Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Trisomy 21 Alters DNA Methylation in Parent-of-Origin-Dependent and -Independent Manners

  • Antônio Francisco Alves da Silva ,

    anttonio.allves@gmail.com (AFAS); filipebma@yahoo.com.br (FBM); quique@uenf.br (EM-A)

    Affiliations Laboratory of Biotechnology, Universidade Estadual do Norte Fluminense Darcy Ribeiro, Campos do Goytacazes, Rio de Janeiro, Brazil, Molecular Identification and Diagnostics Unit, Hospital Escola Álvaro Alvim, Campos dos Goytacazes, Rio de Janeiro, Brazil, Graduate Program in Biosciences and Biotechnology, Center for Biosciences and Biotechnology, Universidade Estadual do Norte Fluminense Darcy Ribeiro, Rio de Janeiro, Brazil

  • Filipe Brum Machado ,

    anttonio.allves@gmail.com (AFAS); filipebma@yahoo.com.br (FBM); quique@uenf.br (EM-A)

    Affiliations Laboratory of Biotechnology, Universidade Estadual do Norte Fluminense Darcy Ribeiro, Campos do Goytacazes, Rio de Janeiro, Brazil, Molecular Identification and Diagnostics Unit, Hospital Escola Álvaro Alvim, Campos dos Goytacazes, Rio de Janeiro, Brazil, Postgraduate Program in Biosciences and Biotechnology, Center for Biosciences and Biotechnology, Universidade Estadual do Norte Fluminense Darcy Ribeiro, Rio de Janeiro, Brazil

  • Érika Cristina Pavarino,

    Affiliation Faculdade de Medicina de São José do Rio Preto, São Paulo, Brazil

  • Joice Matos Biselli-Périco,

    Affiliation Faculdade de Medicina de São José do Rio Preto, São Paulo, Brazil

  • Bruna Lancia Zampieri,

    Affiliation Faculdade de Medicina de São José do Rio Preto, São Paulo, Brazil

  • Ronaldo da Silva Francisco Junior,

    Affiliations Laboratory of Biotechnology, Universidade Estadual do Norte Fluminense Darcy Ribeiro, Campos do Goytacazes, Rio de Janeiro, Brazil, Molecular Identification and Diagnostics Unit, Hospital Escola Álvaro Alvim, Campos dos Goytacazes, Rio de Janeiro, Brazil

  • Pedro Thyago Mozer Rodrigues,

    Affiliations Laboratory of Biotechnology, Universidade Estadual do Norte Fluminense Darcy Ribeiro, Campos do Goytacazes, Rio de Janeiro, Brazil, Molecular Identification and Diagnostics Unit, Hospital Escola Álvaro Alvim, Campos dos Goytacazes, Rio de Janeiro, Brazil

  • Douglas Terra Machado,

    Affiliations Laboratory of Biotechnology, Universidade Estadual do Norte Fluminense Darcy Ribeiro, Campos do Goytacazes, Rio de Janeiro, Brazil, Molecular Identification and Diagnostics Unit, Hospital Escola Álvaro Alvim, Campos dos Goytacazes, Rio de Janeiro, Brazil

  • Cíntia Barros Santos-Rebouças,

    Affiliation Department of Genetics, Universidade do Estado do Rio de Janeiro, Rio de Janeiro, Rio de Janeiro, Brazil

  • Maria Gomes Fernandes,

    Affiliation Department of Anatomy and Embryology, Leiden University Medical Center, Leiden, South Holland, The Netherlands

  • Susana Marina Chuva de Sousa Lopes,

    Affiliation Department of Anatomy and Embryology, Leiden University Medical Center, Leiden, South Holland, The Netherlands

  • Álvaro Fabricio Lopes Rios,

    Affiliation Laboratory of Biotechnology, Universidade Estadual do Norte Fluminense Darcy Ribeiro, Campos do Goytacazes, Rio de Janeiro, Brazil

  • Enrique Medina-Acosta

    anttonio.allves@gmail.com (AFAS); filipebma@yahoo.com.br (FBM); quique@uenf.br (EM-A)

    Affiliations Laboratory of Biotechnology, Universidade Estadual do Norte Fluminense Darcy Ribeiro, Campos do Goytacazes, Rio de Janeiro, Brazil, Molecular Identification and Diagnostics Unit, Hospital Escola Álvaro Alvim, Campos dos Goytacazes, Rio de Janeiro, Brazil, Graduate Program in Biosciences and Biotechnology, Center for Biosciences and Biotechnology, Universidade Estadual do Norte Fluminense Darcy Ribeiro, Rio de Janeiro, Brazil

Abstract

The supernumerary chromosome 21 in Down syndrome differentially affects the methylation statuses at CpG dinucleotide sites and creates genome-wide transcriptional dysregulation of parental alleles, ultimately causing diverse pathologies. At present, it is unknown whether those effects are dependent or independent of the parental origin of the nondisjoined chromosome 21. Linkage analysis is a standard method for the determination of the parental origin of this aneuploidy, although it is inadequate in cases with deficiency of samples from the progenitors. Here, we assessed the reliability of the epigenetic 5mCpG imprints resulting in the maternally (oocyte)-derived allele methylation at a differentially methylated region (DMR) of the candidate imprinted WRB gene for asserting the parental origin of chromosome 21. We developed a methylation-sensitive restriction enzyme-specific PCR assay, based on the WRB DMR, across single nucleotide polymorphisms (SNPs) to examine the methylation statuses in the parental alleles. In genomic DNA from blood cells of either disomic or trisomic subjects, the maternal alleles were consistently methylated, while the paternal alleles were unmethylated. However, the supernumerary chromosome 21 did alter the methylation patterns at the RUNX1 (chromosome 21) and TMEM131 (chromosome 2) CpG sites in a parent-of-origin-independent manner. To evaluate the 5mCpG imprints, we conducted a computational comparative epigenomic analysis of transcriptome RNA sequencing (RNA-Seq) and histone modification expression patterns. We found allele fractions consistent with the transcriptional biallelic expression of WRB and ten neighboring genes, despite the similarities in the confluence of both a 17-histone modification activation backbone module and a 5-histone modification repressive module between the WRB DMR and the DMRs of six imprinted genes. We concluded that the maternally inherited 5mCpG imprints at the WRB DMR are uncoupled from the parental allele expression of WRB and ten neighboring genes in several tissues and that trisomy 21 alters DNA methylation in parent-of-origin-dependent and -independent manners.

Introduction

Trisomy 21 (Down syndrome) is the most common autosomal aneuploidy that is compatible with life (average rate of 1/400-800 live births; average life expectancy of 55 years) [1]. The supernumerary chromosome 21 results from meiotic nondisjunction errors in approximately 90–95% of cases during oogenesis [2, 3]. Thus, most individuals with Down syndrome inherit two maternal complete and free copies of chromosome 21. Advanced maternal age increases the risk of pregnancy with trisomy 21 [4], while the evidence for an association with paternal age is inconsistent [58]. Therefore, the extremely skewed disparity observed between the maternal and paternal meiotic errors at the origin of chromosome 21 nondisjunction is mainly explained by the effect of advanced maternal age.

Although individuals with Down syndrome share phenotypically distinctive traits including clinical manifestations of atypical and segmental accelerated aging [9], the syndrome exhibits a large variety of physical stigmata that are unevenly represented among probands. Some of the defects may constitute severe pathologies (i.e., cognitive dysfunction, acute lymphoblastic leukemia, congenital heart disease, premature aging and Alzheimer disease-like neuropathology). Despite the considerable number of clinical studies on the intellectual and physical disabilities associated with trisomy 21, and owing to the small proportion of paternally inherited cases, there is still no conclusive information regarding whether parent-of-origin (maternal versus paternal) genetic factors contribute differentially to the observed phenotypic variation. For example, in one study [10], no significant difference was found in the distribution of phenotypic clinical findings between Down syndrome patients with maternal (n = 150) and paternal (n = 8) origin of the nondisjunction errors. In another study [11], congenital heart defects, high arched palate, and short fingers occurred less frequently in cases with a paternally-derived extra chromosome 21 (n = 8) than in cases with a maternally-derived extra chromosome 21 (n = 28). The main caveats in those reports are the use of a limited number of polymorphic markers in the first study mentioned and the use of nucleolar organizer region heteromorphisms in the second study mentioned for the conclusive determination of the paternal origin of the nondisjunction errors.

Considering the scenario of genomic imprinting on chromosome 21 [12], the parental origin of the supernumerary chromosome 21 will contribute differentially to the development of some defects in a parent-dependent manner in Down syndrome. Notably, the uncommon condition of a normal phenotype with maternal [1315] or paternal [16, 17] uniparental disomy (UPD) of chromosome 21 does not rule out the possibility of genomic imprinting of non-essential genes. In fact, UPD involving chromosomes containing imprinted genes does not necessarily reveal an imprinted pathological disorder with clinical consequences [18]. Importantly, the extra copy of chromosome 21 causes a genome-wide [19, 20] domain pattern [20] of dysregulation of gene expression, and it affects the DNA methylation levels differentially at distinct CpG dinucleotide sites [21, 22]. At present, it is also is unknown whether those effects are dependent or independent of the parental origin of the nondisjoined chromosome 21.

