Runx2 expression: A mesenchymal stem marker for cancer

  • Authors:
    • Maria Teresa Valenti
    • Paola Serafini
    • Giulio Innamorati
    • Anna Gili
    • Samuele Cheri
    • Claudio Bassi
    • Luca Dalle Carbonare
  • View Affiliations

  • Published online on: September 23, 2016     https://doi.org/10.3892/ol.2016.5182
  • Pages: 4167-4172
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

The transcription factor runt-related transcription factor 2 (Runx2) is a master gene implicated in the osteogenic differentiation of mesenchymal stem cells, and thus serves a determinant function in bone remodelling and skeletal integrity. Various signalling pathways regulate Runx2 abundance, which requires a number of molecules to finely modulate its expression. Furthermore, this gene may be ectopically‑expressed in cancer cells. Recent studies have reported the involvement of Runx2 in cell proliferation, epithelial‑mesenchymal transition, apoptosis and metastatic processes, suggesting it may represent a useful therapeutic target in cancer treatment. However, studies evaluating this gene as a cancer marker are lacking. In the present study, Runx2 expression was analysed in 11 different cancer cell lines not derived from bone tumour. In addition, the presence of Runx2‑related cell‑free RNA was examined in the peripheral blood of 41 patients affected by different forms of tumours. The results demonstrated high expression levels of Runx2 in the cancer cell lines and identified the presence of Runx2‑related cell‑free RNA in the peripheral blood of patients with cancer. As compared with normal individuals, the expression level was increased by 14.2‑fold in patients with bone metastases and by 4.01‑fold in patients without metastases. The results of the present study therefore opens up the possibility to exploit Runx2 expression as a cancer biomarker allowing the use of minimally invasive approaches for diagnosis and follow-up.

Introduction

The osteogenic differentiation process of mesenchymal stem cells involves either systemic hormones or specific local molecules, including transforming growth factor-β 1/2 (TGF-β), fibroblast growth factor-2 (FGF-2), bone morphogenic proteins (BMPs), insulin-like growth factor (IGF), prostaglandins, vascular endothelial growth factors (VEGFs) and the Wnt/β-catenin pathway (1). As a result, intracellular signalling promotes the expression of transcription factors. Among these, runt-related transcription factor 2 (Runx2) serves a pivotal role and it is considered a master gene for osteogenic differentiation (1). Runx2 induces the expression of specific downstream genes, including collagen type I, bone alkaline phosphatase, osteopontin and osteocalcin (2), and it is essential for terminal chondrocyte differentiation (3). Runx2 knock-out mice are affected by cleidocranial dysplasia syndrome (3), while Runx2 overexpression in mice impairs mineralization, suggesting that this gene affects bone formation in different ways (4). A previous study demonstrated that the expression of Runx2 in circulating mesenchymal stem cells was lower in patients with osteoporosis when compared with normal donors (5). Runx2 expression is modulated by several regulatory pathways. Important negative regulators include histone deacetylases (HDACs), in particular HDAC3, HDAC4, HDAC5, HDAC6 and HDAC7 (6). Twist proteins (7), activator protein 1, transcription factor 4 and osterix are additional regulators of Runx2 expression (2). Furthermore, it has been demonstrated that Runx2 function may be downmodulated by microRNA (miR) action, in particular miR-3960 (8), and phosphorylation induced by the extracellular signal-regulated kinase/mitogen-activated protein kinase pathway results in Runx2 activation (9). The involvement of Runx2 in the oncogenic process has been recently suggested to occur in human osteosarcoma (10), in addition to other forms of malignancy such as pancreatic and thyroid cancer, and increased expression correlates with a poor prognosis (11,12).

Epithelial-mesenchymal transition (EMT) is involved in carcinogenesis and promotes metastatic spreading (1315). Following its recognition as a regulator gene in transformed epithelial cells in breast, lung and thyroid carcinoma (1315), it has been suggested that Runx2 may promote breast cancer metastasis by EMT (13). The cancer caused by EMT is a consequence of complicated reprogramming process involving differentiation, epigenetics and metabolic balance disruption (16). In this scenario, Runx2 has been identified as a regulator gene of transformed epithelial cells in breast, lung and thyroid carcinoma (1315), and it has been suggested that this gene promotes breast cancer metastasis via EMT (13).

