Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2016 Feb 6.
Published in final edited form as: Oncogene. 2014 Oct 27;34(32):4153–4161. doi: 10.1038/onc.2014.349

Hsp70 in cancer: back to the future

Michael Y Sherman 1, Vladimir L Gabai 1
PMCID: PMC4411196  NIHMSID: NIHMS630502  PMID: 25347739

Abstract

Mechanistic studies from cell culture and animal models have revealed critical roles for the heat shock protein Hsp70 in cancer initiation and progression. Surprisingly, many effects of Hsp70 on cancer have not been related to its chaperone activity, but rather to its role(s) in regulating cell signaling. A major factor that directs Hsp70 signaling activity appears to be the co-chaperone Bag3. Here, we review these recent breakthroughs, and how these discoveries drive drug development efforts.

Keywords: BAG-3, cancer transformation and progression, senescence, proteotoxicity

Introduction

In this review we provide an overview of the roles of Hsp70 in cancer, spanning the initial observations that the levels of the heat shock protein Hsp70 are elevated in many human tumors, to dissecting mechanisms of its action in cancer, and finally to development novel anticancer compounds.

Hsp70 is a founding member of the highly conserved Hsp70 family of molecular chaperones, which are found in every membranous organelle of all cells 1. There are two abundant family members found in the cytosol of mammalian cells, a constitutively expressed HSPA8 (Hsc70, Hsc73) and an inducible HSPA1A (Hsp70, Hsp72). Several minor family members also exist (e.g. GRP75 or GRP78), but will not be discussed here. Hsc73 is an indispensable chaperone, which under homeostatic conditions is involved in folding, translocation across membranes, and degradation of proteins 2. Unlike Hsc73, Hsp70 is normally kept at low levels, and is induced under protein damaging conditions such as heat shock, oxidative stress, hypoxia or heave metals, functioning to provide resistance to a variety of proteotoxic stresses 3, 4. Little is known about the functional specificity of Hsp70 family members, and until recently it has been assumed that these proteins function redundantly. Recent observations, however, suggest that these proteins differ in their ability to interact with the ubiquitin ligase CHIP, leading to differential effects on protein degradation, such as the degradation of tau, an important factor contributing to Alzheimer’s disease 5. With cancer there also appears to be specificity for Hsp70 family members, as depletion or knockout of Hsp70 leads to profound anti-cancer effects, despite of the presence of Hsc73 (see below), though the molecular mechanisms underlying this specificity remain obscure.

Hsp70 is overexpressed in cancers, but not always

About two decades ago Daniel Ciocca and colleagues reported that expression of Hsp70 is highly elevated in breast cancer 6. Later this observation was extended for other types of tumors, including colon 7, liver 8, 9, prostate 10,11, esophagus 12, cervix 13,14 (see also 15 for review). These findings led to numerous attempts to develop Hsp70 as a biomarker of tumor stage, metastases, or prognosis. In several human cancers, prognostic significance of Hsp70 has proven quite remarkable, and independent on others prognostic factors 16,16,17,18,19,20. For example, in breast cancer without regional metastases at the time of diagnoses, 70% of patients with low levels of Hsp70 expression survive for 5 years, comparing with 30% survival of patients with high levels of Hsp70 expression 6. Despite substantial work in this field, Hsp70 has not yet been accepted as a clinical biomarker of any cancer parameter. Nevertheless, these studies have boosted investigations into the role(s) of Hsp70 in cancer, and have led to important developments in our understanding of etiology of malignancies, defining new avenues of drug design.

Though Hsp70 is overexpressed in most human cancers, there are some tumor types were lower expression of Hsp70 is observed compared to adjusted normal tissues, and correlation between Hsp70 levels and survival is lacking (e.g., in renal cancer 21, or cervix 14). Moreover, an inverse correlation between Hsp70 expression and patient’s prognosis in observed in some cancers, such as squamous cell carcinoma 22, cholangiocarcinoma 23, oral 24, or lung cancer 25.

One possible explanation for the negative correlation is that Hsp70 can suppress inflammation, likely by inhibiting the NF-kB pathway 2730. Accordingly, in cancers where the inflammatory component is critical for their development, a decrease rather than increase in the levels of Hsp70 would promote cancer. Indeed, in mouse colon carcinoma caused by a carcinogen DSS, which is accompanied by severe inflammation, knockout of Hsp70 increased inflammatory cytokines and aggravated cancer 31

Another possible role for Hsp70 in suppression of cancer is related to its immune-modulating properties 32. For instance, in squamous cell carcinoma, where high expression of Hsp70 associates with a good prognosis, there was a significant positive correlation between Hsp70 levels and lymphocyte infiltration which usually indicates higher anti-tumor immune response and is a favorable prognostic factor 22. The immune-modulating activity of Hsp70 is related to its extracellular function as immune stimulator (see 32 for review). Indeed, extracellular Hsp70 may activate innate immune system and can be used as adjuvant for tumor antigens, the property which is currently widely used in development of anticancer vaccines 33, 34. Thus, one may suggest that in tumors with inverse correlation between expression Hsp70 and prognosis, stronger immune response to Hsp70-tumor antigen complexes (released from cells or presented on the cell surface) is involved. Discussion of immune-modulating functions of Hsp70, including the role of extracellular Hsp70 in innate immunity and membrane-associated Hsp70 in sensitivity to NK cells, is out of the scope of this review, and here we will discuss cancers whose progression associates with elevated levels of Hsp70.

Dependence of tumor cells on Hsp70

Observations that certain types of tumors have elevated levels of Hsp70, which correlates with cancer grade and prognosis, suggested that Hsp70 could be involved in critical biochemical or genetic alterations that take place upon malignant transformation and further cancer development. In the mid-90s, when the phenomenon of apoptosis became the focus of biomedical research, it was discovered that Hsp70 potently suppresses apoptosis 3638. Since tumor cells live under conditions of continuous stress, e.g. hypoxia, nutrient deprivation or low pH, all of which are potent inducers of apoptosis, development of tumor must require adaptations that suppress apoptosis 39. Accordingly, it was suggested that elevated levels of Hsp70 are critical to cancer cells to combat these harsh conditions and suppress apoptosis. These early studies demonstrated that, surprisingly, besides its molecular chaperone function, Hsp70 also plays a special role in apoptotic signal transduction. Indeed, Hsp70 could prevent apoptosis in response to a variety of conditions that do not cause protein damage, such as cisplatin 40, 41 or TNF 42. Moreover, mutants of Hsp70 lacking chaperone function still retained their ability to effectively suppress the TNF-induced apoptosis 42 indicating that there must be a special function of Hsp70 in the apoptotic signal transduction. Indeed, it was found that Hsp70 suppresses apoptotic signaling at several points, including suppression of stress-activated kinases JNK and p38 37, 38, prevention of translocation of Bax or Bid to mitochondria 42,43, and suppression of the apoptosome formation 44. In line with these observations, depletion of Hsp70 sensitized to drug-induced apoptosis in myc-expressing lymphoid cells 45.

Apoptosis is a form of cell death which can be easily induced in lymphoid cells, but it plays a lesser role in epithelial cells from which most of human tumors originate. In some of these cancer cells of epithelial origin Hsp70 suppresses autophagic cell death which is independent on caspases and is not suppressed by Bcl-2 46. This form of cell death apparently associates with permeabilization of lysosomal membranes and release of lysosomal enzymes such as cathepsin 47. In spite of these conceptual advances, the lack of animal modeling makes it hard to conclude whether the anti-apoptotic or anti-autophagic death function of Hsp70 is important for cancer development.

While protection from harsh conditions of tumor microenviroment was initially considered the main reason for Hsp70 overexpression in tumors, a surprising observation was later made by the Jaattela group, that elevated levels of Hsp70 in unstressed cancer cells are required for their growth even under normal conditions 46, 48. This observation was further extended to several cancer cell lines, and now it is firmly established that tumor cells, in contrast to normal cells, require Hsp70 for their survival and growth (see ref 49, 50 for recent review). In searching for the mechanism of this dependence, we found that knockdown of Hsp70 in certain tumor epithelial cell lines can cause senescence. Senescence is an irreversible growth arrest with specific cell morphology (e.g., enlargement and flattening) and biochemistry (e.g., senescence-associated secretory phenotype and expression of SA-β-galactosidase) 51, 52. While apoptotic cells commit suicide, senescent cells stay alive for a long time in vitro, but they can be effectively eliminated in vivo by the innate immune system 53. Senescence is considered to evolve as an intrinsic mechanism to prevent propagation of cells with unrepaired DNA or cells that express major oncogenes (oncogene-induced senescence, OIS) 54. We hypothesized that Hsp70, being upregulated in tumor cells, may allow cells to a bypass the OIS barrier 55. This idea was based on the observation that whereas knockdown of Hsp70 in normal human mammary epithelial cells does not affect their growth, it causes senescence in the cells transformed with RAS, Her2, or PIK3CA oncogenes (Fig. 1) 55.

Fig. 1.

Fig. 1

Tumor cells overexpress Hsp70 to prevent oncogene-induced senescence.

The bypass of OIS modulated by Hsp70 in transformed cells may involve the p53-p21 pathway. Indeed, in tumor cells with wild type p53, Hsp70 knockdown stabilized p53 and increased p21 levels, while in cells with p53 mutation there was no increase in p21 56. Since p53 is frequently mutated in a variety of tumor types, it is possible that disabling p53-mediated senescence may release tumors from their Hsp70 dependence. However, even tumors with mutant p53 retain their dependence on Hsp70 because of the p53-independent pathways of senescence 55. Our finding suggested that these alternative senescence pathways activated upon Hsp70 depletion in cancer cells with mutant p53 could associate with decrease in the cell cycle regulator survivin or activation of the ERK pathway 55.

