HSP70 in Cancer: History
Please note this is an old version of this entry, which may differ significantly from the current revision.
Subjects: Cell Biology
Contributor:

The 70-kDa heat shock proteins (HSP70s) are abundantly present in cancer, providing malignant cells selective advantage by suppressing multiple apoptotic pathways, regulating necrosis, bypassing cellular senescence program, interfering with tumor immunity, promoting angiogenesis and supporting metastasis. This direct involvement of HSP70 in most of the cancer hallmarks explains the phenomenon of cancer “addiction” to HSP70, tightly linking tumor survival and growth to the HSP70 expression. HSP70 operates in different states through its catalytic cycle, suggesting that it can multi-function in malignant cells in any of these states. Clinically, tumor cells intensively release HSP70 in extracellular microenvironment, resulting in diverse outcomes for patient survival. Given its clinical significance, small molecule inhibitors were developed to target different sites of the HSP70 machinery. Furthermore, several HSP70-based immunotherapy approaches were assessed in clinical trials.

  • HSP70
  • cancer
  • immunotherapy
  • apoptosis
  • authophagy
  • senescence
  • metastasis
  • exosomes

1. Introduction

The discovery of impact of heat on chromosome puffing patterns of Drosophila in 1962 by Ritossa and the identification of heat-inducible genes and proteins opened a new field of research on heat shock response [1][2][3][4]. Later it became clear that high expression of heat shock proteins (HSPs), and especially 70 kDa proteins (HSP70), protects cells against stress [4][5][6]. Further research revealed the chaperoning function of HSP70 being responsible for its ability to enhance cell survival through its ability to catalyze reassembly of damaged ribonucleoproteins [4][7]. Serving as a molecular chaperone, HSP70 mediates a wide range of house-keeping activities (reviewed in [8]). House-keeping and stress-related functions of HSP70 include de novo protein folding and refolding, prevention of protein aggregation, degradation of proteins, transport of proteins across membranes, assembly and disassembly of protein complexes.
The HSP70 family is highly conserved in evolution [9][10][11]. Multiple HSP70s present in both prokaryotes and eukaryotes. In humans, 13 HSP70 homologues are found in different compartments (cytosol, nucleus, lysosomes, ER and mitochondria), suggesting individual and organelle-specific biological roles (reviewed in [4]) [8]. Long lines of experimental evidence suggest a crucial role of HSP70 in cancer [12]. It is highly expressed in malignant tumors and typically serves as a biomarker for poor prognosis [13][14].

2. The HSP70 Machinery

The central function of the HSP70 chaperones is that they do not work alone, but rather as machinery of HSP70 and (co)chaperones collaborating with each other [15]. To perform such collaborations, HSP70 can operate in different states during its functional cycle. Co-chaperones that are involved in HSP70 functional cycle form an internal HSP70 network. During its functional cycle and within its internal network HSP70 directly interacts with the client proteins to perform its chaperone function. Concomitantly, HSP70 can handover client proteins to other (co)chaperone machines for further folding or degradation and this will be further referred to as an external HSP70 network.

2.1. HSP70 Structure

A full-length crystal structure of human HSP70 in either its free or closed conformation has not yet been obtained. Structures of its two major domains, namely N-terminal nucleotide-binding domain (NBD, ~45kDa), responsible for ATPase activity, and C-terminal substrate-binding domain (SBD, ~25kDa), required for peptide binding, have been determined independently of each other in free or bound states [16][17][18][19][20][21][22][23][24][25][26][27][28]. The two HSP70 domains are connected by the linker (13aa) (Figure 1A) [27]. HSP70 binds to exposed hydrophobic residues on unfolded proteins and unlike HSP90, HSP70 does not have specific clients for binding [29]. From the HSP70-structural point of view, further studies should be performed to fulfill the currently missing communication between its two domains.

2.2. HSP70 Functional Cycle

In 1995 Ha and McKay demonstrated that constitutive isoform of HSP70, Heat shock cognate 70 (HSC70), undergoes several conformational changes through its ATPase cycle (Figure 1) [30]. The binding of polypeptide to SBD is tightly regulated by the NBD nucleotide states either being ADP-, ATP- bound or nucleotide-free. In the ADP-bound state, SBD binds to polypeptides with high affinity and slow association and dissociation rates [8]. In ADP-bound state SBDα forms the lid over the SBDβ, resulting in a closed conformation of SBD (Figure 1C) [17][27][30]. ATP binding to NBD decreases the affinity of SBD for substrate, increasing association and dissociation rates by 100-fold and 1000-fold, respectively (Figure 1A) [8][30][31][32][33]. This mechanism of rapid association of peptide and its timely release allows HSP70 to prevent peptide aggregation and perform its folding function [8][34]. Therefore, ATP-ADP exchange is crucial for HSP70 polypeptide uptake and release.

2.3. The Internal HSP70 Network

HSP70 functional cycle is regulated by co-chaperone HSP40 and nucleotide exchange factors (NEFs) (Figure 1) [35][36]. HSP40 chaperones, also known as J-domain proteins (JDPs), play important roles in HSP70 machinery. J-domain of HSP40 is involved in the stimulation of ATPase activity of the HSP70-NBD domain, whereas HSP40 C-terminal peptide-binding domain 1 (CTD1) interacts with C-tail of HSP70 (Figure 1B) [15][37][38]. HSP40 binds to non-native clients and engages HSP70 in its low-affinity ATP state. Binding of HSP40 to HSP70 displaces the client to HSP70 C-tail, causing HSP40 dissociation from the HSP70-HSP40 complex and leaving the client bound to HSP70 in its ADP state [38]. Therefore, HSP40 delivers clients to HSP70 and stimulates ATPase domain, shifting HSP70 towards its high-affinity (ADP) state.
Interestingly, there are much more HSP40 isoforms found in cells than HSP70s [15][39][40]. Moreover, much lower concentration of J-protein is required to that of HSP70 to stimulate client folding [41]. Increase in cellular level of J-protein without rising up HSP70 level results in reduced refolding of client proteins [15]. For some cellular functions full structure of J-protein is not required for its J-domain is sufficient to stimulate HSP70 ATPase activity [15]. In such cases, there will be no direct J-protein interaction with the client itself, while increase in just J-domain level in a specific compartment will direct HSP70 to particular clients at this site [15]. Several HSP40 family members were described in tumor growth, however, their role is still not clearly defined (reviewed in [42]). Hence, understanding the effects of diversity of HSP40 members on HSP70 and how HSP40s manipulate HSP70 functions can be advantageous for the development of HSP70 modulators targeting specific locations in cancer research.
Contrary to HSP40, which triggers HSP70 transition to ADP-bound state, NEFs facilitate ADP dissociation from NBD, resulting in rebinding of ATP and substrate release (Figure 1D) [8]. Eukaryotic NEFs include Bcl2-associated anthanogen (BAG), HSP110, HspBP1, each using distinct mechanism for nucleotide exchange (reviewed in [43]). It is important to note that NEFs do not exchange nucleotides, but rather use their bulk mass to stabilize open conformation of HSP70 [43][44]. Binding of ATP displaces NEF from NBD for the next round of peptide binding [43]. Intriguingly, NEF function can be attenuated by the HSP70-interacting protein (Hip). Hip preferentially binds to ADP-bound state of HSP70-peptide complexes, thereby, slowing ADP dissociation from NBD [45]. Hip showed to promote protein ubiquitination and degradation [46]. Several antitumor compounds were described to successfully mimic Hip function [46][47]. Therefore, Hip is involved in prevention of protein aggregation by competing with NEF for proceeding to the next stage of the HSP70 cycle or transfer of proteins to degradation.
Figure 1. Functional cycle of HSP70 chaperones. (A) Structure of HSP70 in low-affinity (ATP) state. ATP binds to NBD, resulting in an open conformation of SBD, ready for client binding [8]. (B) HSP70-HSP40 complex. HSP40 presents non-native clients to HSP70. J-domain of HSP40 binds to HSP70-NBD stimulating its ATPase activity. Binding of HSP70 C-tail to HSP40-CTD1 displaces the client to HSP70, shifting HSP70 to ADP-bound state [15][37][38]. (C) High-affinity (ADP) state. ADP binds to NBD, SBDα forms the lid over SBDβ, locking substrate in SBD [17][27]. (D) HSP70-NEF complex. NEF displaces ADP from NBD, allowing ATP to bind NBD, shifting the HSP70 to low-affinity (ATP) state [8].NBD, nucleotide-binding domain; SBDα/β, substrate-binding domain; NEF, nucleotide exchange factor; CTD1/2, C-terminal peptide-binding domain of HSP40; J; J domain of HSP40.

2.4. The External HSP70 Network

Newly synthesized peptides will be either released from HSP70 for spontaneous folding, transferred further to HSP90 chaperone or will be degraded [8]. The interaction between HSP70 and HSP90 is mediated by HSP70-HSP90 organizing protein (HOP) (Figure 2A) [48][49][50]. HOP primarily consists of tetratrico-peptide repeat (TPR) domains, acting as a bridge between HSP70 and HSP90 for passing a client protein [48]. In cases when substrates spend too long time bound to HSP70 and cannot be refolded or handed over to HSP90, these substrates are targeted for degradation mediated by C-terminus of HSP70 interacting protein (CHIP) (Figure 2B) [51][52][53][54]. CHIP is an E3 ubiquitin ligase which works as a co-chaperone, competing with HOP for binding to HSP70 C-terminus and SBDα through its TPR-domain [55][56].
Figure 2. The external HSP70 network. (A) HSP70-HSP90-HOP complex. HOP acting as an adaptor molecule mediates the association between HSP70 and HSP90. HSP70 C-terminal EEVD motif binds TPR1 on HOP, whereas HSP90 C-terminal EEVD binds TPR2A [48]. This association via HOP allows handing over the substrate from HSP70 to HSP90 for further folding. (B) HSP70-CHIP complexes. CHIP binds to C-terminus of HSP70-substrate complex as well as to SBDα lid through its TPR domain, ubiquitylates HSP70-bound peptides and targets them for proteasomal degradation [54][56]. HOP, HSP70/HSP90-organizing protein; TPR, tetratrico-peptide repeats; CHIP, C-terminus of HSP70 interacting protein.
There is a growing interest in both academia and pharmaceutical industry in targeting HSP90 in cancer research [57]. HSP90 has a wide range of clients, some of which are oncogenic proteins, responsible for driving tumor transformation and progression [58][59]. Notably, HSP70-HSP90 multi-chaperone complex termed ‘epichaperome’ was found to be expressed in more than half of all tumors [60]. It is interesting to point out that this tight association in the epichaperome complex promotes tumor survival simultaneously making malignant cells more sensitive to HSP90 inhibitors [60].

