Hsp90 inhibitor as a sensitizer of cancer cells to different therapies (Review)

  • Authors:
    • Zuzana Solárová
    • Ján Mojžiš
    • Peter Solár
  • View Affiliations

  • Published online on: December 10, 2014     https://doi.org/10.3892/ijo.2014.2791
  • Pages: 907-926
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Hsp90 is a molecular chaperone that maintains the structural and functional integrity of various client proteins involved in signaling and many other functions of cancer cells. The natural inhibitors, ansamycins influence the Hsp90 chaperone function by preventing its binding to client proteins and resulting in their proteasomal degradation. N- and C-terminal inhibitors of Hsp90 and their analogues are widely tested as potential anticancer agents in vitro, in vivo as well as in clinical trials. It seems that Hsp90 competitive inhibitors target different tumor types at nanomolar concentrations and might have therapeutic benefit. On the contrary, some Hsp90 inhibitors increased toxicity and resistance of cancer cells induced by heat shock response, and through the interaction of survival signals, that occured as side effects of treatments, could be very effectively limited via combination of therapies. The aim of our review was to collect the data from experimental and clinical trials where Hsp90 inhibitor was combined with other therapies in order to prevent resistance as well as to potentiate the cytotoxic and/or antiproliferative effects.

1. Introduction

The cell responds to environmental stress by increasing synthesis of several molecular chaperons known as heat shock proteins (Hsp). These proteins are cathegorized according to their molecular weight into five classes: small Hsp, Hsp60, Hsp70, Hsp90 and Hsp100. One of the most abundant molecular chaperones is Hsp90, which is a highly conserved protein, whose association is required for the stability and function of multiple-mutated, chimeric- and overexpressed-signaling proteins that promote the growth and/or survival of cancer cells (1). Hsp90 client proteins lack specific Hsp90 binding motifs and vary in terms of intracellular localization, structure, and function. Their main common denominator is their role in promoting cell proliferation and protection from apoptosis (2). Hsp90 is involved in the maturation and stabilization of a wide range of oncogenic client proteins which are crucial for oncogenesis and malignant progression (3), such as signal transduction molecules (Src, Raf1 and cdk4) and steroid receptors and in enhancing the activity of telomerase and nitric-oxide synthase (4). For the updated list of Hsp90 client proteins see http://www.picard.ch/download. Under normal growth conditions, Hsp90 plays a major role in various aspects of the secretory pathway, cellular transport and during environmental stress. Hsp90 is required for the cell cycle, meiosis and cytokinesis (5), however, Hsp90 does not participate in general protein folding (6). Recent studies also indicate that many types of cells express Hsp90 on the cell surface and secrete Hsp90 into the extracellular space to carry out important extracellular functions (7).

2. Hsp90 inhibitors - potential therapeutic agents

There are several reasons why Hsp90 should be considered an important target for cancer therapy. One is that Hsp90 protein is involved in the maturation and stabilization of a wide range of oncogenic proteins (Bcr-Abl, HER-2, EGFR, C-Raf, B-Raf, Akt, Met, VEGFR, FLT3, AR and ER, HIF-1α) and telomerase (8), which are included in tumor formation and therefore causing certain dependence of cancer cells on Hsp90 function. Secondly, unsuitable environmental conditions in cancer cells (hypoxia, low pH, poor nutritional conditions) may result in the destabilization of proteins, and therefore the function of Hsp90 for such cells is necessary. In addition, Hsp90 protein exists as a multi-chaperone complex with unusually high affinity for ATP and some substances in tumor cells (8), whereas in normal cells a latent form of Hsp90 protein is present. In this regard, Hsp90 protein of tumor cells has higher affinity for inhibitors (9), caused by a change in binding of ATP to Hsp90 as an effect of co-chaperones, that modulate the function of Hsp90, as well as by higher ATPase activity of Hsp90 protein, whereas in normal cells Hsp90 is found mostly in a latent or inactive form (10). Furthermore, Hsp90 is constitutively, and 2–10 times more expressed in tumor cells than in normal ones, which means, that the correct function of Hsp90 is essential for the growth and survival of tumor cells (11). Hsp90 inhibition may lead to the degradation of proteins involved in the so-called six main characteristics of cancer: i) the ability to produce growth factors, ii) resistance to anticancer agents, iii) avoidance of apoptosis, iv) unlimited replicative potential, v) uninterrupted angiogenesis and vi) invasiveness and metastasis (8,12,13).

The Hsp90 monomer is composed of four domains: a highly conserved N- and C-terminal domain, a middle domain and a charged linker region that connects to both N-terminal as well as middle domain (14). A long, highly charged linker between the N-terminal domain and the middle domain is a special characteristic of Hsp90s in eukaryotic organisms (15). N-terminal domain contains ATP-binding site, thus blocking of ATP binding can serve as a single well known mechanism of Hsp90 inhibitors. Middle domain is responsible for co-chaperone and clients binding and its function is probably discrimination between various types of client proteins during their processes of activation (16). C-terminal domain is used for dimerization of Hsp90 and it is supposed, that it contains a second nucleotide-binding site, which has been shown to bind to novobiocin, epigallocatechin (ECGC) and taxol. However, neither the apo-form crystal structure, nor any complex structure verified this prediction (17).

3. Inhibitors of the N-teminal ATP binding pocket

Hsp90 inhibitors are divided into several groups according to the form of inhibition, which includes: i) blocking the binding of ATP, ii) decoupling co-chaperon/Hsp90 and iii) antagonism of client/Hsp90 association. The benzoquinone ansamycins mimic ATP and bind to the nucleotide-binding pocket on the N-terminus of Hsp90, blocking the natural substrate ATP binding (18). Geldanamycin (GA) and its analogs specifically inhibit only Hsp90 (and its endoplasmic reticulum homolog grp94), but not other chaperones or other intracellular molecular targets (19). Hsp90 inhibitor 17-allylamino-17-demethoxygeldanamycin (17-AAG) administration was shown to induce a pronounced downregulation of multiple Hsp90 protein clients and other downstream effectors, such as IGF-IR, Akt, IKK-α, IKK-β, FOXO1, Erk1/2 and c-Met, resulting in sequestration-mediated inactivation of NF-κB, reduced cell proliferation and decline of cell motility (20). The inhibition of Hsp90 with 17-AAG and another ansamycin derivative 17-(dimethylaminoethylamino)-17-demethoxygeldanamycin (17-DMAG) downregulated B-Raf, decreased cell proliferation, reduced Mek/Erk signaling activation and induced a significant inhibition of telomerase activity in melanoma cells (21,22). On the contrary, treatment with 17-AAG triggers a stress response in cells that leads to activation of the transcription factor of Hsp-1 (HSF-1), resulting in elevation of anti-apoptotic Hsp90, Hsp70, and Hsp27 chaperons (23).

Although GA exhibits very strongly in vitro antitumor activities, its clinical potential is unfortunately low, due to poor solubility in aqueous solutions and its hepatotoxicity (24). Therefore, GA derivatives with comparable antitumor effects and even better toxicological properties have been synthesized. Following derivatives of GA have entered into the clinical trials (per August 2014): 17-AAG (tanespimycin; CNF1010 - an oil-in-water nanoemulsion of 17-AAG; 39 trials), 17-DMAG (alvespimycin; 7 trials), IPI-504 (retaspimycin hydrochloride; 13 trials) and most recently 17-amino-17-demethoxygeldanamycin (17-AG or IPI-493; 2 trials). 17-AAG is one of the derivatives which has entered to the third phase of the clinical trials. Furthermore, there are phenol derivatives produced by Kosan and Biotica, which demonstrated cell growth inhibition comparable to 17-AAG and 17-DMAG (25). These compounds are in preclinical stages of development.

Another intensively studied N-terminal domain binder of Hsp90 is radicicol, a macrocyclic natural antibiotic isolated from the fungus Monocillium nordinii and from the plant-associated fungus Chaetomium chiversii (26). Despite the potent antitumoral activity in vitro, radicicol has little or no activity in vivo (27,28). On the contrary, newly created stable oxime derivatives of radicicol and cykloproparadicicol have shown antitumor activity with tolerable toxicity in animal models (28). Moreover, synthetic Hsp90 inhibitors with resorcinol moiety NVP-AUY922 (27 trials) and AT-13387 (6 trials) are now in phase I–II and phase I clinical trials, respectively.

With the aim to improve pharmacological properties and even increase the specificity of Hsp90 inhibitors, many phamacological companies focused on the synthesis of small molecules. Chiosis et al (29) demonstrated the effect of the first identified purine scaffold inhibitor PU3 comparable to GA in breast cancer cells. Kasibhatla et al (30) modified such a purine inhibitor by rotating the C8-attached aryl ring to the N9 position. This resulted in the compound CNF2024/BIIB021 identified as a stronger inhibitor compared to 17-AAG in vitro and in vivo (31), which is currently involved in phase II clinical trials (7 trials). Interestingly, purine-scaffold inhibitors have better solubility in aqueous solutions, metabolic stability and can be administered orally compared to ansamycin Hsp90 inhibitors (32). PU-H71 derivative is now in phase I clinical trials (2 trials).

Using a chemical proteomics approach, Serenex company identified small molecules, that are distinct from GA (structures not disclosed). Water soluble and orally active compound SNX-5422, administered in xenograft models, resulted in partial tumor regression (33) and is currently involved in phase I clinical trials (10 trials).

Synta Pharmaceuticals Corporation developed another second-generation Hsp90 inhibitor ganetespib (formerly STA-9090), a unique resorcinolic triazolone (chemical structure distinct from 17-AAG), undergoing now phases II–III clinical trials (32 trials).

Couple of additional Hsp90 inhibitors have been described and entered phase I–II of clinical trials. Details as follows: KW-2478 (2 trials), Hsp990 (2 trials), MPC-3100 (1 trial), Debio 0932 (2 trials), XL888 (2 trials), BIIB028 (1 trial) and DS-2248 (1 trial). The outcome of phase I clinical trials with Hsp90 inhibitors are well summarized in reviews of Jhaveri et al (34,35). For detailed published results of phase II studies see Table I.

Table I

Published phase II study results of Hsp90 inhibitors.

Table I

Published phase II study results of Hsp90 inhibitors.

Hsp90 inhibitorResponses of cancer patientsAuthors (Ref.)
17-AAGNo objective response in the treatment of clear cell or papillary renal cell carcinomaRonnen et al, 2006 (163)
17-AAGNo activity with regard to prostate-specific antigen responseHeath et al, 2008 (164)
17-AAGNo objective antimelanoma responsesSolit et al, 2008 (90)
IPI-504Activity in NSCLS patients, particularly among those with ALK rearrangementsSequist et al, 2010 (165)
IPI-504Not warranted in patients with castration-resistant prostate cancerOh et al, 2011 (166)
17-AAG + bortezomibSignificant and durable responses with acceptable toxicity (relapsed or relapsed and refractory multiple myeloma)Richardson et al, 2011 (167)
17-AAG + trastuzumabSignificant anticancer activity in patients with HER-2-positive, metastatic breast cancer previously progressing on trastuzumabModi et al, 2011, (168)
17-AAGFurther study not recommended (metastatic or locally advanced, unresectable breast cancer)Gartner et al, 2012 (169)
17-AAGStudy endpoints were not reached because of early termination (further study in melanoma should be considered)Pacey et al, 2012 (170)
STA-9090Activity in NSCLS patients, particularly among those with ALK rearrangementsSocinski et al, 2013 (171)
IPI-504 + trastuzumabModest clinical activity in patients with advanced or metastatic HER-2-positive breast cancerModi et al, 2013 (172)
STA-9090Activity was observed in trastuzumab-refractory HER2-positive and TNBC cancer patientsJhaveri et al, 2014 (173)

4. Inhibitors of the C-terminal ATP binding pocket

Unsatisfactory, results with N-terminal inhibitors in experimental and clinical trials arising from the increased toxicity and resistance induced by heat shock response led to development and characterization of a novel C-terminal inhibitors of Hsp90. Novobiocin (antiobiotic of coumarin group) and its analogues (clorobiocin and coumermycin A1) have been generated as agents for the treatment of bacterial infections with multi-resistant gram-positive bacteria such as Staphylococcus aureus and S. epidermidis (36,37). Coumarin antibiotics were reported to bind strongly to gyrase B (DNA topoisomerase II) (38) but also weakly to the Hsp90 C-terminal nucleotide-binding site (39) followed by induced degradation of Hsp90 client proteins such as v-Src, Raf-1, ErbB2 and mutant p53 (39). Moreover, novobiocin disrupted the interaction of both the cochaperons p23 and Hsp70 with the Hsp90 complex (40). Structural modification of this compound has led to analogues with 1,000-fold greater efficacy in antiproliferative assays against various cancer cell lines (4143). In this regard, 3-aminocoumarin analogue with introduced tosyl substituent on C-4 position of coumarin nucleus (4TCNA) promoted apoptosis through activation of caspases 7 and 8 in MCF-7 human breast cancer cells. Furthermore, 4TCNA induced cleavage of p23 which plays an important role in activity of number of transcription factors of steroids/thyroid receptors family (40). The same group of scientists identified a new family of novobiocin analogues in which the coumarin unit was replaced by a 2-quinoleinone moiety. Compound 4-tosyl-3[(chroman-6-yl)carboxylamino]-2-quinolon (4TCCQ) was 100-fold more potent than the parental natural novobiocin and 6-fold more active than the synthetic analogue 4TCNA. Additionally, 4TCCQ induced the degradation of estrogen receptor α and strongly induced the cell death in MCF-7 cells (40). Compound A4, lacking the 4-hydroxyl of the coumarin moiety and containing an N-acetyl side chain in lieu of the benzamide, was the most active compound, which induced degradation of Hsp90-dependent client proteins at 70-fold lower concentration than novobiocin (40). 4-hydroxy moiety of the coumarin ring and the 3′-carbamate of the noviose appendage were detrimental to Hsp90 inhibitory activity. In an effort to confirm these findings, 4-deshydroxy novobiocin (DHN1) and 3′-descarbamoyl-4-deshydroxynovobiocin (DHN2) were prepared and evaluated against Hsp90. Both compounds were significantly more potent than the natural product (44).

Novel novobiocin-derived C-terminal Hsp90 inhibitor KU135 induced antiproliferative effects in Jurkat T-lymphocytes, caused the degradation of known Hsp90 client proteins and revealed more potent antiproliferative effects than the known N-terminal Hsp90 inhibitor 17-AAG. Moreover, KU135 in comparison to 17-AAG was found to be a potent inducer of mitochondria-mediated apoptosis as evidenced, in part, by the fact that cell death was inhibited to a similar extent by Bcl-2/Bcl-xL overexpression or the depletion of apoptotic protease-activating factor-1 (Apaf-1). Intriguingly, KU135 did not induce upregulation of Hsp70 and Hsp90 in Jurkat T-lymphocytes (45). The study of Samadi et al (46) demonstrated similar effect of KU135 in most melanoma cells, where it reduced cell survival, proliferation, and induced apoptosis more significantly than 17-AAG. Additionally, levels of Hsp90 and Hsp70 did not increase, while the levels of phosphorylated HSF-1 levels decreased after KU135 treatment of melanoma cells. KU135 inhibited cell proliferation by regulating signaling pathways, which are different from those targeted by 17-AAG and as such represents a novel opportunity for Hsp90 inhibition. In this regard, KU135 seems to be a potential candidate for cancer therapy against melanoma. Recently developed C-terminal Hsp90 inhibitor KU174 exhibited very robust antiproliferative and cytotoxic activity together with client protein degradation and disruption of Hsp90 native complexes without induction of an HSR in prostate cancer cells. In a pilot in vivo proof-of-concept study KU174 demonstrated efficacy at 75 mg/kg in a PC3-MM2 rat tumor model (47).

Furthermore, novel novobiocin analogue F-4 as a putative C-terminal inhibitor of Hsp90 also demonstrated improved potency and efficacy compared to novobiocin in antiproliferative assays and decreased the expression of client proteins. Prostate specific antigen (PSA) secretion was inhibited in a dose-dependent manner that paralleled a decrease in androgen expression (AR). In addition, superior efficacy was demonstrated by F-4 compared to 17-AAG in experiments measuring cytotoxicity and apoptosis (48). Recently, assisted by molecular docking studies, a scaffold containing a biphenyl moiety in lieu of the coumarin ring system found in novobiocin, was identified for development of new Hsp90 C-terminal inhibitors. A library of small molecules containing the biphenyl moiety was designed, synthetised, and evaluated against two breast cancer cell lines with good antiproliferative activity manifested (49).

The most abundant polyphenolic catechin with chemopreventive and anticancer activities is epigallocatechin-3-gallate (EGCG) (50). Palermo et al (51) demonstrated that EGCG can inhibit the transcriptional activity of aryl hydrocarbon receptor (AhR) through a mechanism involving direct binding to the C-terminal region of Hsp90. A novel Hsp90 inhibitor impaired the association of Hsp90/Hsc70 and Hsp90/p23, by directly binding to the C-terminal region of Hsp90, inhibited Hsp90 chaperoning function, and simultaneously degraded multiple cancer-related Hsp90 client proteins (52,53). EGCG specifically inhibited the expression of Hsp70 and Hsp90 by inhibiting the promoter activity of Hsp70 and Hsp90. Pretreatment with EGCG increased the stress sensitivity of MCF-7 cells upon heat shock (44°C for 1 h) or oxidative stress (H2O2, 500 μM for 24 h). Moreover, treatment with EGCG (10 mg/kg) in a xenograft model resulted in delayed tumor incidence and reduced tumor size, as well as the inhibition of Hsp70 and Hsp90 expression (54).