Hsp70 is critical for cancer development

Following correlative human population studies and cell culture experiments, the ultimate question remained whether Hsp70 is mechanistically involved in tumor development. Recent studies with animal cancer models provided experimental evidence in favor of this possibility. For example, expression Her2 (NeuT) oncogene in mammary epithelial cells triggered invasive breast cancer with 100% penetrance in control animals, while in Hsp70 knockout mice expression of Her2 rarely led to development of cancer 57. Lack of tumors in the knockout animals indicated that Hsp70 is essential for early stages of tumor development, consistent with its role in suppression of the oncogene-induced senescence. Further supporting this notion was the finding that Hsp70 knockout suppressed hyperplasia of mammary epithelium induced by Her2, which is normally seen in young mice long before tumor development. Furthermore, ducts and alveoli in Her2-expressing Hsp70 knockout mice were filled with senescent cells, while without Her2 expression these mice had normal mammary tissue 57. This experiment indicated that endogenous Hsp70 prevents OIS in the model of Her2-positive breast cancer and thus allows tumor emergence.

The role of Hsp70 in regulation of OIS was further corroborated by a recent publication that did not find significant effects of Hsp70 knockout on either survival or tumor emergence in a model that combined Ras and E6 oncogenes 58. In fact, E6 downregulates p53 and effectively reduces expression of its downstream target p21 59. Accordingly OIS was also suppressed and could not be triggered by the Hsp70 knockout in this model, and therefore such a knockout could not inhibit tumor emergence. Similarly in polyoma middle T antigen (PyMT)–induced breast cancer model, in which minimal OIS was seen 60, Hsp70 knockout did not reduce tumor emergence (Gong, personal communication).

Since bypassing OIS is critical during the early steps of tumor initiation, Hsp70 should be seen at elevated levels in low grade tumors. While this has been reported, curiously, the levels of Hsp70 increase further in high grade cancers 15, 61 suggesting that Hsp70 may play distinct role(s) in later stages of cancer progression that may require even higher levels of the chaperone. Since in a model of Her2-positive breast cancer Hsp70 is essential for tumor initiation, to assess the effect on metastasis, a distinct breast cancer model lacking OIS must be used. Indeed in the PyMT model that shows little OIS, Hsp70 knockout dramatically suppressed metastasis (Gong et al., submitted). Two independent factors appear to contribute to the effects of the Hsp70 knockout. First, the lack of Hsp70 leads to a depletion of cancer stem cells, and tumors emerging in PyMT/hsp70−/− mice are poorly transplanted into other host mice. Second, the overall population of tumor cells demonstrates reduced motility and invasion (Gong et al., submitted). These genetics data indicate that Hsp70 is critical for both tumor initiation and metastasis.

The problems with “non-oncogene addiction” hypothesis

Recent studies on sequencing the genomes of diverse human tumors did not reveal mutations or amplifications of Hsp70 (though certain SNPs in the Hsp70 gene HspA1A show association with cancer 62,63,64). Thus, these observations imply that regulation of Hsp70 expressions should occur mainly at transcriptional and/or translational levels. Accordingly, data from the human cancer microarrays database Oncomine show that in many tumors Hsp70 mRNA levels are increased compared to corresponding normal tissues. This induction of Hsp70 may be associated with increase in levels and activity of the heat shock transcription factor Hsf1, the master regulator of transcription of heat shock proteins, which is also overexpressed in many cancers 65. In line with these considerations various noxious factors of tumor microenvironment, like hypoxia or lack of energy, can cause proteotoxic stress, activate Hsf1, and induce Hsp70 (see ref 66 for review). For example, during progressive growth of ascites tumor cells in vivo, which associates with development of hypoxia and decrease in ATP, we observed increased levels of major Hsps, including Hsp70 67). However, the idea that proteotoxic stress in cancer can be responsible for upregulation of Hsp70, being plausible during in vivo tumor growth, cannot explain why tumor cell lines cultured without a stress in vitro also have elevated levels of Hsp70.

An attractive hypothesis of “non-oncogene addiction” has been proposed a number of years ago to explain the dependence of cancer on high levels of chaperones 68. According to this idea, cancer cells, which are often aneuploid, have a disbalance in protein complexes, and therefore increased demand for molecular chaperones. In other words, cancer cells experience intrinsic proteotoxic stress which leads to permanent activation of Hsf1 that enhances expression of chaperones. Such permanent proteotoxic stress leads to “addiction” of cancer cells to Hsf1, Hsp70 and other Hsps, which become essential for their survival. Thus, it was suggested that Hsp70, not being an oncogene by itself, is essential because it prevents an intrinsic proteotoxicity in tumor cells. This idea is widely accepted in the field and is supported by certain evidences. For example, aneuploid strains of yeast are more sensitive to conditions that interfere with protein translation, folding and degradation 69,70. Also, mammalian cell strains that have certain extra chromosomes demonstrate higher levels of Hsp70 and increased sensitivity to inhibitors of the chaperone Hsp90 71.

The concept of the requirement for Hsp70 in cancer cells due to aneuploidy and associated increased demand for chaperones directly predicts that upon knockdown of Hsp70, cancer cells must experience proteotoxic stress. However, as we have recently found, depletion from cultured breast or cervix human cancer cells of the major chaperone Hsp70 did not generate proteotoxic stress 72. As a measure of proteotoxic stress, several criteria were used in these experiments, including (a) levels of ubiquitinated proteins, (b) activation of Hsf1, (c) activation of Nrf2, a transcription factor that responds to oxidative and proteasome stress, and (d) luciferase refolding in vivo 72. Similar data were reported with depletion of another major chaperone the member of the small heat shock protein family Hsp27 73. Importantly, upon depletion of either Hsp70 or Hsp27, a dramatic reduction of cell growth 56 and massive senescence were seen 73. As noted previously, these effects were specific to transformed cells and were not seen in normal breast epithelial MCF10A cells 55, 57. These data directly indicate that higher levels of Hsp70 and “addiction” to this and other chaperones is unrelated to intrinsic proteotoxicity in cancer cells and must have a different underlying mechanism, e.g. regulation of cancer cell signaling (see next section).

What could be the mechanisms of chaperone induction in tumor cells beside intrinsic proteotoxic stress? Recent data demonstrate that such induction could be directly associated with oncogene signaling. For example, in early studies it was demonstrated that the major oncogene c-myc can induce transcription of Hsp70 74. Induction of Hsp70 was also seen upon expression of RET oncogene involved in thyroid cancer 75 and BCR-Abl responsible for chronic myelogenous leukemia 76. Similarly, the Her2 oncogene can activate Hsf1 and induce Hsp70 57. Interestingly, heregulin, a ligand of related tyrosine kinase receptors ErbB1 and ErbB3, can also activate Hsf1 77. Phosphopylation of Hsf1 by mTOR, a downstream component of the ErbB pathway, was found to activate Hsf1, indicating a direct role for signaling (rather than proteotoxicity) in Hsf1 regulation 78. The mTOR pathway is also frequently upregulated in cancer by oncogenic PIK3CA and PTEN mutations 79, which may also contribute to Hsf1 activity. Of note, it was shown that mutations in the p53 tumor suppressor also lead to activation of Hsf1 via the EGFR/ErbB2 pathway, suggesting that this pathway may integrate various signals to regulate Hsf1, and subsequently expression of Hsp70 80, It would be interesting to test whether these in vitro results on the role of oncogene signaling have relevance to human tumors, considering that data regarding expression of oncogenes and Hsp70 in a variety of human cancers became recently available.

Hsp70 regulates multiple signaling pathways

Original findings that Hsp70 suppresses apoptosis and oncogene-induced senescence suggested that it can interfere with signaling pathways that govern these processes. Indeed, a large number of publications from many labs indicated that Hsp70 affects cellular signaling pathways. For example, Hsp70 suppresses activation of multiple kinase cascades, including JNK, p38 and ERK 37, 38, 81. Via regulation of these cascades Hsp70 influences many physiological and pathological processes, like inflammatory processes 2729 or development of type II diabetes 82. Several cancer-related signaling pathways regulated by Hsp70 have been reported beside above-mentioned JNK and ERK. For example, Hsp70 was found to control expression of the major regulator of OIS, the cell cycle inhibitor p21 56, which is critical for establishing senescence upon Hsp70 depletion 56 (Fig. 2). In part, increase in p21 levels resulted from stabilization of p53, which in turn was mediated by effects of Hsp70 on the E3 ligase Mdm2 56. On the other hand, Hsp70 could also exert its effects on p21 in a p53-independent manner, which involves the transcription factor FoxM1 83 (Fig. 2). Overall, these pathways modulate effects of Hsp70 on OIS, and thus control the process of cancer initiation.

Fig. 2.

Fig. 2

Hsp70 regulates multiple pathways in cancer cells via interaction with BAG3 cochaperone.

In relation to cancer progression and metastasis, Hsp70 was shown to regulate a number of pathways, including transcription factors Hif1α, NF-kB 83(Fig. 2). Effects of Hsp70 on FoxM1 also contribute to cancer progression and metastasis 83. In addition, depletion of Hsp70 causes downregulation of a major translation modulator HuR, which controls expression of multiple proteins involved in tumor growth, invasion and metastasis 83. These findings most likely represent just a tip of the iceberg, since Gene Set Enrichment analysis of microarray data suggest effects of Hsp70 on a large number of signaling pathways, the majority of which yet to be validated (Colvin et al, unpublished data). Therefore, Hsp70 appears to control a complex network of pathways that regulate multiple steps in cancer development both at transcriptional and translational level.

What are the mechanism(s) by which Hsp70 affects cell signaling? One mode of Hsp70 action appears to involve its interaction with Hsp90 and other co-chaperones. This complex is involved in stabilization of a multitude of client proteins, many of which are components of major signaling pathways, e.g. c-Raf or Akt 84. It was demonstrated that upon inhibition of Hsp90, it dissociates from the complexes, and Hsp70 (or cognate Hsc73) links the clients with E3 ligases, like CHIP, thus promoting their ubiquitination and. It was reported that depletion of Hsp70 together with the cognate Hsc73 mimics effects of inhibition of Hsp90 on degradation of multiple client proteins 85. The mechanistic basis for these effects is unclear since according to common paradigm Hsp70/Hsc73 is necessary for degradation of clients, and therefore its depletion should stabilize them, while the report demonstrated the opposite. In line with this observation, a recently developed inhibitor YK5, which targets a novel allosteric site on Hsp70 and Hsc73, had similar effects, indicating that involvement in degradation of Hsp90 clients significantly contributes to effects of Hsp70 on cancer 86. On the other hand, unlike double Hsp70/Hsc73 depletion, depletion of Hsp70 alone does not trigger downregulation of Hsp90 clients (our unpublished data), and therefore knockdown of Hsp70 alone on signaling pathways must be based on a distinct mechanism.

