Targeting heat shock transcription factor 1 for novel hyperthermia therapy (Review)

  • Authors:
    • Yoshiaki Tabuchi
    • Takashi Kondo
  • View Affiliations

  • Published online on: May 1, 2013     https://doi.org/10.3892/ijmm.2013.1367
  • Pages: 3-8
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Hyperthermia (HT) has shown promising antitumor effects against various types of malignant tumors, and its pleiotropic effects support its combined use with radiotherapy and/or chemotherapy. However, HT is rendered less effective by the acquisition of thermoresistance in tumors, which arises through the elevation of heat shock proteins (HSPs) or other tumor responses. In mammals, the induction of HSPs is principally regulated at the transcriptional level by the activation of heat shock transcription factor 1 (HSF1). This transactivator has been shown to be abundantly expressed in a wide variety of tumors in humans. In addition, HSF1 participates in the initiation, proliferation and maintenance of tumors. Of note, HSF1 silencing has been shown to prevent the progression of tumors and to enhance their sensitivity to HT. Here, we review the physiological and pathological roles of HSF1 in cancer cells, and discuss its potential as a therapeutic target for HT therapy.

Contents

Introduction

Function of HSF1

HSF1 and cancer

HSF1 and hyperthermia

Discussion

Introduction

Hyperthermia (HT) is considered to have potential as a cancer treatment modality (1). HT in combination with radiotherapy and/or chemotherapy has been used for various types of cancer, and the anticancer effects of such combinations have been verified in many clinical trials (17). One of the problems with HT therapy is the acquisition of thermoresistance against heat stress (812). In general, cells have protective functions for various stressors that occur from outside of the cell. Heat shock proteins (HSPs), molecular chaperones with strong cytoprotective and antiapoptotic properties, are induced by a wide variety of stresses, particularly heat stress (1316); they are also induced by treatment with HT, and, thus, it has been considered that HSPs play a role in the acquisition of thermoresistance in cells (1012). In mammals, the expression of HSPs is mainly regulated by heat shock transcription factor 1 (HSF1) (17,18). Elevation of HSF1 has been observed in several types of human tumor tissues (1926), and it has been shown to participate in the initiation, proliferation and maintenance of tumors (12,2529). Notably, both inhibition of the expression of HSF1 and the functional loss of HSF1 have been suggested to enhance the sensitivity to HT under basic experimental conditions (2938). In the present review, the physiological and pathological roles of HSF1 in cancer cells are summarized, and the potential of HSF1 as a therapeutic target for HT therapy is discussed.

Function of HSF1

The heat-shock response, which is a universal cellular response to heat, is a critical cellular event for cell adaptation. Heat stress elicits a wide spectrum of stress responses, including an induction of HSPs, protein aggregation, an imbalance of protein homeostasis, DNA and RNA damage, reactive oxygen species production, cell growth arrest and cell death in mammalian cells. HSPs are characterized as molecular chaperones and they exert strong cytoprotective effects against stress-induced proteotoxic damage (1316). Their functions and amino acid sequences are well conserved in a wide range of species from yeast to humans. The expression of HSPs is primarily regulated by heat shock transcription factors (HSFs). In humans, three HSFs (HSF1, HSF2 and HSF4) have been identified and, among them, HSF1 plays the most important role against stress responses (17,18). Under nonstress conditions, HSF1 is localized in the cytoplasm as an inactive monomer. Upon exposure to stresses, especially heat stress, a series of events is triggered: the HSF1 inactive monomer is converted to a DNA-binding homotrimer that translocates from the cytoplasm to the nucleus, binds to the heat shock elements (HSEs) that are located in the regions of inducible HSP genes including Hsp27, Hsp70 and Hsp90, undergoes a hyperphosphorylation reaction and activates transcription (17,18,39) (Fig. 1). In mammalian cells, heat activates the transcription of heat responsive genes, including HSPs, coincident with a bulk decrease in mRNA and protein syntheses, and this overall reprogramming of gene expression permits the selective synthesis of HSPs (4042). In our previous studies using global-scale microarrays, we also detected a number of genes that were downregulated and participated in biological functions, including post-transcriptional modification and gene expression (4345). This bulk decrease has been considered to represent a form of transcriptional suppression due to the binding of activated HSF1 to the promoter region of the gene (40). Mariner et al (41) recently reported that non-corded RNA is very closely related to the mechanism by which heat induces a decrease in gene expression.

Numerous studies targeting HSFs in gene-modified animals have enhanced our understanding of the pleiotropic roles of HSFs in mammals (46). A previous study with HSF1-knockout (KO) mice demonstrated that the HSF1 protein was dispensable for the maintenance of the mice, since the KO mice remained alive under normal environmental conditions (30). On the other hand, HSF1 is reported to be required for extra-embryonic development, postnatal growth and fertility of the female in the HSF1-KO mice (47). Homma et al (48) observed hallmarks of central nervous system disease, such as demyelination, astrogliosis and accumulation of ubiquitinated proteins in HSF1-deficient mice. Furthermore, previous studies with HSF1 KO animals have shown that HSF1 is involved in the development and maintenance of several organs, such as the heart (49), trachea (50), nose (50) and placenta (47). It has also been demonstrated that both HSF2 (51) and HSF4 (52) play critical roles in the neural functions or lens development by using HSF2- and HSF4-KO mice. Thus, accumulating evidence suggests that HSFs are closely associated with longevity (53), neurodegenerative diseases (54) and cancer (12,2529). Indeed, HSFs play a versatile role in the organism, from homeostatic maintenance to disease.

