Open Access

Gene expression profiling analysis of ovarian cancer

  • Authors:
    • Ji‑Gang Yin
    • Xian‑Ying Liu
    • Bin Wang
    • Dan‑Yang Wang
    • Man Wei
    • Hua Fang
    • Mei Xiang
  • View Affiliations

  • Published online on: June 1, 2016     https://doi.org/10.3892/ol.2016.4663
  • Pages: 405-412
  • Copyright: © Yin et al. This is an open access article distributed under the terms of Creative Commons Attribution License.

Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

As a gynecological oncology, ovarian cancer has high incidence and mortality. To study the mechanisms of ovarian cancer, the present study analyzed the GSE37582 microarray. GSE37582 was downloaded from Gene Expression Omnibus and included data from 74 ovarian cancer cases and 47 healthy controls. The differentially‑expressed genes (DEGs) were screened using linear models for microarray data package in R and were further screened for functional annotation. Next, Gene Ontology and pathway enrichment analysis of the DEGs was conducted. The interaction associations of the proteins encoded by the DEGs were searched using the Search Tool for the Retrieval of Interacting Genes, and the protein‑protein interaction (PPI) network was visualized by Cytoscape. Moreover, module analysis of the PPI network was performed using the BioNet analysis tool in R. A total of 284 DEGs were screened, consisting of 145 upregulated genes and 139 downregulated genes. In particular, downregulated FBJ murine osteosarcoma viral oncogene homolog (FOS) was an oncogene, while downregulated cyclin‑dependent kinase inhibitor 1A (CDKN1A) was a tumor suppressor gene and upregulated cluster of differentiation 44 (CD44) was classed as an ‘other’ gene. The enriched functions included collagen catabolic process, stress‑activated mitogen‑activated protein kinases cascade and insulin receptor signaling pathway. Meanwhile, FOS (degree, 15), CD44 (degree, 9), B‑cell CLL/lymphoma 2 (BCL2; degree, 7), CDKN1A (degree, 7) and matrix metallopeptidase 3 (MMP3; degree, 6) had higher connectivity degrees in the PPI network for the DEGs. These genes may be involved in ovarian cancer by interacting with other genes in the module of the PPI network (e.g., BCL2‑FOS, BCL2‑CDKN1A, FOS‑CDKN1A, FOS‑CD44, MMP3‑MMP7 and MMP7‑CD44). Overall, BCL2, FOS, CDKN1A, CD44, MMP3 and MMP7 may be correlated with ovarian cancer.

Introduction

Ovarian cancer is a common malignant tumor of the female reproductive organs, with an incidence ranked third of all malignant tumors, and the majority of these tumors are epitheliomas (1). With not only non-typical early symptoms but also complex ovarian embryonic development, tissue dissection and endocrine functions, only 19% of all ovarian cancers can be diagnosed at an early stage, and the mortality rate for this disease is highest in the field of gynecological oncology (2).

In recent years, a number of studies have been performed to investigate the molecular mechanisms of ovarian cancer. For example, as a member of the protein kinase B family, v-akt murine thymoma viral oncogene homolog 2 (AKT2) can be frequently activated and induce the apoptosis of human primary ovarian cancer by inhibiting the phosphoinositide 3-kinase (PI3K)/Akt pathway (3). With elevated expression levels in patients with epithelial ovarian cancer, ovarian cancer carbohydrate antigen 125 can be used for monitoring patients with ovarian cancer (4). As a chemokine receptor, chemokine (C-X-C motif) receptor 4 can be expressed on a subset of ovarian cancer cells and may have important roles in the development of site-specific metastasis (5,6). Chemokine (C-X-C motif) ligand 12 may affect ovarian cancer through several methods, such as stimulating cell migration and invasion, inducing establishment of a cytokine network and DNA synthesis in conditions that are suboptimal for tumor cell growth (7). PI3Ks are lipid kinases associated with neoplasia (8), and phosphoinositide-3-kinase, catalytic subunit α (which encodes the p110alpha catalytic subunit of PI3-kinase) is an oncogene that plays a critical role in ovarian cancer (9).

In 2012, Zhao et al (10) used a whole-genome microarray approach to analyze genes that differed between ovarian cancer cases and healthy controls, and obtained a total of 10,435 mRNA genes that could be used for downstream analysis. Using the data obtained by Zhao et al (10), the present study aimed to obtain the differentially-expressed genes (DEGs) and investigate their possible functions by Gene Ontology (GO) and pathway enrichment analyses. In addition, the interaction associations between these DEGs were searched using the protein-protein interaction (PPI) network and modules of the PPI network.

Materials and methods

Microarray data

The expression profile of GSE37582, deposited by Zhao et al (10), was downloaded from Gene Expression Omnibus (http://www.ncbi.nlm.nih.gov/geo/) and was based on the platform of the GPL6947 Illumina HumanHT-12 V3.0 expression beadchip. GSE37582 included data from 74 ovarian cancer cases and 47 healthy controls.