3. Diverse Functions of HSP70 in the Hallmarks of Cancer

Being abundantly expressed in cancer, HSP70s have a wide range of activities. With a particular focus on a major stress-inducible HSP70 member, we will examine here HSP70 involvement in the hallmarks of cancer—a concept of multistep tumor development originally proposed by Hanahan and Weinberg (Figure 4) [61][62].
Figure 4. HSP70 in the Hallmarks of Cancer. Taking into account emerging role of HSP70 in cancer, it is important to assess its association with the currently established Hanahan and Weinberg model of the Hallmarks of Cancer [61]. Even though there are a lot of unanswered questions concerning HSP70 functions in tumor progression, current model can serve for the better understanding of diverse HSP70 roles in cancer for future development of anti-cancer therapeutics.WASF3, Wiskott-Aldridge syndrome family member 3; MMP, Matrix metalloproteinase; HIF-1, hypoxia-inducible factor-1; AIMP2-DX2,Aminoacyl-transfer RNA synthetase-interacting multifunctional protein 2 lacking exon 2

3.1. HSP70 and Tumor Immunity

Cancer progresses over years continuously shaping its interaction with an immune system. The role of HSP70 in that interaction is not yet completely understood making it one of the major endeavors in cancer immune biology. Immunogenic nature of HSP70 comes from its ability to bind antigenic peptides derived from tumors [63]. Blachere and colleagues showed that HSP70- peptide complex resulted in antigen-specific CD8+ T cell response [64]. This experiment has opened new perspectives for the use of HSP70 as an adjuvant [63].
HSP70 is highly expressed on the surface of tumor cells [65]. These HSP70-membrane positive tumor cells actively release HSP70 surface-positive exosomes leading to the stimulation of NK cells [66]. Multhoff and colleagues showed that surface-bound HSP70s are recognized by NK cells pre-stimulated with HSP70 protein or 14-mer HSP70 peptide (TKD) in the presence of interleukin-2 (IL-2) or IL-15 cytokines [67][68][69][70][71]. Notably, TKD peptide alone is not sufficient to stimulate NK cells, requiring the presence of IL-2 or IL-15 [72][67][68]. In 2015 phase II clinical trial was initiated to study the efficacy of autologous NK cells pre-activated with TKD and IL-2 following radiochemotherapy on patients with non-small cell lung carcinoma [71].
Considering results from the clinic, it is important to point out that there are two types of HSP70 circulating in the serum of cancer patients. One of the types is exosomal HSP70 released by viable tumor cells, whereas another is HSP70 released by dying cancer cells serving as damage-associated molecular patterns (DAMPs) [73][74][75].
Acting as a DAMP released by necrotic cells, HSP70 has great immunogenic potential for eliciting strong anti-tumor T cell response either being bound to tumor antigen or antigen-free [76][77]. However, long-term exposure of immune cells to free HSP70 following radiotherapy showed to induce immune tolerance and promote tumor growth [78]. These findings are consistent with report showing that low dose of HSP70-peptide complex is sufficient to stimulate antitumor immunity [64]. Translating this to the endogenous HSP70 released from tumor cells, it can be proposed that initial release of HSP70 acts as tumor suppressant, while HSP70 overload leads to tumor progression.
HSP70-membrane positive phenotype (mHSP70) showed controversial results in predicting patients’ prognosis where positive clinical outcome was achieved for gastric and colon cancer, supporting the idea of NK targeting mHSP70, and poor prognosis in lower rectal and squamous cell carcinomas [79]. Authors deduced that differences in clinical outcomes can be explained by the differences in metastatic routes of these cancers [79]. Therefore, HSP70 release in extracellular microenvironment may lead to killing of mHSP70 tumor cells by NK cells or result in immune tolerance to tumor antigens.
HSP70 showed to stimulate both innate and adaptive immune responses [80]. In this regard, extracellular HSP70 (eHSP70) activated T regulatory cells resulting in downregulation of interferon-γ (IFN-γ) and tumor necrosis factor-α (TNF-α) and upregulation of IL-10 and transforming growth factor-β (TGF-β) cytokines [81]. The eHSP70 interaction with antigen-presenting cells (APC) led to the stimulation of inflammatory cytokines such as TNF-α, IL-1β, IL-6 and IL-12 via activation of nuclear factor kappa B (NF-kB) [82][83][84][85]. Furthermore, HSP70 showed to negatively regulate critical component of innate immune response – the Nod-like receptor protein-3 (NLRP3) inflammasome [86]. Taking into account cytosolic location of NLRP3 inflammasome, its ability to sense DAMPs and its dysregulation in cancer, further studies should be performed to understand the effect of intracellular HSP70 on NLRP3 inflammasome in the context of cancer [87]. In light of the reported, other innate immunity molecule Tag7 showed to form a complex with HSP70 [88]. This Tag7-HSP70 complex interacts with Tumor necrosis factor receptor 1 (TNFR1) leading to permeabilization of lysosomal membrane and subsequent tumor cell death [89][88]. T lymphocytes also showed to utilize Tag7-HSP70 cytotoxic complex [90].
HSP70-positive exosomes showed to activate myeloid-derived suppressor cells (MDSCs) leading to IL-6 production [91]. Intriguingly, Bausero and co-workers (2004) demonstrated that IL-10 and IFN-γ cytokines lead to active release of HSP70 in exosomes by increasing intracellular level of HSP70 [92]. Since high levels of IL-10 and IFN-γ have been found in the serum of cancer patients, it can be hypothesized that these cytokines initially mark tumor cells for immune recognition in the form of HSP70-bound exosomes, while chronic activation of these cytokines and their subsequent effect on promoting HSP70 overexpression may lead to immune tolerance [93][94]. However, further studies should be performed for a better understanding of relationships between cytokines and HSP70 in shaping tumor immunity.
The concept of HSP70-based vaccines was proposed utilizing the ability of HSP-peptide complexes to elicit cytotoxic T lymphocyte (CTL) response by cross-priming. In that context, HSP70-peptide complexes extracted from tumor were used to immunize the host, thus, delivering these complexes to antigen-presenting cells (APCs) for T cell priming through MHC class I pathway [95]. Notably, HSP70-peptide complexes isolated from fusion of tumor and dendritic cells (DCs) showed the most promising results [96]. Other HSP70-based immunotherapy approaches which have entered clinical trials will be discussed later in this review.
In light of that, one of the top 10 challenges described by Hegde and Chen (2020) in cancer immunotherapy is to determine dominant drivers of tumor immunity [97]. In this regard, HSP70 showed to be clinically important in different types of cancer and tumor immunology, however, further research is needed to determine its role in driving tumor immunity.

3.2. HSP70 and Tumor Resistance to Cell Death

Investigators exploring the role of HSPs on malignant cells proposed that tumor cells are “addicted to chaperones” [98]. Indeed, tumor cells survive under conditions of continuous stress such as hypoxia and nutrient starvation which make them susceptible to apoptosis [61]. HSP70, which is abundantly found in cancers, showed to suppress apoptosis, allowing cancer to progress [99]. Notably, decrease in the level of HSP70 showed to induce cancer cell death [100].
Evidently, HSP70 showed to affect both intrinsic and extrinsic apoptotic pathways (Figure 5). In the model of TNF-α induced apoptosis it was reported that the HSP70-CHIP complex promotes proteasomal degradation of apoptosis signal-regulating kinase 1 (ASK1) resulting in inhibition of c-Jun N-terminal kinase (JNK) and p38 (Figure 5A) [101]. JNK activity showed to be important for the release of cytochrome c from the mitochondria for the induction of intrinsic apoptotic pathway [102]. This JNK-dependent apoptosis involves p53 for the upregulation of Bcl-2-associated X protein (Bax) and BH3 interacting-domain death agonist (Bid) [103][104]. In addition, HSP70 inhibits apoptosis-inducing factor (AIF) required for DNA fragmentation [105]. In extrinsic pathway, HSP70 showed to block the formation of death-inducing signalling complex (DISC) through its interaction with TNF-related apoptosis-inducing ligand receptor 1 (TRAIL-R1) and TRAIL-R2 (Figure 5B) [106]. This may also result in an impaired cleavage of Bid by caspase-8 [107]. Therefore, HSP70 suppresses apoptosis by caspase-dependent mechanism via inhibition of JNK, caspase-independent mechanism via inhibition of AIF and through the interaction with death receptors.
Apart from apoptosis, HSP70 plays an important role in alternative forms of cell death including autophagy and necrosis. In contrast to normal cells, HSP70 was found to be present in the lysosomes of cancer cells [108]. Several studies have demonstrated that HSP70 stabilizes lysosomal membranes allowing cancer cells to escape cell death (Figure 5A) [108][109][110]. HSP70 inhibitor apoptozole showed to be effective for promoting apoptosis in cancer cells by triggering lysosomal membrane permeabilization and cathepsin release [111].
HSP70 protection from necrotic cell death is mediated through inhibition of JNK and does not require its chaperone activity [112]. As discussed earlier, HSP70 acts as DAMP and can initially elicit strong immunogenic response which with chronic exposure leads to immune tolerance. Therefore, it can be concluded that HSP70 plays dual role in necrosis of malignant cells. On the one hand, following necrosis HSP70 may induce anti-tumor response. On the other hand, HSP70 protects cancer cells from the necrotic cell death, thus, allowing for tumor progression which appears to be the case with long-term anti-cancer treatment accompanied by the sustained release of necrotic signals, leading to tumor growth and resistance to therapy [113].
Figure 5. HSP70 in apoptotic signaling pathways. (A) HSP70 in intrinsic apoptosis. In TNF-α induced apoptosis, HSP70 inhibits JNK and p38, thus, preventing the translocation of Bax and Bid to the mitochondria [101][107][114]. This blocks the release of cytochrome c to the cytosol and its formation of apoptosome complex with Apaf-1 and caspase-9. HSP70 may also disrupt the formation of apoptosome by binding to Apaf-1, preventing its association with procaspase-9 [115]. In addition, HSP70 stabilizes lysosomal membrane, inhibiting cathepsin release to the cytosol for the activation of intrinsic apoptotic pathway. (B) HSP70 in extrinsic apoptosis. HSP70 interacts with TRAIL-R1 and TRAIL-R2 and prevents the formation of DISC complex required for the initiation of extrinsic apoptosis. JNK, c-Jun N-terminal kinase; Cyt C, cytochrome c; Bax, Bcl-2-associated X protein; Bid, BH3 interacting-domain death agonist; p38, p38 mitogen-activated protein kinase; Apaf-1, apoptotic protease-activating factor 1; ASK, apoptosis signal-regulating kinase 1; TRAIL, TNF-related apoptosis-inducing ligand; TRAIL-R1/R2, TNF-related apoptosis-inducing ligand receptor 1/2; FADD, FAS-associated death domain (FADD).

3.3. HSP70 and Senescence Program

High expression of HSP70 is critical for the survival and growth of cancer cells as HSP70 depletion in cancer showed to activate oncogene-induced senescence program. HSP70 suppresses senescence by regulating p53 and cell cycle kinase Cdc2 [116]. Furthermore, depletion of HSP70 triggers the activation of p53-dependent or ERK-dependent senescence pathways depending on the expression of different oncogenes. For example, cells that express PIK3CA oncogene led to p53-dependent senescence, while cells with Ras oncogene expression activated ERK-dependent senescence [117]. Along this line, HSP70 downregulation showed to inhibit the transformation of Her2-expressing non-malignant mammary epithelial cells (MCF10A) leading to cellular senescence by upregulation of p21 and depletion of survivin [118]. Interestingly, results from another Her2 –positive cell line SK-BR-3 showed that p21 and survivin pathways function either in p53-dependent or p53-independent manner, respectively [118]. In addition, nucleotide-exchange factor BAG3 also showed to play a role in regulation of survivin and p21 in several cancer lines, suggesting potential role of HSP70-BAG3 complex in suppression of senescence [119].
It has been shown that downregulation of heat shock transcription factor 1 (HSF1) inhibits tumor growth and promotes cellular senescence by regulating HSP70 [120]. However, recent report showed that HSF1 can independently induce cellular senescence via p53-p21 pathway [121].

3.4. HSP70 in Sustained Proliferation

Recent reports have demonstrated that HSP70 in complex with HSP40 co-chaperone inactivates tumor suppressor protein p53 by interacting with its DNA binding domain, leading to p53 unfolded state [122][123]. These findings shed the light on why the depletion of HSP70 resulted in upregulation of p53 target molecule p21 and subsequent activation of senescence program discussed earlier in this review [118][122]. In contrast, transfer of p53 from HSP70 to HSP90 via HOP showed to stabilize p53 in its native conformation [122][123]. Therefore, HSP70-HSP40 and HSP70-HOP-HSP90 interplay is important in maintaining p53 conformational equilibrium [122].
Along this line, HSP70 showed to stabilize aminoacyl-transfer RNA synthetase-interacting multifunctional protein 2 (AIMP2) lacking exon 2 (AIMP2-DX2), the splicing variant of AIMP2. AIMP2 acts as pro-apoptotic factor positively regulating p53 in contrast to its splicing variant AIMP-DX2 which expression showed to correlate with aggressive tumor phenotype and chemoresistance [22][124]. AIMP2-DX2 escapes the Siah1- dependent ubiquitination by binding to HSP70 SBD domain [22]. Taken together, these data provide a good example on how HSP70 can exert its actions by manipulating the conformational state of its client proteins or protecting them from further degradation. Therefore, further studies are required for identifying other cancer-related client proteins of HSP70 machinery.