A number of researchers have focused on identifying cancer markers that may provide clinical information a less invasive way. A previous study reported that Runx2 expression was elevated in the tissue, serum and circulating cells of patients with thyroid cancer suggesting that Runx2 may serve as a useful biomarker for thyroid malignancies (17).

On the basis of these findings, the present study speculated that Runx2 may be expressed in cells derived from malignancies other than bone tumours. Therefore, the expression of this gene was analysed in pancreatic, melanoma, breast and prostate cancer cell lines. In addition, in order to evaluate potential applications in oncological malignancies, Runx2 cell-free RNA was examined in sera obtained from patients affected by various forms of cancer.

Materials and methods

Cell culture

A total of 4 pancreatic, 2 breast, 3 prostate and 2 bone human cancer cell lines, purchased by American Type Culture Collection (Rockville, MD, USA), were used in the present study (Table I). Table I specified the previous applications of these cell lines studies (1828). The pancreatic cancer cell lines were cultured in RPMI 1640 (Sigma-Aldrich; Merck Millipore, Darmstadt, Germany) with 10% foetal bovine serum (FBS) (Sigma-Aldrich; Merck Millipore), whereas the breast, prostate and bone cell lines were cultured as previously described (2933). For cell synchronization, cell cycles were arrested at G1 phase by adding 400 mM mimosine (Sigma-Aldrich; Merck Millipore) for 24 h as previously described (34). Cells subsequently underwent three washes with PBS (Sigma-Aldrich; Merck Millipore) and were cultured in serum-free RPMI 1640 medium for 3 days. Finally, cells were cultured in fresh RPMI 1640 medium with 10% FBS (plus 2 mM L-glutamine and penicillin/streptomycin) until they reached 70% confluence. Adherent cells and supernatants for each cell line were harvested to perform expression analyses. For each cell line, three different cultures were tested.

Table I.

Cancer cell lines.

Table I.

Cancer cell lines.

Author, yearCell lineSourceTumourRefs.
Morgan et al, 1980Colo357MetastaticPancreatic(18)
Kim et al, 1989HPAFMetastaticPancreatic(19)
Lieber et al, 1975Panc1PrimaryPancreatic(20)
Parekh et al, 1994BONMetastaticPancreatic(21)
Soule et al, 1973T47DMetastaticBreast(22)
Keydar et al, 1979MCF7MetastaticBreast(23)
Stone et al, 1978DU145MetastaticProstatic(24)
Tai et al, 2011PC3PrimaryProstatic(25)
Horoszewicz et al, 1983LNCaPMetastaticProstatic(26)
Niforou et al, 2008U2OSPrimaryOsteosarcoma(27)
Billiau et al, 1977MG63PrimaryOsteosarcoma(28)
Patients

Characteristics of the population analysed are presented in Table II. A total of 41 patients with cancer were positively diagnosed from 2010 to 2013 by pathologists (Pancreas Institute; Integrated University Hospital of Verona, Verona, Italy) prior to providing blood samples, and 41 age-matched donors, who were hospitalized in Clinic of Internal Medicine, Integrated University Hospital of Verona for cardiovascular or metabolic diseases, were recruited as controls. Bone metastases were present in 17 patients. All subjects had provided written informed consent and the study was approved by the local Institutional Ethics Committee of the Integrated University Hospital of Verona.

Table II.

Characteristics of the study population.

Table II.

Characteristics of the study population.

NGenderAge, yearsDiagnosisTNM
  1M65Neuroendocrine adenocarcinomaTxN1M1
  2M80Intestinal adenocarcinomaT1-2N0M0
  3M71 HepatocarcinomaT3N1M0
  4M55Prostatic adenocarcinoma   T2N0M0
  5M83Prostatic adenocarcinomaT3N0M1
  6M87Prostatic adenocarcinomaT4N1M0
  7M66Prostatic adenocarcinomaT2N0M0
  8M67Kidney adenocarcinomaT4N0M0
  9M73Intestinal adenocarcinomaTxN0M0
10M94Gastric adenocarcinomaT3N0M0
11M81Gastric adenocarcinomaT3N2M0
12M70Lung carcinomaT1N1M1
13M70Mesenchymal cancerT4NxM1
14M81Prostatic adenocarcinomaT1N1M1
15M92Breast carcinomaT2N1M1
16M60Intestinal adenocarcinomaT1-2N0M0
17M75Pancreatic adenocarcinomaT3N1M1
18M60Pancreatic adenocarcinomaT3N1M1
19M64Pancreatic adenocarcinomaT3N0M0
20M67Pancreatic adenocarcinomaT1N0M0
21M78Bladder carcinomaT1N0M0
22F87 HepatocarcinomaT3N0M0
23F72Intestinal adenocarcinomaT1N0M0
24F27Adrenal carcinomaT3-4N1M1
25F82Intestinal adenocarcinomaTxN0M0
26F82Lung carcinomaT2N0M0
27F52Esophageal carcinomaT4N1M1
28F68Ovarian carcinomaT3N2M1
29F80Breast carcinomaT2N1M1
30F86Breast carcinomaT0N1M1
31F75Lung carcinomaT1NxM0
32F81Pancreatic adenocarcinomaT3N0M0
33F71Bladder carcinomaT3aN1M0
34F62Pancreatic adenocarcinomaT3N1M1
35F71Pancreatic adenocarcinomaT4N0M0
36F75Pancreatic adenocarcinomaT3N0M0
37F70Pancreatic adenocarcinomaT1N0M0
38F49Pancreatic adenocarcinomaT1N0M0
39M78Prostatic adenocarcinomaT2N0M1
40M80Prostatic adenocarcinomaT2N0M1
41M75Prostatic adenocarcinomaT3N1M1

[i] TNM, tumour-node-metastasis.

Serum preparation

Serum samples were obtained following three rounds of centrifugation (800 × g, 1,000 × g and 1,500 × g at 4°C) of collected blood to keep lymphocyte contamination to a minimum as previously described (35).

RNA extraction and reverse transcription

RNA from cancer cell lines was extracted using the RNeasy® Mini kit (Qiagen, Hilden, Germany), and RNA extraction from sera and culture supernatants was performed using the QIAamp® UltraSens® Virus kit (Qiagen) with DNAse I treatment according to the manufacturer's protocol. First-strand cDNA was generated using the High-Capacity cDNA Archive kit with random hexamers (Applied Biosystems; Thermo Fisher Scientific, Inc., Waltham, MA, USA) according to the manufacturer's protocol. cDNA products were stored at −80°C until use.

Quantitative polymerase chain reaction (qPCR)

PCR was performed in a total volume of 50 µl containing 1X Taqman Universal PCR Master mix, No AmpErase® UNG and 5 µl cDNA. The real time amplifications included 10 min at 95°C, followed by 40 cycles at 95°C for 15 sec and at 60°C for 1 min. Predesigned, gene-specific primers and a probe set for Runx2 were obtained from Assay-on-Demand™ Gene Expression products (Applied Biosystems; Thermo Fisher Scientific, Inc.). In order to normalize the results, the following three housekeeping genes were used: β-actin (structural gene), glyceraldehyde 3-phosphate dehydrogenase (GAPDH; metabolism-related gene) and β-2 microglobulin (component of major histocompatibility complex class I gene). The primer sequences were pre-designed by the supplier (Applied Biosystems; Thermo Fisher Scientific, Inc.). The relative expression levels of the Runx2 gene were calculated for each sample following normalization using the 2-∆∆Ct method for comparing differences in relative fold expression (36). The data are reported as mRNA fold expression.

Western blot analysis

Cells were lysed on ice for 45 min in a buffer containing protease inhibitor cocktail [1% IGEPAL®, 1% sodium dodecyl sulfate (SDS), 10% glycerol, 1 mM ethylenediaminetetraacetic acid, 5% b-mercaptoethanol, 1.5% Triton X-100 and 4% Protease Inhibitor Cocktail (Sigma-Aldrich; Merck Millipore)]. Cell lysates were then centrifuged (10,000 × g) for 15 min at 4°C to remove insoluble materials. Protein concentration in the supernatants was measured using the Coomassie Protein assay kit (Pierce Biotechnology, Inc., Rockford, IL, USA). Proteins (70 µg) were separated by 10% SDS-polyacrylamide gel electrophoresis and electrotransferred onto a polyvinylidene fluoride membrane. The membrane was subsequently blocked for 30 min with 3% bovine serum albumin (Sigma-Aldrich; Merck Millipore) in 0.05% Tween 20 with Tris-buffered saline (t-TBS) at room temperature. For immunodetection, blots were incubated for 2 h at room temperature on titer plate agitator with anti-Runx2 antibodies (cat no. 05-1478; dilution 1:500; clone AS110; EMD Millipore, Billerica, MA, USA). The membranes were washed three times in t-TBS, incubated at room temperature with horseradish peroxidase-conjugated anti-rabbit secondary antibodies (dilution, 1:2,500) in TBS for 1 h and washed in fresh t-TBS three times for a total of 20 min. Bands were detected using Luminata™ Forte Western HRP Substrate (Merck Millipore) and a G:BOX Chemi XX6 (Syngene, Frederick, MD, USA).

Statistical analysis

Results are expressed as the mean ± standard error. The Wilcoxon signed-ranked test was used for non-parametric data. Analysis of variance followed by Bonferroni correction was performed as a post-hoc analysis and the results are expressed as the mean ± standard error of the mean. P<0.05 was considered to indicate a statistically significant difference. Analyses were applied to experiments carried out at least three times, and statistical analyses were performed using SPSS v16.0 (SPSS, Inc., Chicago, IL, USA).

Results

Runx2 expression in cancer cell lines

Runx2 gene expression was analysed in adherent cells and in culture supernatants, and the MG63 cell line was used as a calibrator (fold of expression). It was observed that Runx2 mRNA was expressed in adherent cells and supernatants of the cancer cell lines, although expression was largely varied across the different cell types (Fig. 1A). In order to analyse the expression of Runx2 protein in adherent cells, immunoblotting using anti-Runx2 antibodies was performed. The results demonstrated that the protein was also expressed in all cell lines (Fig. 1B).

Runx2 gene expression in patients with cancer

The expression data of patients with cancer was reported as fold of expression in respect to a calibrator (40 normal donors). Patients with cancer and normal donors each expressed Runx2 mRNA; however, their expression levels were different. Notably, the expression of Runx2 in the patients with cancer was 8.74 (±3.5)-fold higher than the normal donors (P<0.01; Fig. 2A). In addition, Runx2 mRNA expression in patients with bone metastases was higher than in patients without metastases. Runx2 expression in patients with metastases was 14.12 (±4.2)-fold higher than the normal donors (P<0.01), whereas in patients without metastases Runx2 expression was 4.01 (±2.01)-fold higher than the normal donors (P<0.05) (Fig. 2B).

Discussion

In order to establish less invasive methods for the diagnosis and follow-up of patients with cancer, research has aimed to identify cell-free RNA encoding for genes upregulated in cancer malignancies (17,35,37). Previous studies primarily focused on osteosarcoma and metastatic breast and prostate cancer have linked Runx2 to neoplastic transformation (3841). The present study enrolled patients affected by various types of tumours, including pancreatic, prostatic, intestinal, lung, breast, gastric, liver, neuroendocrine, kidney, mesenchymal, adrenal gland, oesophageal and ovarian cancer. Notably, the results of the current study demonstrated an increase in Runx2 circulating mRNA in multiple forms of cancer, thus opening the possibility to investigate it as a relatively comprehensive biomarker.

The Runx gene family is comprised of three related transcription factors, which are involved in the differentiation of multiple haematopoietic lineages (Runx1), cartilage and bone (Runx2) and epithelial tissues (Runx3). However, all three genes are implicated in cancer by promoting (Runx1 and Runx2) or suppressing (Runx3) neoplastic transformation (42). Multiple mechanisms contribute to Runx2 functional modulation, including post-translational modification, in addition to protein-protein interaction and direct stimulation (11). Several hypotheses, such as the involvement of integrin alpha5 (39), p53 (43) or microRNA-205 (40) have been put forward to describe the molecular process of Runx2 in carcinogenesis. In osteosarcoma, loss of p53 upregulates Runx2 expression (43); this cause-effect relationship may explain Runx2 ectopic expression in various forms of cancer.

P53 and Runx2 have been demonstrated to be part of the regulatory network controlling EMT (44). P53 controls miRNAs, major EMT-related signalling pathways (TGF-β, Wnt, IGF, and signal transducer and activator of transcription), and EMT-associated transcription factors that promote a chemoresistant phenotype, invasion and loss of cell polarity (44). The direct involvement of Runx2 in cancer was demonstrated by downmodulation experiments in thyroid carcinoma cells (15) and upregulation experiments in breast cancer (45). EMT represents an early event of tumour progression and is mediated by well-characterized transcription factors (e.g. Snail and Twist family and helix-loop-helix factors) (46). The present study speculates that Runx2 participates in these events to promote invasion and metastasis in a larger number of cancer forms than previously anticipated. The data from the current study demonstrated an increase in the concentration of circulating cell-free Runx2 cell-free mRNA in patients with metastasis. In agreement with these results, Runx2 has been repeatedly identified as a regulator of bone metastases in breast and prostate cancer in previous studies (4749). Bone is particularly recurrent as a target of metastasizing cells, thus a master skeletal transcription factor like Runx2 may be extremely relevant in potentiating tumour cell invasiveness of bone marrow, among others, contributing directly to the osteolysis process (38). Further studies with a larger number of patients should be performed in order to validate the predictive value of minimally invasive tests based on Runx2 cell-free mRNA.

In conclusion, the present study demonstrated that Runx2 is expressed at high levels in osteosarcoma and expanded this finding to non-osseous cells, thus supporting the possible use of Runx2-related cell-free RNA as a cancer marker for screening purposes. In addition, this useful, less invasive method may allow clinicians to monitor the development of metastases in patients with cancer.

References

1 

Carbonare L Dalle, Innamorati G and Valenti MT: Transcription factor Runx2 and its application to bone tissue engineering. Stem Cell Rev. 8:891–897. 2012. View Article : Google Scholar : PubMed/NCBI

2 

Cohen MM Jr: Perspectives on RUNX genes: An update. Am J Med Genet A 149A. 2629–2646. 2009. View Article : Google Scholar

3 

Otto F, Thornell AP, Crompton T, Denzel A, Gilmour KC, Rosewell IR, Stamp GW, Beddington RS, Mundlos S, Olsen BR, et al: Cbfa1, a candidate gene for cleidocranial dysplasia syndrome, is essential for osteoblast differentiation and bone development. Cell. 89:765–771. 1997. View Article : Google Scholar : PubMed/NCBI

4 

Otto F, Lübbert M and Stock M: Upstream and downstream targets of RUNX proteins. J Cell Biochem. 89:9–18. 2003. View Article : Google Scholar : PubMed/NCBI

5 

Valenti MT, Garbin U, Pasini A, Zanatta M, Stranieri C, Manfro S, Zucal C and Carbonare L Dalle: Role of ox-PAPCs in the differentiation of mesenchymal stem cells (MSCs) and Runx2 and PPARγ2 expression in MSCs-like of osteoporotic patients. PloS One. 6:e203632011. View Article : Google Scholar : PubMed/NCBI

6 

Jensen ED, Schroeder TM, Bailey J, Gopalakrishnan R and Westendorf JJ: Histone deacetylase 7 associates with Runx2 and represses its activity during osteoblast maturation in a deacetylation-independent manner. J Bone Miner Res. 23:361–372. 2008. View Article : Google Scholar : PubMed/NCBI

7 

Yousfi M, Lasmoles F and Marie PJ: TWIST inactivation reduces CBFA1/RUNX2 expression and DNA binding to the osteocalcin promoter in osteoblasts. Biochem Biophys Res Commun. 297:641–644. 2002. View Article : Google Scholar : PubMed/NCBI

8 

Hu R, Liu W, Li H, Yang L, Chen C, Xia ZY, Guo LJ, Xie H, Zhou HD, Wu XP and Luo XH: A Runx2/miR-3960/miR-2861 regulatory feedback loop during mouse osteoblast differentiation. J Biol Chem. 286:12328–12339. 2011. View Article : Google Scholar : PubMed/NCBI

9 

Ge C, Xiao G, Jiang D, Yang Q, Hatch NE, Roca H and Franceschi RT: Identification and functional characterization of ERK/MAPK phosphorylation sites in the Runx2 transcription factor. J Biol Chem. 284:32533–32543. 2009. View Article : Google Scholar : PubMed/NCBI

10 

Lau CC, Harris CP, Lu XY, Perlaky L, Gogineni S, Chintagumpala M, Hicks J, Johnson ME, Davino NA, Huvos AG, et al: Frequent amplification and rearrangement of chromosomal bands 6p12-p21 and 17p11.2 in osteosarcoma. Genes Chromosomes Cancer. 39:11–21. 2004. View Article : Google Scholar : PubMed/NCBI

11 

Kayed H, Jiang X, Keleg S, Jesnowski R, Giese T, Berger MR, Esposito I, Löhr M, Friess H and Kleeff J: Regulation and functional role of the Runt-related transcription factor-2 in pancreatic cancer. Br J Cancer. 97:1106–1115. 2007. View Article : Google Scholar : PubMed/NCBI

12 

Endo T, Ohta K and Kobayashi T: Expression and function of Cbfa-1/Runx2 in thyroid papillary carcinoma cells. J Clin Endocrinol Metab. 93:2409–2412. 2008. View Article : Google Scholar : PubMed/NCBI

13 

Owens TW, Rogers RL, Best SA, Ledger A, Mooney AM, Ferguson A, Shore P, Swarbrick A, Ormandy CJ, Simpson PT, et al: Runx2 is a novel regulator of mammary epithelial cell fate in development and breast cancer. Cancer Res. 74:5277–5286. 2014. View Article : Google Scholar : PubMed/NCBI

14 

Hsu YL, Huang MS, Yang CJ, Hung JY, Wu LY and Kuo PL: Lung tumor-associated osteoblast-derived bone morphogenetic protein-2 increased epithelial-to-mesenchymal transition of cancer by Runx2/Snail signaling pathway. J Biol Chem. 286:37335–37346. 2011. View Article : Google Scholar : PubMed/NCBI

15 

Niu DF, Kondo T, Nakazawa T, Oishi N, Kawasaki T, Mochizuki K, Yamane T and Katoh R: Transcription factor Runx2 is a regulator of epithelial-mesenchymal transition and invasion in thyroid carcinomas. Lab Invest. 92:1181–1190. 2012. View Article : Google Scholar : PubMed/NCBI

16 

Li L and Li W: Epithelial-mesenchymal transition in human cancer: Comprehensive reprogramming of metabolism, epigenetics, and differentiation. Pharmacol Ther. 150:33–46. 2015. View Article : Google Scholar : PubMed/NCBI

17 

Carbonare L Dalle, Frigo A, Francia G, Davì MV, Donatelli L, Stranieri C, Brazzarola P, Zatelli MC, Menestrina F and Valenti MT: Runx2 mRNA expression in the tissue, serum, and circulating non-hematopoietic cells of patients with thyroid cancer. J Clin Endocrinol Metab. 97:E1249–E1256. 2012. View Article : Google Scholar : PubMed/NCBI

18 

Morgan RT, Woods LK, Moore GE, Quinn LA, McGavran L and Gordon SG: Human cell line (COLO 357) of metastatic pancreatic adenocarcinoma. Int J Cancer. 25:591–598. 1980. View Article : Google Scholar : PubMed/NCBI

19 

Kim YW, Kern HF, Mullins TD, Koriwchak MJ and Metzgar RS: Characterization of clones of a human pancreatic adenocarcinoma cell line representing different stages of differentiation. Pancreas. 4:353–362. 1989. View Article : Google Scholar : PubMed/NCBI

20 

Lieber M, Mazzetta J, Nelson-Rees W, Kaplan M and Todaro G: Establishment of a continuous tumor-cell line (panc-1) from a human carcinoma of the exocrine pancreas. Int J Cancer. 15:741–747. 1975. View Article : Google Scholar : PubMed/NCBI

21 

Parekh D, Ishizuka J, Townsend CM Jr, Haber B, Beauchamp RD, Karp G, Kim SW, Rajaraman S, Greeley G Jr and Thompson JC: Characterization of a human pancreatic carcinoid in vitro: Morphology, amine and peptide storage, and secretion. Pancreas. 9:83–90. 1994. View Article : Google Scholar : PubMed/NCBI

22 

Soule HD, Vazguez J, Long A, Albert S and Brennan M: A human cell line from a pleural effusion derived from a breast carcinoma. J Natl Cancer Inst. 51:1409–1416. 1973.PubMed/NCBI

23 

Keydar I, Chen L, Karby S, Weiss FR, Delarea J, Radu M, Chaitcik S and Brenner HJ: Establishment and characterization of a cell line of human breast carcinoma origin. Eur J Cancer. 15:659–670. 1979. View Article : Google Scholar : PubMed/NCBI

24 

Stone KR, Mickey DD, Wunderli H, Mickey GH and Paulson DF: Isolation of a human prostate carcinoma cell line (DU 145). Int J Cancer. 21:274–281. 1978. View Article : Google Scholar : PubMed/NCBI

25 

Tai S, Sun Y, Squires JM, Zhang H, Oh WK, Liang CZ and Huang J: PC3 is a cell line characteristic of prostatic small cell carcinoma. Prostate. 71:1668–1679. 2011. View Article : Google Scholar : PubMed/NCBI

26 

Horoszewicz JS, Leong SS, Kawinski E, Karr JP, Rosenthal H, Chu TM, Mirand EA and Murphy GP: LNCaP model of human prostatic carcinoma. Cancer Res. 43:1809–1818. 1983.PubMed/NCBI

27 

Niforou KM, Anagnostopoulos AK, Vougas K, Kittas C, Gorgoulis VG and Tsangaris GT: The proteome profile of the human osteosarcoma U2OS cell line. Cancer Genomics Proteomics. 5:63–78. 2008.PubMed/NCBI

28 

Billiau A, Edy VG, Heremans H, Van Damme J, Desmyter J, Georgiades JA and De Somer P: Human interferon: Mass production in a newly established cell line, MG-63. Antimicrob Agents Chemother. 12:11–15. 1977. View Article : Google Scholar : PubMed/NCBI

29 

Carbonare L Dalle, Valenti MT, Bertoldo F, Fracalossi A, Balducci E, Azzarello G, Vinante O and Lo Cascio V: Amino-bisphosphonates decrease hTERT gene expression in breast cancer in vitro. Aging Clin Exp Res. 19:91–96. 2007. View Article : Google Scholar : PubMed/NCBI

30 

Valenti MT, Carbonare L Dalle, Bertoldo F, Donatelli L and Lo Cascio V: The effects on hTERT gene expression is an additional mechanism of amino-bisphosphonates in prostatic cancer cells. Eur J Pharmacol. 580:36–42. 2008. View Article : Google Scholar : PubMed/NCBI

31 

Gatta V, Drago D, Fincati K, Valenti MT, Carbonare L Dalle, Sensi SL and Zatta P: Microarray analysis on human neuroblastoma cells exposed to aluminum, β(1–42)-amyloid or the beta (1–42)-amyloid aluminum complex. PloS One. 6:e159652011. View Article : Google Scholar : PubMed/NCBI

32 

Murayama T, Kawasoe Y, Yamashita Y, Ueno Y, Minami S, Yokouchi M and Komiya S: Efficacy of the third-generation bisphosphonate risedronate alone and in combination with anticancer drugs against osteosarcoma cell lines. Anticancer Res. 28:2147–2154. 2008.PubMed/NCBI

33 

Valenti MT, Zanatta M, Donatelli L, Viviano G, Cavallini C, Scupoli MT and Carbonare L Dalle: Ascorbic acid induces either differentiation or apoptosis in MG-63 osteosarcoma lineage. Anticancer Res. 34:1617–1627. 2014.PubMed/NCBI

34 

Galindo M, Pratap J, Young DW, Hovhannisyan H, Im HJ, Choi JY, Lian JB, Stein JL, Stein GS and van Wijnen AJ: The bone-specific expression of Runx2 oscillates during the cell cycle to support a G1-related antiproliferative function in osteoblasts. J Biol Chem. 280:20274–20285. 2005. View Article : Google Scholar : PubMed/NCBI

35 

Valenti MT, Carbonare L Dalle, Donatelli L, Bertoldo F, Giovanazzi B, Caliari F and Lo Cascio V: STEAP mRNA detection in serum of patients with solid tumours. Cancer Lett. 273:122–126. 2009. View Article : Google Scholar : PubMed/NCBI

36 

Livak KJ and Schmittgen TD: Analysis of relative gene expression data using real-time quantitative PCR and the 2(−Delta Delta C (T)) method. Methods. 25:402–408. 2001. View Article : Google Scholar : PubMed/NCBI

37 

Carbonare L Dalle, Gasparetto A, Donatelli L, Dellantonio A and Valenti MT: Telomerase mRNA detection in serum of patients with prostate cancer. Urol Oncol. 31:205–210. 2013. View Article : Google Scholar : PubMed/NCBI

38 

Pratap J, Lian JB and Stein GS: Metastatic bone disease: Role of transcription factors and future targets. Bone. 48:30–36. 2011. View Article : Google Scholar : PubMed/NCBI

39 

Li XQ, Lu JT, Tan CC, Wang QS and Feng YM: RUNX2 promotes breast cancer bone metastasis by increasing integrin α5-mediated colonization. Cancer Lett. 380:78–86. 2016. View Article : Google Scholar : PubMed/NCBI

