Skip to main content

Hypoxia and hypoglycaemia in Ewing's sarcoma and osteosarcoma: regulation and phenotypic effects of Hypoxia-Inducible Factor

Abstract

Background

Hypoxia regulates gene expression via the transcription factor HIF (Hypoxia-Inducible Factor). Little is known regarding HIF expression and function in primary bone sarcomas. We describe HIF expression and phenotypic effects of hypoxia, hypoglycaemia and HIF in Ewing's sarcoma and osteosarcoma.

Methods

HIF-1α and HIF-2α immunohistochemistry was performed on a Ewing's tumour tissue array. Ewing's sarcoma and osteosarcoma cell lines were assessed for HIF pathway induction by Western blot, luciferase assay and ELISA. Effects of hypoxia, hypoglycaemia and isoform-specific HIF siRNA were assessed on proliferation, apoptosis and migration.

Results

17/56 Ewing's tumours were HIF-1α-positive, 15 HIF-2α-positive and 10 positive for HIF-1α and HIF-2α. Expression of HIF-1α and cleaved caspase 3 localised to necrotic areas. Hypoxia induced HIF-1α and HIF-2α in Ewing's and osteosarcoma cell lines while hypoglycaemia specifically induced HIF-2α in Ewing's. Downstream transcription was HIF-1α-dependent in Ewing's sarcoma, but regulated by both isoforms in osteosarcoma. In both cell types hypoglycaemia reduced cellular proliferation by ≥ 45%, hypoxia increased apoptosis and HIF siRNA modulated hypoxic proliferation and migration.

Conclusions

Co-localisation of HIF-1α and necrosis in Ewing's sarcoma suggests a role for hypoxia and/or hypoglycaemia in in vivo induction of HIF. In vitro data implicates hypoxia as the primary HIF stimulus in both Ewing's and osteosarcoma, driving effects on proliferation and apoptosis. These results provide a foundation from which to advance understanding of HIF function in the pathobiology of primary bone sarcomas.

Peer Review reports

Background

Hypoxia is a fundamental micro-environmental component of solid tumour tissue which is associated with resistance to therapy, poor survival and a malignant phenotype [1]. Hypoxia induces stabilisation of the Hypoxia-Inducible Factor transcription factors, HIF-1 and HIF-2, which direct responses central to survival and expansion of the malignant cell population.

HIF comprises a hypoxia-inducible alpha subunit and a constitutively expressed beta subunit. Regulation of the active transcription factor occurs via enzymatic control of the abundance and activity of HIFα subunits. Under normoxia HIFα is post-translationally hydroxylated by the prolyl hydroxylase domain enzymes, targeting it for proteasomal degradation [2, 3]. These enzymes are absolutely dependent on O2 and limitation of activity under hypoxia allows stabilisation of HIFα. Binding of the active complex to the hypoxia-response element (HRE) of target genes results in activation of pathways regulating processes such as angiogenesis, apoptosis and metabolic adaptation [4].

HIF over-expression is an independent prognostic factor in many carcinomas [5], although limited data is available in primary bone sarcomas. Nuclear over-expression of HIF-1α has been reported in approximately 60% of clinical osteosarcomas where it correlates with disease grade, stage, recurrence and survival [6–8]. In Ewing's sarcoma the presence of tumour cell-lined blood lakes correlates with clinical outcome. Cells surrounding these lakes also express HIF-1α and the hypoxia marker pimonidazole [9]. Expression of both HIF-1α and HIF-2α has been reported in chondrosarcoma [10] and giant cell tumour of bone [11], HIF-1α expression correlating with reduced disease-free survival in chondrosarcoma. Levels of serum VEGF, a HIF target gene, are significantly higher in Ewing's patients than healthy controls [12] and are an independent prognostic factor for survival [13]. In osteosarcoma VEGF levels were higher in the tumour and serum of patients who subsequently relapsed, tumour VEGF being predictive of pulmonary metastasis and poor prognosis [14, 15]. In Ewing's sarcoma the presence of necrotic, non-perfused and presumably hypoxic tumour areas correlates with survival and frequency of metastatic spread [16–18].

This data implies that hypoxia and/or HIF contribute substantially to the pathobiology of primary bone sarcomas. Hypoxia induces expression of HIF-1α and VEGF in the osteosarcoma cell lines Saos2, 143B, U2-OS and MG-63 [11, 19–21], with MG-63 also expressing HIF-2α [11, 22]. The Ewing's sarcoma cell lines A673, SK-ES-1, SK-N-MC and TC-71 also demonstrate hypoxic induction of HIF-1α and downstream genes [23–25]. Despite such evidence for hypoxic activation of the HIF transcriptional cascade in osteosarcoma and Ewing's sarcoma cells, little is known regarding the effect of either HIF-1α or HIF-2α on the hypoxic phenotype of these cells. We have therefore analysed characteristics of the induction of HIF-1α, HIF-2α and HIF target genes in a panel of osteosarcoma and Ewing's sarcoma cell lines and investigated effects of isoform-specific HIF siRNA on the hypoxic phenotype of these cells.

Methods

Reagents

Tissue culture reagents were from Lonza (Wokingham, UK), except FBS (Invitrogen, Paisley, UK). Unless otherwise stated, reagents were from Sigma (Poole, UK). This study was approved by the Oxford Clinical Research Ethics Committee (C01.071).

Immunohistochemistry

A tissue array comprising 47 Ewing's sarcomas was constructed at the University of Dusseldorf. Additional sections were from the Nuffield Orthopaedic Centre. Formalin fixed sections were stained for HIF-1α (BD Biosciences, Oxford, UK), HIF-2α (EP190b, Abcam, Cambridge, UK), Glut-1 (Abcam) and cleaved caspase 3 (Cell Signalling Technology, Danvers, USA). Staining was visualized with the EnVision™ Peroxidase/DAB Rabbit/Mouse detection kit (Dako, Ely, UK). Image acquisition was performed using an Olympus BX40 microscope with 20× or 40× objective, Olympus DP70 camera and CellF. Tumours were considered positive for an antigen when ≥ 3 positive cells/field of view were observed at 20× magnification.