HSF1 and cancer

In humans, tumorigenesis is a multistep development including genome instabilities (mutation and chromosomal deletion) and epigenetic changes (abnormal histone acetylation and DNA methylation). During the multistep tumorigenesis, mutated nonnative proteins are synthesized and deregulated and abnormal signal transduction pathways exist in cancer cells. Moreover, the conditions within the tumor microenvironment where cancer cells are present include hypoxia, acidity, and low glucose levels, and these differ from the conditions in normal tissue. Therefore, cancer cells are generally exposed to more stresses compared with normal cells (55). Several types of tumors contain high expression levels of one or more HSPs (Hsp27, Hsp40, Hsp60, Hsp70, Hsp90 and/or Hsp110) compared with adjacent normal tissues (12,56,57). A significant elevation of HSF1 has been reported in a wide variety of cancer cells or tumor tissues in nonclinical studies (1926) (Table I). In human prostate carcinoma cell lines, the expression levels of HSPs and HSF1 are elevated in cells with more highly malignant features (19,20). An abundant expression of HSF1 protein has been observed in the tumor tissues of human prostate (19) and pancreas (22) or oral squamous cell carcinoma (OSCC) cells (24) compared to their normal counterparts. In addition, Khaleque et al (21) have shown that the protein level of HSF1 is upregulated in breast carcinoma sections and that HSF1 co-localizes and binds to the corepressor metastasis-associated 1 (MTA1) protein in this cancer.

Table I.

Examples of nonclinical and clinical studies elevating HSF1.

Table I.

Examples of nonclinical and clinical studies elevating HSF1.

Cancer typeObservationsAuthor/(Ref.)
Nonclinical studies
  Prostate cancer specimensProtein level is upregulated in the malignant prostate epithelial cellsHoang et al (19)
  Prostate carcinoma cellsProtein level is elevated in the more highly malignant carcinoma cellsHoang et al (19)
Tang et al (20)
  Breast carcinoma sectionsProtein level is upregulated and it co-localizes with MTA1 proteinKhaleque et al (21)
  Pancreatic cancer tissuesProtein level is upregulated in cancer tissuesDudeja et al (22)
  OSCC cellsProtein and mRNA levels are elevated in carcinoma cellsIshiwata et al (24)
Clinical studies
  Breast cancerNuclear level is associated with poor prognosisSantagata et al (23)
  OSCCNuclear level is related to tumor sizeIshiwata et al (24)
  Breast, colon or lung cancerActive HSF1 transcriptional program is strongly associated with metastasis and mortalityMendillo et al (25)
  Breast cancermRNA level is correlated with grade, metastasis and poor prognosisGabai et al (26)

[i] HSF1, heat shock transcription factor 1; OSCC, oral squamous cell carcinoma; MTA1, metastasis-associated 1.

In 2007, Dai et al (27) found that HSF1, a non-oncogene, plays a major role in enabling the initiation and maintenance of cancer in several mouse tumor models. Knockdown of HSF1 using short hairpin RNA (shRNA) technology effectively inhibited cell viability in several cancer cell lines, whereas it had no effect on normal human fibroblasts. They concluded that HSF1 function helps to maintain the growth and survival of human cancer cells (27). Similar results were reported when other in vitro experiments were performed using melanoma cells (29,36), OSCC cells (37), pancreatic cancer cells (22) and human epidermal growth factor receptor 2 (HER2)-positive cancer cells (58) (Table II). Although the molecular mechanisms by which HSF1 regulates cellular proliferation and survival in cancer cells are less understood, genome-wide transcriptome studies have indicated that HSF1 regulates numerous other targets in addition to HSPs (25,5962).

Table II.

Examples of experimental anticancer effects targeting HSF1.

Table II.

Examples of experimental anticancer effects targeting HSF1.