Mechanistically important, deletion of C-terminal EEVD sequence, which completely inhibited the chaperone activity of Hsp70, did not affect its function in suppression of JNK pathway and protection from TNF-induced apoptosis 42. Similarly, with regulation of Raf-1 kinase, the chaperone activity of Hsp70 was dispensable 87. These data indicated that Hsp70 affects various signaling pathways in a mode that does not involve the chaperone function either in folding or degradation.

Bag3 mediates effects of Hsp70 on signaling

In search for a potential mediator of effects of Hsp70 on signaling, we turned to a co-chaperone Bag3, a unique member of the Bag-protein family. Beside the Bag region that can interact with a motif in the ATPase domain of Hsp70 proteins, Bag3 also contains functional PxxP and WW domains 88 (Fig. 3), which may connect it to SH3 domains and PPxY motifs of signaling proteins 89,90. Notably, Bag3 has been implicated in tumor development 91. Briefly, expression of Bag3 is upregulated in a number of cancers such as glioblastoma 92, melanoma 93, 94, pancreatic adenocarcinoma 95, 96, and hepatocellular carcinoma (HCC) 97. Furthermore, overexpression of Bag3 correlates with dismal prognosis in melanoma 93, pancreatic carcinoma and HCC 97. Moreover, its serum level was proposed to serve as a biomarker of pancreatic adenocarcinoma 98.

Fig. 3.

Fig. 3

Domain structure of Bag3 cochaperone and its involvement in cell signaling.

Like Hsp70, Bag3 inhibits apoptosis by interfering with cytochrome c release, apoptosome assembly and other events in the cellular death program 99. Moreover, it takes part in the processes of cancer cell adhesion and migration 91, 100. In human cancer cells, including lymphocytic and myeloblastic leukemic cells, Bag3 promotes cell survival and enhances resistance to chemotherapy 101, 102. Importantly, expression of Bag3 is co-regulated with Hsp70 in Hsf1-dependent manner 103. Based on these functional data together with structural analysis of Bag3, we proposed that this factor may serve as a scaffold/mediator that transmits effects of Hsp70 to multiple signaling pathways that control cancer development.

So far, the main focus in studies of Bag3 has been on its role in autophagy and aggresome formation, two major protective responses to protein aggregation. Two models have been proposed to explain the requirement of Bag3 and Hsp70 for these processes. According to one model, Bag3 is physically involved in recruitment of aggregated proteins to microtubules for further transport to aggresome or formation of autophagic vacuoles 104, 105. There is a significant circumstantial evidence to support this model; however, it has not been proven directly. According to another model, Bag3 transmits signals that regulate autophagy by mechanical stress, in part by binding to the LATS1 kinase via the WW domain to affect Hippo signaling pathway 106. These finding also implicate Bag3 in cancer signaling since Hippo pathway plays a major role in cancer cell stemness and epithelial-mesenchymal transition 107, 108. Another cancer-related signaling factor regulated by Bag3 is PLCγ 109. Bag3 interacts with the SH3 domain of this protein via its PXXP region 109 (Fig. 3). While implicating Bag3 in cancer signaling, these data also suggested that Bag3 can serve as a direct mediator of effects of Hsp70 on cancer, and our recent work addressed this possibility.

Using several cell lines, we showed that depletion of Bag3 has similar effects on multiple signaling pathways as depletion of Hsp70. Indeed, knockdown of of either Bag3 or Hsp70 downregulated FoxM1 and survivin and upregulated p21 and p27, and both depletions suppressed Hif1 and HuR 83 (Fig. 2) Importantly, overexpression of Bag3 mutants with substitutions in Bag domain that cannot interact with Hsp70 mimicked effects of depletions, indicating that Hsp70-Bag3 interaction is critical for regulation of these pathways 83. Interestingly, effects of Bag3 and Hsp70 knockdowns on NF-kB differed significantly. Unlike Hsp70 depletion, Bag3 depletion did not activate NF-kB, but at the same time it prevented activation of NF-kB by depletion of Hsp70 83. In other words, Bag3 was critical for regulation of NF-kB by Hsp70. Taken together, it appears that a network of cancer signaling pathways is regulated by Bag3-Hsp70 module and the effects may vary mechanistically with different pathways.

Direct biochemical test whether Bag3 mediates effects of Hsp70 on signaling pathways came from investigation of Src. As with NF-kB, Hsp70 regulates Src activity in a Bag3-dependent manner 83. Importantly, PXXP region of Bag3 directly interacts with the SH3 domain of Src, and this interaction requires association of Hsp70 with Bag domain 83 (Fig. 3).

In summary, Hsp70 is involved in regulating a complex signaling network that involves major cancer-related factors, including FoxM1, p21, p27, survivin, HuR, Hif1, NF-kB, Src, as well as other pathways (s.2). When individually considered, these pathways could have cancer-promoting or cancer-repressing effects. Nevertheless, we know that at least in breast cancer models, knockout of Hsp70 suppresses both tumor initiation and progression, indicating that the net outcome of this complex Hsp70 signaling network to be pro-oncogenic.

The network of pathways regulated by Hsp70 may be even more complex since distinct Bag-family members may also mediate effects of Hsp70 on signaling. For example, originally it was reported that Bag1 mediates effects of Hsp70 on Raf1 87. Bag1 has ubiquitin-like domain that is involved in its interaction with various components of the ubiquitin-proteasome system, which increases putative complexity of the system 110. In another example, Bag6 may also be involved in signaling since it has a proline-rich domain 111 which may also be involved in interaction with SH3-domain proteins.

Targeting Hsp70 for cancer treatment

Overall, concerted efforts of multiple labs uncovered that Hsp70 satisfies major requirements of an anti-cancer drug target. Indeed, (1) Hsp70 levels are strongly elevated in a large fraction of various cancers compared to normal corresponding tissues, (2) Hsp70 is essential for growth and survival of cancer but not normal cells, (3) Hsp70 knockout mice are viable, and therefore Hsp70 is dispensable for normal tissues and even organism development, (4) Hsp70 is “drugable”. Accordingly, significant efforts have been undertaken to develop Hsp70 inhibitors for cancer treatment (see recent review by Jason Gestwicki and colleagues 112, 113). These studies indicated that (a) Hsp70-Bag3 module can be targeted by small molecules and used for cancer treatment, and (b) Hsp70-Bag3 interaction is not the only mechanism that defines effects of Hsp70 on cancer, and its other activities and interactions with distinct factors should also be considered for cancer targeting.

First molecule that showed effect on interaction between Hsp70 and Bag-family proteins was PES, which was originally developed by the group of Andrei Gudkov as inhibitor of mitochondrial p53114. PES can bind to Hsp70/Hsc70 and, and showed potent anti-cancer effects in cell culture and a transgenic model of myc-induced lymphoma development 115 (Table 1). Recently the next generation of PES-related compound (PES-Cl) with higher affinity was developed and demonstrated increased potency in this mouse model 116. Interestingly, though PES interacts with the substrate-binding domain of Hsp70 116, it effectively inhibits interaction of Hsp70 with various co-factors that interact with the ATPase domain, including J-domain and Bag-domain proteins 115. Therefore, it is possible that interference with the Hsp70-Bag3 interaction may be a significant contributor to the anti-cancer effects of these compounds.

Table 1.

Inhibitors of Hsp70

Name Target Pathway inhibited Effect in vitro Effect in vivo Refs
Apaptomer Hsp70 Subsrate-binding, ATPase Anti-apoptotic Sensitzation to drug-induced apoptosis Inhibition of B16 melanoma growth 41
MAL3-101 DNAJ Anti-apoptotic Apoptosis, growth inhibition Inhibition of nultiple myeloma and Mercell cell carcinoma growth 132, 133
Myricetin DNA J/Hsp(c)70 Proteasome, topo-I, thioredoxin reductase etc Apoptosis Inhibition of pancreatic cancer growth 125, 127, 128, 138
PES (pithifrin u) Hsp(c)70 Autophagy Growth inhibition Delay in myc-induced lymphoma development 115
VER-155008 Hsp(c)70 ATPase Hsp90 clients Apoptosis, growth inhibition ND 124
YM-1 Bag3-Hsp(c)70 FoxM1, HuR Inhibition of growth Inhibition of growth of breast carcinoma and B16 melanoma 117
YK5 Hsp(c)70-Hsp90 Hsp90 clients Inhibition of growth, apoptosis ND 86

Another series of compounds that affect the Hsp70-Bag3 module is YM-1 and its analogs (Table 1). This series of compounds was developed by the Gestwicki group in their screen for molecules that affect the ATPase activity of Hsp70 117. YM-1 prevents the nucleotide exchange, and thus freezes Hsp70 in the ADP-bound state 118, 119. Interestingly, in such a state, Hsp70 tightly associates with its polypeptide substrates, which facilitates their ubiquitination by the Hsp70-associated E3 ligase CHIP, and further proteasome-dependent degradation 119. This property of YM-1 and related compounds could be used for treatment of neurodegenerative protein misfolding disorders, since administration of these compounds leads to rapid degradation of the disease proteins, e.g. tau 120 or androgen receptor with expanded polyglutamine domain 119. Via stabilization of Hsp70 in the ADP-bound form, YM-1 can also potently inhibit interaction of Hsp70 with Bag3 and other Bag-domain proteins both in a test tube and in cells 83. Because of this property of YM-1, we tested its effects on various signaling pathways controlled by Hsp70, including FoxM1, p21, p27, survivin Hif1, HuR, and Src 83. YM-1 mimicked effects of Hsp70 depletion on these pathways both in cell culture and in a cancer xenograft mouse model 83. Furthermore, it selectively killed cancer cells in culture, and potently inhibited cancer growth in mouse xenografts 83. Importantly, a close chemical analog of YM-1, which was significantly less potent in inhibition of the Hsp70-Bag3 interaction, did not affect these signaling pathways (manuscript in preparation), further suggesting that targeting the Hsp70-Bag3 module is responsible for the control of cancer-related signaling and anti-cancer activity of YM-1. Currently, the next generation of YM-1 analogs with higher potency and improved pharmacological properties is under development (Gestwicki, personal communication).

