoxin-1 redox signaling regulates cell survival in response to hyperoxia. Free Radic Biol Med. 2014;75:167–77. [Europe PMC free article] [Abstract] [Google Scholar]
  • Forafonov F, Toogun OA, Grad I, Suslova E, Freeman BC, Picard D. p23/Sba1p protects against Hsp90 inhibitors independently of its intrinsic chaperone activity. Mol Cell Biol. 2008;28:3446–56. [Europe PMC free article] [Abstract] [Google Scholar]
  • Fornace A, Jr, Alamo I, Jr, Hollander MC. DNA damage-inducible transcripts in mammalian cells. Proc Natl Acad Sci U S A. 1988;85:8800–4. [Europe PMC free article] [Abstract] [Google Scholar]
  • Forsythe HL, Jarvis JL, Turner JW, Elmore LW, Holt SE. Stable association of hsp90 and p23, but Not hsp70, with active human telomerase. J Biol Chem. 2001;276:15571–4. [Abstract] [Google Scholar]
  • Forthun RB, Sengupta T, Skjeldam HK, Lindvall JM, McCormack E, Gjertsen BT, Nilsen H. Cross-species functional genomic analysis identifies resistance genes of the histone deacetylase inhibitor valproic acid. PLoS One. 2012;7:e48992. [Europe PMC free article] [Abstract] [Google Scholar]
  • Freeman BC, Felts SJ, Toft DO, Yamamoto KR. The p23 molecular chaperones act at a late step in intracellular receptor action to differentially affect ligand efficacies. Genes Dev. 2000;14:422–34. [Europe PMC free article] [Abstract] [Google Scholar]
  • Fromm-Dornieden C, von der Heyde S, Lytovchenko O, Salinas-Riester G, Brenig B, Beissbarth T, Baumgartner BG. Novel polysome messages and changes in translational activity appear after induction of adipogenesis in 3T3-L1 cells. BMC Mol Biol. 2012;13:9. [Europe PMC free article] [Abstract] [Google Scholar]
  • Fukumoto R, Kiang JG. Geldanamycin analog 17-DMAG limits apoptosis in human peripheral blood cells by inhibition of p53 activation and its interaction with heat-shock protein 90 kDa after exposure to ionizing radiation. Radiat Res. 2011;176:333–45. [Europe PMC free article] [Abstract] [Google Scholar]
  • Fusco D, Vargiolu M, Vidone M, Mariani E, Pennisi LF, Bonora E, Capellari S, Dirnberger D, Baumeister R, Martinelli P, et al. The RET51/FKBP52 complex and its involvement in Parkinson disease. Hum Mol Genet. 2010;19:2804–16. [Abstract] [Google Scholar]
  • Galigniana MD. Steroid receptor coupling becomes nuclear. Chem Biol. 2012;19:662–3. [Abstract] [Google Scholar]
  • Galigniana MD, Echeverria PC, Erlejman AG, 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. 2010a;1:299–308. [Europe PMC free article] [Abstract] [Google Scholar]
  • Galigniana MD, Harrell JM, O’Hagen HM, Ljungman M, Pratt WB. Hsp90-binding immunophilins link p53 to dynein during p53 transport to the nucleus. J Biol Chem. 2004;279:22483–9. [Abstract] [Google Scholar]
  • Galigniana MD, Erlejman AG, Monte M, Gomez-Sanchez C, Piwien-Pilipuk G. The hsp90-FKBP52 complex links the mineralocorticoid receptor to motor proteins and persists bound to the receptor in early nuclear events. Mol Cell Biol. 2010b;30:1285–98. [Europe PMC free article] [Abstract] [Google Scholar]
  • Galigniana MD, Harrell JM, Murphy PJ, Chinkers M, Radanyi C, Renoir JM, Zhang M, Pratt WB. Binding of hsp90-associated immunophilins to cytoplasmic dynein: direct binding and in vivo evidence that the peptidylprolyl isomerase domain is a dynein interaction domain. Biochemistry. 2002;41:13602–10. [Abstract] [Google Scholar]
  • Gallo LI, Ghini AA, Piwien Pilipuk G, Galigniana MD. Differential recruitment of tetratricorpeptide repeat domain immunophilins to the mineralocorticoid receptor influences both heat-shock protein 90-dependent retrotransport and hormone-dependent transcriptional activity. Biochemistry. 2007;46:14044–57. [Abstract] [Google Scholar]
  • Gallo LI, Lagadari M, Piwien-Pilipuk G, Galigniana MD. The 90-kDa heat-shock protein (Hsp90)-binding immunophilin FKBP51 is a mitochondrial protein that translocates to the nucleus to protect cells against oxidative stress. J Biol Chem. 2011;286:30152–60. [Europe PMC free article] [Abstract] [Google Scholar]
  • Gano JJ, Simon JA. A proteomic investigation of ligand-dependent HSP90 complexes reveals CHORDC1 as a novel ADP-dependent HSP90-interacting protein. Mol Cell Proteomics. 2010;9:255–70. [Europe PMC free article] [Abstract] [Google Scholar]
  • Gao Y, Yechikov S, Vazquez AE, Chen D, Nie L. Distinct roles of molecular chaperones HSP90alpha and HSP90beta in the biogenesis of KCNQ4 channels. PLoS One. 2013;8:e57282. [Europe PMC free article] [Abstract] [Google Scholar]
  • Garrido C, Schmitt E, Cande C, Vahsen N, Parcellier A, Kroemer G. HSP27 and HSP70: potentially oncogenic apoptosis inhibitors. Cell Cycle. 2003;2:579–84. [Abstract] [Google Scholar]
  • Georgopoulos CP, Hendrix RW, Casjens SR, Kaiser AD. Host participation in bacteriophage lambda head assembly. J Mol Biol. 1973;76:45–60. [Abstract] [Google Scholar]
  • Gerard M, Deleersnijder A, Daniels V, Schreurs S, Munck S, Reumers V, Pottel H, Engelborghs Y, Van den Haute C, Taymans JM, et al. Inhibition of FK506 binding proteins reduces alpha-synuclein aggregation and Parkinson’s disease-like pathology. J Neurosci. 2010;30:2454–63. [Abstract] [Google Scholar]
  • Giessrigl B, Krieger S, Rosner M, Huttary N, Saiko P, Alami M, Messaoudi S, Peyrat JF, Maciuk A, Gollinger M, et al. Hsp90 stabilizes Cdc25A and counteracts heat shock-mediated Cdc25A degradation and cell-cycle attenuation in pancreatic carcinoma cells. Hum Mol Genet. 2012;21:4615–27. [Abstract] [Google Scholar]
  • Giustiniani J, Daire V, Cantaloube I, Durand G, Pous C, Perdiz D, Baillet A. Tubulin acetylation favors Hsp90 recruitment to microtubules and stimulates the signaling function of the Hsp90 clients Akt/PKB and p53. Cell Signal. 2009;21:529–39. [Abstract] [Google Scholar]
  • Golden T, Aragon IV, Rutland B, Tucker JA, Shevde LA, Samant RS, Zhou G, Amable L, Skarra D, Honkanen RE. Elevated levels of Ser/Thr protein phosphatase 5 (PP5) in human breast cancer. Biochim Biophys Acta. 2008;1782:259–70. [Europe PMC free article] [Abstract] [Google Scholar]
  • Gong Y, Kakihara Y, Krogan N, Greenblatt J, Emili A, Zhang Z, Houry WA. An atlas of chaperone-protein interactions in Saccharomyces cerevisiae: implications to protein folding pathways in the cell. Mol Syst Biol. 2009;5:275. [Europe PMC free article] [Abstract] [Google Scholar]
  • Grabek KR, Karimpour-Fard A, Epperson LE, Hindle A, Hunter LE, Martin SL. Multistate proteomics analysis reveals novel strategies used by a hibernator to precondition the heart and conserve ATP for winter heterothermy. Physiol Genomics. 2011;43:1263–75. [Europe PMC free article] [Abstract] [Google Scholar]
  • Grenert JP, Sullivan WP, Fadden P, Haystead TA, Clark J, Mimnaugh E, Krutzsch H, Ochel HJ, Schulte TW, Sausville E, et al. The amino-terminal domain of heat shock protein 90 (hsp90) that binds geldanamycin is an ATP/ADP switch domain that regulates hsp90 conformation. J Biol Chem. 1997;272:23843–50. [Abstract] [Google Scholar]