The parental origin of the nondisjoined chromosome 21 can be established experimentally in nuclear trios (mother, father, and proband) by linkage analysis using highly polymorphic DNA markers (i.e., genotyping with short tandem repeats, STRs) [23]. Such expert analysis is difficult when DNA samples from the progenitors are unavailable. Here, we assessed the dependability of reversible epigenetic molecular imprints resulting in germline-specific 5mCpG for the discrimination of the parental origin of chromosome 21 nondisjunction. For that purpose, we developed a PCR assay based on maternally derived allele methylation at a differentially methylated region (DMR) located in the tryptophan-rich basic protein WRB gene. Two research groups recently identified the target maternal-of-origin 5mCpG imprints at the WRB DMR using genome-wide methylation chip technology [24, 25]. The occurrence of a DMR in the WRB gene warranted campaigning it as the first candidate maternally imprinted gene (i.e., paternally expressed) on the human chromosome 21. In contrast to the uniparental inheritance pattern of allele expression determined by imprinting, one study reported alternate (i.e., opposing) monoallelic expression of the maternal or the paternal WRB alleles in different human fetal tissues from the same embryo [25], a pattern consistent with random monoallelic expression, rather than with imprinting. Thus, we also reappraised the candidate imprinting status of the WRB gene. We performed single-nucleotide polymorphism primer extension (SNuPE) at informative 3´-UTR SNP variants in DNA from blood cells, and in human embryonic stem cell lines (hESCs). Furthermore, we conducted an integrative and comparative epigenomic computational analysis using transcriptome RNA sequencing (RNA-Seq) public repositories, essentially adopting the frameworks recently described to explore enrichment-based sequencing data [26, 27].

Finally, we also evaluated the impact of the parental origin of the supernumerary chromosome 21 on the methylation statuses at CpG sites in the RUNX1 (chromosome 21) and TMEM131 (chromosome 2) genes, the methylation levels of which are altered with trisomy 21 [21, 22, 28].

Materials and Methods

Ethics Statement

Peripheral blood samples from participating control and Down syndrome nuclear families were collected with written informed consent. For infants and children with Down syndrome, a surrogate consent procedure was used, whereby the next of kin or a legally authorized representative consented in writing on the behalf of the participants. The subjects were included from projects approved by the Ethics Committee of the Faculdade de Medicina de Campos, Brazil (approval code FR-278769), the Faculdade de Medicina de São José do Rio Preto, Brazil (HCRP 5810/2009) and the Universidade do Estado do Rio de Janeiro, Brazil (040/2005). The main objectives of those projects were to develop molecular genetic tests for the determination of the parental origin of the nondisjunction of human chromosome 21 and to screen for parent-of-origin genetic risk factors for trisomy 21. This study was conducted according to the principles expressed in the Declaration of Helsinki.

Subjects

We included 70 nuclear families (mother, father, and child) of either male or female index cases with the full free trisomy 21 status established by conventional karyotyping (G band) in at least 20 cells. We also included 20 families of children with no trisomy (controls). The average maternal age of index cases was 31.6 years (ranging from 16 to 44 years). After determining the parental origin of the nondisjunction error for at least three informative STRs and selecting for the informative SNPs in the target loci, 15 trios with a maternal origin (MT21), six with a paternal origin (PT21) and 13 control trios (N21) met the criteria for the investigation of parental-of-origin allele methylation and allele expression.

Cell Lines

Cell pellets from the commercial, established human embryonic stem cell lines HUES 1, HUES 3, HUES 5, HUES 7, and HUES 15 were kindly provided by Christine L. Mummery, from the Department of Anatomy and Embryology, Leiden University Medical Center, Leiden, The Netherlands. General information regarding these cell lines is available at the NIH Human Embryonic Stem Cell Registry (http://grants.nih.gov/stem_cells/registry/current.htm)

DNA and RNA Extraction

Human genomic DNA and total RNA from freshly drawn peripheral blood samples were extracted using phenol-chloroform and ethanol precipitation [29] and TRIzol reagent (Invitrogen, Carlsbad, CA, USA), respectively, as previously described [30].

Determination of the Parental Origin of Chromosome 21 Nondisjunction through Linkage Analysis

To select cases of trisomy 21 of either maternal or paternal origin of the nondisjunction error, we performed linkage analysis by quantitative fluorescence PCR genotyping nuclear trios at highly polymorphic STRs using fluorochrome-labeled primers and separating the amplimers by high-resolution capillary electrophoresis. The physical locations and primer sequences are shown in S1 Table. We determined the parental origin of the extra copy of chromosome 21 by comparing the genotype of the index case with the parental genotypes at the informative STR loci. We identified the parental alleles according to the following possible scenarios of allele segregation: index case presenting with a triallelic pattern exhibiting allele ratios of approximately 1:1:1 or a biallelic pattern exhibiting a consistent allele ratio of 2:1. The observed genotypes are shown in S2 Table.

Assessment of Heritable Epigenetic Methylation Imprints

To infer the methylation status at a given CpG island (CGI), we developed CGI-specific methylation-sensitive restriction enzyme-based PCR triplex assays (MSRE-PCR). In one tube reaction, each assay amplifies the target CGI and two other genomic regions as a control for the efficiency of the restriction enzyme digestion and to normalize the estimate ratios of the restriction enzyme-resistant 5mCpG sites. Allele-specific methylation was determined by interrogating informative SNPs neighboring the target CGIs. Information regarding the physical location of the selected target CGIs, 5mCpG-sensitive restriction enzymes, SNPs and primer sequences used in the assays is included in S1 Table. Because of the potential diagnostic value of the WRB DMR (CGI-2) for the discrimination of the parental origin of nondisjunction of the supernumerary chromosome 21, we describe its specific MSRE-PCR assay here. On untreated gDNA, the assay generates three FAM-labeled amplimers of different lengths. On HhaI treated gDNA, the combined pattern of the amplimers is used to determine one of three possible statuses of methylation: hypomethylated, hemimethylated or hypermethylated. The genomic amplimers are: (1) The WRB DMR target region (observed amplicon size: 254 bp), encompassing the recognition sites for HhaI, and the rs2299739 and rs2244352 SNPs (S1 Table). (2) The known unmethylated ESCO2 core promoter CGI (observed amplicon size ranging 248 to 250 bp), encompassing many HhaI recognition sites. The ESCO2 core promoter CGI was originally featured as a query target in the SALSA MS-MLPA ME030 probemix commercial kit (MRC Holland). This genomic amplimer provides information regarding the efficiency of the restriction enzyme reaction. (3) A region (observed amplicon size ranging 269 to 271 bp) genetically linked to the WRB DMR but lacking HhaI recognition sites. This amplimer serves as a normalization reference. Parental allele methylation statuses at the WRB DMR were assessed by interrogating neighboring SNPs using a non-fluorescent uniplex version of the MSRE-PCR. Next, we carried out SNuPE assays to genotype the methylated alleles, refractory to digestion with the restriction enzyme, using SNaPshot technology (Thermo Fisher Scientific, Waltham, MA, USA). To validate the parental allele-specific methylation statuses, we used the McrBC restriction endonuclease that acts upon DNA containing methylcytosine on one or both strands [31]. We used the same experimental design to profile the methylation statuses at 5mCpG-restriction enzyme sensitive sites located at the WRB CGI-1 and CGI-3. To evaluate the impact of the parental origin of the extra copy of chromosome 21 in the methylation of CpG sites at the RUNX1 and TMEM131 genes, we developed gene-specific methylation-sensitive restriction enzyme MSRE-PCR assays (S1 Table).

We calculated the ratio of restriction enzyme-resistant 5mCpG sites at the queried CGIs using the following equation: where, C is the value of the area under the peak (AUP) of the queried CGI amplimer from the restriction enzyme digested DNA sample; D is the AUP of the negative amplimer from the restriction enzyme digested DNA sample, and fAB is the correction factor that accounts for the lack of symmetry and imbalance observed between the AUP for the queried GCI amplimer (A) and the negative amplimer (B) obtained from the undigested DNA sample.

An Androgenetic Reference Sample of Methylation Statuses in Male Germline Derivatives

The homogenous androgenetic nature of the hydatidiform mole was determined by comparing the genotypes (S3 Table) of the two samples of the same specimen with that of a peripheral blood sample from the donor by quantitative fluorescence PCR using the AmpFlSTR Identifiler system (Thermo Fisher Scientific, Waltham, MA, USA).

Assessment of Transcriptional Allele Expression

To discriminate between the possible biallelic and the monoallelic patterns of transcriptional expression of the WRB gene, we interrogated 3´-UTR SNPs in cDNAs obtained from heterozygous blood donors and hESCs using SNaPshot technology. For comparison, we genotyped SNPs at two paternally imprinted reference genes: H19 [32] and ATP10A [33]. Primer sequences are shown in S1 Table.

High-Resolution Capillary Electrophoresis for Separation of Fluorochrome-Labeled Amplimers

Amplimers were analyzed on an automated laser fluorescent ABI PRISM 310 Genetic Analyzer (Thermo Fisher Scientific, Waltham, MA, USA). The electropherograms were generated using the dedicated GeneScan® Analysis and Genotyper® software version 3.7 packages and GeneMapper® ID version 3.2 (Applied Biosystems ® from Thermo Fisher Scientific, Waltham, MA, USA).

The Methylation Statuses at the WRB CpG Islands in Methylome Public Datasets

To examine the methylation patterns at the CGI mapped to the reference WRB locus, we performed an integrative and comparative epigenomic computational analysis by viewing the Smith Lab Public DNA methylation track hub [34] and the 111 distinct epigenomes from the Roadmap Consortium [35] at the UCSC Genome Browser [36, 37]. The Smith Lab Public DNA methylation track hub comprises a pre-loaded set of 183 analyzed human methylomes from bisulfite sequencing experiments from brain [3840], hESCs [41], induced pluripotent stem cells (iPSCs) [42] and blood cells [4345]. The rates of 5mCpG along the chromosome 21 in human oocytes [46] were displayed as a custom track in the UCSC Genome Browser. We graphically displayed the predicted promoters and CTCF and POL2 binding sites at the UCSC interface using the Ensembl Regulatory Build and Transcription Factor Binding track resource [47].