5. Combination of Hsp90 inhibitor with second therapy

The interaction of survival signals that occured as side effects of treatments could be very effectively limited through combination of therapies. In this regard, Hsp90 inhibitors may be a promising strategy e.g., to sensitize cisplatin-resistant Fancony anemia pathway-proficient tumor cells to cisplatin (55), to overcome autophagy, a mechanism protecting rhabdomyosarcoma cells from drug-induced cytotoxicity (56), to sensitize human tumors to irradiation (57) or simply to enhance the effect of therapy when combined with the non-genotoxic Hsp90 inhibitor (58).

Oral, esophageal, gastrointestinal, colon and pancreatic cancer

The Hsp90 inhibitor 17-AAG enhanced significantly the radiation sensitivity and induced apoptosis in oral squamous cell carcinoma cell line SAS/neo which has a wild-type p53. On the other hand, the radiation sensitizing effect of 17-AAG was limited in the SAS/Trp248 cells which has a mutated p53 (59). 17-AAG was also shown to inhibit ATP-binding cassette sub-family G member 2 (ABCG2) upregulation, thereby reversing the ABCG2-mediated multidrug resistance (MDR) and in this regard can be used as a chemosensitizer for the treatment of esophageal cancer (60). Furthermore, 17-AAG sensitized esophageal cancer cells to γ-photon radiation, probably through its ability to affect a range of signalling components simultaneously, for example both EGF and IGF-1 (61).

Hsp90 inhibitors, IPI-493 as well as IPI-504, revealed antitumor activity and induced receptor tyrosine kinase KIT downregulation in xenograft model of gastrointestinal stromal tumor (GIST) with heterogeneous KIT mutations (62,63). IPI-493 synergized with tyrosine kinase inhibitors, that are commonly used for the treatment of advanced or imatinib-resistant GIST (63). Treatment effects of IPI-504 were also enhanced and/or more potent in combination with imatinib or sunitinib (62). Similarly a combination of imatinib and Hsp90 inhibitor AT13387 treatment in the imatinib-resistant GIST430 model significantly enhanced tumor growth inhibition over the monotherapies. These results, as well as good tolerance of AT13387, suggest this Hsp90 inhibitor as an excellent candidate for clinical testing in gastrointestinal stromal tumors in combination with imatinib (64). Moreover, by blocking Hsp90, Lang et al (65) disrupted rapamycin-induced activation of alternative signaling pathways in hepatocellular carcinoma and substantially improved the growth-inhibitory effects of mTOR inhibition in vivo.

Hsp90 inhibition may be also an attractive target for innovative treatment of gastroenteropancreatic neuroendocrine tumors alone or in combination with either IGF-1R or mTOR inhibitors (66). Unusually, Hsp90 inhibitor 17-AAG was more responsive in colon cancer cells with BRAFV600E mutation than selective BRAFV600E inhibitor PLX4720, probably due to the multiple oncogenic proteins that are Hsp90 clients (67). Indeed, the inhibition of Hsp90 with 17-DMAG abrogated the invasive properties of colon cancer cells and modulated the expression of the antimetastatic factor activating transcription factor-3 and improved the efficacy of oxaliplatin (68). Furthermore, 17-AAG interventions directed to cell cycle arrest in G1 may potentiate oxaliplatin activity in colon cancer cells with impaired p53 (69) and may be critical for its ability to reverse cisplatin resistance (70). Similarly, GA reversed topotecan induced upregulation of the anti-apoptotic protein Bcl-2 in both p53+/+ and p53−/− HCT116 cells and sensitised human colon cancer cells to topoisomerase I poisons via depletion of key anti-apoptotic and cell cycle checkpoint proteins (71). On the other hand, the combination of an active metabolite of irinotecan SN-38 and 17-AAG was shown to be synergistic in p53-null, but not in parental HCT116 cells by median effect. Taken together, 17-AAG specifically inhibited the G(2)/M checkpoint in p53-defective cells by downregulation of checkpoint kinases Chk1 and Wee1 (72). Similarly, GA enhanced the effect of ionizing radiation in p53 and p21 null cells (73).

The inhibition of Hsp90 with GA destabilised Cdc25A independent of Chk1/2, whereas the standard drug for pancreas carcinoma treatment, gemcitabine caused Cdc25A degradation through the activation of Chk2. Both agents applied together additively inhibited the expression of Cdc25A and the proliferation of pancreatic carcinoma cells, as well as reduced the resistance of pancreatic carcinoma cells to treatment with gemcitabine (74). Indeed, the combined therapy of 17-DMAG and Akt inhibitor perifosine overcame tumor growth and resistance induced by bone marrow stromal cells and endothelial cells as well as the proliferative effect of exogenous interleukin-6, insulin-like growth factor-I, and vascular endothelial growth factor. The combination induced also apoptosis and growth inhibition in endothelial cells and inhibited angiogenesis (75). Moreover, short interfering RNA (siRNA) was used to knockdown ErbB3 in the resistant cell line AsPC1. Whereas individual treatments with siRNA to ErbB3 or 17DMAG had no effect on radiosensitivity, the combination, which reduced both ErbB2 and ErbB3, resulted in a significant enhancement in AsPC1 radiosensitivity (76). Interestingly, the dietary component from broccoli and broccoli sprouts sulforaphane also potentiated the efficacy of 17-AAG against pancreatic cancer through enhanced abrogation of Hsp90 function (77). For the complete list of oral, esophageal, pancreatic, gastrointestinal and colon cancer cell response to combination of Hsp90 inhibitor and other therapy (treatment) in vitro and/or in vivo see Table II.

Table II

The response of oral, esophageal, pancreatic, gastrointestinal and colon cancer cells to combination of Hsp90 inhibitor with another therapy in vitro and/or in vivo.

Table II

The response of oral, esophageal, pancreatic, gastrointestinal and colon cancer cells to combination of Hsp90 inhibitor with another therapy in vitro and/or in vivo.

CancerHsp90 inhibitorSecond therapyConditionsTumor response(Ref.)
Oral squamous17-AAGIonizing radiationIn vitro↑ p53-dependent(59)
Esophageal17-AAGIonizing radiationIn vitro(61)
Esophageal17-AAGCisplatinIn vitro(174)
EsophagealRadicicolHeatIn vitro(175)
PancreaticGA5-FluorouracilIn vitro↑ in presence of serum(176)
PancreaticGAGemcitabineIn vitro(74)
PancreaticGA, 17-AAG 3-BromopyruvateIn vitro, in vivo(177,178)
Pancreatic17-AAGU0126In vitro(179)
Pancreatic17-AAGSulforaphaneIn vitro, in vivo(77)
Pancreatic and other17-DMAGIonizing radiationIn vitro↑ predicted by ErbB3 expression(76)
Hepatocellular17-DMAGRapamycinIn vitro, in vivo↑ + ECs, and VSMCs(65)
Pancreatic, glioblastomaNVP-HSP990Ionizing radiation, hyperthermiaIn vitro(180)
Gastrointestinal stromalIPI-504Imatinib, sunitinibIn vivo(62)
Gastroentero-pancreaticIP-504IGF1R and mTOR inhibitorsIn vitro, in vivo(66)
Gastrointestinal stromalIP-493Imatinib, sunitinibIn vivo(63)
Gastrointestinal stromalAT13387ImatinibIn vitro, in vivo(64)
ColonGAIonizing radiationIn vitro↑ in p53 compromised cells(73)
ColonGATopotecanIn vitro↑ p53-independent(71)
Colon and otherGAIonizing radiationIn vitro↑ more in SQ-5 than in DLD-1 cells(181)
Colon17-AAGTRAILIn vitro↑ greater in TRAIL-resistant cells(182)
Colon17-AAGAHA1 siRNAIn vitro(183)
Colon17-AAGCisplatinIn vitro↑ in cells with sustained JNK(70)
Colon17-AAGOxaliplatinIn vitro↑ in p53 deficient cells(69)
Colon17-AAGIrinotecan SN-38In vitro↑ in p53 defective cells(72)
Colon17-AAGTRAILIn vitro↑ in BRAF (V600E) mutant cells(67,184)
Colon and other17-AAGNutlinIn vitro, in vivo(58)
Colon17-DMAGOxaliplatinIn vitro, in vivo↑ in p53 deficient cells(68)
ColonSTA-9090Ionizing radiation, 5-fluorouracilIn vitro, in vivo(185)

[i] The symbol ↑ means at least the additive effect of the combination compared to single therapies.

Lung cancer

Hsp90 inhibitor IPI-504, induced tumor regression in aggressive Nf1-deficient malignancies and KRas/p53-mutant lung cancer mouse models, but only when combined with rapamycin (78). Combinations of low-dose of non-ansamycin inhibitor of Hsp90, ganetespib (STA-9090 a unique resorcinolic triazolone) with MEK or PI3K/mTOR inhibitors, resulted in superior cytotoxic activity than single agents alone in a subset of mutant KRAS cells, and the anti-tumor efficacy of STA-9090 was potentiated by cotreatment with the PI3K/mTOR inhibitor BEZ235 in A549 xenografts in vivo (79). Similar combination of STA-9090 with paclitaxel, docetaxel or vincristine resulted in synergistic antiproliferative effects in the H1975 cells in vitro as wells as enhanced tumor growth inhibition observed in combination with paclitaxel and tumor regressions seen with docetaxel (80). Furthermore, Hsp90 inhibitors have overcame the resistance to EGFR tyrosine kinase inhibitors in subcutaneous xenograft models of H1975 and A549 (81) and enhanced the antitumor activity of paclitaxel and camptothecin, respectively, in breast and colorectal tumor models (82). Hsp90 inhibition overcame the hepatocyte growth factor-triggered resitance to EGFR tyrosine kinase inhibitors and resulted in more successful treatment of patients with EGFR-mutant lung cancers (83). The effectiveness of combined therapy of EGFR tyrosine kinase inhibitors and Hsp90 inhibitors was also confirmed in the treatment of lung cancers co-driven by mutant EGFR containing T790M and MET (84,85). Moreover, 17-AAG decreased the etoposide-induced p38 MAPK-mediated ERCC1 expression (DNA repair capacity) (86) and downregulated the cisplatin-induced thymidine phosphorylase (a key enzyme in pyrimidine nucleoside salvage pathway) expression and Erk1/2 and Akt activation (87) what sensitized lung cancer cells to etoposide and cisplatin, respectively.

Hsp90 inhibitors were also effective in treating crizotinib-resistant tumors harboring secondary gatekeeper mutations within the ALK TK domain (88). Furthermore, Normant et al (89) observed partial responses to administration of IPI-504 as a single agent in a phase II clinical trial in patients with NSCLC, specifically in patients that carry an ALK rearrangement. The experimental data of Solit et al (90) suggested that Hsp90 inhibition in combination with chemotherapy may represent an effective treatment strategy for patients, whose tumors express EGFR kinase domain mutations, including those with de novo and acquired resistance to EGFR tyrosine kinase inhibitors. Acquired resistance limits the efficacy of ALK inhibitors, such as crizotinib and TAE684, however, combining inhibitors of ALK and Hsp90 resulted in synergistic cytotoxicity (91,92).

Hsp90 inhibitor 17-DMAG, as well as purine-scaffold inhibitor PU-H71, were shown to be the most effective as radiosensitizer of lung cancer cells, when administered before radiation, because of its suppression of DNA repair at multiple levels, including BER and ATM-regulated pathways (93,94). Recent results of Kim and Pyo (95) show that combined treatment of 17-AAG and celecoxib at clinically relevant concentrations may significantly enhance the therapeutic efficacy of ionizing radiation therapy. Moreover, very recent combination of 17-DMAG and histone deacetylase inhibitor belinostat synergistically inhibited in vitro proliferation of selected panel of 12 NSCLC cell lines. Importantly, both agents and their combination almost completely prevented TKI-resistant tumor formation (EGFR T790M mutation) in a xenograft model (96). For the above, and even further studies with the response of lung cancer cells to combination of Hsp90 inhibitor and another therapy see Table III.

Table III

The response of lung cancer cells to combination of Hsp90 inhibitor with another therapy in vitro and/or in vivo.

Table III

The response of lung cancer cells to combination of Hsp90 inhibitor with another therapy in vitro and/or in vivo.

CancerHsp90 inhibitorSecond therapyConditionsTumor response(Ref.)
NSCLC17-AAGPaclitaxelIn vitro↑ in cells with high levels of p185(186)
NSCLC17-AAGCisplatinIn vitro(87)
NSCLC17-AAGEtoposideIn vitro(86)
NSCLC17-AAGVER-155008In vitro(187)
NSCLC17-DMAGIonizing radiationIn vitro(93)
NSCLC17-DMAGPTACH, belinostatIn vitro, in vivo(96,188)
NSCLC17-DMAGWZ4002in vivo↑ in EGFR (T790M) mutant cells(84)
NSCLC17-DMAGVelcadeIn vitro(189)
NSCLCSTA-9090Paclitaxel, docetaxel, vincristineIn vitro, in vivo(80)
NSCLCSTA-9090BEZ235In vitro, in vivo↑ in KRAS mutant cells(79)
NSCLCSTA-9090CrizotinibIn vitro, in vivo↑ in ALK- and MET-driven cells(92,190,191)
NSCLCSNX-2112
STA-9090
ErlotinibIn vitro, in vivo↑ in wild-type or mutant
EGFR (L858R, T790M) cells
(85,192)
NSCLCCH5164840ErlotinibIn vitro, in vivo↑ in EGFR overexpressed and mutant cells(193)
NSCLCCUDC-305PaclitaxelIn vitro, in vivo↑ in EGFR (T790M) or KRAS mutant cells(81)
NSCLCWK88-1GefitinibIn vitro, in vivo↑ in MET-driven cells(194)
LungGATRAILIn vitro↑ in TRAIL-resistant cells; adverse effects of high-dose GA(195)
Lung and otherGA, 17-AAGIonizing radiation, celecoxibIn vitro, in vivo↑ in transformed cells only(95,196198)
Lung17-AAGPaclitaxelIn vitro, in vivo↑ in EGFR mutant cells(199)
Lung and other17-AAGSmac mimetic compound 3In vitro(200)
Lung17-DMAGCrizotinibIn vitro↑ EML4-ALK-expressing cells(91)
LungIP-504RapamycinIn vitro, in vivo↑ in RAS-driven cells(78)
LungPU-H71Ionizing radiationIn vitro(194)

[i] The symbol ↑ means at least the additive effect of the combination compared to single therapies. NSCLC, non-small cell lung cancer.

Breast, ovary and prostate cancer

Hsp90 inhibitor IPI-504 resulted in a marked decrease in the levels of HER-2, Akt, p-Akt, and p-MAPK in trastuzumab-resistant xenografts as early as 12 h after a single dose of IPI-504 inhibitor. IPI-504-mediated Hsp90 inhibition may represent, together with other Hsp90 inhibitors, a novel therapeutic approach in HER-2 overexpressing and/or trastuzumab refractory HER-2-positive cancer (97100), and provide a promising strategy to overcome the development of resistance against trastuzumab. Very promising strategy to overcome the development of trastuzumab resistance in breast cancer patients seems to be a combination of the Hsp90 inhibitor and histone deacetylase 6 inhibitor carbamazepine (101). GA synergized with carbamazepine to promote HER-2 degradation and inhibited breast cancer cell proliferation. Solit et al (102) found that the inhibition of Hsp90 function sensitized breast cancer xenografts to Taxol through downregulation of Akt kinase. Hsp90 inhibitors may be an effective therapy also to treat aromatase inhibitor-resistant breast cancers and improved efficacy can be achieved by combined use of an Hsp90 inhibitor and an Akt inhibitor (103,104). Moreover, Pashtan et al (105) demonstrated that Hsp90 inhibitor 17-AAG inhibiting multiple signaling activators including ErbB and Src kinases, it does not permit oncogene switching and resulted in a more prolonged and robust inhibition of downstream signaling pathways in breast cancer cells, than the individual tyrosine kinase inhibitors. In this way, Hsp90 inhibitors can prevent breast cancer cells from tyrosine kinase inhibition (105). Another way, to sensitize cancer cells is elimination of RIP1 protein, which can support cell survival due to activation of NF-κB. In this regard, 17-DMAG markedly reduced RIP1 expression and sensitized breast tumor cells to tumor necrosis factor-related apoptosis-inducing ligand (TRAIL)-induced apoptosis (eliminate the resistance to TRAIL) (106). A recent study of Stecklein et al (107) demonstrated the role of BRCA1, which is associated with the resistance to platinum-based chemotherapeutics and poly(ADP ribose) polymerase (PARP) inhibitors in regulating damage-associated checkpoint and repair responses to Hsp90 inhibitors. The above study pointed out BRCA1 as a clinically relevant target for enhancing sensitivity in refractory and/or resistant malignancies. Furthermore, complete regression of aggressive breast tumors in mice was observed after the combination of anti-angiogenic stress imposed by loss of Id protein and the inactivation of the HIF-1α chaperone Hsp90 (108).