  • Grossmann C, Ruhs S, Langenbruch L, Mildenberger S, Stratz N, Schumann K, Gekle M. Nuclear shuttling precedes dimerization in mineralocorticoid receptor signaling. Chem Biol. 2012;19:742–51. [Abstract] [Google Scholar]
  • Gusarova V, Caplan AJ, Brodsky JL, Fisher EA. Apoprotein B degradation is promoted by the molecular chaperones hsp90 and hsp70. J Biol Chem. 2001;276:24891–900. [Abstract] [Google Scholar]
  • Haase MG, Geyer P, Fitze G, Baretton GB. Down-regulation of heat shock protein HSP90ab1 in radiation-damaged lung cells other than mast cells. J Histochem Cytochem. 2014;62:355–368. [Europe PMC free article] [Abstract] [Google Scholar]
  • Hahn JS. Regulation of Nod1 by Hsp90 chaperone complex. FEBS Lett. 2005;579:4513–9. [Abstract] [Google Scholar]
  • Hainzl O, Lapina MC, Buchner J, Richter K. The charged linker region is an important regulator of Hsp90 function. J Biol Chem. 2009;284:22559–67. [Europe PMC free article] [Abstract] [Google Scholar]
  • Hamamoto R, Toyokawa G, Nakakido M, Ueda K, Nakamura Y. SMYD2-dependent HSP90 methylation promotes cancer cell proliferation by regulating the chaperone complex formation. Cancer Lett. 2014;351:126–33. [Abstract] [Google Scholar]
  • Harlow L, Rosas IO, Gochuico BR, Mikuls TR, Dellaripa PF, Oddis CV, Ascherman DP. Identification of citrullinated hsp90 isoforms as novel autoantigens in rheumatoid arthritis-associated interstitial lung disease. Arthritis Rheum. 2013;65:869–79. [Abstract] [Google Scholar]
  • Harrell JM, Kurek I, Breiman A, Radanyi C, Renoir JM, Pratt WB, Galigniana MD. All of the protein interactions that link steroid receptor.hsp90.immunophilin heterocomplexes to cytoplasmic dynein are common to plant and animal cells. Biochemistry. 2002;41:5581–7. [Abstract] [Google Scholar]
  • Harrell JM, Murphy PJ, Morishima Y, Chen H, Mansfield JF, Galigniana MD, Pratt WB. Evidence for glucocorticoid receptor transport on microtubules by dynein. J Biol Chem. 2004;279:54647–54. [Abstract] [Google Scholar]
  • Hartl FU. Molecular chaperones in cellular protein folding. Nature. 1996;381:571–9. [Abstract] [Google Scholar]
  • Hartson SD, Matts RL. Approaches for defining the Hsp90-dependent proteome. Biochim Biophys Acta. 2012;1823:656–67. [Europe PMC free article] [Abstract] [Google Scholar]
  • Hendrix RW. Purification and properties of groE, a host protein involved in bacteriophage assembly. J Mol Biol. 1979;129:375–92. [Abstract] [Google Scholar]
  • Hildenbrand ZL, Molugu SK, Herrera N, Ramirez C, Xiao C, Bernal RA. Hsp90 can accommodate the simultaneous binding of the FKBP52 and HOP proteins. Oncotarget. 2011;2:43–58. [Europe PMC free article] [Abstract] [Google Scholar]
  • Hinds TD, Jr, Sanchez ER. Protein phosphatase 5. Int J Biochem Cell Biol. 2008;40:2358–62. [Europe PMC free article] [Abstract] [Google Scholar]
  • Hinz M, Broemer M, Arslan SC, Otto A, Mueller EC, Dettmer R, Scheidereit C. Signal responsiveness of IkappaB kinases is determined by Cdc37-assisted transient interaction with Hsp90. J Biol Chem. 2007;282:32311–9. [Abstract] [Google Scholar]
  • Hoffmann T, Hovemann B. Heat-shock proteins, Hsp84 and Hsp86, of mice and men: two related genes encode formerly identified tumour-specific transplantation antigens. Gene. 1988;74:491–501. [Abstract] [Google Scholar]
  • Hong DS, Banerji U, Tavana B, George GC, Aaron J, Kurzrock R. Targeting the molecular chaperone heat shock protein 90 (HSP90): lessons learned and future directions. Cancer Treat Rev. 2013;39:375–87. [Abstract] [Google Scholar]
  • Horwich AL, Farr GW, Fenton WA. GroEL-GroES-mediated protein folding. Chem Rev. 2006;106:1917–30. [Abstract] [Google Scholar]
  • Hubler TR, Denny WB, Valentine DL, Cheung-Flynn J, Smith DF, Scammell JG. The FK506-binding immunophilin FKBP51 is transcriptionally regulated by progestin and attenuates progestin responsiveness. Endocrinology. 2003;144:2380–7. [Abstract] [Google Scholar]
  • Hunt CR, Dix DJ, Sharma GG, Pandita RK, Gupta A, Funk M, Pandita TK. Genomic instability and enhanced radiosensitivity in Hsp70.1- and Hsp70.3-deficient mice. Mol Cell Biol. 2004;24:899–911. [Europe PMC free article] [Abstract] [Google Scholar]
  • Hunter MC, O’Hagan KL, Kenyon A, Dhanani KC, Prinsloo E, Edkins AL. Hsp90 binds directly to fibronectin (FN) and inhibition reduces the extracellular fibronectin matrix in breast cancer cells. PLoS One. 2014;9:e86842. [Europe PMC free article] [Abstract] [Google Scholar]
  • Huntoon CJ, Nye MD, Geng L, Peterson KL, Flatten KS, Haluska P, Kaufmann SH, Karnitz LM. Heat shock protein 90 inhibition depletes LATS1 and LATS2, two regulators of the mammalian hippo tumor suppressor pathway. Cancer Res. 2010;70:8642–50. [Europe PMC free article] [Abstract] [Google Scholar]
  • Iki T, Yoshikawa M, Nishikiori M, Jaudal MC, Matsumoto-Yokoyama E, Mitsuhara I, Meshi T, Ishikawa M. In vitro assembly of plant RNA-induced silencing complexes facilitated by molecular chaperone HSP90. Mol Cell. 2010;39:282–91. [Abstract] [Google Scholar]
  • Iwasaki S, Kobayashi M, Yoda M, Sakaguchi Y, Katsuma S, Suzuki T, Tomari Y. Hsc70/Hsp90 chaperone machinery mediates ATP-dependent RISC loading of small RNA duplexes. Mol Cell. 2010;39:292–9. [Abstract] [Google Scholar]
  • Jahn M, Rehn A, Pelz B, Hellenkamp B, Richter K, Rief M, Buchner J, Hugel T. The charged linker of the molecular chaperone Hsp90 modulates domain contacts and biological function. Proc Natl Acad Sci U S A. 2014;111:17881–6. [Europe PMC free article] [Abstract] [Google Scholar]
  • Jarosz DF, Lindquist S. Hsp90 and Environmental Stress Transform the Adaptive Value of Natural Genetic Variation. Science. 2010;330:1820–1824. [Europe PMC free article] [Abstract] [Google Scholar]
  • Jarosz DF, Taipale M, Lindquist S. Protein homeostasis and the phenotypic manifestation of genetic diversity: principles and mechanisms. Annu Rev Genet. 2010;44:189–216. [Abstract] [Google Scholar]
  • Jhaveri K, Taldone T, Modi S, Chiosis G. Advances in the clinical development of heat shock protein 90 (Hsp90) inhibitors in cancers. Biochim Biophys Acta. 2012;1823:742–55. [Europe PMC free article] [Abstract] [Google Scholar]
  • Jinwal UK, Koren J, 3rd, Dickey CA. Reconstructing the Hsp90/Tau Machine. Curr Enzym Inhib. 2013;9:41–45. [Europe PMC free article] [Abstract] [Google Scholar]
  • Jinwal UK, Trotter JH, Abisambra JF, Koren J, 3rd, Lawson LY, Vestal GD, O’Leary JC, 3rd, Johnson AG, Jin Y, Jones JR, et al. The Hsp90 kinase co-chaperone Cdc37 regulates tau stability and phosphorylation dynamics. J Biol Chem. 2011;286:16976–83. [Europe PMC free article] [Abstract] [Google Scholar]
  • Jinwal UK, Koren J, 3rd, Borysov SI, Schmid AB, Abisambra JF, Blair LJ, Johnson AG, Jones JR, Shults CL, O’Leary JC, 3rd, et al. The Hsp90 cochaperone, FKBP51, increases Tau stability and polymerizes microtubules. J Neurosci. 2010;30:591–9. [Europe PMC free article] [Abstract] [Google Scholar]
  • Johnson JL, Toft DO. Binding of p23 and hsp90 during assembly with the progesterone receptor. Mol Endocrinol. 1995;9:670–8. [Abstract] [Google Scholar]
  • Johnson JL, Craig EA. Protein folding in vivo: unraveling complex pathways. Cell. 1997;90:201–4. [Abstract] [Google Scholar]