Cancer typeTreatmentEffectAuthor/(Ref.)
HSF1 silencing
  Pancreatic cancer cellsQuercetinQuercetin increases the apoptosis of cellsAghdassi et al (64)
  Pancreatic cancer xenograftsQuercetinQuercetin reduces the growth of tumorsAghdassi et al (64)
  Colon cancer xenograftsKRIBB11KRIBB11 inhibits the growth of tumorsYoon et al (65)
  Pancreatic cancer xenograftsLy101-4BLy101-4B inhibits the growth of tumorsXia et al (66)
  Mammary epithelial cell xenograftsshRNAshRNA inhibits HER2-induced cellular transformation and tumorigenesisMeng et al (59)
  Melanoma cell xenograftsshRNAshRNA inhibits the growth of melanomasFujimoto et al (29)
  Cancer cellsshRNAshRNA inhibits the growth of cancer cellsDai et al (27)
  HER2-positive cancer cellsshRNAshRNA decreases proliferation of cancer cellsMeng et al (59)
  Melanoma cellsshRNAshRNA inhibits the growth of melanoma cellsNakamura et al (36)
Fujimoto et al (29)
  Pancreatic cancer cellssiRNAsiRNA decreases the viability of cancer cellsDudeja et al (22)
  OSCC cellssiRNAsiRNA inhibits the growth of OSCC cellsTabuchi et al (37)
Genetically modified mice
  Chemical-induced skin carcinomasHSF1 KO mouseKO inhibits carcinogen-induced tumorigenesisDai et al (27)
  p53-deficient spontaneous tumorsHSF1 KO mouse x p53R172H mouseaKO increases survival and alters the distribution of tumor typesDai et al (27)
  p53-deficient spontaneous tumorsHSF1 KO mouse x p53 KO mousebKO selectively decreases lymphomasMin et al (63)
  Chemical-induced hepatocellular carcinomasHSF1 KO mouseKO inhibits carcinogen-induced tumorigenesisJin et al (28)
  NeuT-induced breast tumorsHSF1 KO mouse x Her2/NeuT mousecKO suppresses tumor progressionGabai et al (26)

a HSF1 KO mice were crossed with knockin mice caring the p53R172H hot spot mutation.

b HSF1 KO mice were crossed with p53 KO mice.

c HSF1 KO mice were crossed with transgenic mice expressing HER2/NeuT (a rodent homolog of HER2 caring an activating mutation). HSF1, heat shock transcription factor 1; OSCC, oral squamous cell carcinoma; shRNA, short hairpin RNA; HER2, human epidermal growth factor receptor 2; KO, knockout.

Basic experiments using HSF1 KO mice clearly demonstrate that this protein participates in the development of skin tumors (27) and hepatocellular carcinomas (28) induced by 7,12-dimethylbenzanthracene (DMBA) and diethylnitrosamine (DEN), respectively. In line with these results, different animal experiments with genetically modified mice showed an essential role of HSF1 in tumor development (26,27,63) (Table II).

Interest in the role of HSF1 in cancer has grown gradually over the last decade, and notable papers in clinical research have recently been published (2326) (Table I). Santagata et al (23) indicated that nuclear localization and increased levels of HSF1 are well associated with poor prognosis in estrogen receptor (ER)-positive breast carcinomas. Ishiwata et al (24) reported that higher nuclear HSF1 expression was closely related to tumor size and histopathologic types in OSCCs. Moreover, high expression of HSF1 mRNA in human breast cancer was correlated with grade, metastasis, and poor prognosis, suggesting that HSF1 is involved in tumor progression (26). Mendillo et al (25) identified an HSF1-regulated transcriptional program specific to highly malignant cells and distinct from heat shock by using chromatin immunoprecipitation coupled with massively parallel DNA sequencing. They established an ‘HSF1-cancer signature’ of 456 genes that were bound by HSF1 near their transcription start sites, and found that this HSF1 cancer program is active in several types of tumors and is strongly associated with metastasis and mortality (25) (Table I).

To date, several types of HSF1 inhibitors have been developed, and their inhibitory effects to cancer have been reported (35,6468) (Table II). Quercetin, a plant-derived flavonoid, is shown to increase apoptosis and to reduce the growth of pancreatic cancer cells (64). Meanwhile, this flavonoid inhibits multiple target proteins, such as HSF1, Hsp70 and NF-κB (6769). In a colon cancer xenograft mouse model, a newly synthesized KRIBB11 [N2-(1H-indazole-5-yl)-N6-methyl-3-nitropyridine-2,6-diamine] significantly inhibited the growth of colon cancer, and its mechanism is considered to inhibit Hsp70 synthesis through suppression of HSF1 function by impairing the recruitment of positive transcription elongation factor b (p-TEFb) to the Hsp70 promoter (65). Ly101-4B, a novel triazole nucleoside analog, targets the heat shock response pathway by downregulation of HSF1 and consequential inhibition of HSPs, Hsp27, Hsp70 and Hsp90 (66). This compound caused the shutdown of several oncogenic pathways and caspase-dependent apoptosis, resulting in a potent anticancer effect in pancreatic cancer cell xenografts (66,68). Similar results were obtained when shRNA for HSF1 was used (29,59). These findings have helped to elucidate the specific role of HSF1 in the pathogenesis of cancer.