Identification of DNA Motifs Mapping at the WRB DMR

To search for cis-acting motifs that may be implicated in the differential epigenetic statuses of the WRB DMR, we queried published libraries of DNA-binding proteins with the FASTA reference sequence for the WRB DMR using online search DNA motif programs [48, 49]. We also used Tandem Repeat Finder [50] to scan the WRB DMR DNA reference sequence for the presence of ungapped (regular, fixed-length patterns) DNA repeat elements. The downstream analysis for the identified array is based on the rationale that cis elements constitute platforms on which the interactions with site-specific DNA-binding factors are built to establish and maintain epigenomic modifications [51]. To investigate whether the identified DNA motifs correspond to transcription factor binding sites, we compared the DNA motifs against a database of known motifs using the TOMTOM tool [52], which ranks each suitable match to the query and displays motif web logos. To address whether the motifs found in the WRB DMR are present in other loci, we searched the entire reference genome sequence using the FIMO tool [53] from the MEME program suite [48] available online at http://meme-suite.org/. The conservation of the array of DNA motifs was investigated by lifting over the coordinates in the reference genomes of vertebrates using the LiftOver tool from the UCSC Genome Browser [36] and submitting the lifted FASTA sequences to the motifs analysis. The coordinates of the tandem repeat motif array were then graphically displayed in the UCSC Genome Browser [36, 37] using the custom track tool.

Combinatorial Histone Modification Expression Signatures

We compared the common combinatorial histone modification expression patterns across the DMRs in 29 known maternally and two paternally imprinted genes [46] with that in the candidate imprinted WRB gene by displaying the sequence features and the activating or repressive histone marks in the UCSC genome browser using annotation and track hubs. As testable predictions of the epigenetic status, we used the following two modification modules defined by Wang and collaborators in human CD4+ T cells [54]. The 17-histone modification activation backbone module: H2A.Z, H2BK5ac, H2BK12ac, H2BK20ac, H2BK120ac, H3K4ac, H3K4me1, H3K4me2, H3K4me3, H3K9ac, H3K9me1, H3K18ac, H3K27ac, H3K36ac, H4K5ac, H4K8ac, and H4K91ac. The 5-histone modification repressive module: H3K27me3, H3K27me2, H3K9me3, H3K9me2, and H4K20me3.

Reappraisal of the Candidate Maternally Imprinting Status of the WRB Gene Using Secondary Analysis of Massive Parallel RNA Sequences

We measured allelic imbalance at likely heterozygous loci that map within a 4-Mb chromosomal region centered at the WRB gene by querying RNA sequence read archive (SRA) public data repositories. We included exon, 5´-UTR, 3´-UTR and ncRNA SNPs being 159 SNPs with MAF > 0.1 (i.e., expected global heterozygosity rate of 0.18) as testable predictions and 4 SNPs with MAF <0.05 as a control for monoallelic expression (S4 Table), mapping to within 35 genes and ncRNAs. As a reference, we used SNPs at the SNURF and H19 genes (S1 Table), which are both imprinted in every human tissue tested so far [55]. The public databases are available at NCBI Gene Expression Omnibus GEO (http://www.ncbi.nlm.nih.gov/geo/) [56], EMBL-EBI ArrayExpress (https://www.ebi.ac.uk/arrayexpress/) [57] and the Human Protein Atlas (http://www.proteinatlas.org/). We used the SRA nucleotide search expression online tool from the NCBI Browser (http://trace.ncbi.nlm.nih.gov/Traces/sra/) to download FASTQ filtered reads. The SRA Nucleotide Search Expression tool returns biological reads for an RNA-Seq spot containing a sequence substring encompassing a given SNP. Each sequence substring was limited to 29 characters in 4NA alphabet including the International Union of Pure and Applied Chemistry (IUPAC) substitution codes for the query SNPs (S1 and S4 Tables). We used IUPAC genomic reference to correct for reference allele preference during alignment [58]. We restricted the analysis to two series of experiments. The first included 1,012 human RNA-Seq sample runs (S5 Table) selected from the 4,978 unsorted accessions recently analyzed by Deelen et al. (2015) [59]. The second included 212 SRA accessions from public repositories and sorted into 15 primary tissue sources (S6 Table) by their reported Biosample and Bioproject unicity (i.e., one sample, not a mixture, from a single donor). We quantified the allele fractions only in runs that yielded a depth of at least 80 reads, with a quality of base calls > Q30, increasing the variant confidence detection to a probability of correct SNP call of 0.999 provided a read depth (coverage) of 40 for a theoretical heterozygous position [60]. We categorized the patterns of allele expression as monoallelic, biallelic or biallelic imbalance. The criteria were: monoallelic if the allele fractions were < 0.15 or > 0.85; biallelic if the allele fractions ranged from 0.35 to 0.65; and biallelic imbalance if the allele fractions ranged from > 0.15 to <0.35 or > 0.65 to <0.85. We used a chi-square test to evaluate whether the allele-specific read counts deviated from the expected proportions (50/50) [58]. To map FASTQ formatted, filtered spot RNA raw sequence data to the hg19 reference sequence we used a flexible workflow created with the web-based Galaxy tool suite (https://usegalaxy.org/) [61]. The workflow incorporates the use of the following software tools: FastQC (read quality control check) [62], FASTQ Groomer, Filter by quality (Phred score quality cut-off: 25; minimum percentage: 90), removal of sequencing artifacts, Bowtie2, BWA, BAM-to-SAM, Filter SAM or BAM, output SAM or BAM (reads with maximum 01 variant or reads with no variant), and FreeBayes (Bayesian genetic variant detector). We visualized the filtered and aligned reads using the UCSC graphical interface.

Results

The Methylation Statuses at the WRB CpG Islands

The WRB reference sequence locus contains five annotated CpG islands, CGI-1, CGI-2, CGI-3, CGI-4, and CGI-5 comprising 64, 27, 19, 17 and 18 CpG dinucleotide sites, respectively (Fig 1). CGI-1 encompasses the 5´-UTR and exon 1 of the ENST00000333781.8 reference WRB transcript variant 1 (long variant), and maps to a predicted promoter region. In public methylome databases, CGI-1 is essentially unmethylated in gametes and in somatic and embryonic cells and tissues (Fig 1). CGI-2 maps to a second predicted promoter region, located upstream of the TSS of the ENST00000380708.4 reference WRB transcript variant 2 (short variant) (Fig 1). CGI-2 is located within the differentially methylated region (DMR) reported by Court and collaborators [24] and Docherty and collaborators [25] for which WRB was classified as a candidate maternally imprinted gene (i.e., paternal-origin allele expressed). CGI-2 is differentially methylated in oocytes versus sperm [46], partially methylated in adult somatic tissues, and hypermethylated in embryonic cells and tissues (Fig 1). CGI-3 is differentially methylated in female and male gametes, essentially unmethylated in adult cells and tissues, and hypomethylated in embryonic stem cells (Fig 1). CGI-4 is hypermethylated in gametes, in all somatic cells, and in tissues (Fig 1). CGI-5 is differentially methylated in the gametes and is ubiquitously hypermethylated in somatic cells and tissues (Fig 1).

thumbnail
Fig 1. The methylation statuses at the CpG islands of the WRB gene in methylome public datasets.

Chromosomal, physical map positions and sequence features of the reference WRB gene locus showing the methylation profiles across the region, with an emphasis on the five annotated CpG islands (CGI-1 to CGI-5). The features depicted are from custom and public tracks for (from top to bottom) the exon organization of the principal (ENST00000333781.8) and alternative (ENST00000380708.4) splice WRB isoforms (variants 1 and 2 in dark blue) (UCSC Genes [36]), the species-conserved principal transcript (ENST00000333781 in pink) (APPRIS [75]), the Ensembl Regulatory Build CTCF and POL2 activity and predicted promoters [76], CpG islands, CpG sites, and regulation and methylome studies indicated in the Materials and Methods section. The CGI-2 is located in the differentially methylated region (DMR) reported by Court and collaborators [24] and Docherty and collaborators [25] (depicted in red in the custom track named Court/Docherty DMR), from which WRB was classified originally as a novel candidate, maternally imprinted gene (i.e., paternal-origin allele expressed). The custom tracks containing the sperm and oocyte DNA methylation signals correspond to the supplementary data reported by Okae et al. [46]. Screenshot generated using the UCSC Genome Browser hg19 (http://genome.ucsc.edu).

https://doi.org/10.1371/journal.pone.0154108.g001

Using target-specific MSRE-PCR assays, we experimentally replicated the above methylation statuses at the WRB CGI-2 (Fig 2A), CGI-1 (Fig 2B), and CGI-3 (Fig 2C) in genomic DNA from disomic individuals. The overall average ratio of restriction enzyme-resistant 5mCpG sites at the WRB CGI-2 was 47.4% in disomic blood cells from adult donors (S7 Table). In hESCs, the overall average ratio of 5mCpG sites was 82.1% (S7 Table). In contrast, the assay revealed a consistent unmethylated pattern of CpG sites at the WRB CGI-1 (Fig 2B) and CGI-3 (Fig 2C).

thumbnail
Fig 2. Experimental validation of the 5mCpG statuses at the WRB CGIs.

(A) A consistent hemimethylated pattern at the WRB CGI-2 revealed in a representative control disomic DNA sample (blood) using the HhaI methylation-sensitive restriction enzyme-based PCR triplex assay developed in this study. Electropherograms of the amplimers generated from either undigested genomic DNA (upper panel) or DNA digested with HhaI (lower panel) genotyped via quantitative fluorescent PCR. The positive amplimer refers to a locus in the ESCO2 gene with constitutively hypomethylated CpG dinucleotides at the target restriction enzyme sites (100% susceptible to HhaI digestion). The negative amplimer refers to a WRB region that lacks HhaI sites, and is, therefore, refractory to enzymatic digestion. The numbers in the upper boxes correspond to the amplimer lengths in base pairs while those in the lower boxes refer to the areas under the peak of the amplimer. In this representative DNA sample, the ratio of 5mCpG sites at the WRB CGI-2 was 50.6%. In contrast, the assay revealed a consistent unmethylated pattern of CpG sites at the WRB CGI-1 (B) and WRB CGI-3 (C).

https://doi.org/10.1371/journal.pone.0154108.g002

The allele-specific methylation at the WRB CGI-2 DMR was assessed experimentally across the neighboring rs2244352 SNP in genomic DNA from peripheral blood cells of disomic heterozygous subjects. The maternal alleles were consistently methylated (i.e., refractory to digestion with HhaI and susceptible to digestion with McrBC) while the paternal alleles were unmethylated (i.e., susceptible to digestion with HhaI and refractory to digestion with McrBC) (Fig 3). Thus, the underlying genetic basis for the hemimethylated profiles observed at the WRB CGI-2 DMR is unequivocally due to the inheritance of specific maternal-allele 5mCpG imprints.

thumbnail
Fig 3. Maternal-of-origin-dependent imprinted methylation marks at the WRB CGI-2 DMR.