  • Johnson JL, Beito TG, Krco CJ, Toft DO. Characterization of a novel 23-kilodalton protein of unactive progesterone receptor complexes. Mol Cell Biol. 1994;14:1956–63. [Europe PMC free article] [Abstract] [Google Scholar]
  • Joshi P, Stoddart CA. Impaired infectivity of ritonavir-resistant HIV is rescued by heat shock protein 90AB1. J Biol Chem. 2011;286:24581–92. [Europe PMC free article] [Abstract] [Google Scholar]
  • Joshi P, Sloan B, Torbett BE, Stoddart CA. Heat shock protein 90AB1 and hyperthermia rescue infectivity of HIV with defective cores. Virology. 2013;436:162–72. [Europe PMC free article] [Abstract] [Google Scholar]
  • Kampinga HH, Hageman J, Vos MJ, Kubota H, Tanguay RM, Bruford EA, Cheetham ME, Chen B, Hightower LE. Guidelines for the nomenclature of the human heat shock proteins. Cell Stress Chaperones. 2009;14:105–11. [Europe PMC free article] [Abstract] [Google Scholar]
  • Kang SA, Cho HS, Yoon JB, Chung IK, Lee ST. Hsp90 rescues PTK6 from proteasomal degradation in breast cancer cells. Biochem J. 2012;447:313–20. [Abstract] [Google Scholar]
  • Karagoz GE, Duarte AM, Ippel H, Uetrecht C, Sinnige T, van Rosmalen M, Hausmann J, Heck AJ, Boelens R, Rudiger SG. N-terminal domain of human Hsp90 triggers binding to the cochaperone p23. Proc Natl Acad Sci U S A. 2011;108:580–5. [Europe PMC free article] [Abstract] [Google Scholar]
  • Karagoz GE, Duarte AM, Akoury E, Ippel H, Biernat J, Moran Luengo T, Radli M, Didenko T, Nordhues BA, Veprintsev DB, et al. Hsp90-Tau complex reveals molecular basis for specificity in chaperone action. Cell. 2014;156:963–74. [Europe PMC free article] [Abstract] [Google Scholar]
  • Karin M, Lin A. NF-kappaB at the crossroads of life and death. Nat Immunol. 2002;3:221–7. [Abstract] [Google Scholar]
  • Kasioulis I, Syred HM, Tate P, Finch A, Shaw J, Seawright A, Fuszard M, Botting CH, Shirran S, Adams IR, et al. Kdm3a lysine demethylase is an Hsp90 client required for cytoskeletal rearrangements during spermatogenesis. Mol Biol Cell. 2014;25:1216–33. [Europe PMC free article] [Abstract] [Google Scholar]
  • Kawauchi K, Ihjima K, Yamada O. IL-2 increases human telomerase reverse transcriptase activity transcriptionally and posttranslationally through phosphatidylinositol 3′-kinase/Akt, heat shock protein 90, and mammalian target of rapamycin in transformed NK cells. J Immunol. 2005;174:5261–9. [Abstract] [Google Scholar]
  • Kim JH, Park SM, Kang MR, Oh SY, Lee TH, Muller MT, Chung IK. Ubiquitin ligase MKRN1 modulates telomere length homeostasis through a proteolysis of hTERT. Genes Dev. 2005;19:776–81. [Europe PMC free article] [Abstract] [Google Scholar]
  • Kim RH, Kim R, Chen W, Hu S, Shin KH, Park NH, Kang MK. Association of hsp90 to the hTERT promoter is necessary for hTERT expression in human oral cancer cells. Carcinogenesis. 2008;29:2425–31. [Europe PMC free article] [Abstract] [Google Scholar]
  • Kim WY, Oh SH, Woo JK, Hong WK, Lee HY. Targeting heat shock protein 90 overrides the resistance of lung cancer cells by blocking radiation-induced stabilization of hypoxia-inducible factor-1alpha. Cancer Res. 2009;69:1624–32. [Europe PMC free article] [Abstract] [Google Scholar]
  • King FW, Wawrzynow A, Hohfeld J, Zylicz M. Co-chaperones Bag-1, Hop and Hsp40 regulate Hsc70 and Hsp90 interactions with wild-type or mutant p53. EMBO J. 2001;20:6297–305. [Europe PMC free article] [Abstract] [Google Scholar]
  • Kochhar A, Kopelovich L, Sue E, Guttenplan JB, Herbert BS, Dannenberg AJ, Subbaramaiah K. p53 modulates Hsp90 ATPase activity and regulates aryl hydrocarbon receptor signaling. Cancer Prev Res (Phila) 2014;7:596–606. [Europe PMC free article] [Abstract] [Google Scholar] Retracted
  • Koshimizu TA, Tsuchiya H, Tsuda H, Fujiwara Y, Shibata K, Hirasawa A, Tsujimoto G, Fujimura A. Inhibition of heat shock protein 90 attenuates adenylate cyclase sensitization after chronic morphine treatment. Biochem Biophys Res Commun. 2010;392:603–7. [Abstract] [Google Scholar]
  • Koyasu S, Nishida E, Kadowaki T, Matsuzaki F, Iida K, Harada F, Kasuga M, Sakai H, Yahara I. Two mammalian heat shock proteins, HSP90 and HSP100, are actin-binding proteins. Proc Natl Acad Sci U S A. 1986;83:8054–8. [Europe PMC free article] [Abstract] [Google Scholar]
  • Krishna KA, Rao GV, Rao KR. Chaperonin GroEL: structure and reaction cycle. Curr Protein Pept Sci. 2007;8:418–25. [Abstract] [Google Scholar]
  • Kubo N, Wu D, Yoshihara Y, Sang M, Nakagawara A, Ozaki T. Co-chaperon DnaJC7/TPR2 enhances p53 stability and activity through blocking the complex formation between p53 and MDM2. Biochem Biophys Res Commun. 2013;430:1034–9. [Abstract] [Google Scholar]
  • Laederich MB, Degnin CR, Lunstrum GP, Holden P, Horton WA. Fibroblast growth factor receptor 3 (FGFR3) is a strong heat shock protein 90 (Hsp90) client: implications for therapeutic manipulation. J Biol Chem. 2011;286:19597–604. [Europe PMC free article] [Abstract] [Google Scholar]
  • Lambert JP, Ivosev G, Couzens AL, Larsen B, Taipale M, Lin ZY, Zhong Q, Lindquist S, Vidal M, Aebersold R, et al. Mapping differential interactomes by affinity purification coupled with data-independent mass spectrometry acquisition. Nat Methods. 2013;10:1239–45. [Europe PMC free article] [Abstract] [Google Scholar]
  • Lamphere L, Fiore F, Xu X, Brizuela L, Keezer S, Sardet C, Draetta GF, Gyuris J. Interaction between Cdc37 and Cdk4 in human cells. Oncogene. 1997;14:1999–2004. [Abstract] [Google Scholar]
  • Lawrence CW, Christensen R. UV mutagenesis in radiation-sensitive strains of yeast. Genetics. 1976;82:207–32. [Europe PMC free article] [Abstract] [Google Scholar]
  • Lees-Miller SP, Anderson CW. 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. 1989a;264:17275–80. [Abstract] [Google Scholar]
  • Lees-Miller SP, Anderson CW. 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. 1989b;264:2431–7. [Abstract] [Google Scholar]
  • Li J, Richter K, Buchner J. Mixed Hsp90-cochaperone complexes are important for the progression of the reaction cycle. Nat Struct Mol Biol. 2011;18:61–6. [Abstract] [Google Scholar]
  • Li R, Soosairajah J, Harari D, Citri A, Price J, Ng HL, Morton CJ, Parker MW, Yarden Y, Bernard O. Hsp90 increases LIM kinase activity by promoting its homo-dimerization. Faseb J. 2006;20:1218–20. [Abstract] [Google Scholar]
  • Li WH, Miao XH, Qi ZT, Ni W, Zhu SY, Fang F. Proteomic analysis of differently expressed proteins in human hepatocellular carcinoma cell lines HepG2 with transfecting hepatitis B virus X gene. Chin Med J (Engl) 2009;122:15–23. [Abstract] [Google Scholar]
  • Lindquist S. The heat-shock response. Annu Rev Biochem. 1986;55:1151–91. [Abstract] [Google Scholar]
  • Liu T, Daniels CK, Cao S. Comprehensive review on the HSC70 functions, interactions with related molecules and involvement in clinical diseases and therapeutic potential. Pharmacol Ther. 2012;136:354–74. [Abstract] [Google Scholar]