HSF1 and hyperthermia

HT has been considered a possible mode of cancer treatment (1). HT in combination with radiotherapy and/or chemotherapy has been used for various types of cancer, and the anticancer effects of such combinations have been verified in many clinical trials (17). One of the problems with HT therapy is the acquisition of thermoresistance against heat stress (812). As the expression of HSPs is primarily mediated at the transcription level by HSF1 (17,18), it is predicted that downregulation of HSPs by the targeting of HSF1 will render tumors more sensitive to HT. Examples of the enhancement of HT sensitivity targeting HSF1 are summarized in Table III. It is well established that preconditioning of the cells with mild-HT results in an increase in HSP levels and leads to the acquisition of thermoresistance to HT. However, McMillan et al (30) were the first show that targeted disruption of HSF1 abolished thermoresistance against lethal HT at 45°C for 40 min in preheated (43°C, 30 min) mouse embryonic fibroblast (MEF) cells from HSF1-KO mice. In the MEF cells, silencing HSF1 prevented induction of the expression of two inducible HSPs, Hsp27 and Hsp70, whereas the expression of the constitutive HSPs, Hsp60 and Hsc70, remained constant, suggesting that inducible HSPs are involved in acquired cellular resistance to HT (30). Consistent with this result, functional silencing by using KO mice (31,32) or a dominant-negative (DN) mutant (33) prevented the thermoresistance to HT in MEF and bone marrow progenitor (BMP) cells or breast cancer cells, respectively (Table III).

Table III.

Examples of the enhancement of HT sensitivity targeting HSF1.

Table III.

Examples of the enhancement of HT sensitivity targeting HSF1.

Cell type targetingHeat treatmentEffectAuthor/(Ref.)
HT and HSF1 silencing
  MEF cells KO mouse43°C for 30 min followed by 45°C for 40 minKO prevents resistance to HTMcMillan et al (30)
  MEF cells KO mouse43°C for 30 min followed by 45°C for 60 minKO prevents resistance to HTLuft et al (31)
  MEF and BMP cells KO mouse43°C for 20 min followed by 44°C for 10–40 minKO prevents resistance to HTZhang et al (32)
  Breast cancer cells DN mutant42°C for 30 min followed by 43°C for 20–100 minBlocking prevents resistance to HTWang et al (33)
  Cervical cancer cells triptolide42°C for 60 min followed by 45°C for 40 minInhibition enhances HT sensitivityWesterheide et al (35)
  Cervical cancer cells DN mutant48°C for 10 minDN mutant increases HT sensitivityXia et al (34)
  Melanoma cells shRNA42°C or 43°C for 60 min or 45°C for 60–180 minSilencing enhances HT sensitivityNakamura et al (36)
Fujimoto et al (29)
  OSCC cells siRNA42°C or 44°C for 90 minSilencing enhances HT sensitivityTabuchi et al (37)
  Breast cancer cell xenografts DN mutant43°C for 30 min every day for a weekBlocking inhibits growth of cancer cellsWang et al (33)
HT, compound and HSF1 silencing
  Cervical cancer cells siRNA and cisplatin43°C for 60 minSilencing sensitizes hyperthermochemotherapyRossi et al (38)
  Melanoma cells shRNA and dacarbazine42°C for 60 minSilencing sensitizes hyperthermochemotherapyNakamura et al (36)

[i] HSF1, heat shock transcription factor 1; OSCC, oral squamous cell carcinoma; HT, hyperthermia; shRNA, short hairpin RNA; KO, knockout; BMP, bone marrow progenitor; MEF, mouse embryonic fibroblast.

Of note, targeting of HSF1 has been suggested to enhance the sensitivity to HT under basic experimental conditions (2938). We recently showed that silencing HSF1 using small interfering RNA (siRNA) enhances the mild HT (42°C, 90 min) and HT (44°C, 90 min) sensitivity in human OSCC cells (37). Significant enhancement of the HT sensitivity has also been observed in HSF1-silenced cervical cancer (35) and melanoma cells (29,36) treated with triptolide, an inhibitor of the human heat shock response, and shRNA for HSF1, respectively. In addition, although HT had only a slight effect on tumor growth in breast cancer cell xenografts, functional defect of HSF1 by its DN mutant in combination with HT had a lower tumorigenic capacity than the control tumors (33).

The effects of inhibition of HSH1 on the hyperthermochemotherapy have been reported (36,38). When HSF1 of cervical cancer cells was silenced by siRNA technology, the silencing in combination with hyperthermochemotherapy (treatment at 43°C for 60 min plus cisplatin) markedly increased the level of apoptosis compared to hyperthermochemotherapy alone (38). In human melanoma cells, although HSF1 silencing did not sensitize to chemotherapy with dacarbazine, it significantly enhanced the sensitivity to this chemotherapy in combination with HT (42°C, 60 min) (36). These results strongly suggest that silencing HSF1 can enhance the sensitivity to either HT or hyperthermochemotherapy. It is also noteworthy that combined treatment with HT and cisplatin effectively suppresses the activation of HSF1 in human glioblastoma cells (70).