3.5. HSP70 in Angiogenesis

HSP70 and HSP90 showed to stabilize the accumulation of hypoxia-inducible factor-1α (HIF-1α), one of the two subunits of hypoxia-inducible factor-1(HIF-1) [125]. Importantly, HIF-1 which is responsible for sensing low level of oxygen showed to be involved in tumor angiogenesis, invasion and metastasis [125][126][127]. It has been reported that HSP70 binds and stabilizes HIF-1α in cancer cell lines [128]. Crucially, HIF-1 showed to be regulated by HSF1 via HuR, an RNA-binding protein overexpressed in cancer [129][130]. The effects of HSF1 on HIF-1 via HuR showed to be HSP70-independent, suggesting that HSP70 and HSF1 may affect HIF-1 through distinct mechanisms [129]. Kim and colleagues showed that extracellular HSP70 can bind to the surface of endothelial cells and induce angiogenesis by ERK-dependent mechanism [131]. In addition, it has been reported that HSP70 enhances IL-5-induced angiogenesis via endothelial nitric oxide synthase (eNOS) pathway, suggesting an inflammatory role of HSP70 in promoting angiogenesis [132].

3.6. HSP70 in Metastasis

High expression of HSP70 was shown to be associated with metastases to the lymph nodes in the breast cancer models [133]. Further research has found that inactivation of HSP70 showed to reduce tumor invasiveness and metastatic potential of breast, cervical and bladder cancer cell lines [134]. High HSP70 expression protects tumor cells against anoikis and amorphosis, forms of cell death that occur when cell loses its contact with extracellular matrix (ECM) [135][136]. In this regard, molecules involved in anoikis and amorphosis such as focal adhesion kinase (FAK) and Akt showed to be affected by HSP70 [137][138][139].
Epithelial-mesenchymal transition (EMT) is a complex process in which cancer cell acquires migratory and metastatic potential. Li and colleagues demonstrated that HSP70-peptide complexes (HSP70-PC) extracted from hepatocarcinoma samples induce EMT in human hepatoma (Huh-7) cell line acting through p38 MAPK signalling pathway [140]. In contrast, recent studies using small interfering RNA (siRNA) and short hairpin RNA (shRNA) have showed that HSP70 inactivation promotes cell transition to mesenchymal phenotype [136][141]. HSP70 downregulation destabilizes E-cadherin-catenin complexes allowing migration of tumor cells [136]. In addition, several reports have demonstrated that HSP70 inhibits TGF-β-induced EMT by decreasing Smad2 phosphorylation [142][143][144]. Furthermore, it has been shown that HSP70 prevents high glucose-induced EMT of peritoneal mesothelial cells by inhibiting Smad3 and Smad4 phosphorylation and reactive oxygen species (ROS) production [145]. These findings suggest dual role of HSP70 in EMT, where HSP70-peptide complexes support transition of tumor cells towards acquiring mesenchymal phenotype. Intriguingly, HSP70 (co)chaperones also showed to be involved in tumor metastasis. One of the key EMT regulators Met is downregulated by C-terminus of HSP70 interacting protein (CHIP) [138][146]. High expression of HSP70’s co-chaperone HOP has been found in ovarian cancer and glioblastoma cells lines [147][148]. HOP inactivation has been reported to inhibit pseudopodia formation and migration of breast cancer cells [149]. Furthermore, HOP knockdown downregulates matrix metalloproteinase-2 (MMP-2), resulting in reduced invasion of pancreatic cancer cells [150]. Another HSP70’s co-chaperone BAG3 enables MMP-2 to promote invasion of ovarian cancer cells [151]. Extracellular HSP70 induces the release of MMP-9, a processing enzyme for cleavage of CD44 receptor involved in adhesion and migration [152][153]. In addition, HSP70 and HSP90 proteins showed to stabilize Wiskott-Aldridge syndrome family member 3 (WASF3), protein required for tumor migration, invasion and metastasis [154][155][156].
Clinically, circulating tumor cells (CTCs) that underwent EMT serve as a prognostic marker of a metastatic spread and can be isolated from patient’s blood using epithelial cell adhesion molecule (EpCAM) as a cell surface marker [157]. Interestingly, in contrast to EpCAM, the EMT does not downregulate HSP70 expression as higher membrane-bound HSP70 (mHSP70) expression has been found on metastatic tumors compared to its primary counterparts [79][157][158][159][160]. In this regard, Stangl and colleagues used HSP70-TKD peptide to develop monoclonal antibody directed against mHSP70 for specific targeting mHSP70-positive tumors by NK cells and efficient isolation of CTCs [157][161]. Taken together, on the one hand, high surface expression of HSP70 is found in metastatic tumors and mHSP70 may serve as a biomarker for assessing the metastatic spread. On the other hand, HSP70 plays dual role in EMT. The possible explanation can be that HSP70 works in a network where interaction of HSP70 either with BAG3 or HOP has a positive effect on tumor progression and this is consistent with the idea of HSP70-PC to promote EMT as HSP70 either bound to BAG3 or HOP may still carry a client protein. Conversely, interaction with CHIP suppresses EMT. Therefore, it can be hypothesized that different HSP70 roles in metastasis are specified through HSP70 interactions with its internal and external network of its (co)chaperones. However, further research is required in elucidating the role of HSP70 and its (co)chaperones in metastatic process.

4. HSP70 Therapies Targeting Cancer

4.1. Small Molecule Inhibitors of HSP70 Cycle

Attempts were made to develop adenosine-derived HSP70 inhibitors using structure-based design approach [162][163]. Non-nucleotide ligands mimicking HSP70-NBD in ADP-bound state were proposed as well, however, from drug discovery perspective, it remains challenging to develop compounds targeting nucleotide-binding domain [18]. HSP70 small molecule inhibitors targeting cancer are summarized in Table 1. Interesting results were gained by Li and colleagues with JG-98 inhibitor. JG-98 disrupts HSP70-BAG3 complex, thereby, freezing HSP70 in its ADP state [164][165]. Similar effects were reported with YM-1, an inhibitor that targets HSP70-NEF complex, indicating that freezing of HSP70 in ADP-bound state can be the most promising strategy for the future developments of anti-cancer drugs [119]. In such state of tight association of HSP70 with peptide, this complex will be ubiquitinated by CHIP and further targeted for proteasomal degradation [46][166].
ATPase inhibitors proved to be effective in vitro and in vivo as one molecule can target multiple HSP70 isoforms and this strategy might be beneficial for cancer types with overexpression of multiple HSP70 isoforms [167][168][169]. However, “one-size-fits-all” approach may not work for distinct HSP70 family members involved in specific cancer type, in such cases it is important to consider isoform-specific inhibitors for future developments. Therefore, it is critical to gain better understanding of HSP70 functional cycle, including conformational changes and interaction between its two domains, the roles (co)chaperones play in HSP70 machinery as well as roles of HSP70 isoforms in different cancer types for future design of efficient HSP70 inhibitors for cancer therapy.
                       Table 1. Small molecules targeting HSP70 functional cycle in cancer.
HSP70-Targeting Molecules Effects Refs.
SBD-targeting inhibitors PES (Pifithrin-µ) Suppressed tumor development in a mouse model of Myc-lymphoma. [170]
ADD70 Increased sensitivity of colon cancer and melanoma cells to apoptosis; showed antimetastatic effects in vivo. [171]
Acridizinium derivative 1 Induced apoptosis in HeLa cells. [172]
NBD-targeting inhibitors Synthetic peptide P17 Decreased melanoma growth in vivo. [173]
JG-98 inhibitor
(MKT-077 analog)
Decreased tumor growth in MCF7 xenograft model. [164]
VER-155008 Inhibited proliferation of human breast and colon cancer cell lines. [168]
YK-5 Induced apoptosis and degradation of HSP70/HSP90 client proteins (HER2,Raf-1, Akt kinases) in breast cancer cells. [174]
Apoptozole Induced apoptosis in ovarian, colon and lung cancer cell lines. [175][176]
HSP70-HSP40 complex inhibitors MAL3-101 Inhibited proliferation of multiple myeloma cells derived from patients. [177]
DMT3132
(MAL3-101 analog)
Reduced proliferation in breast cancer cells. [178][179]
Myricetin Inhibited tumor growth in pancreatic cancer. [180]
HSP70-NEF complex inhibitors YM-1 Inhibited tumor growth in mammary and melanoma xenograft models. [119]
Allosteric HSP70 inhibitors HS-72 Inhibited growth in HER2-positive breast cancel model. [181]

4.2. Other HSP70 Inhibitors

Triptolide which was originally isolated from Tripterygium wilfordii and its water-soluble pro-drug derivative minnelide showed to inhibit HSP70 expression in pancreatic cancer, neuroblastoma, mesothelioma, osteosarcoma and non-small cell lung carcinoma by inducing microRNA miR-142-3p binding to 3′UTR and downregulation of HSP70 mRNA level [182][183][184][185][186][187][188]. Notably, minnelide has entered phase II clinical trial for the treatment of refractory pancreatic cancer [189]. Quercetin and demethoxycurcumin also showed to inhibit HSP70 in breast and prostate cancer cells, respectively [190][191][192].

4.3. HSP70-Based Therapies in Cancer Clinical Trials

There is an enormous challenge in developing safe and effective drug that can pass all phases of clinical trials to further be accessible for the treatment of cancer patients. In this regard, HSP70 showed to be effective in vitro and in vivo, however, the success of HSP70-based therapies in cancer can only be verified by assessing their efficacy and safety in clinical trials (Table 2). In that regard it is encouraging that HSP70-based therapies showed to be well tolerated among patients [71][193][194][195][196][197][198]. However, since most of these clinical trials, except phase II trial of Specht and colleagues, were conducted in 2003-2015 years and have not been progressed through drug development timeline and further pharmacovigilance, assessment of the efficacy and safety of HSP70-based therapies for the treatment of cancer patients still remains a major challenge.
Table 2. HSP70-based therapies in cancer clinical trials and case studies.
HSP70-Based Therapies Condition Clinical Trial Phase/Case Study Refs.
Autologous HSP70-peptide complex in combination with imatinib mesylate Chronic myeloid leukemia Phase I [198]
Autologous HSP70-peptide complex (AG-858) in combination with Gleevec Chronic Myelogenous Leukemia Phase II [199]
Autologous peripheral blood mononuclear cells (PBMC) pre-activated with TKD peptide/IL-2 Colon carcinoma Case study [193]
Autologous NK cells pre-activated with TKD and IL-2 following radiochemotherapy Advanced colorectal carcinoma and non- small cell lung carcinoma (NSCLC) Phase I; Phase II [71][194]
DNA vaccine expressing HPV16 E7 mutant form and HSP70 Cervical intraepithelial neoplasia Phase I [195]
Dendritic cells transfected with HSP70 mRNA Hepatocellular carcinoma Phase I [196]
Autologous NK cells pre-treated with TKD/IL-2 in combination with radiochemotherapy and nivolumab (PD-1 antibody) NSCLC stage IIIb Case study [197]
Intratumoral injection of recombinant oncolytic type 2 adenovirus overexpressing HSP70 (H103) Advanced solid tumors Phase I [200]