  • Liu Y, Liu M, Liu J, Zhang H, Tu Z, Xiao X. KLF4 is a novel regulator of the constitutively expressed HSP90. Cell Stress Chaperones. 2010;15:211–7. [Europe PMC free article] [Abstract] [Google Scholar]
  • Lobert VH, Brech A, Pedersen NM, Wesche J, Oppelt A, Malerod L, Stenmark H. Ubiquitination of alpha 5 beta 1 integrin controls fibroblast migration through lysosomal degradation of fibronectin-integrin complexes. Dev Cell. 2010;19:148–59. [Abstract] [Google Scholar]
  • Lovgren AK, Kovarova M, Koller BH. cPGES/p23 is required for glucocorticoid receptor function and embryonic growth but not prostaglandin E2 synthesis. Mol Cell Biol. 2007;27:4416–30. [Europe PMC free article] [Abstract] [Google Scholar]
  • Lowrie AG, Salter DM, Ross JA. Latent effects of fibronectin, alpha5beta1 integrin, alphaVbeta5 integrin and the cytoskeleton regulate pancreatic carcinoma cell IL-8 secretion. Br J Cancer. 2004;91:1327–34. [Europe PMC free article] [Abstract] [Google Scholar]
  • Ma Y, Bogatcheva NV, Gusev NB. Heat shock protein (hsp90) interacts with smooth muscle calponin and affects calponin-binding to actin. Biochim Biophys Acta. 2000;1476:300–10. [Abstract] [Google Scholar]
  • Maxwell PH, Wiesener MS, Chang GW, Clifford SC, Vaux EC, Cockman ME, Wykoff CC, Pugh CW, Maher ER, Ratcliffe PJ. The tumour suppressor protein VHL targets hypoxia-inducible factors for oxygen-dependent proteolysis. Nature. 1999;399:271–5. [Abstract] [Google Scholar]
  • McClellan AJ, Scott MD, Frydman J. Folding and quality control of the VHL tumor suppressor proceed through distinct chaperone pathways. Cell. 2005;121:739–48. [Abstract] [Google Scholar]
  • McDowell CL, Bryan Sutton R, Obermann WM. Expression of Hsp90 chaperone [corrected] proteins in human tumor tissue. Int J Biol Macromol. 2009;45:310–4. [Abstract] [Google Scholar]
  • McKeen HD, McAlpine K, Valentine A, Quinn DJ, McClelland K, Byrne C, O’Rourke M, Young S, Scott CJ, McCarthy HO, et al. A novel FK506-like binding protein interacts with the glucocorticoid receptor and regulates steroid receptor signaling. Endocrinology. 2008;149:5724–34. [Abstract] [Google Scholar]
  • McLaughlin SH, Sobott F, Yao ZP, Zhang W, Nielsen PR, Grossmann JG, Laue ED, Robinson CV, Jackson SE. The co-chaperone p23 arrests the Hsp90 ATPase cycle to trap client proteins. J Mol Biol. 2006;356:746–58. [Abstract] [Google Scholar]
  • Mendel DB, Orti E. Isoform composition and stoichiometry of the approximately 90-kDa heat shock protein associated with glucocorticoid receptors. J Biol Chem. 1988;263:6695–702. [Abstract] [Google Scholar]
  • Metchat A, Akerfelt M, Bierkamp C, Delsinne V, Sistonen L, Alexandre H, Christians ES. Mammalian heat shock factor 1 is essential for oocyte meiosis and directly regulates Hsp90alpha expression. J Biol Chem. 2009;284:9521–8. [Europe PMC free article] [Abstract] [Google Scholar]
  • Meyer P, Prodromou C, Liao C, Hu B, Mark Roe S, Vaughan CK, Vlasic I, Panaretou B, Piper PW, Pearl LH. Structural basis for recruitment of the ATPase activator Aha1 to the Hsp90 chaperone machinery. Embo J. 2004;23:511–9. [Europe PMC free article] [Abstract] [Google Scholar]
  • Millson SH, Vaughan CK, Zhai C, Ali MM, Panaretou B, Piper PW, Pearl LH, Prodromou C. Chaperone ligand-discrimination by the TPR-domain protein Tah1. Biochem J. 2008;413:261–8. [Europe PMC free article] [Abstract] [Google Scholar]
  • 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. 1991;266:10099–103. [Abstract] [Google Scholar]
  • Miwa D, Sakaue T, Inoue H, Takemori N, Kurokawa M, Fukuda S, Omi K, Goishi K, Higashiyama S. Protein kinase D2 and heat shock protein 90 beta are required for BCL6-associated zinc finger protein mRNA stabilization induced by vascular endothelial growth factor-A. Angiogenesis. 2013;16:675–88. [Abstract] [Google Scholar]
  • Miyata Y, Chambraud B, Radanyi C, Leclerc J, Lebeau MC, Renoir JM, Shirai R, Catelli MG, Yahara I, Baulieu EE. Phosphorylation of the immunosuppressant FK506-binding protein FKBP52 by casein kinase II: regulation of HSP90-binding activity of FKBP52. Proc Natl Acad Sci U S A. 1997;94:14500–5. [Europe PMC free article] [Abstract] [Google Scholar]
  • Miyoshi T, Takeuchi A, Siomi H, Siomi MC. A direct role for Hsp90 in pre-RISC formation in Drosophila. Nat Struct Mol Biol. 2010;17:1024–6. [Abstract] [Google Scholar]
  • Mok D, Allan RK, Carrello A, Wangoo K, Walkinshaw MD, Ratajczak T. The chaperone function of cyclophilin 40 maps to a cleft between the prolyl isomerase and tetratricopeptide repeat domains. FEBS Lett. 2006;580:2761–8. [Abstract] [Google Scholar]
  • Moore SK, Kozak C, Robinson EA, Ullrich SJ, Appella E. Cloning and nucleotide sequence of the murine hsp84 cDNA and chromosome assignment of related sequences. Gene. 1987;56:29–40. [Abstract] [Google Scholar]
  • 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. 2004;279:48846–54. [Abstract] [Google Scholar]
  • Murphy PJ, Kanelakis KC, Galigniana MD, Morishima Y, Pratt WB. Stoichiometry, abundance, and functional significance of the hsp90/hsp70-based multiprotein chaperone machinery in reticulocyte lysate. J Biol Chem. 2001;276:30092–8. [Abstract] [Google Scholar]
  • Nenasheva VV, Kovaleva GV, Khaidarova NV, Novosadova EV, Manuilova ES, Antonov SA, Tarantul VZ. Trim14 overexpression causes the same transcriptional changes in mouse embryonic stem cells and human HEK293 cells. In Vitro Cell Dev Biol Anim. 2014;50:121–8. [Abstract] [Google Scholar]
  • New M, Olzscha H, Liu G, Khan O, Stimson L, McGouran J, Kerr D, Coutts A, Kessler B, Middleton M, et al. A regulatory circuit that involves HR23B and HDAC6 governs the biological response to HDAC inhibitors. Cell Death Differ. 2013;20:1306–16. [Europe PMC free article] [Abstract] [Google Scholar]
  • Noh H, Kim HJ, Yu MR, Kim WY, Kim J, Ryu JH, Kwon SH, Jeon JS, Han DC, Ziyadeh F. Heat shock protein 90 inhibitor attenuates renal fibrosis through degradation of transforming growth factor-beta type II receptor. Lab Invest. 2012;92:1583–96. [Abstract] [Google Scholar]
  • Novosadova EV, Manuilova ES, Arsenieva EL, Lebedev AN, Khaidarova NV, Tarantul VZ, Grivennikov IA. Influence of pub Gene Expression on Differentiation of Mouse Embryonic Stem Cells into Derivatives of Ecto-, Meso-, and Endoderm in vitro. Acta Naturae. 2009;1:93–7. [Europe PMC free article] [Abstract] [Google Scholar]
  • O’Keeffe B, Fong Y, Chen D, Zhou S, Zhou Q. Requirement for a kinase-specific chaperone pathway in the production of a Cdk9/cyclin T1 heterodimer responsible for P-TEFb-mediated tat stimulation of HIV-1 transcription. J Biol Chem. 2000;275:279–87. [Abstract] [Google Scholar]
  • Obermann WM, Sondermann H, Russo AA, Pavletich NP, Hartl FU. In vivo function of Hsp90 is dependent on ATP binding and ATP hydrolysis. J Cell Biol. 1998;143:901–10. [Europe PMC free article] [Abstract] [Google Scholar]
  • Ogiso H, Kagi N, Matsumoto E, Nishimoto M, Arai R, Shirouzu M, Mimura J, Fujii-Kuriyama Y, Yokoyama S. Phosphorylation analysis of 90 kDa heat shock protein within the cytosolic arylhydrocarbon receptor complex. Biochemistry. 2004;43:15510–9. [Abstract] [Google Scholar]