Though targeting Hsp70-Bag3 interaction appears to be most relevant to known cancer-promoting activities of Hsp70, its other functions may also contribute to cancer development and thus should be considered for targeting with small molecules. For example, there have been reports that Hsp70 is involved in chaperoning certain tumor suppressors, e.g. p53 121123 and therefore targeting the chaperone activity could be beneficial. Unfortunately, extremely strong binding to ATP makes it difficult to develop competitive inhibitors. Though one such inhibitor (VER-155008) has been developed by Vernalis 124 (Table 1), this company has abandoned the project because of the problems with further development and achieving reasonable specificity due to the similarity of the ATPase domain of Hsp70 to domains in actin and some other proteins. There were also attempts to target the chaperone activity by aptamers that bind substrate-binding or ATPase domains of Hsp70 41. Such aptamers sensitized cancer cell to drug-induced apoptotic cell death in culture and in mice 41(Table 1). On the other hand, because of the common delivery problems, the aptamers were expressed endogenously, and therefore their anti-cancer activity in a relevant cancer model has not been evaluated.

A distinct alternative approach has been to target interaction of Hsp70 with DnaJ co-chaperones. For example, a flavonoid myricetin was shown to bind to the ATPase domain of Hsp70 and inhibit DnaJ binding 125. Unlike Bag-domain proteins that serve as nucleotide-exchange factors, DnaJ proteins facilitate interaction of Hsp70 with certain substrates and stimulate ATP hydrolysis 126. These proteins bind to a distinct surface on the ATPase domain of Hsp70, and accordingly myricetin that interferes with the Hsp70-DnaJ interaction 125 did not affect interaction between Hsp70 and Bag3. In line with this finding, myricetin did not affect signaling pathways regulated by the Hsp70-Bag3 module 125. Nevertheless, myricetin demonstrated potent anti-cancer activities 117,127 (Table 1). However, as with other flavonoids, myricetin interacts with many proteins, e.g. thioredoxin reductase 128, DNA polymerase, topoisomerase I 129, and proteasome 130 which may define its anti-cancer effects (Table 1).

Another series of inhibitors of the Hsp70-J-domain interaction were developed using the screen for ATPase activity of Hsp70 stimulated by DnaJ 131. A compound MAL3-101 potently inhibited the DnaJ-stimulated ATPase, but not the intrinsic ATPase activity. Though its effect on the Hsp70-Bag3 interaction has not been tested, this series of compounds is unlikely to affect such interaction because J-domains bind to the opposite side of the ATPase domain of Hsp70. MAL3-101 demonstrated potent activity against multiple myeloma and Merkel cell carcinoma both in cell culture and animal models 132,133 (Table 1), strongly suggesting that DnaJ-related activities of Hsp70 unrelated to regulation of signaling via Hsp70-Bag3 module may be important for cancer development. Overall these experiments indicate that targeting the chaperone activity of Hsp70 either by interference with binding to substrates, or ATP, or DnaJ co-chaperones may be useful for cancer treatment. Recently a new allosteric site was identified in the ATPase domain of Hsp70, and a small molecule that binds to this domain has been developed 86 (Table 1). This molecule YK5 was shown to affect function of Hsp70 in Hsp90 complexes. Conventional inhibitors of Hsp90 destabilize client proteins, including factors involved in cancer, and facilitate their degradation, which defines anti-cancer effects of such inhibitors 134. Unfortunately, alone with depletion of cancer factors, they activate Hsf1 and enhance transcription of Hsp70 135. In turn, induction of Hsp70 suppresses apoptotic signaling and protects cells from Hsp90 inhibitors, which jeopardizes their anti-cancer activity 136. Interestingly, while mimicking effects of Hsp90 inhibitors on degradation of Hsp90 client signaling proteins, YK5 did not trigger induction of Hsp70, and therefore it may eventually demonstrate a much better potency than Hsp90 inhibitors 86. Notably, YM-1 showed only a very minor effect on major Hsp90 clients, further indicating that its anti-cancer effect is distinct and related to inhibition of the Hsp70-Bag3 module (manuscript in preparation).

A critical question in development of Hsp70 inhibitors as drugs is their specificity towards members of the Hsp70 family. The ideal drug, of course, would target Hsp70 only, but not the cognate Hsc73 or other family members. Due to the structural similarity of Hsp70 protein family this goal appears unrealistic, and so far existing drugs do not show significant specificity towards Hsp70 over other family members. A recent discovery, however, suggests an interesting approach towards development of specific drugs. Gestwicki’s group demonstrated that methylene blue is a mild oxidizer that can specifically oxidize a cystein in the ATPase domain of Hsp70, and inhibit its ATPase activity 137. This cystein residue is absent in Hsc73, thus providing a specificity of methylene blue against Hsp70. Though this molecule does not affect the Hsp70-Bag3 module (Gestwicki, personal communication), it affects the chaperone activity, and therefore may target oncogene folding or stability thus exerting anti-cancer effects.

Overall Hsp70 has matured as an attractive target for drug design, and targeting Hsp70-Bag3 interaction or the chaperone activity of Hsp70 could prove to be highly beneficial.

Contributor Information

Michael Y. Sherman, Email: sherma1@bu.edu.

Vladimir L. Gabai, Email: gabai@bu.edu.