This entry is adapted from the peer-reviewed paper 10.3390/cells9030587

References

  1. Ritossa, F. A new puffing pattern induced by temperature shock and DNP in drosophila. Cell. Mol. Life Sci. 1962, 18, 571–573.
  2. Tissières, A.; Mitchell, H.K.; Tracy, U.M. Protein synthesis in salivary glands of Drosophila melanogaster: Relation to chromosome puffs. J. Mol. Biol. 1974, 84, 389–398.
  3. Moran, L.; Mirault, M.-E.; Arrigo, A.P.; Goldschmidt-Clermont, M.; Tissières, A. Heat shock of Drosophila melanogaster induces the synthesis of new messenger RNAs and proteins. Philos. Trans. R. Soc. B Biol. Sci. 1978, 283, 391–406.
  4. Daugaard, M.; Rohde, M.; Jäättelä, M. The heat shock protein 70 family: Highly homologous proteins with overlapping and distinct functions. FEBS Lett. 2007, 581, 3702–3710.
  5. Beere, H. Stress management—Heat shock protein-70 and the regulation of apoptosis. Trends Cell Biol. 2001, 11, 6–10.
  6. Yang, X.; Wang, J.; Zhou, Y.; Wang, Y.; Wang, S.; Zhang, W.-Y. Hsp70 promotes chemoresistance by blocking Bax mitochondrial translocation in ovarian cancer cells. Cancer Lett. 2012, 321, 137–143.
  7. Pelham, H. Hsp70 accelerates the recovery of nucleolar morphology after heat shock. EMBO J. 1984, 3, 3095–3100.
  8. Rosenzweig, R.; Nillegoda, N.B.; Mayer, M.P.; Bukau, B. The Hsp70 chaperone network. Nat. Rev. Mol. Cell Biol. 2019, 20, 665–680.
  9. Hunt, C.; Morimoto, R.I. Conserved features of eukaryotic hsp70 genes revealed by comparison with the nucleotide sequence of human hsp70. Proc. Natl. Acad. Sci. USA 1985, 82, 6455–6459.
  10. Lindquist, S.; Craig, E.A. The heat-shock proteins. Annu. Rev. Genet. 1988, 22, 631–677.
  11. Gupta, R.S.; Singh, B. Phylogenetic analysis of 70 kD heat shock protein sequences suggests a chimeric origin for the eukaryotic cell nucleus. Curr. Biol. 1994, 4, 1104–1114.
  12. Mosser, D.D.; Morimoto, R.I. Molecular chaperones and the stress of oncogenesis. Oncogene 2004, 23, 2907–2918.
  13. Ciocca, D.R.; Calderwood, S.K. Heat shock proteins in cancer: Diagnostic, prognostic, predictive, and treatment implications. Cell Stress Chaperones 2005, 10, 86–103.
  14. Calderwood, S.; Khaleque, A.; Sawyer, U.B.; Ciocca, D.R. Heat shock proteins in cancer: Chaperones of tumorigenesis. Trends Biochem. Sci. 2006, 31, 164–172.
  15. Kampinga, H.H.; Craig, E. The HSP70 chaperone machinery: J proteins as drivers of functional specificity. Nat. Rev. Mol. Cell Biol. 2010, 11, 579–592.
  16. Arakawa, A.; Handa, N.; Ohsawa, N.; Shida, M.; Kigawa, T.; Hayashi, F.; Shirouzu, M.; Yokoyama, S. The C-Terminal BAG Domain of BAG5 Induces Conformational Changes of the Hsp70 Nucleotide- Binding Domain for ADP-ATP Exchange. Structure 2010, 18, 309–319.
  17. Arakawa, A.; Handa, N.; Shirouzu, M.; Yokoyama, S. Biochemical and structural studies on the high affinity of Hsp70 for ADP. Protein Sci. 2011, 20, 1367–1379.
  18. Cheeseman, M.; Westwood, I.M.; Barbeau, O.; Rowlands, M.; Dobson, S.; Jones, A.M.; Jeganathan, F.; Burke, R.; Kadi, N.; Workman, P.; et al. Exploiting Protein Conformational Change to Optimize Adenosine-Derived Inhibitors of HSP70. J. Med. Chem. 2016, 59, 4625–4636.
  19. Gao, X.-C.; Zhou, C.-J.; Zhou, Z.-R.; Wu, M.; Cao, C.-Y.; Hu, H.-Y. The C-terminal Helices of Heat Shock Protein 70 Are Essential for J-domain Binding and ATPase Activation*. J. Biol. Chem. 2012, 287, 6044–6052.
  20. Hassan, A.Q.; Kirby, C.A.; Zhou, W.; Schuhmann, T.; Kityk, R.; Kipp, D.R.; Baird, J.; Chen, J.; Chen, Y.; Chung, F.; et al. The Novolactone Natural Product Disrupts the Allosteric Regulation of Hsp70. Chem. Biol. 2015, 22, 87–97.
  21. Jones, A.M.; Westwood, I.M.; Osborne, J.D.; Matthews, T.P.; Cheeseman, M.; Rowlands, M.G.; Jeganathan, F.; Burke, R.; Lee, D.; Kadi, N.; et al. A fragment-based approach applied to a highly flexible target: Insights and challenges towards the inhibition of HSP70 isoforms. Sci. Rep. 2016, 6, 34701.
  22. Lim, S.; Cho, H.Y.; Kim, D.G.; Roh, Y.; Son, S.-Y.; Mushtaq, A.U.; Kim, M.; Bhattarai, D.; Sivaraman, A.; Lee, Y.; et al. Targeting the interaction of AIMP2-DX2 with HSP70 suppresses cancer development. Nat. Methods 2019, 16, 31–41.
  23. Osipiuk, J.; Walsh, M.A.; Freeman, B.C.; Morimoto, R.I.; Joachimiak, A. Structure of a new crystal form of human Hsp70 ATPase domain. Acta Crystallogr. Sect. D Biol. Crystallogr. 1999, 55, 1105–1107.
  24. Ravalin, M.; Theofilas, P.; Basu, K.; Opoku-Nsiah, K.A.; Assimon, V.A.; Medina-Cleghorn, D.; Chen, Y.-F.; Bohn, M.F.; Arkin, M.; Grinberg, L.T.; et al. Specificity for latent C termini links the E3 ubiquitin ligase CHIP to caspases. Nat. Methods 2019, 15, 786–794.
  25. Shida, M.; Arakawa, A.; Ishii, R.; Kishishita, S.; Takagi, T.; Kukimoto-Niino, M.; Sugano, S.; Tanaka, A.; Shirouzu, M.; Yokoyama, S. Direct inter-subdomain interactions switch between the closed and open forms of the Hsp70 nucleotide-binding domain in the nucleotide-free state. Acta Crystallogr. Sect. D Biol. Crystallogr. 2010, 66, 223–232.
  26. Sriram, M.; Osipiuk, J.; Freeman, B.; Morimoto, R.; Joachimiak, A. Human Hsp70 molecular chaperone binds two calcium ions within the ATPase domain. Structure 1997, 5, 403–414.
  27. Umehara, K.; Hoshikawa, M.; Tochio, N.; Tate, S.-I. Substrate Binding Switches the Conformation at the Lynchpin Site in the Substrate-Binding Domain of Human Hsp70 to Enable Allosteric Interdomain Communication. Molecules 2018, 23, 528.
  28. Zhang, P.; Leu, J.I.-J.; Murphy, M.E.; George, D.L.; Marmorstein, R. Crystal Structure of the Stress-Inducible Human Heat Shock Protein 70 Substrate-Binding Domain in Complex with Peptide Substrate. PLoS ONE 2014, 9, e103518.
  29. Murphy, M.E. The HSP70 family and cancer. Carcinogen 2013, 34, 1181–1188.
  30. Ha, J.-H.; McKay, D.B. Kinetics of Nucleotide-Induced Changes in the Tryptophan Fluorescence of the Molecular Chaperone Hsc70 and Its Subfragments Suggest the ATP-Induced Conformational Change Follows Initial ATP Binding. Biochemistry 1995, 34, 11635–11644.
  31. Palleros, D.R.; Raid, K.L.; Shi, L.; Welch, W.J.; Fink, A.L. ATP-induced protein Hsp70 complex dissociation requires K+ but not ATP hydrolysis. Nature 1993, 365, 664–666.
  32. Schmid, D.; Baici, A.; Gehring, H.; Christen, P. Kinetics of molecular chaperone action. Science 1994, 263, 971–973.
  33. Prasad, K.; Heuser, J.; Eisenberg, E.; Greene, L. Complex formation between clathrin and uncoating ATPase. J. Biol. Chem. 1994, 269, 6931–6939.
  34. Hartl, F.U. Molecular chaperones in cellular protein folding. Nature 1996, 381, 571–580.
  35. Laufen, T.; Mayer, M.P.; Beisel, C.; Klostermeier, D.; Mogk, A.; Reinstein, J.; Bukau, B. Mechanism of regulation of Hsp70 chaperones by DnaJ cochaperones. Proc. Natl. Acad. Sci. USA 1999, 96, 5452–5457.
  36. Mayer, M.P.; Bukau, B. Hsp70 chaperones: Cellular functions and molecular mechanism. Cell. Mol. Life Sci. 2005, 62, 670–684.
  37. Suzuki, H.; Noguchi, S.; Arakawa, H.; Tokida, T.; Hashimoto, M.; Satow, Y. Peptide-Binding Sites As Revealed by the Crystal Structures of the Human Hsp40 Hdj1 C-Terminal Domain in Complex with the Octapeptide from Human Hsp70. Biochemistry 2010, 49, 8577–8584.
  38. Jiang, Y.; Rossi, P.; Kalodimos, C.G. Structural basis for client recognition and activity of Hsp40 chaperones. Science 2019, 365, 1313–1319.
  39. Craig, E.; Huang, P.; Aron, R.; Andrew, A. The diverse roles of J-proteins, the obligate Hsp70 co-chaperone. Ergeb. der Physiol. 2006, 156.
  40. Hageman, J.; Kampinga, H.H. Computational analysis of the human HSPH/HSPA/DNAJ family and cloning of a human HSPH/HSPA/DNAJ expression library. Cell Stress Chaperones 2008, 14, 1–21.
  41. Diamant, S.; Goloubinoff, P. Temperature-Controlled Activity of DnaK−DnaJ−GrpE Chaperones: Protein-Folding Arrest and Recovery during and after Heat Shock Depends on the Substrate Protein and the GrpE Concentration†. Biochemistry 1998, 37, 9688–9694.
  42. Mitra, A.; Shevde, L.A.; Samant, R.S. Multi-faceted role of HSP40 in cancer. Clin. Exp. Metastasis 2009, 26, 559–567.
  43. Bracher, A.; Verghese, J. The nucleotide exchange factors of Hsp70 molecular chaperones. Front. Mol. Biosci. 2015, 2, 10.
  44. Brodsky, J.L.; Bracher, A. Nucleotide Exchange Factors for HSP70 Molecular Chaperones. In Networking of Chaperones by Co-Chaperones; Molecular Biology Intelligence Unit; Springer: New York, NY, USA, 2007; pp. 1–12.
  45. Li, Z.; Hartl, F.U.; Bracher, A. Structure and function of Hip, an attenuator of the Hsp70 chaperone cycle. Nat. Struct. Mol. Biol. 2013, 20, 929–935.
  46. Wang, A.M.; Miyata, Y.; Klinedinst, S.; Peng, H.-M.; Chua, J.P.; Komiyama, T.; Li, X.; Morishima, Y.; Merry, D.E.; Pratt, W.B.; et al. Activation of Hsp70 reduces neurotoxicity by promoting polyglutamine protein degradation. Nat. Methods 2012, 9, 112–118.