  • Ohmori H, Friedberg EC, Fuchs RP, Goodman MF, Hanaoka F, Hinkle D, Kunkel TA, Lawrence CW, Livneh Z, Nohmi T, et al. The Y-family of DNA polymerases. Mol Cell. 2001;8:7–8. [Abstract] [Google Scholar]
  • Ohtsuka K, Hata M. Molecular chaperone function of mammalian Hsp70 and Hsp40–a review. Int J Hyperthermia. 2000;16:231–45. [Abstract] [Google Scholar]
  • Okayama S, Kopelovich L, Balmus G, Weiss RS, Herbert BS, Dannenberg AJ, Subbaramaiah K. p53 protein regulates Hsp90 ATPase activity and thereby Wnt signaling by modulating Aha1 expression. J Biol Chem. 2014;289:6513–25. [Europe PMC free article] [Abstract] [Google Scholar] Retracted
  • Okumura F, Okumura AJ, Matsumoto M, Nakayama KI, Hatakeyama S. TRIM8 regulates Nanog via Hsp90beta-mediated nuclear translocation of STAT3 in embryonic stem cells. Biochim Biophys Acta. 2011;1813:1784–92. [Abstract] [Google Scholar]
  • Owens-Grillo JK, Czar MJ, Hutchison KA, Hoffmann K, Perdew GH, Pratt WB. A model of protein targeting mediated by immunophilins and other proteins that bind to hsp90 via tetratricopeptide repeat domains. J Biol Chem. 1996;271:13468–75. [Abstract] [Google Scholar]
  • Oxelmark E, Roth JM, Brooks PC, Braunstein SE, Schneider RJ, Garabedian MJ. The cochaperone p23 differentially regulates estrogen receptor target genes and promotes tumor cell adhesion and invasion. Mol Cell Biol. 2006;26:5205–13. [Europe PMC free article] [Abstract] [Google Scholar]
  • Pare JM, LaPointe P, Hobman TC. Hsp90 cochaperones p23 and FKBP4 physically interact with hAgo2 and activate RNA interference-mediated silencing in mammalian cells. Mol Biol Cell. 2013;24:2303–10. [Europe PMC free article] [Abstract] [Google Scholar]
  • Park MS, Chu F, Xie J, Wang Y, Bhattacharya P, Chan WK. Identification of cyclophilin-40-interacting proteins reveals potential cellular function of cyclophilin-40. Anal Biochem. 2011;410:257–65. [Europe PMC free article] [Abstract] [Google Scholar]
  • Park SJ, Suetsugu S, Takenawa T. Interaction of HSP90 to N-WASP leads to activation and protection from proteasome-dependent degradation. EMBO J. 2005;24:1557–70. [Europe PMC free article] [Abstract] [Google Scholar]
  • Park SJ, Suetsugu S, Sagara H, Takenawa T. HSP90 cross-links branched actin filaments induced by N-WASP and the Arp2/3 complex. Genes Cells. 2007;12:611–22. [Abstract] [Google Scholar]
  • Parrilla-Castellar ER, Arlander SJ, Karnitz L. Dial 9–1-1 for DNA damage: the Rad9-Hus1-Rad1 (9–1-1) clamp complex. DNA Repair (Amst) 2004;3:1009–14. [Abstract] [Google Scholar]
  • Pearl LH, Prodromou C. Structure and in vivo function of Hsp90. Curr Opin Struct Biol. 2000;10:46–51. [Abstract] [Google Scholar]
  • Pearl LH, Prodromou C. Structure and mechanism of the Hsp90 molecular chaperone machinery. Annu Rev Biochem. 2006;75:271–94. [Abstract] [Google Scholar]
  • Peng M, Emmadi R, Wang Z, Wiley EL, Gann PH, Khan SA, Banerji N, McDonald W, Asztalos S, Pham TN, et al. PTK6/BRK is expressed in the normal mammary gland and activated at the plasma membrane in breast tumors. Oncotarget. 2014;5:6038–48. [Europe PMC free article] [Abstract] [Google Scholar]
  • Pires ES, Khole VV. A block in the road to fertility: autoantibodies to heat-shock protein 90-beta in human ovarian autoimmunity. Fertil Steril. 2009;92:1395–409. [Abstract] [Google Scholar]
  • Pires ES, Meherji PK, Vaidya RR, Parikh FR, Ghosalkar MN, Khole VV. Specific and sensitive immunoassays detect multiple anti-ovarian antibodies in women with infertility. J Histochem Cytochem. 2007;55:1181–90. [Abstract] [Google Scholar]
  • Pirkl F, Buchner J. Functional analysis of the Hsp90-associated human peptidyl prolyl cis/trans isomerases FKBP51, FKBP52 and Cyp40. J Mol Biol. 2001;308:795–806. [Abstract] [Google Scholar]
  • Piwien Pilipuk G, Vinson GP, Sanchez CG, Galigniana MD. Evidence for NL1-independent nuclear translocation of the mineralocorticoid receptor. Biochemistry. 2007;46:1389–97. [Abstract] [Google Scholar]
  • Pozo FM, Oda T, Sekimoto T, Murakumo Y, Masutani C, Hanaoka F, Yamashita T. Molecular chaperone Hsp90 regulates REV1-mediated mutagenesis. Mol Cell Biol. 2011;31:3396–409. [Europe PMC free article] [Abstract] [Google Scholar]
  • Pratt WB, Silverstein AM, Galigniana MD. A model for the cytoplasmic trafficking of signalling proteins involving the hsp90-binding immunophilins and p50cdc37. Cell Signal. 1999;11:839–51. [Abstract] [Google Scholar]
  • Pratt WB, Galigniana MD, Harrell JM, DeFranco DB. Role of hsp90 and the hsp90-binding immunophilins in signalling protein movement. Cell Signal. 2004;16:857–72. [Abstract] [Google Scholar]
  • Prodromou C, Roe SM, Piper PW, Pearl LH. A molecular clamp in the crystal structure of the N-terminal domain of the yeast Hsp90 chaperone. Nat Struct Biol. 1997a;4:477–82. [Abstract] [Google Scholar]
  • Prodromou C, Roe SM, O’Brien R, Ladbury JE, Piper PW, Pearl LH. Identification and structural characterization of the ATP/ADP-binding site in the Hsp90 molecular chaperone. Cell. 1997b;90:65–75. [Abstract] [Google Scholar]
  • Prodromou C, Panaretou B, Chohan S, Siligardi G, O’Brien R, Ladbury JE, Roe SM, Piper PW, Pearl LH. The ATPase cycle of Hsp90 drives a molecular ‘clamp’ via transient dimerization of the N-terminal domains. EMBO J. 2000;19:4383–92. [Europe PMC free article] [Abstract] [Google Scholar]
  • Qin S, Ni M, Wang X, Maurier-Mahe F, Shurland DL, Rodrigues GA. Inhibition of RPE cell sterile inflammatory responses and endotoxin-induced uveitis by a cell-impermeable HSP90 inhibitor. Exp Eye Res. 2011;93:889–97. [Abstract] [Google Scholar]
  • Quinta HR, Galigniana MD. The neuroregenerative mechanism mediated by the Hsp90-binding immunophilin FKBP52 resembles the early steps of neuronal differentiation. Br J Pharmacol. 2012;166:637–49. [Europe PMC free article] [Abstract] [Google Scholar]
  • Quinta HR, Maschi D, Gomez-Sanchez C, Piwien-Pilipuk G, Galigniana MD. Subcellular rearrangement of hsp90-binding immunophilins accompanies neuronal differentiation and neurite outgrowth. J Neurochem. 2010;115:716–34. [Europe PMC free article] [Abstract] [Google Scholar]
  • Quinta HR, Galigniana NM, Erlejman AG, Lagadari M, Piwien-Pilipuk G, Galigniana MD. Management of cytoskeleton architecture by molecular chaperones and immunophilins. Cell Signal. 2011;23:1907–20. [Europe PMC free article] [Abstract] [Google Scholar]
  • Radovanac K, Morgner J, Schulz JN, Blumbach K, Patterson C, Geiger T, Mann M, Krieg T, Eckes B, Fassler R, et al. Stabilization of integrin-linked kinase by the Hsp90-CHIP axis impacts cellular force generation, migration and the fibrotic response. Embo J. 2013;32:1409–24. [Europe PMC free article] [Abstract] [Google Scholar]
  • Ramalingam S, Goss G, Rosell R, Schmid-Bindert G, Zaric B, Andric Z, Bondarenko I, Komov D, Ceric T, Khuri F, et al. A randomized phase II study of ganetespib, a heat shock protein 90 inhibitor, in combination with docetaxel in second-line therapy of advanced non-small cell lung cancer (GALAXY-1) Ann Oncol 2015 [Abstract] [Google Scholar]