Interestingly, ovarian cancer cells overexpressing HER2 showed 5-fold increased sensitivity to the Hsp90 molecular chaperone inhibitor GA. In contrast, HER2-overexpressing cells showed statistically significant resistance to cisplatin, PI3K inhibitor LY294002 and tyrosine kinase inhibitor emodin (109). On the other hand, GA and 17-AAG may sensitize ovarian cancer cells to cisplatin (110) and paclitaxel (111), respectively, particularly those tumors in which resistance is driven by HER-2 and/or p-Akt.

Pharmacological inhibition of the cell cycle-dependent kinase (essential for G2-M checkpoint) Wee1 and its molecular silencing with siRNA uniformly increased apoptotic activity of the Hsp90 inhibitor 17-AAG in vitro (112). Moreover, Wee1 inhibition synergized with any one of the clinically evaluated Hsp90 inhibitors to inhibit an androgen-independent and invasive human prostate carcinoma cell line PC3 in vitro and in vivo (113). Hsp90 inhibitors trigger a heat shock response associated with increased expression of Hsp90, Hsp70, Hsp27, and clusterin that attenuates drug effectiveness in the treatment of castrate-resistant prostate cancer. However, clusterin inhibitor OGX-011 revealed markedly potentiated antitumor efficacy in xenograft models of human castrate-resistant prostate cancer, leading to an 80% inhibition of tumor growth with prolonged survival compared with Hsp90 inhibitor monotherapy (114). Furthermore, administration of 17-AAG maintained androgen-sensitivity, prevented nuclear localization of endogenous androgen receptor and delayed the progression of LuCaP35 xenograft tumors to castration resistance. Thus, it appears that targeting Hsp90 may lead to new approaches to prevent and/or treat castration-resistant prostate cancer (115). Alternatively, co-administration of hyperthermia and 17-AAG released from plasmonic matrices was a powerful approach for the ablation of malignant cells and can be used in the form of similar combinations of nanoparticles and chemotherapeutic drugs for a variety of malignancies (116). For the complete list of breast, ovarian and prostate cancer cell response to combination of Hsp90 inhibitor and another therapy (treatment) in vitro and/or in vivo see Table IV.

Table IV

The response of breast, ovarian and prostate cancer cells to combination of Hsp90 inhibitor with another therapy in vitro and/or in vivo.

Table IV

The response of breast, ovarian and prostate cancer cells to combination of Hsp90 inhibitor with another therapy in vitro and/or in vivo.

CancerHsp90 inhibitorSecond therapyConditionsTumor response(Ref.)
BreastGACarbamazepineIn vitro(101)
BreastGABortezomibIn vitro(201)
Breast17-AAGTrastuzumabIn vitro↑ in high ErbB2(202,203)
Breast17-AAGRapamycinIn vitro(204)
Breast17-AAGTaxol, doxorubicinIn vitro, in vivo↑ Rb-dependent(102,205)
Breast17-AAGGefitinibIn vitro(105)
Breast17-AAGLoss of Id proteinIn vivoComplete regression(108)
Breast17-AAGPDT-Photofrin, PDT-NPe6In vitro(206)
Breast and other17-AAGPaclitaxel + rapamycinIn vitro, in vivo(207)
Breast17-DMAGTRAILIn vitro(106)
Breast17-DMAGTriciribineIn vitro(103)
Breast17-DMAGPDT-hypericinIn vitro(208)
BreastIP-504TrastuzumabIn vitro, in vivo(99)
BreastSNX-2112TrastuzumabIn vitro, in vivo(100)
Breast and otherCUDC-305Paclitaxel, camptothecinIn vitro, in vivo(82)
Breast and otherCH5164840
NVP-AUY922
Trastuzumab, lapatinibIn vitro, in vivo(97,98)
BreastNVP-AUY922AZD8055In vitro(209)
OvarianGACisplatinIn vitro(110)
Ovarian17-AAGCarboplatinIn vitro, in vivo↑ beneficial in vivo(210)
Ovarian17-AAGPaclitaxelIn vitro↑ ErbB2, p-Akt-dependent(111)
ProstateGATRAILIn vitro(195)
Prostate17-AAGCdc37 siRNAIn vitro(211)
Prostate17-AAGSurvivin siRNAIn vitro(212)
Prostate17-AAGWee1 siRNAIn vitro(112)
Prostate17-AAG, SNX-2112, STA-9090Wee1 inhibitorIn vitro, in vivo(113)
Prostate17-AAGHyperthermiaIn vitro(116)
Prostate17-AAG
PF04928473
PF04929113
Clusterin siRNA, OGX-011, OGX-427In vitro, in vivo(114,159)
Prostate17-AAGCastrationIn vivo(115)
ProstateNVP-AUY922Ionizing radiationIn vitro, in vivo(213)
ProstateNVP-AUY922DocetaxelIn vivo(214)

[i] The symbol ↑ means at least the additive effect of the combination compared to single therapies.

Multiple myeloma, lymphocytic and myelogenous leukemia

Lamottke et al (117) demonstrated that inhibition of Hsp90 by small molecular mass inhibitors induced cell death in multiple myeloma. They analyzed the effects of the novel, orally bioavailable Hsp90 inhibitor NVP-Hsp990 on multiple myeloma cell proliferation and survival. NVP-Hsp990 inhibitor led to a significant reduction in myeloma cell viability and induced G2 cell cycle arrest, degradation of caspase-8 and caspase-3, and induction of apoptosis. Inhibition of the Hsp90 ATPase activity was accompanied by the degradation of multiple myeloma phospho-Akt and phospho-Erk1/2 and upregulation of Hsp70. In this regard, PI3K/mTOR inhibitor in combination with NVP-Hsp990 mentioned above, rendered the Hsp90 blockade-mediated stress response ineffective and considerably increased the anti-multiple myeloma toxicity (118). Similarly, Chatterjee et al (119) demonstrated that the knockdown of Hsp72 and/or Hsp73 or treatment with VER-155008, decreased protein levels of Hsp90-chaperone clients affecting multiple oncogenic signaling pathways, and acted synergistically with the novel Hsp90 inhibitor NVP-AUY922 in myeloma cell death induction. Inhibition of the PI3K/Akt/GSK3β pathway with siRNA or PI103 decreased expression of the heat shock transcription factor 1 and downregulated constitutive and inducible Hsp70 expression (119). Moreover, NVP-AUY922 in combination with histone deacetylase inhibitors SAHA, NVP-LBH589, melphalan or doxorubicin resulted in synergistic inhibition of viability, with strong synergy (combination index <0.3) in the case of melphalan. Importantly, resistance of the RPMI-8226 cell line and relative resistance of some primary myeloma cells against NVP-AUY922 could be overcome by combination treatment (120). In this manner, the p38 MAPK inhibitor BIRB 796, inhibited protein expression and phosphorylation of Hsp27 induced by Hsp90 inhibitor 17-AAG and enhanced its cytotoxicity (121). Transcription induction of Hsps in response to 17-AAG was also inhibited by the transcription inhibitor actinomycin D. Thus, the combination strategy of 17-AAG and actinomycin D represents a potential approach for the treatment of hematological malignancies (23).

Furthermore, Hsp90 is a promising therapeutic target also in JAK-2 driven cancer, e.g., myeloproliferative neoplasms, B cell acute lymphoblastoid leukemia, including those with genetic resistance to JAK enzymatic inhibitors (122,123). Intriguingly, Hsp90 inhibitor-based antileukemia therapy may override de novo or acquired resistance of acute myelogenous leukemia cells to pan-histone deacetylase inhibitors (124). On the contrary, clinically tolerable doses of tanespimycin (17-AAG) had little effect on resistance-mediating client proteins in relapsed leukemia and exhibits limited activity in combination with cytarabine (Clinicaltrials.gov identifier: NCT00098423) (125).

The Hsp90 inhibitor SNX-7081, which is significantly more potent than 17-AAG synergized and restored sensitivity to fludarabine in seven lymphocytic leukemia cell lines and 23 patient samples, including TP53 mutated cell lines and TP53 or ATM disfunctional patient cells (126). Similar sensitizing effect of Hsp90 inhibitor on chronic lymphocytic leukemia cells was demonstrated by Jones et al (127) as well as Walsby et al (128) who combined GA and NVP-AUY922, respectively, in combination with fludarabine. In this regard, Hsp90 inhibitors represent a potential treatment strategy for fludarabine refractory diseases. A novel strategy to enhance therapeutic response in chronic lymphocytic leukemia may also represent the combination of Hsp90 inhibitor and therapeutic antibody rituximab (129).

Both HL-60 cells transfected with Bcr-Abl and naturally Ph1-positive K562 leukemia cells are resistant to most cytotoxic drugs, but were found to be sensitive to GA. In this regard, GA sensitized Bcr-Abl-expressing cells to doxorubicin and paclitaxel. However, in parental HL-60 cells, GA antagonized the cytotoxic and cell death effects of doxorubicin included inhibition of PARP cleavage and nuclear fragmentation (130). Elimination of mutant BCR-ABL (T315I) kinase seems to be more effective therapeutic strategy for treating BCR-ABL-induced leukemia than the inhibition of BCR-ABL (T315I) kinase activity. Indeed, combination treatment with Hsp90 inhibitor IPI-504 and imatinib was more effective than either treatment alone in prolonging survival of mice simultaneously bearing both wild-type and T315I leukemic cells (131). Simultaneous exposure of BaF3 cells expressing BCR-ABL mutants including T315I to AUY922 and nilotinib was more effective at reducing the outgrowth of resistant cell clones as well as at survival prolongation of mice transplanted with mixture of BaF3 cells expressing wild-type BCR-ABL and mutant forms (132). Similarly, flavopiridol (semisynthetic flavone and selective inhibitor of CDKs) sensitized imatinib-resistant HL/Bcr-Abl cells to GA and it seems that a cocktail of flavopiridol, 17-AAG (or its analogs) and imatinib may be an attractive approach for further investigations (133).

Despite the promising introduction of the proteasome inhibitor bortezomib in the treatment of mantle cell lymphoma (MCL), not all patients respond, and resistance often appears after initial treatment. In bortezomib-resistant cells, cell pretreatment with Hsp90 inhibitor IPI-504, led to synergistic induction of apoptotic cell death when combined with bortezomib (134). For the above, and further studies with a response of multiple myeloma, lymphocytic and myelogenous leukemia cells to combination of Hsp90 inhibitor and another therapy see Table V.

Table V

The response of multiple myeloma, lymphocytic and myelogenous leukemia cells to combination of Hsp90 inhibitor with another therapy in vitro and/or in vivo.

Table V

The response of multiple myeloma, lymphocytic and myelogenous leukemia cells to combination of Hsp90 inhibitor with another therapy in vitro and/or in vivo.

CancerHsp90 inhibitorSecond therapyConditionsTumor response(Ref.)
MyelomaGAApigeninIn vitro(215)
Myeloma17-AAGGRN163LIn vitro, in vivo(216)
Myeloma17-AAGBortezomibIn vitro(217)
Myeloma17-AAGRapamycinIn vitro(218)
Myeloma17-AAG 8-chloro-adenosineIn vitro(219)
Myeloma17-AAGActinomycin DIn vitro(23)
Myeloma17-AAGBIRB 796In vitro(121)
Myeloma17-DMAGPerifosineIn vitro↑ + BM microenvir.(75)
MyelomaIPI-504BortezomibIn vitro, in vivo(220)
MyelomaNVP-AUY922NVP-LBH589 and SAHA, melphalan, doxorubicinIn vitro(120)
MyelomaNVP-AUY922Hsp72/73 siRNA, PI103In vitro(119)
MyelomaNVP-AUY922TriptolideIn vitro(221)
MyelomaNVP-AUY922TG101209In vitro(123)
MyelomaNVP-AUY922NVP_BVB808In vitro, in vivo(122)
MyelomaNVP-HSP990MelphalanIn vitro(117)
MyelomaNVP-HSP990PI3K/mTOR inhibitorIn vitro(118)
LeukemiaGADoxorubicinIn vitro↑ Bcr-Abl-dependent(130,222)
LeukemiaGAFlavopiridolIn vitro(133)
LeukemiaGAHerbimycin A, chlorambucil, fludarabineIn vitro(127)
LeukemiaGANutlin-3In vitro↑ p53 wild-type(223)
Leukemia17-AAGSAHA, SBIn vitro(224)
Leukemia17-AAGUCN-01In vitro(225)
Leukemia17-AAGGTP14564In vitro(226)
Leukemia17-AAGCytarabineIn vitro(227)
Leukemia17-AAGPKC412, LBH589In vitro(228,229)
Leukemia17-AAGImatinibIn vitro(230)
Leukemia17-AAGKNK437, Hsp70 siRNAIn vitro(160)
Leukemia17-AAGHistone deacetylase 6 siRNA, tubacinIn vitro(231)
Leukemia17-AAGRituximabIn vitro(129)
Leukemia17-DMAGDasatinibIn vitro(232)
LeukemiaIP-504ImatinibIn vivo↑ + leukemia stem cells(131)
LeukemiaNVP-AUY922Cytarabine, fludarabineIn vitro, in vivo(128,233)
LeukemiaNVP-AUY922NilotinibIn vitro, in vivo(132)
Leukemia NVP-AUY922
SNX-7081
FludarabineIn vitro(126,234)
Lymphoma17-DMAGDoxorubicinIn vitro↑ when 17-DMAG after doxorubicin(235)
Lymphoma17-DMAGVorinostatIn vitro(236)
LymphomaIPI-504Akt/PI3K inhibitor, doxorubicinIn vitro(237)
LymphomaIPI-504BortezomibIn vitro(134)

[i] The symbol ↑ means at least the additive effect of the combination compared to single therapies.

Glioblastoma multiforme

17-AAG and PARP inhibitor olaparib had modest, replication-dependent radiosensitizing effects on T98G glioma cells. Additive radiosensitization was observed through the combination treatment, mirrored by increases in gammaH2AX foci in G(2)-phase cells. Unlike olaparib, 17-AAG did not increase radiation sensitivity of Chinese hamster ovary cells, indicating tumor specificity (135). In addition, 17-AAG sensitized xenografted U87MG cells to cisplatin in nude mice. Indeed, Hsp90-targeted therapy seems to be an effective strategy for potentiating chemotherapy using DNA-crosslinking agents for temozolomide-refractory gliomas (136). Moreover, treatment with subtoxic doses of 17-AAG in combination with tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) induced rapid apoptosis in TRAIL-resistant glioma cells, suggesting an attractive strategy for treating gliomas (137). Interestingly, 17-AAG inhibited the growth of intracranial tumors and synergized with radiation both in tissue culture and in intracranial tumors. On the contrary, 17-AAG did not synergize with temozolomide in any of glioma models (138). The response of glioblastoma cells to combination of Hsp90 inhibitor with other therapy is summarized in Table VI.

Table VI

The response of glioblastoma cells to combination of Hsp90 inhibitor with another therapy in vitro and/or in vivo.

Table VI

The response of glioblastoma cells to combination of Hsp90 inhibitor with another therapy in vitro and/or in vivo.

CancerHsp90 inhibitorSecond therapyConditionsTumor response(Ref.)
Glioblastoma17-AAGZD1839, LY294002In vitro(238,239)
Glioblastoma17-AAGEnzastaurinIn vitro(240)
Glioblastoma17-AAGTRAILIn vitro(137)
Glioblastoma17-AAGCisplatin, 1,3-bis(2-chloroethyl)-1-nitrosoureaIn vitro, in vivo(136)
Glioblastoma17-AAG
17-DMAG
Ionizing radiation, temozolomideIn vitro, in vivo↑ to ionizing radiation, but not to temozolomide
↑ in glioma stem cells
(136,138,241)
Glioblastoma17-AAGIonizing radiation, olaparibIn vitro(135)
Glioblastoma and other17-DMAGIonizing radiationIn vitro, in vivo(242)
GlioblastomaNVP-AUY922PI3K/mTOR inhibitor, Mek inhibitorIn vitro(155)
GlioblastomaHSP990BKM120In vitro↑ independent on PTEN/p53(243)

[i] The symbol ↑ means at least the additive effect of the combination compared to single therapies.

Melanoma

17-AAG induced in uveal melanoma cells compared to cutaneous melanoma cells (with low tyrosine kinase activity of WTB-Raf) downregulation of WTB-Raf protein, followed by Mek/Erk inactivation and a decrease in cyclin D1. In this regard, the combination of 17-DMAG and imatinib mesylate was more efficient, than individual treatments of c-Kit positive uveal melanoma cells (21). The Hsp90 inhibitor XL888 increased Bcl-2 interacting mediator of cell death expression, decreased Mcl-1 expression, and induced apoptosis more effectively than dual mitogen-activated protein-extracellular signal-regulated in most resistant melanoma models. The reversal of the resistance phenotype was associated with the degradation of PDGFRβ, COT, IGFR1, CRAF, ARAF, S6, cyclin D1 and Akt (139). Inhibition of Hsp90 by GA sensitized melanoma cells also to thermosensitive ferromagnetic particle-mediated by hyperthermia with temperature 43°C (140). The Hsp90 inhibitor ganetespib was reported to exhibit robust antitumor efficacy in preclinical model of BRAFV600E melanoma and could thus readily overcome mechanisms of intrinsic and acquired resistance to selective BRAF inhibitors. The data mentioned above suggest that ganetespib may offer an alternative, and potentially complementary, strategy for therapeutic intervention in mutant BRAF-driven disease (141). The above studies are summarized in Table VII.