DEG screening and functional annotation

Once GSE37582 had been downloaded, microarray data was preprocessed by robust multi-array average (11) background correction. Next, quantile normalization was conducted. The average value of multiple probes mapped with one gene was obtained as the ultimate gene expression value. The linear models for microarray data package in R (12) was used to analyze the DEGs between ovarian cancer cases and healthy controls. A false discovery rate (FDR) of <0.05 and |log2fold-change|>1 were used as the cut-off criteria.

To predict genes with functions of transcription factors, the DEGs were screened in combination with the transcription factors database. In combination with the tumor suppressor gene (TSG) (13) and tumor-associated gene (14) databases, the TSGs and oncogenes were then further screened from the DEGs.

Functional and pathway enrichment analysis

Through use of controlled and structured vocabularies, GO (www.geneontology.org/) can act as a community-based bioinformatics resource and classify gene product functions (15). As an integrated database resource, the Kyoto Encyclopedia of Genes and Genomes (KEGG; www.genome.jp/kegg/) consists of systems information, genomic information and chemical information (16). GO and KEGG pathway enrichment analyses were performed for the DEGs. P<0.05 was used as the cut-off criterion.

PPI network and module construction

The Search Tool for the Retrieval of Interacting Genes online software (17) was used to search interaction associations of the proteins encoded by the DEGs, and the required confidence (combined score) of >0.4 was used as the cut-off criterion. Cytoscape (18) was used to visualize the PPI network. Next, connectivity degree analysis was conducted and key nodes (i.e., hub proteins) (19) were searched. The proteins in the network were defined as the nodes. Next, the BioNet (20) analysis tool in R was used to screen modules of the PPI network. The FDR was set to 3.

Results

DEG analysis and functional annotation

A total of 284 DEGs were screened, consisting of 145 upregulated genes and 139 downregulated genes. Only 1 transcription factor was significantly upregulated, while 8 transcription factors were significantly downregulated. For the upregulated genes, there was 1 oncogene, 4 TSGs and 3 ‘other’ genes [the influence of ‘other’ genes on tumors was unclear, and could therefore not be distinguished as oncogenes or tumor suppressor genes, including cluster of differentiation 44 (CD44)]. For the downregulated genes, there were 4 oncogenes [such as FBJ murine osteosarcoma viral oncogene homolog (FOS)], 11 TSGs [such as cyclin-dependent kinase inhibitor 1A (CDKN1A)] and 5 ‘other’ genes (Table I).

Table I.

Functional annotations of the differentially-expressed genes.

Table I.

Functional annotations of the differentially-expressed genes.

RegulationTF numbersTF symbolsTAG numbersTAG symbols
Up1KLF12  8TCL1B, CTNND1, MUC1, PTPRF, STARD13, CD44, FES, MIAT
Down8ASCL1, ETV4, HSF1, LMO3, PML, RUNX3, TCF7, USF220AKT2, FOS, GNA12, PVT1, CD82, CDKN1A, FBXO32, GADD45G, MAL, PML, PPP1R3C, RUNX3, SCGB3A1, SOCS1, VHL, ASCL1, BCAS1, OGG1, PMS2, RHOB

[i] TF, transcription factor; TAG, tumor-associated gene.

Functional and pathway enrichment analysis

The enriched GO functions for the DEGs are listed in Table II. For the upregulated genes, the enriched GO functions included the collagen catabolic process (P=1.86×10−2), extracellular matrix disassembly (P=4.50×10−3) and negative regulation of intrinsic apoptotic signaling pathway in response to DNA damage by p53 class mediator (P=3.07×10−3). For the downregulated genes, the enriched GO functions included the negative regulation of cell growth (P=1.39×10−4), stress-activated mitogen-activated protein kinases (MAPK) cascade (P=1.73×10−3) and insulin receptor signaling pathway (P=1.19×10−2).

Table II.

Enriched GO functions for the differentially-expressed genes.

Table II.

Enriched GO functions for the differentially-expressed genes.