Discussion

Although HT is considered to be a promising approach in cancer therapy, the thermoresistance due to the increase in HSPs in cancer cells remains a disadvantage, reducing the effects of HT in clinical treatment. The expression of HSPs is mainly regulated by HSF1 (17,18). HSPs (12,56,57) as well as HSF1 (1926) are abundantly expressed in human cancer cells of various origins, and HSPs (12,56,57) and HSF1 (12,2529) play critical roles in the initiation, proliferation and maintenance of cancer. Thus, the combination of HT and HSF1-targeting may be an attractive option and worthy of further investigation. As expected, the inhibition of functions of HSF1 by using gene targeting or HSF1 inhibitors was shown to reduce the acquisition of thermoresistance (3033) and to sensitize tumors to HT-induced cell death (2938). In the near future, targeting HSF1 (67,68) in combination with HT may come to be a promising approach for the treatment of cancer.

Abbreviations:

BMP

bone marrow progenitor;

DN

dominant-negative;

ER

estrogen receptor;

HER2

epidermal growth factor receptor-2;

HSE

heat shock element;

HSP

heat shock protein;

HSF

heat shock transcription factor;

HSF1

heat shock transcription factor 1;

HT

hyperthermia;

KO

knockout;

MTA1

metastasis-associated 1;

MEF

mouse embryonic fibroblast;

OSCC

oral squamous cell carcinoma;

p-TEFb

positive transcription elongation factor b;

shRNA

short hairpin RNA;

siRNA

small interfering RNA

References

1. 

Hall EJ: Hyperthermia. Radiobiology for the Radiologist. 5th edition. Lippincott Williams and Wilkins; Philadelphia, PA: pp. 495–520. 2000

2. 

van der Zee J, González González D, van Rhoon GC, van Dijk JD, van Putten WL and Hart AA: Comparison of radiotherapy alone with radiotherapy plus hyperthermia in locally advanced pelvic tumours: a prospective, randomised, multicentre trial. Dutch Deep Hyperthermia Group Lancet. 355:1119–1125. 2000.PubMed/NCBI

3. 

Harima Y, Nagata K, Harima K, Ostapenko VV, Tanaka Y and Sawada S: A randomized clinical trial of radiation therapy versus thermoradiotherapy in stage IIIB cervical carcinoma. Int J Hyperthermia. 17:97–105. 2001. View Article : Google Scholar

4. 

Wust P, Hildebrandt B, Sreenivasa G, et al: Hyperthermia in combined treatment of cancer. Lancet Oncol. 3:487–497. 2002. View Article : Google Scholar : PubMed/NCBI

5. 

Issels RD: Hyperthermia adds to chemotherapy. Eur J Cancer. 44:2546–2554. 2008. View Article : Google Scholar : PubMed/NCBI

6. 

Zagar TM, Oleson JR, Vujaskovic Z, et al: Hyperthermia combined with radiation therapy for superficial breast cancer and chest wall recurrence: a review of the randomised data. Int J Hyperthermia. 26:612–617. 2010. View Article : Google Scholar : PubMed/NCBI

7. 

Westermann A, Mella O, Van Der Zee J, et al: Long-term survival data of triple modality treatment of stage IIB-III-IVA cervical cancer with the combination of radiotherapy, chemotherapy and hyperthermia - an update. Int J Hyperthermia. 28:549–553. 2012. View Article : Google Scholar : PubMed/NCBI

8. 

Rhee JG, Schuman VL, Song CW and Levitt SH: Difference in the thermotolerance of mouse mammary carcinoma cells in vivo and in vitro. Cancer Res. 47:2571–2575. 1987.PubMed/NCBI

9. 

Dings RP, Loren ML, Zhang Y, et al: Tumour thermotolerance, a physiological phenomenon involving vessel normalisation. Int J Hyperthermia. 27:42–52. 2011. View Article : Google Scholar : PubMed/NCBI

10. 

Li GC, Mivechi NF and Weitzel G: Heat shock proteins, thermotolerance, and their relevance to clinical hyperthermia. Int J Hyperthermia. 11:459–488. 1995. View Article : Google Scholar : PubMed/NCBI

11. 

Cheng L, Smith DJ, Anderson RL and Nagley P: Human neuroblastoma SH-SY5Y cells show increased resistance to hyperthermic stress after differentiation, associated with elevated levels of Hsp72. Int J Hyperthermia. 27:415–426. 2011. View Article : Google Scholar

12. 

Mosser DD and Morimoto RI: Molecular chaperones and the stress of oncogenesis. Oncogene. 23:2907–2918. 2004. View Article : Google Scholar : PubMed/NCBI

13. 

Lindquist S and Craig EA: The heat-shock proteins. Annu Rev Genet. 22:631–677. 1988. View Article : Google Scholar

14. 

Hartl FU: Molecular chaperones in cellular protein folding. Nature. 381:571–579. 1996. View Article : Google Scholar : PubMed/NCBI

15. 

Beere HM: ‘The stress of dying’: the role of heat shock proteins in the regulation of apoptosis. J Cell Sci. 117:2641–2651. 2004.

16. 

Richter K, Haslbeck M and Buchner J: The heat shock response: life on the verge of death. Mol Cell. 40:253–266. 2010. View Article : Google Scholar : PubMed/NCBI

17. 

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. 12:3788–3796. 1998. View Article : Google Scholar : PubMed/NCBI

18. 