References

  • 1.Bukau B, Weissman J, Horwich A. Molecular Chaperones and Protein Quality Control. Cell. 2006;125:443–451. doi: 10.1016/j.cell.2006.04.014. [DOI] [PubMed] [Google Scholar]
  • 2.Buchberger A, Bukau B, Sommer T. Protein Quality Control in the Cytosol and the Endoplasmic Reticulum: Brothers in Arms. Molecular Cell. 2010;40:238–252. doi: 10.1016/j.molcel.2010.10.001. [DOI] [PubMed] [Google Scholar]
  • 3.Morimoto RI. Regulation of the heat shock transcriptional response: cross talk between a family of heat shock factors, molecular chaperones, and negative regulators. Genes Dev. 1998;12:3788–3796. doi: 10.1101/gad.12.24.3788. [DOI] [PubMed] [Google Scholar]
  • 4.Voellmy R. Sensing stress and responding to stress. In: Feige U, Morimoto I, Yahara I, Polla B, editors. Stress-Inducible Cellular Responses. Vol. 77. Birkhauser Verlag; Basel: 1996. pp. 121–137. [Google Scholar]
  • 5.Jinwal UK, Akoury E, Abisambra JF, O’Leary JC, 3rd, Thompson AD, Blair LJ, et al. Imbalance of Hsp70 family variants fosters tau accumulation. FASEB J. 2013;27:1450–1459. doi: 10.1096/fj.12-220889. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Ciocca DR, Clark GM, Tandon AK, Fuqua SA, Welch WJ, McGuire WL. Heat shock protein hsp70 in patients with axillary lymph node-negative breast cancer: prognostic implications. Journal of the National Cancer Institute. 1993;85:570–574. doi: 10.1093/jnci/85.7.570. [DOI] [PubMed] [Google Scholar]
  • 7.Hwang TS, Han HS, Choi HK, Lee YJ, Kim YJ, Han MY, et al. Differential, stage-dependent expression of Hsp70, Hsp110 and Bcl-2 in colorectal cancer. J Gastroenterol Hepatol. 2003;18:690–700. doi: 10.1046/j.1440-1746.2003.03011.x. [DOI] [PubMed] [Google Scholar]
  • 8.Joo M, Chi JG, Lee H. Expressions of HSP70 and HSP27 in hepatocellular carcinoma. J Korean Med Sci. 2005;20:829–834. doi: 10.3346/jkms.2005.20.5.829. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Luk JM, Lam CT, Siu AF, Lam BY, Ng IO, Hu MY, et al. Proteomic profiling of hepatocellular carcinoma in Chinese cohort reveals heat-shock proteins (Hsp27, Hsp70, GRP78) up-regulation and their associated prognostic values. Proteomics. 2006;6:1049–1057. doi: 10.1002/pmic.200500306. [DOI] [PubMed] [Google Scholar]
  • 10.Alaiya AA, Oppermann M, Langridge J, Roblick U, Egevad L, Brindstedt S, et al. Identification of proteins in human prostate tumor material by two-dimensional gel electrophoresis and mass spectrometry. Cell Mol Life Sci. 2001;58:307–311. doi: 10.1007/PL00000858. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Tang D, Khaleque MA, Jones EL, Theriault JR, Li C, Wong WH, et al. Expression of heat shock proteins and heat shock protein messenger ribonucleic acid in human prostate carcinoma in vitro and in tumors in vivo. Cell Stress Chaperones. 2005;10:46–58. doi: 10.1379/CSC-44R.1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Wang XP, Wang QX, Li HY, Chen RF. Heat shock protein 70 chaperoned alpha-fetoprotein in human hepatocellular carcinoma cell line BEL-7402. World J Gastroenterol. 2005;11:5561–5564. doi: 10.3748/wjg.v11.i35.5561. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Hellman K, Alaiya AA, Schedvins K, Steinberg W, Hellstrom AC, Auer G. Protein expression patterns in primary carcinoma of the vagina. Br J Cancer. 2004;91:319–326. doi: 10.1038/sj.bjc.6601944. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Park CS, Joo IS, Song SY, Kim DS, Bae DS, Lee JH. An immunohistochemical analysis of heat shock protein 70, p53, and estrogen receptor status in carcinoma of the uterine cervix. Gynecol Oncol. 1999;74:53–60. doi: 10.1006/gyno.1999.5429. [DOI] [PubMed] [Google Scholar]
  • 15.Ciocca DR, Calderwood SK. Heat shock proteins in cancer: diagnostic, prognostic, predictive, and treatment implications. Cell Stress Chaperones. 2005;10:86–103. doi: 10.1379/CSC-99r.1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Bauer K, Nitsche U, Slotta-Huspenina J, Drecoll E, von Weyhern CH, Rosenberg R, et al. High HSP27 and HSP70 expression levels are independent adverse prognostic factors in primary resected colon cancer. Cell Oncol (Dordr) 2012;35:197–205. doi: 10.1007/s13402-012-0079-3. [DOI] [PubMed] [Google Scholar]
  • 17.Yu H-J, Chang Y-H, Pan C-C. Prognostic significance of heat shock proteins in urothelial carcinoma of the urinary bladder. Histopathology. 2013;62:788–798. doi: 10.1111/his.12087. [DOI] [PubMed] [Google Scholar]
  • 18.Syrigos KN, Harrington KJ, Karayiannakis AJ, Sekara E, Chatziyianni E, Syrigou EI, et al. Clinical significance of heat shock protein-70 expression in bladder cancer. Urology. 2003;61:677–680. doi: 10.1016/s0090-4295(02)02289-6. [DOI] [PubMed] [Google Scholar]
  • 19.Lee HW, Lee EH, Kim SH, Roh MS, Jung SB, Choi YC. Heat shock protein 70 (HSP70) expression is associated with poor prognosis in intestinal type gastric cancer. Virchows Arch. 2013;463:489–495. doi: 10.1007/s00428-013-1461-x. [DOI] [PubMed] [Google Scholar]
  • 20.Shin E, Ryu HS, Kim SH, Jung H, Jang JJ, Lee K. The clinicopathological significance of heat shock protein 70 and glutamine synthetase expression in hepatocellular carcinoma. J Hepatobiliary Pancreat Sci. 2011;18:544–550. doi: 10.1007/s00534-010-0367-0. [DOI] [PubMed] [Google Scholar]
  • 21.Ramp U, Mahotka C, Heikaus S, Shibata T, Grimm MO, Willers R, et al. Expression of heat shock protein 70 in renal cell carcinoma and its relation to tumor progression and prognosis. Histol Histopathol. 2007;22:1099–1107. doi: 10.14670/HH-22.1099. [DOI] [PubMed] [Google Scholar]
  • 22.Nakajima M, Kuwano H, Miyazaki T, Masuda N, Kato H. Significant correlation between expression of heat shock proteins 27, 70 and lymphocyte infiltration in esophageal squamous cell carcinoma. Cancer Lett. 2002;178:99–106. doi: 10.1016/s0304-3835(01)00825-4. [DOI] [PubMed] [Google Scholar]
  • 23.Boonjaraspinyo S, Boonmars T, Kaewkes S, Laummaunwai P, Pinlaor S, Loilome W, et al. Down-regulated expression of HSP70 in correlation with clinicopathology of cholangiocarcinoma. Pathol Oncol Res. 2012;18:227–237. doi: 10.1007/s12253-011-9432-5. [DOI] [PubMed] [Google Scholar]
  • 24.Tavassol F, Starke OF, Kokemuller H, Wegener G, Muller-Tavassol CC, Gellrich NC, et al. Prognostic significance of heat shock protein 70 (HSP70) in patients with oral cancer. Head Neck Oncol. 2011;3:10. doi: 10.1186/1758-3284-3-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Malusecka E, Krzyzowska-Gruca S, Gawrychowski J, Fiszer-Kierzkowska A, Kolosza Z, Krawczyk Z. Stress proteins HSP27 and HSP70i predict survival in non-small cell lung carcinoma. Anticancer Res. 2008;28:501–506. [PubMed] [Google Scholar]
  • 26.Ricaniadis N, Kataki A, Agnantis N, Androulakis G, Karakousis CP. Long-term prognostic significance of HSP-70, c-myc and HLA-DR expression in patients with malignant melanoma. Eur J Surg Oncol. 2001;27:88–93. doi: 10.1053/ejso.1999.1018. [DOI] [PubMed] [Google Scholar]
  • 27.Tang D, Kang R, Xiao W, Wang H, Calderwood SK, Xiao X. The anti-inflammatory effects of heat shock protein 72 involve inhibition of high-mobility-group box 1 release and proinflammatory function in macrophages. J Immunol. 2007;179:1236–1244. doi: 10.4049/jimmunol.179.2.1236. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Zheng Z, Kim JY, Ma H, Lee JE, Yenari MA. Anti-inflammatory effects of the 70 kDa heat shock protein in experimental stroke. J Cereb Blood Flow Metab. 2008;28:53–63. doi: 10.1038/sj.jcbfm.9600502. [DOI] [PubMed] [Google Scholar]
  • 29.Tanaka K, Namba T, Arai Y, Fujimoto M, Adachi H, Sobue G, et al. Genetic evidence for a protective role for heat shock factor 1 and heat shock protein 70 against colitis. J Biol Chem. 2007;282:23240–23252. doi: 10.1074/jbc.M704081200. [DOI] [PubMed] [Google Scholar]
  • 30.Chen H, Wu Y, Zhang Y, Jin L, Luo L, Xue B, et al. Hsp70 inhibits lipopolysaccharide-induced NF-kappaB activation by interacting with TRAF6 and inhibiting its ubiquitination. FEBS Lett. 2006;580:3145–3152. doi: 10.1016/j.febslet.2006.04.066. [DOI] [PubMed] [Google Scholar]
  • 31.Tao Y, Hart J, Lichtenstein L, Joseph LJ, Ciancio MJ, Hu S, et al. Inducible heat shock protein 70 prevents multifocal flat dysplastic lesions and invasive tumors in an inflammatory model of colon cancer. Carcinogenesis. 2009;30:175–182. doi: 10.1093/carcin/bgn256. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Calderwood SK, Theriault JR, Gong J. Message in a bottle: role of the 70-kDa heat shock protein family in anti-tumor immunity. Eur J Immunol. 2005;35:2518–2527. doi: 10.1002/eji.200535002. [DOI] [PubMed] [Google Scholar]
  • 33.Bolhassani A, Rafati S. Heat-shock proteins as powerful weapons in vaccine development. Expert Review of Vaccines. 2008;7:1185–1199. doi: 10.1586/14760584.7.8.1185. [DOI] [PubMed] [Google Scholar]
  • 34.Calderwood SK, Ciocca DR. Heat shock proteins: Stress proteins with Janus-like properties in cancer. International Journal of Hyperthermia. 2008;24:31–39. doi: 10.1080/02656730701858305. [DOI] [PubMed] [Google Scholar]