5mCpG-sensitive restriction endonuclease sites at the WRB CGI-2 DMR are differentially methylated on the maternal allele versus the paternal allele in a manner consistent with imprinting. Electropherograms of the genotype profiles in a control nuclear family, informative for the rs2244352 (C>A) SNP neighboring the WRB CGI-2 DMR. In the child, the maternal-derived C allele is 100% resistant to HhaI digestion (fully methylated), whereas the paternal-derived A allele is 100% susceptible to HhaI digestion (i.e., unmethylated). The parental allele-specific methylation statuses were validated using the McrBC restriction endonuclease that cleaves methylated DNA and, therefore, the unmethylated paternal allele, but not the maternally methylated allele, remains undigested.

https://doi.org/10.1371/journal.pone.0154108.g003

Discrimination of the Parental Origin of Chromosome 21 Nondisjunction Using Heritable 5mCpG Imprints at the WRB CGI-2 DMR

We assessed the dependability of the known heritable epigenetic methylation [24, 25] resulting in maternal allele methylation imprints at the WRB CGI-2 DMR to ascertain the parental origin of chromosome 21 nondisjunctional events in trisomy 21 probands. For these experiments, the parental origin of the nondisjunctional events was unambiguously established by linkage analysis in nuclear families using highly polymorphic STRs. The genotypes and the segregating alleles are shown in S2 Table.

The ratios of restriction enzyme-resistant 5mCpG sites at the WRB CGI-2 DMR were significantly different between the MT21 probands (average ratio of 68.4%) and the PT21 probands (average ratio of 33.5%) (Fig 4 and S7 Table). As it follows from the inheritance of maternal-allele derived 5mCpG imprints, the WRB CGI-2 was unmethylated in DNA from a complete androgenetic mole (Figs 4 and 5A). In hESCs (Figs 4 and 5B, and S7 Table), the WRB CGI-2 was hypermethylated, indicating the loss of the imprinted pattern of 5mCpG marks at this developmental stage.

thumbnail
Fig 4. The maternal (oocyte)-derived allele methylation at the WRB CGI-2 distinguishes the parental origin of chromosome 21 nondisjunction events in Down syndrome probands.

The proportion of HhaI-resistant 5mCpG sites at the WRB CGI-2 DMR in trisomic samples with a maternal origin (MT21) of the nondisjoined chromosome 21 is consistently and statistically higher than in trisomic samples with a paternally (PT21) derived extra copy of chromosome 21. In genomic DNA from an androgenetic hydatidiform mole (ACHM), the WRB CGI-2 is unmethylated, whereas it is hypermethylated in hESCs. In control disomic samples (N21) the locus is partially methylated, consistent with a hemimethylated state. All differences were statistically significant.

https://doi.org/10.1371/journal.pone.0154108.g004

thumbnail
Fig 5. 5mCpG statuses at the WRB CGI-2 DMR in a complete androgenetic mole and a human embryonic stem cell line.

A consistent unmethylated pattern of CpG sites at the WRB CGI-2 DMR revealed in a sample of an androgenetic complete hydatidiform mole (A) contrast with the hypermethylated pattern observed in the representative HUES 3 embryonic cell line (B). Electropherograms of the amplimers (see details of the assay in Fig 2) generated from either undigested DNA or HhaI-digested DNA. The numbers in the upper boxes correspond to the amplimer lengths in base pairs while those in the lower boxes refer to the areas under the peak of the amplimer.

https://doi.org/10.1371/journal.pone.0154108.g005

Parent-of-Origin-Independent Altered Methylation Effects Due to the Extra Copy of Chromosome 21

Because the supernumerary copy of chromosome 21 differentially affects the methylation statuses at distinct loci in the genome [21, 22], we tested whether the levels of methylation at loci in either cis- or trans-configuration with regards to the WRB CGI-2 DMR change in a parent-of-origin-dependent fashion. Using the MSRE-PCR approach, we assessed the methylation statuses at CpG sites within the RUNX1 (cis) and TMEM131 (trans) genes that are differentially epigenetically perturbed in trisomic samples versus disomic samples [21, 22, 28]. Compared with disomic individuals, in trisomic subjects there is a gain of methylation at RUNX1 and a loss of methylation at TMEM131. We replicated the reported perturbation effects of the extra copy of chromosome 21 on the methylation patterns at these loci. Importantly, and in contrast to the situation at the WRB CGI-2 DMR, the epigenetic perturbations observed at the RUNX1 (Fig 6A) and TMEM131 (Fig 6B) genes were independent of the parental origin of the supernumerary chromosome 21.

thumbnail
Fig 6. Parent-of-origin-independent methylation effects of the supernumerary chromosome 21 on both syntenic and non-syntenic genes.

(A) A statistically significant gain of methylation at CpG dinucleotide sites within the syntenic RUNX1 gene is observed in trisomic individuals compared with control disomic samples (N21). The gain is independent of the maternal (MT21) or the paternal (PT21) origin of the nondisjunction event. (B) A statistically significant loss of methylation at CpG dinucleotide sites within the non-syntenic TMEM131 gene is observed in trisomic individuals compared with control disomic samples (N21). The loss is independent of the maternal (MT21) or the paternal (PT21) origin of the nondisjunction event.

https://doi.org/10.1371/journal.pone.0154108.g006

The Maternal 5mCpG Imprints at the WRB CGI-2 DMR Do Not Dictate a Paternal Monoallelic Expression

To assess whether the maternal allele-specific methylation pattern observed at the intragenic WRB CGI-2 DMR is indicative of functional gene regulation differences in the form of genomic imprinting, we qualitatively tested for allele-specific gene RNA expression by interrogating 3´-UTR SNP variants in cDNAs of RNA from blood samples and hESCs. We noted that both the maternal and paternal alleles were expressed in heterozygous samples independently of the methylation statuses observed at the WRB CGI-2 DMR (Fig 7). In hESCs, in which the WRB CGI-2 DMR is hypermethylated, the WRB transcriptional expression profile was biallelic (Fig 8).

thumbnail
Fig 7. The maternally derived, imprinted 5mCpG marks at the WRB CGI-2 DMR do not dictate a paternal monoallelic expression in blood.

Qualitative SNuPE allele-specific profiling of the WRB 3´-UTR rs1060180, rs13230 and rs60490159 SNPs in a disomic, informative nuclear family (DNA). The assay reveals a pattern consistent with biallelic transcriptional expression (cDNA) in the child, who is heterozygous for all three SNP variants.

https://doi.org/10.1371/journal.pone.0154108.g007

thumbnail
Fig 8. The maternally derived, imprinted 5mCpG marks at the WRB CGI-2 DMR do not dictate a paternal monoallelic expression in a human embryonic stem cell line.

Allele-specific transcriptional profiling of the WRB 3´-UTR rs1060180 SNP in the informative hESC HUES 1 sample (DNA; upper panel) reveals a pattern consistent with biallelic expression (cDNA; lower panel). In contrast, for the known paternally imprinted H19 and ATP10A genes, the expression profiles for the informative rs2839702 and rs2076743 SNPs, respectively, are monoallelic.

https://doi.org/10.1371/journal.pone.0154108.g008

Genomic imprinting can either completely silence one parental allele or significantly reduce its expression. We queried RNA-Seq public archives using sequence substrings for each exon 1 of the two major WRB reference transcripts, variants 1 (ENST00000333781.8) and 2 (ENST00000380708.4). We found evidence of expression (> 80 reads) of the WRB transcript variant 1 in RNA-Seq experiments in the brain, fallopian tube, liver, muscle, ovary, skin, and testis samples (S8 Table). However, using a sequence substring specific for the exon 1 of the WRB transcript variant 2, the number of reads filtered in each SRA accession from primary tissues was consistently ≤ 16, thus impairing further analysis of this short transcript variant (S8 Table) according to our stringent criteria.

Because relevant genotypes are unfortunately lacking in the entire RNA-Seq datasets analyzed (i.e., samples are not DNA/RNA exome sequencing pairs), we next aimed to provide RNA-Seq evidence for either monoallelic or biallelic expression queried with the rs1060180 and rs13230 WRB 3´-UTR SNPs. These SNPs map to both the long and short WRB transcript variants. In the 1,012 unsorted RNA-Seq experiments, we found allele fractions consistent with a biallelic pattern of expression (mean allele fraction ranging from 0.49 to 0.51) for both SNPs in the brain, epidermal keratinocytes, fetal large and small intestine, large airway epithelial cells, ovary, skin and testis samples (S9 Table). In the analysis of the RNA-Seq public databases sorted by tissue, we also observed allele fractions consistent with biallelic expression across the WRB 3´-UTR SNPs in brain, fallopian tube, thyroid, muscle, ovary, skin, and testis samples (S10 Table). Altogether, we observed no suppression or expression bias effects on either allele (i.e., the evidence against a maternally suppressed, imprinting effect) in ten biosamples (brain, fallopian tube, fetal large and small intestine, large airway epithelial cells, thyroid muscle, epidermal keratinocytes, ovary, skin, and testis). In contrast, for the known imprinted SNURF gene (tested across the rs705 SNP in the brain, fallopian tube, liver, muscle, ovary, skin, and testis) both alleles were expressed monoallelically (i.e., passed the “flip test” required for genomic imprinting) (S10 Table). Similarly, for the H19 rs217727 and H19 rs10840159 SNPs we observed monoallelic expression of both alleles in the fallopian tube, ovary, and testis (S10 Table).