Cell culture

Cell lines were obtained from the EuroBoNeT cell line biobank, comprising recently characterised bone tumour cell lines [26], and maintained in culture for < 30 passages. All cell lines were maintained in RPMI (except TC-71; IMDM) with 10% FBS, L-glutamine (2 mM), penicillin (50 IU/ml) and streptomycin sulphate (50 μg/ml) in a humidified atmosphere at 37°C (5% CO2 in air). Hypoxic exposures were performed in 0.1% O2, 5% CO2, balance N2 in a MiniGalaxy incubator (RS Biotech, Irvine, UK). Low glucose conditions were achieved using RPMI media without glucose, supplemented as for normal media, under normoxic conditions.

Western blotting

Cells were homogenized in lysis buffer (6.2 M urea, 10% glycerol, 5 M DTT, 1% SDS, protease inhibitors). Whole cell extract was separated by 8% SDS-PAGE and transferred to PVDF membrane. Primary antibodies were against HIF-1α, HIF-2α, Glut-1, Ki-67 (MIB-1, Dako) and β-tubulin. Immuno-reactivity was visualised with HRP-linked goat serum and chemiluminescence. Densitometric quantification of Ki-67, normalised to β-tubulin, was performed on scanned blots using ImageJ.

Luciferase assay

Cells at 60% confluence were transfected with a PGK-HRE-firefly luciferase reporter construct and the control pHRG-TK renilla luciferase plasmid (Promega, Southampton, UK) using Lipofectamine 2000 (Invitrogen). 24 h post-transfection cells were exposed to experimental conditions then lysed in Passive Lysis Buffer. Firefly and renilla luciferase were assayed using the Dual-Luciferase Reporter Assay System (Promega), with firefly luciferase values normalized to the renilla transfection control.

ELISA

Secretion of VEGF and TGFα was measured using the Human VEGF DuoSet and Quantikine Human TGFα Immunoassay (R&D Systems, Abingdon, UK) respectively.

Cell number, proliferation and apoptosis

Cellular apoptosis was assessed using the Apo-One Homogeneous Caspase-3/7 Assay (Promega), with apoptosis levels normalised to total cell number (CellTiter 96 Aqueous One Solution Cell Proliferation Assay, Promega). For quantification of mitotic index cells were fixed in formalin, stained for H&E and the number of mitotic figures was expressed as a percentage of total cell number in 4 random fields of view.

siRNA transfection

Small interfering RNA (siRNA) sequences against HIF-1α (H1), HIF-2α (H2) and HIF-1α scrambled control (scr) were as described [27] and obtained pre-annealed from Eurogentec. Cells at 40% confluence were transfected with 50 nM siRNA duplex using Lipofectamine 2000. Duplexes were removed after 24 h and cells treated as required.

Scratch migration assay

A scratch was made through cells at 90% confluence using a 20 ul pipette tip. Specific points of the scratch were photographed before and after exposure to experimental conditions. Wound width was measured in CellF and migration expressed as fraction wound closure.

Statistical significance

Results are expressed as mean ± SD of at least three independent experiments. Statistical analysis comprised one-way ANOVA using Bonferroni's multiple comparison as a post-hoc test, with results considered significant at p < 0.05.

Results

HIF-1α and HIF-2α are expressed in Ewing's sarcoma

As HIF-1α has already been described to correlate with clinical and survival parameters in osteosarcoma, immunohistochemistry focussed on Ewing's sarcoma. Of 56 Ewing's tumours assessed for expression of HIF; 30% expressed only HIF-1α, 27% only HIF-2α, 18% both HIF-1α and HIF-2α and 25% neither HIF isoform. HIF-1α was predominantly localised in the nucleus of Ewing's tumour cells, whereas HIF-2α expression was mainly cytoplasmic (Figure 1A, B). In tumours expressing both HIF-1α and HIF-2α there was no evidence of co-localisation and no correlation in intensity of staining. In 17 of the 27 HIF-1α-positive cases expression of HIF-1α was associated with regions of necrosis, areas of tissue also immunoreactive for the apoptosis marker cleaved caspase 3 (Figure 1C, D). No such relationship was evident for HIF-2α. With the clinical data available, no correlation was observed between HIF expression and clinical parameters including tumour volume, metastasis and survival.

Figure 1
figure 1

Expression of HIF in Ewing's sarcoma. (a) HIF-1α and (b) HIF-2α expression in Ewing's tumour cells; scale bar = 50 μm. Expression of HIF-1α (c) co-localises with that for cleaved caspase 3 (d); scale bar = 100 μm.

Differential HIF response of ES and OS cells to hypoxia and low glucose

HIF-1α protein was stabilised in response to hypoxia in all Ewing's (ES) and osteosarcoma (OS) cell lines tested. Hypoxia induced the HIF-2α isoform in all OS and in 4/5 ES cell lines (Figure 2A). Low glucose conditions were also assessed as an additional micro-environmental characteristic of regions of necrosis. Low glucose produced no HIF response in any OS cell line, while ES cells displayed strong induction of HIF-2α protein and a moderate increase in HIF-1α in SK-ES-1 and SK-N-MC (Figure 2B). This pattern of HIF induction was reflected in transcriptional events downstream of HIF. Both ES and OS cells strongly transactivated the HRE-promoter element in response to hypoxia and induced expression of the HIF target genes, Glut-1 and VEGF (Figure 2A, C, D). ES cells induced expression of both Glut-1 and VEGF in response to low glucose, although the magnitude of induction was less than for hypoxia and no transcriptional activation was evident in the HRE promoter-luciferase assay (Figure 2B-D). OS cells showed no HIF transcriptional response to low glucose conditions.