  • 35.Cai MB, Wang XP, Zhang JX, Han HQ, Liu CC, Bei JX, et al. Expression of heat shock protein 70 in nasopharyngeal carcinomas: different expression patterns correlate with distinct clinical prognosis. Journal of Translational Medicine. 2012:10. doi: 10.1186/1479-5876-10-96. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Gabai VL, Zamulaeva IV, Mosin AF, Makarova YM, Mosina VA, Budagova KR, et al. Resistance of Ehrlich tumor cells to apoptosis can be due to accumulation of heat shock proteins. FEBS Letters. 1995;375:21–26. doi: 10.1016/0014-5793(95)01152-5. [DOI] [PubMed] [Google Scholar]
  • 37.Gabai VL, Meriin AB, Mosser DD, Caron AW, Rits S, Shifrin VI, Sherman MY. HSP70 prevent activation of stress kinases: a novel pathway of cellular thermotolerance. Journal of Biological Chemistry. 1997;272:18033–18037. doi: 10.1074/jbc.272.29.18033. [DOI] [PubMed] [Google Scholar]
  • 38.Mosser DD, Caron AW, Bourget L, Denis-Larose C, Massie B. Role of the human heat shock protein hsp70 in protection against stress-induced apoptosis. Molecular & Cellular Biology. 1997;17:5317–5327. doi: 10.1128/mcb.17.9.5317. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Hanahan D, Weinberg Robert A. Hallmarks of Cancer: The Next Generation. Cell. 2011;144:646–674. doi: 10.1016/j.cell.2011.02.013. [DOI] [PubMed] [Google Scholar]
  • 40.Schmitt E, Maingret L, Puig P-E, Rerole A-L, Ghiringhelli F, Hammann A, et al. Heat Shock Protein 70 Neutralization Exerts Potent Antitumor Effects in Animal Models of Colon Cancer and Melanoma. [10.1158/0008-5472];CAN-05-3778 Cancer Res. 2006 66:4191–4197. doi: 10.1158/0008-5472.CAN-05-3778. [DOI] [PubMed] [Google Scholar]
  • 41.Rérole A-L, Gobbo J, De Thonel A, Schmitt E, Pais de Barros JP, Hammann A, et al. Peptides and Aptamers Targeting HSP70: A Novel Approach for Anticancer Chemotherapy. Cancer Research. 2011;71:484–495. doi: 10.1158/0008-5472.CAN-10-1443. [DOI] [PubMed] [Google Scholar]
  • 42.Gabai VL, Mabuchi K, Mosser DD, Sherman MY. Hsp72 and Stress Kinase c-jun N-Terminal Kinase Regulate the Bid-Dependent Pathway in Tumor Necrosis Factor-Induced Apoptosis. Mol Cell Biol. 2002;22:3415–3424. doi: 10.1128/MCB.22.10.3415-3424.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Stankiewicz AR, Lachapelle G, Foo CPZ, Radicioni SM, Mosser DD. Hsp70 Inhibits Heat-induced Apoptosis Upstream of Mitochondria by Preventing Bax Translocation. J Biol Chem. 2005;280:38729–38739. doi: 10.1074/jbc.M509497200. [DOI] [PubMed] [Google Scholar]
  • 44.Beere HM, Wolf BB, Cain K, Mosser DD, Mahboubi A, Kuwana T, et al. Heat-shock protein 70 inhibits apoptosis by preventing recruitment of procaspase-9 to the Apaf-1 apoptosome. Nat Cell Biol. 2000;2:469–475. doi: 10.1038/35019501. [DOI] [PubMed] [Google Scholar]
  • 45.Afanasyeva EA, Komarova EY, Larsson L-G, Bahram F, Margulis BA, Guzhova IV. Drug-induced Myc-mediated apoptosis of cancer cells is inhibited by stress protein Hsp70. International Journal of Cancer. 2007;121:2615–2621. doi: 10.1002/ijc.22974. [DOI] [PubMed] [Google Scholar]
  • 46.Nylandsted J, Rohde M, Brand K, Bastholm L, Elling F, Jaattela M. Selective depletion of heat shock protein 70 (Hsp70) activates a tumor- specific death program that is independent of caspases and bypasses Bcl- 2. Proc Natl Acad Sci U S A. 2000;97:7871–7876. doi: 10.1073/pnas.97.14.7871. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Nylandsted J, Gyrd-Hansen M, Danielewicz A, Fehrenbacher N, Lademann U, Hoyer-Hansen M, et al. Heat Shock Protein 70 Promotes Cell Survival by Inhibiting Lysosomal Membrane Permeabilization. J Exp Med. 2004;200:425–435. doi: 10.1084/jem.20040531. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Nylandsted J, Wick W, Hirt UA, Brand K, Rohde M, Leist M, et al. Eradication of Glioblastoma, and Breast and Colon Carcinoma Xenografts by Hsp70 Depletion. Cancer Res. 2002;62:7139–7142. [PubMed] [Google Scholar]
  • 49.Murphy ME. The HSP70 family and cancer. Carcinogenesis. 2013;34:1181–1188. doi: 10.1093/carcin/bgt111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Powers MV, Jones K, Barillari C, Westwood I, Montfort RLMv, Workman P. Targeting HSP70: The second potentially druggable heat shock protein and molecular chaperone? Cell Cycle. 2010;9:1542–1550. doi: 10.4161/cc.9.8.11204. [DOI] [PubMed] [Google Scholar]
  • 51.Collado M, Serrano M. The power and the promise of oncogene-induced senescence markers. 2006;6:472–476. doi: 10.1038/nrc1884. [DOI] [PubMed] [Google Scholar]
  • 52.Coppe JP, Patil CK, Rodier F, Sun Y, Munoz DP, Goldstein J, et al. Senescence-associated secretory phenotypes reveal cell-nonautonomous functions of oncogenic RAS and the p53 tumor suppressor. PLoS Biol. 2008;6:2853–2868. doi: 10.1371/journal.pbio.0060301. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Raulet DH, Guerra N. Oncogenic stress sensed by the immune system: role of natural killer cell receptors. Nat Rev Immunol. 2009;9:568–580. doi: 10.1038/nri2604. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Braig M, Schmitt CA. Oncogene-Induced Senescence: Putting the Brakes on Tumor Development. [10.1158/0008-5472];CAN-05-4006 Cancer Res. 2006 66:2881–2884. doi: 10.1158/0008-5472.CAN-05-4006. [DOI] [PubMed] [Google Scholar]
  • 55.Gabai VL, Yaglom JA, Waldman T, Sherman MY. Heat Shock Protein Hsp72 Controls Oncogene-Induced Senescence Pathways in Cancer Cells. [10.1128/MCB. 01041-08];Mol Cell Biol. 2009 29:559–569. doi: 10.1128/MCB.01041-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Yaglom JA, Gabai VL, Sherman MY. High Levels of Heat Shock Protein Hsp72 in Cancer Cells Suppress Default Senescence Pathways. [10.1158/0008-5472];CAN-06-3796 Cancer Res. 2007 67:2373–2381. doi: 10.1158/0008-5472.CAN-06-3796. [DOI] [PubMed] [Google Scholar]
  • 57.Meng L, Hunt C, Yaglom JA, Gabai VL, Sherman MY. Heat shock protein Hsp72 plays an essential role in Her2-induced mammary tumorigenesis. Oncogene. 2011;30:2836–2845. doi: 10.1038/onc.2011.5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Dodd K, Nance S, Quezada M, Janke L, Morrison JB, Williams RT, et al. Tumor-derived inducible heat-shock protein 70 (HSP70) is an essential component of anti-tumor immunity. Oncogene. 2014 doi: 10.1038/onc.2014.63. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Goodwin EC, DiMaio D. Repression of human papillomavirus oncogenes in HeLa cervical carcinoma cells causes the orderly reactivation of dormant tumor suppressor pathways. Proceedings of the National Academy of Sciences. 2000;97:12513–12518. doi: 10.1073/pnas.97.23.12513. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Weng D, Penzner J, Song B, Koido S, Calderwood S, Gong J. Metastasis is an early event in mouse mammary carcinomas and is associated with cells bearing stem cell markers. Breast Cancer Research. 2012;14:R18. doi: 10.1186/bcr3102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Calderwood SK, Khaleque MA, Sawyer DB, Ciocca DR. Heat shock proteins in cancer: chaperones of tumorigenesis. Trends Biochem Sci. 2006;31:164–172. doi: 10.1016/j.tibs.2006.01.006. [DOI] [PubMed] [Google Scholar]
  • 62.Partida-Rodríguez O, Torres J, Flores-Luna L, Camorlinga M, Nieves-Ramírez M, Lazcano E, et al. Polymorphisms in TNF and HSP-70 show a significant association with gastric cancer and duodenal ulcer. International Journal of Cancer. 2010;126:1861–1868. doi: 10.1002/ijc.24773. [DOI] [PubMed] [Google Scholar]
  • 63.Szondy K, Rusai K, Szabó AJ, Nagy A, Gal K, Fekete A, et al. Tumor Cell Expression of Heat Shock Protein (HSP) 72 is Influenced by HSP72 [HSPA1B A(1267)G] Polymorphism and Predicts Survival in Small Cell Lung Cancer (SCLC) Patients. Cancer Investigation. 2012;30:317–322. doi: 10.3109/07357907.2012.657815. [DOI] [PubMed] [Google Scholar]
  • 64.Wang Y, Zhou F, Wu Y, Xu D, Li W, Liang S. The relationship between three heat shock protein 70 gene polymorphisms and susceptibility to lung cancer. Clinical Chemistry and Laboratory Medicine. 2010;48:1657. doi: 10.1515/CCLM.2010.304. [DOI] [PubMed] [Google Scholar]
  • 65.Ciocca D, Arrigo A, Calderwood S. Heat shock proteins and heat shock factor 1 in carcinogenesis and tumor development: an update. Archives of Toxicology. 2013;87:19–48. doi: 10.1007/s00204-012-0918-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Kabakov AE, Gabai VL. Heat Shock proteins and Cytoprotection: ATP-deprived Mammalian Cells. R.G. Landes Co; Austin: 1997. [Google Scholar]
  • 67.Kabakov AE, Molotkov AO, Budagova KR, Makarova Yu M, Mosin AF, Gabai VL. Adaptation of Ehrlich ascites carcinoma cells to energy deprivation in vivo can be associated with heat shock protein accumulation. Journal of Cellular Physiology. 1995;165:1–6. doi: 10.1002/jcp.1041650102. [DOI] [PubMed] [Google Scholar]
  • 68.Luo J, Solimini NL, Elledge SJ. Principles of Cancer Therapy: Oncogene and Non-oncogene Addiction. Cell. 2009;136:823–837. doi: 10.1016/j.cell.2009.02.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Torres EM, Sokolsky T, Tucker CM, Chan LY, Boselli M, Dunham MJ, et al. Effects of Aneuploidy on Cellular Physiology and Cell Division in Haploid Yeast. Science. 2007;317:916–924. doi: 10.1126/science.1142210. [DOI] [PubMed] [Google Scholar]