Akerfelt M, Morimoto RI and Sistonen L: Heat shock factors: integrators of cell stress, development and lifespan. Nat Rev Mol Cell Biol. 11:545–555. 2010. View Article : Google Scholar : PubMed/NCBI

19. 

Hoang AT, Huang J, Rudra-Ganguly N, et al: A novel association between the human heat shock transcription factor 1 (HSF1) and prostate adenocarcinoma. Am J Pathol. 156:857–864. 2000. View Article : Google Scholar : PubMed/NCBI

20. 

Tang D, Khaleque MA, Jones EL, 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. 10:46–58. 2005. View Article : Google Scholar : PubMed/NCBI

21. 

Khaleque MA, Bharti A, Gong J, et al: Heat shock factor 1 represses estrogen-dependent transcription through association with MTA1. Oncogene. 27:1886–1893. 2008. View Article : Google Scholar : PubMed/NCBI

22. 

Dudeja V, Chugh RK, Sangwan V, et al: Prosurvival role of heat shock factor 1 in the pathogenesis of pancreatobiliary tumors. Am J Physiol Gastrointest Liver Physiol. 300:G948–G955. 2011. View Article : Google Scholar : PubMed/NCBI

23. 

Santagata S, Hu R, Lin NU, et al: High levels of nuclear heat-shock factor 1 (HSF1) are associated with poor prognosis in breast cancer. Proc Natl Acad Sci USA. 108:18378–18383. 2011. View Article : Google Scholar : PubMed/NCBI

24. 

Ishiwata J, Kasamatsu A, Sakuma K, et al: State of heat shock factor 1 expression as a putative diagnostic marker for oral squamous cell carcinoma. Int J Oncol. 40:47–52. 2012.PubMed/NCBI

25. 

Mendillo ML, Santagata S, Koeva M, et al: HSF1 drives a transcriptional program distinct from heat shock to support highly malignant human cancers. Cell. 150:549–562. 2012. View Article : Google Scholar : PubMed/NCBI

26. 

Gabai VL, Meng L, Kim G, Mills TA, Benjamin IJ and Sherman MY: 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. 32:929–940. 2012. View Article : Google Scholar : PubMed/NCBI

27. 

Dai C, Whitesell L, Rogers AB and Lindquist S: Heat shock factor 1 is a powerful multifaceted modifier of carcinogenesis. Cell. 130:1005–1018. 2007. View Article : Google Scholar : PubMed/NCBI

28. 

Jin X, Moskophidis D and Mivechi NF: Heat shock transcription factor 1 is a key determinant of HCC development by regulating hepatic steatosis and metabolic syndrome. Cell Metab. 14:91–103. 2011. View Article : Google Scholar : PubMed/NCBI

29. 

Fujimoto M, Takaki E, Takii R, et al: RPA Assists HSF1 access to nucleosomal DNA by recruiting histone chaperone FACT. Mol Cell. 48:182–194. 2012. View Article : Google Scholar : PubMed/NCBI

30. 

McMillan DR, Xiao X, Shao L, Graves K and Benjamin IJ: Targeted disruption of heat shock transcription factor 1 abolishes thermotolerance and protection against heat-inducible apoptosis. J Biol Chem. 273:7523–7528. 1998. View Article : Google Scholar : PubMed/NCBI

31. 

Luft JC, Benjamin IJ, Mestril R and Dix DJ: Heat shock factor 1-mediated thermotolerance prevents cell death and results in G2/M cell cycle arrest. Cell Stress Chaperones. 6:326–336. 2001. View Article : Google Scholar : PubMed/NCBI

32. 

Zhang Y, Huang L, Zhang J, Moskophidis D and Mivechi NF: Targeted disruption of hsf1 leads to lack of thermotolerance and defines tissue-specific regulation for stress-inducible Hsp molecular chaperones. J Cell Biochem. 86:376–393. 2002. View Article : Google Scholar : PubMed/NCBI

33. 

Wang JH, Yao MZ, Gu JF, Sun LY, Shen YF and Liu XY: Blocking HSF1 by dominant-negative mutant to sensitize tumor cells to hyperthermia. Biochem Biophys Res Commun. 290:1454–1461. 2002. View Article : Google Scholar : PubMed/NCBI

34. 

Xia W, Vilaboa N, Martin JL, Mestril R, Guo Y and Voellmy R: Modulation of tolerance by mutant heat shock transcription factors. Cell Stress Chaperones. 4:8–18. 1999. View Article : Google Scholar : PubMed/NCBI

35. 

Westerheide SD, Kawahara TL, Orton K and Morimoto RI: Triptolide, an inhibitor of the human heat shock response that enhances stress-induced cell death. J Biol Chem. 281:9616–9622. 2006. View Article : Google Scholar : PubMed/NCBI

36. 

Nakamura Y, Fujimoto M, Hayashida N, Takii R, Nakai A and Muto M: Silencing HSF1 by short hairpin RNA decreases cell proliferation and enhances sensitivity to hyperthermia in human melanoma cell lines. J Dermatol Sci. 60:187–192. 2010. View Article : Google Scholar : PubMed/NCBI

37. 