Because some known DMRs are physically distant up to 2-Mb from the target imprinted gene [24, 25], and therefore function as imprinted control regions (ICRs), we extended the query to 162 SNPs (S4 Table), mapping within a 4-Mb chromosomal region centered at the WRB gene. Most SNP substrings yielded < 20 reads and, therefore, were unsuitable for further analysis according to the inclusion criteria. Importantly, in addition to WRB, we found SNPs with > 80 reads yielding allele fractions that were consistent with the biallelic expression in 15 biosamples in 10 genes: DYRK1A, KCNJ15, ETS2, PSMG1, BRWD1, HMGN1, LOC102724757, LCA5L, SH3BGR, and BACE2 (Fig 9, and S10 and S11 Tables).

thumbnail
Fig 9. The allele expression of WRB and ten neighboring genes is uncoupled from the control of the maternally inherited 5mCpG imprints at the WRB DMR.

(A) UCSC Genome Browser screenshot of custom tracks for the chromosomal and physical map positions of (top to bottom) the 11 genes that map within the 4-Mb chromosomal region centered at the WRB gene, for which evidence for biallelic transcriptional expression was unveiled in this study, and the relative locations of the 162 SNPs queried in RNA-seq public repositories for the determination of allele expression fractions. (B) Density distribution plot, by primary tissue, of the number of SRA accessions of RNA-seq experiments that yielded evidence consistent with biallelic expression. For each gene column, the intensity of the green color of each cell is proportional to the indicated numbers of informative SRA experiments. For each gene, the number of primary tissues with evidence is represented in the bottom row in blue scale.

https://doi.org/10.1371/journal.pone.0154108.g009

Common Combinatorial Patterns of Histone Acetylation and Methylation Marks at the Predicted WRB Promoter Regions

We next intended to identify epigenetic histone marks around the WRB CGI-2 DMR that conflate traceable combinatorial patterns, which possibly regulate the expression patterns of the WRB transcript variants 1 and 2. We conducted the comparative analysis of DMRs classified as germline DMRs (gDMRs) [46] and associated with patterns of imprinted methylation or expression. We found similarities in the confluence of both the 17-histone modification activation backbone module and the 5-histone modification repressive module at the CGIs mapping to the long (ENST00000333781.8) and short (ENST00000380708.4) transcript variants of the WRB gene and at the gDMRs of the BLCAP, GNAS, IGF2R, GRB10, and RB1 genes, which were most striking with the MEST gDMR (Fig 10 and S1 Fig).

thumbnail
Fig 10. Common combinatorial histone modification expression patterns around the WRB CGI-2 DMRs.

Graphical representation of the confluence of activating and repressive epigenetics histone modification marks for the known imprinted MEST gene (A) and the candidate imprinted WRB gene (B). Shown are the 17-histone modification activation backbone module and the 5-histone modification repressive module found in human CD4+ T cells [54]. Highlighted in light blue is the DMR in each gene. Composite of screenshots of the dataset viewed at the UCSC Genome Browser hg18 (http://genome.ucsc.edu).

https://doi.org/10.1371/journal.pone.0154108.g010

DNA Motifs at the WRB CGI-2 DMR

Based on a computational comparative analysis of genomic reference sequences, we observed that the nucleotide sequence encompassing the CGI-2 is conserved among primates (Fig 11A). We analyzed the region for the occurrence of arrays of cis elements (other than the CpG dinucleotide sites) that may be specific to this DMR or common to known DMRs by searching for sequence motifs as testable predictions of the differential epigenetic status. We found the DNA repeat motif [AGGYGBYSYAGGACT] (Fig 11B). In humans, the motif occurs in a cluster of seven tandem units. The motif repeat units are spaced at regular chromosomal intervals through the WRB DMR (CGI-2), encompassing 388 bp (Fig 11C). We then searched the whole genome reference sequence and were unable to find other loci containing the array that occurs at the WRB CGI-2 DMR. The [AGGYGBYSYAGGACT] motif repeat unit has sequence similarity (i.e., match overlap of thirteen nucleotides in the optimal alignment) to the putative multimer SPDEF_DBD_2 DNA-binding specificity consensus site ([gtggTCCCGGATYAT]) of the transcriptional factor SPDEF (HumanTF 1.0 DNA binding motif library) [49, 63]. The number of [AGGYGBYSYAGGACT] DNA motif units has varied since its evolutionary occurrence in marmosets (Fig 11C).

thumbnail
Fig 11. The WRB CGI-2 DMR contains a primate-conserved cluster array of a DNA motif.

(A) Molecular phylogenetic relationship of eight WRB CGI-2 DNA primate reference sequences inferred using multiple sequence alignment by CLUSTALW. (B) Sequence logo of the primate-conserved [AGGYGBYSYAGGACT] DNA repeat unit that occurs at the WRB CGI-2. (C) Physical map of the primate-conserved [AGGYGBYSYAGGACT] DNA repeat unit array.

https://doi.org/10.1371/journal.pone.0154108.g011

Discussion

In this multidisciplinary study, we used complementary experimental and computational approaches to address the challenging biological questions of whether the extra copy of chromosome 21 in Down syndrome affects the methylation patterns at distinct CpG dinucleotide sites and whether the epigenetic alterations occur dependently or independently of the parental origin of the nondisjoined chromosome 21.

Although prior studies had shown that the extra copy of chromosome 21 differentially affects the DNA methylation levels at distinct CpG dinucleotide sites [21, 22], it was unknown whether the effects are dependent (i.e., genomic imprinting) and/or independent of the parental origin of the nondisjoined chromosome 21. We showed these epigenetic effects on three genes, two located on chromosome 21 (WRB and RUNX1) and one located on chromosome 2 (TMEM131).

We established that the supernumerary chromosome 21 altered the methylation patterns at distinct CpG dinucleotide sites in the RUNX1 and TMEM131 genes in a parent-of-origin-independent manner. From a genome-wide perspective, this finding suggests that the epigenetic effect of the extra copy of chromosome 21 does not vary greatly with the parental origin of the supernumerary copy of chromosome 21; however, the extra copy does affect the methylation statuses of genes located on the same and other chromosomes.

We explored the differentially heritable epigenetic methylation imprints at the WRB CGI-2 DMR to develop a simple PCR assay based on the maternally imprinted 5mCpG marks to ascertain the parental origin of chromosome 21 nondisjunctional events in Down syndrome probands. The assay does not require bisulfite conversion and improves on the current linkage analysis approach by not requiring genomic DNA from the progenitors. This epigenetic feature will be an asset when parental samples are unavailable as in the situation of cryopreserved banked specimens. In addition to the notorious scarcity (incidence of approximately 5%) of the nondisjunctional events of paternal origin, the dependability of their identification from DNA available from the progenitors has precluded studies on phenotype-(epi)genotype correlations for Down syndrome of patrilineal descent. The WRB CGI-2 DMR-based assay will greatly facilitate the identification of Down syndrome cases of paternal origin and the establishment of representative cohorts for studies on the variation in phenotypic outcomes. The assay will also be useful in the prenatal molecular diagnosis of the parental origin, without collecting parental samples, in triploid pregnancies where only the conceptuses with two paternal sets have the potential to cause maternal complications [64].

The initial evidence suggesting a functional link between the maternal-of-origin-specific imprinted methylation at the WRB CGI-2 DMR and the WRB monoallelic expression came from the study by Docherty and collaborators [25]. By sequencing cDNA across the rs1060180 SNP, they observed alternate (i.e., maternal or paternal) monoallelic transcript expression in skeletal muscle and aorta tissues and biallelic transcript expression in the spinal cord from the same embryo. Unfortunately, that study did not conclusively establish whether the paternal allele is unmethylated. Moreover, a distinctive feature of imprinted genes is the preference for the full expression of one (and always the same) of the two parental alleles [65]. Importantly, alternate monoallelic transcript expression occurs only between different genes regulated by the same imprinting control region. For example, between the paternally expressed Peg13 non-coding RNA and the maternally expressed Kcnk9 gene in the mouse brain [66], and between the maternally expressed H19 and the paternally expressed Igf2 genes in mouse hematopoietic stem and progenitor cells [67].

We evaluated the candidate imprinting status of the WRB gene. We showed that the maternal heritable epigenetic 5mCpG imprints at a CGI-2 DMR were uncoupled to the predicted monoallelic expression of the paternal WRB allele. We replicated this observation in twelve biosamples (brain, blood, fallopian tube, fetal large and small intestine, hESCs, large airway epithelial cells, thyroid, muscle, epidermal keratinocytes, ovary, skin, and testis). These results are in agreement with recently reported findings from three genome-wide scale analyses of public RNA-Seq data [26, 59, 68]. These studies reported mean allele fractions showing either non-significant deviation (mean value = 0.504) [26, 59] or significant deviation (average value ranging from 0.553 to 0.803) [26, 68] for the WRB 3´-UTR rs1060180 and rs13230 SNPs. Furthermore, we provide evidence from 15 tissue repositories of allele rates that are consistent with the biallelic transcript expression of 10 genes that map to within 2-Mb around the WRB gene. Thus, the allele transcript expression of 11 genes in at least one primary human tissue tested is uncoupled from the control of maternally inherited imprints at the WRB DMR.

Steyaert and collaborators [26] used a SNP-guided analytical framework to identify monoallelic DNA methylation events from enrichment-based sequencing data. In the WRB gene, they functionally annotated the locus comprising the SNP rs2299739 with significant monoallelic DNA methylation. Using the rs2244352 SNP, here we demonstrated that the locus undergoes maternal-of-origin-specific differential methylation. We showed that in hESCs the WRB CGI-2 is hypermethylated, in contrast to the differential maternally inherited allele methylation profile observed in blood cells. A hypermethylated state for the WRB CGI-2 DMR occurs in several methylome studies of hESCs [35, 41, 42]. Significantly, we observed a biallelic transcriptional expression pattern for WRB, despite the hypermethylated state in hESCs.