  47. Rousaki, A.; Miyata, Y.; Jinwal, U.K.; Dickey, C.A.; Gestwicki, J.E.; Zuiderweg, E.R.P. Allosteric Drugs: The Interaction of Antitumor Compound MKT-077 with Human Hsp70 Chaperones. J. Mol. Biol. 2011, 411, 614–632.
  48. Scheufler, C.; Brinker, A.; Bourenkov, G.; Pegoraro, S.; Moroder, L.; Bartunik, H.; Hartl, F.; Moarefi, I. Structure of TPR Domain–Peptide Complexes. Cell 2000, 101, 199–210.
  49. Smith, D.F.; Sullivan, W.P.; Marion, T.N.; Zaitsu, K.; Madden, B.; McCormick, D.J.; O Toft, D. Identification of a 60-kilodalton stress-related protein, p60, which interacts with hsp90 and hsp70. Mol. Cell. Biol. 1993, 13, 869–876.
  50. Honoré, B.; Leffers, H.; Madsen, P.; Rasmussen, H.H.; Vandekerckhove, J.; E Celis, J. Molecular cloning and expression of a transformation-sensitive human protein containing the TPR motif and sharing identity to the stress-inducible yeast protein STI1. J. Biol. Chem. 1992, 267, 8485–8491.
  51. Stankiewicz, M.; Nikolay, R.; Rybin, V.; Mayer, M.P. CHIP participates in protein triage decisions by preferentially ubiquitinating Hsp70-bound substrates. FEBS J. 2010, 277, 3353–3367.
  52. Höhfeld, J.; Cyr, D.M.; Patterson, C. From the cradle to the grave: Molecular chaperones that may choose between folding and degradation. EMBO Rep. 2001, 2, 885–890.
  53. Connell, P.; Ballinger, C.A.; Jiang, J.; Wu, Y.; Thompson, L.J.; Höhfeld, J.; Patterson, C. The co-chaperone CHIP regulates protein triage decisions mediated by heat-shock proteins. Nature 2000, 3, 93–96.
  54. Meacham, G.C.; Patterson, C.; Zhang, W.; Younger, J.M.; Cyr, D.M. The Hsc70 co-chaperone CHIP targets immature CFTR for proteasomal degradation. Nature 2000, 3, 100–105.
  55. Ballinger, C.A.; Connell, P.; Wu, Y.; Hu, Z.; Thompson, L.J.; Yin, L.-Y.; Patterson, C. Identification of CHIP, a Novel Tetratricopeptide Repeat-Containing Protein That Interacts with Heat Shock Proteins and Negatively Regulates Chaperone Functions. Mol. Cell. Biol. 1999, 19, 4535–4545.
  56. Zhang, H.; Amick, J.; Chakravarti, R.; Santarriaga, S.; Schlanger, S.; McGlone, C.; Dare, M.; Nix, J.C.; Scaglione, K.M.; Stuehr, D.J.; et al. A Bipartite Interaction between Hsp70 and CHIP Regulates Ubiquitination of Chaperoned Client Proteins. Structure 2015, 23, 472–482.
  57. Neckers, L.; Workman, P. Hsp90 molecular chaperone inhibitors: Are we there yet? Clin. Cancer Res. 2012, 18, 64–76.
  58. Schopf, F.H.; Biebl, M.M.; Buchner, J. The HSP90 chaperone machinery. Nat. Rev. Mol. Cell. Biol. 2017, 18, 345–360.
  59. Miyata, Y.; Nakamoto, H.; Neckers, L. The therapeutic target Hsp90 and cancer hallmarks. Curr. Pharm. Des. 2013, 19, 347–365.
  60. Rodina, A.; Wang, T.; Yan, P.; Gomes, E.D.; Dunphy, M.P.S.; Pillarsetty, N.; Koren, J.; Gerecitano, J.F.; Taldone, T.; Zong, H.; et al. The epichaperome is an integrated chaperome network that facilitates tumour survival. Nature 2016, 538, 397–401.
  61. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674.
  62. Hanahan, D.; Weinberg, R.A. The Hallmarks of Cancer. Cell 2000, 100, 57–70.
  63. Srivastava, P.K. Interaction of Heat Shock Proteins with Peptides and Antigen Presenting Cells: Chaperoning of the Innate and Adaptive Immune Responses. Annu. Rev. Immunol. 2002, 20, 395–425.
  64. Blachere, N.E.; Li, Z.; Chandawarkar, R.Y.; Suto, R.; Jaikaria, N.S.; Basu, S.; Udono, H.; Srivastava, P.K. Heat Shock Protein–Peptide Complexes, Reconstituted In Vitro, Elicit Peptide-specific Cytotoxic T Lymphocyte Response and Tumor Immunity. J. Exp. Med. 1997, 186, 1315–1322.
  65. Multhoff, G.; Botzler, C.; Wiesnet, M.; Müller, E.; Meier, T.; Wilmanns, W.; Issels, R.D. A stress-inducible 72-kDa heat-shock protein (HSP72) is expressed on the surface of human tumor cells, but not on normal cells. Int. J. Cancer 1995, 61, 272–279.
  66. Gastpar, R.; Gehrmann, M.; Bausero, M.A.; Asea, A.; Gross, C.; Schroeder, J.A.; Multhoff, G. Heat shock protein 70 surface-positive tumor exosomes stimulate migratory and cytolytic activity of natural killer cells. Cancer Res. 2005, 65, 5238–5247.
  67. Stangl, S.; Gross, C.; Pockley, A.G.; Asea, A.; Multhoff, G. Influence of Hsp70 and HLA-E on the killing of leukemic blasts by cytokine/Hsp70 peptide-activated human natural killer (NK) cells. Cell Stress Chaperones 2008, 13, 221–230.
  68. Multhoff, G.; Pfister, K.; Gehrmann, M.; Hantschel, M.; Gross, C.; Hafner, M.; Hiddemann, W. A 14-mer Hsp70 peptide stimulates natural killer (NK) cell activity. Cell Stress Chaperons 2001, 6, 337–344.
  69. Multhoff, G.; Botzler, C.; Jennen, L.; Schmidt, J.; Ellwart, J.; Issels, R. Heat shock protein 72 on tumor cells: A recognition structure for natural killer cells. J. Immunol. 1997, 158, 4341–4350.
  70. Multhoff, G. Heat shock protein 70 (Hsp70) stimulates proliferation and cytolytic activity of natural killer cells. Exp. Hematol. 1999, 27, 1627–1636.
  71. Specht, H.M.; Ahrens, N.; Blankenstein, C.; Duell, T.; Fietkau, R.; Gaipl, U.S.; Günther, C.; Gunther, S.; Habl, G.; Hautmann, H.; et al. Heat Shock Protein 70 (Hsp70) Peptide Activated Natural Killer (NK) Cells for the Treatment of Patients with Non-Small Cell Lung Cancer (NSCLC) after Radiochemotherapy (RCTx) – From Preclinical Studies to a Clinical Phase II Trial. Front. Immunol. 2015, 6.
  72. Hromadníková, I.; Li, S.; Kotlabova, K.; Dickinson, A.M. Influence of In Vitro IL-2 or IL-15 Alone or in Combination with Hsp 70 Derived 14-Mer Peptide (TKD) on the Expression of NK Cell Activatory and Inhibitory Receptors on Peripheral Blood T Cells, B Cells and NKT Cells. PLoS ONE 2016, 11, e0151535.
  73. Rückert, M.; Deloch, L.; Fietkau, R.; Frey, B.; Hecht, M.; Gaipl, U.S. Immune modulatory effects of radiotherapy as basis for well-reasoned radioimmunotherapies. Strahlenther. und Onkol. 2018, 194, 509–519.
  74. Günther, S.; Ostheimer, C.; Stangl, S.; Specht, H.M.; Mozes, P.; Jesinghaus, M.; Vordermark, D.; Combs, S.E.; Peltz, F.; Jung, M.P.; et al. Correlation of Hsp70 Serum Levels with Gross Tumor Volume and Composition of Lymphocyte Subpopulations in Patients with Squamous Cell and Adeno Non-Small Cell Lung Cancer. Front. Immunol. 2015, 6, 128.
  75. Ostheimer, C.; Gunther, S.; Bache, M.; Vordermark, D.; Multhoff, G. Dynamics of Heat Shock Protein 70 Serum Levels As a Predictor of Clinical Response in Non-Small-Cell Lung Cancer and Correlation with the Hypoxia-Related Marker Osteopontin. Front. Immunol. 2017, 8.
  76. A Daniels, G.; Sanchez-Perez, L.; Diaz, R.M.; Kottke, T.; Thompson, J.; Lai, M.; Gough, M.; Karim, M.; Bushell, A.; Chong, H.; et al. A simple method to cure established tumors by inflammatory killing of normal cells. Nat. Biotechnol. 2004, 22, 1125–1132.
  77. Breloer, M.; Fleischer, B.; Von Bonin, A. In vivo and in vitro activation of T cells after administration of Ag-negative heat shock proteins. J. Immunol. 1999, 162.
  78. Rothammer, A.; Sage, E.K.; Werner, C.; Combs, S.E.; Multhoff, G. Increased heat shock protein 70 (Hsp70) serum levels and low NK cell counts after radiotherapy - potential markers for predicting breast cancer recurrence? Radiat. Oncol. 2019, 14, 78.
  79. Pfister, K.; Radons, J.; Busch, R.; Tidball, J.G.; Pfeifer, M.; Freitag, L.; Feldmann, H.-J.; Milani, V.; Issels, R.; Multhoff, G. Patient survival by Hsp70 membrane phenotype. Cancer 2007, 110, 926–935.
  80. Zininga, T.; Ramatsui, L.; Shonhai, A. Heat Shock Proteins as Immunomodulants. Molecules 2018, 23, 2846.
  81. Wachstein, J.; Tischer, S.; Figueiredo, C.; Limbourg, A.; Falk, C.; Immenschuh, S.; Blasczyk, R.; Eiz-Vesper, B. HSP70 Enhances Immunosuppressive Function of CD4+CD25+FoxP3+ T Regulatory Cells and Cytotoxicity in CD4+CD25− T Cells. PLoS ONE 2012, 7, e51747.
  82. Asea, A.; Kraeft, S.-K.; Kurt-Jones, E.A.; Stevenson, M.A.; Chen, L.B.; Finberg, R.W.; Koo, G.C.; Calderwood, S. HSP70 stimulates cytokine production through a CD14-dependant pathway, demonstrating its dual role as a chaperone and cytokine. Nat. Med. 2000, 6, 435–442.
  83. Asea, A. Novel Signal Transduction Pathway Utilized by Extracellular HSP70. Role of Toll-Like Receptor (TLR) 2 and TLR4. J. Biol. Chem. 2002, 277, 15028–15034.
  84. Asea, A. Initiation of the Immune Response by Extracellular Hsp72: Chaperokine Activity of Hsp72. Curr. Immunol. Rev. 2006, 2, 209–215.
  85. Asea, A. Chaperokine-induced signal transduction pathways. Exerc. Immunol. Rev. 2003, 9, 25–33.
  86. Martine, P.; Chevriaux, A.; Derangère, V.; Apetoh, L.; Garrido, C.; Ghiringhelli, F.; Rebe, C. HSP70 is a negative regulator of NLRP3 inflammasome activation. Cell Death Dis. 2019, 10, 256.
  87. Moossavi, M.; Parsamanesh, N.; Bahrami, A.; Atkin, S.; Sahebkar, A. Role of the NLRP3 inflammasome in cancer. Mol. Cancer 2018, 17, 158.