Tabuchi Y, Furusawa Y, Wada S, Ohtsuka K and Kondo T: Silencing heat shock transcription factor 1 using small interfering RNA enhances mild hyperthermia and hyperthermia sensitivity in human oral squamous cell carcinoma cells. Thermal Med. 27:99–108. 2011. View Article : Google Scholar

38. 

Rossi A, Ciafrè S, Balsamo M, Pierimarchi P and Santoro MG: Targeting the heat shock factor 1 by RNA interference: a potent tool to enhance hyperthermochemotherapy efficacy in cervical cancer. Cancer Res. 66:7678–7685. 2006. View Article : Google Scholar : PubMed/NCBI

39. 

Sakurai H and Enoki Y: Novel aspects of heat shock factors: DNA recognition, chromatin modulation and gene expression. FEBS J. 277:4140–4149. 2010. View Article : Google Scholar : PubMed/NCBI

40. 

Westwood JT, Clos J and Wu C: Stress-induced oligomerization and chromosomal relocalization of heat-shock factor. Nature. 353:822–827. 1991. View Article : Google Scholar : PubMed/NCBI

41. 

Mariner PD, Walters RD, Espinoza CA, et al: Human Alu RNA is a modular transacting repressor of mRNA transcription during heat shock. Mol Cell. 29:499–509. 2008. View Article : Google Scholar : PubMed/NCBI

42. 

Spriggs KA, Bushell M and Willis AE: Translational regulation of gene expression during conditions of cell stress. Mol Cell. 40:228–237. 2010. View Article : Google Scholar : PubMed/NCBI

43. 

Furusawa Y, Tabuchi Y, Wada S, Takasaki I, Ohtsuka K and Kondo T: Identification of biological functions and gene networks regulated by heat stress in U937 human lymphoma cells. Int J Mol Med. 28:143–151. 2011.PubMed/NCBI

44. 

Tabuchi Y, Wada S, Furusawa Y, Ohtsuka K and Kondo T: Gene networks related to the cell death elicited by hyperthermia in human oral squamous cell carcinoma HSC-3 cells. Int J Mol Med. 29:380–386. 2012.PubMed/NCBI

45. 

Tabuchi Y, Furusawa Y, Kariya A, Wada S, Ohtsuka K and Kondo T: Common gene expression patterns responsive to mild temperature hyperthermia in normal human fibroblastic cells. Int J Hyperthermia. 29:38–50. 2013. View Article : Google Scholar : PubMed/NCBI

46. 

Jin X, Eroglu B, Moskophidis D and Mivechi NF: Targeted deletion of Hsf1, 2, and 4 genes in mice. Methods Mol Biol. 787:1–20. 2011. View Article : Google Scholar : PubMed/NCBI

47. 

Xiao X, Zuo X, Davis AA, et al: HSF1 is required for extra-embryonic development, postnatal growth and protection during inflammatory responses in mice. EMBO J. 18:5943–5952. 1999. View Article : Google Scholar : PubMed/NCBI

48. 

Homma S, Jin X, Wang G, et al: Demyelination, astrogliosis, and accumulation of ubiquitinated proteins, hallmarks of CNS disease in hsf1-deficient mice. J Neurosci. 27:7974–7986. 2007. View Article : Google Scholar : PubMed/NCBI

49. 

Zou Y, Zhu W, Sakamoto M, et al: Heat shock transcription factor 1 protects cardiomyocytes from ischemia/reperfusion injury. Circulation. 108:3024–3030. 2003. View Article : Google Scholar : PubMed/NCBI

50. 

Takaki E, Fujimoto M, Sugahara K, et al: Maintenance of olfactory neurogenesis requires HSF1, a major heat shock transcription factor in mice. J Biol Chem. 281:4931–4937. 2006. View Article : Google Scholar : PubMed/NCBI

51. 

Kallio M, Chang Y, Manuel M, et al: Brain abnormalities, defective meiotic chromosome synapsis and female subfertility in HSF2 null mice. EMBO J. 21:2591–2601. 2002. View Article : Google Scholar : PubMed/NCBI

52. 

Fujimoto M, Izu H, Seki K, et al: HSF4 is required for normal cell growth and differentiation during mouse lens development. EMBO J. 23:4297–4306. 2004. View Article : Google Scholar : PubMed/NCBI

53. 

Bishop NA and Guarente L: Genetic links between diet and lifespan: Shared mechanisms from yeast to humans. Nat Rev Genet. 8:835–844. 2007. View Article : Google Scholar : PubMed/NCBI

54. 

Neef DW, Jaeger AM and Thiele DJ: Heat shock transcription factor 1 as a therapeutic target in neurodegenerative diseases. Nat Rev Drug Discov. 10:930–944. 2011. View Article : Google Scholar : PubMed/NCBI

55. 

Hanahan D and Weinberg RA: Hallmarks of cancer: the next generation. Cell. 144:646–674. 2011. View Article : Google Scholar : PubMed/NCBI

56. 