  • 70.Pavelka N, Rancati G, Zhu J, Bradford WD, Saraf A, Florens L, et al. Aneuploidy confers quantitative proteome changes and phenotypic variation in budding yeast. Nature. 2010;468:321–325. doi: 10.1038/nature09529. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Tang Y-C, Williams BR, Siegel JJ, Amon A. Identification of Aneuploidy-Selective Antiproliferation Compounds. Cell. 2011;144:499–512. doi: 10.1016/j.cell.2011.01.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Colvin TA, Gabai VL, Sherman MY. Proteotoxicity is not the reason for the dependence of cancer cells on the major chaperone Hsp70. Cell Cycle. 2014 doi: 10.4161/cc.29296. in press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.O’Callaghan-Sunol C, Gabai VL, Sherman MY. Hsp27 Modulates p53 Signaling and Suppresses Cellular Senescence. [10.1158/0008-5472.CAN-07-2441];Cancer Res. 2007 67:11779–11788. doi: 10.1158/0008-5472.CAN-07-2441. [DOI] [PubMed] [Google Scholar]
  • 74.Menssen A, Hermeking H. Characterization of the c-MYC-regulated transcriptome by SAGE: Identification and analysis of c-MYC target genes. Proceedings of the National Academy of Sciences. 2002;99:6274–6279. doi: 10.1073/pnas.082005599. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Myers SM, Mulligan LM. The RET Receptor Is Linked to Stress Response Pathways. Cancer Research. 2004;64:4453–4463. doi: 10.1158/0008-5472.CAN-03-3605. [DOI] [PubMed] [Google Scholar]
  • 76.Ray S, Lu Y, Kaufmann SH, Gustafson WC, Karp JE, Boldogh I, et al. Genomic Mechanisms of p210BCR-ABL Signaling: INDUCTION OF HEAT SHOCK PROTEIN 70 THROUGH THE GATA RESPONSE ELEMENT CONFERS RESISTANCE TO PACLITAXEL-INDUCED APOPTOSIS. Journal of Biological Chemistry. 2004;279:35604–35615. doi: 10.1074/jbc.M401851200. [DOI] [PubMed] [Google Scholar]
  • 77.Khaleque M, Bharti A, Sawyer D, Gong J, Benjamin IJ, Stevenson MA, Calderwood SK. Induction of heat shock proteins by heregulin beta1 leads to protection from apoptosis and anchorage-independent growth. Oncogene. 2005;24:6564–6573. doi: 10.1038/sj.onc.1208798. [DOI] [PubMed] [Google Scholar]
  • 78.Chou SD, Prince T, Gong J, Calderwood SK. mTOR is essential for the proteotoxic stress response, HSF1 activation and heat shock protein synthesis. PLoS ONE. 2012;7:e39679. doi: 10.1371/journal.pone.0039679. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.McCubrey JA, Steeman L, Chappell W, Abrams S, Montalto G, Cervello M, et al. Mutations and Deregulation of Ras/Raf/MEK/ERK and PI3K/PTEN/Akt/mTOR Cascades. 2012;3 [Google Scholar]
  • 80.Li D, Yallowitz A, Ozog L, Marchenko N. A gain-of-function mutant p53-HSF1 feed forward circuit governs adaptation of cancer cells to proteotoxic stress. Cell Death Dis. 2014;5:e1194. doi: 10.1038/cddis.2014.158. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Yaglom J, O’Callaghan-Sunol C, Gabai V, Sherman MY. Inactivation of Dual-Specificity Phosphatases Is Involved in the Regulation of Extracellular Signal-Regulated Kinases by Heat Shock and Hsp72. Mol Cell Biol. 2003;23:3813–3824. doi: 10.1128/MCB.23.11.3813-3824.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Chung J, Nguyen A-K, Henstridge DC, Holmes AG, Chan MHS, Mesa JL, et al. HSP72 protects against obesity-induced insulin resistance. Proceedings of the National Academy of Sciences. 2008;105:1739–1744. doi: 10.1073/pnas.0705799105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Colvin TA, Gabai VL, Gong J, Calderwood SK, Li H, Gummuluru S, et al. Hsp70-Bag3 module regulates cancer-related signaling networks. Cancer Research. 2014 doi: 10.1158/0008-5472.CAN-14-0747. in press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Wandinger SK, Richter K, Buchner J. The Hsp90 Chaperone Machinery. Journal of Biological Chemistry. 2008;283:18473–18477. doi: 10.1074/jbc.R800007200. [DOI] [PubMed] [Google Scholar]
  • 85.Powers MV, Clarke PA, Workman P. Dual Targeting of HSC70 and HSP72 Inhibits HSP90 Function and Induces Tumor-Specific Apoptosis. Cancer Cell. 2008;14:250–262. doi: 10.1016/j.ccr.2008.08.002. [DOI] [PubMed] [Google Scholar]
  • 86.Rodina A, Patel Pallav D, Kang Y, Patel Y, Baaklini I, Wong Michael JH, et al. Identification of an Allosteric Pocket on Human Hsp70 Reveals a Mode of Inhibition of This Therapeutically Important Protein. Chemistry & Biology. 2013;20:1469–1480. doi: 10.1016/j.chembiol.2013.10.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Song J, Takeda M, Morimoto RI. Bag1-Hsp70 mediates a physiological stress signalling pathway that regulates Raf-1/ERK and cell growth. Nat Cell Biol. 2001;3:276–282. doi: 10.1038/35060068. [DOI] [PubMed] [Google Scholar]
  • 88.Takayama S, Reed JC. Molecular chaperone targeting and regulation by BAG family proteins. Nat Cell Biol. 2001;3:E237–E241. doi: 10.1038/ncb1001-e237. [DOI] [PubMed] [Google Scholar]
  • 89.Ingham RJ, Colwill K, Howard C, Dettwiler S, Lim CSH, Yu J, et al. WW Domains Provide a Platform for the Assembly of Multiprotein Networks. Molecular and Cellular Biology. 2005;25:7092–7106. doi: 10.1128/MCB.25.16.7092-7106.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Li SS-C. Specificity and versatility of SH3 and other proline-recognition domains: structural basis and implications for cellular signal transduction. Biochem J. 2005;390:641–653. doi: 10.1042/BJ20050411. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Iwasaki M, Homma S, Hishiya A, Dolezal SJ, Reed JC, Takayama S. BAG3 Regulates Motility and Adhesion of Epithelial Cancer Cells. Cancer Research. 2007;67:10252–10259. doi: 10.1158/0008-5472.CAN-07-0618. [DOI] [PubMed] [Google Scholar]
  • 92.Festa M, Del Valle L, Khalili K, Franco R, Scognamiglio G, Graziano V, et al. BAG3 Protein Is Overexpressed in Human Glioblastoma and Is a Potential Target for Therapy. The American Journal of Pathology. 2011;178:2504–2512. doi: 10.1016/j.ajpath.2011.02.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Guerriero L, Chong K, Franco R, Rosati A, De Caro F, Capunzo M, et al. BAG3 protein expression in melanoma metastatic lymph nodes correlates with patients/’ survival. Cell Death Dis. 2014;5:e1173. doi: 10.1038/cddis.2014.143. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Franco R, Scognamiglio G, Salerno V, Sebastiani A, Cennamo G, Ascierto PA, et al. Expression of the Anti-Apoptotic Protein BAG3 in Human Melanomas. J Invest Dermatol. 2012;132:252–254. doi: 10.1038/jid.2011.257. [DOI] [PubMed] [Google Scholar]
  • 95.Rosati A, Bersani S, Tavano F, Dalla Pozza E, De Marco M, Palmieri M, et al. Expression of the Antiapoptotic Protein BAG3 Is a Feature of Pancreatic Adenocarcinoma and Its Overexpression Is Associated With Poorer Survival. The American Journal of Pathology. 2012;181:1524–1529. doi: 10.1016/j.ajpath.2012.07.016. [DOI] [PubMed] [Google Scholar]
  • 96.Liao Q, Ozawa F, Friess H, Zimmermann A, Takayama S, Reed JC, et al. The anti-apoptotic protein BAG-3 is overexpressed in pancreatic cancer and induced by heat stress in pancreatic cancer cell lines. FEBS Letters. 2001;503:151–157. doi: 10.1016/s0014-5793(01)02728-4. [DOI] [PubMed] [Google Scholar]
  • 97.Chuma M, Sakamoto N, Nakai A, Hige S, Nakanishi M, Natsuizaka M, et al. Heat shock factor 1 accelerates hepatocellular carcinoma development by activating nuclear factor-κB/mitogen-activated protein kinase. Carcinogenesis. 2014;35:272–281. doi: 10.1093/carcin/bgt343. [DOI] [PubMed] [Google Scholar]
  • 98.Falco A, Rosati A, Festa M, Basile A, De Marco M, d’Avenia M, et al. BAG3 Is a Novel Serum Biomarker for Pancreatic Adenocarcinomas. Am J Gastroenterol. 2013;108:1178–1180. doi: 10.1038/ajg.2013.128. [DOI] [PubMed] [Google Scholar]
  • 99.Rosati A, Ammirante M, Gentilella A, Basile A, Festa M, Pascale M, et al. Apoptosis inhibition in cancer cells: A novel molecular pathway that involves BAG3 protein. The International Journal of Biochemistry & Cell Biology. 2007;39:1337–1342. doi: 10.1016/j.biocel.2007.03.007. [DOI] [PubMed] [Google Scholar]
  • 100.Kassis JN, Guancial EA, Doong H, Virador V, Kohn EC. CAIR-1/BAG-3 modulates cell adhesion and migration by downregulating activity of focal adhesion proteins. Experimental Cell Research. 2006;312:2962–2971. doi: 10.1016/j.yexcr.2006.05.023. [DOI] [PubMed] [Google Scholar]
  • 101.Romano MF, Festa M, Petrella A, Pascale M, Bisogni R, Poggi V, et al. BAG3 Protein Regulates Cell Survival in Childhood Acute Lymphoblastic Leukemia Cells. Cancer Biology & Therapy. 2003;2:508–510. doi: 10.4161/cbt.2.5.524. [DOI] [PubMed] [Google Scholar]
  • 102.Romano MF, Festa M, Pagliuca G, Lerose R, Bisogni R, Chiurazzi F, et al. BAG3 protein controls B-chronic lymphocytic leukaemia cell apoptosis. Cell Death Differ. 2003;10:383–385. doi: 10.1038/sj.cdd.4401167. [DOI] [PubMed] [Google Scholar]
  • 103.Franceschelli S, Rosati A, Lerose R, De Nicola S, Turco MC, Pascale M. bag3 gene expression is regulated by heat shock factor 1. Journal of Cellular Physiology. 2008;215:575–577. doi: 10.1002/jcp.21397. [DOI] [PubMed] [Google Scholar]
  • 104.Gamerdinger M, Carra S, Behl C. Emerging roles of molecular chaperones and co-chaperones in selective autophagy: focus on BAG proteins. J Mol Med. 2011;89:1175–1182. doi: 10.1007/s00109-011-0795-6. [DOI] [PubMed] [Google Scholar]