The hemimethylated status characteristic of the known imprinted DMRs is not a sufficient epigenetic signature to determine the uniparental monoallelic expression of genes. For example, the gDMR in the GRB10 gene, which is isoform-specifically imprinted only in the brain, is hemimethylated in all tissues analyzed [55, 69]. We observed a comparable confluence of both the 17-histone modification activation backbone module and the 5-histone modification repressive module between six maternally imprinted genes (MEST, BLCAP, GNAS, IGF2R, GRB10, and RB1) and the WRB gene. Therefore, there is still the possibility of the WRB gene being an isoform- and tissue-specific imprinted gene. Due to the unavailability of isoform-specific SNPs and the small levels of expression of the WRB transcript variant 2 in RNA-Seq experiments, unfortunately, we could not test that hypothesis with our current approach.

In a recent methylome study in human oocytes, Okae and collaborators [46] used the criteria of DNA methylation levels at the DMRs of known imprinted genes to classify the WRB CGI-2 as a secondary DMR (i.e., rather than a gDMR). At the WRB CGI-2, they observed an average methylation of 17.5% in oocytes (hg19 coordinates chr21:40757510–40758276; 11.5% in oocytes pool 1 PCR and 49.9% in oocyte pool 2) and of 0.45% in sperm, but with 35–65% methylation levels in blood cells. Importantly, a genome-wide scan of the methylomes of oocytes and blastocysts from that study revealed that the methylation profiles are highly comparable, with the methylation levels of maternal germline DMRs in blastocysts being on average half the levels found in oocytes (data not shown). Because the reported level of methylation at the WRB CGI-2 in blastocyst was 37.72% [46], one would expect a methylation level of approximately 75% in oocytes. In fact, the average methylation level at a core region with hg19 coordinates chr21: 40757603–40757721 in oocyte pool 2 is 70.3% (reanalysis of data from S1 Table of ref. [46]). Therefore, we consider the WRB CGI-2 to be a maternal gDMR.

We demonstrated the maintenance of the DNA sequence context of the human WRB CGI-2 DMR in primates. Within the WRB CGI-2 DMR, we identified a cis DNA motif array. In primates, the array consists of 2 to 8 consensus repeat units in primates that bear sequence similarity to the DNA-binding specificity consensus site of the transcription factor SPDEF. We speculate that the cis DNA motif array is a distinctive feature in the establishment and/or maintenance of the hemimethylated state of this particular DMR, rather than being involved in the differential expression of WRB transcript variants or alleles.

WRB is essentially a housekeeping gene. The human WRB protein corresponds to the Get1 protein in yeast. The WRB/Get1 proteins form a conserved family (IPR028945) in eukaryotes. These proteins function as transmembrane receptors for ASNA1/TRC40 (Get3 in yeast)-mediated insertion of tail-anchored (TA) proteins into the endoplasmic reticulum membrane (GO:0071816) [70]. RNA-Seq profiling revealed that WRB transcription occurs in all individual tissue categories investigated, although higher levels (i.e., FPKM/TPM) are displayed in the brain, testis, ovary and kidney [71]. Most tissues displayed moderate to strong nuclear and cytoplasmic positivity by WRB-specific antibody staining [72]. RNAi-mediated knockdown of WRB expression increases the rate of homologous recombination DNA double-strand break repair [73]. Importantly, WRB transcription levels are compensated in trisomic T21 lymphoblastoid cells [74], and the WRB gene lies outside a chromosomal domain dysregulated by the presence of the supernumerary chromosome 21 [20]. Therefore, the biological significance of the WRB gene perhaps being imprinted in an isoform, tissue and/or developmental-stage-specific manner is rather intriguing.

Supporting Information

S1 Fig. Comparison of the common combinatorial histone modification expression patterns around the WRB CGI-2 DMRs and known imprinted genes.

Graphical representation of the confluence of activating and repressive epigenetics histone modification marks for the known imprinted genes (A) BLCAP, (B) GNAS, (C) IGF2R, (D) GRB10 and (E) RB1. Shown are the 17-histone modification activation backbone module, and the 5-histone modification repressive module found in human CD4+ T cells [54]. Highlighted in light blue is the DMR in each gene. Composite of screenshots of the dataset viewed at the UCSC Genome Browser hg18 (http://genome.ucsc.edu).

https://doi.org/10.1371/journal.pone.0154108.s001

(TIF)

S1 Table. List of target loci, primer sequences, and restriction enzymes interrogated for parent-of-origin- dependent and -independent effects on DNA methylation.

https://doi.org/10.1371/journal.pone.0154108.s002

(XLSX)

S2 Table. STR genotypes and allele segregation analysis used in the parental origin determination of the extra chromosome 21 in trisomy 21 nuclear families.

https://doi.org/10.1371/journal.pone.0154108.s003

(XLSX)

S3 Table. STR genotypes and allele segregation analysis used in the genetic characterization of the androgenetic complete hydatidiform mole (ACHM) and the peripheral blood of the donor.

https://doi.org/10.1371/journal.pone.0154108.s004

(XLSX)

S4 Table. Consolidated information regarding the 162 SNPs that map within the 4-Mb chromosomal region centered at the WRB gene queried in public RNA-Seq repositories of fifteen primary tissues to determine the status of allele expression.

https://doi.org/10.1371/journal.pone.0154108.s005

(XLS)

S5 Table. List of the 1,012 RNA-Seq public datasets queried to investigate locus and allele-specific expression.

The list is a manually curated compilation of tables downloaded from the NCBI Sequence Read Archive (SRA).

https://doi.org/10.1371/journal.pone.0154108.s006

(XLSX)

S6 Table. List of the RNA-Seq public data experiments configuring the Atlas of primary human tissues queried to replicate the imprinting-dependent and independent patterns of allele-specific expression of target genes.

The list is a manually curated compilation of tables downloaded from the NCBI Sequence Read Archive (SRA), Biosample and Bioproject browsers.

https://doi.org/10.1371/journal.pone.0154108.s007

(XLS)

S7 Table. Ratios of the restriction enzyme-resistant 5mCpG sites in the WRB CGI-2 estimated in control disomic and trisomic study subjects and in human embryonic stem cell lines.

https://doi.org/10.1371/journal.pone.0154108.s008

(XLSX)

S8 Table. RNA-Seq reads filtered with sequence substrings specific for each exon 1 of either WRB transcript variants 1 (ENST00000333781.8, long) or 2 (ENST00000380708.4, short).

https://doi.org/10.1371/journal.pone.0154108.s009

(XLS)

S9 Table. Allele expression fractions across the WRB 3´-UTR rs1060180 and rs13230 SNPs found using unsorted 1,012 RNA-Seq public datasets.

https://doi.org/10.1371/journal.pone.0154108.s010

(XLSX)

S10 Table. Allele expression fractions across the 162 SNPs mapping within the 4-Mb chromosomal region centered at the candidate imprinted WRB gene and across SNPs in the SNURF and H19 known imprinted genes.

Shown in the different worksheets are the number of reads for each SNP filtered in the RNA-Seq public datasets, sorted by informative tissue. Worksheet labels correspond to the series “tissue gene SNP” (i.e., Adrenal BACE2 rs11701157), grouped alphabetically, and highlighted in different colors by tissue.

https://doi.org/10.1371/journal.pone.0154108.s011

(XLS)

S11 Table. Summary of the RNA-seq evidence of biallelic expression of eleven genes mapping to a 4-Mb chromosomal region centered at the WRB gene in fifteen primary human tissues (Fig 9 in the main text is a graphical representation of these data).

https://doi.org/10.1371/journal.pone.0154108.s012

(XLS)

Acknowledgments

We would like thank Dr. Francisco Augusto Colucci Coelho (Centro de Infertilidade e Medicina Fetal do Norte Fluminense, Hospital Escola Álvaro Alvim, Campos dos Goytacazes, RJ, Brazil) and Dr. Jorge de Rezende Filho (Santa Casa de Misericórdia do Rio de Janeiro, RJ, Brazil) for the biological sample from the androgenetic hydatidiform mole. We would also like to thank Dr. Christine L Mummery (Leiden University Medical Center, Leiden, and The Netherlands) for the gift of the frozen pellets from human embryonic stem cell lines.

Web Resources

The URLs for the data presented herein are as follows:

UCSC Genome Bioinformatics tools, http://genome.ucsc.edu/

Combinatorial patterns of histone acetylation and methylation marks, http://dir.nhlbi.nih.gov/papers/lmi/epigenomes/hgtcellacetylation.aspx http://dir.nhlbi.nih.gov/papers/lmi/epigenomes/hgtcell.aspx

dbSNP, http://www.ncbi.nlm.nih.gov/snp/

Ensembl, http://www.ensembl.org/index.html

NCBI (National Center for Biotechnology), http://www.ncbi.nlm.nih.gov/guide/

REBASE, http://rebase.neb.com/cgi-bin/mslist

The MEME Suite (http://meme.nbcr.net/meme/)

Author Contributions

Conceived and designed the experiments: AFAS FMB EM-A. Performed the experiments: AFAS FMB RSFJ PTMR DTM EM-A. Analyzed the data: AFAS FMB RSFJ PTMR DTM EM-A. Contributed reagents/materials/analysis tools: ECP JMBP BLZ CBSR MGF SMCSL AFLR EM-A. Wrote the paper: AFAS FMB EM-A.

References

  1. 1. Megarbane A, Ravel A, Mircher C, Sturtz F, Grattau Y, Rethore MO, et al. The 50th anniversary of the discovery of trisomy 21: the past, present, and future of research and treatment of Down syndrome. Genet Med. 2009;11(9):611–6. pmid:19636252.
  2. 2. Antonarakis SE. Parental origin of the extra chromosome in trisomy 21 as indicated by analysis of DNA polymorphisms. Down Syndrome Collaborative Group. N Engl J Med. 1991;324(13):872–6. pmid:1825697.
  3. 3. Oliver TR, Feingold E, Yu K, Cheung V, Tinker S, Yadav-Shah M, et al. New insights into human nondisjunction of chromosome 21 in oocytes. PLoS Genet. 2008;4(3):e1000033. pmid:18369452; PubMed Central PMCID: PMCPMC2265487.
  4. 4. Gaulden ME. Maternal age effect: the enigma of Down syndrome and other trisomic conditions. Mutat Res. 1992;296(1–2):69–88. pmid:1279409.
  5. 5. Erickson JD, Bjerkedal TO. Down syndrome associated with father's age in Norway. J Med Genet. 1981;18(1):22–8. pmid:6454784; PubMed Central PMCID: PMCPMC1048652.
  6. 6. Hatch M, Kline J, Levin B, Hutzler M, Warburton D. Paternal age and trisomy among spontaneous abortions. Hum Genet. 1990;85(3):355–61. pmid:2394449.
  7. 7. Kazaura MR, Lie RT. Down's syndrome and paternal age in Norway. Paediatr Perinat Epidemiol. 2002;16(4):314–9. pmid:12445147.
  8. 8. De Souza E, Morris JK, Group EW. Case-control analysis of paternal age and trisomic anomalies. Arch Dis Child. 2010;95(11):893–7. pmid:20584846.
  9. 9. Zigman WB. Atypical aging in Down syndrome. Dev Disabil Res Rev. 2013;18(1):51–67. pmid:23949829.
  10. 10. Stoll C, Alembik Y, Dott B, Feingold J. No evidence for genomic imprinting in liver-born Down syndrome patients. Acta Genet Med Gemellol (Roma). 1996;45(1–2):265–71.
  11. 11. Muranjan M, Chaudhari T, Vundinti BR. Phenotypic heterogeneity and parental origin of extra chromosome 21 in Down syndrome. Indian Pediatr. 2010;47(5):429–32. pmid:19671944.
  12. 12. Henderson DJ, Sherman LS, Loughna SC, Bennett PR, Moore GE. Early embryonic failure associated with uniparental disomy for human chromosome 21. Hum Mol Genet. 1994;3(8):1373–6. pmid:7987317.
  13. 13. Creau-Goldberg N, Gegonne A, Delabar J, Cochet C, Cabanis MO, Stehelin D, et al. Maternal origin of a de novo balanced t(21q21q) identified by ets-2 polymorphism. Hum Genet. 1987;76(4):396–8. pmid:2886422.
  14. 14. Rogan PK, Sabol DW, Punnett HH. Maternal uniparental disomy of chromosome 21 in a normal child. Am J Med Genet. 1999;83(1):69–71. pmid:10076888.
  15. 15. Mansuet-Lupo A, Henke J, Henke L, Blank C, Ernsting A, Kozlowski P, et al. A paternity case with three genetic incompatibilities between father and child due to maternal uniparental disomy 21 and a mutation at the Y chromosome. Forensic Sci Int Genet. 2009;3(2):141–3. pmid:19215885.
  16. 16. Blouin JL, Avramopoulos D, Pangalos C, Antonarakis SE. Normal phenotype with paternal uniparental isodisomy for chromosome 21. Am J Hum Genet. 1993;53(5):1074–8. pmid:8213833; PubMed Central PMCID: PMCPMC1682298.
  17. 17. Robinson WP, Bernasconi F, Basaran S, Yuksel-Apak M, Neri G, Serville F, et al. A somatic origin of homologous Robertsonian translocations and isochromosomes. Am J Hum Genet. 1994;54(2):290–302. pmid:8304346; PubMed Central PMCID: PMCPMC1918173.
  18. 18. Liehr T. Uniparental Disomy (UPD) in Clinical Genetics. Berlin: Springer Heidelberg 2014. 201 p.
  19. 19. Conti A, Fabbrini F, D'Agostino P, Negri R, Greco D, Genesio R, et al. Altered expression of mitochondrial and extracellular matrix genes in the heart of human fetuses with chromosome 21 trisomy. BMC Genomics. 2007;8:268. pmid:17683628; PubMed Central PMCID: PMCPMC1964766.
  20. 20. Letourneau A, Santoni FA, Bonilla X, Sailani MR, Gonzalez D, Kind J, et al. Domains of genome-wide gene expression dysregulation in Down's syndrome. Nature. 2014;508(7496):345–50. pmid:24740065.
  21. 21. Kerkel K, Schupf N, Hatta K, Pang D, Salas M, Kratz A, et al. Altered DNA methylation in leukocytes with trisomy 21. PLoS Genet. 2010;6(11):e1001212. pmid:21124956; PubMed Central PMCID: PMCPMC2987931.
  22. 22. Jones MJ, Farre P, McEwen LM, Macisaac JL, Watt K, Neumann SM, et al. Distinct DNA methylation patterns of cognitive impairment and trisomy 21 in Down syndrome. BMC Med Genomics. 2013;6:58. pmid:24373378; PubMed Central PMCID: PMCPMC3879645.
  23. 23. Oliver TR, Bhise A, Feingold E, Tinker S, Masse N, Sherman SL. Investigation of factors associated with paternal nondisjunction of chromosome 21. Am J Med Genet A. 2009;149A(8):1685–90. pmid:19606484; PubMed Central PMCID: PMCPMC4111419.
  24. 24. Court F, Tayama C, Romanelli V, Martin-Trujillo A, Iglesias-Platas I, Okamura K, et al. Genome-wide parent-of-origin DNA methylation analysis reveals the intricacies of human imprinting and suggests a germline methylation-independent mechanism of establishment. Genome Res. 2014;24(4):554–69. pmid:24402520; PubMed Central PMCID: PMCPMC3975056.
  25. 25. Docherty LE, Rezwan FI, Poole RL, Jagoe H, Lake H, Lockett GA, et al. Genome-wide DNA methylation analysis of patients with imprinting disorders identifies differentially methylated regions associated with novel candidate imprinted genes. J Med Genet. 2014;51(4):229–38. pmid:24501229; PubMed Central PMCID: PMCPMC3963529.
  26. 26. Steyaert S, Van Criekinge W, De Paepe A, Denil S, Mensaert K, Vandepitte K, et al. SNP-guided identification of monoallelic DNA-methylation events from enrichment-based sequencing data. Nucleic Acids Res. 2014;42(20):e157. pmid:25237057; PubMed Central PMCID: PMCPMC4227762.
  27. 27. Wang X, Clark AG. Using next-generation RNA sequencing to identify imprinted genes. Heredity (Edinb). 2014;113(2):156–66. pmid:24619182; PubMed Central PMCID: PMCPMC4105452.
  28. 28. Bacalini MG, Gentilini D, Boattini A, Giampieri E, Pirazzini C, Giuliani C, et al. Identification of a DNA methylation signature in blood cells from persons with Down Syndrome. Aging (Albany NY). 2015;7(2):82–96. pmid:25701644; PubMed Central PMCID: PMCPMC4359691.
  29. 29. Sambrook J, Russell DW. Molecular cloning: a laboratory manual. 3rd ed. Cold Spring Harbor: Cold Spring Harbor Laboratory Press; 2001. 999 p p.
  30. 30. Machado FB, Machado FB, Faria MA, Lovatel VL, Alves da Silva AF, Radic CP, et al. 5meCpG epigenetic marks neighboring a primate-conserved core promoter short tandem repeat indicate X-chromosome inactivation. PLoS One. 2014;9(7):e103714. pmid:25078280; PubMed Central PMCID: PMCPMC4117532.
  31. 31. Sutherland E, Coe L, Raleigh EA. McrBC: a multisubunit GTP-dependent restriction endonuclease. J Mol Biol. 1992;225(2):327–48. pmid:1317461.
  32. 32. Rachmilewitz J, Goshen R, Ariel I, Schneider T, de Groot N, Hochberg A. Parental imprinting of the human H19 gene. FEBS Lett. 1992;309(1):25–8. pmid:1380925.
  33. 33. Meguro M, Kashiwagi A, Mitsuya K, Nakao M, Kondo I, Saitoh S, et al. A novel maternally expressed gene, ATP10C, encodes a putative aminophospholipid translocase associated with Angelman syndrome. Nat Genet. 2001;28(1):19–20. pmid:11326269.
  34. 34. Song Q, Decato B, Hong EE, Zhou M, Fang F, Qu J, et al. A reference methylome database and analysis pipeline to facilitate integrative and comparative epigenomics. PLoS One. 2013;8(12):e81148. pmid:24324667; PubMed Central PMCID: PMCPMC3855694.
  35. 35. Roadmap Epigenomics C, Kundaje A, Meuleman W, Ernst J, Bilenky M, Yen A, et al. Integrative analysis of 111 reference human epigenomes. Nature. 2015;518(7539):317–30. pmid:25693563; PubMed Central PMCID: PMCPMC4530010.
  36. 36. Kent WJ, Sugnet CW, Furey TS, Roskin KM, Pringle TH, Zahler AM, et al. The human genome browser at UCSC. Genome Res. 2002;12(6):996–1006. Article published online before print in May 2002. pmid:12045153; PubMed Central PMCID: PMCPMC186604.
  37. 37. Raney BJ, Dreszer TR, Barber GP, Clawson H, Fujita PA, Wang T, et al. Track data hubs enable visualization of user-defined genome-wide annotations on the UCSC Genome Browser. Bioinformatics. 2014;30(7):1003–5. pmid:24227676; PubMed Central PMCID: PMCPMC3967101.