Gene regulationGO termDescriptionGene numberGene symbolP-value
Upregulated genesGO:0042158Lipoprotein biosynthetic process5ABCA1, LCAT, MTTP, PIGF, PPM1B 1.30×10−4
GO:0007204Elevation of cytosolic calcium ion concentration5CCR1, CXCR3, GALR2, LIME1, TPCN2 1.34×10−2
GO:0022617Extracellular matrix disassembly4CTSK, KLKB1, MMP3, MMP7 4.50×10−3
GO:0051592Response to calcium ion4ANXA11, KCNMB1, MTTP, SPARC 5.28×10−3
GO:0030574Collagen catabolic process3CTSK, MMP3, MMP7 1.86×10−2
GO:0010875Positive regulation of cholesterol efflux2ABCA1, ABCG1 3.07×10−3
GO:1902166Negative regulation of intrinsic apoptotic signaling pathway in response to DNA damage by p53 class mediator2CD44, MUC1 3.07×10−3
GO:003443Cholesterol esterification2ABCG1, LCAT 3.67×10−3
GO:0034616Response to laminar fluid shear stress2ABCA1, TGFB3 3.67×10−3
GO:0034375High-density lipoprotein particle remodeling2ABCG1, LCAT 5.01×10−3
Downregulated genesGO:0006469Negative regulation of protein kinase activity8CDKN1A, DUSP1, DUSP10, DUSP4, DVL1, GADD45G, PDPK1, SOCS1 4.97×10−5
GO:0030308Negative regulation of cell growth7BCL2, CDKN1A, HSPA1B, PML, PPP1R9B, SCGB3A1, SERTAD3 1.39×10−4
GO:0051403Stress-activated MAPK cascade7AKT2, CCM2, CDC42EP5, DUSP10, DUSP4, FOS, MAP2K3 1.73×10−3
GO:0046777Protein autophosphorylation6CAMKK2, ERN1, PDGFA, PDPK1, PIM3, STK24 3.65×10−3
GO:0006986Response to unfolded protein5DNAJB1, ERN1, HSPA1B, HSPA2, HSPA6 4.61×10−3
GO:0008286Insulin receptor signaling pathway5AKT2, EIF4EBP1, PDPK1, SOCS1, TCIRG1 1.19×10−2
GO:0048146Positive regulation of fibroblast proliferation3CDKN1A, PDGFA, S100A6 4.29×10−3
GO:0045069Regulation of viral genome replication3BCL2, OASL, PPID 9.26×10−3
GO:0030833Regulation of actin filament polymerization3ARPC1B, CDC42EP5, PFN2 3.98×10−2
GO:0008295Spermidine biosynthetic process2AGMAT, AMD1 3.78×10−4

[i] GO, Gene Ontology; MAPK, mitogen-activated protein kinases.

The enriched KEGG pathways for the DEGs are listed in Table III. For the upregulated genes, the enriched KEGG pathways included the pentose phosphate pathway (P=2.43×10−2), hematopoietic cell lineage (P=4.46×10−2) and rheumatoid arthritis (P=4.84×10−2). Meanwhile, for the downregulated genes, the enriched KEGG pathways included the MAPK signaling pathway (P=3.77×10−6), pathways in cancer (P=7.47×10−3) and toxoplasmosis (P=7.24×10−5).

Table III.

Enriched KEGG pathways for the differentially-expressed genes.

Table III.

Enriched KEGG pathways for the differentially-expressed genes.

Gene regulationTermDescriptionGene numberGene symbolP-value
Upregulated genes  30Pentose phosphate pathway  2H6PD, RBKS 2.43×10−2
4640Hematopoietic cell lineage  3CD1C, CD44, CR1 4.46×10−2
5323Rheumatoid arthritis  3CTSK, MMP3, TGFB3 4.84×10−2
Downregulated genes4010MAPK signaling pathway13AKT2, DUSP1, DUSP10, DUSP4, FOS, GADD45G, GNA12, HSPA1A, HSPA1B, HSPA2, HSPA6, MAP2K3, PDGFA 3.77×10−6
5200Pathways in cancer  9AKT2, BCL2, CDKN1A, DVL1, FOS, PDGFA, PML, TCF7, VHL 7.47×10−3
5145Toxoplasmosis  8AKT2, BCL2, HSPA1A, HSPA1B, HSPA2, HSPA6, MAP2K3, SOCS1 7.24×10−5
4141Protein processing in endoplasmic reticulum  8BCL2, DNAJB1, ERN1, HSPA1A, HSPA1B, HSPA2, HSPA6, HSPBP1 3.41×10−4
5215Prostate cancer  6AKT2, BCL2, CDKN1A, PDGFA, PDPK1, TCF7 3.48×10−4
4910Insulin signaling pathway  6AKT2, EIF4EBP1, HK1, PDPK1, PPP1R3C, SOCS1 3.42×10−3
4144Endocytosis  6CHMP6, HSPA1A, HSPA1B, HSPA2, HSPA6, PML 2.01×10−2
3040Spliceosome  5HSPA1A, HSPA1B, HSPA2, HSPA6, PRPF6 1.13×10−2
5221Acute myeloid leukemia  4AKT2, EIF4EBP1, PML, TCF7 3.11×10−3
5210Colorectal cancer  4AKT2, BCL2, FOS, TCF7 4.22×10−3

[i] KEGG, Kyoto Encyclopedia of Genes and Genomes; MAPK, mitogen-activated protein kinases.

PPI network and module construction

The PPI network consisted of 110 nodes and 136 interactions. There were 45 upregulated genes and 65 downregulated genes in the PPI network of the DEGs (Fig. 1). The number of upregulated genes was less than that of the downregulated genes. The DEGs, which encoded proteins with connectivity degrees of ≥5, included FOS (degree, 15), CD44 (degree, 9), B-cell CLL/lymphoma 2 (BCL-2; degree, 7), CDKN1A (degree, 7), DnaJ heat shock protein family (Hsp40) member B1 (degree, 7), AKT2 (degree, 7) and matrix metalloproteinase 3 (MMP3; degree, 6).

A module was obtained from the PPI network of the DEGs (Fig. 2). The module had 62 nodes and 93 interactions. In this module, the number of upregulated genes was less than the number of downregulated genes. Moreover, BCL2 could interact with FOS and CDKN1A, FOS could interact with CDKN1A and CD44, and MMP7 could interact with MMP3 and CD44. The enriched GO functions for the DEGs in the module included extracellular matrix organization (P=2.86×10−6), response to drug (P=3.36×10−4) and regulation of cell motility (P=1.37×10−3) (Table IV). The enriched KEGG pathways for the DEGs in the module are listed in Table V, including pathways in cancer (P=9.01×10−5), prostate cancer (P=3.98×10−5), colorectal cancer (P=7.89×10−5) and acute myeloid leukemia (P=7.44×10−4).

Table IV.

Top 10 enriched GO functions for the differentially-expressed genes in the module.

Table IV.

Top 10 enriched GO functions for the differentially-expressed genes in the module.

GO termDescriptionGene numberGene symbolP-value
GO:0030198Extracellular matrix organization9CTSK, MMP3, CD44, CD151, SPARC, PDGFA, MMP7, VHL, TGFB3 2.86×10−6
GO:0006954Inflammatory response9LTB4R, CD44, MAP2K3, FOS, CR1, DUSP10, PTAFR, CXCR3, TNFAIP6 1.41×10−4
GO:0022612Gland morphogenesis7BCL2, CD44, CTNND1, ETV4, PML, PDGFA, TGFB3 2.37×10−7
GO:0006469Negative regulation of protein kinase activity7DVL1, CDKN1A, GADD45G, SOCS1, DUSP10, DUSP1, PDPK1 2.82×10−6
GO:0042493Response to drug7BCL2, CDKN1A, MCM7, FOS, PDGFA, MMP7, GNA12 3.36×10−4
GO:2000145Regulation of cell motility7BCL2, MMP3, FES, AKT2, PDGFA, CXCR3, PDPK1 1.37×10−3
GO:0043583Ear development6BCL2, DVL1, CCM2, SPARC, PDGFA, TGFB3 9.35×10−5
GO:0051403Stress-activated MAPK cascade6MAP2K3, AKT2, FOS, CCM2, DUSP10, TAOK1 1.69×10−4
GO:0043068Positive regulation of programmed cell death6CDKN1A, PML, RHOB, DUSP1, PPID, TGFB3 1.71×10−3
GO:0060333 Interferon-γ-mediated signaling pathway5CD44, OASL, SOCS1, PML, PTAFR 9.81×10−6

[i] GO, Gene Ontology; MAPK, mitogen-activated protein kinases.

Table V.

Enriched KEGG pathways for the differentially-expressed genes in the module.

Table V.

Enriched KEGG pathways for the differentially-expressed genes in the module.

TermDescriptionGene numberGene symbolP-value
4010MAPK signaling pathway14GADD45G, MAP2K3, HSPA2, AKT2, HSPA1A, FOS, HSPA1B, PDGFA, DUSP10, DUSP1, GNA12, TAOK1, HSPA6, TGFB3 2.91×10−9
5200Pathways in cancer10BCL2, DVL1, CDKN1A, AKT2, FOS, PML, PDGFA, VHL, TCF7, TGFB3 9.01×10−5
5145Toxoplasmosis  9BCL2, MAP2K3, HSPA2, AKT2, HSPA1A, SOCS1, HSPA1B, HSPA6, TGFB3 3.30×10−7
4141Protein processing in endoplasmic reticulum  7BCL2, DNAJB1, HSPA2, HSPA1A, HSPA1B, HSPBP1, HSPA6 1.68×10−4
5215Prostate cancer  6BCL2, CDKN1A, AKT2, PDGFA, TCF7, PDPK1 3.98×10−5
4144Endocytosis  6HSPA2, HSPA1A, HSPA1B, PML, HSPA6, TGFB3 3.17×10−3
5210Colorectal cancer  5BCL2, AKT2, FOS, TCF7, TGFB3 7.89×10−5
4910Insulin signaling pathway  5AKT2, SOCS1, EIF4EBP1, PPP1R3C, PDPK1 3.14×10−3
5221Acute myeloid leukemia  4AKT2, PML, EIF4EBP1, TCF7 7.44×10−4
4612Antigen processing and presentation  4HSPA2, HSPA1A, HSPA1B, HSPA6 2.18×10−3

[i] KEGG, Kyoto Encyclopedia of Genes and Genomes; MAPK, mitogen-activated protein kinases.

Discussion

In the present study, a total of 284 DEGs were screened, consisting of 145 upregulated genes and 139 downregulated genes. In particular, downregulated FOS was an oncogene, while downregulated CDKN1A was a tumor suppressor gene and upregulated CD44 was classed as an ‘other’ gene. The enriched functions included the collagen catabolic process, stress-activated MAPK cascade and insulin receptor signaling pathway. Meanwhile, FOS (degree, 15), CD44 (degree, 9), BCL2 (degree, 7), CDKN1A (degree, 7) and MMP3 (degree, 6) had higher connectivity degrees in the PPI network for the DEGs. Also, a module was obtained from the PPI network of the DEGs.

Bcl-2 belongs to an family of pro- and antiapoptotic proteins (21), able to enhance cell survival by inhibiting apoptosis (22). BCL2 is frequently expressed in primary ovarian carcinoma and may be an important determinant of drug-induced apoptosis in ovarian cancer (23,24). By deregulating programmed cell death and changing sensitivity to chemotherapy, alteration of the Bax/Bcl-2 balance may affect the clinical course of ovarian cancer (25). As a member of the Bcl-2 family, ABT-737 can significantly improve the activity of carboplatin and thus may be of great use in the treatment of patients with ovarian cancer (26). These data may indicate that the expression level of BCL2 is associated with ovarian cancer. The c-fos protein may act as a tumor suppressor and may be involved in apoptosis (27). Expression of Fos can affect the progression of ovarian carcinoma and may function as a prognostic factor (28). Thus, we speculated that FOS may play a role in ovarian cancer. In the module, the present study found that BCL2 exhibited interaction associations with FOS, indicating that BCL2 may also play a role in ovarian cancer by mediating FOS.

As one of the most important CDK inhibitors, p27Kip1 can function in the G1 cell cycle arrest induced by anti-human epidermal growth factor receptor type 2 antibody and inhibit tumor growth through post-translational regulation (29). As the product of the CDKN2A gene at chromosome 9p2, the CDK inhibitor p16 is associated with the progression and unfavorable prognosis of ovarian cancer (30,31). p21, also known as CDKN1A, regulates several processes, including cell growth (32), differentiation (33) and apoptosis (34). Overexpressed CDK inhibitor p21 can reduce the phosphorylation of retinoblastoma protein in trichostatin A-treated ovarian cancer cells, and therefore p21 may be involved in the cell cycle blockade and the differentiation of the cells (35). These data may indicate that the expression level of CDKN1A exhibits an association with ovarian cancer. In the module in the present study, it was found that CDKN1A could interact with FOS and BCL2, indicating that CDKN1A may also affect ovarian cancer by mediating the two genes.

CD44 is regarded as the main cell surface receptor for hyaluronan (36), an extracellular polysaccharide associated with the progression of ovarian cancer (37). Significant overexpression of CD44 in ovarian cancer cells can induce strong adhesion to the peritoneal mesothelium (38,39), and CD44 has a close association with the implantation of ovarian cancer metastasis (40). The CD44+/CD117+ status can be a molecular marker of ovarian cancer-initiating cells (41). As a stem cell marker, CD44 combined with higher aldehyde dehydrogenase 1 family member A1 activity may be used to define ovarian cancer stem cells (42). These data indicate that CD44 may play a role in ovarian cancer. In the module, the present study found that CD44 had interaction associations with FOS, indicating that CD44 may also play a role in ovarian cancer by mediating FOS.

MMPs are a family of zinc-dependent neutral endopeptidases that are overexpressed in the tumor microenvironment and are able to degrade essentially all extracellular matrix components (43,44). Members of the MMP family are involved in tumor cell metastasis and invasion, including the progression of ovarian cancer (45). As the smallest member of the MMP family, MMP7 can be overexpressed in malignant ovarian epithelium and may assist tumor cell invasion by inducing progelatinase activation (46). Meanwhile, MMP7 may act as a marker for detection and as a target for treatment in ovarian cancer (47). These data suggest that MMP7 may have a close association with ovarian cancer. In the module, the present study found there were several interaction associations (e.g., MMP3-MMP7 and MMP7-CD44), indicating that MMP7 may also play a role in ovarian cancer by mediating MMP3 and CD44.

In conclusion, a comprehensive bioinformatics analysis of DEGs that may affect ovarian cancer was conducted in the present study. A total of 284 DEGs were screened, and it was found that BCL2, FOS, CDKN1A, CD44, MMP3 and MMP7 may have a correlation with ovarian cancer. However, further research is required to unravel their roles in ovarian cancer.

References

1 

Scully RE: Pathology of ovarian cancer precursors. J Cell Biochem Suppl. 23:208–218. 1995. View Article : Google Scholar : PubMed/NCBI

2 

Iorio MV and Croce CM: MicroRNA profiling in ovarian cancer. Methods Mol Biol. 1049:187–197. 2013. View Article : Google Scholar : PubMed/NCBI

3 

Yuan ZQ, Sun M, Feldman RI, Wang G, Ma X, Jiang C, Coppola D, Nicosia SV and Cheng JQ: Frequent activation of AKT2 and induction of apoptosis by inhibition of phosphoinositide-3-OH kinase/Akt pathway in human ovarian cancer. Oncogene. 19:2324–2330. 2000. View Article : Google Scholar : PubMed/NCBI

4 

Yin BW and Lloyd KO: Molecular cloning of the CA125 ovarian cancer antigen: Identification as a new mucin, Muc16. J Biol Chem. 276:27371–27375. 2001. View Article : Google Scholar : PubMed/NCBI

5 

Scotton CJ, Wilson JL, Milliken D, Stamp G and Balkwill FR: Epithelial cancer cell migration: A role for chemokine receptors? Cancer Res. 61:4961–4965. 2001.PubMed/NCBI

6 

Strieter RM: Chemokines: Not just leukocyte chemoattractants in the promotion of cancer. Nat Immunol. 2:285–286. 2001. View Article : Google Scholar : PubMed/NCBI

7 

Scotton CJ, Wilson JL, Scott K, Stamp G, Wilbanks GD, Fricker S, Bridger G and Balkwill FR: Multiple actions of the chemokine CXCL12 on epithelial tumor cells in human ovarian cancer. Cancer Res. 62:5930–5938. 2002.PubMed/NCBI

8 

Campbell IG, Russell SE, Choong DY, Montgomery KG, Ciavarella ML, Hooi CS, Cristiano BE, Pearson RB and Phillips WA: Mutation of the PIK3CA gene in ovarian and breast cancer. Cancer Res. 64:7678–7681. 2004. View Article : Google Scholar : PubMed/NCBI

9 

Karakas B, Bachman K and Park B: Mutation of the PIK3CA oncogene in human cancers. Br J Cancer. 94:455–459. 2006. View Article : Google Scholar : PubMed/NCBI

10 

Zhao H, Shen J, Wang D, Guo Y, Gregory S, Medico L, Hu Q, Yan L, Odunsi K, Lele S and Liu S: Associations between gene expression variations and ovarian cancer risk alleles identified from genome wide association studies. PloS One. 7:e479622012. View Article : Google Scholar : PubMed/NCBI

11 

Irizarry RA, Hobbs B, Collin F, Beazer-Barclay YD, Antonellis KJ, Scherf U and Speed TP: Exploration, normalization and summaries of high density oligonucleotide array probe level data. Biostatistics. 4:249–264. 2003. View Article : Google Scholar : PubMed/NCBI

12 

Smyth GK: Limma: Linear models for microarray data. Bioinformatics and Computational Biology Solutions using R and Bioconductor. Springer. (New York). 397–420. 2005. View Article : Google Scholar

13 

Zhao M, Sun J and Zhao Z: TSGene: A web resource for tumor suppressor genes. Nucleic Acids Res. 41:D970–D976. 2013. View Article : Google Scholar : PubMed/NCBI

14 

Chen JS, Hung WS, Chan HH, Tsai SJ and Sun HS: In silico identification of oncogenic potential of fyn-related kinase in hepatocellular carcinoma. Bioinformatics. 29:420–427. 2013. View Article : Google Scholar : PubMed/NCBI

15 

Palla G, Derényi I, Farkas I and Vicsek T: Uncovering the overlapping community structure of complex networks in nature and society. Nature. 435:814–818. 2005. View Article : Google Scholar : PubMed/NCBI

16 

Lu M, Zhang Q, Deng M, Miao J, Guo Y, Gao W and Cui Q: An analysis of human microRNA and disease associations. PloS One. 3:e34202008. View Article : Google Scholar : PubMed/NCBI

17 

Von Mering C, Huynen M, Jaeggi D, Schmidt S, Bork P and Snel B: STRING: A database of predicted functional associations between proteins. Nucleic Acids Res. 31:258–261. 2003. View Article : Google Scholar : PubMed/NCBI

18 

Smoot ME, Ono K, Ruscheinski J, Wang PL and Ideker T: Cytoscape 2.8: New features for data integration and network visualization. Bioinformatics. 27:431–432. 2011. View Article : Google Scholar : PubMed/NCBI

19 

He X and Zhang J: Why do hubs tend to be essential in protein networks? PLoS Genet. 2:e882006. View Article : Google Scholar : PubMed/NCBI

20 

Beisser D, Klau GW, Dandekar T, Müller T and Dittrich MT: BioNet: An R-Package for the functional analysis of biological networks. Bioinformatics. 26:1129–1130. 2010. View Article : Google Scholar : PubMed/NCBI

21 

Baekelandt M, Kristensen GB, Nesland JM, Tropé CG and Holm R: Clinical significance of apoptosis-related factors p53, Mdm2 and Bcl-2 in advanced ovarian cancer. J Clin Oncol. 17:20611999.PubMed/NCBI

22 

White E: Life, death and the pursuit of apoptosis. Genes Dev. 10:1–15. 1996. View Article : Google Scholar : PubMed/NCBI

23 

Eliopoulos AG, Kerr DJ, Herod J, Hodgkins L, Krajewski S, Reed JC and Young LS: The control of apoptosis and drug resistance in ovarian cancer: Influence of p53 and Bcl-2. Oncogene. 11:1217–1228. 1995.PubMed/NCBI

24 

Miyashita T and Reed JC: Bcl-2 gene transfer increases relative resistance of S49. 1 and WEHI7. 2 lymphoid cells to cell death and DNA fragmentation induced by glucocorticoids and multiple chemotherapeutic drugs. Cancer Res. 52:5407–5411. 1992.PubMed/NCBI

25 

Marx D, Binder C, Meden H, Lenthe T, Ziemek T, Hiddemann T, Kuhn W and Schauer A: Differential expression of apoptosis associated genes bax and bcl-2 in ovarian cancer. Anti Cancer Res. 17:2233–2240. 1997.

26 

Witham J, Valenti MR, De-Haven-Brandon AK, Vidot S, Eccles SA, Kaye SB and Richardson A: The Bcl-2/Bcl-XL family inhibitor ABT-737 sensitizes ovarian cancer cells to carboplatin. Clin Cancer Res. 13:7191–7198. 2007. View Article : Google Scholar : PubMed/NCBI

27 

Teng CS: Protooncogenes as mediators of apoptosis. Int Rev Cytol. 197:137–202. 2000. View Article : Google Scholar : PubMed/NCBI

28 

Mahner S, Baasch C, Schwarz J, Hein S, Wölber L, Jänicke F and Milde-Langosch K: C-Fos expression is a molecular predictor of progression and survival in epithelial ovarian carcinoma. Br J Cancer. 99:1269–1275. 2008. View Article : Google Scholar : PubMed/NCBI

29 

Le XF, Claret FX, Lammayot A, Tian L, Deshpande D, LaPushin R, Tari AM and Bast RC Jr: The role of cyclin-dependent kinase inhibitor p27Kip1 in anti-HER2 antibody-induced G1 cell cycle arrest and tumor growth inhibition. J Biol Chem. 278:23441–23450. 2003. View Article : Google Scholar : PubMed/NCBI

30 

Kamb A, Gruis NA, Weaver-Feldhaus J, Liu Q, Harshman K, Tavtigian SV, Stockert E, Day RS III, Johnson BE and Skolnick MH: A cell cycle regulator potentially involved in genesis of many tumor types. Science. 264:436–440. 1994. View Article : Google Scholar : PubMed/NCBI

31 

Dong Y, Walsh MD, McGuckin MA, Gabrielli BG, Cummings MC, Wright RG, Hurst T, Khoo SK and Parsons PG: Increased expression of cyclin-dependent kinase inhibitor 2 (CDKN2A) gene product P16INK4A in ovarian cancer is associated with progression and unfavourable prognosis. Int J Cancer. 74:57–63. 1997. View Article : Google Scholar : PubMed/NCBI

32 

Gartel AL, Serfas MS and Tyner AL: p21–negative regulator of the cell cycle. Proc Soc Exp Biol Med. 213:138–149. 1996. View Article : Google Scholar : PubMed/NCBI

33 

Wu H, Wade M, Krall L, Grisham J, Xiong Y and Van Dyke T: Targeted in vivo expression of the cyclin-dependent kinase inhibitor p21 halts hepatocyte cell-cycle progression, postnatal liver development and regeneration. Genes Dev. 10:245–260. 1996. View Article : Google Scholar : PubMed/NCBI

34 

Ostrovsky O and Bengal E: The mitogen-activated protein kinase cascade promotes myoblast cell survival by stabilizing the cyclin-dependent kinase inhibitor, p21WAF1 protein. J Biol Chem. 278:21221–21231. 2003. View Article : Google Scholar : PubMed/NCBI

35 

Strait KA, Dabbas B, Hammond EH, Warnick CT, Ilstrup SJ and Ford CD: Cell cycle blockade and differentiation of ovarian cancer cells by the histone deacetylase inhibitor trichostatin A are associated with changes in p21, Rb and Id proteins. Mol Cancer Ther. 1:1181–1190. 2002.PubMed/NCBI

36 

Aruffo A, Stamenkovic I, Melnick M, Underhill CB and Seed B: CD44 is the principal cell surface receptor for hyaluronate. Cell. 61:1303–1313. 1990. View Article : Google Scholar : PubMed/NCBI

37 

Fraser JR, Laurent TC and Laurent UB: Hyaluronan: Its nature, distribution, functions and turnover. J Intern Med. 242:27–33. 1997. View Article : Google Scholar : PubMed/NCBI

38 

Cannistra SA, Kansas GS, Niloff J, DeFranzo B, Kim Y and Ottensmeier C: Binding of ovarian cancer cells to peritoneal mesothelium in vitro is partly mediated by CD44H. Cancer Res. 53:3830–3838. 1993.PubMed/NCBI

39 

Cannistra SA, Abu-Jawdeh G, Niloff J, Strobel T, Swanson L, Andersen J and Ottensmeier C: CD44 variant expression is a common feature of epithelial ovarian cancer: Lack of association with standard prognostic factors. J Clin Oncol. 13:1912–1921. 1995.PubMed/NCBI

40 

Bourguignon LY, Zhu H, Chu A, Iida N, Zhang L and Hung MC: Interaction between the adhesion receptor, CD44 and the oncogene product, p185 HER2, promotes human ovarian tumor cell activation. J Biol Chem. 272:27913–27918. 1997. View Article : Google Scholar : PubMed/NCBI

41 

Zhang S, Balch C, Chan MW, Lai HC, Matei D, Schilder JM, Yan PS, Huang TH and Nephew KP: Identification and characterization of ovarian cancer-initiating cells from primary human tumors. Cancer Res. 68:4311–4320. 2008. View Article : Google Scholar : PubMed/NCBI

42 

Wang YC, Yo YT, Lee HY, Liao YP, Chao TK, Su PH and Lai HC: ALDH1-bright epithelial ovarian cancer cells are associated with CD44 expression, drug resistance and poor clinical outcome. Am J Pathol. 180:1159–1169. 2012. View Article : Google Scholar : PubMed/NCBI

43 

Overall CM and López-Otín C: Strategies for MMP inhibition in cancer: Innovations for the post-trial era. Nat Rev Cancer. 2:657–672. 2002. View Article : Google Scholar : PubMed/NCBI

44 

Vihinen P and Kähäri VM: Matrix metalloproteinases in cancer: Prognostic markers and therapeutic targets. Int J Cancer. 99:157–166. 2002. View Article : Google Scholar : PubMed/NCBI

45 

Schmalfeldt B, Prechtel D, Härting K, Späthe K, Rutke S, Konik E, Fridman R, Berger U, Schmitt M, Kuhn W and Lengyel E: Increased expression of matrix metalloproteinases (MMP)-2, MMP-9 and the urokinase-type plasminogen activator is associated with progression from benign to advanced ovarian cancer. Clin Cancer Res. 7:2396–2404. 2001.PubMed/NCBI

46 

Wang FQ, So J, Reierstad S and Fishman DA: Matrilysin (MMP-7) promotes invasion of ovarian cancer cells by activation of progelatinase. Int J Cancer. 114:19–31. 2005. View Article : Google Scholar : PubMed/NCBI

47 

Tanimoto H, Underwood LJ, Shigemasa K, Parmley TH, Wang Y, Yan Y, Clarke J and O'Brien TJ: The matrix metalloprotease pump-1 (MMP-7, Matrilysin): A candidate marker/target for ovarian cancer detection and treatment. Tumor Biol. 20:88–98. 1999. View Article : Google Scholar

Related Articles

Journal Cover

July-2016
Volume 12 Issue 1

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
Yin JG, Liu XY, Wang B, Wang DY, Wei M, Fang H and Xiang M: Gene expression profiling analysis of ovarian cancer. Oncol Lett 12: 405-412, 2016
APA
Yin, J., Liu, X., Wang, B., Wang, D., Wei, M., Fang, H., & Xiang, M. (2016). Gene expression profiling analysis of ovarian cancer. Oncology Letters, 12, 405-412. https://doi.org/10.3892/ol.2016.4663
MLA
Yin, J., Liu, X., Wang, B., Wang, D., Wei, M., Fang, H., Xiang, M."Gene expression profiling analysis of ovarian cancer". Oncology Letters 12.1 (2016): 405-412.
Chicago
Yin, J., Liu, X., Wang, B., Wang, D., Wei, M., Fang, H., Xiang, M."Gene expression profiling analysis of ovarian cancer". Oncology Letters 12, no. 1 (2016): 405-412. https://doi.org/10.3892/ol.2016.4663