Figure 2
figure 2

HIF induction in ES and OS cell lines. Western blots showing induction of HIF-1α (120 kDa), HIF-2α (120 kDa) and Glut-1 (50 kDa) in a panel of ES (A673, RD-ES, SK-ES-1, SK-N-MC, TC-71) and OS (MG-63, Saos2, G292, 143B) cell lines in response to 24 h (a) hypoxia (H; 0.1% O2) or (b) low glucose (G; over-exposed in comparison to hypoxia blots) compared with the untreated/normoxic (N) control. (c) Activation of the HRE-luciferase reporter construct and (d) secretion of VEGF in response to hypoxia (light grey bars) or low glucose (dark grey bars). *, p < 0.05; **, p < 0.01; ***, p < 0.001.

Distinct transcriptional effects of HIF-1α and HIF-2α in ES and OS cells

Isoform-specific HIF siRNA was used to distinguish the relative contribution of HIF-1α and HIF-2α to the HIF transcriptional response. In ES cell lines, HIF-2α siRNA had no effect on Glut-1 expression (Figure 3A), VEGF secretion (Figure 3B) or HRE-luciferase activity (Figure 3C). HIF-1α siRNA significantly inhibited hypoxic induction of Glut-1, VEGF and HRE transactivation (Figure 3A-C). Normoxic levels of Glut-1 and VEGF were also reduced by HIF-1α siRNA, even in the absence of immuno-detectable HIF (Figure 3A, B). However, hypoxia and low glucose both induced secretion of HIF-2-specific TGFα [28, 29] in ES cell lines which express the HIF-2 isoform (Figure 3D). In contrast, HIF transactivation in OS cell lines was regulated by a combination of both transcription factors. Suppressed hypoxic induction of Glut-1 (Figure 3A), VEGF (Figure 3B) and HRE-luciferase activity (Figure 3C) was evident with siRNA targeting either HIF-1α or HIF-2α, combined 'HIF-1α plus HIF-2α' siRNA generally exhibiting an additive effect.

Figure 3
figure 3

HIF siRNA in ES and OS cell lines. (a) Western blots showing expression of HIF-1α (120 kDa), HIF-2α (120 kDa) and Glut-1 (50 kDa) in response to siRNA targeting HIF-1α (H1), HIF-2α (H2), HIF-1α and HIF-2α (H12) or control siRNA (scr) in ES (SK-N-MC) and OS (143B) cells. Effects of HIF siRNA on (b) secretion of VEGF and (c) activation of the HRE-luciferase reporter construct and in response to hypoxia or low glucose in ES and OS cell lines. Bar graph data is normalised to the siRNA mock control (-) for each condition and represents pooled data from the entire panel of ES or OS cell lines. *, p < 0.05; **, p < 0.01; ***, p < 0.001 versus scr siRNA control. (d) Secretion of TGFα from ES cell lines in response to hypoxia (light grey bars) or low glucose (dark grey bars). *, p < 0.05; **, p < 0.01; ***, p < 0.001.

Phenotypic effects of hypoxia and low glucose in ES and OS cell lines

Phenotypic effects of 72 h exposure to hypoxia or low glucose conditions were assessed in a reduced panel of 3 ES and 3 OS cell lines. Hypoxia had no effect on total cell number in any cell line except 143B, whereas low glucose consistently reduced cell number by > 45% (Figure 4A). Caspase 3/7 activation was assessed as a marker of cellular apoptosis. Hypoxia increased apoptosis in both cell panels, having a 2 to 3-fold greater effect in OS cell lines. Glucose deprivation only had a significant effect on apoptosis in SK-ES-1 (Figure 4B). Proliferation was measured in OS cells by mitotic index and in ES cells by immuno-detection of Ki-67. The percentage of mitoses generally did not change during exposure to hypoxia, but was reduced by low glucose conditions (Figure 4C). Similarly, Ki-67 expression was inhibited by low glucose conditions but maintained on exposure to hypoxia (Figure 4D).

Figure 4
figure 4

Phenotypic effects of hypoxia and low glucose. Effect of hypoxia (light grey bars) or low glucose (dark grey bars) on (a) cell number and (b) apoptosis in ES and OS cell lines and (c) mitotic index in OS cell lines. Bar graph data is normalised to the normoxic control (white bars) for each condition: *, p < 0.05; **, p < 0.01; ***, p < 0.001 versus control. (d) Proliferation assessed in ES cell lines by Western blot for Ki-67. N, normoxia; H, hypoxia; G, low glucose. Numerical values represent mean fold change in Ki-67 as assessed by densitometry:*, p < 0.05; **, p < 0.01 versus normoxia.

Effect of HIF siRNA on the hypoxic phenotype

Since no differences were observed between ES and OS cell lines regarding phenotypic effects of glucose deprivation, and as OS cells have no HIF transcriptional response to this stimulus, effects of HIF siRNA were only investigated under hypoxia. No effect was observed on cell number or apoptosis under hypoxia with the exception of a minor increase in OS cell number with HIF-2α siRNA, the magnitude of which questions its biological significance (Figure 5A, B). However, in ES cells HIF-1α siRNA consistently increased cell proliferation under hypoxia as assessed by Western blotting for Ki-67 (Figure 5C). Conversely, siRNA targeting either HIF-1α or HIF-2α resulted in reduced hypoxic proliferation in 2/3 OS cell lines, although siRNA targeting either HIF-1α or HIF-2α in 143B cells produced the opposite effect (Figure 5D).

Figure 5
figure 5

Phenotypic effects of HIF siRNA. Effect of HIF siRNA on (a) cell number, (b) apoptosis and (c, d) proliferation under hypoxia in ES and OS cell lines. Bar graph data is normalised to the relevant scrambled siRNA control (white bars): *, p < 0.05; **, p < 0.01; ***, p < 0.001; NS, not significant. (c) Proliferation assessed in ES cell lines by Western blot for Ki-67. N, normoxia; H, hypoxia. Numerical values represent mean fold change in Ki-67 as assessed by densitometry:*, p < 0.05; **, p < 0.01 versus hypoxic scrambled (scr) control.