  • 105.Gamerdinger M, Kaya AM, Wolfrum U, Clement AM, Behl C. BAG3 mediates chaperone-based aggresome-targeting and selective autophagy of misfolded proteins. 2011;12 doi: 10.1038/embor.2010.203. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Ulbricht A, Eppler Felix J, Tapia Victor E, van der Ven Peter FM, Hampe N, Hersch N, et al. Cellular Mechanotransduction Relies on Tension-Induced and Chaperone-Assisted Autophagy. Current Biology. 2013;23:430–435. doi: 10.1016/j.cub.2013.01.064. [DOI] [PubMed] [Google Scholar]
  • 107.Zhang J, Ji J-Y, Yu M, Overholtzer M, Smolen GA, Wang R, et al. YAP-dependent induction of amphiregulin identifies a non-cell-autonomous component of the Hippo pathway. Nat Cell Biol. 2009;11:1444–1450. doi: 10.1038/ncb1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Mo JS, Park HW, Guan KL. The Hippo signaling pathway in stem cell biology and cancer. 2014;15 doi: 10.15252/embr.201438638. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Doong H, Price J, Kim YS, Gasbarre C, Probst J, Liotta LA, et al. CAIR-1/BAG-3 forms an EGF-regulated ternary complex with phospholipase C-gamma and Hsp70/Hsc70. Oncogene. 2000;19:4385–4395. doi: 10.1038/sj.onc.1203797. [DOI] [PubMed] [Google Scholar]
  • 110.Tsukahara F, Maru Y. Bag1 directly routes immature BCR-ABL for proteasomal degradation. Blood. 2010;116:3582–3592. doi: 10.1182/blood-2009-10-249623. [DOI] [PubMed] [Google Scholar]
  • 111.Leznicki P, Roebuck QP, Wunderley L, Clancy A, Krysztofinska EM, Isaacson RL, et al. The Association of BAG6 with SGTA and Tail-Anchored Proteins. PLoS ONE. 2013;8:e59590. doi: 10.1371/journal.pone.0059590. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Assimon VA, Gillies AT, Rauch JN, Gestwicki JE. Hsp70 protein complexes as drug targets. Curr Pharm Des. 2013;19:404–417. doi: 10.2174/138161213804143699. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Patury S, Miyata Y, Gestwicki JE. Pharmacological targeting of the Hsp70 chaperone. Curr Top Med Chem. 2009;9:1337–1351. doi: 10.2174/156802609789895674. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Strom E, Sathe S, Komarov PG, Chernova OB, Pavlovska I, Shyshynova I, et al. Small-molecule inhibitor of p53 binding to mitochondria protects mice from gamma radiation. Nat Chem Biol. 2006;2:474–479. doi: 10.1038/nchembio809. [DOI] [PubMed] [Google Scholar]
  • 115.Leu JIJ, Pimkina J, Frank A, Murphy ME, George DL. A Small Molecule Inhibitor of Inducible Heat Shock Protein 70. Molecular Cell. 2009;36:15–27. doi: 10.1016/j.molcel.2009.09.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Balaburski GM, Leu JI-J, Beeharry N, Hayik S, Andrake MD, Zhang G, et al. A Modified HSP70 Inhibitor Shows Broad Activity as an Anticancer Agent. Molecular Cancer Research. 2013;11:219–229. doi: 10.1158/1541-7786.MCR-12-0547-T. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Koren J, III, Miyata Y, Kiray J, O’Leary JC, III, Nguyen L, Guo J, et al. Rhodacyanine Derivative Selectively Targets Cancer Cells and Overcomes Tamoxifen Resistance. PLoS ONE. 2012;7:e35566. doi: 10.1371/journal.pone.0035566. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Rousaki A, Miyata Y, Jinwal UK, Dickey CA, Gestwicki JE, Zuiderweg ERP. Allosteric Drugs: The Interaction of Antitumor Compound MKT-077 with Human Hsp70 Chaperones. Journal of Molecular Biology. 2011;411:614–632. doi: 10.1016/j.jmb.2011.06.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Wang AM, Miyata Y, Klinedinst S, Peng H-M, Chua JP, Komiyama T, et al. Activation of Hsp70 reduces neurotoxicity by promoting polyglutamine protein degradation. Nat Chem Biol. 2013;9:112–118. doi: 10.1038/nchembio.1140. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Miyata Y, Li X, Lee H-F, Jinwal UK, Srinivasan SR, Seguin SP, et al. Synthesis and Initial Evaluation of YM-08, a Blood-Brain Barrier Permeable Derivative of the Heat Shock Protein 70 (Hsp70) Inhibitor MKT-077, Which Reduces Tau Levels. ACS Chemical Neuroscience. 2013;4:930–939. doi: 10.1021/cn300210g. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Wiech M, Olszewski MB, Tracz-Gaszewska Z, Wawrzynow B, Zylicz M, Zylicz A. Molecular Mechanism of Mutant p53 Stabilization: The Role of HSP70 and MDM2. PLoS ONE. 2012;7:e51426. doi: 10.1371/journal.pone.0051426. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Muller P, Hrstka R, Coomber D, Lane DP, Vojtesek B. Chaperone-dependent stabilization and degradation of p53 mutants. Oncogene. 2008;27:3371–3383. doi: 10.1038/sj.onc.1211010. [DOI] [PubMed] [Google Scholar]
  • 123.Walerych D, Olszewski MB, Gutkowska M, Helwak A, Zylicz M, Zylicz A. Hsp70 molecular chaperones are required to support p53 tumor suppressor activity under stress conditions. Oncogene. 2009;28:4284–4294. doi: 10.1038/onc.2009.281. [DOI] [PubMed] [Google Scholar]
  • 124.Massey A, Williamson D, Browne H, Murray J, Dokurno P, Shaw T, et al. A novel, small molecule inhibitor of Hsc70/Hsp70 potentiates Hsp90 inhibitor induced apoptosis in HCT116 colon carcinoma cells. Cancer Chemother Pharmacol. 2010;66:535–545. doi: 10.1007/s00280-009-1194-3. [DOI] [PubMed] [Google Scholar]
  • 125.Chang L, Miyata Y, Ung PMU, Bertelsen EB, McQuade TJ, Carlson HA, et al. Chemical Screens against a Reconstituted Multiprotein Complex: Myricetin Blocks DnaJ Regulation of DnaK through an Allosteric Mechanism. Chemistry & Biology. 2011;18:210–221. doi: 10.1016/j.chembiol.2010.12.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Vos MJ, Hageman J, Carra S, Kampinga HH. Structural and Functional Diversities between Members of the Human HSPB, HSPH, HSPA, and DNAJ Chaperone Families†. Biochemistry. 2008;47:7001–7011. doi: 10.1021/bi800639z. [DOI] [PubMed] [Google Scholar]
  • 127.Phillips PA, Sangwan V, Borja-Cacho D, Dudeja V, Vickers SM, Saluja AK. Myricetin induces pancreatic cancer cell death via the induction of apoptosis and inhibition of the phosphatidylinositol 3-kinase (PI3K) signaling pathway. Cancer Letters. 2011;308:181–188. doi: 10.1016/j.canlet.2011.05.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Lu J, Papp LV, Fang J, Rodriguez-Nieto S, Zhivotovsky B, Holmgren A. Inhibition of Mammalian Thioredoxin Reductase by Some Flavonoids: Implications for Myricetin and Quercetin Anticancer Activity. Cancer Research. 2006;66:4410–4418. doi: 10.1158/0008-5472.CAN-05-3310. [DOI] [PubMed] [Google Scholar]
  • 129.López-Lázaro M, Martín-Cordero C, Toro MV, Ayuso MJ. Flavonoids As DNA Topoisomerase I Poisons. Journal of Enzyme Inhibition and Medicinal Chemistry. 2002;17:25–29. doi: 10.1080/14756360290011744. [DOI] [PubMed] [Google Scholar]
  • 130.Chen D, Daniel KG, Chen MS, Kuhn DJ, Landis-Piwowar KR, Dou QP. Dietary flavonoids as proteasome inhibitors and apoptosis inducers in human leukemia cells. Biochemical Pharmacology. 2005;69:1421–1432. doi: 10.1016/j.bcp.2005.02.022. [DOI] [PubMed] [Google Scholar]
  • 131.Huryn DM, Brodsky JL, Brummond KM, Chambers PG, Eyer B, Ireland AW, et al. Chemical methodology as a source of small-molecule checkpoint inhibitors and heat shock protein 70 (Hsp70) modulators. Proceedings of the National Academy of Sciences. 2011;108:6757–6762. doi: 10.1073/pnas.1015251108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Braunstein MJ, Scott SS, Scott CM, Behrman S, Walter P, Wipf P, et al. Antimyeloma Effects of the Heat Shock Protein 70 Molecular Chaperone Inhibitor MAL3-101. J Oncol. 2011;2011:232037. doi: 10.1155/2011/232037. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Adam C, Baeurle A, Brodsky JL, Wipf P, Schrama D, Becker JC, et al. The HSP70 Modulator MAL3-101 Inhibits Merkel Cell Carcinoma. PLoS ONE. 2014;9:e92041. doi: 10.1371/journal.pone.0092041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Neckers L, Workman P. Hsp90 Molecular Chaperone Inhibitors: Are We There Yet? Clinical Cancer Research. 2012;18:64–76. doi: 10.1158/1078-0432.CCR-11-1000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Zou JY, Guo YL, Guettouche T, Smith DF, Voellmy R. Repression Of Heat Shock Transcription Factor Hsf1 Activation By Hsp90 (Hsp90 Complex) That Forms a Stress-Sensitive Complex With Hsf1. Cell. 1998;94:471–480. doi: 10.1016/s0092-8674(00)81588-3. [DOI] [PubMed] [Google Scholar]
  • 136.Zaarur NG, VL, Porco J, Calderwood S, Sherman M. Targeting Heat Shock Response to Sensitize Cancer Cell to Proteasome and Hsp90 Inhibitors. Cancer Research. 2006;66:1783–1791. doi: 10.1158/0008-5472.CAN-05-3692. [DOI] [PubMed] [Google Scholar]
  • 137.Miyata Y, Rauch Jennifer N, Jinwal Umesh K, Thompson Andrea D, Srinivasan S, Dickey Chad A, et al. Cysteine Reactivity Distinguishes Redox Sensing by the Heat-Inducible and Constitutive Forms of Heat Shock Protein 70. Chemistry & Biology. 2012;19:1391–1399. doi: 10.1016/j.chembiol.2012.07.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Koren J, Jinwal UK, Jin Y, O’Leary J, Jones JR, Johnson AG, et al. Facilitating Akt Clearance via Manipulation of Hsp70 Activity and Levels. Journal of Biological Chemistry. 2010;285:2498–2505. doi: 10.1074/jbc.M109.057208. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES