Downregulation of OCLN and GAS1 in clear cell renal cell carcinoma

  • Authors:
    • André Luis Giacometti Conceição
    • Camila Tainah Da Silva
    • Rodolfo Miglioli Badial
    • Marina Curado Valsechi
    • Bruna Stuqui
    • Jéssica Domingues Gonçalves
    • Miriam Galvonas Jasiulionis
    • Marilia De Freitas Calmon
    • Paula Rahal
  • View Affiliations

  • Published online on: January 31, 2017     https://doi.org/10.3892/or.2017.5414
  • Pages: 1487-1496
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Clear cell renal cell carcinoma (ccRCC) is the most common histological subtype of kidney cancer. This carcinoma is histologically characterized by the presence of clear and abundant cytoplasm. In the present study, we sought to identify genes differentially expressed in ccRCC and build a molecular profile of this cancer. We selected genes described in the literature related to cellular differentiation and proliferation. We analyzed the gene and protein expression by quantitative PCR (qPCR) and immunohistochemistry, respectively, and examined possible epigenetic mechanisms that regulate their expression in ccRCC samples and cell lines. Occludin (OCLN) and growth arrest-specific 1 (GAS1) genes were underexpressed in ccRCC, and we report that miR-122 and miR-34a, respectively, may regulate their expression in this cancer. Furthermore, we showed by qPCR and immunohistochemistry that solute carrier family 2 member 1 (SLC2A1) was significantly overexpressed in ccRCC. The set of genes identified in the present study furthers our understanding of the molecular basis and development of ccRCC.

Introduction

Clear cell renal cell carcinoma (ccRCC) is the most common histological subtype of kidney cancer in adults (70–75% of cases) (1). Approximately 30% of early-stage patients develop metastases after surgery for localized disease (2). This carcinoma is histologically characterized by the presence of clear and abundant cytoplasm (2,3). The aggressive characteristics of renal clear cell carcinoma interfere in the efficiency of conventional treatment approaches, such as chemotherapy and radiotherapy (4).

Mutations or deletions associated with tumor suppressor genes may cause the loss or inactivation of these negative regulators. However, the loss of function may be caused by epigenetic changes (5), including acetylation, methylation, phosphorylation, ubiquitinylation, sumoylation, carbonylation and regulation of microRNAs (6). Epigenetic alterations occur at a high frequency, are reversible upon treatment with pharmacological agents and arise at defined regions within a gene (7,8). These features make epigenetics a very attractive field for cancer detection and several studies have shown multiple genes with different types of epigenetic alterations in tumor tissue (911).

Analyzing the gene expression and regulatory mechanisms in tumor samples has shown great efficiency in providing new information on signaling pathway activation of several processes, such as evasion of apoptosis, angiogenesis and metastatic potential.

The occludin (OCLN) gene, responsible for producing the protein occludin, is known to be present in all the tight junctions (TJs). TJ formation and expression has been described as a hallmark of epithelial cells and their subsequent loss from cancer cells tends to be considered as a direct effect of epithelial-mesenchymal transition (EMT) (12). Furthermore, occludin is structurally connected to the TGF-β receptor contributing to essential changes in EMT cell phenotype (13).

Growth arrest-specific 1 (GAS1) gene is associated with cell arrest and the protein is linked to the outer cell membrane by a glycosyl phosphatydylinositol. GAS1 protein binds to Hedgehog ligands for the potentiation of Hedgehog signaling (14).

Solute carrier family 2 member 1 (SLC2A1) gene encodes the solute carrier family 2 facilitated glucose transporter member 1 protein essential for cellular energy metabolism pathways (1517). This protein is able to develop bidirectional flow of glucose according to the gradient of the substrate. The overexpression of the SLC2A1 protein has been found to be involved in increased uptake of glucose in human breast adenocarcinoma cell lines (18).

In the present study, we selected genes described in the literature related to cellular differentiation and proliferation. We analyzed their expression by quantitative PCR (qPCR) and immunohistochemistry, and examined the possible epigenetic mechanisms that regulate genes in ccRCC samples and cell lines. OCLN and GAS1 were underexpressed in ccRCC and miR-122 and miR-34a may regulate their expression in this type of cancer. Furthermore, we showed by qPCR and immunohistochemistry that SLC2A1 was significantly overexpressed in ccRCC samples. These genes may be involved in the molecular biology and development of the disease and analysis of them may contribute to insights for new prognoses, treatments and understanding this type of tumor.

Materials and methods

ccRCC samples

The samples were collected from 101 patients diagnosed with ccRCC, including 45 fresh samples of ccRCC, 56 paraffin-embedded samples of ccRCC, 2 fresh-frozen histologically healthy renal tissue and 24 paraffin-embedded samples of histologically healthy renal tissues, all of which were confirmed by pathologists (Tables I and II). The samples were obtained from the Hospital de Base at Faculdade de Medicina de São José do Rio Preto (Sao Paulo, Brazil). The use of patient-derived material was approved by the Research Ethics Committee of the IBILCE-UNESP and FAMERP (Sao Paulo, Brazil) and written consent was obtained from all the patients. Tissues were obtained from patients undergoing tumor resection surgery, and a diagnosis of ccRCC was determined post-operatively using histopathology. The samples were classified according to the criteria provided by the International Union against Cancer (19).

Table I.

Epidemiological and clinicopathological characteristics of the patients diagnosed with ccRCC (fresh samples).

Table I.

Epidemiological and clinicopathological characteristics of the patients diagnosed with ccRCC (fresh samples).

SampleGenderAge (years)AlcoholSmokingTNM
  F1M58NoNoT1N1M1
  F2F39NoNoT2N0M0
  F3F71NoNoT4N0M1
  F4F64NoNoT2N0M0
  F5M46NoNoT1bN0M0
  F6M70NoNoT2N0M0
F7M70YesYesT3aN0M0
F8F58YesYesT3N3M1
F9M36NoNoT1aN0M0
F10M48YesNoT2N0M0
F11M52NoNoT3aN0M0
F12F62NoNoT1aN0M0
F13M75NoYesT1N0M0
F14F49NoNoT1bN0M0
F15F71NoNoT1N0M0
F16F79NoNoT1aN0M0
F17F29NoNoT2N0M0
F18M70NoNoT2N0M0
F19M39NoYesT2N0M0
F20M66YesNoT2N0M0
F21M77NoNoT2N0M1
F22F23NoNoT2N0M0
F23F53NoNoT2N0M0
F24M69NoNoT3N0M0
F25F60NoNoT3bN0M1
F26F79NoNoT3aN0M0
F27F59NoNoT2N0M0
F28F64NoNoT1N0M0
F29M53YesNoT2N0M0
F30M56YesNoT1N1M1
F31M54NoNoT2N0M0
F32M52YesNoT1aN0M0
F33M67YesNoT1aN0M0
F34M41NoYesT1N0M0
F35F55NoNoT3aN2M1
F36F52YesYesT3aNxMx
F37F25NoYesNI
F38M56NoNoT1bN0Mx
F39F83NoNoT1N0M0
F40M68YesYesNI
F41F71NoYesT1bNxMx
F42M73NoYesT3NxM1
F43M63YesNoT3aN0M0
F44M62YesYesNI
F45M39YesNoT2NxM0

[i] M, male; F, female; NI, no information was obtained. ccRCC, clear cell renal cell carcinoma; TNM, tumor-node-metastasis.

Table II.

Epidemiological and clinicopathological characteristics of the patients diagnosed with ccRCC (paraffin-embedded samples).

Table II.

Epidemiological and clinicopathological characteristics of the patients diagnosed with ccRCC (paraffin-embedded samples).

SampleGenderAge (years)AlcoholSmokingTNM
  P1M80NoNoT4N1M1
  P2M50NoNoT2N0M0
  P3M61NoNoT3bN0M0
  P4F69NoNoT1bN0M0
  P5M52NoYesT2N0M0
  P6M88NoYesT2N0M0
  P7M60NoNoT1bN0M0
  P8M58NoNoT1bN1M1
  P9M71NoYesT1N0M0
P10M56NoNoT3bN0M1
P11M77NoNoT2N1M1
P12F32YesNoT1aN0M0
P13M68NoNoT3bN0M0
P14M83YesYesT2N0M1
P15F65NoNoT2N0M0
P16F54NoNoT1bN0M0
P17M44NININI
P18M65NoNoT3aN0M1
P19F72NoYesT2N0M1
P20F60NoNoT2N0M0
P21M77NoNoT1bN0M0
P22M52NoNoT1bN0M0
P23M74NoNoT3aN0M1
P24M75NoNoT2N0M0
P25M78NoNoT2N0M0
P26F66NoNoT2N0M0
P27M64NoNoT1bN0M0
P28M67YesYesT2N0M0
P29M49NoNoT2N0M0
P30F58NoNoT1bN0M0
P31M84NoYesT1bN0M0
P32F48NoYesT2N0M0
P33F55NoNoT1bN0M0
P34F50NoNoT2N0M0
P35F46NoNoT1aN0M0
P36M43NoNoT2N0M0
P37F54NoYesT3aN0M0
P38F86NoYesT3N0M0
P39M51NoNoT2N0M0
P40M74NoNoT2N0M0
P41F51NoYesT1aN0M0
P42M63YesYesT3aN0M0
P43M56NoNoT2N0M0
P44F76NoYesT1bNxMx
P45M70NoYesT1aNxMx
P46M58NoNoT1bN0M0
P47M56NoNoT1bN0M0
P48M76NoNoT3N0M0
P49F61NoNoNI
P50M81YesNoNI
P51M76YesYesT2N0M0
P52F72NoNoT1N0M0
P53F79NoNoT2N0M0
P54M59NoYesT2N0M0
P55M46NoNoT2N0M1
P56F89NoYesT1N0M0

[i] M, male; F, female; NI, no information was obtained. ccRCC, clear cell renal cell carcinoma; TNM, tumor-node-metastasis.

Cell lines and treatments

The 786-O (primary clear cell adenocarcinoma) cell line was obtained from the American Type Culture Collection (ATCC; Manassas, VA, USA) and HaCaT (an immortal human keratinocyte cell line) was kindly provided by Luisa Lina Villa (Department of Radiology, Center of Translational Oncology Investigation, São Paulo State Cancer Institute, Universidade de São Paulo, Brazil). 786-O cells were cultured in RPMI-1640 medium, and HaCaT cells were cultured in Dulbeccos modified Eagles medium (DMEM) (both from Gibco by Life Technologies, Grand Island, NY, USA). The media were supplemented with 10% fetal bovine serum (Cultilab, Sao Paulo, Brazil), 100 U/ml of penicillin and 100 µg/ml of streptomycin (both from Invitrogen, Grand Island, NY, USA). Cell cultures were grown at 37°C in a 5% CO2 atmosphere. 786-O cells were plated into 6-well plates, cultured for 24 h, and treated for 72 h with fresh 5, 2 or 1 µM of 5-Aza-dC (Sigma-Aldrich, St. Louis, MO, USA) dissolved in phosphate-buffered saline. Due to its chemical instability, fresh medium with 5-Aza-dC was added every 24 h. For the trichostatin A (TSA) experiments, the cells were seeded in 6-well plates, cultured for 24 h, and treated for 24 h with fresh 1,000, 500, 300, 200 or 100 nM of TSA (Sigma-Aldrich) dissolved in dimethyl sulfoxide (DMSO). To assess the combined effect of both drugs, we performed the co-treatment of cells with 5 µM of 5-Aza-dC and 1,000 nM of TSA using the following protocol: 5-Aza-dC was added for 72 h, after which it was removed and TSA was added for an additional 24 h. All experiments were performed in triplicate.

Cytotoxicity assay

786-O cells were seeded into 96-well plates at a density of 103 cells for 5-Aza-dC treatment, 104 cells for TSA and 5×103 cells for co-treatment. After the 5-Aza-dC and TSA treatments, 1 mg/ml of 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT) (Sigma-Aldrich) was added to each well and incubated for 30 min at 37°C. Then, the MTT was removed and 100 µl of 100% DMSO (Sigma-Aldrich) was added to each well and the absorbance was measured at 562 nm. Each experiment was performed in triplicate and in 2 independent assays.

RNA and DNA extraction

Genomic DNA and total RNA extraction were performed using TRIzol® reagent (Life Technologies) using both cell lines, the 45 fresh samples of ccRCC and 2 fresh-frozen histologically healthy renal tissue samples, following the manufacturer's instructions. The DNA and RNA quality was verified by electrophoresis through an agarose gel and stained with ethidium bromide. For quantitative real-time PCR, 5 µg of total RNA from each sample was used to synthesize cDNA with a High-Capacity cDNA Archive kit (Applied Biosystems, Foster City, CA, USA), according to the manufacturer's instructions. The cDNA quality was ascertained by PCR using the housekeeping gene β-actin. The PCR products were analyzed by electrophoresis on a 1% agarose gel and stained with ethidium bromide.

Selection of genes

We analyzed the differential expression of the following genes by qPCR: E-cadherin (CDH1), N-cadherin (CDH2), claudin 1 (CLDN1), cathepsin B (CTSB), platelet-activating factor acetylhydrolase 2 (PAFAH2), growth arrest-specific 1 (GAS1), laminin β1 (LAMβ1), occludin (OCLN), phosphoglycerate dehydrogenase (PHGDH), solute carrier family 2 member 1 (SLC2A1), transforming growth factor β1 (TGFβ1), tight junction protein 1 (TJP1) and vimentin (VIM). The genes were selected according to their role in cellular differentiation and proliferation.

Quantitative real-time polymerase chain reaction

The qPCR amplification was performed with Power SYBR-Green and an ABI 7300 Real-Time PCR System (Applied Biosystems, Warrington, UK). In brief, the reaction mixture (20 µl of total volume) contained 25 ng of cDNA, gene-specific forward and reverse primers and 10 µl of 2X Quantitative SYBR-Green PCR Master Mix. The samples were tested in triplicate. The relative expression of each specific gene was calculated using the following previously published (20) formula: R = (E target)ΔCt target (control - sample)/(E endogenous ΔCt endogenous (control - sample). A 2-fold-change was set as the cut-off. Gene expression was analyzed in 45 ccRCC samples, in the 786-O cells after treatment and the HaCaT cell line. Two healthy renal fresh-frozen tissue samples were used as the healthy reference (control group). A non-treated 786-O cell line was used as a reference sample and GAPDH was selected as the endogenous control gene.

Bisulfite sequencing

Genomic DNA was subjected to a sodium bisulfite treatment to modify unmethylated cytosine to uracil, as described by Calmon et al (21). Bisulfite-treated DNA was amplified by PCR, using primers designed to amplify CpG-rich regions located at the 5′ region of OCLN (−438 to +1232, relative to the transcription start site (TSS) and encompassing 164 CpG dinucleotides] and GAS1 [-485 to +1518, relative to the TSS and encompassing 269 CpG dinucleotides). The amplified products were cloned using a CloneJET PCR Cloning kit (Thermo Scientific, Carlsbad, CA, USA). Five positive clones from the 786-O cell line before and after treatment with 5-Aza-dC, one ccRCC sample and one healthy renal sample were sequenced using a BigDye Terminator v3.1 Cycle Sequencing kit and an ABI3130XL sequencer, according to the manufacturer's instructions (Applied Biosystems). The methylation percentage for each sample was calculated as the proportion of unconverted CpG dinucleotides among all of the CpG dinucleotides analyzed from the 5 positive clones.

Chromatin immunoprecipitation (ChIP)

ChIP was performed using a ChIP-IT® High Sensitivity kit according to the manufacturers instructions (Active Motif, Carlsbad, CA, USA). The 786-O and HaCaT cells were cross-linked with 1% formaldehyde for 10 min at room temperature and nuclei were isolated. Isolated nuclei were sonicated on ice to break the chromatin into 8,200–1,200 bp fragments. Soluble chromatin was used in immunoprecipitation with H3K4me3, H3K27me3, H3K4me2 and H3K9ac antibodies (Thermo Scientific, Rockford, IL, USA). IgG was used as a negative control (Abcam, Cambridge, MA, USA) and immune complexes were absorbed with protein G magnetic beads. After reversing the cross-linking and treating with proteinase K, the immunoprecipitated DNA was analyzed on a gene-specific basis using qPCR. The quality control ChIP was performed using a ChIP-IT® qPCR Analysis kit (Active Motif) with Human Negative Control Primer Set 1 and Human Positive Control Primer Set GAPDH-2, according to the manufacturers instructions.

MicroRNA analysis

Two microRNA databases, miRBase (www.mirbase.org) and TargetScanHuman (www.targetscan.org), were used to select microRNAs that target the OCLN and GAS1 mRNA. For microRNA expression analysis, total RNA was reverse transcribed into cDNA using miR-34a, miR-122 and U48-specific steam-loop primers. Expression of mature miR-34a and miR-122 was quantified as described in the qPCR methods. The microRNA expression was analyzed in 30 ccRCC samples and in the cell lines 786-O and HaCaT. Two healthy renal fresh-frozen tissue samples were used as a healthy reference. Each sample was run in triplicate, and the miRNA expression levels were normalized to U48.

Tissue array construction and immunohistochemistry

Tissue microarrays were constructed using a tissue microarrayer (Beecher Instruments, Silver Spring, MD, USA). Briefly, the slides were reviewed by a pathologist, and areas containing each category were annotated on the hematoxylin and eosin (H&E) slides. Fifty-six 1.0 mm in diameter cylindrical tissue cores from the ccRCC tumor samples and 24 from the tumor margins were taken from the corresponding regions of the paraffin blocks and transplanted into a recipient paraffin block.

For immunohistochemical staining, all of the paraffin-embedded sections were cut into 4-µm in thickness and the endogenous peroxidase activity was blocked with 3% hydrogen peroxide for 30 min. Antigen retrieval was performed using sodium citrate buffer (pH 6,0) for 20 min at 96°C in a steam pan. The sections were then incubated with SLC2A1 polyclonal primary antibody (1:200) (Abcam) diluted in 1% BSA at 4°C overnight. After incubation with a primary antibody, the sections were incubated with a biotinylated secondary antibody (Santa Cruz Biotechnology, Paso Robles, CA, USA) and then exposed to a streptavidin complex (HRP ready-to-use; DakoCytomation, Carpinteria, CA, USA). Positive staining was identified using DAB substrate color detection (Dako, Cambridge, UK). The sections were counterstained with hematoxylin.

The SLC2A1 densitometric analyses were conducted with an Axioskop II microscope using the Axiovision™ software (both from Zeiss, Oberkochen, Germany). For these analyses, 20 different points were analyzed from 3 different fields of each tumor fragment and the average immunoreactivity intensity [arbitrary units (AU)] was recorded.

Statistical analysis

Statistical analyses were performed using GraphPad Prism 5 Software (GraphPad Software Inc., San Diego, CA, USA). The relative expression levels detected by qPCR in ccRCC were transformed into natural logarithms. The Wilcoxon signed-rank test was applied to compare the gene expression levels in tumor and healthy tissues. The Mann-Whitney U test and Wilcoxon signed-ranks tests were used to compare the protein expression levels detected by immunohistochemistry.

Results

Evaluation of EMT gene expression in ccRCC

To evaluate gene expression in 45 ccRCC fresh samples and the 786-O cell line, we performed qPCR for the following target genes: CDH1, CDH2, CLDN1, CTSB, PAFAH2, GAS1, LAMβ1, OCLN, PHGDH, SLC2A1, TGFβ1, TJP1 and VIM.

CDH2, CTSB, GAS1, LAMβ1, OCLN, PHGDH, PAFAH2 and TJP1 were significantly downregulated in the ccRCC samples when compared to the reference samples. SLC2A1 was significantly overexpressed in tumor tissues when compared to healthy renal tissues. CDH1, TGFβ1, VIM and CLDN1 were not differentially expressed (Fig. 1).

CDH1, CTSB, GAS1 and OCLN were significantly downregulated in 786-O cells. SLC2A1 and VIM were significantly overexpressed in 786-O cells. CDH2, CLDN1, LAMβ1, PAFAH2, PHGDH, TGFβ1 and TJP1 were not differentially expressed (Fig. 1).

OCLN and GAS1 were chosen for epigenetic analysis due to the presence of CpG islands in their promoter regions and the observed downregulation in both the ccRCC samples and 786-O cells. SLC2A1 was selected for immunohistochemical analysis since it was overexpressed in both the ccRCC samples and 786-O cells.

DNA methylation and histone acetylation analysis

The methylation status of CpG islands at the 5′ region of the OCLN and GAS1 genes was investigated by bisulfite sequencing in ccRCC samples, healthy renal tissue and 786-O cells before and after 5-Aza-dC treatment. The bisulfite sequencing was performed in the −438 to +1232 region of OCLN and the −485 to +1518 region of GAS1. The methylation levels of OCLN and GAS1 were not significantly reduced in 786-O cells before and after 5 µM of 5-Aza-dC treatment (24.39–22.19% for OCLN and 22.5–22.23% for GAS1). The methylation level of the genes compared between the ccRCC samples and the healthy renal samples did not significantly differ, 4–1% for OCLN and 1.8–0% for GAS1 (Fig. 2).

An MTT assay was performed to detect the possible effects of 5-Aza-dC and TSA treatment on 786-O cell viability. All of the drug concentrations tested showed viability >75%, including the co-treatment (Fig. 3).

The cell growth was inhibited in a dose-dependent manner by 5-Aza-dC treatment in 786-O cells. The expression of OCLN mRNA was higher in the cells treated with 5 and 2 µM of 5-Aza-dC than in the untreated cells (Fig. 4A). There was no difference in the GAS1 mRNA expression after treatment with any 5-Aza-dC concentration (Fig. 4A).

Similar to the 5-Aza-dC treatment, TSA treatment showed a better response at the highest concentration (1,000 nM). The OCLN gene was upregulated in the 786-O cells after 1,000 nM TSA treatment when compared to the untreated cells (Fig. 4B). Other TSA concentrations were not able to significantly upregulate OCLN expression. The GAS1 expression levels were unchanged after TSA treatment in 786-O cells when compared to the untreated cells (Fig. 4B).

The 5-Aza-dC (5 µM) and TSA (1,000 nM) co-treatment efficiently upregulated OCLN expression in 786-O cells. However, GAS1 gene expression was not altered after co-treatment (Fig. 4C).

ChIP

Since there was alteration in the OCLN gene expression after 5-Aza-dC treatment without any difference in the DNA methylation level in its promoter region and it is known that 5-Aza-dC treatment may alter both DNA and histone methylation (22), we investigated whether the loss of OCLN expression in 786-O cells could be due to histone modifications. Regarding the GAS1 gene, we could not observe any change in its expression after the 5-Aza-dC and TSA treatments, and therefore, we also investigated if histone modifications could contribute to GAS1 supression in the absence of DNA hypermethylation or histone deacetylation.

In both cell lines, ChIP followed by qPCR analysis showed that all of the histone modifications were reduced or absent in the promoter regions of OCLN and GAS1. It was not possible to differentiate the relative occupancy of the marks studied between the healthy and tumor cell lines. The Human Negative Control Primer Set 1 and Human Positive Control Primer Set GAPDH-2 served as negative and positive controls, respectively (Fig. 5).

microRNA analysis

We performed stem-loop reverse transcription PCR (RT-qPCR) to quantify miR-122 and miR-34a expression in the 786-O and HaCaT cell lines. The relative expression of miR-122 and miR-34a in the 786-O cells was higher than in the HaCaT cells (Fig. 6). Both microRNAs were significantly overexpressed in tumor tissues when compared to healthy renal tissues. These results suggest that upregulation of miR-122 and miR-34a may be, at least in part, involved in the downregulation of OCLN and GAS1 in ccRCC.

Immunohistochemistry

The SLC2A1 protein expression was investigated in ccRCC tissue array by immunohistochemistry.The ccRCC tissues showed a higher expression of the SLC2A1 protein (p<0.0001) relative to the healthy tissues. Fig. 7 shows the SLC2A1 immunoreactivity and the densitometric data.

Discussion

Several prognostic models are used to predict the survival of ccRCC patients, including clinicopathological parameters and protein, genetic and epigenetic biomarkers (2325). In the present study, we investigated the epigenetic regulatory mechanisms of OCLN and GAS1, genes that may contribute to epithelial-mesenchymal transition (EMT), cellular differentiation and proliferation in renal cell carcinoma.

The OCLN gene is responsible for producing the occludin protein, which is present in all tight junctions. Occludin has been implicated in tight junction permeability and, in particular, to the regulation of size-selective diffusion (2628). The transmembrane proteins occludin and claudin are present in tight junctions, which are connected to the cytoskeleton via a network of proteins, such as zonula occludens-1 (29). This epithelial junctional complex, along with adherens junctions and desmosomes, is located at the most apical part of the complex (30). Tight junctions determine epithelial cell polarity and undergo dissolution during cancer progression that is accompanied by cytoskeletal changes (31).

Downregulation of OCLN has been observed in liver tumors, breast cancer, endometrial and lung carcinomas (3235). A study of human cutaneous squamous cell carcinoma observed the loss of OCLN and related this loss to cell-cell adhesion, apoptosis and proliferation (36). OCLN was found to be partially methylated in the promoter and endogenous region of a breast cancer cell line, and treatment with 5-Aza-dC and TSA could induce re-expression of this gene (37). In the present study, we observed hypermethylation in ~20% of the endogenous region, although, treatment with 5-Aza-dC did not change the global methylation profile of the OCLN gene in the ccRCC cell line. However, we demonstrated that the effect of 5-Aza-dC treatment and the synergistic effect of 5-Aza-dC and TSA treatment in 786-O cells were effective in reactivating OCLN gene expression. Thus, 786-O cells showed a methylation profile in the OCLN promoter, despite the suggestion of alternate modes of reactivation based upon OCLN expression. To further investigate whether increased OCLN expression in 786-O cells could be due to chromatin modifications, we performed ChIP for H3K27me3, H3K4me3, H3K4me2 and H3K9ac marks in the OCLN promoter in 786-O cells. ChIP analysis did not show any differences between H3K27me3 silencing marks or changes in H3K4me3, H3K4me2 and H3K9ac activation marks in the OCLN gene promoter region. H3K4me1, H3K4me2 and H3K4me3 marks inform us on the potential prognosis since their increase was found to be inversely correlated with lymph node involvement and distant metastasis in patients with renal cell carcinoma (38).

Another potential regulatory mechanism that may influence OCLN expression is microRNAs. miR-122 has been reported to regulate OCLN expression in intestinal tissue, which contributes to increased intestinal permeability (39). In the present study, miR-122 expression levels were higher in the ccRCC cell line and tumor samples, suggesting that the increase in microRNA expression may contribute to the downregulation of the OCLN gene. miR-122 expression was upregulated in 15 cases of clear cell papillary renal cell carcinoma and in 80 samples of ccRCC (4042). In breast cancer, high miR-122 levels have been associated with metastasis, likely by increasing nutrient availability in the pre-metastatic niche (43).

GAS1 is a membrane-associated protein; it is tethered to the membrane through a glycosyl-phosphatidylinositol anchor (44,45). GAS1 contributes to inhibitory signal transduction and interferes with cell proliferation due to the inhibition of DNA synthesis during the G0 to S phase (44,46). The Hedgehog pathway is an essential regulator of carcinogenesis, and the GAS1 protein participates in the Hedgehog signaling pathway through the binding to Hedgehog ligands and enhancing the signaling, thus, indirectly inducing EMT (14,47,48).

Reduced GAS1 expression was found to be positively associated with the survival time in gastric cancer cells (49). The downregulation of the GAS1 gene has been described in colorectal cancer, melanoma and prostate cancer metastasis (5052).

In the present study, the global methylation profile of GAS1 in ccRCC cells was unchanged before and after 5-Aza-dC treatment. Similarly, there were no gene expression modifications after treatment with TSA or co-treatment with TSA and 5-Aza-dC. Sacilotto et al (53) suggested that GAS1 gene transcription is modulated by chromatin remodeling and histone modifying complex recruitment to the GAS1 promoter in hepatic cells. However, our chromatin immunoprecipitation assays did not show a difference in the H3K27me3 silencing mark, nor were there differences in the H3K4me3, H3K4me2 and H3K9ac activation marks in the GAS1 gene promoter region in ccRCC cells. It is possible that other regulatory mechanisms are altering the expression of GAS1 in ccRCC, such as microRNAs.

GAS1 gene regulation in papillary thyroid carcinoma occurs through miR-34a, leading to cell proliferation and apoptotic suppression (54). miR-34a regulates mesangial proliferation and glomerular hypertrophy by directly inhibiting GAS1 in early diabetic nephropathies (55). In the present study, we measured miR-34a expression levels, which target the GAS1 gene, in ccRCC cell lines and tumor samples and suggest that the increase in this microRNA may contribute to the downregulation of GAS1. miR-34a was upregulated in 8 ccRCC samples and renal carcinoma cell lines, establishing an association between this miR and multiple targets (56,57). miR-34a overexpression supports cell proliferation in the majority of cancers, suggesting an unexpected link between cancer and neuronal and endocrine cell metabolism (58).

The SLC2A1 gene is overexpressed in different types of cancers, including lung, liver and breast (5961). SLC2A1 facilitates glucose transport and is present in the epithelium and endothelium, contributing to proliferation (59,62,63). In the present study, SLC2A1 was upregulated in ccRCC samples and 786-O cells as determined by both qPCR and immunohistochemistry. Cifuentes et al (64) demonstrated that SLC2A1 and glucose transport are regulated by insulin and therefore may be an important target for conventional pharmaceutical therapy and gene therapy in the treatment of osteosarcoma. The SLC2A1 protein belongs to a class of markers that can detect breast cancer by molecular imaging with more common immunoreactivity in ductal carcinoma in situ than in invasive breast cancer (65,66).

In summary, the silencing of OCLN and GAS1 may contribute to tumor progression due to a possible relationship between these genes and cell-cell adhesion, apoptosis and proliferation. We suggest that the increase in miR-122 and miR-34a expression may contribute to the downregulation of the OCLN and GAS1 genes, respectively. Overexpression of the SLC2A1 gene was higher in ccRCC than that in the non-neoplastic samples. The expression of these genes may be involved in ccRCC progression in addition to other cellular processes that contribute to tumor development. Once these genes have been characterized in ccRCC, molecular genetic tools may be used to explore the biological process involved, thus improving diagnosis, treatment and patient outcomes. More studies should be conducted to assess the overall importance of epigenetic events in the regulation of OCLN and GAS1 expression in ccRCC and in others tumors, as well as its relevancce in the analysis of possible transcription factors that may also contribute to the regulation of the expression of these genes.

Acknowledgements

The present study was supported by the São Paulo State Research Foundation, FAPESP (2012/08853-0), and the National Council for Research, CNPq.

References

1 

Shuch B, Amin A, Armstrong AJ, Eble JN, Ficarra V, Lopez-Beltran A, Martignoni G, Rini BI and Kutikov A: Understanding pathologic variants of renal cell carcinoma: Distilling therapeutic opportunities from biologic complexity. Eur Urol. 67:85–97. 2015. View Article : Google Scholar : PubMed/NCBI

2 

Cohen HT and McGovern FJ: Renal-cell carcinoma. N Engl J Med. 353:2477–2490. 2005. View Article : Google Scholar : PubMed/NCBI

3 

Kuroda N, Hosokawa T, Michal M, Hes O, Sima R, Ohe C and Lee GH: Clear cell renal cell carcinoma with focal renal angiomyoadenomatous tumor-like area. Ann Diagn Pathol. 15:202–206. 2011. View Article : Google Scholar : PubMed/NCBI

4 

Nagata M, Sakurai-Yageta M, Yamada D, Goto A, Ito A, Fukuhara H, Kume H, Morikawa T, Fukayama M, Homma Y, et al: Aberrations of a cell adhesion molecule CADM4 in renal clear cell carcinoma. Int J Cancer. 130:1329–1337. 2012. View Article : Google Scholar : PubMed/NCBI

5 

Baylin SB: DNA methylation and gene silencing in cancer. Nat Clin Pract Oncol. 2 Suppl 1:S4–S11. 2005. View Article : Google Scholar : PubMed/NCBI

6 

Tang J and Zhuang S: Epigenetics in acute kidney injury. Curr Opin Nephrol Hypertens. 24:351–358. 2015.PubMed/NCBI

7 

Baylin SB and Ohm JE: Epigenetic gene silencing in cancer - a mechanism for early oncogenic pathway addiction? Nat Rev Cancer. 6:107–116. 2006. View Article : Google Scholar : PubMed/NCBI

8 

Feinberg AP and Tycko B: The history of cancer epigenetics. Nat Rev Cancer. 4:143–153. 2004. View Article : Google Scholar : PubMed/NCBI

9 

Benard A, Goossens-Beumer IJ, van Hoesel AQ, De Graaf W, Horati H, Putter H, Zeestraten EC, van de Velde CJ and Kuppen PJ: Histone trimethylation at H3K4, H3K9 and H4K20 correlates with patient survival and tumor recurrence in early-stage colon cancer. BMC Cancer. 14:5312014. View Article : Google Scholar : PubMed/NCBI

10 

Khakpour G, Pooladi A, Izadi P, Noruzinia M and Bazzaz J Tavakkoly: DNA methylation as a promising landscape: A simple blood test for breast cancer prediction. Tumour Biol. 36:4905–4912. 2015. View Article : Google Scholar : PubMed/NCBI

11 

Li J and Mansmann UR: A microRNA molecular modeling extension for prediction of colorectal cancer treatment. BMC Cancer. 15:4722015. View Article : Google Scholar : PubMed/NCBI

12 

De Wever O, Pauwels P, De Craene B, Sabbah M, Emami S, Redeuilh G, Gespach C, Bracke M and Berx G: Molecular and pathological signatures of epithelial-mesenchymal transitions at the cancer invasion front. Histochem Cell Biol. 130:481–494. 2008. View Article : Google Scholar : PubMed/NCBI

13 

Pierucci-Alves F, Yi S and Schultz BD: Transforming growth factor beta 1 induces tight junction disruptions and loss of transepithelial resistance across porcine vas deferens epithelial cells. Biol Reprod. 86:362012. View Article : Google Scholar : PubMed/NCBI

14 

Katoh Y and Katoh M: Hedgehog signaling, epithelial-to-mesenchymal transition and miRNA (Review). Int J Mol Med. 22:271–275. 2008.PubMed/NCBI

15 

Kunkel M, Reichert TE, Benz P, Lehr HA, Jeong JH, Wieand S, Bartenstein P, Wagner W and Whiteside TL: Overexpression of Glut-1 and increased glucose metabolism in tumors are associated with a poor prognosis in patients with oral squamous cell carcinoma. Cancer. 97:1015–1024. 2003. View Article : Google Scholar : PubMed/NCBI

16 

Sun L, Zeng X, Yan C, Sun X, Gong X, Rao Y and Yan N: Crystal structure of a bacterial homologue of glucose transporters GLUT1-4. Nature. 490:361–366. 2012. View Article : Google Scholar : PubMed/NCBI

17 

Sasaki H, Shitara M, Yokota K, Hikosaka Y, Moriyama S, Yano M and Fujii Y: Overexpression of GLUT1 correlates with Kras mutations in lung carcinomas. Mol Med Rep. 5:599–602. 2012.PubMed/NCBI

18 

Li W, Wei Z, Liu Y, Li H, Ren R and Tang Y: Increased 18F-FDG uptake and expression of Glut1 in the EMT transformed breast cancer cells induced by TGF-beta. Neoplasma. 57:234–240. 2010. View Article : Google Scholar : PubMed/NCBI

19 

Fisseler-Eckhoff A: New TNM classification of malignant lung tumors 2009 from a pathology perspective. Pathologe. 30 Suppl 2:S193–S199. 2009.(In German). View Article : Google Scholar

20 

Pfaffl MW: A new mathematical model for relative quantification in real-time RT-PCR. Nucleic Acids Res. 29:e452001. View Article : Google Scholar : PubMed/NCBI

21 

Calmon MF, Colombo J, Carvalho F, Souza FP, Filho JF, Fukuyama EE, Camargo AA, Caballero OL, Tajara EH, Cordeiro JA, et al: Methylation profile of genes CDKN2Ap14 and p16), DAPK1CDH1, and ADAM23 in head and neck cancer. Cancer Genet Cytogenet. 173:31–37. 2007. View Article : Google Scholar : PubMed/NCBI

22 

Nguyen CT, Weisenberger DJ, Velicescu M, Gonzales FA, Lin JC, Liang G and Jones PA: Histone H3-lysine 9 methylation is associated with aberrant gene silencing in cancer cells and is rapidly reversed by 5-aza-2′-deoxycytidine. Cancer Res. 62:6456–6461. 2002.PubMed/NCBI

23 

Zigeuner R, Hutterer G, Chromecki T, Imamovic A, Kampel-Kettner K, Rehak P, Langner C and Pummer K: External validation of the Mayo Clinic stage, size, grade, and necrosis (SSIGN) score for clear-cell renal cell carcinoma in a single European centre applying routine pathology. Eur Urol. 57:102–109. 2010. View Article : Google Scholar : PubMed/NCBI

24 

Gulati S, Martinez P, Joshi T, Birkbak NJ, Santos CR, Rowan AJ, Pickering L, Gore M, Larkin J, Szallasi Z, et al: Systematic evaluation of the prognostic impact and intratumour heterogeneity of clear cell renal cell carcinoma biomarkers. Eur Urol. 66:936–948. 2014. View Article : Google Scholar : PubMed/NCBI

25 

Fisel P, Kruck S, Winter S, Bedke J, Hennenlotter J, Nies AT, Scharpf M, Fend F, Stenzl A, Schwab M, et al: DNA methylation of the SLC16A3 promoter regulates expression of the human lactate transporter MCT4 in renal cancer with consequences for clinical outcome. Clin Cancer Res. 19:5170–5181. 2013. View Article : Google Scholar : PubMed/NCBI

26 

Furuse M, Hirase T, Itoh M, Nagafuchi A, Yonemura S and Tsukita S and Tsukita S: Occludin: A novel integral membrane protein localizing at tight junctions. J Cell Biol. 123:1777–1788. 1993. View Article : Google Scholar : PubMed/NCBI

27 

Balda MS, Whitney JA, Flores C, González S, Cereijido M and Matter K: Functional dissociation of paracellular permeability and transepithelial electrical resistance and disruption of the apical-basolateral intramembrane diffusion barrier by expression of a mutant tight junction membrane protein. J Cell Biol. 134:1031–1049. 1996. View Article : Google Scholar : PubMed/NCBI

28 

Balda MS, Flores-Maldonado C, Cereijido M and Matter K: Multiple domains of occludin are involved in the regulation of paracellular permeability. J Cell Biochem. 78:85–96. 2000. View Article : Google Scholar : PubMed/NCBI

29 

Tsukita S, Furuse M and Itoh M: Multifunctional strands in tight junctions. Nat Rev Mol Cell Biol. 2:285–293. 2001. View Article : Google Scholar : PubMed/NCBI

30 

Farquhar MG and Palade GE: Junctional complexes in various epithelia. J Cell Biol. 17:375–412. 1963. View Article : Google Scholar : PubMed/NCBI

31 

Thiery JP and Sleeman JP: Complex networks orchestrate epithelial-mesenchymal transitions. Nat Rev Mol Cell Biol. 7:131–142. 2006. View Article : Google Scholar : PubMed/NCBI

32 

Orbán E, Szabó E, Lotz G, Kupcsulik P, Páska C, Schaff Z and Kiss A: Different expression of occludin and ZO-1 in primary and metastatic liver tumors. Pathol Oncol Res. 14:299–306. 2008. View Article : Google Scholar : PubMed/NCBI

33 

Martin TA, Mansel RE and Jiang WG: Loss of occludin leads to the progression of human breast cancer. Int J Mol Med. 26:723–734. 2010. View Article : Google Scholar : PubMed/NCBI

34 

Tobioka H, Isomura H, Kokai Y, Tokunaga Y, Yamaguchi J and Sawada N: Occludin expression decreases with the progression of human endometrial carcinoma. Hum Pathol. 35:159–164. 2004. View Article : Google Scholar : PubMed/NCBI

35 

Tobioka H, Tokunaga Y, Isomura H, Kokai Y, Yamaguchi J and Sawada N: Expression of occludin, a tight-junction-associated protein, in human lung carcinomas. Virchows Arch. 445:472–476. 2004. View Article : Google Scholar : PubMed/NCBI

36 

Rachow S, Zorn-Kruppa M, Ohnemus U, Kirschner N, Vidal-y-Sy S, von den Driesch P, Börnchen C, Eberle J, Mildner M, Vettorazzi E, et al: Occludin is involved in adhesion, apoptosis, differentiation and Ca2+-homeostasis of human keratinocytes: Implications for tumorigenesis. PLoS One. 8:e551162013. View Article : Google Scholar : PubMed/NCBI

37 

Osanai M, Murata M, Nishikiori N, Chiba H, Kojima T and Sawada N: Epigenetic silencing of occludin promotes tumorigenic and metastatic properties of cancer cells via modulations of unique sets of apoptosis-associated genes. Cancer Res. 66:9125–9133. 2006. View Article : Google Scholar : PubMed/NCBI

38 

Ellinger J, Kahl P, Mertens C, Rogenhofer S, Hauser S, Hartmann W, Bastian PJ, Büttner R, Müller SC and von Ruecker A: Prognostic relevance of global histone H3 lysine 4 (H3K4) methylation in renal cell carcinoma. Int J Cancer. 127:2360–2366. 2010. View Article : Google Scholar : PubMed/NCBI

39 

Ye D, Guo S, Al-Sadi R and Ma TY: MicroRNA regulation of intestinal epithelial tight junction permeability. Gastroenterology. 141:1323–1333. 2011. View Article : Google Scholar : PubMed/NCBI

40 

Munari E, Marchionni L, Chitre A, Hayashi M, Martignoni G, Brunelli M, Gobbo S, Argani P, Allaf M, Hoque MO, et al: Clear cell papillary renal cell carcinoma: micro-RNA expression profiling and comparison with clear cell renal cell carcinoma and papillary renal cell carcinoma. Hum Pathol. 45:1130–1138. 2014. View Article : Google Scholar : PubMed/NCBI

41 

White NM, Bao TT, Grigull J, Youssef YM, Girgis A, Diamandis M, Fatoohi E, Metias M, Honey RJ, Stewart R, et al: miRNA profiling for clear cell renal cell carcinoma: Biomarker discovery and identification of potential controls and consequences of miRNA dysregulation. J Urol. 186:1077–1083. 2011. View Article : Google Scholar : PubMed/NCBI

42 

Zhou L, Chen J, Li Z, Li X, Hu X, Huang Y, Zhao X, Liang C, Wang Y, Sun L, et al: Integrated profiling of microRNAs and mRNAs: microRNAs located on Xq27.3 associate with clear cell renal cell carcinoma. PLoS One. 5:e152242010. View Article : Google Scholar : PubMed/NCBI

43 

Fong MY, Zhou W, Liu L, Alontaga AY, Chandra M, Ashby J, Chow A, O'Connor ST, Li S, Chin AR, et al: Breast-cancer-secreted miR-122 reprograms glucose metabolism in premetastatic niche to promote metastasis. Nat Cell Biol. 17:183–194. 2015. View Article : Google Scholar : PubMed/NCBI

44 

Ruaro ME, Stebel M, Vatta P, Marzinotto S and Schneider C: Analysis of the domain requirement in Gas1 growth suppressing activity. FEBS Lett. 481:159–163. 2000. View Article : Google Scholar : PubMed/NCBI

45 

Stebel M, Vatta P, Ruaro ME, Del Sal G, Parton RG and Schneider C: The growth suppressing gas1 product is a GPI-linked protein. FEBS Lett. 481:152–158. 2000. View Article : Google Scholar : PubMed/NCBI

46 

Del Sal G, Ruaro ME, Philipson L and Schneider C: The growth arrest-specific gene, gas1, is involved in growth suppression. Cell. 70:595–607. 1992. View Article : Google Scholar : PubMed/NCBI

47 

Beachy PA, Karhadkar SS and Berman DM: Tissue repair and stem cell renewal in carcinogenesis. Nature. 432:324–331. 2004. View Article : Google Scholar : PubMed/NCBI

48 

Allen BL, Tenzen T and McMahon AP: The Hedgehog-binding proteins Gas1 and Cdo cooperate to positively regulate Shh signaling during mouse development. Genes Dev. 21:1244–1257. 2007. View Article : Google Scholar : PubMed/NCBI

49 

Wang H, Zhou X, Zhang Y, Zhu H, Zhao L, Fan L, Wang Y, Gang Y, Wu K, Liu Z, et al: Growth arrest-specific gene 1 is downregulated and inhibits tumor growth in gastric cancer. FEBS J. 279:3652–3664. 2012. View Article : Google Scholar : PubMed/NCBI

50 

Jiang Z, Xu Y and Cai S: Down-regulated GAS1 expression correlates with recurrence in stage II and III colorectal cancer. Hum Pathol. 42:361–368. 2011. View Article : Google Scholar : PubMed/NCBI

51 

Gobeil S, Zhu X, Doillon CJ and Green MR: A genome-wide shRNA screen identifies GAS1 as a novel melanoma metastasis suppressor gene. Genes Dev. 22:2932–2940. 2008. View Article : Google Scholar : PubMed/NCBI

52 

Scaltriti M, Brausi M, Amorosi A, Caporali A, DArca D, Astancolle S, Corti A and Bettuzzi S: Clusterin (SGP-2, ApoJ) expression is downregulated in low- and high-grade human prostate cancer. Int J Cancer. 108:23–30. 2004. View Article : Google Scholar : PubMed/NCBI

53 

Sacilotto N, Espert A, Castillo J, Franco L and López-Rodas G: Epigenetic transcriptional regulation of the growth arrest-specific gene 1Gas1) in hepatic cell proliferation at mononucleosomal resolution. PLoS One. 6:e233182011. View Article : Google Scholar : PubMed/NCBI

54 

Ma Y, Qin H and Cui Y: MiR-34a targets GAS1 to promote cell proliferation and inhibit apoptosis in papillary thyroid carcinoma via PI3K/Akt/Bad pathway. Biochem Biophys Res Commun. 441:958–963. 2013. View Article : Google Scholar : PubMed/NCBI

55 

Zhang L, He S, Guo S, Xie W, Xin R, Yu H, Yang F, Qiu J, Zhang D, Zhou S, et al: Down-regulation of miR-34a alleviates mesangial proliferation in vitro and glomerular hypertrophy in early diabetic nephropathy mice by targeting GAS1. J Diabetes Complications. 28:259–264. 2014. View Article : Google Scholar : PubMed/NCBI

56 

Liu H, Brannon AR, Reddy AR, Alexe G, Seiler MW, Arreola A, Oza JH, Yao M, Juan D, Liou LS, et al: Identifying mRNA targets of microRNA dysregulated in cancer: With application to clear cell renal cell carcinoma. BMC Syst Biol. 4:512010. View Article : Google Scholar : PubMed/NCBI

57 

Yu G, Li H, Wang J, Gumireddy K, Li A, Yao W, Tang K, Xiao W, Hu J, Xiao H, et al: miRNA-34a suppresses cell proliferation and metastasis by targeting CD44 in human renal carcinoma cells. J Urol. 192:1229–1237. 2014. View Article : Google Scholar : PubMed/NCBI

58 

Dutta KK, Zhong Y, Liu YT, Yamada T, Akatsuka S, Hu Q, Yoshihara M, Ohara H, Takehashi M, Shinohara T, et al: Association of microRNA-34a overexpression with proliferation is cell type-dependent. Cancer Sci. 98:1845–1852. 2007. View Article : Google Scholar : PubMed/NCBI

59 

Yan S, Wang Y, Chen M, Li G and Fan J: Deregulated SLC2A1 promotes tumor cell proliferation and metastasis in gastric cancer. Int J Mol Sci. 16:16144–16157. 2015. View Article : Google Scholar : PubMed/NCBI

60 

Amann T, Maegdefrau U, Hartmann A, Agaimy A, Marienhagen J, Weiss TS, Stoeltzing O, Warnecke C, Schölmerich J, Oefner PJ, et al: GLUT1 expression is increased in hepatocellular carcinoma and promotes tumorigenesis. Am J Pathol. 174:1544–1552. 2009. View Article : Google Scholar : PubMed/NCBI

61 

Krzeslak A, Wojcik-Krowiranda K, Forma E, Jozwiak P, Romanowicz H, Bienkiewicz A and Brys M: Expression of GLUT1 and GLUT3 glucose transporters in endometrial and breast cancers. Pathol Oncol Res. 18:721–728. 2012. View Article : Google Scholar : PubMed/NCBI

62 

Manel N, Kim FJ, Kinet S, Taylor N, Sitbon M and Battini JL: The ubiquitous glucose transporter GLUT-1 is a receptor for HTLV. Cell. 115:449–459. 2003. View Article : Google Scholar : PubMed/NCBI

63 

Augustin R: The protein family of glucose transport facilitators: It's not only about glucose after all. IUBMB Life. 62:315–333. 2010.PubMed/NCBI

64 

Cifuentes M, García MA, Arrabal PM, Martínez F, Yañez MJ, Jara N, Weil B, Domínguez D, Medina RA and Nualart F: Insulin regulates GLUT1-mediated glucose transport in MG-63 human osteosarcoma cells. J Cell Physiol. 226:1425–1432. 2011. View Article : Google Scholar : PubMed/NCBI

65 

Vermeulen JF, van Brussel AS, van der Groep P, Morsink FH, Bult P, van der Wall E and van Diest PJ: Immunophenotyping invasive breast cancer: Paving the road for molecular imaging. BMC Cancer. 12:2402012. View Article : Google Scholar : PubMed/NCBI

66 

Vermeulen JF, van der Wall E, Witkamp AJ and van Diest PJ: Analysis of expression of membrane-bound tumor markers in ductal carcinoma in situ of the breast: Paving the way for molecular imaging. Cell Oncol. 36:333–340. 2013. View Article : Google Scholar

Related Articles

Journal Cover

March-2017
Volume 37 Issue 3

Print ISSN: 1021-335X
Online ISSN:1791-2431

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Conceição AG, Da Silva CT, Badial RM, Valsechi MC, Stuqui B, Gonçalves JD, Jasiulionis MG, De Freitas Calmon M and Rahal P: Downregulation of OCLN and GAS1 in clear cell renal cell carcinoma. Oncol Rep 37: 1487-1496, 2017
APA
Conceição, A.G., Da Silva, C.T., Badial, R.M., Valsechi, M.C., Stuqui, B., Gonçalves, J.D. ... Rahal, P. (2017). Downregulation of OCLN and GAS1 in clear cell renal cell carcinoma. Oncology Reports, 37, 1487-1496. https://doi.org/10.3892/or.2017.5414
MLA
Conceição, A. G., Da Silva, C. T., Badial, R. M., Valsechi, M. C., Stuqui, B., Gonçalves, J. D., Jasiulionis, M. G., De Freitas Calmon, M., Rahal, P."Downregulation of OCLN and GAS1 in clear cell renal cell carcinoma". Oncology Reports 37.3 (2017): 1487-1496.
Chicago
Conceição, A. G., Da Silva, C. T., Badial, R. M., Valsechi, M. C., Stuqui, B., Gonçalves, J. D., Jasiulionis, M. G., De Freitas Calmon, M., Rahal, P."Downregulation of OCLN and GAS1 in clear cell renal cell carcinoma". Oncology Reports 37, no. 3 (2017): 1487-1496. https://doi.org/10.3892/or.2017.5414