Effect of hypoxia and HIF on cell migration

ES and OS cell lines responded in an opposing manner with respect to cell migration during exposure to 18 h hypoxia. ES cell migration was inhibited by 65-80% under hypoxia, whereas OS cells showed a trend to increased hypoxic migration (Figure 6A). Despite contrasting responses to the initial stimulus, both ES and OS cells demonstrated similar responses to HIF siRNA, with targeted inhibition of HIF-2α resulting in increased cell migration under hypoxia (Figure 6B).

Figure 6
figure 6

Effect of hypoxia on migration of ES and OS cell lines. (a) Fraction closure of a scratch wound due to cell migration during 18 h normoxia (white bars) or hypoxia (grey bars). (b) Effect of siRNA targeting HIF-1α (H1) or HIF-2α (H2) on migration of ES (combined data from A673, SK-ES-1 and SK-N-MC) or OS (combined data from MG63, Saos2 and 143B) cell lines under hypoxia. N, normoxia; H, hypoxia. Bar graph data is normalised to the (a) normoxic or (b) scrambled hypoxic siRNA control: *, p < 0.05; **, p < 0.01; ***, p < 0.001; NS, not significant.

Discussion

These results provide the first detailed description of HIF induction by micro-environmental conditions relevant to primary bone tumours in Ewing's sarcoma and osteosarcoma. We have described phenotypic effects of these conditions, outlining the HIF dependence of these phenotypes and differences between the two tumour types regarding these responses.

Work in many cell types describes HIF-1α as constitutively hypoxia-regulated and HIF-2α as more cell type specific, its transcriptional activation being influenced by micro-environmental conditions other than hypoxia [30, 31]. Despite both isoforms sharing an identical core binding motif, the majority of HIF-regulated hypoxia-inducible genes are induced by HIF-1α alone or by a combination of both HIF-1α and HIF-2α [27, 31–33]. Our results in ES cells reflect this, with HIF-regulated expression of VEGF, Glut-1 and PGK (the luciferase assay HRE) being mediated solely by HIF-1α. Also in accordance with previous results [32, 34], inducible HIF-2α stimulated expression of HIF-2-specific genes, e.g. TGFα, suggesting that HIF-2α is functional in ES cells but that its effects are normally masked by HIF-1α to regulate the majority of HIF-dependent genes. In contrast, gene expression in OS cells was regulated by both HIF isoforms. Indeed, the observed shift towards relative transcriptional domination of HIF-2α places OS cells in a small group, including 786-0 renal cancer cells and the MDA-MB-435 melanoma cell line [27, 28], displaying preferential utilisation of HIF-2α.

Hypoglycaemic induction of HIF-2α is the second transcriptional feature distinguishing ES and OS cell lines. Reports of HIF pathway induction under low glucose conditions suggest a specific role for HIF-2α, for example in neuroblastoma cell lines [35]. Similarly, embryonic stem cells and pancreatic cancer cells under hypoglycaemia induced HIF target gene expression in a manner dependent on the HIFα dimerisation partner ARNT [36], but independent of HIF-1α [37]. Our results in ES cell lines also suggest a dominant role for HIF-2α in regulation of the HIF response to hypoglycaemia. Hypoxic stabilisation of HIF-1α is dependent on the presence of glucose [38], which regulates translation via effects on the mTOR/Akt pathway [39, 40]. The lack of a HIF transcriptional response to low glucose in OS cells may represent a differential sensitivity to glucose deprivation. This could be a consequence of the high levels of PAS+ glycogen present in ES cells [41] that might provide sufficient intracellular available glucose for translation of HIF under hypoglycaemic conditions.

Phenotypic effects of prolonged exposure to hypoxia and low glucose were similar in both cell types. Hypoxia did not affect cell number, despite evidence of increased apoptosis. Hypoglycaemia produced a ≥ 45% reduction in cell number, predominantly due to reduced cellular proliferation. The only previous reports on apoptosis in ES cell lines described A673 as resistant [23] and SK-N-MC as sensitive [42] to hypoxia-induced apoptosis. Our results for A673 and SK-N-MC show a 2 to 3-fold increase in apoptosis after 72 h at 0.1% O2, longer exposure to more severe hypoxia potentially explaining the increased apoptosis in A673. Hypoxia-induced apoptosis of OS cells has not previously been reported, but was generally 2-fold greater than that in ES cell lines. There is considerably less data regarding hypoglycaemia-induced apoptosis [43], suggesting that this is potentially a cell type-specific response which was largely absent from either ES or OS cell lines.

It is intriguing that both ES and OS cells were sensitive to hypoxia-induced apoptosis but that HIF siRNA had no effect on the apoptotic phenotype. In the majority of cells, hypoxia-induced apoptosis is mediated by HIF-1α [23, 44, 45]. However there is an increasing literature suggesting that, in some cell types, apoptosis induced by severe hypoxia is HIF-1α-independent [46, 47]. However, in ES in vivo expression of HIF-1α and cleaved caspase 3 localised to areas adjacent to necrotic, and presumably hypoxic, tissue and HIF-1α has been correlated with the percentage of dead cells in OS tumours [6]., This suggests some in vivo relationship between HIF and apoptosis which we have been unable to dissect in the in vitro situation.

The first phenotypic difference between ES and OS cell lines was the effect of HIF siRNA on hypoxic proliferation. Although overall proliferation rates were similar in normoxia and hypoxia in both cell types, HIF-1α siRNA enhanced proliferation 2 to 3-fold in ES cells whereas in 2/3 OS cell lines both HIF-1α and HIF-2α siRNA inhibited hypoxic proliferation. It is not uncommon for HIF-1α and HIF-2α to exert opposing phenotypic effects in different cell lines. For example, HIF-2α is necessary to maintain tumour growth in RCC4 renal carcinoma cells [48], potentially via regulation of the cell cycle regulatory proteins TGFα and cyclin D1. It may also specifically promote proliferation by enhancing activity of the c-Myc oncoprotein [49], in contrast to inhibition of cell cycle progression by HIF-1α counteracting c-Myc transcriptional activity [50]. Conversely, in SW480 colon cancer cells, HIF-1α siRNA inhibits proliferation whereas HIF-2α siRNA increases anchorage independent growth [51].

The second phenotypic difference between the two cell types encompassed hypoxic migration, which was inhibited in ES cells and maintained or increased in OS cell lines. In both cell types, however, HIF-2α siRNA stimulated hypoxic migration, suggesting an effect of HIF-2α to inhibit cell migration under hypoxia. There are many reports demonstrating migration to be dependent on either HIF-1α and/or HIF-2α in different cell types [27, 51, 52], again demonstrating the ability of HIF-1α and HIF-2α to exert distinct phenotypic effects in different cell lines.

This data implicates hypoxia as the primary micro-environmental stimulus inducing expression of HIF and downstream genes in both ES and OS cells in culture. However, phenotypic results in both cell types suggest that hypoxia does not significantly increase either cell number or migration and, in addition, enhances tumour cell apoptosis. This does not correlate with clinical reports that hypoxia [9, 16–18] and/or HIF-1α [6–8] are predictive of poor outcome in these tumours. This may be due to antagonistic effects of genes downstream of HIF, which cancel each other out when manipulating the whole pathway in vitro. For example, VEGF blockade decelerates the growth of experimental osteosarcoma [53] and Ewing's sarcoma [54], demonstrating the viablility of specific targeting of HIF downstream genes. Despite being currently unable to dissect specific mechanisms by which HIF promotes ES and OS progression in vivo, we have identified features of the HIF pathway which distinguish the two cell types. This provides a solid platform from which to advance our understanding of the role of hypoxia and HIF in the biology of primary bone sarcomas.

Conclusions

To our knowledge, this study represents the first comprehensive report of effects of hypoxia and hypoglycaemia on HIF expression, transcriptional activation and phenotype of ES and OS cells. The two cell types differed with respect to their HIF isoform-specific transcriptional response to hypoxia and hypoglycaemia. Phenotypically, ES and OS cells demonstrated opposing migratory responses to hypoxia and opposing HIF-specific effects on hypoxic proliferation. Although the specific mechanism(s) whereby HIF promotes ES and OS progression in vivo remains to be identified, this data provides a comprehensive characterisation from which to advance our understanding of HIF function in the pathobiology of primary bone sarcomas.

References

  1. Vaupel P, Mayer A: Hypoxia in cancer: significance and impact on clinical outcome. Cancer Metastasis Rev. 2007, 26: 225-239. 10.1007/s10555-007-9055-1.

    Article  CAS  PubMed  Google Scholar 

  2. Epstein AC, Gleadle JM, McNeill LA, Hewitson KS, O'Rourke J, Mole DR, Mukherji M, Metzen E, Wilson MI, Dhanda A, Tian YM, Masson N, Hamilton DL, Jaakkola P, Barstead R, Hodgkin J, Maxwell PH, Pugh CW, Schofield CJ, Ratcliffe PJ: C. elegans EGL-9 and mammalian homologs define a family of dioxygenases that regulate HIF by prolyl hydroxylation. Cell. 2001, 107: 43-54. 10.1016/S0092-8674(01)00507-4.

    Article  CAS  PubMed  Google Scholar 

  3. Bruick RK, McKnight SL: A conserved family of prolyl-4-hydroxylases that modify HIF. Science. 2001, 294: 1337-1340. 10.1126/science.1066373.

    Article  CAS  PubMed  Google Scholar 

  4. Harris AL: Hypoxia--a key regulatory factor in tumour growth. Nat Rev Cancer. 2002, 2: 38-47. 10.1038/nrc704.

    Article  CAS  PubMed  Google Scholar 

  5. Maynard MA, Ohh M: The role of hypoxia-inducible factors in cancer. Cell Mol Life Sci. 2007, 64: 2170-2180. 10.1007/s00018-007-7082-2.

    Article  CAS  PubMed  Google Scholar 

  6. Yang QC, Zeng BF, Dong Y, Shi ZM, Jiang ZM, Huang J: Overexpression of Hypoxia-Inducible Factor-1{alpha} in Human Osteosarcoma: Correlation with Clinicopathological Parameters and Survival Outcome. Jap J Clin Oncol. 2007, 37: 127-134. 10.1093/jjco/hyl137.

    Article  Google Scholar 

  7. Mizobuchi H, Garcia-Castellano JM, Philip S, Healey JH, Gorlick R: Hypoxia markers in human osteosarcoma: An exploratory study. Clin Orthopaed Rel Res. 2008, 466: 2052-2059. 10.1007/s11999-008-0328-y.

    Article  Google Scholar 

  8. Chen WL, Feng HJ, Li HG: Expression and significance of hypoxemia-inducible factor-1 alpha in osteosarcoma of the jaws. Oral Surg Oral Med Oral Pathol Oral Radiol Endodontol. 2008, 106: 254-257. 10.1016/j.tripleo.2008.01.029.

    Article  Google Scholar 

  9. van der Schaft DW, Hillen F, Pauwels P, Kirschmann DA, Castermans K, Egbrink MG, Tran MG, Sciot R, Hauben E, Hogendoorn PC, Delattre O, Maxwell PH, Hendrix MJ, Griffioen AW: Tumor cell plasticity in Ewing sarcoma an alternative circulatory system stimulated by hypoxia. Cancer Res. 2005, 65: 11520-11528. 10.1158/0008-5472.CAN-05-2468.

    Article  CAS  PubMed  Google Scholar 

  10. Kubo T, Sugita T, Shimose S, Matsuo T, Arihiro K, Ochi M: Expression of hypoxia-inducible factor-1 alpha and its relationship to tumour angiogenesis and cell proliferation in cartilage tumours. J Bone Joint Surg-Br Vol. 2008, 90B: 364-370. 10.1302/0301-620X.90B3.19806.

    Article  Google Scholar 

  11. Knowles HJ, Athanasou NA: Hypoxia-inducible factor is expressed in giant cell tumour of bone and mediates paracrine effects of hypoxia on monocyte-osteoclast differentiation via induction of VEGF. J Pathol. 2008, 215: 56-66. 10.1002/path.2319.

    Article  CAS  PubMed  Google Scholar 

  12. Holzer G, Obermair A, Koschat M, Preyer O, Kotz R, Trieb K: Concentration of vascular endothelial growth factor (VEGF) in the serum of patients with malignant bone tumors. Med Pediat Oncol. 2001, 36: 601-604. 10.1002/mpo.1136.

    Article  CAS  Google Scholar 

  13. Kreuter M, Paulussen M, Boeckeler J, Gerss J, Buerger H, Liebscher C, Kessler T, Jurgens H, Berdel WE, Mesters RM: Clinical significance of vascular endothelial growth factor - a expression in Ewing's sarcoma. Eur J Cancer. 2006, 42: 1904-1911. 10.1016/j.ejca.2006.01.063.

    Article  CAS  PubMed  Google Scholar 

  14. Kaya M, Wada T, Akatsuka T, Kawaguchi S, Nagoya S, Shindoh M, Higashino F, Mezawa F, Okada F, Ishii S: Vascular endothelial growth factor expression in untreated osteosarcoma is predictive of pulmonary metastasis and poor prognosis. Clin Cancer Res. 2000, 6: 572-577.

    CAS  PubMed  Google Scholar 

  15. Kaya M, Wada T, Kawaguchi S, Nagoya S, Yamashita T, Abe Y, Hiraga H, Isu K, Shindoh M, Higashino F, Okada F, Tada M, Yamawaki S, Ishii S: Increased pre-therapeutic serum vascular endothelial growth factor in patients with early clinical relapse of osteosarcoma. Brit J Cancer. 2002, 86: 864-869. 10.1038/sj.bjc.6600201.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Dunst J, Ahrens S, Paulussen M, Burdach S, Jurgens H: Prognostic impact of tumor perfusion in MR-imaging studies in Ewing tumors. Strahlenther Onkol. 2001, 177: 153-159.

    Article  CAS  PubMed  Google Scholar 

  17. Jenkin RD, Al-Fawaz I, Al-Shabanah M, Allam A, Ayas M, Khafaga Y, Memon M, Rifai S, Schultz H, Younge D: Localised Ewing Sarcoma/PNET of bone - Prognostic factors and international data comparison. Med Pediat Oncol. 2002, 39: 586-593. 10.1002/mpo.10212.

    Article  Google Scholar 

  18. Kavalar R, Marinsek ZP, Jereb B, Cagran B, Golouh R: Prognostic value of CD99, CD117, p53 and bcl-2 in Ewing sarcoma family tumours. Eur J Oncol. 2007, 12: 243-253.

    Google Scholar 

  19. Mayes PA, Campbell L, Ricci MS, Plastaras JP, Dicker DT, El-Deiry WS: Modulation of TRAIL-induced tumor cell apoptosis in a hypoxic environment. Cancer Biol Ther. 2005, 4: 1068-1074. 10.4161/cbt.4.10.2255.

    Article  CAS  PubMed  Google Scholar 

  20. Kim HH, Lee SE, Chung WJ, Choi Y, Kwack K, Kim SW, Kim MS, Park H, Lee ZH: Stabilization of hypoxia-inducible factor-1 alpha is involved in the hypoxic stimuli-induced expression of vascular endothelial growth factor in osteoblastic cells. Cytokine. 2002, 17: 14-27. 10.1006/cyto.2001.0985.

    Article  CAS  PubMed  Google Scholar 

  21. Doege K, Heine S, Jensen I, Jelkmann W, Metzen E: Inhibition of mitochondrial respiration elevates oxygen concentration but leaves regulation of hypoxia-inducible factor (HIF) intact. Blood. 2005, 106: 2311-2317. 10.1182/blood-2005-03-1138.

    Article  CAS  PubMed  Google Scholar 

  22. Akeno N, Czyzyk-Krzeska MF, Gross TS, Clemens TL: Hypoxia induces vascular endothelial growth factor gene transcription in human osteoblast-like cells through the hypoxia-inducible factor-2alpha. Endocrinol. 2001, 142: 959-962. 10.1210/en.142.2.959.

    CAS  Google Scholar 

  23. Kilic M, Kasperczyk H, Fulda S, Debatin KM: Role of hypoxia inducible factor-1 alpha in modulation of apoptosis resistance. Oncogene. 2007, 26: 2027-2038. 10.1038/sj.onc.1210008.

    Article  CAS  PubMed  Google Scholar 

  24. Yeo EJ, Chun YS, Cho YS, Kim JH, Lee JC, Kim MS, Park JW: YGI: A potential anticancer drug targeting hypoxia-inducible factor 1. J Natl Cancer Inst. 2003, 95: 516-525. 10.1093/jnci/95.7.516.

    Article  CAS  PubMed  Google Scholar 

  25. Guan H, Jia SF, Zhou Z, Stewart J, Kleinerman ES: Herceptin down-regulates HER-2/neu and vascular endothelial growth factor expression and enhances taxol-induced cytotoxicity of human Ewing's sarcoma cells in vitro and in vivo. Clin Cancer Res. 2005, 11: 2008-2017. 10.1158/1078-0432.CCR-04-0777.

    Article  CAS  PubMed  Google Scholar 

  26. Ottaviano L, Schaefer KL, Gajewski M, Huckenbeck W, Baldus S, Rogel U, Mackintosh C, de Alava E, Myklebost O, Kresse SH, Meza-Zepeda LA, Serra M, Cleton-Jansen AM, Hogendoorn PC, Buerger H, Aigner T, Gabbert HE, Poremba C: Molecular characterization of commonly used cell lines for bone tumor research: a trans-European EuroBoNet effort. Genes Chromosomes Cancer. 49: 40-51. 10.1002/gcc.20717.

  27. Sowter HM, Raval RR, Moore JW, Ratcliffe PJ, Harris AL: Predominant role of hypoxia-inducible transcription factor (Hif)-1alpha versus Hif-2alpha in regulation of the transcriptional response to hypoxia. Cancer Res. 2003, 63: 6130-6134.

    CAS  PubMed  Google Scholar 

  28. Raval RR, Lau KW, Tran MGB, Sowter HM, Mandriota SJ, Li JL, Pugh CW, Maxwell PH, Harris AL, Ratcliffe PJ: Contrasting properties of hypoxia-inducible factor 1 (HIF-1) and HIF-2 in von Hippel-Lindau-associated renal cell carcinoma. Mol Cell Biol. 2005, 25: 5675-5686. 10.1128/MCB.25.13.5675-5686.2005.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Covello KL, Simon MC, Keith B: Targeted replacement of hypoxia-inducible factor-1alpha by a hypoxia-inducible factor-2alpha knock-in allele promotes tumor growth. Cancer Res. 2005, 65: 2277-2286. 10.1158/0008-5472.CAN-04-3246.

    Article  CAS  PubMed  Google Scholar 

  30. Qing GL, Simon MC: Hypoxia inducible factor-2 alpha: a critical mediator of aggressive tumor phenotypes. Curr Opin Genet Dev. 2009, 19: 60-66. 10.1016/j.gde.2008.12.001.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Warnecke C, Weidemann A, Volke M, Schietke R, Wu XQ, Knaup KX, Hackenbeck T, Bernhardt W, Willam C, Eckardt KU, Wiesener MS: The specific contribution of hypoxia-inducible factor-2 alpha to hypoxic gene expression in vitro is limited and modulated by cell type-specific and exogenous factors. Exp Cell Res. 2008, 314: 2016-2027. 10.1016/j.yexcr.2008.03.003.

    Article  CAS  PubMed  Google Scholar 

  32. Warnecke C, Zaborowska Z, Kurreck J, Erdmann VA, Frei U, Wiesener M, Eckardt KU: Differentiating the functional role of hypoxia-inducible factor (HIF)-1 alpha and HIF-2 alpha (EPAS-1) by the use of RNA interference: erythropoietin is a HIF-2 alpha target gene in Hep3B and Kelly cells. Faseb J. 2004, 18: 1462-

    CAS  PubMed  Google Scholar 

  33. Mole DR, Blancher C, Copley RR, Pollard PJ, Gleadle JM, Ragoussis J, Ratcliffe PJ: Genome-wide Association of Hypoxia-inducible Factor (HIF)-1 alpha and HIF-2 alpha DNA Binding with Expression Profiling of Hypoxia-inducible Transcripts. J Biol Chem. 2009, 284: 16767-16775. 10.1074/jbc.M901790200.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Batra S, Perelman N, Luck LR, Shimada H, Malik P: Pediatric tumor cells express erythropoietin and a functional erythropoietin receptor that promotes angiogenesis and tumor cell survival. Lab Invest. 2003, 83: 1477-1487. 10.1097/01.LAB.0000090156.94795.48.

    Article  CAS  PubMed  Google Scholar 

  35. Nilsson H, Jogi A, Beckman S, Harris AL, Poellinger L, Pahlman S: HIF-2 alpha expression in human fetal paraganglia and neuroblastoma: relation to sympathetic differentiation glucose deficiency and hypoxia. Exp Cell Res. 2005, 303: 447-456. 10.1016/j.yexcr.2004.10.003.

    Article  CAS  PubMed  Google Scholar 

  36. Maltepe E, Schmidt JV, Baunoch D, Bradfield CA, Simon MC: Abnormal angiogenesis and responses to glucose and oxygen deprivation in mice lacking the protein ARNT. Nature. 1997, 386: 403-407. 10.1038/386403a0.

    Article  CAS  PubMed  Google Scholar 

  37. Natsuizaka M, Ozasa M, Darmanin S, Miyamoto M, Kondo S, Kamada S, Shindoh M, Higashino F, Suhara W, Koide H, Aita K, Nakagawa K, Kondo T, Asaka M, Okada F, Kobayashi M: Synergistic up-regulation of Hexokinase-2, glucose transporters and angiogenic factors in pancreatic cancer cells by glucose deprivation and hypoxia. Exp Cell Res. 2007, 313: 3337-3348. 10.1016/j.yexcr.2007.06.013.

    Article  CAS  PubMed  Google Scholar 

  38. Vordermark D, Kraft P, Katzer A, Bolling T, Willner J, Flentje M: Glucose requirement for hypoxic accumulation of hypoxia-inducible factor-1 alpha (HIF-1 alpha). Cancer Lett. 2005, 230: 122-133. 10.1016/j.canlet.2004.12.040.

    Article  CAS  PubMed  Google Scholar 

  39. Kwon SJ, Lee YJ: Effect of low glutamine/glucose on hypoxia-induced elevation of hypoxia-inducible factor-1 alpha in human pancreatic cancer MiaPaCa-2 and human prostatic cancer DU-145 cells. Clin Cancer Res. 2005, 11: 4694-4700. 10.1158/1078-0432.CCR-04-2530.

    Article  CAS  PubMed  Google Scholar 

  40. Harada H, Itasaka S, Kizaka-Kondoh S, Shibuya K, Morinibu A, Shinomiya K, Hiraoka M: The Akt/mTOR Pathway Assures the Synthesis of HIF-1 alpha Protein in a Glucose- and Reoxygenation-dependent Manner in Irradiated Tumors. J Biol Chem. 2009, 284: 5332-5342. 10.1074/jbc.M806653200.

    Article  CAS  PubMed  Google Scholar 

  41. Kissane JM, Askin FB, Foulkes M, Stratton LB, Shirley SF: Ewing's sarcoma of bone: clinicopathologic aspects of 303 cases from the Intergroup Ewing's Sarcoma Study. Hum Pathol. 1983, 14: 773-779. 10.1016/S0046-8177(83)80300-1.

    Article  CAS  PubMed  Google Scholar 

  42. Araya R, Uehara T, Nomura Y: Hypoxia induces apoptosis in human neuroblastoma SK-N-MC cells by caspase activation accompanying cytochrome c release from mitochondria. FEBS letters. 1998, 439: 168-172. 10.1016/S0014-5793(98)01363-5.

    Article  CAS  PubMed  Google Scholar 

  43. Brusselmans K, Bono F, Maxwell P, Dor Y, Dewerchin M, Collen D, Herbert JM, Carmeliet P: Hypoxia-inducible factor-2 alpha (HIF-2 alpha) is involved in the apoptotic response to hypoglycemia but not to hypoxia. J Biol Chem. 2001, 276: 39192-39196. 10.1074/jbc.C100428200.

    Article  CAS  PubMed  Google Scholar 

  44. Carmeliet P, Dor Y, Herbert JM, Fukumura D, Brusselmans K, Dewerchin M, Neeman M, Bono F, Abramovitch R, Maxwell P, Koch CJ, Ratcliffe P, Moons L, Jain RK, Collen D, Keshet E: Role of HIF-1 alpha or in hypoxia-mediated apoptosis cell proliferation and tumour angiogenesis. Nature. 1998, 394: 485-490. 10.1038/28867.

    Article  CAS  PubMed  Google Scholar 

  45. Halterman MW, Miller CC, Federoff HJ: Hypoxia-inducible factor-1 alpha mediates hypoxia-induced delayed neuronal death that involves p53. J Neurosci. 1999, 19: 6818-6824.

    CAS  PubMed  Google Scholar 

  46. Papandreou I, Krishna C, Kaper F, Cai DL, Giaccia AJ, Denko NC: Anoxia is necessary for tumor cell toxicity caused by a low-oxygen environment. Cancer Res. 2005, 65: 3171-3178.

    CAS  PubMed  Google Scholar 

  47. Brunelle JK, Shroff EH, Perlman H, Strasser A, Moraes CT, Flavell RA, Danial NN, Keith B, Thompson CB, Chandel NS: Loss of Mcl-1 protein and inhibition of electron transport chain together induce anoxic cell death. Mol Cell Biol. 2007, 27: 1222-1235. 10.1128/MCB.01535-06.

    Article  CAS  PubMed  Google Scholar 

  48. Kondo K, Kim WY, Lechpammer M, Kaelin WG: Inhibition of HIF2 alpha is sufficient to suppress pVHL-defective tumor growth. Plos Biol. 2003, 1: 439-444. 10.1371/journal.pbio.0000083.

    Article  CAS  Google Scholar 

  49. Gordan JD, Bertout JA, Hu CJ, Diehl JA, Simon MC: HIF-2alpha promotes hypoxic cell proliferation by enhancing c-myc transcriptional activity. Cancer Cell. 2007, 11: 335-347. 10.1016/j.ccr.2007.02.006.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Koshiji M, Kageyama Y, Pete EA, Horikawa I, Barrett JC, Huang LE: HIF-1 alpha induces cell cycle arrest by functionally counteracting Myc. Embo J. 2004, 23: 1949-1956. 10.1038/sj.emboj.7600196.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Imamura T, Kikuchi H, Herraiz MT, Park DY, Mizukami Y, Mino-Kenduson M, Lynch MP, Rueda BR, Benita Y, Xavier RJ, Chung DC: HIF-1 alpha and HIF-2 alpha have divergent roles in colon cancer. Int J Cancer. 2009, 124: 763-771. 10.1002/ijc.24032.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Ben-Shoshan J, Schwartz S, Luboshits G, Maysel-Auslender S, Barzelay A, Polak-Charcon S, Tzahor E, Barshack I, Barak A, Levkovitch-Verbin H, Keren G, George J: Constitutive Expression of HIF-1 alpha and HIF-2 alpha in Bone Marrow Stromal Cells Differentially Promotes Their Proangiogenic Properties. Stem Cells. 2008, 26: 2634-2643. 10.1634/stemcells.2008-0369.

    Article  CAS  PubMed  Google Scholar 

  53. Yin DZ, Jia TH, Gong WM, Yu HY, Wooley PH, Mott MP, Yang SY: VEGF blockade decelerates the growth of a murine experimental osteosarcoma. Int J Oncol. 2008, 33: 253-259.

    PubMed  Google Scholar 

  54. Guan H, Zhou ZC, Wang H, Jia SF, Liu WB, Kleinerman ES: A small interfering RNA targeting vascular endothelial growth factor inhibits Ewing's sarcoma growth in a xenograft mouse model. Clin Cancer Res. 2005, 11: 2662-2669. 10.1158/1078-0432.CCR-04-1206.

    Article  CAS  PubMed  Google Scholar 

Pre-publication history

Download references

Acknowledgements

This study was funded by EuroBoNeT, a European Commission-granted European Network of Excellence for studying the pathology and genetics of bone tumours (LSHCCT-2006-018 814). HJK and NAA are additionally supported by the Bone Cancer Research Trust (BCRT/06/08) and acknowledge support from the Oxford NIHR BRU. UD is also supported by Deutsche Krebshilfe: 50-2551 Jü3 und 50-2551-Jü4, the Federal Ministry of Education and Research Germany, BMBF (TranSaRNet) and Deutsches Zentrum für Luft und Raumfahrt e.V.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Helen J Knowles.

Additional information

Competing interests

The authors declare that they have no competing interests.

Authors' contributions

HJK designed the study, performed the experimental work and wrote the manuscript. KLS constructed the ES tissue microarray. UD collected samples and patient information for the tissue microarray. NAA conceived the study, contributed to the study design and writing of the manuscript and provided additional ES tissue sections. All authors read and approved the final manuscript.

Authors’ original submitted files for images

Rights and permissions

This article is published under license to BioMed Central Ltd. This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

Knowles, H.J., Schaefer, KL., Dirksen, U. et al. Hypoxia and hypoglycaemia in Ewing's sarcoma and osteosarcoma: regulation and phenotypic effects of Hypoxia-Inducible Factor. BMC Cancer 10, 372 (2010). https://doi.org/10.1186/1471-2407-10-372

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1471-2407-10-372

Keywords