  88. Yashin, D.V.; Romanova, E.A.; Ivanova, O.K.; Sashchenko, L.P. The Tag7–Hsp70 cytotoxic complex induces tumor cell necroptosis via permeabilisation of lysosomes and mitochondria. Biochimie 2016, 123, 32–36.
  89. Balogi, Z.; Multhoff, G.; Jensen, T.K.; Lloyd-Evans, E.; Yamashima, T.; Jäättelä, M.; Harwood, J.L.; Vígh, L.; Multhoff, G. Hsp70 interactions with membrane lipids regulate cellular functions in health and disease. Prog. Lipid Res. 2019, 74, 18–30.
  90. Sashchenko, L.P.; Dukhanina, E.A.; Yashin, D.V.; Shatalov, Y.V.; Romanova, E.A.; Korobko, E.V.; Demin, A.V.; Lukyanova, T.I.; Kabanova, O.D.; Khaidukov, S.V.; et al. Peptidoglycan Recognition Protein Tag7 Forms a Cytotoxic Complex with Heat Shock Protein 70 in Solution and in Lymphocytes. J. Biol. Chem. 2003, 279, 2117–2124.
  91. Chalmin, F.; Ladoire, S.; Mignot, G.; Vincent, J.; Bruchard, M.; Remy-Martin, J.-P.; Boireau, W.; Rouleau, A.; Simon, B.; Lanneau, D.; et al. Membrane-associated Hsp72 from tumor-derived exosomes mediates STAT3-dependent immunosuppressive function of mouse and human myeloid-derived suppressor cells. J. Clin. Investig. 2010, 120, 457–471.
  92. Bausero, M.A.; Gastpar, R.; Multhoff, G.; Asea, A. Alternative Mechanism by which IFN-γ Enhances Tumor Recognition: Active Release of Heat Shock Protein 721. J. Immunol. 2005, 175, 2900–2912.
  93. De Vita, F.; Orditura, M.; Galizia, G.; Romano, C.; Infusino, S.; Auriemma, A.; Lieto, E.; Catalano, G. Serum interleukin-10 levels in patients with advanced gastrointestinal malignancies. Cancer 1999, 86, 1936–1943.
  94. Lebrecht, A.; Hefler, L.; Tempfer, C.B.; Koelbl, H. Serum Cytokine Concentrations in Patients with Cervical Cancer: Interleukin-4, Interferon-γ, and Monocyte Chemoattractant Protein-1. Gynecol. Oncol. 2001, 83, 170–171.
  95. Udono, H.; Srivastava, P.K. Comparison of tumor-specific immunogenicities of stress-induced proteins gp96, hsp90, and hsp70. J. Immunol. 1994, 152, 5398–5403.
  96. Weng, D.; Calderwood, S.K.; Gong, J. Preparation of a heat-shock protein 70-based vaccine from DC-tumor fusion cells. Breast Cancer 2011, 787, 255–265.
  97. Hegde, P.S.; Chen, D.S. Top 10 Challenges in Cancer Immunotherapy. Immunity 2020, 52, 17–35.
  98. Calderwood, S.K.; Murshid, A. Molecular Chaperone Accumulation in Cancer and Decrease in Alzheimer’s Disease: The Potential Roles of HSF1. Front. Neurosci. 2017, 11, 192.
  99. Mosser, D.D.; Caron, A.W.; Bourget, L.; Denis-Larose, C.; Massie, B. Role of the human heat shock protein hsp70 in protection against stress-induced apoptosis. Mol. Cell. Biol. 1997, 17, 5317–5327.
  100. Nylandsted, J.; Rohde, M.; Brand, K.; Bastholm, L.; Elling, F.; Jäättelä, 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. USA 2000, 97, 7871–7876.
  101. Gao, Y.; Han, C.; Huang, H.; Xin, Y.; Xu, Y.; Luo, L.; Yin, Z. Heat shock protein 70 together with its co-chaperone CHIP inhibits TNF-α induced apoptosis by promoting proteasomal degradation of apoptosis signal-regulating kinase1. Apoptosis 2010, 15, 822–833.
  102. Tournier, C. Requirement of JNK for Stress- Induced Activation of the Cytochrome c-Mediated Death Pathway. Science 2000, 288, 870–874.
  103. Dhanasekaran, D.N.; Reddy, E.P. JNK signaling in apoptosis. Oncogene 2008, 27, 6245–6251.
  104. Sax, J.K.; Fei, P.; Murphy, M.E.; Bernhard, E.; Korsmeyer, S.J.; El-Deiry, W.S. BID regulation by p53 contributes to chemosensitivity. Nature 2002, 4, 842–849.
  105. Ravagnan, L.; Gurbuxani, S.; Susin, S.A.; Maisse, C.; Daugas, E.; Zamzami, N.; Mak, T.; Jäättelä, M.; Penninger, J.M.; Garrido, C.; et al. Heat-shock protein 70 antagonizes apoptosis-inducing factor. Nature 2001, 3, 839–843.
  106. Guo, F.; Sigua, C.; Bali, P.; George, P.; Fiskus, W.; Scuto, A.; Annavarapu, S.; Mouttaki, A.; Sondarva, G.; Wei, S.; et al. Mechanistic role of heat shock protein 70 in Bcr-Abl–mediated resistance to apoptosis in human acute leukemia cells. Blood 2005, 105, 1246–1255.
  107. Gabai, V.; Mabuchi, K.; Mosser, D.D.; Sherman, M. 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.
  108. Nylandsted, J.; Gyrd-Hansen, M.; Danielewicz, A.; Fehrenbacher, N.; Lademann, U.; Høyer-Hansen, M.; Weber, E.; Multhoff, G.; Rohde, M.; Jäättelä, M. Heat Shock Protein 70 Promotes Cell Survival by Inhibiting Lysosomal Membrane Permeabilization. J. Exp. Med. 2004, 200, 425–435.
  109. Bivik, C.; Rosdahl, I.; Öllinger, K. Hsp70 protects against UVB induced apoptosis by preventing release of cathepsins and cytochrome c in human melanocytes. Carcinogen 2006, 28, 537–544.
  110. Daugaard, M.; Kirkegaard-Sorensen, T.; Ostenfeld, M.S.; Aaboe, M.; Hoyer-Hansen, M.; Orntoft, T.F.; Rohde, M.; Jäättelä, M. Lens Epithelium-Derived Growth Factor Is an Hsp70-2 Regulated Guardian of Lysosomal Stability in Human Cancer. Cancer Res. 2007, 67, 2559–2567.
  111. Park, S.-H.; Baek, K.-H.; Shin, I.; Shin, I. Subcellular Hsp70 Inhibitors Promote Cancer Cell Death via Different Mechanisms. Cell Chem. Biol. 2018, 25, 1242–1254.e8.
  112. Yaglom, J.A.; Ekhterae, D.; Gabai, V.; Sherman, M. Regulation of Necrosis of H9c2 Myogenic Cells upon Transient Energy Deprivation. J. Biol. Chem. 2003, 278, 50483–50496.
  113. Karsch-Bluman, A.; Feiglin, A.; Arbib, E.; Stern, T.; Shoval, H.; Schwob, O.; Berger, M.; Benny, O. Tissue necrosis and its role in cancer progression. Oncogene 2018, 38, 1920–1935.
  114. Stankiewicz, A.R.; Lachapelle, G.; Foo, C.P.Z.; Radicioni, S.M.; Mosser, D.D. Hsp70 Inhibits Heat-induced Apoptosis Upstream of Mitochondria by Preventing Bax Translocation. J. Biol. Chem. 2005, 280, 38729–38739.
  115. Beere, H.M.; Wolf, B.B.; Cain, K.; Mosser, D.D.; Mahboubi, A.; Kuwana, T.; Tailor, P.; Morimoto, R.I.; Cohen, G.M.; Green, D.R. Heat-shock protein 70 inhibits apoptosis by preventing recruitment of procaspase-9 to the Apaf-1 apoptosome. Nature 2000, 2, 469–475.
  116. Yaglom, J.A.; Gabai, V.; Sherman, M. High Levels of Heat Shock Protein Hsp72 in Cancer Cells Suppress Default Senescence Pathways. Cancer Res. 2007, 67, 2373–2381.
  117. Gabai, V.; Yaglom, J.A.; Waldman, T.; Sherman, M.Y. Heat Shock Protein Hsp72 Controls Oncogene-Induced Senescence Pathways in Cancer Cells. Mol. Cell. Biol. 2008, 29, 559–569.
  118. Meng, L.; Hunt, C.; Yaglom, J.A.; Gabai, V.; Sherman, M.Y. Heat shock protein Hsp72 plays an essential role in Her2-induced mammary tumorigenesis. Oncogene 2011, 30, 2836–2845.
  119. Colvin, T.A.; Gabai, V.; Gong, J.; Calderwood, S.K.; Li, H.; Gummuluru, S.; Matchuk, O.; Smirnova, S.G.; Orlova, N.V.; Zamulaeva, I.A.; et al. Hsp70-Bag3 interactions regulate cancer-related signaling networks. Cancer Res. 2014, 74, 4731–4740.
  120. Meng, L.; Gabai, V.; Sherman, M.Y. Heat-shock transcription factor HSF1 has a critical role in human epidermal growth factor receptor-2-induced cellular transformation and tumorigenesis. Oncogene 2010, 29, 5204–5213.
  121. Oda, T.; Sekimoto, T.; Kurashima, K.; Fujimoto, M.; Nakai, A.; Yamashita, T. Acute HSF1 depletion induces cellular senescence through the MDM2-p53-p21 pathway in human diploid fibroblasts. J. Cell Sci. 2018, 131, jcs210724.
  122. Boysen, M.; Kityk, R.; Mayer, M.P. Hsp70- and Hsp90-Mediated Regulation of the Conformation of p53 DNA Binding Domain and p53 Cancer Variants. Mol. Cell 2019, 74, 831–843.e4.
  123. Dahiya, V.; Agam, G.; Lawatscheck, J.; Rutz, D.A.; Lamb, D.C.; Buchner, J. Coordinated Conformational Processing of the Tumor Suppressor Protein p53 by the Hsp70 and Hsp90 Chaperone Machineries. Mol. Cell 2019, 74, 816–830.e7.
  124. Han, J.M.; Park, B.-J.; Park, S.G.; Oh, Y.S.; Choi, S.J.; Lee, S.W.; Hwang, S.-K.; Chang, S.-H.; Cho, M.-H.; Kim, S. AIMP2/p38, the scaffold for the multi-tRNA synthetase complex, responds to genotoxic stresses via p53. Proc. Natl. Acad. Sci. USA 2008, 105, 11206–11211.
  125. Huang, L.E.; Bunn, H.F. Hypoxia-inducible Factor and Its Biomedical Relevance. J. Biol. Chem. 2003, 278, 19575–19578.
  126. Ryan, H.E.; Lo, J.; Johnson, R.S. HIF-1 alpha is required for solid tumor formation and embryonic vascularization. EMBO J. 1998, 17, 3005–3015.
  127. Gordan, J.D.; Simon, M.C. Hypoxia-inducible factors: Central regulators of the tumor phenotype. Curr. Opin. Genet. Dev. 2007, 17, 71–77.
  128. Zhou, J.; Schmid, T.; Frank, R.; Brüne, B. PI3K/Akt Is Required for Heat Shock Proteins to Protect Hypoxia-inducible Factor 1 from pVHL-independent Degradation. J. Biol. Chem. 2004, 279, 13506–13513.
  129. Gabai, V.; Meng, L.; Kim, G.; Mills, T.A.; Benjamin, I.J.; Sherman, M.Y. Heat Shock Transcription Factor Hsf1 Is Involved in Tumor Progression via Regulation of Hypoxia-Inducible Factor 1 and RNA-Binding Protein HuR. Mol. Cell. Biol. 2012, 32, 929–940.