Table VII

The response of melanoma cells to combination of Hsp90 inhibitor with other therapy in vitro and/or in vivo.

Table VII

The response of melanoma cells to combination of Hsp90 inhibitor with other therapy in vitro and/or in vivo.

CancerHsp90 inhibitorSecond therapyConditionsTumor response(Ref.)
MelanomaGAHyperthermiaIn vitro, in vivo(140)
Melanoma17-DMAGImatinib mesylateIn vitro↑ in WT B-Raf uveal melanoma cells(21)
MelanomaSTA-9090TAK-733In vitro, in vivo↑ in melanoma cells with BRAF(V600E) mutation(141)
Melanoma and otherPU-H71Cisplatin, melphalanIn vitro(244)
MelanomaXL888Vemurafenib, doxycyclineIn vitro, 3D spheroids(139)

[i] The symbol ↑ means at least the additive effect of the combination compared to single therapies.

Bladder and renal cancer

Although up to 70% of advanced bladder cancer patients initially show good tumor response to systemic cisplatin-based combination chemotherapy, >90% of good responders relapse and eventually die of the disease. This is attributable to the re-growth of bladder cancer-initiating cells, that have survived chemotherapy. These cancer-initiating cells were more resistant to cisplatin and exhibited more activity in the Akt and Erk oncogenic signaling pathways when compared with their CD44 counterparts. On the other hand, 17-DMAG simultaneously inactivated both Akt and Erk signaling and synergistically potentiated the cytotoxicity of cisplatin against bladder cancer-initiating cells by enhancing cisplatin-induced apoptosis in vitro and in vivo (142). Similarly, low doses of 17-DMAG potentiated antitumor activity of chemoradiotherapy (cisplatin + ionizing radiation) and supported clinical trial of Hsp90 inhibitors to overcome chemoradiotherapy resistance in patients with muscle-invasive bladder cancer (143). The combination of 17-AAG and HIV protease inhibitor ritonavir inhibited renal cancer growth through the inhibition of HSF-1 expression and downregulation of Hsp27, Hsp70 and Hsp90 induced by 17-AAG (144). Recently, Ma et al (145) demonstrated dual targeting of two chaperones Hsp90 and Hsp70 and their combination with conventional anticancer drugs as promising therapeutic option for patients with advanced bladder cancer. The response of bladder and renal cancer cells to combination of Hsp90 inhibitor with another therapy is summarized in Table VIII.

Table VIII

The response of bladder and renal cancer cells to combination of Hsp90 inhibitor with another therapy in vitro and/or in vivo.

Table VIII

The response of bladder and renal cancer cells to combination of Hsp90 inhibitor with another therapy in vitro and/or in vivo.

CancerHsp90 inhibitorSecond therapyConditionsTumor response(Ref.)
BladderGADNA vaccine against p185 and MetIn vivo(245)
Bladder17-AAGPifithrin + chemotherapyIn vitro(145)
Renal17-AAGRitonavirIn vitro, in vivo(144)
Bladder17-AAG, 17-DMAGCisplatinIn vitro, in vivo(142)
Bladder17-DMAG ChemoradiotherapyIn vitro, in vivo(143)

[i] The symbol ↑ means at least the additive effect of the combination compared to single therapies.

Head and neck cancer

The synthetic Hsp90 inhibitor BIIB021 revealed a strong antitumor effect in head and neck squamous cell carcinoma cells in vitro and in vivo. Its antitumor effect was higher than the effect of ansamycin-based Hsp90 inhibitor 17-AAG and when combined with radiation sensitized the efficacy of radiation therapy (146). The activation of multi-chaperone complex could be a resistance mechanism to the antiproliferative and apoptotic effects induced by tipifarnib (farnesyl transferase inhibitor) and that the combination of tipifarnib with GA is possibly a suitable therapeutic strategy in the treatment of Hsp90-overexpressing head and neck squamous cell cancers (147). Interestingly, the combination of GA with docetaxel was not able to induce any synergistic effects on cell growth inhibition of head and neck squamous cell cancers independently of Hsp90 expression and activation status. On the other hand, the combination of GA with tipifarnib was potentially able to give synergistic effects on the growth arrest of head and neck squamous cell cancers especially when Hsp90 was overexpressed and even more when it was over-activated (147). Sensitizing effect of Hsp90 inhibitor was demonstrated in many other transformed cells, the results of which are shown together with head and neck cancer cells in Table IX.

Table IX

The response of head and neck, neuroblastoma, osteosarcoma and other transformed cells to combination of Hsp90 inhibitor with another therapy in vitro and/or in vivo.

Table IX

The response of head and neck, neuroblastoma, osteosarcoma and other transformed cells to combination of Hsp90 inhibitor with another therapy in vitro and/or in vivo.

CancerHsp90 inhibitorSecond therapyConditionsTumor response(Ref.)
Head and neck squamous cells overexpressing Hsp90GATipifarnib, docetaxelIn vitro↑ after GA + tipifarnib, not after
GA + docetaxel
(147)
Head and neck squamous cells17-AAGCisplatin, nutlin-3aIn vitro, in vivo↑ depended on p53 and MDMX(246)
Head and neck squamous cellsBIIB021Ionizing radiationIn vitro, in vivo(146)
Neuroblastoma, osteosarcomaGACisplatinIn vitro(247)
OsteosarcomaNVP-AUY922CDK inhibitorIn vitro(248)
SV40-transformed COS-7 cellsGAVelcadeIn vitro(249)
Ewing’s sarcoma17-AAGIGF1R inhibitorIn vitro, in vivo↑ in IGF1R-positive cells(250)
HeLa, B16, MCA205 cells17-DMAGQuinacrine, 9-aminoacridineIn vitro, in vivo(251)
HeLa cells17-AAGIonizing radiationIn vitro(252)
Rhabdomyosarcoma17-DMAGBortezomibIn vitro(56)
PediatricSNX-2112CisplatinIn vitro(253)

[i] The symbol ↑ means at least the additive effect of the combination compared to single therapies.

Stem cells

Recent studies of Lee et al (148) showed that the combination of GA with quercetin or KNK437 sensitized breast cancer stem-like cells characterized by high intracellular aldehyde dehydrogenase activity and higher expression of Hsp90α toward antiproliferation and anti-migration effects of GA. Lee et al (148) also found that knockdown of Hsp27 could mimic the effect of Hsp inhibitors to potentiate the breast cancer stem-like cell targeting effect of GA. Newman et al (149) also supported the use of Hsp90 inhibitor as a cancer stem cell targeting agent and showed, that HSF-1 is an important target for elimination of both cancer and non cancer stem cells in cancer. They found, that low concentrations of the Hsp90 inhibitor 17-AAG eliminate lymphoma cancer stem cells in vitro and in vivo by disrupting the transcriptional function of HIF-1α (149).

6. Resistance to Hsp90 inhibitors

Kelland et al (150) observed a positive relationship between NAD(P)H:quinone oxidoreductase (NQO1, DT-diaphorase) expression level and growth inhibition by 17-AAG. Stable, high-level expression of the NQO1 gene transfected into the DT-diaphorase-deficient (by NQO1 mutation) BE human colon carcinoma cell line resulted in a 32-fold increase in 17-AAG growth-inhibition activity. Increased sensitivity to 17-AAG in the transfected cell line was also confirmed in xenografts (150). Low NQO1 activity could be a likely mechanism of acquired resistance to 17-AAG in glioblastoma and melanoma. Indeed, Hsp90 inhibitors which avoid the liability of NQO1 metabolism are able to avoid the resistance due to decreased NQO1 activity and provide additional support for the clinical development of such structurally novel Hsp90 inhibitors (151). In this regard, the synthetic purine-scaffold Hsp90 inhibitor BIIB021, resorcinylic diaryl pyrazole CCT018159 as well as resorcinylic diaryl pyrazole/isoxazole amides VER-49009, VER-50589 or VER-52296 (NVP-AUY922 and AUY922) are promising Hsp90 inhibitors with several advantages compared to 17-AAG, including better solubility and independence from the effects of P-glycoprotein and NQO1 (151153). Alongside depletion of client proteins, however, the BIIB021, VER-50589, VER-49009 and NVP-AUY922 all caused induction of Hsp72 and Hsp27 (31,154,155). On the other hand, the upregulation of Hsp27 has a significant role in 17-AAG resistance, which may be mediated in part through GSH regulation. The modulation of GSH was able to enhance the efficacy of Hsp90-directed therapy in clinic (156). Sharp et al (157) demonstrated that thiadiazole inhibitors display a more limited core set of interactions relative to the clinical trial candidate NVP-AUY922 and consequently may be less susceptible to resistance derived through mutations in Hsp90, but these inhibitors also caused induction of the heat shock response with the upregulation of Hsp72 and Hsp27 protein expression.

7. Combination of Hsp inhibitors

The key, how to solve the problem of induced expression of heat shock proteins especially Hsp27 and Hsp70, after the treatment with Hsp90 inhibitors could be combined therapy of Hsp90 inhibitor and other Hsp inhibitors. Targeting multiple Hsps could likely shut down the heat shock response and associated rescue mechanisms, leading to low chemotherapeutic resistance (158). Dual therapy of combined Hsp inhibitors showed that Hsp27 inhibitors potentiate the effects of Hsp90 inhibitors and delay castrate-resistant prostate cancer (114,159) as well as sensitize breast cancer stem-like cells (148). Moreover, KNK437 a benzylidine lactam inhibitor of Hsp70, attenuated 17-AAG-mediated Hsp70 induction and increased 17-AAG-induced apoptosis and loss of clonogenic survival of HL-60 cells (160). The overexpression of Hsp70 induced by 17-AAG or NVP-AUY922 inhibitors in cancer cells was counteracted considerably when Hsp90 inhibitors are combined with the new type of chaperone inhibitor 1,6-dimethyl-3-propylpyrimido[5,4-e][1,2,4]triazine-5,7-dione (C9), which prevents Hsp90 from interacting with the cochaperone HOP (161). Very promising therapeutic option for cancer patients was reported by Ma et al (145), who targeted two chaperones Hsp90 and Hsp70 together with conventional anticancer drugs. Very similar combination of Hsp90 and Hsp70 inhibition with hyperthermia represented by ferromagnetic particles increased very effectively their antitumor effects in vitro as well as in vivo (162). For the future, it seems that the combinations of Hsp inhibitors applied with other therapies could serve as a potential solution to prevent the drug resistance and avoid the toxicity of high dose of Hsp90 inhibitors in clinical application.

Acknowledgements

This study was partly supported by the Scientific Grant Agency of the Ministry of Education of the Slovak Republic under contract no. VEGA 1/0733/12. Professional consulting service was provided by company K-TRADE spol. s r.o. (http://www.ktrade.sk/).

References

1 

Neckers L: Hsp90 inhibitors as novel cancer chemotherapeutic agents. Trends Mol Med. 8:S55–S61. 2002. View Article : Google Scholar : PubMed/NCBI

2 

Mitsiades CS, Mitsiades NS, McMullan CJ, et al: Antimyeloma activity of heat shock protein-90 inhibition. Blood. 107:1092–1100. 2006. View Article : Google Scholar

3 

Chiosis G, Caldas Lopes E and Solit D: Heat shock protein-90 inhibitors: a chronicle from geldanamycin to today’s agents. Curr Opin Investig Drugs. 7:534–541. 2006.PubMed/NCBI

4 

Teng SC, Chen YY, Su YN, et al: Direct activation of HSP90A transcription by c-Myc contributes to c-Myc-induced transformation. J Biol Chem. 279:14649–14655. 2004. View Article : Google Scholar : PubMed/NCBI

5 

McClellan AJ, Xia Y, Deutschbauer AM, Davis RW, Gerstein M and Frydman J: Diverse cellular functions of the Hsp90 molecular chaperone uncovered using systems approaches. Cell. 131:121–135. 2007. View Article : Google Scholar : PubMed/NCBI

6 

Okamoto J, Mikami I, Tominaga Y, et al: Inhibition of Hsp90 leads to cell cycle arrest and apoptosis in human malignant pleural mesothelioma. J Thorac Oncol. 3:1089–1095. 2008. View Article : Google Scholar : PubMed/NCBI

7 

Trepel J, Mollapour M, Giaccone G and Neckers L: Targeting the dynamic HSP90 complex in cancer. Nat Rev Cancer. 10:537–549. 2010. View Article : Google Scholar : PubMed/NCBI

8 

Banerji U: Heat shock protein 90 as a drug target: some like it hot. Clin Cancer Res. 15:9–14. 2009. View Article : Google Scholar : PubMed/NCBI

9 

Li Y, Zhang T, Schwartz SJ and Sun D: New developments in Hsp90 inhibitors as anti-cancer therapeutics: mechanisms, clinical perspective and more potential. Drug Resist Updat. 12:17–27. 2009. View Article : Google Scholar : PubMed/NCBI

10 

Kamal A, Thao L, Sensintaffar J, et al: A high-affinity conformation of Hsp90 confers tumour selectivity on Hsp90 inhibitors. Nature. 425:407–410. 2003. View Article : Google Scholar : PubMed/NCBI

11 

Isaacs JS, Xu W and Neckers L: Heat shock protein 90 as a molecular target for cancer therapeutics. Cancer Cell. 3:213–217. 2003. View Article : Google Scholar : PubMed/NCBI

12 

Holzbeierlein JM, Windsperger A and Vielhauer G: Hsp90: a drug target? Curr Oncol Rep. 12:95–101. 2010. View Article : Google Scholar : PubMed/NCBI

13 

Amolins MW and Blagg BS: Natural product inhibitors of Hsp90: potential leads for drug discovery. Mini Rev Med Chem. 9:140–152. 2009. View Article : Google Scholar : PubMed/NCBI

14 

Krishna P and Gloor G: The Hsp90 family of proteins in Arabidopsis thaliana. Cell Stress Chaperones. 6:238–246. 2001. View Article : Google Scholar : PubMed/NCBI

15 

Hainzl O, Lapina MC, Buchner J and Richter K: The charged linker region is an important regulator of Hsp90 function. J Biol Chem. 284:22559–22567. 2009. View Article : Google Scholar : PubMed/NCBI

16 

Hawle P, Siepmann M, Harst A, Siderius M, Reusch HP and Obermann WM: The middle domain of Hsp90 acts as a discriminator between different types of client proteins. Mol Cell Biol. 26:8385–8395. 2006. View Article : Google Scholar : PubMed/NCBI

17 

Lee CC, Lin TW, Ko TP and Wang AH: The hexameric structures of human heat shock protein 90. PLoS One. 6:e199612011. View Article : Google Scholar : PubMed/NCBI

18 

Prodromou C, Roe SM, O’Brien R, Ladbury JE, Piper PW and Pearl LH: Identification and structural characterization of the ATP/ADP-binding site in the Hsp90 molecular chaperone. Cell. 90:65–75. 1997. View Article : Google Scholar : PubMed/NCBI

19 

Whitesell L, Mimnaugh EG, De Costa B, Myers CE and Neckers LM: Inhibition of heat shock protein HSP90-pp60v-src heteroprotein complex formation by benzoquinone ansamycins: essential role for stress proteins in oncogenic transformation. Proc Natl Acad Sci USA. 91:8324–8328. 1994. View Article : Google Scholar : PubMed/NCBI

20 

Karkoulis PK, Stravopodis DJ, Margaritis LH and Voutsinas GE: 17-Allylamino-17-demethoxygeldanamycin induces downregulation of critical Hsp90 protein clients and results in cell cycle arrest and apoptosis of human urinary bladder cancer cells. BMC Cancer. 10:4812010. View Article : Google Scholar : PubMed/NCBI

21 

Babchia N, Calipel A, Mouriaux F, Faussat AM and Mascarelli F: 17-AAG and 17-DMAG-induced inhibition of cell proliferation through B-Raf downregulation in WT B-Raf-expressing uveal melanoma cell lines. Invest Ophthalmol Vis Sci. 49:2348–2356. 2008. View Article : Google Scholar : PubMed/NCBI

22 

Villa R, Folini M, Porta CD, et al: Inhibition of telomerase activity by geldanamycin and 17-allylamino, 17-demethoxygeldanamycin in human melanoma cells. Carcinogenesis. 24:851–859. 2003. View Article : Google Scholar : PubMed/NCBI

23 

Cervantes-Gomez F, Nimmanapalli R and Gandhi V: Transcription inhibition of heat shock proteins: a strategy for combination of 17-allylamino-17-demethoxygeldanamycin and actinomycin D. Cancer Res. 69:3947–3954. 2009. View Article : Google Scholar : PubMed/NCBI

24 

Messaoudi S, Peyrat JF, Brion JD and Alami M: Heat-shock protein 90 inhibitors as antitumor agents: a survey of the literature from 2005 to 2010. Expert Opin Ther Pat. 21:1501–1542. 2011. View Article : Google Scholar : PubMed/NCBI

25 

Menzella HG, Tran TT, Carney JR, et al: Potent non-benzoquinone ansamycin heat shock protein 90 inhibitors from genetic engineering of Streptomyces hygroscopicus. J Med Chem. 52:1518–1521. 2009. View Article : Google Scholar : PubMed/NCBI

26 

Turbyville TJ, Wijeratne EM, Liu MX, et al: Search for Hsp90 inhibitors with potential anticancer activity: isolation and SAR studies of radicicol and monocillin I from two plant-associated fungi of the Sonoran desert. J Nat Prod. 69:178–184. 2006. View Article : Google Scholar : PubMed/NCBI

27 

Pearl LH, Prodromou C and Workman P: The Hsp90 molecular chaperone: an open and shut case for treatment. Biochem J. 410:439–453. 2008. View Article : Google Scholar : PubMed/NCBI

28 

Soga S, Shiotsu Y, Akinaga S and Sharma SV: Development of radicicol analogues. Curr Cancer Drug Targets. 3:359–369. 2003. View Article : Google Scholar : PubMed/NCBI

29 

Chiosis G, Timaul MN, Lucas B, et al: A small molecule designed to bind to the adenine nucleotide pocket of Hsp90 causes Her2 degradation and the growth arrest and differentiation of breast cancer cells. Chem Biol. 8:289–299. 2001. View Article : Google Scholar : PubMed/NCBI

30 

Kasibhatla SR, Hong K, Biamonte MA, et al: Rationally designed high-affinity 2-amino-6-halopurine heat shock protein 90 inhibitors that exhibit potent antitumor activity. J Med Chem. 50:2767–2778. 2007. View Article : Google Scholar : PubMed/NCBI

31 

Lundgren K, Zhang H, Brekken J, et al: BIIB021, an orally available, fully synthetic small-molecule inhibitor of the heat shock protein Hsp90. Mol Cancer Ther. 8:921–929. 2009. View Article : Google Scholar : PubMed/NCBI

32 

Taldone T and Chiosis G: Purine-scaffold Hsp90 inhibitors. Curr Top Med Chem. 9:1436–1446. 2009. View Article : Google Scholar : PubMed/NCBI

33 

Chandarlapaty S, Sawai A, Ye Q, et al: SNX2112, a synthetic heat shock protein 90 inhibitor, has potent antitumor activity against HER kinase-dependent cancers. Clin Cancer Res. 14:240–248. 2008. View Article : Google Scholar : PubMed/NCBI

34 

Jhaveri K, Taldone T, Modi S and Chiosis G: Advances in the clinical development of heat shock protein 90 (Hsp90) inhibitors in cancers. Biochim Biophys Acta. 1823:742–755. 2012. View Article : Google Scholar :

35 

Jhaveri K, Ochiana SO, Dunphy MP, et al: Heat shock protein 90 inhibitors in the treatment of cancer: current status and future directions. Expert Opin Investig Drugs. 23:611–628. 2014. View Article : Google Scholar : PubMed/NCBI

36 

Raad II, Hachem RY, Abi-Said D, et al: A prospective crossover randomized trial of novobiocin and rifampin prophylaxis for the prevention of intravascular catheter infections in cancer patients treated with interleukin-2. Cancer. 82:403–411. 1998. View Article : Google Scholar : PubMed/NCBI

37 

Rappa G, Shyam K, Lorico A, Fodstad O and Sartorelli AC: Structure-activity studies of novobiocin analogs as modulators of the cytotoxicity of etoposide (VP-16). Oncol Res. 12:113–119. 2000. View Article : Google Scholar

38 

Gormley NA, Orphanides G, Meyer A, Cullis PM and Maxwell A: The interaction of coumarin antibiotics with fragments of DNA gyrase B protein. Biochemistry. 35:5083–5092. 1996. View Article : Google Scholar : PubMed/NCBI

39 

Marcu MG, Schulte TW and Neckers L: Novobiocin and related coumarins and depletion of heat shock protein 90-dependent signaling proteins. J Natl Cancer Inst. 92:242–248. 2000. View Article : Google Scholar : PubMed/NCBI

40 

Peyrat JF, Messaoudi S, Brion JD and Alami M: Inhibitors of the heat shock protein 90: from cancer clinical trials to neurodegenerative diseases. Atlas Genet Cytogenet Oncol Haematol. 15:89–105. 2011.

41 

Le Bras G, Radanyi C, Peyrat JF, et al: New novobiocin analogues as antiproliferative agents in breast cancer cells and potential inhibitors of heat shock protein 90. J Med Chem. 50:6189–6200. 2007. View Article : Google Scholar : PubMed/NCBI

42 

Donnelly AC, Mays JR, Burlison JA, et al: The design, synthesis, and evaluation of coumarin ring derivatives of the novobiocin scaffold that exhibit antiproliferative activity. J Org Chem. 73:8901–8920. 2008. View Article : Google Scholar : PubMed/NCBI

43 

Radanyi C, Le Bras G, Marsaud V, et al: Antiproliferative and apoptotic activities of tosylcyclonovobiocic acids as potent heat shock protein 90 inhibitors in human cancer cells. Cancer Lett. 274:88–94. 2009. View Article : Google Scholar

44 

Burlison JA and Blagg BS: Synthesis and evaluation of coumermycin A1 analogues that inhibit the Hsp90 protein folding machinery. Org Lett. 8:4855–4858. 2006. View Article : Google Scholar : PubMed/NCBI

45 

Shelton SN, Shawgo ME, Matthews SB, et al: KU135, a novel novobiocin-derived C-terminal inhibitor of the 90-kDa heat shock protein, exerts potent antiproliferative effects in human leukemic cells. Mol Pharmacol. 76:1314–1322. 2009. View Article : Google Scholar : PubMed/NCBI

46 

Samadi AK, Zhang X, Mukerji R, Donnelly AC, Blagg BS and Cohen MS: A novel C-terminal HSP90 inhibitor KU135 induces apoptosis and cell cycle arrest in melanoma cells. Cancer Lett. 312:158–167. 2011. View Article : Google Scholar : PubMed/NCBI

47 

Eskew JD, Sadikot T, Morales P, et al: Development and characterization of a novel C-terminal inhibitor of Hsp90 in androgen dependent and independent prostate cancer cells. BMC Cancer. 11:4682011. View Article : Google Scholar : PubMed/NCBI

48 

Matthews SB, Vielhauer GA, Manthe CA, et al: Characterization of a novel novobiocin analogue as a putative C-terminal inhibitor of heat shock protein 90 in prostate cancer cells. Prostate. 70:27–36. 2010. View Article : Google Scholar

49 

Zhao H, Moroni E, Colombo G and Blagg BS: Identification of a new scaffold for hsp90 C-terminal inhibition. ACS Med Chem Lett. 5:84–88. 2013. View Article : Google Scholar

50 

Fujiki H, Suganuma M, Okabe S, et al: Mechanistic findings of green tea as cancer preventive for humans. Proc Soc Exp Biol Med. 220:225–228. 1999. View Article : Google Scholar : PubMed/NCBI

51 

Palermo CM, Westlake CA and Gasiewicz TA: Epigallocatechin gallate inhibits aryl hydrocarbon receptor gene transcription through an indirect mechanism involving binding to a 90 kDa heat shock protein. Biochemistry. 44:5041–5052. 2005. View Article : Google Scholar : PubMed/NCBI

52 

Li Y, Zhang T, Jiang Y, Lee HF, Schwartz SJ and Sun D: (−)-Epigallocatechin-3-gallate inhibits Hsp90 function by impairing Hsp90 association with cochaperones in pancreatic cancer cell line Mia Paca-2. Mol Pharm. 6:1152–1159. 2009. View Article : Google Scholar : PubMed/NCBI

53 

Yin Z, Henry EC and Gasiewicz TA: (−)-Epigallocatechin-3-gallate is a novel Hsp90 inhibitor. Biochemistry. 48:336–345. 2009. View Article : Google Scholar :

54 

Tran PL, Kim SA, Choi HS, Yoon JH and Ahn SG: Epigallocatechin-3-gallate suppresses the expression of HSP70 and HSP90 and exhibits anti-tumor activity in vitro and in vivo. BMC Cancer. 10:2762010. View Article : Google Scholar : PubMed/NCBI

55 

Jacquemont C, Simon JA, D’Andrea AD and Taniguchi T: Non-specific chemical inhibition of the Fanconi anemia pathway sensitizes cancer cells to cisplatin. Mol Cancer. 11:262012. View Article : Google Scholar : PubMed/NCBI

56 

Peron M, Bonvini P and Rosolen A: Effect of inhibition of the ubiquitin-proteasome system and Hsp90 on growth and survival of rhabdomyosarcoma cells in vitro. BMC Cancer. 12:2332012. View Article : Google Scholar : PubMed/NCBI

57 

Kabakov AE, Kudryavtsev VA and Gabai VL: Hsp90 inhibitors as promising agents for radiotherapy. J Mol Med (Berl). 88:241–247. 2010. View Article : Google Scholar

58 

Vaseva AV, Yallowitz AR, Marchenko ND, Xu S and Moll UM: Blockade of Hsp90 by 17AAG antagonizes MDMX and synergizes with Nutlin to induce p53-mediated apoptosis in solid tumors. Cell Death Dis. 2:e1562011. View Article : Google Scholar : PubMed/NCBI

59 

Shintani S, Zhang T, Aslam A, Sebastian K, Yoshimura T and Hamakawa H: P53-dependent radiosensitizing effects of Hsp90 inhibitor 17-allylamino-17-demethoxygeldanamycin on human oral squamous cell carcinoma cell lines. Int J Oncol. 29:1111–1117. 2006.PubMed/NCBI

60 

To KK, Yu L, Liu S, Fu J and Cho CH: Constitutive AhR activation leads to concomitant ABCG2-mediated multidrug resistance in cisplatin-resistant esophageal carcinoma cells. Mol Carcinog. 51:449–464. 2012. View Article : Google Scholar

61 

Wu X, Wanders A, Wardega P, et al: Hsp90 is expressed and represents a therapeutic target in human oesophageal cancer using the inhibitor 17-allylamino-17-demethoxygeldanamycin. Br J Cancer. 100:334–343. 2009. View Article : Google Scholar : PubMed/NCBI

62 

Floris G, Debiec-Rychter M, Wozniak A, et al: The heat shock protein 90 inhibitor IPI-504 induces KIT degradation, tumor shrinkage, and cell proliferation arrest in xenograft models of gastrointestinal stromal tumors. Mol Cancer Ther. 10:1897–1908. 2011. View Article : Google Scholar : PubMed/NCBI

63 

Floris G, Sciot R, Wozniak A, et al: The novel HSP90 inhibitor, IPI-493, is highly effective in human gastrostrointestinal stromal tumor xenografts carrying heterogeneous KIT mutations. Clin Cancer Res. 17:5604–5614. 2011. View Article : Google Scholar : PubMed/NCBI

64 

Smyth T, Van Looy T, Curry JE, et al: The HSP90 inhibitor, AT13387, is effective against imatinib-sensitive and -resistant gastrointestinal stromal tumor models. Mol Cancer Ther. 11:1799–1808. 2012. View Article : Google Scholar : PubMed/NCBI

65 

Lang SA, Moser C, Fichnter-Feigl S, et al: Targeting heat-shock protein 90 improves efficacy of rapamycin in a model of hepatocellular carcinoma in mice. Hepatology. 49:523–532. 2009. View Article : Google Scholar

66 

Gloesenkamp C, Nitzsche B, Lim AR, et al: Heat shock protein 90 is a promising target for effective growth inhibition of gastrointestinal neuroendocrine tumors. Int J Oncol. 40:1659–1667. 2012.PubMed/NCBI

67 

Oikonomou E, Koc M, Sourkova V, Andera L and Pintzas A: Selective BRAFV600E inhibitor PLX4720, requires TRAIL assistance to overcome oncogenic PIK3CA resistance. PLoS One. 6:e216322011. View Article : Google Scholar : PubMed/NCBI

68 

Moser C, Lang SA, Kainz S, et al: Blocking heat shock protein-90 inhibits the invasive properties and hepatic growth of human colon cancer cells and improves the efficacy of oxaliplatin in p53-deficient colon cancer tumors in vivo. Mol Cancer Ther. 6:2868–2878. 2007. View Article : Google Scholar : PubMed/NCBI

69 

Rakitina TV, Vasilevskaya IA and O’Dwyer PJ: Inhibition of G1/S transition potentiates oxaliplatin-induced cell death in colon cancer cell lines. Biochem Pharmacol. 73:1715–1726. 2007. View Article : Google Scholar : PubMed/NCBI

70 

Vasilevskaya IA, Rakitina TV and O’Dwyer PJ: Quantitative effects on c-Jun N-terminal protein kinase signaling determine synergistic interaction of cisplatin and 17-allylamino-17-demethoxygeldanamycin in colon cancer cell lines. Mol Pharmacol. 65:235–243. 2004. View Article : Google Scholar : PubMed/NCBI

71 

McNamara AV, Barclay M, Watson AJ and Jenkins JR: Hsp90 inhibitors sensitise human colon cancer cells to topoisomerase I poisons by depletion of key anti-apoptotic and cell cycle checkpoint proteins. Biochem Pharmacol. 83:355–367. 2012. View Article : Google Scholar

72 

Tse AN, Sheikh TN, Alan H, Chou TC and Schwartz GK: 90-kDa heat shock protein inhibition abrogates the topoisomerase I poison-induced G2/M checkpoint in p53-null tumor cells by depleting Chk1 and Wee1. Mol Pharmacol. 75:124–133. 2009. View Article : Google Scholar :

73 

Moran DM, Gawlak G, Jayaprakash MS, Mayar S and Maki CG: Geldanamycin promotes premature mitotic entry and micro-nucleation in irradiated p53/p21 deficient colon carcinoma cells. Oncogene. 27:5567–5577. 2008. View Article : Google Scholar : PubMed/NCBI

74 

Giessrigl B, Krieger S, Rosner M, et al: Hsp90 stabilizes Cdc25A and counteracts heat shock-mediated Cdc25A degradation and cell-cycle attenuation in pancreatic carcinoma cells. Hum Mol Genet. 21:4615–4627. 2012. View Article : Google Scholar : PubMed/NCBI

75 

Huston A, Leleu X, Jia X, et al: Targeting Akt and heat shock protein 90 produces synergistic multiple myeloma cell cytotoxicity in the bone marrow microenvironment. Clin Cancer Res. 14:865–874. 2008. View Article : Google Scholar : PubMed/NCBI

76 

Dote H, Cerna D, Burgan WE, Camphausen K and Tofilon PJ: ErbB3 expression predicts tumor cell radiosensitization induced by Hsp90 inhibition. Cancer Res. 65:6967–6975. 2005. View Article : Google Scholar : PubMed/NCBI

77 

Li Y, Zhang T, Schwartz SJ and Sun D: Sulforaphane potentiates the efficacy of 17-allylamino 17-demethoxygeldanamycin against pancreatic cancer through enhanced abrogation of Hsp90 chaperone function. Nutr Cancer. 63:1151–1159. 2011. View Article : Google Scholar : PubMed/NCBI

78 

De Raedt T, Walton Z, Yecies JL, et al: Exploiting cancer cell vulnerabilities to develop a combination therapy for ras-driven tumors. Cancer Cell. 20:400–413. 2011. View Article : Google Scholar : PubMed/NCBI

79 

Acquaviva J, Smith DL, Sang J, et al: Targeting KRAS-mutant non-small cell lung cancer with the Hsp90 inhibitor ganetespib. Mol Cancer Ther. 11:2633–2643. 2012. View Article : Google Scholar : PubMed/NCBI

80 

Proia DA, Sang J, He S, et al: Synergistic activity of the Hsp90 inhibitor ganetespib with taxanes in non-small cell lung cancer models. Invest New Drugs. 30:2201–2209. 2012. View Article : Google Scholar : PubMed/NCBI

81 

Bao R, Lai CJ, Wang DG, et al: Targeting heat shock protein 90 with CUDC-305 overcomes erlotinib resistance in non-small cell lung cancer. Mol Cancer Ther. 8:3296–3306. 2009. View Article : Google Scholar : PubMed/NCBI

82 

Bao R, Lai CJ, Qu H, et al: CUDC-305, a novel synthetic HSP90 inhibitor with unique pharmacologic properties for cancer therapy. Clin Cancer Res. 15:4046–4057. 2009. View Article : Google Scholar : PubMed/NCBI

83 

Koizumi H, Yamada T, Takeuchi S, et al: Hsp90 inhibition overcomes HGF-triggering resistance to EGFR-TKIs in EGFR-mutant lung cancer by decreasing client protein expression and angiogenesis. J Thorac Oncol. 7:1078–1085. 2012. View Article : Google Scholar : PubMed/NCBI

84 

Xu L, Kikuchi E, Xu C, et al: Combined EGFR/MET or EGFR/HSP90 inhibition is effective in the treatment of lung cancers codriven by mutant EGFR containing T790M and MET. Cancer Res. 72:3302–3311. 2012. View Article : Google Scholar : PubMed/NCBI

85 

Rice JW, Veal JM, Barabasz A, et al: Targeting of multiple signaling pathways by the Hsp90 inhibitor SNX-2112 in EGFR resistance models as a single agent or in combination with erlotinib. Oncol Res. 18:229–242. 2009. View Article : Google Scholar : PubMed/NCBI

86 

Tsai MS, Weng SH, Chen HJ, et al: Inhibition of p38 MAPK-dependent excision repair cross-complementing 1 expression decreases the DNA repair capacity to sensitize lung cancer cells to etoposide. Mol Cancer Ther. 11:561–571. 2012. View Article : Google Scholar

87 

Weng SH, Tseng SC, Huang YC, Chen HJ and Lin YW: Inhibition of thymidine phosphorylase expression by using an HSP90 inhibitor potentiates the cytotoxic effect of cisplatin in non-small-cell lung cancer cells. Biochem Pharmacol. 84:126–136. 2012. View Article : Google Scholar : PubMed/NCBI

88 

Katayama R, Khan TM, Benes C, et al: Therapeutic strategies to overcome crizotinib resistance in non-small cell lung cancers harboring the fusion oncogene EML4-ALK. Proc Natl Acad Sci USA. 108:7535–7540. 2011. View Article : Google Scholar : PubMed/NCBI

89 

Normant E, Paez G, West KA, et al: The Hsp90 inhibitor IPI-504 rapidly lowers EML4-ALK levels and induces tumor regression in ALK-driven NSCLC models. Oncogene. 30:2581–2586. 2011. View Article : Google Scholar : PubMed/NCBI

90 

Solit DB, Osman I, Polsky D, et al: Phase II trial of 17-allylamino-17-demethoxygeldanamycin in patients with metastatic melanoma. Clin Cancer Res. 14:8302–8307. 2008. View Article : Google Scholar : PubMed/NCBI

91 

Heuckmann JM, Balke-Want H, Malchers F, et al: Differential protein stability and ALK inhibitor sensitivity of EML4-ALK fusion variants. Clin Cancer Res. 18:4682–4690. 2012. View Article : Google Scholar : PubMed/NCBI

92 

Sang J, Acquaviva J, Friedland JC, et al: Targeted inhibition of the molecular chaperone Hsp90 overcomes ALK inhibitor resistance in non-small cell lung cancer. Cancer Discov. 3:430–443. 2013. View Article : Google Scholar : PubMed/NCBI

93 

Koll TT, Feis SS, Wright MH, et al: HSP90 inhibitor, DMAG, synergizes with radiation of lung cancer cells by interfering with base excision and ATM-mediated DNA repair. Mol Cancer Ther. 7:1985–1992. 2008. View Article : Google Scholar : PubMed/NCBI

94 

Segawa T, Fujii Y, Tanaka A, et al: Radiosensitization of human lung cancer cells by the novel purine-scaffold Hsp90 inhibitor, PU-H71. Int J Mol Med. 33:559–564. 2014.

95 

Kim YM and Pyo H: Cooperative enhancement of radiosensitivity after combined treatment of 17-(allylamino)-17-demethoxygeldanamycin and celecoxib in human lung and colon cancer cell lines. DNA Cell Biol. 31:15–29. 2012. View Article : Google Scholar :

96 

Sudo M, Chin TM, Mori S, et al: Inhibiting proliferation of gefitinib-resistant, non-small cell lung cancer. Cancer Chemother Pharmacol. 71:1325–1334. 2013. View Article : Google Scholar : PubMed/NCBI

97 

Wainberg ZA, Anghel A, Rogers AM, et al: Inhibition of HSP90 with AUY922 induces synergy in HER2-amplified trastuzumab-resistant breast and gastric cancer. Mol Cancer Ther. 12:509–519. 2013. View Article : Google Scholar : PubMed/NCBI

98 

Ono N, Yamazaki T, Nakanishi Y, et al: Preclinical antitumor activity of the novel heat shock protein 90 inhibitor CH5164840 against human epidermal growth factor receptor 2 (HER2)-overexpressing cancers. Cancer Sci. 103:342–349. 2012. View Article : Google Scholar

99 

Scaltriti M, Serra V, Normant E, et al: Antitumor activity of the Hsp90 inhibitor IPI-504 in HER2-positive trastuzumab-resistant breast cancer. Mol Cancer Ther. 10:817–824. 2011. View Article : Google Scholar : PubMed/NCBI

100 

Chandarlapaty S, Scaltriti M, Angelini P, et al: Inhibitors of HSP90 block p95-HER2 signaling in Trastuzumab-resistant tumors and suppress their growth. Oncogene. 29:325–334. 2010. View Article : Google Scholar

101 

Meng Q, Chen X, Sun L, Zhao C, Sui G and Cai L: Carbamazepine promotes Her-2 protein degradation in breast cancer cells by modulating HDAC6 activity and acetylation of Hsp90. Mol Cell Biochem. 348:165–171. 2011. View Article : Google Scholar

102 

Solit DB, Basso AD, Olshen AB, Scher HI and Rosen N: Inhibition of heat shock protein 90 function down-regulates Akt kinase and sensitizes tumors to Taxol. Cancer Res. 63:2139–2144. 2003.PubMed/NCBI

103 

Wong C and Chen S: Heat shock protein 90 inhibitors: new mode of therapy to overcome endocrine resistance. Cancer Res. 69:8670–8677. 2009. View Article : Google Scholar : PubMed/NCBI

104 

Wong C, Wang X, Smith D, Reddy K and Chen S: AKT-aro and HER2-aro, models for de novo resistance to aromatase inhibitors; molecular characterization and inhibitor response studies. Breast Cancer Res Treat. 134:671–681. 2012. View Article : Google Scholar : PubMed/NCBI

105 

Pashtan I, Tsutsumi S, Wang S, Xu W and Neckers L: Targeting Hsp90 prevents escape of breast cancer cells from tyrosine kinase inhibition. Cell Cycle. 7:2936–2941. 2008. View Article : Google Scholar : PubMed/NCBI

106 

Palacios C, Lopez-Perez AI and Lopez-Rivas A: Down-regulation of RIP expression by 17-dimethylaminoethylamino-17-demethoxygeldanamycin promotes TRAIL-induced apoptosis in breast tumor cells. Cancer Lett. 287:207–215. 2010. View Article : Google Scholar

107 

Stecklein SR, Kumaraswamy E, Behbod F, et al: BRCA1 and HSP90 cooperate in homologous and non-homologous DNA double-strand-break repair and G2/M checkpoint activation. Proc Natl Acad Sci USA. 109:13650–13655. 2012. View Article : Google Scholar : PubMed/NCBI

108 

Benezra R, Henke E, Ciarrocchi A, et al: Induction of complete regressions of oncogene-induced breast tumors in mice. Cold Spring Harb Symp Quant Biol. 70:375–381. 2005. View Article : Google Scholar

109 

Smith V, Hobbs S, Court W, Eccles S, Workman P and Kelland LR: ErbB2 overexpression in an ovarian cancer cell line confers sensitivity to the HSP90 inhibitor geldanamycin. Anticancer Res. 22:1993–1999. 2002.PubMed/NCBI

110 

Solar P, Horvath V, Kleban J, et al: Hsp90 inhibitor geldanamycin increases the sensitivity of resistant ovarian adenocarcinoma cell line A2780cis to cisplatin. Neoplasma. 54:127–130. 2007.PubMed/NCBI

111 

Sain N, Krishnan B, Ormerod MG, et al: Potentiation of paclitaxel activity by the HSP90 inhibitor 17-allylamino-17-demethoxygeldanamycin in human ovarian carcinoma cell lines with high levels of activated AKT. Mol Cancer Ther. 5:1197–1208. 2006. View Article : Google Scholar : PubMed/NCBI

112 

Mollapour M, Tsutsumi S, Donnelly AC, et al: Swe1 Wee1-dependent tyrosine phosphorylation of Hsp90 regulates distinct facets of chaperone function. Mol Cell. 37:333–343. 2010. View Article : Google Scholar : PubMed/NCBI

113 

Iwai A, Bourboulia D, Mollapour M, et al: Combined inhibition of Wee1 and Hsp90 activates intrinsic apoptosis in cancer cells. Cell Cycle. 11:3649–3655. 2012. View Article : Google Scholar : PubMed/NCBI

114 

Lamoureux F, Thomas C, Yin MJ, et al: Clusterin inhibition using OGX-011 synergistically enhances Hsp90 inhibitor activity by suppressing the heat shock response in castrate-resistant prostate cancer. Cancer Res. 71:5838–5849. 2011. View Article : Google Scholar : PubMed/NCBI

115 

O’Malley KJ, Langmann G, Ai J, Ramos-Garcia R, Vessella RL and Wang Z: Hsp90 inhibitor 17-AAG inhibits progression of LuCaP35 xenograft prostate tumors to castration resistance. Prostate. 72:1117–1123. 2012. View Article : Google Scholar

116 

Huang HC, Yang Y, Nanda A, Koria P and Rege K: Synergistic administration of photothermal therapy and chemotherapy to cancer cells using polypeptide-based degradable plasmonic matrices. Nanomedicine (Lond). 6:459–473. 2011. View Article : Google Scholar

117 

Lamottke B, Kaiser M, Mieth M, et al: The novel, orally bioavailable HSP90 inhibitor NVP-HSP990 induces cell cycle arrest and apoptosis in multiple myeloma cells and acts synergistically with melphalan by increased cleavage of caspases. Eur J Haematol. 88:406–415. 2012. View Article : Google Scholar : PubMed/NCBI

118 

Stuhmer T, Iskandarov K, Gao Z, et al: Preclinical activity of the novel orally bioavailable HSP90 inhibitor NVP-HSP990 against multiple myeloma cells. Anticancer Res. 32:453–462. 2012.PubMed/NCBI

119 

Chatterjee M, Andrulis M, Stuhmer T, et al: The PI3K/Akt signaling pathway regulates the expression of Hsp70, which critically contributes to Hsp90-chaperone function and tumor cell survival in multiple myeloma. Haematologica. 98:1132–1141. 2012. View Article : Google Scholar : PubMed/NCBI

120 

Kaiser M, Lamottke B, Mieth M, et al: Synergistic action of the novel HSP90 inhibitor NVP-AUY922 with histone deacetylase inhibitors, melphalan, or doxorubicin in multiple myeloma. Eur J Haematol. 84:337–344. 2010. View Article : Google Scholar

121 

Yasui H, Hideshima T, Ikeda H, et al: BIRB 796 enhances cytotoxicity triggered by bortezomib, heat shock protein (Hsp) 90 inhibitor, and dexamethasone via inhibition of p38 mitogen-activated protein kinase/Hsp27 pathway in multiple myeloma cell lines and inhibits paracrine tumour growth. Br J Haematol. 136:414–423. 2007. View Article : Google Scholar

122 

Weigert O, Lane AA, Bird L, et al: Genetic resistance to JAK2 enzymatic inhibitors is overcome by HSP90 inhibition. J Exp Med. 209:259–273. 2012. View Article : Google Scholar : PubMed/NCBI

123 

Fiskus W, Verstovsek S, Manshouri T, et al: Heat shock protein 90 inhibitor is synergistic with JAK2 inhibitor and overcomes resistance to JAK2-TKI in human myeloproliferative neoplasm cells. Clin Cancer Res. 17:7347–7358. 2011. View Article : Google Scholar : PubMed/NCBI

124 

Fiskus W, Rao R, Fernandez P, et al: Molecular and biologic characterization and drug sensitivity of pan-histone deacetylase inhibitor-resistant acute myeloid leukemia cells. Blood. 112:2896–2905. 2008. View Article : Google Scholar : PubMed/NCBI

125 

Kaufmann SH, Karp JE, Litzow MR, et al: Phase I and pharmacological study of cytarabine and tanespimycin in relapsed and refractory acute leukemia. Haematologica. 96:1619–1626. 2011. View Article : Google Scholar : PubMed/NCBI

126 

Best OG, Che Y, Singh N, Forsyth C, Christopherson RI and Mulligan SP: The Hsp90 inhibitor SNX-7081 synergizes with and restores sensitivity to fludarabine in chronic lymphocytic leukemia cells with lesions in the TP53 pathway: a potential treatment strategy for fludarabine refractory disease. Leuk Lymphoma. 53:1367–1375. 2012. View Article : Google Scholar

127 

Jones DT, Addison E, North JM, et al: Geldanamycin and herbimycin A induce apoptotic killing of B chronic lymphocytic leukemia cells and augment the cells’ sensitivity to cytotoxic drugs. Blood. 103:1855–1861. 2004. View Article : Google Scholar

128 

Walsby E, Pearce L, Burnett AK, Fegan C and Pepper C: The Hsp90 inhibitor NVP-AUY922-AG inhibits NF-kappaB signaling, overcomes microenvironmental cytoprotection and is highly synergistic with fludarabine in primary CLL cells. Oncotarget. 3:525–534. 2012.PubMed/NCBI

129 

Johnson AJ, Wagner AJ, Cheney CM, et al: Rituximab and 17-allylamino-17-demethoxygeldanamycin induce synergistic apoptosis in B-cell chronic lymphocytic leukaemia. Br J Haematol. 139:837–844. 2007. View Article : Google Scholar : PubMed/NCBI

130 

Blagosklonny MV, Fojo T, Bhalla KN, et al: The Hsp90 inhibitor geldanamycin selectively sensitizes Bcr-Abl-expressing leukemia cells to cytotoxic chemotherapy. Leukemia. 15:1537–1543. 2001. View Article : Google Scholar : PubMed/NCBI

131 

Peng C, Brain J, Hu Y, et al: Inhibition of heat shock protein 90 prolongs survival of mice with BCR-ABL-T315I-induced leukemia and suppresses leukemic stem cells. Blood. 110:678–685. 2007. View Article : Google Scholar : PubMed/NCBI

132 

Tauchi T, Okabe S, Ashihara E, Kimura S, Maekawa T and Ohyashiki K: Combined effects of novel heat shock protein 90 inhibitor NVP-AUY922 and nilotinib in a random mutagenesis screen. Oncogene. 30:2789–2797. 2011. View Article : Google Scholar : PubMed/NCBI

133 

Demidenko ZN, An WG, Lee JT, Romanova LY, McCubrey JA and Blagosklonny MV: Kinase-addiction and bi-phasic sensitivity-resistance of Bcr-Abl- and Raf-1-expressing cells to imatinib and geldanamycin. Cancer Biol Ther. 4:484–490. 2005. View Article : Google Scholar : PubMed/NCBI

134 

Roue G, Perez-Galan P, Mozos A, et al: The Hsp90 inhibitor IPI-504 overcomes bortezomib resistance in mantle cell lymphoma in vitro and in vivo by down-regulation of the prosur-vival ER chaperone BiP/Grp78. Blood. 117:1270–1279. 2011. View Article : Google Scholar

135 

Dungey FA, Caldecott KW and Chalmers AJ: Enhanced radiosensitization of human glioma cells by combining inhibition of poly(ADP-ribose) polymerase with inhibition of heat shock protein 90. Mol Cancer Ther. 8:2243–2254. 2009. View Article : Google Scholar : PubMed/NCBI

136 

Ohba S, Hirose Y, Yoshida K, Yazaki T and Kawase T: Inhibition of 90-kD heat shock protein potentiates the cytotoxicity of chemotherapeutic agents in human glioma cells. J Neurosurg. 112:33–42. 2010. View Article : Google Scholar

137 

Siegelin MD, Habel A and Gaiser T: 17-AAG sensitized malignant glioma cells to death-receptor mediated apoptosis. Neurobiol Dis. 33:243–249. 2009. View Article : Google Scholar

138 

Sauvageot CM, Weatherbee JL, Kesari S, et al: Efficacy of the HSP90 inhibitor 17-AAG in human glioma cell lines and tumorigenic glioma stem cells. Neuro Oncol. 11:109–121. 2009. View Article : Google Scholar :

139 

Paraiso KH, Haarberg HE, Wood E, et al: The HSP90 inhibitor XL888 overcomes BRAF inhibitor resistance mediated through diverse mechanisms. Clin Cancer Res. 18:2502–2514. 2012. View Article : Google Scholar : PubMed/NCBI

140 

Ito A, Saito H, Mitobe K, et al: Inhibition of heat shock protein 90 sensitizes melanoma cells to thermosensitive ferromagnetic particle-mediated hyperthermia with low Curie temperature. Cancer Sci. 100:558–564. 2009. View Article : Google Scholar : PubMed/NCBI

141 

Acquaviva J, Smith DL, Jimenez JP, et al: Overcoming acquired BRAF inhibitor resistance in melanoma via targeted inhibition of Hsp90 with ganetespib. Mol Cancer Ther. 13:353–363. 2014. View Article : Google Scholar : PubMed/NCBI

142 

Tatokoro M, Koga F, Yoshida S, et al: Potential role of Hsp90 inhibitors in overcoming cisplatin resistance of bladder cancer-initiating cells. Int J Cancer. 131:987–996. 2012. View Article : Google Scholar

143 

Yoshida S, Koga F, Tatokoro M, et al: Low-dose Hsp90 inhibitors tumor-selectively sensitize bladder cancer cells to chemoradiotherapy. Cell Cycle. 10:4291–4299. 2011. View Article : Google Scholar : PubMed/NCBI

144 

Sato A and Asano T, Ito K and Asano T: 17-Allylamino-17-demethoxygeldanamycin and ritonavir inhibit renal cancer growth by inhibiting the expression of heat shock factor-1. Int J Oncol. 41:46–52. 2012.PubMed/NCBI

145 

Ma L, Sato F, Sato R, et al: Dual targeting of heat shock proteins 90 and 70 promotes cell death and enhances the anticancer effect of chemotherapeutic agents in bladder cancer. Oncol Rep. 31:2482–2492. 2014.PubMed/NCBI

146 

Yin X, Zhang H, Lundgren K, Wilson L, Burrows F and Shores CG: BIIB021, a novel Hsp90 inhibitor, sensitizes head and neck squamous cell carcinoma to radiotherapy. Int J Cancer. 126:1216–1225. 2010.

147 

Misso G, Giuberti G, Lombardi A, et al: Pharmacological inhibition of HSP90 and ras activity as a new strategy in the treatment of HNSCC. J Cell Physiol. 228:130–141. 2013. View Article : Google Scholar

148 

Lee CH, Hong HM, Chang YY and Chang WW: Inhibition of heat shock protein (Hsp) 27 potentiates the suppressive effect of Hsp90 inhibitors in targeting breast cancer stem-like cells. Biochimie. 94:1382–1389. 2012. View Article : Google Scholar : PubMed/NCBI

149 

Newman B, Liu Y, Lee HF, Sun D and Wang Y: HSP90 inhibitor 17-AAG selectively eradicates lymphoma stem cells. Cancer Res. 72:4551–4561. 2012. View Article : Google Scholar : PubMed/NCBI

150 

Kelland LR, Sharp SY, Rogers PM, Myers TG and Workman P: DT-Diaphorase expression and tumor cell sensitivity to 17-allylamino, 17-demethoxygeldanamycin, an inhibitor of heat shock protein 90. J Natl Cancer Inst. 91:1940–1949. 1999. View Article : Google Scholar : PubMed/NCBI

151 

Gaspar N, Sharp SY, Pacey S, et al: Acquired resistance to 17-allylamino-17-demethoxygeldanamycin (17-AAG, tanespimycin) in glioblastoma cells. Cancer Res. 69:1966–1975. 2009. View Article : Google Scholar : PubMed/NCBI

152 

Zhang H, Neely L, Lundgren K, et al: BIIB021, a synthetic Hsp90 inhibitor, has broad application against tumors with acquired multidrug resistance. Int J Cancer. 126:1226–1234. 2010.

153 

Sharp SY, Boxall K, Rowlands M, et al: In vitro biological characterization of a novel, synthetic diaryl pyrazole resorcinol class of heat shock protein 90 inhibitors. Cancer Res. 67:2206–2216. 2007. View Article : Google Scholar : PubMed/NCBI

154 

Sharp SY, Prodromou C, Boxall K, et al: Inhibition of the heat shock protein 90 molecular chaperone in vitro and in vivo by novel, synthetic, potent resorcinylic pyrazole/isoxazole amide analogues. Mol Cancer Ther. 6:1198–1211. 2007. View Article : Google Scholar : PubMed/NCBI

155 

Gaspar N, Sharp SY, Eccles SA, et al: Mechanistic evaluation of the novel HSP90 inhibitor NVP-AUY922 in adult and pediatric glioblastoma. Mol Cancer Ther. 9:1219–1233. 2010. View Article : Google Scholar : PubMed/NCBI

156 

McCollum AK, Teneyck CJ, Sauer BM, Toft DO and Erlichman C: Up-regulation of heat shock protein 27 induces resistance to 17-allylamino-demethoxygeldanamycin through a glutathione-mediated mechanism. Cancer Res. 66:10967–10975. 2006. View Article : Google Scholar : PubMed/NCBI

157 

Sharp SY, Roe SM, Kazlauskas E, et al: Co-crystalization and in vitro biological characterization of 5-aryl-4-(5-substituted-2-4-dihydroxyphenyl)-1,2,3-thiadiazole Hsp90 inhibitors. PLoS One. 7:e446422012. View Article : Google Scholar : PubMed/NCBI

158 

McConnell JR and McAlpine SR: Heat shock proteins 27, 40, and 70 as combinational and dual therapeutic cancer targets. Bioorg Med Chem Lett. 23:1923–1928. 2013. View Article : Google Scholar : PubMed/NCBI

159 

Lamoureux F, Thomas C, Yin MJ, Fazli L, Zoubeidi A and Gleave ME: Suppression of heat shock protein 27 using OGX-427 induces endoplasmic reticulum stress and potentiates heat shock protein 90 inhibitors to delay castrate-resistant prostate cancer. Eur Urol. 66:145–155. 2013. View Article : Google Scholar

160 

Guo F, Rocha K, Bali P, et al: Abrogation of heat shock protein 70 induction as a strategy to increase antileukemia activity of heat shock protein 90 inhibitor 17-allylamino-demethoxy geldanamycin. Cancer Res. 65:10536–10544. 2005. View Article : Google Scholar : PubMed/NCBI

161 

Pimienta G, Herbert KM and Regan L: A compound that inhibits the HOP-Hsp90 complex formation and has unique killing effects in breast cancer cell lines. Mol Pharm. 8:2252–2261. 2011. View Article : Google Scholar : PubMed/NCBI

162 

Miyagawa T, Saito H, Minamiya Y, et al: Inhibition of Hsp90 and 70 sensitizes melanoma cells to hyperthermia using ferromagnetic particles with a low Curie temperature. Int J Clin Oncol. 19:722–730. 2014. View Article : Google Scholar

163 

Ronnen EA, Kondagunta GV, Ishill N, et al: A phase II trial of 17-(Allylamino)-17-demethoxygeldanamycin in patients with papillary and clear cell renal cell carcinoma. Invest New Drugs. 24:543–546. 2006. View Article : Google Scholar : PubMed/NCBI

164 

Heath EI, Hillman DW, Vaishampayan U, et al: A phase II trial of 17-allylamino-17-demethoxygeldanamycin in patients with hormone-refractory metastatic prostate cancer. Clin Cancer Res. 14:7940–7946. 2008. View Article : Google Scholar : PubMed/NCBI

165 

Sequist LV, Gettinger S, Senzer NN, et al: Activity of IPI-504, a novel heat-shock protein 90 inhibitor, in patients with molecularly defined non-small-cell lung cancer. J Clin Oncol. 28:4953–4960. 2010. View Article : Google Scholar : PubMed/NCBI

166 

Oh WK, Galsky MD, Stadler WM, et al: Multicenter phase II trial of the heat shock protein 90 inhibitor, retaspimycin hydrochloride (IPI-504), in patients with castration-resistant prostate cancer. Urology. 78:626–630. 2011. View Article : Google Scholar : PubMed/NCBI

167 

Richardson PG, Chanan-Khan AA, Lonial S, et al: Tanespimycin and bortezomib combination treatment in patients with relapsed or relapsed and refractory multiple myeloma: results of a phase 1/2 study. Br J Haematol. 153:729–740. 2011. View Article : Google Scholar : PubMed/NCBI

168 

Modi S, Stopeck A, Linden H, et al: HSP90 inhibition is effective in breast cancer: a phase II trial of tanespimycin (17-AAG) plus trastuzumab in patients with HER2-positive metastatic breast cancer progressing on trastuzumab. Clin Cancer Res. 17:5132–5139. 2011. View Article : Google Scholar : PubMed/NCBI

169 

Gartner EM, Silverman P, Simon M, et al: A phase II study of 17-allylamino-17-demethoxygeldanamycin in metastatic or locally advanced, unresectable breast cancer. Breast Cancer Res Treat. 131:933–937. 2012. View Article : Google Scholar

170 

Pacey S, Gore M, Chao D, et al: A phase II trial of 17-allylamino, 17-demethoxygeldanamycin (17-AAG, tanespimycin) in patients with metastatic melanoma. Invest New Drugs. 30:341–349. 2012. View Article : Google Scholar

171 

Socinski MA, Goldman J, El-Hariry I, et al: A multicenter phase II study of ganetespib monotherapy in patients with genotypically defined advanced non-small cell lung cancer. Clin Cancer Res. 19:3068–3077. 2013. View Article : Google Scholar : PubMed/NCBI

172 

Modi S, Saura C, Henderson C, et al: A multicenter trial evaluating retaspimycin HCL (IPI-504) plus trastuzumab in patients with advanced or metastatic HER2-positive breast cancer. Breast Cancer Res Treat. 139:107–113. 2013. View Article : Google Scholar : PubMed/NCBI

173 

Jhaveri K, Chandarlapaty S, Lake D, et al: A phase II open-label study of ganetespib, a novel heat shock protein 90 inhibitor for patients with metastatic breast cancer. Clin Breast Cancer. 14:154–160. 2014. View Article : Google Scholar : PubMed/NCBI

174 

Ui T, Morishima K, Saito S, et al: The HSP90 inhibitor 17-N-allylamino-17-demethoxy geldanamycin (17-AAG) synergizes with cisplatin and induces apoptosis in cisplatin-resistant esophageal squamous cell carcinoma cell lines via the Akt/XIAP pathway. Oncol Rep. 31:619–624. 2014.

175 

Akimoto T, Nonaka T, Harashima K, Sakurai H, Ishikawa H and Mitsuhashi N: Radicicol potentiates heat-induced cell killing in a human oesophageal cancer cell line: the Hsp90 chaperone complex as a new molecular target for enhancement of thermosensitivity. Int J Radiat Biol. 80:483–492. 2004. View Article : Google Scholar : PubMed/NCBI

176 

Burkitt M, Magee C, O’Connor D, Campbell F, Cornford P and Greenhalf W: Potentiation of chemotherapeutics by the Hsp90 antagonist geldanamycin requires a steady serum condition. Mol Carcinog. 46:466–475. 2007. View Article : Google Scholar : PubMed/NCBI

177 

Cao X, Bloomston M, Zhang T, et al: Synergistic antipancreatic tumor effect by simultaneously targeting hypoxic cancer cells with HSP90 inhibitor and glycolysis inhibitor. Clin Cancer Res. 14:1831–1839. 2008. View Article : Google Scholar : PubMed/NCBI

178 

Cao X, Jia G, Zhang T, et al: Non-invasive MRI tumor imaging and synergistic anticancer effect of HSP90 inhibitor and glycolysis inhibitor in RIP1-Tag2 transgenic pancreatic tumor model. Cancer Chemother Pharmacol. 62:985–994. 2008. View Article : Google Scholar : PubMed/NCBI

179 

Zhang T, Li Y, Zhu Z, Gu M, Newman B and Sun D: MEK inhibition potentiates the activity of Hsp90 inhibitor 17-AAG against pancreatic cancer cells. Mol Pharm. 7:1576–1584. 2010. View Article : Google Scholar : PubMed/NCBI

180 

Milanovic D, Firat E, Grosu AL and Niedermann G: Increased radiosensitivity and radiothermosensitivity of human pancreatic MIA PaCa-2 and U251 glioblastoma cell lines treated with the novel Hsp90 inhibitor NVP-HSP990. Radiat Oncol. 8:422013. View Article : Google Scholar : PubMed/NCBI

181 

Machida H, Matsumoto Y, Shirai M and Kubota N: Geldanamycin, an inhibitor of Hsp90, sensitizes human tumour cells to radiation. Int J Radiat Biol. 79:973–980. 2003. View Article : Google Scholar

182 

Vasilevskaya IA and O’Dwyer PJ: 17-Allylamino-17-demethoxygeldanamycin overcomes TRAIL resistance in colon cancer cell lines. Biochem Pharmacol. 70:580–589. 2005. View Article : Google Scholar : PubMed/NCBI

183 

Holmes JL, Sharp SY, Hobbs S and Workman P: Silencing of HSP90 cochaperone AHA1 expression decreases client protein activation and increases cellular sensitivity to the HSP90 inhibitor 17-allylamino-17-demethoxygeldanamycin. Cancer Res. 68:1188–1197. 2008. View Article : Google Scholar : PubMed/NCBI

184 

Saturno G, Valenti M, De Haven Brandon A, et al: Combining trail with PI3 kinase or HSP90 inhibitors enhances apoptosis in colorectal cancer cells via suppression of survival signaling. Oncotarget. 4:1185–1198. 2013.PubMed/NCBI

185 

He S, Smith DL, Sequeira M, Sang J, Bates RC and Proia DA: The HSP90 inhibitor ganetespib has chemosensitizer and radiosensitizer activity in colorectal cancer. Invest New Drugs. 32:577–586. 2014. View Article : Google Scholar : PubMed/NCBI

186 

Nguyen DM, Chen A, Mixon A and Schrump DS: Sequence-dependent enhancement of paclitaxel toxicity in non-small cell lung cancer by 17-allylamino 17-demethoxygeldanamycin. J Thorac Cardiovasc Surg. 118:908–915. 1999. View Article : Google Scholar : PubMed/NCBI

187 

Wen W, Liu W, Shao Y and Chen L: VER-155008, a small molecule inhibitor of HSP70 with potent anti-cancer activity on lung cancer cell lines. Exp Biol Med (Maywood). 239:638–645. 2014. View Article : Google Scholar

188 

Zismanov V, Drucker L and Gottfried M: ER homeostasis and motility of NSCLC cell lines can be therapeutically targeted with combined Hsp90 and HDAC inhibitors. Pulm Pharmacol Ther. 26:388–394. 2013. View Article : Google Scholar : PubMed/NCBI

189 

Zismanov V, Drucker L and Gottfried M: Combined inhibition of Hsp90 and the proteasome affects NSCLC proteostasis and attenuates cell migration. Anticancer Drugs. 25:998–1006. 2014. View Article : Google Scholar : PubMed/NCBI

190 

Chen Z, Akbay E, Mikse O, et al: Co-clinical trials demonstrate superiority of crizotinib to chemotherapy in ALK-rearranged non-small cell lung cancer and predict strategies to overcome resistance. Clin Cancer Res. 20:1204–1211. 2013. View Article : Google Scholar : PubMed/NCBI

191 

Miyajima N, Tsutsumi S, Sourbier C, et al: The HSP90 inhibitor ganetespib synergizes with the MET kinase inhibitor crizotinib in both crizotinib-sensitive and -resistant MET-driven tumor models. Cancer Res. 73:7022–7033. 2013. View Article : Google Scholar : PubMed/NCBI

192 

Smith DL, Acquaviva J, Sequeira M, et al: The HSP90 inhibitor ganetespib potentiates the antitumor activity of EGFR tyrosine kinase inhibition in mutant and wild-type non-small cell lung cancer. Target Oncol. Aug 1–2014.(Epub ahead of print). View Article : Google Scholar : PubMed/NCBI

193 

Ono N, Yamazaki T, Tsukaguchi T, et al: Enhanced antitumor activity of erlotinib in combination with the Hsp90 inhibitor CH5164840 against non-small-cell lung cancer. Cancer Sci. 104:1346–1352. 2013. View Article : Google Scholar : PubMed/NCBI

194 

Jang WJ, Jung SK, Kang JS, et al: Anti-tumor activity of WK88-1, a novel geldanamycin derivative, in gefitinib-resistant NSCLCs with Met amplification. Cancer Sci. Aug 13–2014.(Epub ahead of print). View Article : Google Scholar

195 

Ma Y, Lakshmikanthan V, Lewis RW and Kumar MV: Sensitization of TRAIL-resistant cells by inhibition of heat shock protein 90 with low-dose geldanamycin. Mol Cancer Ther. 5:170–178. 2006. View Article : Google Scholar : PubMed/NCBI

196 

Senju M, Sueoka N, Sato A, et al: Hsp90 inhibitors cause G2/M arrest associated with the reduction of Cdc25C and Cdc2 in lung cancer cell lines. J Cancer Res Clin Oncol. 132:150–158. 2006. View Article : Google Scholar

197 

Russell JS, Burgan W, Oswald KA, Camphausen K and Tofilon PJ: Enhanced cell killing induced by the combination of radiation and the heat shock protein 90 inhibitor 17-allylamino-17-demethoxygeldanamycin: a multitarget approach to radiosensitization. Clin Cancer Res. 9:3749–3755. 2003.PubMed/NCBI

198 

Bisht KS, Bradbury CM, Mattson D, et al: Geldanamycin and 17-allylamino-17-demethoxygeldanamycin potentiate the in vitro and in vivo radiation response of cervical tumor cells via the heat shock protein 90-mediated intracellular signaling and cytotoxicity. Cancer Res. 63:8984–8995. 2003.PubMed/NCBI

199 

Sawai A, Chandarlapaty S, Greulich H, et al: Inhibition of Hsp90 down-regulates mutant epidermal growth factor receptor (EGFR) expression and sensitizes EGFR mutant tumors to paclitaxel. Cancer Res. 68:589–596. 2008. View Article : Google Scholar : PubMed/NCBI

200 

Bai L, Xu S, Chen W, et al: Blocking NF-kappaB and Akt by Hsp90 inhibition sensitizes Smac mimetic compound 3-induced extrinsic apoptosis pathway and results in synergistic cancer cell death. Apoptosis. 16:45–54. 2011. View Article : Google Scholar :

201 

Mimnaugh EG, Xu W, Vos M, et al: Simultaneous inhibition of hsp 90 and the proteasome promotes protein ubiquitination, causes endoplasmic reticulum-derived cytosolic vacuolization, and enhances antitumor activity. Mol Cancer Ther. 3:551–566. 2004.PubMed/NCBI

202 

Raja SM, Clubb RJ, Bhattacharyya M, et al: A combination of Trastuzumab and 17-AAG induces enhanced ubiquitinylation and lysosomal pathway-dependent ErbB2 degradation and cytotoxicity in ErbB2-overexpressing breast cancer cells. Cancer Biol Ther. 7:1630–1640. 2008. View Article : Google Scholar : PubMed/NCBI

203 

Zsebik B, Citri A, Isola J, Yarden Y, Szollosi J and Vereb G: Hsp90 inhibitor 17-AAG reduces ErbB2 levels and inhibits proliferation of the trastuzumab resistant breast tumor cell line JIMT-1. Immunol Lett. 104:146–155. 2006. View Article : Google Scholar

204 

Roforth MM and Tan C: Combination of rapamycin and 17-allylamino-17-demethoxygeldanamycin abrogates Akt activation and potentiates mTOR blockade in breast cancer cells. Anticancer Drugs. 19:681–688. 2008. View Article : Google Scholar : PubMed/NCBI

205 

Munster PN, Basso A, Solit D, Norton L and Rosen N: Modulation of Hsp90 function by ansamycins sensitizes breast cancer cells to chemotherapy-induced apoptosis in an RB- and schedule-dependent manner. See: E A Sausville, Combining cytotoxics and 17-allylamino, 17-demethoxygeldanamycin: sequence and tumor biology matters. Clin Cancer Res. 7:2155–2158. 2001.

Clin Cancer Res. 7:2228–2236. 2001.

206 

Ferrario A, Rucker N, Wong S, Luna M and Gomer CJ: Survivin, a member of the inhibitor of apoptosis family, is induced by photodynamic therapy and is a target for improving treatment response. Cancer Res. 67:4989–4995. 2007. View Article : Google Scholar : PubMed/NCBI

207 

Hasenstein JR, Shin HC, Kasmerchak K, Buehler D, Kwon GS and Kozak KR: Antitumor activity of Triolimus: a novel multidrug-loaded micelle containing Paclitaxel, Rapamycin, and 17-AAG. Mol Cancer Ther. 11:2233–2242. 2012. View Article : Google Scholar : PubMed/NCBI

208 

Solar P, Ferenc P, Koval J, et al: Photoactivated hypericin induces downregulation of HER2 gene expression. Radiat Res. 175:51–56. 2011. View Article : Google Scholar

209 

Chen SM, Guo CL, Shi JJ, et al: HSP90 inhibitor AUY922 abrogates up-regulation of RTKs by mTOR inhibitor AZD8055 and potentiates its antiproliferative activity in human breast cancer. Int J Cancer. 135:2462–2474. 2014. View Article : Google Scholar : PubMed/NCBI

210 

Banerji U, Sain N, Sharp SY, et al: An in vitro and in vivo study of the combination of the heat shock protein inhibitor 17-allylamino-17-demethoxygeldanamycin and carboplatin in human ovarian cancer models. Cancer Chemother Pharmacol. 62:769–778. 2008. View Article : Google Scholar : PubMed/NCBI

211 

Gray PJ Jr, Stevenson MA and Calderwood SK: Targeting Cdc37 inhibits multiple signaling pathways and induces growth arrest in prostate cancer cells. Cancer Res. 67:11942–11950. 2007. View Article : Google Scholar : PubMed/NCBI

212 

Paduano F, Villa R, Pennati M, et al: Silencing of survivin gene by small interfering RNAs produces supra-additive growth suppression in combination with 17-allylamino-17-demethoxygeldanamycin in human prostate cancer cells. Mol Cancer Ther. 5:179–186. 2006. View Article : Google Scholar : PubMed/NCBI

213 

Gandhi N, Wild AT, Chettiar ST, et al: Novel Hsp90 inhibitor NVP-AUY922 radiosensitizes prostate cancer cells. Cancer Biol Ther. 14:347–356. 2013. View Article : Google Scholar : PubMed/NCBI

214 

Ku S, Lasorsa E, Adelaiye R, Ramakrishnan S, Ellis L and Pili R: Inhibition of hsp90 augments docetaxel therapy in castrate resistant prostate cancer. PLoS One. 9:e1036802014. View Article : Google Scholar : PubMed/NCBI

215 

Zhao M, Ma J, Zhu HY, et al: Apigenin inhibits proliferation and induces apoptosis in human multiple myeloma cells through targeting the trinity of CK2, Cdc37 and Hsp90. Mol Cancer. 10:1042011. View Article : Google Scholar : PubMed/NCBI

216 

Shammas MA, Koley H, Bertheau RC, et al: Telomerase inhibitor GRN163L inhibits myeloma cell growth in vitro and in vivo. Leukemia. 22:1410–1418. 2008. View Article : Google Scholar : PubMed/NCBI

217 

Duus J, Bahar HI, Venkataraman G, et al: Analysis of expression of heat shock protein-90 (HSP90) and the effects of HSP90 inhibitor (17-AAG) in multiple myeloma. Leuk Lymphoma. 47:1369–1378. 2006. View Article : Google Scholar : PubMed/NCBI

218 

Francis LK, Alsayed Y, Leleu X, et al: Combination mammalian target of rapamycin inhibitor rapamycin and HSP90 inhibitor 17-allylamino-17-demethoxygeldanamycin has synergistic activity in multiple myeloma. Clin Cancer Res. 12:6826–6835. 2006. View Article : Google Scholar : PubMed/NCBI

219 

Cervantes-Gomez F, Nimmanapalli R and Gandhi V: ATP analog enhances the actions of a heat shock protein 90 inhibitor in multiple myeloma cells. J Pharmacol Exp Ther. 339:545–554. 2011. View Article : Google Scholar : PubMed/NCBI

220 

Sydor JR, Normant E, Pien CS, et al: Development of 17-allylamino-17-demethoxygeldanamycin hydroquinone hydrochloride (IPI-504), an anti-cancer agent directed against Hsp90. Proc Natl Acad Sci USA. 103:17408–17413. 2006. View Article : Google Scholar : PubMed/NCBI

221 

Heimberger T, Andrulis M, Riedel S, et al: The heat shock transcription factor 1 as a potential new therapeutic target in multiple myeloma. Br J Haematol. 160:465–476. 2012. View Article : Google Scholar : PubMed/NCBI

222 

Sugimoto K, Sasaki M, Isobe Y, et al: Hsp90-inhibitor geldanamycin abrogates G2 arrest in p53-negative leukemia cell lines through the depletion of Chk1. Oncogene. 27:3091–3101. 2008. View Article : Google Scholar

223 

Haaland I, Opsahl JA, Berven FS, et al: Molecular mechanisms of nutlin-3 involve acetylation of p53, histones and heat shock proteins in acute myeloid leukemia. Mol Cancer. 13:1162014. View Article : Google Scholar : PubMed/NCBI

224 

Rahmani M, Yu C, Dai Y, et al: Coadministration of the heat shock protein 90 antagonist 17-allylamino-17-demethoxygeldanamycin with suberoylanilide hydroxamic acid or sodium butyrate synergistically induces apoptosis in human leukemia cells. Cancer Res. 63:8420–8427. 2003.PubMed/NCBI

225 

Jia W, Yu C, Rahmani M, et al: Synergistic antileukemic interactions between 17-AAG and UCN-01 involve interruption of RAF/MEK- and AKT-related pathways. Blood. 102:1824–1832. 2003. View Article : Google Scholar : PubMed/NCBI

226 

Yao Q, Nishiuchi R, Kitamura T and Kersey JH: Human leukemias with mutated FLT3 kinase are synergistically sensitive to FLT3 and Hsp90 inhibitors: the key role of the STAT5 signal transduction pathway. Leukemia. 19:1605–1612. 2005. View Article : Google Scholar : PubMed/NCBI

227 

Mesa RA, Loegering D, Powell HL, et al: Heat shock protein 90 inhibition sensitizes acute myelogenous leukemia cells to cytarabine. Blood. 106:318–327. 2005. View Article : Google Scholar : PubMed/NCBI

228 

George P, Bali P, Annavarapu S, et al: Combination of the histone deacetylase inhibitor LBH589 and the hsp90 inhibitor 17-AAG is highly active against human CML-BC cells and AML cells with activating mutation of FLT-3. Blood. 105:1768–1776. 2005. View Article : Google Scholar

229 

George P, Bali P, Cohen P, et al: Cotreatment with 17-allylamino-demethoxygeldanamycin and FLT-3 kinase inhibitor PKC412 is highly effective against human acute myelogenous leukemia cells with mutant FLT-3. Cancer Res. 64:3645–3652. 2004. View Article : Google Scholar : PubMed/NCBI

230 

Radujkovic A, Schad M, Topaly J, et al: Synergistic activity of imatinib and 17-AAG in imatinib-resistant CML cells overexpressing BCR-ABL - Inhibition of P-glycoprotein function by 17-AAG. Leukemia. 19:1198–1206. 2005. View Article : Google Scholar : PubMed/NCBI

231 

Rao R, Fiskus W, Yang Y, et al: HDAC6 inhibition enhances 17-AAG--mediated abrogation of hsp90 chaperone function in human leukemia cells. Blood. 112:1886–1893. 2008. View Article : Google Scholar : PubMed/NCBI

232 

McCaig AM, Cosimo E, Leach MT and Michie AM: Dasatinib inhibits B cell receptor signalling in chronic lymphocytic leukaemia but novel combination approaches are required to overcome additional pro-survival microenvironmental signals. Br J Haematol. 153:199–211. 2011. View Article : Google Scholar : PubMed/NCBI

233 

Walsby EJ, Lazenby M, Pepper CJ, Knapper S and Burnett AK: The HSP90 inhibitor NVP-AUY922-AG inhibits the PI3K and IKK signalling pathways and synergizes with cytarabine in acute myeloid leukaemia cells. Br J Haematol. 161:57–67. 2013. View Article : Google Scholar : PubMed/NCBI

234 

Best OG and Mulligan SP: Heat shock protein-90 inhibitor, NVP-AUY922, is effective in combination with fludarabine against chronic lymphocytic leukemia cells cultured on CD40L-stromal layer and inhibits their activated/proliferative phenotype. Leuk Lymphoma. 53:2314–2320. 2012. View Article : Google Scholar : PubMed/NCBI

235 

Robles AI, Wright MH, Gandhi B, et al: Schedule-dependent synergy between the heat shock protein 90 inhibitor 17-(dimethylaminoethylamino)-17-demethoxygeldanamycin and doxorubicin restores apoptosis to p53-mutant lymphoma cell lines. Clin Cancer Res. 12:6547–6556. 2006. View Article : Google Scholar : PubMed/NCBI

236 

Rao R, Lee P, Fiskus W, et al: Co-treatment with heat shock protein 90 inhibitor 17-dimethylaminoethylamino-17-demethoxygeldanamycin (DMAG) and vorinostat: a highly active combination against human mantle cell lymphoma (MCL) cells. Cancer Biol Ther. 8:1273–1280. 2009. View Article : Google Scholar : PubMed/NCBI

237 

Abramson JS, Chen W, Juszczynski P, et al: The heat shock protein 90 inhibitor IPI-504 induces apoptosis of AKT-dependent diffuse large B-cell lymphomas. Br J Haematol. 144:358–366. 2009. View Article : Google Scholar

238 

Premkumar DR, Arnold B, Jane EP and Pollack IF: Synergistic interaction between 17-AAG and phosphatidylinositol 3-kinase inhibition in human malignant glioma cells. Mol Carcinog. 45:47–59. 2006. View Article : Google Scholar

239 

Premkumar DR, Arnold B and Pollack IF: Cooperative inhibitory effect of ZD1839 (Iressa) in combination with 17-AAG on glioma cell growth. Mol Carcinog. 45:288–301. 2006. View Article : Google Scholar : PubMed/NCBI

240 

Jane EP and Pollack IF: The heat shock protein antagonist 17-AAG potentiates the activity of enzastaurin against malignant human glioma cells. Cancer Lett. 268:46–55. 2008. View Article : Google Scholar : PubMed/NCBI

241 

Choi EJ, Cho BJ, Lee DJ, et al: Enhanced cytotoxic effect of radiation and temozolomide in malignant glioma cells: targeting PI3K-AKT-mTOR signaling, HSP90 and histone deacetylases. BMC Cancer. 14:172014. View Article : Google Scholar : PubMed/NCBI

242 

Bull EE, Dote H, Brady KJ, et al: Enhanced tumor cell radiosensitivity and abrogation of G2 and S phase arrest by the Hsp90 inhibitor 17-(dimethylaminoethylamino)-17-demethoxygeldanamycin. Clin Cancer Res. 10:8077–8084. 2004. View Article : Google Scholar : PubMed/NCBI

243 

Wachsberger PR, Lawrence YR, Liu Y, et al: Hsp90 inhibition enhances PI-3 kinase inhibition and radiosensitivity in glioblastoma. J Cancer Res Clin Oncol. 140:573–582. 2014. View Article : Google Scholar : PubMed/NCBI

244 

Gallerne C, Prola A and Lemaire C: Hsp90 inhibition by PU-H71 induces apoptosis through endoplasmic reticulum stress and mitochondrial pathway in cancer cells and overcomes the resistance conferred by Bcl-2. Biochim Biophys Acta. 1833:1356–1366. 2013. View Article : Google Scholar : PubMed/NCBI

245 

Lin CC, Tu CF, Yen MC, et al: Inhibitor of heat-shock protein 90 enhances the antitumor effect of DNA vaccine targeting clients of heat-shock protein. Mol Ther. 15:404–410. 2007. View Article : Google Scholar : PubMed/NCBI

246 

Roh JL, Kim EH, Park HB and Park JY: The Hsp90 inhibitor 17-(allylamino)-17-demethoxygeldanamycin increases cisplatin antitumor activity by inducing p53-mediated apoptosis in head and neck cancer. Cell Death Dis. 4:e9562013. View Article : Google Scholar : PubMed/NCBI

247 

Bagatell R, Beliakoff J, David CL, Marron MT and Whitesell L: Hsp90 inhibitors deplete key anti-apoptotic proteins in pediatric solid tumor cells and demonstrate synergistic anticancer activity with cisplatin. Int J Cancer. 113:179–188. 2005. View Article : Google Scholar

248 

Fu W, Sharma SS, Ma L, et al: Apoptosis of osteosarcoma cultures by the combination of the cyclin-dependent kinase inhibitor SCH727965 and a heat shock protein 90 inhibitor. Cell Death Dis. 4:e5662013. View Article : Google Scholar : PubMed/NCBI

249 

Mimnaugh EG, Xu W, Vos M, Yuan X and Neckers L: Endoplasmic reticulum vacuolization and valosin-containing protein relocalization result from simultaneous hsp90 inhibition by geldanamycin and proteasome inhibition by velcade. Mol Cancer Res. 4:667–681. 2006. View Article : Google Scholar : PubMed/NCBI

250 

Martins AS, Ordonez JL, Garcia-Sanchez A, et al: A pivotal role for heat shock protein 90 in Ewing sarcoma resistance to anti-insulin-like growth factor 1 receptor treatment: in vitro and in vivo study. Cancer Res. 68:6260–6270. 2008. View Article : Google Scholar : PubMed/NCBI

251 

Neznanov N, Gorbachev AV, Neznanova L, et al: Anti-malaria drug blocks proteotoxic stress response: anti-cancer implications. Cell Cycle. 8:3960–3970. 2009. View Article : Google Scholar : PubMed/NCBI

252 

Enomoto A, Fukasawa T, Takamatsu N, et al: The HSP90 inhibitor 17-allylamino-17-demethoxygeldanamycin modulates radiosensitivity by downregulating serine/threonine kinase 38 via Sp1 inhibition. Eur J Cancer. 49:3547–3558. 2013. View Article : Google Scholar : PubMed/NCBI

253 

Chinn DC, Holland WS, Yoon JM, Zwerdling T and Mack PC: Anti-tumor activity of the HSP90 inhibitor SNX-2112 in pediatric cancer cell lines. Pediatr Blood Cancer. 58:885–890. 2012. View Article : Google Scholar

Related Articles

Journal Cover

March-2015
Volume 46 Issue 3

Print ISSN: 1019-6439
Online ISSN:1791-2423

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Solárová Z, Mojžiš J and Solár P: Hsp90 inhibitor as a sensitizer of cancer cells to different therapies (Review). Int J Oncol 46: 907-926, 2015
APA
Solárová, Z., Mojžiš, J., & Solár, P. (2015). Hsp90 inhibitor as a sensitizer of cancer cells to different therapies (Review). International Journal of Oncology, 46, 907-926. https://doi.org/10.3892/ijo.2014.2791
MLA
Solárová, Z., Mojžiš, J., Solár, P."Hsp90 inhibitor as a sensitizer of cancer cells to different therapies (Review)". International Journal of Oncology 46.3 (2015): 907-926.
Chicago
Solárová, Z., Mojžiš, J., Solár, P."Hsp90 inhibitor as a sensitizer of cancer cells to different therapies (Review)". International Journal of Oncology 46, no. 3 (2015): 907-926. https://doi.org/10.3892/ijo.2014.2791