40 

Zhang C, Long F, Wan J, Hu Y and He H: MicroRNA-205 acts as a tumor suppressor in osteosarcoma via targeting RUNX2. Oncol Rep. 35:3275–3284. 2016.PubMed/NCBI

41 

Ge C, Zhao G, Li Y, Li H, Zhao X, Pannone G, Bufo P, Santoro A, Sanguedolce F, Tortorella S, et al: Role of Runx2 phosphorylation in prostate cancer and association with metastatic disease. Oncogene. 35:366–376. 2016. View Article : Google Scholar : PubMed/NCBI

42 

Blyth K, Vaillant F, Jenkins A, McDonald L, Pringle MA, Huser C, Stein T, Neil J and Cameron ER: Runx2 in normal tissues and cancer cells: A developing story. Blood Cells Mol Dis. 45:117–123. 2010. View Article : Google Scholar : PubMed/NCBI

43 

He Y, de Castro LF, Shin MH, Dubois W, Yang HH, Jiang S, Mishra PJ, Ren L, Gou H, Lal A, et al: p53 loss increases the osteogenic differentiation of bone marrow stromal cells. Stem Cells. 33:1304–1319. 2015. View Article : Google Scholar : PubMed/NCBI

44 

Engelmann D and Pützer BM: Emerging from the shade of p53 mutants: N-terminally truncated variants of the p53 family in EMT signaling and cancer progression. Sci Signal. 7:re92014. View Article : Google Scholar : PubMed/NCBI

45 

Chimge NO, Baniwal SK, Little GH, Chen YB, Kahn M, Tripathy D, Borok Z and Frenkel B: Regulation of breast cancer metastasis by Runx2 and estrogen signaling: The role of SNAI2. Breast Cancer Res. 13:R1272011. View Article : Google Scholar : PubMed/NCBI

46 

Lee JY and Kong G: Roles and epigenetic regulation of epithelial-mesenchymal transition and its transcription factors in cancer initiation and progression. Cell Mol Life Sci. Jul 26–2016.(Epub ahead of print). View Article : Google Scholar

47 

Akech J, Wixted JJ, Bedard K, van der Deen M, Hussain S, Guise TA, van Wijnen AJ, Stein JL, Languino LR, Altieri DC, et al: Runx2 association with progression of prostate cancer in patients: Mechanisms mediating bone osteolysis and osteoblastic metastatic lesions. Oncogene. 29:811–821. 2010. View Article : Google Scholar : PubMed/NCBI

48 

Leong DT, Lim J, Goh X, Pratap J, Pereira BP, Kwok HS, Nathan SS, Dobson JR, Lian JB, Ito Y, et al: Cancer-related ectopic expression of the bone-related transcription factor RUNX2 in non-osseous metastatic tumor cells is linked to cell proliferation and motility. Breast Cancer Res. 12:R892010. View Article : Google Scholar : PubMed/NCBI

49 

Baniwal SK, Khalid O, Gabet Y, Shah RR, Purcell DJ, Mav D, Kohn-Gabet AE, Shi Y, Coetzee GA and Frenkel B: Runx2 transcriptome of prostate cancer cells: Insights into invasiveness and bone metastasis. Mol Cancer. 9:2582010. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

November-2016
Volume 12 Issue 5

Print ISSN: 1792-1074
Online ISSN:1792-1082

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Valenti MT, Serafini P, Innamorati G, Gili A, Cheri S, Bassi C and Dalle Carbonare L : Runx2 expression: A mesenchymal stem marker for cancer. Oncol Lett 12: 4167-4172, 2016
APA
Valenti, M.T., Serafini, P., Innamorati, G., Gili, A., Cheri, S., Bassi, C., & Dalle Carbonare, L. . (2016). Runx2 expression: A mesenchymal stem marker for cancer. Oncology Letters, 12, 4167-4172. https://doi.org/10.3892/ol.2016.5182
MLA
Valenti, M. T., Serafini, P., Innamorati, G., Gili, A., Cheri, S., Bassi, C., Dalle Carbonare, L. ."Runx2 expression: A mesenchymal stem marker for cancer". Oncology Letters 12.5 (2016): 4167-4172.
Chicago
Valenti, M. T., Serafini, P., Innamorati, G., Gili, A., Cheri, S., Bassi, C., Dalle Carbonare, L. ."Runx2 expression: A mesenchymal stem marker for cancer". Oncology Letters 12, no. 5 (2016): 4167-4172. https://doi.org/10.3892/ol.2016.5182