  • Rebbe NF, Ware J, Bertina RM, Modrich P, Stafford DW. Nucleotide sequence of a cDNA for a member of the human 90-kDa heat-shock protein family. Gene. 1987;53:235–45. [Abstract] [Google Scholar]
  • Rebbe NF, Hickman WS, Ley TJ, Stafford DW, Hickman S. Nucleotide sequence and regulation of a human 90-kDa heat shock protein gene. J Biol Chem. 1989;264:15006–11. [Abstract] [Google Scholar]
  • Regan PL, Jacobs J, Wang G, Torres J, Edo R, Friedmann J, Tang XX. Hsp90 inhibition increases p53 expression and destabilizes MYCN and MYC in neuroblastoma. Int J Oncol. 2011;38:105–12. [Europe PMC free article] [Abstract] [Google Scholar]
  • Retzlaff M, Stahl M, Eberl HC, Lagleder S, Beck J, Kessler H, Buchner J. Hsp90 is regulated by a switch point in the C-terminal domain. EMBO Rep. 2009;10:1147–53. [Europe PMC free article] [Abstract] [Google Scholar]
  • Rexin M, Busch W, Gehring U. Protein components of the nonactivated glucocorticoid receptor. J Biol Chem. 1991;266:24601–5. [Abstract] [Google Scholar]
  • Riggs DL, Roberts PJ, Chirillo SC, Cheung-Flynn J, Prapapanich V, Ratajczak T, Gaber R, Picard D, Smith DF. The Hsp90-binding peptidylprolyl isomerase FKBP52 potentiates glucocorticoid signaling in vivo. Embo J. 2003;22:1158–67. [Europe PMC free article] [Abstract] [Google Scholar]
  • Riordan JR. Assembly of functional CFTR chloride channels. Annu Rev Physiol. 2005;67:701–18. [Abstract] [Google Scholar]
  • Roe SM, Prodromou C, O’Brien R, Ladbury JE, Piper PW, Pearl LH. Structural basis for inhibition of the Hsp90 molecular chaperone by the antitumor antibiotics radicicol and geldanamycin. J Med Chem. 1999;42:260–6. [Abstract] [Google Scholar]
  • Roe SM, Ali MM, Meyer P, Vaughan CK, Panaretou B, Piper PW, Prodromou C, Pearl LH. The Mechanism of Hsp90 regulation by the protein kinase-specific cochaperone p50(cdc37) Cell. 2004;116:87–98. [Abstract] [Google Scholar]
  • Rutherford SL, Lindquist S. Hsp90 as a capacitor for morphological evolution. Nature. 1998;396:336–42. [Abstract] [Google Scholar]
  • Sadeu JC, Foster WG. The cigarette smoke constituent benzo[a]pyrene disrupts metabolic enzyme, and apoptosis pathway member gene expression in ovarian follicles. Reprod Toxicol. 2013;40:52–9. [Abstract] [Google Scholar]
  • Samadi A, Loo P, Mukerji R, O’Donnell G, Tong X, Timmermann BN, Cohen MS. A novel HSP90 modulator with selective activity against thyroid cancers in vitro. Surgery. 2009;146:1196–207. [Europe PMC free article] [Abstract] [Google Scholar]
  • Sanokawa-Akakura R, Cao W, Allan K, Patel K, Ganesh A, Heiman G, Burke R, Kemp FW, Bogden JD, Camakaris J, et al. Control of Alzheimer’s amyloid beta toxicity by the high molecular weight immunophilin FKBP52 and copper homeostasis in Drosophila. PLoS One. 2010;5:e8626. [Europe PMC free article] [Abstract] [Google Scholar]
  • Scammell JG, Hubler TR, Denny WB, Valentine DL. Organization of the human FK506-binding immunophilin FKBP52 protein gene (FKBP4) Genomics. 2003;81:640–3. [Abstract] [Google Scholar]
  • Scheibel T, Siegmund HI, Jaenicke R, Ganz P, Lilie H, Buchner J. The charged region of Hsp90 modulates the function of the N-terminal domain. Proc Natl Acad Sci U S A. 1999;96:1297–302. [Europe PMC free article] [Abstract] [Google Scholar]
  • Scheufler C, Brinker A, Bourenkov G, Pegoraro S, Moroder L, Bartunik H, Hartl FU, Moarefi I. Structure of TPR domain-peptide complexes: critical elements in the assembly of the Hsp70–Hsp90 multichaperone machine. Cell. 2000;101:199–210. [Abstract] [Google Scholar]
  • Schulte TW, Akinaga S, Soga S, Sullivan W, Stensgard B, Toft D, Neckers LM. Antibiotic radicicol binds to the N-terminal domain of Hsp90 and shares important biologic activities with geldanamycin. Cell Stress Chaperones. 1998;3:100–8. [Europe PMC free article] [Abstract] [Google Scholar]
  • Senftleben U, Karin M. The IKK/NF-kappaB pathway. Crit Care Med. 2002;30:S18–S26. [Abstract] [Google Scholar]
  • Sepehrnia B, Paz IB, Dasgupta G, Momand J. Heat shock protein 84 forms a complex with mutant p53 protein predominantly within a cytoplasmic compartment of the cell. J Biol Chem. 1996;271:15084–90. [Abstract] [Google Scholar]
  • Shang Y, Xu X, Duan X, Guo J, Wang Y, Ren F, He D, Chang Z. Hsp70 and Hsp90 oppositely regulate TGF-beta signaling through CHIP/Stub1. Biochem Biophys Res Commun. 2014;446:387–92. [Abstract] [Google Scholar]
  • Shen HY, Zhao Y, Chen XY, Xiong RP, Lu JL, Chen JF, Chen LY, Zhou YG. Differential alteration of heat shock protein 90 in mice modifies glucocorticoid receptor function and susceptibility to trauma. J Neurotrauma. 2010;27:373–81. [Abstract] [Google Scholar]
  • Shen Y, Liu J, Wang X, Cheng X, Wang Y, Wu N. Essential role of the first intron in the transcription of hsp90beta gene. FEBS Lett. 1997;413:92–8. [Abstract] [Google Scholar]
  • Shibata T, Kimura Y, Mukai A, Mori H, Ito S, Asaka Y, Oe S, Tanaka H, Takahashi T, Uchida K. Transthiocarbamoylation of proteins by thiolated isothiocyanates. J Biol Chem. 2011;286:42150–61. [Europe PMC free article] [Abstract] [Google Scholar]
  • Shirane M, Nakayama KI. Inherent calcineurin inhibitor FKBP38 targets Bcl-2 to mitochondria and inhibits apoptosis. Nat Cell Biol. 2003;5:28–37. [Abstract] [Google Scholar]
  • Silverstein AM, Galigniana MD, Chen MS, Owens-Grillo JK, Chinkers M, Pratt WB. Protein phosphatase 5 is a major component of glucocorticoid receptor.hsp90 complexes with properties of an FK506-binding immunophilin. J Biol Chem. 1997;272:16224–30. [Abstract] [Google Scholar]
  • Sivils JC, Storer CL, Galigniana MD, Cox MB. Regulation of steroid hormone receptor function by the 52-kDa FK506-binding protein (FKBP52) Curr Opin Pharmacol. 2011;11:314–9. [Europe PMC free article] [Abstract] [Google Scholar]
  • Smith DF. Dynamics of heat shock protein 90-progesterone receptor binding and the disactivation loop model for steroid receptor complexes. Mol Endocrinol. 1993;7:1418–29. [Abstract] [Google Scholar]
  • Stebbins CE, Russo AA, Schneider C, Rosen N, Hartl FU, Pavletich NP. Crystal structure of an Hsp90-geldanamycin complex: targeting of a protein chaperone by an antitumor agent. Cell. 1997;89:239–50. [Abstract] [Google Scholar]
  • Stepanova L, Leng X, Parker SB, Harper JW. Mammalian p50Cdc37 is a protein kinase-targeting subunit of Hsp90 that binds and stabilizes Cdk4. Genes Dev. 1996;10:1491–502. [Abstract] [Google Scholar]
  • Sternberg N. Properties of a mutant of Escherichia coli defective in bacteriophage lambda head formation (groE). I. Initial characterization. J Mol Biol. 1973a;76:1–23. [Abstract] [Google Scholar]
  • Sternberg N. Properties of a mutant of Escherichia coli defective in bacteriophage lambda head formation (groE). II. The propagation of phage lambda. J Mol Biol. 1973b;76:25–44. [Abstract] [Google Scholar]
  • Storer CL, Dickey CA, Galigniana MD, Rein T, Cox MB. FKBP51 and FKBP52 in signaling and disease. Trends Endocrinol Metab. 2011;22:481–90. [Europe PMC free article] [Abstract] [Google Scholar]