Ciocca DR and Calderwood SK: Heat shock proteins in cancer: diagnostic, prognostic, predictive, and treatment implications. Cell Stress Chaperones. 10:86–103. 2005. View Article : Google Scholar : PubMed/NCBI

57. 

Ciocca DR, Arrigo AP and Calderwood SK: Heat shock proteins and heat shock factor 1 in carcinogenesis and tumor development: an update. Arch Toxicol. 87:19–48. 2013. View Article : Google Scholar : PubMed/NCBI

58. 

Tabuchi Y, Kariya A, Yunoki T and Kondo T: Genes involved in the cell death induced by knockdown of heat shock transcription factor 1 in human oral squamous cell carcinoma HSC-3 cells. Thermal Med. 28:29–42. 2012. View Article : Google Scholar

59. 

Meng L, Gabai VL and Sherman MY: Heat-shock transcription factor HSF1 has a critical role in human epidermal growth factor receptor-2-induced cellular transformation and tumorigenesis. Oncogene. 29:5204–5213. 2010. View Article : Google Scholar : PubMed/NCBI

60. 

Hahn JS, Hu Z, Thiele DJ and Iyer VR: Genome-wide analysis of the biology of stress responses through heat shock transcription factor. Mol Cell Biol. 24:5249–5256. 2004. View Article : Google Scholar : PubMed/NCBI

61. 

Trinklein ND, Murray JI, Hartman SJ, Botstein D and Myers RM: The role of heat shock transcription factor 1 in the genome-wide regulation of the mammalian heat shock response. Mol Biol Cell. 15:1254–1261. 2004. View Article : Google Scholar : PubMed/NCBI

62. 

Page TJ, Sikder D, Yang L, et al: Genome-wide analysis of human HSF1 signaling reveals a transcriptional program linked to cellular adaptation and survival. Mol Biosyst. 2:627–639. 2006. View Article : Google Scholar : PubMed/NCBI

63. 

Min JN, Huang L, Zimonjic DB, Moskophidis D and Mivechi NF: Selective suppression of lymphomas by functional loss of Hsf1 in a p53-deficient mouse model for spontaneous tumors. Oncogene. 26:5086–5097. 2007. View Article : Google Scholar : PubMed/NCBI

64. 

Aghdassi A, Phillips P, Dudeja V, et al: Heat shock protein 70 increases tumorigenicity and inhibits apoptosis in pancreatic adenocarcinoma. Cancer Res. 67:616–625. 2007. View Article : Google Scholar : PubMed/NCBI

65. 

Yoon YJ, Kim JA, Shin KD, et al: KRIBB11 inhibits HSP70 synthesis through inhibition of heat shock factor 1 function by impairing the recruitment of positive transcription elongation factor b to the hsp70 promoter. J Biol Chem. 286:1737–1747. 2011. View Article : Google Scholar : PubMed/NCBI

66. 

Xia Y, Liu Y, Rocchi P, et al: Targeting heat shock factor 1 with a triazole nucleoside analog to elicit potent anticancer activity on drug-resistant pancreatic cancer. Cancer Lett. 318:145–153. 2012. View Article : Google Scholar : PubMed/NCBI

67. 

Whitesell L and Lindquist S: Inhibiting the transcription factor HSF1 as an anticancer strategy. Expert Opin Ther Targets. 13:469–478. 2009. View Article : Google Scholar : PubMed/NCBI

68. 

Xia Y, Rocchi P, Iovanna JL and Peng L: Targeting heat shock response pathways to treat pancreatic cancer. Drug Discov Today. 17:35–43. 2012. View Article : Google Scholar : PubMed/NCBI

69. 

Chen SS, Michael A and Butler-Manuel SA: Advances in the treatment of ovarian cancer: a potential role of antiinflammatory phytochemicals. Discov Med. 13:7–17. 2012.PubMed/NCBI

70. 

Matsumoto H, Hayashi S, Shioura H, et al: Suppression of heat-induced HSF activation by CDDP in human glioblastoma cells. Int J Radiat Oncol Biol Phys. 41:915–920. 1998. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

July 2013
Volume 32 Issue 1

Print ISSN: 1107-3756
Online ISSN:1791-244X

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Tabuchi Y and Tabuchi Y: Targeting heat shock transcription factor 1 for novel hyperthermia therapy (Review). Int J Mol Med 32: 3-8, 2013
APA
Tabuchi, Y., & Tabuchi, Y. (2013). Targeting heat shock transcription factor 1 for novel hyperthermia therapy (Review). International Journal of Molecular Medicine, 32, 3-8. https://doi.org/10.3892/ijmm.2013.1367
MLA
Tabuchi, Y., Kondo, T."Targeting heat shock transcription factor 1 for novel hyperthermia therapy (Review)". International Journal of Molecular Medicine 32.1 (2013): 3-8.
Chicago
Tabuchi, Y., Kondo, T."Targeting heat shock transcription factor 1 for novel hyperthermia therapy (Review)". International Journal of Molecular Medicine 32, no. 1 (2013): 3-8. https://doi.org/10.3892/ijmm.2013.1367