  130. De Silanes, I.L.; Fan, J.; Yang, X.; Zonderman, A.B.; Potapova, O.; Pizer, E.S.; Gorospe, M. Role of the RNA-binding protein HuR in colon carcinogenesis. Oncogene 2003, 22, 7146–7154.
  131. Kim, T.-K.; Na, H.-J.; Lee, W.R.; Jeoung, M.H.; Lee, S. Heat shock protein 70-1A is a novel angiogenic regulator. Biochem. Biophys. Res. Commun. 2016, 469, 222–228.
  132. Park, S.L.; Chung, T.-W.; Kim, S.; Hwang, B.; Kim, J.M.; Lee, H.M.; Cha, H.-J.; Seo, Y.; Choe, S.Y.; Ha, K.-T.; et al. HSP70-1 is required for interleukin-5-induced angiogenic responses through eNOS pathway. Sci. Rep. 2017, 7, 44687.
  133. Kluger, H.M.; Goel, H.L.; Breen, M.; Zhang, J.; Das, I.; Aznavoorian-Cheshire, S.; Greenberg, N.M.; Elgavish, A.; Languino, L. Using a Xenograft Model of Human Breast Cancer Metastasis to Find Genes Associated with Clinically Aggressive Disease. Cancer Res. 2005, 65, 5578–5587.
  134. Teng, Y.; Ngoka, L.; Mei, Y.; Lesoon, L.; Cowell, J. HSP90 and HSP70 Proteins Are Essential for Stabilization and Activation of WASF3 Metastasis-promoting Protein*. J. Biol. Chem. 2012, 287, 10051–10059.
  135. Frisch, S.M.; Francis, H. Disruption of epithelial cell-matrix interactions induces apoptosis. J. Cell Biol. 1994, 124, 619–626.
  136. Kasioumi, P.; Vrazeli, P.; Vezyraki, P.; Zerikiotis, S.; Katsouras, C.; Damalas, A.; Angelidis, C. Hsp70 (HSP70A1A) downregulation enhances the metastatic ability of cancer cells. Int. J. Oncol. 2018, 54, 821–832.
  137. Mao, H.; Li, F.; Ruchalski, K.; Mosser, D.D.; Schwartz, J.; Wang, Y.; Borkan, S.C.; Stortchevoi, A.; Varshney, U.; Rajbhandary, U.L. hsp72 Inhibits Focal Adhesion Kinase Degradation in ATP-depleted Renal Epithelial Cells. J. Biol. Chem. 2003, 278, 18214–18220.
  138. Jang, K.W.; Lee, J.E.; Kim, S.Y.; Kang, M.-W.; Na, M.H.; Lee, C.S.; Song, K.S.; Lim, S.P. The C-terminus of Hsp70-Interacting Protein Promotes Met Receptor Degradation. J. Thorac. Oncol. 2011, 6, 679–687.
  139. Iii, J.K.; Jinwal, U.K.; Jin, Y.; O’Leary, J.; Jones, J.; Johnson, A.G.; Blair, L.J.; Abisambra, J.F.; Chang, L.; Miyata, Y.; et al. Facilitating Akt Clearance via Manipulation of Hsp70 Activity and Levels*. J. Biol. Chem. 2009, 285, 2498–2505.
  140. Li, H.; Li, Y.; Liu, D.; Sun, H.; Su, N.; Yang, F.; Liu, J. Extracellular HSP70/HSP70-PCs Promote Epithelial-Mesenchymal Transition of Hepatocarcinoma Cells. PLoS ONE 2013, 8, e84759.
  141. Komarova, E.; Marchenko, L.; Zhakhov, A.; Nikotina, A.; Aksenov, N.; Suezov, R.; Ischenko, A.; Margulis, B.; Guzhova, I.V. Extracellular Hsp70 Reduces the Pro-Tumor Capacity of Monocytes/Macrophages Co-Cultivated with Cancer Cells. Int. J. Mol. Sci. 2019, 21, 59.
  142. Mao, H.; Li, Z.; Zhou, Y.; Li, Z.; Zhuang, S.; An, X.; Zhang, B.; Chen, W.; Nie, J.; Wang, Z.; et al. HSP72 attenuates renal tubular cell apoptosis and interstitial fibrosis in obstructive nephropathy. Am. J. Physiol. Physiol. 2008, 295, F202–F214.
  143. Yun, C.-H.; Yoon, S.-Y.; Nguyen, T.T.; Cho, H.-Y.; Kim, T.-H.; Kim, S.-T.; Kim, B.-C.; Hong, Y.-S.; Kim, S.-J.; Lee, H.J. Geldanamycin inhibits TGF-β signaling through induction of Hsp70. Arch. Biochem. Biophys. 2010, 495, 8–13.
  144. Li, Y.; Kang, X.; Wang, Q. HSP70 decreases receptor-dependent phosphorylation of Smad2 and blocks TGF-β-induced epithelial-mesenchymal transition. J. Genet. Genom. 2011, 38, 111–116.
  145. Liu, J.; Bao, J.; Hao, J.; Peng, Y.; Hong, F. HSP70 inhibits high glucose-induced Smad3 activation and attenuates epithelial-to-mesenchymal transition of peritoneal mesothelial cells. Mol. Med. Rep. 2014, 10, 1089–1095.
  146. Jeon, H.-M.; Lee, J. MET: Roles in epithelial-mesenchymal transition and cancer stemness. Ann. Transl. Med. 2017, 5, 5.
  147. Chao, A.; Lai, C.-H.; Tsai, C.-L.; Hsueh, S.; Hsueh, C.; Lin, C.-Y.; Chou, H.-H.; Lin, Y.-J.; Chen, H.-W.; Chang, T.-C.; et al. Tumor Stress-Induced Phosphoprotein1 (STIP1) as a Prognostic Biomarker in Ovarian Cancer. PLoS ONE 2013, 8, e57084.
  148. Da Fonseca, A.C.C.; Wang, H.; Fan, H.; Chen, X.; Zhang, I.; Zhang, L.; Lima, F.R.S.; Badie, B. Increased expression of stress inducible protein 1 in glioma-associated microglia/macrophages. J. Neuroimmunol. 2014, 274, 71–77.
  149. Willmer, T.; Contu, L.; Blatch, G.; Edkins, A.L. Knockdown of Hop downregulates RhoC expression, and decreases pseudopodia formation and migration in cancer cell lines. Cancer Lett. 2013, 328, 252–260.
  150. Walsh, N.; Larkin, A.; Swan, N.; Conlon, K.; Dowling, P.; McDermott, R.; Clynes, M. RNAi knockdown of Hop (Hsp70/Hsp90 organising protein) decreases invasion via MMP-2 down regulation. Cancer Lett. 2011, 306, 180–189.
  151. Suzuki, M.; Iwasaki, M.; Sugio, A.; Hishiya, A.; Tanaka, R.; Endo, T.; Takayama, S.; Saito, T. BAG3 (BCL2-associated athanogene 3) interacts with MMP-2 to positively regulate invasion by ovarian carcinoma cells. Cancer Lett. 2011, 303, 65–71.
  152. Lee, K.-J.; Kim, Y.M.; Kim, D.Y.; Jeoung, O.; Han, K.; Lee, S.-T.; Lee, Y.-S.; Park, K.H.; Park, J.H.; Kim, D.J.; et al. Release of heat shock protein 70 (Hsp70) and the effects of extracellular Hsp70 on matric metalloproteinase-9 expression in human monocytic U937 cells. Exp. Mol. Med. 2006, 38, 364–374.
  153. Chetty, C.; Vanamala, S.K.; Gondi, C.; Dinh, D.H.; Gujrati, M.; Rao, J.S. MMP-9 induces CD44 cleavage and CD44 mediated cell migration in glioblastoma xenograft cells. Cell. Signal. 2011, 24, 549–559.
  154. Sossey-Alaoui, K.; Li, X.; Ranalli, T.A.; Cowell, J. WAVE3-mediated Cell Migration and Lamellipodia Formation Are Regulated Downstream of Phosphatidylinositol 3-Kinase*. J. Biol. Chem. 2005, 280, 21748–21755.
  155. Teng, Y.; Ren, M.Q.; Cheney, R.; Sharma, S.; Cowell, J. Inactivation of the WASF3 gene in prostate cancer cells leads to suppression of tumorigenicity and metastases. Br. J. Cancer 2010, 103, 1066–1075.
  156. Sossey-Alaoui, K.; Safina, A.; Li, X.; Vaughan, M.M.; Hicks, D.G.; Bakin, A.; Cowell, J. Down-Regulation of WAVE3, a Metastasis Promoter Gene, Inhibits Invasion and Metastasis of Breast Cancer Cells. Am. J. Pathol. 2007, 170, 2112–2121.
  157. Breuninger, S.; Stangl, S.; Werner, C.; Sievert, W.; Lobinger, D.; Foulds, G.; Wagner, S.; Pickhard, A.; Piontek, G.; Kokowski, K.; et al. Membrane Hsp70—A Novel Target for the Isolation of Circulating Tumor Cells After Epithelial-to-Mesenchymal Transition. Front. Oncol. 2018, 8, 497.
  158. Farkas, B.; Hantschel, M.; Magyarlaki, M.; Becker, B.; Scherer, K.; Landthaler, M.; Pfister, K.; Gehrmann, M.; Gross, C.; Mackensen, A.; et al. Heat shock protein 70 membrane expression and melanoma-associated marker phenotype in primary and metastatic melanoma. Melanoma Res. 2003, 13, 147–152.
  159. Botzler, C.; Schmidt, J.; Luz, A.; Jennen, L.; Issels, R.; Multhoff, G. Differential Hsp70 plasma-membrane expression on primary human tumors and metastases in mice with severe combined immunodeficiency. Int. J. Cancer 1998, 77, 942–948.
  160. Jolly, M.K.; Boareto, M.; Huang, B.; Jia, D.; Lu, M.; Ben-Jacob, E.; Onuchic, J.N.; Levine, H. Implications of the Hybrid Epithelial/Mesenchymal Phenotype in Metastasis. Front. Oncol. 2015, 5, 155.
  161. Stangl, S.; Gehrmann, M.; Riegger, J.; Kuhs, K.; Riederer, I.; Sievert, W.; Hube, K.; Mocikat, R.; Dressel, R.; Kremmer, E.; et al. Targeting membrane heat-shock protein 70 (Hsp70) on tumors by cmHsp70.1 antibody. Proc. Natl. Acad. Sci. USA USA 2010, 108, 733–738.
  162. Williamson, D.; Borgognoni, J.; Clay, A.; Daniels, Z.; Dokurno, P.; Drysdale, M.J.; Foloppe, N.; Francis, G.L.; Graham, C.J.; Howes, R.; et al. Novel Adenosine-Derived Inhibitors of 70 kDa Heat Shock Protein, Discovered Through Structure-Based Design†. J. Med. Chem. 2009, 52, 1510–1513.
  163. Macias, A.; Williamson, D.; Allen, N.; Borgognoni, J.; Clay, A.; Daniels, Z.; Dokurno, P.; Drysdale, M.J.; Francis, G.L.; Graham, C.J.; et al. Adenosine-Derived Inhibitors of 78 kDa Glucose Regulated Protein (Grp78) ATPase: Insights into Isoform Selectivity. J. Med. Chem. 2011, 54, 4034–4041.
  164. Li, X.; Colvin, T.; Rauch, J.N.; Acosta-Alvear, D.; Kampmann, M.; Dunyak, B.; Hann, B.; Aftab, B.T.; Murnane, M.; Cho, M.; et al. Validation of the Hsp70-Bag3 protein-protein interaction as a potential therapeutic target in cancer. Mol. Cancer Ther. 2015, 14, 642–648.
  165. Yaglom, J.A.; Wang, Y.; Li, A.; Li, Z.; Monti, S.; Alexandrov, I.; Lu, X.; Sherman, M.Y. Cancer cell responses to Hsp70 inhibitor JG-98: Comparison with Hsp90 inhibitors and finding synergistic drug combinations. Sci. Rep. 2018, 8, 3010.
  166. Sherman, M.; Gabai, V. Hsp70 in cancer: Back to the future. Oncogene 2014, 34, 4153–4161.