  • Stricher F, Macri C, Ruff M, Muller S. HSPA8/HSC70 chaperone protein: structure, function, and chemical targeting. Autophagy. 2013;9:1937–54. [Abstract] [Google Scholar]
  • Sun Y, Biscarini F, Bovenhuis H, Parmentier HK, van der Poel JJ. Genetic parameters and across-line SNP associations differ for natural antibody isotypes IgM and IgG in laying hens. Anim Genet. 2013;44:413–24. [Abstract] [Google Scholar]
  • Sunnotel O, Hiripi L, Lagan K, McDaid JR, De Leon JM, Miyagawa Y, Crowe H, Kaluskar S, Ward M, Scullion C, et al. Alterations in the steroid hormone receptor co-chaperone FKBPL are associated with male infertility: a case-control study. Reprod Biol Endocrinol. 2010;8:22. [Europe PMC free article] [Abstract] [Google Scholar]
  • Suzuki S, Kulkarni AB. Extracellular heat shock protein HSP90beta secreted by MG63 osteosarcoma cells inhibits activation of latent TGF-beta1. Biochem Biophys Res Commun. 2010;398:525–31. [Europe PMC free article] [Abstract] [Google Scholar]
  • Synoradzki K, Bieganowski P. Middle domain of human Hsp90 isoforms differentially binds Aha1 in human cells and alters Hsp90 activity in yeast. Biochim Biophys Acta. 2015;1853:445–52. [Abstract] [Google Scholar]
  • Taipale M, Jarosz DF, Lindquist S. HSP90 at the hub of protein homeostasis: emerging mechanistic insights. Nat Rev Mol Cell Biol. 2010;11:515–28. [Abstract] [Google Scholar]
  • Taipale M, Krykbaeva I, Koeva M, Kayatekin C, Westover KD, Karras GI, Lindquist S. Quantitative analysis of HSP90-client interactions reveals principles of substrate recognition. Cell. 2012;150:987–1001. [Europe PMC free article] [Abstract] [Google Scholar]
  • Taipale M, Krykbaeva I, Whitesell L, Santagata S, Zhang J, Liu Q, Gray NS, Lindquist S. Chaperones as thermodynamic sensors of drug-target interactions reveal kinase inhibitor specificities in living cells. Nat Biotechnol. 2013;31:630–7. [Europe PMC free article] [Abstract] [Google Scholar]
  • Taipale M, Tucker G, Peng J, Krykbaeva I, Lin ZY, Larsen B, Choi H, Berger B, Gingras AC, Lindquist S. A quantitative chaperone interaction network reveals the architecture of cellular protein homeostasis pathways. Cell. 2014;158:434–48. [Europe PMC free article] [Abstract] [Google Scholar]
  • Taiyab A, Rao ChM. HSP90 modulates actin dynamics: inhibition of HSP90 leads to decreased cell motility and impairs invasion. Biochim Biophys Acta. 2011;1813:213–21. [Abstract] [Google Scholar]
  • Tanaka K, Eskin A, Chareyre F, Jessen WJ, Manent J, Niwa-Kawakita M, Chen R, White CH, Vitte J, Jaffer ZM, et al. Therapeutic potential of HSP90 inhibition for neurofibromatosis type 2. Clin Cancer Res. 2013;19:3856–70. [Europe PMC free article] [Abstract] [Google Scholar]
  • Tenge VR, Zuehlke AD, Shrestha N, Johnson JL. The Hsp90 cochaperones Cpr6, Cpr7, and Cns1 interact with the intact ribosome. Eukaryot Cell. 2015;14:55–63. [Europe PMC free article] [Abstract] [Google Scholar]
  • Thompson AD, Scaglione KM, Prensner J, Gillies AT, Chinnaiyan A, Paulson HL, Jinwal UK, Dickey CA, Gestwicki JE. Analysis of the tau-associated proteome reveals that exchange of Hsp70 for Hsp90 is involved in tau degradation. ACS Chem Biol. 2012;7:1677–86. [Europe PMC free article] [Abstract] [Google Scholar]
  • Thoreen CC, Chantranupong L, Keys HR, Wang T, Gray NS, Sabatini DM. A unifying model for mTORC1-mediated regulation of mRNA translation. Nature. 2012;485:109–13. [Europe PMC free article] [Abstract] [Google Scholar]
  • Tiedemann RE, Zhu YX, Schmidt J, Yin H, Shi CX, Que Q, Basu G, Azorsa D, Perkins LM, Braggio E, et al. Kinome-wide RNAi studies in human multiple myeloma identify vulnerable kinase targets, including a lymphoid-restricted kinase, GRK6. Blood. 2010;115:1594–604. [Europe PMC free article] [Abstract] [Google Scholar]
  • Tomcik M, Zerr P, Pitkowski J, Palumbo-Zerr K, Avouac J, Distler O, Becvar R, Senolt L, Schett G, Distler JH. Heat shock protein 90 (Hsp90) inhibition targets canonical TGF-beta signalling to prevent fibrosis. Ann Rheum Dis. 2014;73:1215–22. [Abstract] [Google Scholar]
  • Toneatto J, Guber S, Charo NL, Susperreguy S, Schwartz J, Galigniana MD, Piwien-Pilipuk G. Dynamic mitochondrial-nuclear redistribution of the immunophilin FKBP51 is regulated by the PKA signaling pathway to control gene expression during adipocyte differentiation. J Cell Sci. 2013;126:5357–68. [Europe PMC free article] [Abstract] [Google Scholar]
  • Travers MA, Meistertzheim AL, Cardinaud M, Friedman CS, Huchette S, Moraga D, Paillard C. Gene expression patterns of abalone, Haliotis tuberculata, during successive infections by the pathogen Vibrio harveyi. J Invertebr Pathol. 2010;105:289–97. [Abstract] [Google Scholar]
  • Trisciuoglio D, Gabellini C, Desideri M, Ziparo E, Zupi G, Del Bufalo D. Bcl-2 regulates HIF-1alpha protein stabilization in hypoxic melanoma cells via the molecular chaperone HSP90. PLoS One. 2010;5:e11772. [Europe PMC free article] [Abstract] [Google Scholar] Retracted
  • Tsaytler PA, Krijgsveld J, Goerdayal SS, Rudiger S, Egmond MR. Novel Hsp90 partners discovered using complementary proteomic approaches. Cell Stress Chaperones. 2009;14:629–38. [Europe PMC free article] [Abstract] [Google Scholar]
  • Tsumuraya M, Kato H, Miyachi K, Sasaki K, Tsubaki M, Akimoto K, Sunagawa M. Comprehensive analysis of genes involved in the malignancy of gastrointestinal stromal tumors. Anticancer Res. 2010;30:2705–15. [Abstract] [Google Scholar]
  • Ullrich SJ, Moore SK, Appella E. Transcriptional and translational analysis of the murine 84- and 86-kDa heat shock proteins. J Biol Chem. 1989;264:6810–6. [Abstract] [Google Scholar]
  • Urban JD, Budinsky RA, Rowlands JC. An evaluation of single nucleotide polymorphisms in the human heat shock protein 90 kDa alpha and beta isoforms. Drug Metab Pharmacokinet. 2012;27:268–78. [Abstract] [Google Scholar]
  • Valentine A, O’Rourke M, Yakkundi A, Worthington J, Hookham M, Bicknell R, McCarthy HO, McClelland K, McCallum L, Dyer H, et al. FKBPL and peptide derivatives: novel biological agents that inhibit angiogenesis by a CD44-dependent mechanism. Clin Cancer Res. 2011;17:1044–56. [Europe PMC free article] [Abstract] [Google Scholar]
  • van Dongen G, Geilenkirchen WL, van Rijn J, van Wijk R. Increase of thermoresistance after growth stimulation of resting Reuber H35 hepatoma cells. Alteration of nuclear characteristics, non-histone chromosomal protein phosphorylation and basal heat shock protein synthesis. Exp Cell Res. 1986;166:427–41. [Abstract] [Google Scholar]
  • Vaughan CK, Gohlke U, Sobott F, Good VM, Ali MM, Prodromou C, Robinson CV, Saibil HR, Pearl LH. Structure of an Hsp90-Cdc37-Cdk4 complex. Mol Cell. 2006;23:697–707. [Abstract] [Google Scholar]
  • Vaughan CK, Mollapour M, Smith JR, Truman A, Hu B, Good VM, Panaretou B, Neckers L, Clarke PA, Workman P, et al. Hsp90-dependent activation of protein kinases is regulated by chaperone-targeted dephosphorylation of Cdc37. Mol Cell. 2008;31:886–95. [Europe PMC free article] [Abstract] [Google Scholar]
