The prospect of discovering new biomarkers for ovarian cancer based on current knowledge of susceptibility loci and genetic variation (Review)

  • Authors:
    • Mikael Kronborg Christophersen
    • Claus Høgdall
    • Estrid Høgdall
  • View Affiliations

  • Published online on: September 26, 2019     https://doi.org/10.3892/ijmm.2019.4352
  • Pages: 1599-1608
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Ovarian cancer is the most lethal gynaecological malignancy. The cancer initially presents with non‑specific symptoms; thus, it is typically not discovered until the patient has reached the late, considerably more lethal, stages of the disease. Research focus is currently on finding novel biomarkers, especially for early detection and stratification of the disease. One promising approach has been to focus on mutations or variations in the genetic code that are associated with the risk of developing ovarian cancer. A certain heritable component is already known regarding genes such as BRCA1/2, TP53, MSH6, BRIP1 and RAD51C, yet these are estimated to only account for ~3.1% of the total risk. Recent advances in sequencing technologies have enabled the investigation of hundreds of thousands of genetic variants in genome‑wide association studies in tens of thousands of patients, which has led to the discovery of 108 (39 loci with P<5.0x10‑8) novel susceptibility loci for ovarian cancer, presented in this review. Using the published variants in a patient cohort screening, together with variants identified in our ongoing whole exome sequencing project, future aims are to ascertain whether certain of the novel variants could be used as biomarkers for early diagnosis and/or treatment decisions.

1. Introduction

Ovarian cancer (OC) is the 5th most common cancer and the most lethal gynaecological malignancy in European women (1). The International Federation of Gynaecology and Obstetrics characterises four major stages of OC, with stages I and II constituting tumours localised and mainly confined to the ovaries, which are associated with a good prognosis [5-year overall survival for stage I, 87.0-89.5% (2)], and the late stages III and IV, with confirmed spread to the peritoneum and/or distant metastasis, and poorer outcome [5-year overall survival for stage IV, 13.2-17.9% (2)]. Early-stage OC presents with non‑specific symptoms [including pelvic or abdominal pain, loss of appetite, fatigue and unexplained weight loss (3,4)] commonly associated with other diseases or ailments. Additionally, OC is a relatively rare disease, meaning that general practitioners will encounter a small number of OC cases throughout their career (5). Combined, this means that most patients with malignant growth in the pelvic region are diagnosed in the late stages of OC.

2. OC subtypes and biomarkers

OC is commonly divided into two major groups, epithelial and non-epithelial. Epithelial OC (EOC) comprises four main subtypes, based on the tissue of origin: Serous adenocarcinoma [high-grade (HGSC) and low-grade]; endometrioid adeno-carcinoma; ovarian clear cell adenocarcinoma (OCCC); and mucinous adenocarcinoma. Non-epithelial OC is subdivided into germ cell and sex chord/stromal OC. Overall, ~86% of OC cases are epithelial, and of these, 76% are serous histological subtype, with HGSC counting 83% (2). The characteristics of particularly the four main EOC types differ markedly in origin tissue, gene and microRNA (miRNA) expression, and morphology, and there is emerging consensus that they should be recognised as four distinct diseases (6-8).

Established biomarkers for OC include CA-125 (9) and human epididymis protein 4 (10), and various multivariate index assays measuring serum concentrations of these and other proteins, as well as taking ultrasound examination of the pelvic region, menopausal state, patient age and/or family history into account, have also been devised (11-15). While these schemes have increased the likelihood of differentiating malignant OC from a benign growth in the pelvis, they have not proved sufficient to decisively decrease mortality rates (16,17). Consequently, there is still a clear requirement for finding robust biomarkers, especially those capable of detecting OC at the early stages that can be used prognostically and to guide targeted treatment.

OC has a significant heritable component. Mutations in particularly BRCA1 and BRCA2, but also in other genes, including TP53, BRIP1, MSH6 and RAD51C, have been described as risk factors (18,19), yet the known and familial genetic factors are estimated to only account for 3.1% of the risk of developing EOC (20). Therefore, it is proposed that there are additional OC susceptibility loci yet to be discovered, and this has been a major focus area in the past decade, as genomic research and sequencing techniques have improved significantly.

This review will present the advances in applying next-generation sequencing (NGS) in large cohort studies in the search for genetic variants that act as susceptibility loci and/or driver mutations for OC.

3. Literature search and inclusion/exclusion criteria

This review was carried out according to the Preferred Reporting Items for Systematic Reviews and Meta-Analyses guidelines (21). Studies were selected based on the search criteria ‘ovarian cancer' and ‘susceptibility loci' in the biomedical databases Medline, EMBASE and Scopus. Studies reporting genome-wide association studies (GWAS) in large cohorts were preferred. The aim was to cover as much of the published literature as possible; however, studies reporting variants associated with low malignant potential (borderline) OC subtypes were omitted, and only studies reported in English were included. In total, 108 susceptibility loci from 28 studies published from 2008 to 2018 were included (Fig. 1).

4. NGS as a tool in searching for ovarian cancer biomarkers

Only a decade ago, sequencing the genome of a single individual took months, if not years. Subsequent advances in micro-array and sequencing technologies prompted by large-scale sequencing efforts such as the Human Genome Project (22) and the 1,000 Genomes Project (23) have revolutionised the field of genomic research, and today this can be accomplished over ~1 week using high-throughput sequencing (24). Targeted sequencing of only parts of the genome, such as transcriptome or whole exome sequencing, or sequencing of a subset of genes known to be involved in tumorigenesis, have enabled scientists and clinicians to develop and tailor research and treatment to the individual patient, a fundamental premise for precision medicine initiatives, and the overall goal for the treatment of patients with OC (25).

As high-throughput sequencing evolved into NGS (also known by the more appropriate term, massively-parallel sequencing), several hundred thousand genetic variants in thousands of patients can now be investigated in only a fraction of the time (26). Naturally, this has spawned large cohort studies, often with participation of clinics across the world, as well as the invention of specific arrays or chips focused on variants or genes suspected to be the cause of specific diseases. The Cancer Genome Atlas (TCGA) Research Network investigated 33 different cancer forms using high-throughput single nucleotide polymorphism (SNP), exome and genome sequencing, as well as gene expression, copy number variation, DNA methylation and miRNA profiling; these findings were recently summarised (27). OC was one of the three cancer types selected for the pilot project, and a cohort of 489 patients with HGSC were selected for analysis. Among other findings, the researchers found TP53 to be mutated in almost all cases, and were able to classify tumours into several subtypes depending on transcription, miRNA and methylation profiles (28).

NGS studies of OC have been reported in the last five years, mainly stemming from two large global initiatives with significant overlaps: The US‑based OncoArray Network and its eponymous genotyping array chip (29); and the mega-consortium Collaborative Oncological Gene-environment Study (COGS) with the iCOGS array, and updated OncoChip (30). Established in 2005, the Ovarian Cancer Association Consortium (OCAC) is a major collaboration, with contributors from the United States, United Kingdom, Australia, The Netherlands, Denmark, Poland, Germany and numerous other countries, and consists of 25,509 population-based EOC cases and 40,941 controls (31). The consortium was included in COGS together with Breast Cancer Association Consortium (BCAC), Prostate Cancer Association Group to Investigate Cancer-Associated Alterations in the Genome and The Consortium of Investigators of Modifiers of BRCA1/2 (CIMBA), with the aim of studying the genetics and risk factors of these three hormone-related cancers [summarised in (30)]. For this collaboration, a custom genotyping array chip called iCOGS capable of genotyping >211,000 SNPs was developed and used on >250,000 subjects (30). Like COGS, the OncoArray Network's research and the OncoArray chip capable of genotyping 570,000 SNPs have resulted in numerous articles on glioblastoma, breast, ovarian, prostate and lung cancers, using, among others, the OCAC and BCAC cohorts (31-35).

5. GWAS identifies numerous susceptibility loci for ovarian cancer

In total, 108 susceptibility loci for OC were identified following a systematic literature search (summarised in Table I and Table SI for variants with P<5.0×10−8 and P>5.0×10−8, respectively). These loci were mainly found via GWAS, in which genetic variations in a cohort of patients are compared to a cohort of healthy controls to isolate variants that may contribute to developing the disease. Variants are given an odds ratio (OR) score, depending on whether the variant is found predominantly in the patient cohorts (OR >1) or in the healthy controls (OR <1).

Table I

Susceptibility loci identified through 28 published genome‑wide association studies in ovarian cancer cohorts. Subtype.

Table I

Susceptibility loci identified through 28 published genome‑wide association studies in ovarian cancer cohorts. Subtype.

SNPLocusPositionaEAbEAFcOR CI95P−value/BFdGeneeeQTLfRefs.
EnOCrs55502517905q12.366825262delACA0.561.18(1.11-1.26) 4.5×10−8MAST4(31)
CACAC
EnOCrs1295105317p13.17674089C0.181.31(1.05-1.62)109.6gTP53(93)
EOCrs19966126617q11.230854203insT0.280.91(0.88-0.94) 5.4×10−9ATAD5, AC130324.2(20)
EOCrs58777813417q23.261776459insTT0.028.13(4.74-13.95) 2.8×10−14BRIP1BRIP1(86)
EOC-BRCA1rs216510902q13111061081C0.251.09(1.05-1.12) 4.2×10−8ACOXL(31)
EOC-BRCA1rs469113904q32.3164987569G0.471.20(1.17-1.38) 3.4×10−8TRIM61, AC106872.4(90)
EOC-BRCA1rs988665108q24.21127805637G0.461.08(1.05-1.11) 3.5×10−9PVT1(31)
EOC-BRCA1rs795324912q24.31120965921G0.421.08(1.06-1.10) 1.1×10−9HNF1AHNF1A, OASL(31)
EOC-BRCA1rs1763130317q2145439036G0.191.27(1.17-1.38) 1.4×10−8PLEKHM1(90)
EOC-BRCA1rs18321117q2146710944A0.231.25(1.16-1.35) 3.1×10−8PLEKHM1PLEKHM1(90)
HGSCrs207259002q31176177905T0.321.20(1.14-1.25) 3.8×10−14HOXD9HOXD1,(69)
HOXD3,
HOXD9
HGSCrs675577702q31.1176178498T0.681.15NR 9.0×10−14HAGLR, HAGLROS(56)
HGSCrs765144603q25156689208A0.051.44(1.35-1.53) 1.5×10−28TIPARPTIPARP(37)
HGSCrs7856112303q25.31156747145C0.061.45(1.32-1.59) 3.0×10−15LINC00886(77)
HGSCrs6227395903q25.31156852891A0.061.45(1.32-1.59) 3.5×10−14LEKR1(77)
HGSCrs1006969005p15.331279675T0.261.15(1.11-1.20) 1.3×10−11TERTTERT(48)
HGSCrs772615905p15.331282204A0.341.12(1.09-1.16) 3.4×10−9TERTTERT(47)
HGSCrs1178265208q2181741409G0.071.24(1.16-1.33) 7.0×10−10CHMP4CCHMP4C(37)
HGSCrs1008821808q24128531703A0.130.76(0.70-0.81) 8.0×10−15MYCMYC, THEM75(69)
HGSCrs781493708q24128529479C0.111.18(1.13-1.24)NRLINC00824(55)
HGSCrs381411309p2216915023C0.270.77(0.73-0.81) 4.1×10−21BNC2BNC2(68)
(73)
HGSCrs708445410p12.3121532345A0.241.10(1.06-1.14)NRMLLT10(55)
HGSCrs228749817p13.17689242A0.141.30(1,07-1,57)165.7gWRAP53, TP53(93)
HGSCrs1295105317p13.17674089C0.181.19(1,01-1,38)47.8gTP53(93)
HGSCrs75721017q1237736525A0.371.12(1.08-1.17) 8.1×10−10HNF1BHNF1B(37)
(36)
HGSCrs740577617q1237733029A0.371.25(1.17-1.34) 2.3×10−11HNF1BHFN1B(36)
(67)
HGSCrs207760617q21.3145452177A0.081.15(1.12-1.19)NRPLEKHM1(55)
HGSCrs296000017q21.3145456987C0.181.16(1.12-1.20) 3.3×10−10PLEKHM1, AC091132.2(70)
HGSCrs1294266617q21.3145422473G0.221.15(1.11-1.20) 1.0×10−9PLEKHM1, ARHGAP27PLEKHM1(70)
HGSCrs105258717q21.3146025238C0.221.12(1.08-1.17) 4.6×10−8MAPT(70)
HGSCrs721834517q21.3248425805T0.311.12(1.08-1.16)NRSKAP1(55)
HGSCrs236395619p13.1117283315T0.491.16(1.11-1.21) 3.8×10−11ANKLE1ANKLE1(50)
HGSCrs817019p13.1117278895T0.191.18(1.12-1.25) 2.7×10−9BABAM1, USHBP1ANKLE1, BABAM1,(50)
ABHD8
HGSCrs480807519p13.1117279482C0.31.19(1.14-1.23) 9.2×10−20USHBP1, BABAM1ABHD8, BABAM1(54)
MOCrs75259002q13113215368G0.211.34(1.21-1.49) 3.3×10−8PAX8PAX8(38)
MOCrs71183002q31.1176172583A0.321.30(1.20-1.40) 7.5×10−12HOXD3,HOXD3, HOXD9(38)
AC009336.2
MOCrs11207182003q22.3139130269insCCA0.331.29(1.20-1.37) 1.5×10−13BPESC1, MRPS22(31)
GATTCA
GAAT
MOCrs32020309q31.1102180944A0.881.29(1.18-1.41) 1.7×10−8GRIN3A(31)
MOCrs68818719q13.239242112A0.320.67(0.60-0.75) 6.8×10−13IFNL3IFNL3, IFNL4(38)

{ label (or @symbol) needed for fn[@id='tfn1-ijmm-44-05-1599'] } Susceptibility loci with an OR P<5.0×10−8 are included.

a Human genome reference assembly GRCh38/hg38.

b EA is not necessarily on the leading strand. Preferably quoted by original publication, otherwise the minor allele, or with frequency corresponding to EAF.

c EAF, preferably in case cohort, otherwise in the European ancestry population in the 1,000 Genomes Project.

d BF >10 is considered strong confidence that the result will withstand further investigation.

e Gene(s) in closest proximity to the most significant ovarian cancer risk variant at each locus.

f Gene with published functional association with phenotype and thus the likely gene responsible. Preferably quoted by reference; otherwise Ensembl's Biomart database (‘Associated gene with phenotype').

g BF. BF, Bayes factor; CI, confidence interval; EA, effect allele; EAF, EA frequency; EnOC, endometrioid ovarian cancer; EOC, epithelial ovarian cancer (no subtype specified); EOC‑BRCA1, EOC cases including BRCA1 carriers (The Consortium of Investigators of Modifiers of BRCA1/2); eQTL, expression quantitative trait loci; HGSC, high grade serous carcinoma; MOC, mucinous ovarian cancer; NR, not reported; OR, odds ratio; SNP, single nucleotide polymorphism.

In total, >50% of the OC susceptibility loci were found to be involved in HGSC (59/108), which was perhaps expected, as this is by far the most prevalent subtype of OC and thus the one most frequently encountered. Certain variants have been reported in >1 subtype, most notably rs757210, which seems to be linked with poor prognosis in HGSC (odds ratio 1.12), but predicts superior outcomes in OCCC (OR 0.80), demonstrating the importance of stratifying GWAS findings by OC subtype (36-38). rs757210 sits in the promoter region of HFN1B, which is known to be overexpressed in OCCC (39) and downregulated in serous OC (36), as well as being a susceptibility gene for diabetes type II (40), prostate cancer (41,42), uterine corpus cancer (43) and endometrial cancer (44,45). Shen et al (36) hypothesised that the difference in expression levels could be due to promoter methylation of HNF1B in proximity to this variant, which was later confirmed (46).

Mutation hotspots are a common feature in cancer genomics, and some of the identified susceptibility loci were situated in or near genes that are frequently altered in cancer cells. As such, Pooley et al and Bojesen et al (47,48) investigated the telomerase gene TERT, which maintains chromosome telomeres. Somatic mutations, especially in the promoter region of TERT, have been found in cancers of the brain, thyroid gland, bladder and skin (49). Bojesen et al (48) reported a locus associated with HGSC (rs10069690) in intron 4 with the minor allele conferring increased risk of disease and creates an alternative splice site that results in a truncated protein and impaired telomerase function. This reinforces the hypothesis that shorter telomeres increase cancer risk.

Bolton et al (50) examined a known breast cancer locus on chromosome 19p13, and found susceptibility loci that were significantly associated with risk of serous OC. Presumably, the SNP rs8170 located in the BABAM1 (previously MERIT40) gene may explain this risk, as BABAM1 has been shown to interact with, and stabilise, BRCA1 interactions with a complex including RAP80, BRCC45 and CCDC98 (51-53). Although counterintuitive, as breast cancer and OC frequently present with inactivating mutations in BRCA1, the authors speculated that an overexpressed BABAM1 leads to stabilisation of the breast cancer susceptibility protein (BRCA) complex, and thus an increased tolerance to DNA damage. Lawrenson et al (54) fine‑mapped the region near rs8170 in the BCAC, OCAC and CIMBA cohorts, and found rs4808075 to be the strongest candidate causal variant. Further investigation led to the discovery of rs4808616, in strong linkage disequilibrium (LD) with rs4808075, that has significant expression quantitative trait loci (eQTL) association with ABHD8, a gene neighbour of BABAM1, in serous OC. Functional analyses revealed an association between the ABHD8 promoter and rs4808616 via chromatin conformation capture, whereas overexpression of ABHD8, but not BABAM1, affected ovarian epithelial cells in vitro. Therefore, BABAM1 and rs8170 identified by Bolton et al may have been a proxy for the real driver variant, potentially either rs4808075 or rs4808616 in ABHD8. ABHD8 is a notable gene that has recently been speculated to be involved in the migration and invasion of OC tumour cells through a homeobox-containing transcription factor network (55,56).

Ghoussaini et al (57) used cohorts that were later included in COGS (Studies of Epidemiology and Risk Factors in Cancer Heredity, MALignant OVArian cancer study, Family Registry for Ovarian Cancer Study, United Kingdom Ovarian Cancer Population Study and United Kingdom Genetic Prostate Cancer Study) to examine the 8q24 gene desert, where the two closest genes are well-known cancer susceptibility genes (c‑MYC and FAM84B). c‑MYC was functionally validated to be implicated in OC (Table I) and was also identified in the TCGA dataset (28).

6. GWAS statistical significance threshold

GWAS analyses have become the golden standard for finding disease susceptibility loci, but there are certain limitations as well. With hundreds of thousands of variants examined on a single chip, the risk of false positives increases dramatically, and stringent data processing must be employed. In general, GWAS studies favour common variants in the population, meaning that fine‑mapping and additional filtering are required to discover variants with a minor allele frequency (MAF) of <5% (58). Genetic variation occurs semi-randomly and is widespread throughout the genome, and large cohort sizes in the thousands are required to obtain statistically significant and reliable results. There is an ongoing debate regarding which P-value threshold should be the standard for GWAS, or whether Bayesian approaches should be employed instead (59). The generally accepted P‑value is P≤5.0×10−8 for common variants, as first introduced by The International HapMap Project (60) and subsequently by Pe'er et al (61), and which has been recently evaluated and confirmed (62). This latest study concluded that this threshold is too relaxed for rare variants (MAF ≤0.5%), and cut‑offs for these should be: 3×10−8 for MAF ≥1%, 2×10−8 for MAF ≥0.5% and 1×10−8 for MAF ≥0.1% (conditions: Whole-genome sequencing studies in European populations with all variants having an LD r2>0.8). For the present review, and contrary to two recent reviews of GWAS susceptibility loci (63,64), it was determined that all variants reported by the original articles would be included, with a more relaxed cut-off of P≤0.05. Variants meeting the threshold criteria discussed above (P≤5.0×10−8) are presented in Table I; the remaining variants are included in Table SI. For simplicity, for articles fine‑mapping a susceptibility loci region and finding additional SNPs with lower P-values, but in strong LD with the index SNP (48,54), only the novel variant with the strongest association was included.

7. Susceptibility loci region fine‑mapping

Genetic variants are not randomly distributed in the genome, but often aggregate in the same populations (23). Genomic research has taken advantage of this, by examining only those variants or polymorphisms already reported in large population studies. Nevertheless, the number of genetic variants in the human genome amounts to tens of millions, which is not feasible to investigate in a research setting on a large number of patients. Instead, an array of representative or index SNPs are frequently used to cover all variants in a genomic region, utilising the fact that neighbouring SNPs are often in tight LD and thus inherited together (65,66). Data from The 1,000 Genomes Project estimates that any given trait-associated variant in the National Human Genome Research Institute GWAS database will have 56 neighbouring variants in LD with r2≥0.5 (28). It follows that fine‑mapping of the region is required to determine if the index SNP is indeed the causal variant, or merely a proxy for other SNPs in the region. Three examples of this have been described in the subsection ‘Genome-wide association studies identify numerous susceptibility loci for ovarian cancer': Bojesen et al fine-mapped the TERT locus and SNPs in LD with rs10069690 and rs7705526 (48); Lawrenson et al examined the ABHD8/ANKLE1 locus and rs4808075 (54); and Shen et al investigated the HNF1B region and rs7405776 and rs11651755 (36). Following the initial findings of the COGS initiative, Earp et al (67) analysed 11 known susceptibility regions and found novel associated variants with more robust P-values and ORs than those previously reported (Table I) (20,37,48,50,68-71).

Several studies over the last few years have fine‑mapped the 9p22.2 region by rs3814113 first reported by Song et al in 2009 (68). eQTL analyses concluded the nearby zinc finger protein basonuclin-2 (BNC2), which has been implicated in oocyte differentiation (72), to be the most likely causal candidate gene (69,73). Additional SNPs were found to be associated with abnormal ovarian ultrasound results (74) and to modify OC risk in BRCA1/2 mutation carriers (75). BNC2 was reported to contribute to a HOX-centric network of transcription factors associated with serous OC risk (55), and Carter et al (76) found a significant association between germline rs3814113 and tumour formation in OC. Finally, a recent study by Buckley et al (73) reported additional SNPs in LD with rs3814113, as well as SNPs located in the regulatory regions of BNC2, including some in this gene's scaffold/matrix attachment region, suggesting that they influence chromosomal three-dimensional organisational optimization for transcription in an allele‑specific manner.

Finally, the cluster of OC-related variants in the TIPARP/LEKR1 region on chromosome 3q25 is of note in an OC disease setting and has been studied thoroughly (37,67,69,77). TIPARP (also known as PARP7) codes for a poly ADP ribose polymerase (PARP), a group of proteins that have been the target of PARP inhibitor cancer treatments showing great promise in the targeted treatment of patients with breast, prostate and OC carrying BRCA1/2 mutations (78).

8. Several variants cause potentially functionally relevant amino acid changes

The potential impact of a genetic variant is associated with its location in the gene. Only 1% of the human genome codes for proteins; the remaining regions are intra/intergenic, promoters, enhancers and long stretches of ‘gene deserts', where genes are tens or hundreds of kilobases apart (79). It follows that a variant within a protein-coding region is potentially more detrimental to the cell than one located in a gene desert. In the present study, 21 of the 108 identified variants alter amino acid sequences (Table II). Two algorithms have been developed to evaluate the potential damage caused by these changes: SIFT (80) and PolyPhen-2 (81) scores. Both have values between 0 and 1, but the values have reciprocal interpretation. A variant with a SIFT score approaching 0 is considered deleterious, while one with a PolyPhen-2 score approaching 1 is considered damaging. Several variants in Table II are located in notable genes from an OC perspective: ANKLE1 and BRCA2, as discussed earlier in this review; BTD, which has shown promise as a biomarker for breast (82) and cervical (83) cancers; ZFHX3, which is a tumour suppressor gene frequently mutated in prostate (84) and endometrial (85) cancer; and LEKR, which, although the variant rs62273959 is considered benign, was found to be in tight LD (r2=0.90) with rs7651446 in the aforementioned TIPARP gene (77). Finally, it is worth noting that rs587778134 causes a frame shift mutation in the DNA repair gene BRIP1 (also known as FANCJ), which is a well-known OC susceptibility gene that interacts with BRCA1 (19,86). Although it is not found in the SIFT/PolyPhen-2 databases, it does have an entry for OC susceptibility (RCV000409984.1) in the ClinVar database of potentially clinically relevant genetic variants and is designated as ‘Likely pathogenic' (Table II) (87).

Table II

Functional relevance of identified susceptibility loci.Subtype.

Table II

Functional relevance of identified susceptibility loci.Subtype.

SNPLocusGeneaAmino acid change (Human
Genome variation society)
SIFT scorebSIFT
prediction
PolyPhen-2
scoreb
PolyPhen-2
prediction
ClinVar
clinical significance
ClinVar
phenotype
EnOCrs13803146801p36.33AGRN NP_001292204:p.Ala237Ser0.15Tolerated0.805Possibly damagingLikely benignNot specified
EnOCrs6175760402p22.2DHX57 NP_001316892:p.Gly49Ser0.06Tolerated-low confidence0.05benign
EOCrs20033737303p25.1BTD NP_001268654:p.Asp202Asn0.42Tolerated0.895Possibly damagingPathogenicBiotinidase deficiency
EOCrs7375739105q11.2AC025470.2 | ACTBL2 NP_001017992:p.Glu108Lys0.04 Deleterious-low0.989Probably damaging
EOCrs19976123806q25.2SYNE1 NP_149062:p.Asn4519AspConfidence0.517Possibly damagingUncertain significancenot specified
EOCrs14743249715q12ATP10A NP_077816:p.Arg999Cys0deleterious0.99Probably damaging
EOCrs58777813417q23.2BRIP1 NP_114432.2:p.Leu680fsLikely pathogenicNeoplasm of ovary
EOCrs1318119q13.32ERCC2 NP_000391:p.Lys751Gln0.45Tolerated0BenignBenignNSCLC
EOCrs14120030122q11.2MMP11 NP_005931:p.Arg334Cys0.01Deleterious0.984Probably damaging
HGSCrs6227395903q25.31LEKR1 NP_001004316:p.Val58Ile0.29Tolerated0.026Benign
HGSCrs38185205q11.2CDC20B | GPX8 NP_001008398:p.Lys182Arg0.62Tolerated0Benign
HGSCrs229798006p12.1TINAG NP_055279:p.Gln22Arg0.28Tolerated0.001Benign
HGSCrs13007206p21.33 CCHCR1 NP_061925:p.Arg627Gln0.03Deleterious0.998Probably damaging
HGSCrs207372406p21.33TCF19 NP_001070979:p.Pro241Leu0.1Tolerated0.521Possibly damaging
HGSCrs223397606p21.33 C6orf15 NP_054789:p.Gly48Arg0.11Tolerated0.324Benign
HGSCrs14551433311q13.1PYGM NP_005600:p.Arg61His0Deleterious0.947Probably damagingUncertainNot specified
HGSCrs1157183313q13BRCA2 NP_000050.2:p.Lys3326Tersignificance benignHereditary BC and OC
HGSCrs14744584616q22.3ZFHX3 NP_008816:p.Leu379Val0.04Deleterious0.703Possibly damaging
HGSCrs15032180917q21.2KRT13 | AC019349.1 NP_705694:p.Arg429His0.08Tolerated0.043BenignLikely benignWhite sponge nevus of cannon
HGSCrs10489465818p11.21MC2R NP_000520:p.Ser74Ile0Deleterious0.995Probably damagingPathogenicACTH resistance
HGSCrs236395619p13.11ANKLE1 NP_001265373:p.Leu185Trp0.03Deleterious0.162Benign

a Gene(s) in closest proximity to the most significant ovarian cancer risk variant at each locus.

b SIFT and PolyPhen-2 scores are reciprocal-both range from 0 to 1, with SIFT scores near 0 considered deleterious, and PolyPhen-2 scores near 0 considered benign. ACTH, adrenocorticotropic hormone; BC, breast cancer; EnOC, endometrioid ovarian cancer; EOC, Epithelial ovarian cancer (no subtype specified); HGSC, high grade serous carcinoma; NSCLC, non-small cell lung cancer; OC, ovarian cancer; SNP, single nucleotide polymorphism.

9. OC as a hormone‑related disease

OC is suspected to be a hormonal disease and related to breast and prostate cancers (88,89). Three new susceptibility loci were found by investigating a cohort of patients with breast or ovarian cancer harbouring mutations in BRCA1 (90). As part of the COGS initiative, the three cancers were examined in individual GWAS projects (20,32,91), whereas Kar et al (92) combined the results from the three projects in a single three-cancer meta-analysis, as well as one-by-one comparisons. The findings showed clear pleiotropy among the diseases, with three susceptibility loci identified in all three cancers (rs17041869, rs7937840 and rs1469713), and four loci shared between breast and ovarian cancers (rs635634, rs11571833, rs200182588 and rs8037137). No shared loci were found for prostate and ovarian cancer alone.

10. Most susceptibility loci are found outside classic OC causal genes

Surprisingly few of the susceptibility loci were found in genes commonly associated with OC, such as TP53 (93), BRCA2 (92) or HNF1B (36,37). This may be explained by the fact that a GWAS only detects susceptibility loci of a single or few nucleotides, often conferring subtle differences in gene expression, whereas some mutations in the classic causal genes are large deletions or inactivating mutations that are detrimental to normal protein function. Additionally, most of the variants have relatively low ORs (<2.0), meaning that they only have moderate effects on OC risk. Incidentally, the variant with the second-highest OR (rs587778134, OR=8.13) was located in BRIP1, and mutations in this gene have been established to confer a moderate to high risk of developing OC (19,86).

11. Future directions

Much attention has been focused on finding isolated susceptibility loci on a genome-wide scale over the past decade. A large number of the identified variants were situated in intergenic regions far from the genes they potentially affect, and while several GWAS have been performed and analysed, few studies have fine‑mapped and functionally validated any of the findings. With most loci situated in genes not previously associated with OC, including hits in long noncoding RNAs (94), there is an urgent requirement and potential for examining these further, particularly those that are near or in genes implicated in oocyte and ovary development, or tumour progression.

The focus must be on finding candidate causal genes for OC. Promising studies have been released in recent years, including the transcriptome-wide association study by Lu et al (95). In this study, they performed a ‘reverse GWAS' by cross‑matching existing OC‑specific gene expression profiles with all known susceptibility loci and candidate SNPs, and reported the Frizzled gene FZD4 as a novel candidate causal gene. This is an area that complements and overlaps well with the search for novel susceptibility loci.

We are currently performing whole exome sequencing of patients with HGSC and OCCC, to identify variants that are subtype‑ and survival‑specific. Combined with the published variants summarised in this review, a screen of a large number of patients with OC will be performed to identify potential biomarkers for the early detection of OC that may decrease the mortality rates for patients.

Supplementary Data

Funding

This study was supported by the Danish Mermaid III project, who was not involved in the decision to write the paper.

Availability of data and materials

All data generated or analyzed during this study are included in this published article.

Authors' contributions

All authors designed the study. MKC drafted and edited the manuscript, EH supervised and edited, and CH supervised. All authors contributed to and approved the final manuscript.

Ethics approval and consent to participate

Not applicable.

Patient consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Acknowledgements

Not applicable.

Abbreviations:

Breast

Cancer Association Consortium

CIMBA

The Consortium of Investigators of Modifiers of BRCA1/2

COGS

Collaborative Oncological Gene-environment Study

eQTL

expression quantitative trait loci

GWAS

genome-wide association study

HGSC

high-grade serous (ovarian) carcinoma

LD

linkage disequilibrium

NGS

next-generation sequencing

OC

ovarian cancer

OCAC

Ovarian Cancer Association Consortium

OCCC

ovarian clear cell carcinoma

OR

odds ratio

SNP

single nucleotide polymorphism

References

1 

Stewart BW and Wild CP: World Cancer Report 2014. 2014.

2 

Danish Gynecologic Cancer Group: Annual Report of the Danish Gynecologic Cancer Database 2016-17. Danish Gynecol Cancer Database. 2017.

3 

Goff BA, Mandel LS, Drescher CW, Urban N, Gough S, Schurman KM, Patras J, Mahony BS and Andersen MR: Development of an ovarian cancer symptom index: Possibilities for earlier detection. Cancer. 109:221–227. 2007. View Article : Google Scholar

4 

Hamilton W, Peters TJ, Bankhead C and Sharp D: Risk of ovarian cancer in women with symptoms in primary care: Population based case-control study. BMJ. 339:b29982009. View Article : Google Scholar : PubMed/NCBI

5 

Sundar S, Neal RD and Kehoe S: Diagnosis of ovarian cancer. BMJ. 351:h44432015. View Article : Google Scholar : PubMed/NCBI

6 

Köbel M, Kalloger SE, Boyd N, McKinney S, Mehl E, Palmer C, Leung S, Bowen NJ, Ionescu DN, Rajput A, et al: Ovarian carcinoma subtypes are different diseases: Implications for biomarker studies. PLoS Med. 5:e2322008. View Article : Google Scholar : PubMed/NCBI

7 

Prat J: Ovarian carcinomas: Five distinct diseases with different origins, genetic alterations, and clinicopathological features. Virchows Arch. 460:237–249. 2012. View Article : Google Scholar : PubMed/NCBI

8 

Vaughan S, Coward JI, Bast RC Jr, Berchuck A, Berek JS, Brenton JD, Coukos G, Crum CC, Drapkin R, Etemadmoghadam D, et al: Rethinking ovarian cancer: Recommendations for improving outcomes. Nat Rev Cancer. 11:719–725. 2011. View Article : Google Scholar : PubMed/NCBI

9 

Bast RC Jr, Klug TL, St John E, Jenison E, Niloff JM, Lazarus H, Berkowitz RS, Leavitt T, Griffiths CT, Parker L, et al: A radioim-munoassay using a monoclonal antibody to monitor the course of epithelial ovarian cancer. N Engl J Med. 309:883–887. 1983. View Article : Google Scholar : PubMed/NCBI

10 

Hellström I, Raycraft J, Hayden-Ledbetter M, Ledbetter JA, Schummer M, McIntosh M, Drescher C, Urban N and Hellström KE: The HE4 (WFDC2) protein is a biomarker for ovarian carcinoma. Cancer Res. 63:3695–3700. 2003.PubMed/NCBI

11 

Karlsen MA, Fagö-Olsen C, Høgdall E, Schnack TH, Christensen IJ, Nedergaard L, Lundvall L, Lydolph MC, Engelholm SA and Høgdall C: A novel index for preoperative, non-invasive prediction of macro-radical primary surgery in patients with stage IIIC-IV ovarian cancer-a part of the Danish prospective pelvic mass study. Tumor Biol. 37:12619–12626. 2016. View Article : Google Scholar

12 

Jacobs I, Oram D, Fairbanks J, Turner J, Frost C and Grudzinskas JG: A risk of malignancy index incorporating CA 125, ultrasound and menopausal status for the accurate preoperative diagnosis of ovarian cancer. Br J Obstet Gynaecol. 97:922–929. 1990. View Article : Google Scholar : PubMed/NCBI

13 

Moore RG, McMeekin DS, Brown AK, DiSilvestro P, Miller MC, Allard WJ, Gajewski W, Kurman R, Bast RC Jr and Skates SJ: A novel multiple marker bioassay utilizing HE4 and CA125 for the prediction of ovarian cancer in patients with a pelvic mass. Gynecol Oncol. 112:40–46. 2009. View Article : Google Scholar

14 

Skates SJ: Ovarian cancer screening: Development of the risk of ovarian cancer algorithm (ROCA) and ROCA screening trials. Int J Gynecol Cancer. 22(Suppl 1): S24–S26. 2012. View Article : Google Scholar : PubMed/NCBI

15 

Ueland FR, Desimone CP, Seamon LG, Miller RA, Goodrich S, Podzielinski I, Sokoll L, Smith A, van Nagell JR Jr and Zhang Z: Effectiveness of a multivariate index assay in the preoperative assessment of ovarian tumors. Obstet Gynecol. 117:1289–1297. 2011. View Article : Google Scholar : PubMed/NCBI

16 

Karlsen MA, Sandhu N, Høgdall C, Christensen IJ, Nedergaard L, Lundvall L, Engelholm SA, Pedersen AT, Hartwell D, Lydolph M, et al: Evaluation of HE4, CA125, risk of ovarian malignancy algorithm (ROMA) and risk of malignancy index (RMI) as diagnostic tools of epithelial ovarian cancer in patients with a pelvic mass. Gynecol Oncol. 127:379–383. 2012. View Article : Google Scholar : PubMed/NCBI

17 

Jacobs IJ and Menon U: Progress and challenges in screening for early detection of ovarian cancer. Mol Cell Proteomics. 3:355–366. 2004. View Article : Google Scholar : PubMed/NCBI

18 

Norquist BM, Harrell MI, Brady MF, Walsh T, Lee MK, Gulsuner S, Bernards SS, Casadei S, Yi Q, Burger RA, et al: Inherited mutations in women with ovarian carcinoma. JAMA Oncol. 2:482–490. 2016. View Article : Google Scholar : PubMed/NCBI

19 

Ramus SJ, Song H, Dicks E, Tyrer JP, Rosenthal AN, Intermaggio MP, Fraser L, Gentry-Maharaj A, Hayward J, Philpott S, et al: Germline mutations in the BRIP1, BARD1, PALB2, and NBN genes in women with ovarian cancer. J Natl Cancer Inst. 107:djv2142015. View Article : Google Scholar : PubMed/NCBI

20 

Kuchenbaecker KB, Ramus SJ, Tyrer J, Lee A, Shen HC, Beesley J, Lawrenson K, McGuffog L, Healey S, Lee JM, et al: Identification of six new susceptibility loci for invasive epithelial ovarian cancer. Nat Genet. 47:164–171. 2015. View Article : Google Scholar : PubMed/NCBI

21 

Moher D, Liberati A, Tetzlaff J and Altman DG; PRISMA Group: Preferred reporting items for systematic reviews and meta-analyses: The PRISMA statement. PLoS Med. 6:e10000972009. View Article : Google Scholar : PubMed/NCBI

22 

International Human Genome Sequencing Consortium: Finishing the euchromatic sequence of the human genome. Nature. 431:931–945. 2004. View Article : Google Scholar : PubMed/NCBI

23 

The 1000 Genomes Project Consortium; Auton A, Brooks LD, Durbin RM, Garrison EP, Kang HM, Korbel JO, Marchini JL, McCarthy S, McVean GA and Abecasis GR: A global reference for human genetic variation. Nature. 526:68–74. 2015. View Article : Google Scholar : PubMed/NCBI

24 

Metzker ML: Sequencing technologies-the next generation. Nat Rev Genet. 11:31–46. 2010. View Article : Google Scholar

25 

Ashley EA: Towards precision medicine. Nat Rev Genet. 17:507–522. 2016. View Article : Google Scholar : PubMed/NCBI

26 

Goodwin S, McPherson JD and McCombie WR: Coming of age: Ten years of next-generation sequencing technologies. Nat Rev Genet. 17:333–351. 2016. View Article : Google Scholar : PubMed/NCBI

27 

Blum A, Wang P and Zenklusen JC: SnapShot: TCGA-analyzed tumors. Cell. 173:5302018. View Article : Google Scholar : PubMed/NCBI

28 

Cancer Genome Atlas Research Network: Integrated genomic analyses of ovarian carcinoma. Nature. 474:609–615. 2011. View Article : Google Scholar : PubMed/NCBI

29 

Amos CI, Dennis J, Wang Z, Byun J, Schumacher FR, Gayther SA, Casey G, Hunter DJ, Sellers TA, Gruber SB, et al: The oncoarray consortium: A network for understanding the genetic architecture of common cancers. Cancer Epidemiol Biomarkers Prev. 26:126–135. 2017. View Article : Google Scholar :

30 

Sakoda LC, Jorgenson E and Witte JS: Turning of COGS moves forward findings for hormonally mediated cancers. Nat Genet. 45:345–348. 2013. View Article : Google Scholar : PubMed/NCBI

31 

Phelan CM, Kuchenbaecker KB, Tyrer JP, Kar SP, Lawrenson K, Winham SJ, Dennis J, Pirie A, Riggan MJ, Chornokur G, et al: Identification of 12 new susceptibility loci for different histotypes of epithelial ovarian cancer. Nat Genet. 49:680–691. 2017. View Article : Google Scholar : PubMed/NCBI

32 

Michailidou K, Beesley J, Lindstrom S, Canisius S, Dennis J, Lush MJ, Maranian MJ, Bolla MK, Wang Q, Shah M, et al: Genome-wide association analysis of more than 120,000 individuals identifies 15 new susceptibility loci for breast cancer. Nat Genet. 47:373–380. 2015. View Article : Google Scholar : PubMed/NCBI

33 

McKay JD, Hung RJ, Han Y, Zong X, Carreras-Torres R, Christiani DC, Caporaso NE, Johansson M, Xiao X, Li Y, et al: Large-scale association analysis identifies new lung cancer susceptibility loci and heterogeneity in genetic susceptibility across histological subtypes. Nat Genet. 49:1126–1132. 2017. View Article : Google Scholar : PubMed/NCBI

34 

Schumacher FR, Al Olama AA, Berndt SI, Benlloch S, Ahmed M, Saunders EJ, Dadaev T, Leongamornlert D, Anokian E, Cieza-Borrella C, et al: Association analyses of more than 140,000 men identify 63 new prostate cancer susceptibility loci. Nat Genet. 50:928–936. 2018. View Article : Google Scholar : PubMed/NCBI

35 

Melin BS, Barnholtz-Sloan JS, Wrensch MR, Johansen C, Il'yasova D, Kinnersley B, Ostrom QT, Labreche K, Chen Y, Armstrong G, et al: Genome-wide association study of glioma subtypes identifies specific differences in genetic susceptibility to glioblastoma and non-glioblastoma tumors. Nat Genet. 49:789–794. 2017. View Article : Google Scholar : PubMed/NCBI

36 

Shen H, Fridley BL, Song H, Lawrenson K, Cunningham JM, Ramus SJ, Cicek MS, Tyrer J, Stram D, Larson MC, et al: Epigenetic analysis leads to identification of HNF1B as a subtype-specific susceptibility gene for ovarian cancer. Nat Commun. 4:16282013. View Article : Google Scholar : PubMed/NCBI

37 

Pharoah PD, Tsai YY, Ramus SJ, Phelan CM, Goode EL, Lawrenson K, Buckley M, Fridley BL, Tyrer JP, Shen H, et al: GWAS meta‑analysis and replication identifies three new susceptibility loci for ovarian cancer. Nat Genet. 45:362–370. 2013. View Article : Google Scholar

38 

Kelemen LE, Lawrenson K, Tyrer J, Li Q, Lee JM, Seo JH, Phelan CM, Beesley J, Chen X, Spindler TJ, et al: Genome-wide significant risk associations for mucinous ovarian carcinoma. Nat Genet. 47:888–897. 2015. View Article : Google Scholar : PubMed/NCBI

39 

Tsuchiya A, Sakamoto M, Yasuda J, Chuma M, Ohta T, Ohki M, Yasugi T, Taketani Y and Hirohashi S: Expression profiling in ovarian clear cell carcinoma: Identification of hepatocyte nuclear factor-1beta as a molecular marker and a possible molecular target for therapy of ovarian clear cell carcinoma. Am J Pathol. 163:2503–2512. 2003. View Article : Google Scholar : PubMed/NCBI

40 

Gudmundsson J, Sulem P, Steinthorsdottir V, Bergthorsson JT, Thorleifsson G, Manolescu A, Rafnar T, Gudbjartsson D, Agnarsson BA, Baker A, et al: Two variants on chromosome 17 confer prostate cancer risk and the one in TCF2 protects against type 2 diabetes. Nat Genet. 39:977–983. 2007. View Article : Google Scholar : PubMed/NCBI

41 

Sun J, Zheng SL, Wiklund F, Isaacs SD, Purcell LD, Gao Z, Hsu FC, Kim ST, Liu W, Zhu Y, et al: Evidence for two independent prostate cancer risk-associated loci in the HNF1B gene at 17q12. Nat Genet. 40:1153–1155. 2008. View Article : Google Scholar : PubMed/NCBI

42 

Thomas G, Jacobs KB, Yeager M, Kraft P, Wacholder S, Orr N, Yu K, Chatterjee N, Welch R, Hutchinson A, et al: Multiple loci identified in a genome‑wide association study of prostate cancer. Nat Genet. 40:310–315. 2008. View Article : Google Scholar : PubMed/NCBI

43 

Spurdle AB, Thompson DJ, Ahmed S, Ferguson K, Healey CS, O'Mara T, Walker LC, Montgomery SB, Dermitzakis ET; Australian National Endometrial Cancer Study Group; et al: Genome-wide association study identifies a common variant associated with risk of endometrial cancer. Nat Genet. 43:451–455. 2011. View Article : Google Scholar : PubMed/NCBI

44 

Painter JN, O'Mara TA, Batra J, Cheng T, Lose FA, Dennis J, Michailidou K, Tyrer JP, Ahmed S, Ferguson K, et al: Fine‑mapping of the HNF1B multicancer locus identifies candidate variants that mediate endometrial cancer risk. Hum Mol Genet. 24:1478–1492. 2015. View Article : Google Scholar

45 

Setiawan VW, Haessler J, Schumacher F, Cote ML, Deelman E, Fesinmeyer MD, Henderson BE, Jackson RD, Vöckler JS, Wilkens LR, et al: HNF1B and endometrial cancer risk: Results from the PAGE study. PLoS One. 7:e303902012. View Article : Google Scholar : PubMed/NCBI

46 

Ross-Adams H, Ball S, Lawrenson K, Halim S, Russell R, Wells C, Strand SH, Ørntoft TF, Larson M, Armasu S, et al: HNF1B variants associate with promoter methylation and regulate gene networks activated in prostate and ovarian cancer. Oncotarget. 7:74734–74746. 2016. View Article : Google Scholar : PubMed/NCBI

47 

Pooley KA, Bojesen SE, Weischer M, Nielsen SF, Thompson D, Amin Al Olama A, Michailidou K, Tyrer JP, Benlloch S, Brown J, et al: A genome-wide association scan (GWAS) for mean telomere length within the COGS project: Identified loci show little association with hormone-related cancer risk. Hum Mol Genet. 22:5056–5064. 2013. View Article : Google Scholar : PubMed/NCBI

48 

Bojesen SE, Pooley KA, Johnatty SE, Beesley J, Michailidou K, Tyrer JP, Edwards SL, Pickett HA, Shen HC, Smart CE, et al: Multiple independent variants at the TERT locus are associated with telomere length and risks of breast and ovarian cancer. Nat Genet. 45:371–384. 2013. View Article : Google Scholar : PubMed/NCBI

49 

Vinagre J, Almeida A, Pópulo H, Batista R, Lyra J, Pinto V, Coelho R, Celestino R, Prazeres H, Lima L, et al: Frequency of TERT promoter mutations in human cancers. Nat Commun. 4:21852013. View Article : Google Scholar : PubMed/NCBI

50 

Bolton KL, Tyrer J, Song H, Ramus SJ, Notaridou M, Jones C, Sher T, Gentry-Maharaj A, Wozniak E, Tsai YY, et al: Common variants at 19p13 are associated with susceptibility to ovarian cancer. Nat Genet. 42:880–884. 2010. View Article : Google Scholar : PubMed/NCBI

51 

Shao G, Patterson-Fortin J, Messick TE, Feng D, Shanbhag N, Wang Y and Greenberg RA: MERIT40 controls BRCA1-Rap80 complex integrity and recruitment to DNA double-strand breaks. Genes Dev. 23:740–754. 2009. View Article : Google Scholar : PubMed/NCBI

52 

Wang B, Hurov K, Hofmann K and Elledge SJ: NBA1, a new player in the Brca1 A complex, is required for DNA damage resistance and checkpoint control. Genes Dev. 23:729–739. 2009. View Article : Google Scholar : PubMed/NCBI

53 

Feng L, Huang J and Chen J: MERIT40 facilitates BRCA1 localization and DNA damage repair. Genes Dev. 23:719–728. 2009. View Article : Google Scholar : PubMed/NCBI

54 

Lawrenson K, Kar S, McCue K, Kuchenbaeker K, Michailidou K, Tyrer J, Beesley J, Ramus SJ, Li Q, Delgado MK, et al: Functional mechanisms underlying pleiotropic risk alleles at the 19p13.1 breast-ovarian cancer susceptibility locus. Nat Commun. 7:126752016. View Article : Google Scholar : PubMed/NCBI

55 

Kar SP, Tyrer JP, Li Q, Lawrenson K, Aben KK, Anton-Culver H, Antonenkova N, Chenevix-Trench G; Australian Cancer Study; Australian Ovarian Cancer Study Group; et al: Network-based integration of GWAS and gene expression identifies a HOX-Centric network associated with serous ovarian cancer risk. Cancer Epidemiol Biomarkers Prev. 24:1574–1584. 2015. View Article : Google Scholar : PubMed/NCBI

56 

Lawrenson K, Li Q, Kar S, Seo JH, Tyrer J, Spindler TJ, Lee J, Chen Y, Karst A, Drapkin R, et al: Cis-eQTL analysis and functional validation of candidate susceptibility genes for high-grade serous ovarian cancer. Nat Commun. 6:82342015. View Article : Google Scholar : PubMed/NCBI

57 

Ghoussaini M, Song H, Koessler T, Al Olama AA, Kote-Jarai Z, Driver KE, Pooley KA, Ramus SJ, Kjaer SK, Hogdall E, et al: Multiple loci with different cancer specificities within the 8q24 gene desert. J Natl Cancer Inst. 100:962–966. 2008. View Article : Google Scholar : PubMed/NCBI

58 

Panagiotou OA, Evangelou E and Ioannidis JP: Genome-wide significant associations for variants with minor allele frequency of 5% or less-an overview: A HuGE review. Am J Epidemiol. 172:869–889. 2010. View Article : Google Scholar : PubMed/NCBI

59 

Stephens M and Balding DJ: Bayesian statistical methods for genetic association studies. Nat Rev Genet. 10:681–690. 2009. View Article : Google Scholar : PubMed/NCBI

60 

International HapMap Consortium: A haplotype map of the human genome. Nature. 437:1299–1320. 2005. View Article : Google Scholar : PubMed/NCBI

61 

Pe'er I, Yelensky R, Altshuler D and Daly MJ: Estimation of the multiple testing burden for genomewide association studies of nearly all common variants. Genet Epidemiol. 32:381–385. 2008. View Article : Google Scholar : PubMed/NCBI

62 

Fadista J, Manning AK, Florez JC and Groop L: The (in)famous GWAS P-value threshold revisited and updated for low-frequency variants. Eur J Hum Genet. 24:1202–1205. 2016. View Article : Google Scholar : PubMed/NCBI

63 

Kar SP, Berchuck A, Gayther SA, Goode EL, Moysich KB, Pearce CL, Ramus SJ, Schildkraut JM, Sellers TA and Pharoah PDP: Common genetic variation and susceptibility to ovarian cancer: Current insights and future directions. Cancer Epidemiol Biomarkers Prev. 27:395–404. 2018. View Article : Google Scholar

64 

Jones MR, Kamara D, Karlan BY, Pharoah PDP and Gayther SA: Genetic epidemiology of ovarian cancer and prospects for poly-genic risk prediction. Gynecol Oncol. 147:705–713. 2017. View Article : Google Scholar : PubMed/NCBI

65 

Elmas A, Ou Yang TH, Wang X and Anastassiou D: Discovering genome-wide tag SNPs based on the mutual information of the variants. PLoS One. 11:e01679942016. View Article : Google Scholar : PubMed/NCBI

66 

Johnson GC, Esposito L, Barratt BJ, Smith AN, Heward J, Di Genova G, Ueda H, Cordell HJ, Eaves IA, Dudbridge F, et al: Haplotype tagging for the identification of common disease genes. Nat Genet. 29:233–237. 2001. View Article : Google Scholar : PubMed/NCBI

67 

Earp M, Winham SJ, Larson N, Permuth JB, Sicotte H, Chien J, Anton-Culver H, Bandera EV, Berchuck A, Cook LS, et al: A targeted genetic association study of epithelial ovarian cancer susceptibility. Oncotarget. 7:7381–7389. 2016. View Article : Google Scholar : PubMed/NCBI

68 

Song H, Ramus SJ, Tyrer J, Bolton KL, Gentry-Maharaj A, Wozniak E, Anton-Culver H, Chang-Claude J, Cramer DW, DiCioccio R, et al: A genome‑wide association study identifies a new ovarian cancer susceptibility locus on 9p22.2. Nat Genet. 41:996–1000. 2009. View Article : Google Scholar : PubMed/NCBI

69 

Goode EL, Chenevix-Trench G, Song H, Ramus SJ, Notaridou M, Lawrenson K, Widschwendter M, Vierkant RA, Larson MC, Kjaer SK, et al: A genome-wide association study identifies susceptibility loci for ovarian cancer at 2q31 and 8q24. Nat Genet. 42:874–879. 2010. View Article : Google Scholar : PubMed/NCBI

70 

Permuth-Wey J, Lawrenson K, Shen HC, Velkova A, Tyrer JP, Chen Z, Lin HY, Chen YA, Tsai YY, Qu X, et al: Identification and molecular characterization of a new ovarian cancer susceptibility locus at 17q21.31. Nat Commun. 4:16272013. View Article : Google Scholar : PubMed/NCBI

71 

Chen K, Ma H, Li L, Zang R, Wang C, Song F, Shi T, Yu D, Yang M, Xue W, et al: Genome-wide association study identifies new susceptibility loci for epithelial ovarian cancer in Han Chinese women. Nat Commun. 5:46822014. View Article : Google Scholar : PubMed/NCBI

72 

Romano RA, Li H, Tummala R, Maul R and Sinha S: Identification of Basonuclin2, a DNA‑binding zinc‑finger protein expressed in germ tissues and skin keratinocytes. Genomics. 83:821–833. 2004. View Article : Google Scholar : PubMed/NCBI

73 

Buckley MA, Woods NT, Tyrer JP, Mendoza-Fandiño G, Lawrenson K, Hazelett DJ, Najafabadi HS, Gjyshi A, Carvalho RS, Lyra PC Jr, et al: Functional analysis and fine mapping of the 9p22.2 ovarian cancer susceptibility locus. Cancer Res. 79:467–481. 2019. View Article : Google Scholar

74 

Wentzensen N, Black A, Jacobs K, Yang HP, Berg CD, Caporaso N, Peters U, Ragard L, Buys SS, Chanock S and Hartge P: Genetic variation on 9p22 is associated with abnormal ovarian ultrasound results in the prostate, lung, colorectal, and ovarian cancer screening trial. PLoS One. 6:e217312011. View Article : Google Scholar : PubMed/NCBI

75 

Vigorito E, Kuchenbaecker KB, Beesley J, Adlard J, Agnarsson BA, Andrulis IL, Arun BK, Barjhoux L, Belotti M, Benitez J, et al: Fine‑scale mapping at 9p22.2 identifies candidate causal variants that modify ovarian cancer risk in BRCA1 and BRCA2 mutation carriers. PLoS One. 11:e01588012016. View Article : Google Scholar

76 

Carter H, Marty R, Hofree M, Gross AM, Jensen J, Fisch KM, Wu X, DeBoever C, Van Nostrand EL, Song Y, et al: Interaction landscape of inherited polymorphisms with somatic events in cancer. Cancer Discov. 7:410–423. 2017. View Article : Google Scholar : PubMed/NCBI

77 

Permuth JB, Pirie A, Ann Chen Y, Lin HY, Reid BM, Chen Z, Monteiro A, Dennis J, Mendoza-Fandino G; AOCS Study Group; et al: Exome genotyping arrays to identify rare and low frequency variants associated with epithelial ovarian cancer risk. Hum Mol Genet. 25:3600–3612. 2016. View Article : Google Scholar : PubMed/NCBI

78 

Fong PC, Boss DS, Yap TA, Tutt A, Wu P, Mergui-Roelvink M, Mortimer P, Swaisland H, Lau A, O'Connor MJ, et al: Inhibition of poly(ADP-Ribose) polymerase in tumors from BRCA mutation carriers. N Engl J Med. 361:123–134. 2009. View Article : Google Scholar : PubMed/NCBI

79 

Venter JC, Adams MD, Myers EW, Li PW, Mural RJ, Sutton GG, Smith HO, Yandell M, Evans CA, Holt RA, et al: The sequence of the human genome. Science. 291:1304–1351. 2001. View Article : Google Scholar : PubMed/NCBI

80 

Vaser R, Adusumalli S, Leng SN, Sikic M and Ng PC: SIFT missense predictions for genomes. Nat Protoc. 11:1–9. 2016. View Article : Google Scholar

81 

Adzhubei IA, Schmidt S, Peshkin L, Ramensky VE, Gerasimova A, Bork P, Kondrashov AS and Sunyaev SR: A method and server for predicting damaging missense mutations. Nat Methods. 7:248–249. 2010. View Article : Google Scholar : PubMed/NCBI

82 

Kang UB, Ahn Y, Lee JW, Kim YH, Kim J, Yu MH, Noh DY and Lee C: Differential profiling of breast cancer plasma proteome by isotope‑coded affinity tagging method reveals biotinidase as a breast cancer biomarker. BMC Cancer. 10:1142010. View Article : Google Scholar

83 

Huang L, Zheng M, Zhou QM, Zhang MY, Jia WH, Yun JP and Wang HY: Identification of a gene‑expression signature for predicting lymph node metastasis in patients with early stage cervical carcinoma. Cancer. 117:3363–3373. 2011. View Article : Google Scholar : PubMed/NCBI

84 

Sun X, Frierson HF, Chen C, Li C, Ran Q, Otto KB, Cantarel BL, Vessella RL, Gao AC, Petros J, et al: Frequent somatic mutations of the transcription factor ATBF1 in human prostate cancer. Nat Genet. 37:407–412. 2005. View Article : Google Scholar : PubMed/NCBI

85 

Walker CJ, Miranda MA, O'Hern MJ, McElroy JP, Coombes KR, Bundschuh R, Cohn DE, Mutch DG and Goodfellow PJ: Patterns of CTCF and ZFHX3 mutation and associated outcomes in endo-metrial cancer. J Natl Cancer Inst. 107:djv2492015. View Article : Google Scholar

86 

Rafnar T, Gudbjartsson DF, Sulem P, Jonasdottir A, Sigurdsson A, Jonasdottir A, Besenbacher S, Lundin P, Stacey SN, Gudmundsson J, et al: Mutations in BRIP1 confer high risk of ovarian cancer. Nat Genet. 43:1104–1107. 2011. View Article : Google Scholar : PubMed/NCBI

87 

Landrum MJ, Lee JM, Benson M, Brown GR, Chao C, Chitipiralla S, Gu B, Hart J, Hoffman D, Jang W, et al: ClinVar: Improving access to variant interpretations and supporting evidence. Nucleic Acids Res. 46:D1062–D1067. 2018. View Article : Google Scholar :

88 

Henderson BE and Feigelson HS: Hormonal carcinogenesis. Carcinogenesis. 21:427–433. 2000. View Article : Google Scholar : PubMed/NCBI

89 

Henderson BE, Ross RK, Pike MC and Casagrande JT: Endogenous hormones as a major factor in human cancer. Cancer Res. 42:3232–3239. 1982.PubMed/NCBI

90 

Couch FJ, Wang X, McGuffog L, Lee A, Olswold C, Kuchenbaecker KB, Soucy P, Fredericksen Z, Barrowdale D, Dennis J, et al: Genome-wide association study in BRCA1 mutation carriers identifies novel loci associated with breast and ovarian cancer risk. PLoS Genet. 9:e10032122013. View Article : Google Scholar : PubMed/NCBI

91 

Al Olama AA, Kote-Jarai Z, Berndt SI, Conti DV, Schumacher F, Han Y, Benlloch S, Hazelett DJ, Wang Z, Saunders E, et al: A meta‑analysis of 87,040 individuals identifies 23 new susceptibility loci for prostate cancer. Nat Genet. 46:1103–1109. 2014. View Article : Google Scholar : PubMed/NCBI

92 

Kar SP, Beesley J, Amin Al Olama A, Michailidou K, Tyrer J, Kote-Jarai Z, Lawrenson K, Lindstrom S, Ramus SJ, Thompson DJ, et al: Genome-wide meta-analyses of breast, ovarian, and prostate cancer association studies identify multiple new susceptibility loci shared by at least two cancer types. Cancer Discov. 6:1052–1067. 2016. View Article : Google Scholar : PubMed/NCBI

93 

Schildkraut JM, Goode EL, Clyde MA, Iversen ES, Moorman PG, Berchuck A, Marks JR, Lissowska J, Brinton L, Peplonska B, et al: Single nucleotide polymorphisms in the TP53 region and susceptibility to invasive epithelial ovarian cancer. Cancer Res. 69:2349–2357. 2009. View Article : Google Scholar : PubMed/NCBI

94 

Johnatty SE, Tyrer JP, Kar S, Beesley J, Lu Y, Gao B, Fasching PA, Hein A, Ekici AB, Beckmann MW, et al: Genome-wide analysis identifies novel loci associated with ovarian cancer outcomes: Findings from the ovarian cancer association consortium. Clin Cancer Res. 21:5264–5276. 2015. View Article : Google Scholar : PubMed/NCBI

95 

Lu Y, Beeghly-Fadiel A, Wu L, Guo X, Li B, Schildkraut JM, Im HK, Chen YA, Permuth JB, Reid BM, et al: A transcrip-tome-wide association study among 97,898 women to identify candidate susceptibility genes for epithelial ovarian cancer risk. Cancer Res. 78:5419–5430. 2018. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

November-2019
Volume 44 Issue 5

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

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Christophersen MK, Høgdall C and Høgdall E: The prospect of discovering new biomarkers for ovarian cancer based on current knowledge of susceptibility loci and genetic variation (Review). Int J Mol Med 44: 1599-1608, 2019
APA
Christophersen, M.K., Høgdall, C., & Høgdall, E. (2019). The prospect of discovering new biomarkers for ovarian cancer based on current knowledge of susceptibility loci and genetic variation (Review). International Journal of Molecular Medicine, 44, 1599-1608. https://doi.org/10.3892/ijmm.2019.4352
MLA
Christophersen, M. K., Høgdall, C., Høgdall, E."The prospect of discovering new biomarkers for ovarian cancer based on current knowledge of susceptibility loci and genetic variation (Review)". International Journal of Molecular Medicine 44.5 (2019): 1599-1608.
Chicago
Christophersen, M. K., Høgdall, C., Høgdall, E."The prospect of discovering new biomarkers for ovarian cancer based on current knowledge of susceptibility loci and genetic variation (Review)". International Journal of Molecular Medicine 44, no. 5 (2019): 1599-1608. https://doi.org/10.3892/ijmm.2019.4352