  167. Koren, J.; Miyata, Y.; Kiray, J.; O’Leary, J.C.; Nguyen, L.; Guo, J.; Blair, L.J.; Li, X.; Jinwal, U.K.; Cheng, J.Q.; et al. Correction: Rhodacyanine Derivative Selectively Targets Cancer Cells and Overcomes Tamoxifen Resistance. PLoS ONE 2012, 7.
  168. Massey, A.J.; Williamson, D.; Browne, H.; Murray, J.B.; Dokurno, P.; Shaw, T.; Macias, A.; Daniels, Z.; Geoffroy, S.; Dopson, M.; et al. A novel, small molecule inhibitor of Hsc70/Hsp70 potentiates Hsp90 inhibitor induced apoptosis in HCT116 colon carcinoma cells. Cancer Chemother. Pharmacol. 2009, 66, 535–545.
  169. Deocaris, C.C.; Widodo, N.; Shrestha, B.G.; Kaur, K.; Ohtaka, M.; Yamasaki, K.; Kaul, S.C.; Wadhwa, R. Mortalin sensitizes human cancer cells to MKT-077-induced senescence. Cancer Lett. 2007, 252, 259–269.
  170. Leu, J.I.-J.; Pimkina, J.; Frank, A.; Murphy, M.E.; George, D.L. A Small Molecule Inhibitor of Inducible Heat Shock Protein 70. Mol. Cell 2009, 36, 15–27.
  171. Schmitt, E.; Maingret, L.; Puig, P.-E.; Rérole, A.-L.; Hammann, A.; Ghiringhelli, F.; Solary, E.; Kroemer, G.; Garrido, C. Heat Shock Protein 70 Neutralization Exerts Potent Antitumor Effects in Animal Models of Colon Cancer and Melanoma. Cancer Res. 2006, 66, 4191–4197.
  172. Ernst, K.; Liebscher, M.; Mathea, S.; Granzhan, A.; Schmid, J.; Popoff, M.R.; Ihmels, H.; Barth, H.; Schiene-Fischer, C. A novel Hsp70 inhibitor prevents cell intoxication with the actin ADP-ribosylating Clostridium perfringens iota toxin. Sci. Rep. 2016, 6, 20301.
  173. Rérole, A.-L.; Gobbo, J.; De Thonel, A.; Schmitt, E.; De Barros, J.P.P.; Hammann, A.; Lanneau, D.; Fourmaux, E.; Demidov, O.; Micheau, O.; et al. Peptides and Aptamers Targeting HSP70: A Novel Approach for Anticancer Chemotherapy. Cancer Res. 2011, 71, 484–495.
  174. Rodina, A.; Patel, P.D.; Kang, Y.; Patel, Y.; Baaklini, I.; Wong, M.J.; Taldone, T.; Yan, P.; Yang, C.; Maharaj, R.; et al. Identification of an Allosteric Pocket on Human Hsp70 Reveals a Mode of Inhibition of This Therapeutically Important Protein. Chem. Biol. 2013, 20, 1469–1480.
  175. Williams, D.R.; Park, S.; Shin, I.; Ko, S.-K.; Lee, M.-R. An Apoptosis-Inducing Small Molecule That Binds to Heat Shock Protein 70. Angew. Chem. Int. Ed. 2008, 47, 7466–7469.
  176. Ko, S.-K.; Kim, J.; Na, D.C.; Park, S.; Park, S.-H.; Hyun, J.Y.; Baek, K.-H.; Kim, N.D.; Kim, N.-K.; Park, Y.N.; et al. A Small Molecule Inhibitor of ATPase Activity of HSP70 Induces Apoptosis and Has Antitumor Activities. Chem. Biol. 2015, 22, 391–403.
  177. Braunstein, M.J.; Scott, S.S.; Scott, C.M.; Behrman, S.; Walter, P.; Wipf, P.; Coplan, J.D.; Chrico, W.; Joseph, D.; Brodsky, J.L.; et al. Antimyeloma Effects of the Heat Shock Protein 70 Molecular Chaperone Inhibitor MAL3-101. J. Oncol. 2011, 2011, 1–11.
  178. Fewell, S.W.; Smith, C.M.; Lyon, M.A.; Dumitrescu, T.P.; Wipf, P.; Day, B.W.; Brodsky, J.L. Small Molecule Modulators of Endogenous and Co-chaperone-stimulated Hsp70 ATPase Activity. J. Biol. Chem. 2004, 279, 51131–51140.
  179. Huryn, D.; Brodsky, J.L.; Brummond, K.M.; Chambers, P.G.; Eyer, B.; Ireland, A.W.; Kawasumi, M.; Laporte, M.G.; Lloyd, K.; Manteau, B.; et al. Chemical methodology as a source of small-molecule checkpoint inhibitors and heat shock protein 70 (Hsp70) modulators. Proc. Natl. Acad. Sci. USA USA 2011, 108, 6757–6762.
  180. Phillips, P.; Sangwan, V.; Borja-Cacho, D.; Dudeja, V.; Vickers, S.; Saluja, A.K. Myricetin induces pancreatic cancer cell death via the induction of apoptosis and inhibition of the phosphatidylinositol 3-kinase (PI3K) signaling pathway. Cancer Lett. 2011, 308, 181–188.
  181. Howe, M.K.; Bodoor, K.; Carlson, D.A.; Hughes, P.F.; Alwarawrah, Y.; Loiselle, D.R.; Jaeger, A.M.; Darr, D.B.; Jordan, J.L.; Hunter, L.M.; et al. Identification of an allosteric small-molecule inhibitor selective for the inducible form of heat shock protein 70. Chem. Biol. 2014, 21, 1648–1659.
  182. Phillips, P.A.; Dudeja, V.; McCarroll, J.A.; Borja-Cacho, D.; Dawra, R.K.; Grizzle, W.E.; Vickers, S.M.; Saluja, A.K. Triptolide Induces Pancreatic Cancer Cell Death via Inhibition of Heat Shock Protein 70. Cancer Res. 2007, 67, 9407–9416.
  183. Banerjee, S.; Thayanithy, V.; Sangwan, V.; MacKenzie, T.N.; Saluja, A.K.; Subramanian, S. Minnelide reduces tumor burden in preclinical models of osteosarcoma. Cancer Lett. 2013, 335, 412–420.
  184. Antonoff, M.B.; Chugh, R.; Skube, S.J.; Dudeja, V.; Borja-Cacho, D.; Clawson, K.A.; Vickers, S.M.; Saluja, A.K. Role of Hsp-70 in Triptolide-Mediated Cell Death of Neuroblastoma. J. Surg. Res. 2010, 163, 72–78.
  185. MacKenzie, T.N.; Mujumdar, N.; Banerjee, S.; Sangwan, V.; Sarver, A.; Vickers, S.; Subramanian, S.; Saluja, A.K. Triptolide induces the expression of miR-142-3p: A negative regulator of heat shock protein 70 and pancreatic cancer cell proliferation. Mol. Cancer Ther. 2013, 12, 1266–1275.
  186. Jacobson, B.A.; Chen, E.Z.; Tang, S.; Belgum, H.S.; McCauley, J.A.; Evenson, K.A.; Etchison, R.G.; Jay-Dixon, J.; Patel, M.; Raza, A.; et al. Triptolide and its prodrug minnelide suppress Hsp70 and inhibit in vivo growth in a xenograft model of mesothelioma. Genes Cancer 2015, 6, 144–152.
  187. Rousalova, I.; Banerjee, S.; Sangwan, V.; Evenson, K.; McCauley, J.A.; Kratzke, R.; Vickers, S.M.; Saluja, A.K.; D’Cunha, J. Minnelide: A Novel Therapeutic That Promotes Apoptosis in Non-Small Cell Lung Carcinoma In Vivo. PLoS ONE 2013, 8, e77411.
  188. Liu, Q. Triptolide and its expanding multiple pharmacological functions. Int. Immunopharmacol. 2011, 11, 377–383.
  189. Propper, D.; Han, H.; Von Hoff, D.; Borazanci, E.; Reya, T.; Ghergurovich, J.; Pshenichnaya, I.; Antal, C.; Condjella, R.; Sharma, S.; et al. Abstract CT165: Phase II open label trial of minnelide™ in patients with chemotherapy refractory metastatic pancreatic cancer. Clin. Trials 2019.
  190. Hung, C.-M.; Su, Y.-H.; Lin, H.-Y.; Lin, J.-N.; Liu, L.-C.; Ho, C.-T.; Way, T.-D. Demethoxycurcumin Modulates Prostate Cancer Cell Proliferation via AMPK-Induced Down-regulation of HSP70 and EGFR. J. Agric. Food Chem. 2012, 60, 8427–8434.
  191. Jung, J.H.; Lee, J.O.; Kim, J.H.; Lee, S.K.; You, G.Y.; Park, S.H.; Park, J.M.; Kim, E.-K.; Suh, P.-G.; An, J.K.; et al. Quercetin suppresses HeLa cell viability via AMPK-induced HSP70 and EGFR down-regulation. J. Cell. Physiol. 2010, 223, 408–414.
  192. Choi, J.-A.; Kim, J.-Y.; Lee, J.-Y.; Kang, C.-M.; Kwon, H.-J.; Yoo, Y.D.; Kim, T.-W.; Lee, Y.-S.; Lee, S.-J. Induction of cell cycle arrest and apoptosis in human breast cancer cells by quercetin. Int. J. Oncol. 2001, 19, 837–844.
  193. Milani, V.; Stangl, S.; Issels, R.; Gehrmann, M.; Wagner, B.; Hube, K.; Mayr, D.; Hiddemann, W.; Molls, M.; Multhoff, G. Anti-tumor activity of patient-derived NK cells after cell-based immunotherapy—A case report. J. Transl. Med. 2009, 7, 50.
  194. Krause, S.W.; Gastpar, R.; Andreesen, R.; Gross, C.; Ullrich, H.; Thonigs, G.; Pfister, K.; Multhoff, G. Treatment of Colon and Lung Cancer Patients with ex Vivo Heat Shock Protein 70-Peptide-Activated, Autologous Natural Killer Cells: A Clinical Phase I Trial. Clin. Cancer Res. 2004, 10, 3699–3707.
  195. Trimble, C.L.; Peng, S.; Kos, F.; Gravitt, P.; Viscidi, R.; Sugar, E.; Pardoll, E.; Wu, T. A phase I trial of a human papillomavirus DNA vaccine for HPV16+ cervical intraepithelial neoplasia 2/3. Clin. Cancer Res. 2009, 15, 361–367.
  196. Maeda, Y.; Yoshimura, K.; Matsui, H.; Shindo, Y.; Tamesa, T.; Tokumitsu, Y.; Hashimoto, N.; Tokuhisa, Y.; Sakamoto, K.; Sakai, K.; et al. Dendritic cells transfected with heat-shock protein 70 messenger RNA for patients with hepatitis C virus-related hepatocellular carcinoma: A phase 1 dose escalation clinical trial. Cancer Immunol. Immunother. 2015, 64, 1047–1056.
  197. Kokowski, K.; Stangl, S.; Seier, S.; Hildebrandt, M.; Vaupel, P.; Multhoff, G. Radiochemotherapy combined with NK cell transfer followed by second-line PD-1 inhibition in a patient with NSCLC stage IIIb inducing long-term tumor control: A case study. Strahlenther. und Onkol. 2019, 195, 352–361.
  198. Li, Z.; Serrano, D.; Baglietto, L.; Johansson, H.; Mariette, F.; Torrisi, R.; Onetto, M.; Paganuzzi, M.; DeCensi, A. Combination of Imatinib Mesylate with Autologous Leukocyte-Derived Heat Shock Protein and Chronic Myelogenous Leukemia. Clin. Cancer Res. 2005, 11, 4460–4468.
  199. ClinicalTrials.gov. Available online: (accessed on 18 February 2020).
  200. Li, J.-L.; Liu, H.-L.; Zhang, X.-R.; Xu, J.-P.; Hu, W.-K.; Liang, M.; Chen, S.-Y.; Hu, F.; Chu, D.-T. A phase I trial of intratumoral administration of recombinant oncolytic adenovirus overexpressing HSP70 in advanced solid tumor patients. Gene Ther. 2008, 16, 376–382.
More
This entry is offline, you can click here to edit this entry!
Video Production Service