  • Wahl A, Schafer F, Bardet W, Hildebrand WH. HLA class I molecules reflect an altered host proteome after influenza virus infection. Hum Immunol. 2010;71:14–22. [Europe PMC free article] [Abstract] [Google Scholar]
  • Walerych D, Kudla G, Gutkowska M, Wawrzynow B, Muller L, King FW, Helwak A, Boros J, Zylicz A, Zylicz M. Hsp90 chaperones wild-type p53 tumor suppressor protein. J Biol Chem. 2004;279:48836–45. [Abstract] [Google Scholar]
  • Wang G, Gu X, Chen L, Wang Y, Cao B, E Q. Comparison of the expression of 5 heat shock proteins in benign and malignant salivary gland tumor tissues. Oncol Lett. 2013;5:1363–1369. [Europe PMC free article] [Abstract] [Google Scholar]
  • Wang X, Venable J, LaPointe P, Hutt DM, Koulov AV, Coppinger J, Gurkan C, Kellner W, Matteson J, Plutner H, et al. Hsp90 cochaperone Aha1 downregulation rescues misfolding of CFTR in cystic fibrosis. Cell. 2006;127:803–15. [Abstract] [Google Scholar]
  • Wegele H, Muller L, Buchner J. Hsp70 and Hsp90–a relay team for protein folding. Rev Physiol Biochem Pharmacol. 2004;151:1–44. [Abstract] [Google Scholar]
  • Weis F, Moullintraffort L, Heichette C, Chretien D, Garnier C. The 90-kDa heat shock protein Hsp90 protects tubulin against thermal denaturation. J Biol Chem. 2010;285:9525–34. [Europe PMC free article] [Abstract] [Google Scholar]
  • Welch WJ, Feramisco JR. Purification of the major mammalian heat shock proteins. J Biol Chem. 1982;257:14949–59. [Abstract] [Google Scholar]
  • Woo SH, An S, Lee HC, Jin HO, Seo SK, Yoo DH, Lee KH, Rhee CH, Choi EJ, Hong SI, et al. A truncated form of p23 down-regulates telomerase activity via disruption of Hsp90 function. J Biol Chem. 2009;284:30871–80. [Europe PMC free article] [Abstract] [Google Scholar]
  • Xiang SL, Kumano T, Iwasaki SI, Sun X, Yoshioka K, Yamamoto KC. The J domain of Tpr2 regulates its interaction with the proapoptotic and cell-cycle checkpoint protein, Rad9. Biochem Biophys Res Commun. 2001;287:932–40. [Abstract] [Google Scholar]
  • Xu Z, Horwich AL, Sigler PB. The crystal structure of the asymmetric GroEL-GroES-(ADP)7 chaperonin complex. Nature. 1997;388:741–50. [Abstract] [Google Scholar]
  • Yahara I. The role of HSP90 in evolution. Genes Cells. 1999;4:375–9. [Abstract] [Google Scholar]
  • Yakkundi A, McCallum L, O’Kane A, Dyer H, Worthington J, McKeen HD, McClements L, Elliott C, McCarthy HO, Hirst DG, et al. The anti-migratory effects of FKBPL and its peptide derivative, AD-01: regulation of CD44 and the cytoskeletal pathway. PLoS One. 2013;8:e55075. [Europe PMC free article] [Abstract] [Google Scholar]
  • Yamamori T, Yura T. Temperature-induced synthesis of specific proteins in Escherichia coli: evidence for transcriptional control. J Bacteriol. 1980;142:843–51. [Europe PMC free article] [Abstract] [Google Scholar]
  • Yang C, Rahimpour S, Lu J, Pacak K, Ikejiri B, Brady RO, Zhuang Z. Histone deacetylase inhibitors increase glucocerebrosidase activity in Gaucher disease by modulation of molecular chaperones. Proc Natl Acad Sci U S A. 2013;110:966–71. [Europe PMC free article] [Abstract] [Google Scholar]
  • Yang J, Roe SM, Cliff MJ, Williams MA, Ladbury JE, Cohen PT, Barford D. Molecular basis for TPR domain-mediated regulation of protein phosphatase 5. EMBO J. 2005;24:1–10. [Europe PMC free article] [Abstract] [Google Scholar]
  • Yin H, Wang H, Zong H, Chen X, Wang Y, Yun X, Wu Y, Wang J, Gu J. SGT, a Hsp90beta binding partner, is accumulated in the nucleus during cell apoptosis. Biochem Biophys Res Commun. 2006;343:1153–8. [Abstract] [Google Scholar]
  • Young JC, Obermann WM, Hartl FU. Specific binding of tetratricopeptide repeat proteins to the C-terminal 12-kDa domain of hsp90. J Biol Chem. 1998;273:18007–10. [Abstract] [Google Scholar]
  • Young JC, Hoogenraad NJ, Hartl FU. Molecular chaperones Hsp90 and Hsp70 deliver preproteins to the mitochondrial import receptor Tom70. Cell. 2003a;112:41–50. [Abstract] [Google Scholar]
  • Young JC, Barral JM, Ulrich Hartl F. More than folding: localized functions of cytosolic chaperones. Trends Biochem Sci. 2003b;28:541–7. [Abstract] [Google Scholar]
  • Yu S, Xu Z, Zou C, Wu D, Wang Y, Yao X, Ng CF, Chan FL. Ion channel TRPM8 promotes hypoxic growth of prostate cancer cells via an O2-independent and RACK1-mediated mechanism of HIF-1alpha stabilization. J Pathol. 2014;234:514–25. [Abstract] [Google Scholar]
  • Zhang S, Sun Y, Yuan Z, Li Y, Li X, Gong Z, Peng Y. Heat shock protein 90beta inhibits apoptosis of intestinal epithelial cells induced by hypoxia through stabilizing phosphorylated Akt. BMB Rep. 2013;46:47–52. [Europe PMC free article] [Abstract] [Google Scholar]
  • Zhang SL, Yu J, Cheng XK, Ding L, Heng FY, Wu NH, Shen YF. Regulation of human hsp90alpha gene expression. FEBS Lett. 1999;444:130–5. [Abstract] [Google Scholar]
  • Zhang T, Li Y, Yu Y, Zou P, Jiang Y, Sun D. Characterization of celastrol to inhibit hsp90 and cdc37 interaction. J Biol Chem. 2009;284:35381–9. [Europe PMC free article] [Abstract] [Google Scholar]
  • Zhao Y, Tao L, Jiang D, Chen X, Li P, Ning Y, Xiong R, Liu P, Peng Y, Zhou YG. The −144C/A polymorphism in the promoter of HSP90beta is associated with multiple organ dysfunction scores. PLoS One. 2013;8:e58646. [Europe PMC free article] [Abstract] [Google Scholar]
  • Zhong X, Shen Y, Ballar P, Apostolou A, Agami R, Fang S. AAA ATPase p97/valosin-containing protein interacts with gp78, a ubiquitin ligase for endoplasmic reticulum-associated degradation. J Biol Chem. 2004;279:45676–84. [Abstract] [Google Scholar]
  • Zhou Y, Xu L, Song X, Ding L, Chen J, Wang C, Gan Y, Zhu X, Yu Y, Liang Q. The potential role of heat shock proteins in acute spinal cord injury. Eur Spine J. 2014;23:1480–90. [Abstract] [Google Scholar]
  • Zou J, Guo Y, Guettouche T, Smith DF, Voellmy R. Repression of heat shock transcription factor HSF1 activation by HSP90 (HSP90 complex) that forms a stress-sensitive complex with HSF1. Cell. 1998;94:471–80. [Abstract] [Google Scholar]
  • Zurawska A, Urbanski J, Matuliene J, Baraniak J, Klejman MP, Filipek S, Matulis D, Bieganowski P. Mutations that increase both Hsp90 ATPase activity in vitro and Hsp90 drug resistance in vivo. Biochim Biophys Acta. 2010;1803:575–83. [Abstract] [Google Scholar]

Citations & impact 


Impact metrics

Jump to Citations
Jump to Data

Citations of article over time

Smart citations by scite.ai
Smart citations by scite.ai include citation statements extracted from the full text of the citing article. The number of the statements may be higher than the number of citations provided by EuropePMC if one paper cites another multiple times or lower if scite has not yet processed some of the citing articles.
Explore citation contexts and check if this article has been supported or disputed.
https://scite.ai/reports/10.1016/j.gene.2015.08.063

Supporting
Mentioning
Contrasting
0
73
0

Article citations


Go to all (70) article citations

Data 


Data behind the article

This data has been text mined from the article, or deposited into data resources.

Funding 


Funders who supported this work.

NIGMS NIH HHS (2)

National Institutes of Health (1)