  38. 38. Zeng J, Konopka G, Hunt BG, Preuss TM, Geschwind D, Yi SV. Divergent whole-genome methylation maps of human and chimpanzee brains reveal epigenetic basis of human regulatory evolution. Am J Hum Genet. 2012;91(3):455–65. pmid:22922032; PubMed Central PMCID: PMCPMC3511995.
  39. 39. Lister R, Mukamel EA, Nery JR, Urich M, Puddifoot CA, Johnson ND, et al. Global epigenomic reconfiguration during mammalian brain development. Science. 2013;341(6146):1237905. pmid:23828890; PubMed Central PMCID: PMCPMC3785061.
  40. 40. Ziller MJ, Gu H, Muller F, Donaghey J, Tsai LT, Kohlbacher O, et al. Charting a dynamic DNA methylation landscape of the human genome. Nature. 2013;500(7463):477–81. pmid:23925113; PubMed Central PMCID: PMCPMC3821869.
  41. 41. Xie W, Schultz MD, Lister R, Hou Z, Rajagopal N, Ray P, et al. Epigenomic analysis of multilineage differentiation of human embryonic stem cells. Cell. 2013;153(5):1134–48. pmid:23664764; PubMed Central PMCID: PMCPMC3786220.
  42. 42. Lister R, Pelizzola M, Kida YS, Hawkins RD, Nery JR, Hon G, et al. Hotspots of aberrant epigenomic reprogramming in human induced pluripotent stem cells. Nature. 2011;471(7336):68–73. pmid:21289626; PubMed Central PMCID: PMCPMC3100360.
  43. 43. Li Y, Zhu J, Tian G, Li N, Li Q, Ye M, et al. The DNA methylome of human peripheral blood mononuclear cells. PLoS Biol. 2010;8(11):e1000533. pmid:21085693; PubMed Central PMCID: PMCPMC2976721.
  44. 44. Hodges E, Molaro A, Dos Santos CO, Thekkat P, Song Q, Uren PJ, et al. Directional DNA methylation changes and complex intermediate states accompany lineage specificity in the adult hematopoietic compartment. Mol Cell. 2011;44(1):17–28. pmid:21924933; PubMed Central PMCID: PMCPMC3412369.
  45. 45. Heyn H, Li N, Ferreira HJ, Moran S, Pisano DG, Gomez A, et al. Distinct DNA methylomes of newborns and centenarians. Proc Natl Acad Sci U S A. 2012;109(26):10522–7. pmid:22689993; PubMed Central PMCID: PMCPMC3387108.
  46. 46. Okae H, Chiba H, Hiura H, Hamada H, Sato A, Utsunomiya T, et al. Genome-wide analysis of DNA methylation dynamics during early human development. PLoS Genet. 2014;10(12):e1004868. pmid:25501653; PubMed Central PMCID: PMCPMC4263407.
  47. 47. Flicek P, Amode MR, Barrell D, Beal K, Billis K, Brent S, et al. Ensembl 2014. Nucleic Acids Res. 2014;42(Database issue):D749–55. pmid:24316576; PubMed Central PMCID: PMCPMC3964975.
  48. 48. Bailey TL, Boden M, Buske FA, Frith M, Grant CE, Clementi L, et al. MEME SUITE: tools for motif discovery and searching. Nucleic Acids Res. 2009;37(Web Server issue):W202–8. pmid:19458158; PubMed Central PMCID: PMCPMC2703892.
  49. 49. Sebastian A, Contreras-Moreira B. footprintDB: a database of transcription factors with annotated cis elements and binding interfaces. Bioinformatics. 2014;30(2):258–65. pmid:24234003.
  50. 50. Benson G. Tandem repeats finder: a program to analyze DNA sequences. Nucleic Acids Res. 1999;27(2):573–80. pmid:9862982; PubMed Central PMCID: PMCPMC148217.
  51. 51. Whitaker JW, Chen Z, Wang W. Predicting the human epigenome from DNA motifs. Nat Methods. 2015;12(3):265–72, 7 p following 72. pmid:25240437; PubMed Central PMCID: PMCPMC4344378.
  52. 52. Gupta S, Stamatoyannopoulos JA, Bailey TL, Noble WS. Quantifying similarity between motifs. Genome Biol. 2007;8(2):R24. pmid:17324271; PubMed Central PMCID: PMCPMC1852410.
  53. 53. Grant CE, Bailey TL, Noble WS. FIMO: scanning for occurrences of a given motif. Bioinformatics. 2011;27(7):1017–8. pmid:21330290; PubMed Central PMCID: PMCPMC3065696.
  54. 54. Wang Z, Zang C, Rosenfeld JA, Schones DE, Barski A, Cuddapah S, et al. Combinatorial patterns of histone acetylations and methylations in the human genome. Nat Genet. 2008;40(7):897–903. pmid:18552846; PubMed Central PMCID: PMCPMC2769248.
  55. 55. Baran Y, Subramaniam M, Biton A, Tukiainen T, Tsang EK, Rivas MA, et al. The landscape of genomic imprinting across diverse adult human tissues. Genome Res. 2015;25(7):927–36. pmid:25953952; PubMed Central PMCID: PMCPMC4484390.
  56. 56. Barrett T, Edgar R. Gene expression omnibus: microarray data storage, submission, retrieval, and analysis. Methods Enzymol. 2006;411:352–69. pmid:16939800; PubMed Central PMCID: PMCPMC1619900.
  57. 57. Parkinson H, Sarkans U, Kolesnikov N, Abeygunawardena N, Burdett T, Dylag M, et al. ArrayExpress update—an archive of microarray and high-throughput sequencing-based functional genomics experiments. Nucleic Acids Res. 2011;39(Database issue):D1002–4. pmid:21071405; PubMed Central PMCID: PMCPMC3013660.
  58. 58. Smith RM, Webb A, Papp AC, Newman LC, Handelman SK, Suhy A, et al. Whole transcriptome RNA-Seq allelic expression in human brain. BMC Genomics. 2013;14:571. pmid:23968248; PubMed Central PMCID: PMCPMC3765493.
  59. 59. Deelen P, Zhernakova DV, de Haan M, van der Sijde M, Bonder MJ, Karjalainen J, et al. Calling genotypes from public RNA-sequencing data enables identification of genetic variants that affect gene-expression levels. Genome Med. 2015;7(1):30. pmid:25954321; PubMed Central PMCID: PMCPMC4423486.
  60. 60. Bentley DR, Balasubramanian S, Swerdlow HP, Smith GP, Milton J, Brown CG, et al. Accurate whole human genome sequencing using reversible terminator chemistry. Nature. 2008;456(7218):53–9. pmid:18987734; PubMed Central PMCID: PMCPMC2581791.
  61. 61. Blankenberg D, Gordon A, Von Kuster G, Coraor N, Taylor J, Nekrutenko A, et al. Manipulation of FASTQ data with Galaxy. Bioinformatics. 2010;26(14):1783–5. pmid:20562416; PubMed Central PMCID: PMCPMC2894519.
  62. 62. Krueger F, Kreck B, Franke A, Andrews SR. DNA methylome analysis using short bisulfite sequencing data. Nat Methods. 2012;9(2):145–51. pmid:22290186.
  63. 63. Jolma A, Yan J, Whitington T, Toivonen J, Nitta KR, Rastas P, et al. DNA-binding specificities of human transcription factors. Cell. 2013;152(1–2):327–39. pmid:23332764.
  64. 64. Joergensen MW, Rasmussen AA, Niemann I, Hindkjaer J, Agerholm I, Bolund L, et al. Methylation-specific multiplex ligation-dependent probe amplification: utility for prenatal diagnosis of parental origin in human triploidy. Prenat Diagn. 2013;33(12):1131–6. pmid:23881788.
  65. 65. Proudhon C, Bourc'his D. Identification and resolution of artifacts in the interpretation of imprinted gene expression. Brief Funct Genomics. 2010;9(5–6):374–84. pmid:20829207; PubMed Central PMCID: PMCPMC3080772.
  66. 66. Court F, Camprubi C, Garcia CV, Guillaumet-Adkins A, Sparago A, Seruggia D, et al. The PEG13-DMR and brain-specific enhancers dictate imprinted expression within the 8q24 intellectual disability risk locus. Epigenetics Chromatin. 2014;7(1):5. pmid:24667089; PubMed Central PMCID: PMCPMC3986935.
  67. 67. Venkatraman A, He XC, Thorvaldsen JL, Sugimura R, Perry JM, Tao F, et al. Maternal imprinting at the H19-Igf2 locus maintains adult haematopoietic stem cell quiescence. Nature. 2013;500(7462):345–9. pmid:23863936; PubMed Central PMCID: PMCPMC3896866.
  68. 68. Wood DL, Nones K, Steptoe A, Christ A, Harliwong I, Newell F, et al. Recommendations for Accurate Resolution of Gene and Isoform Allele-Specific Expression in RNA-Seq Data. PLoS One. 2015;10(5):e0126911. pmid:25965996; PubMed Central PMCID: PMCPMC4428808.
  69. 69. Blagitko N, Mergenthaler S, Schulz U, Wollmann HA, Craigen W, Eggermann T, et al. Human GRB10 is imprinted and expressed from the paternal and maternal allele in a highly tissue- and isoform-specific fashion. Hum Mol Genet. 2000;9(11):1587–95. pmid:10861285.
  70. 70. Vilardi F, Lorenz H, Dobberstein B. WRB is the receptor for TRC40/Asna1-mediated insertion of tail-anchored proteins into the ER membrane. J Cell Sci. 2011;124(Pt 8):1301–7. pmid:21444755; PubMed Central PMCID: PMCPMC3115773.
  71. 71. Petryszak R, Burdett T, Fiorelli B, Fonseca NA, Gonzalez-Porta M, Hastings E, et al. Expression Atlas update—a database of gene and transcript expression from microarray- and sequencing-based functional genomics experiments. Nucleic Acids Res. 2014;42(Database issue):D926–32. pmid:24304889; PubMed Central PMCID: PMCPMC3964963.
  72. 72. Uhlen M, Fagerberg L, Hallstrom BM, Lindskog C, Oksvold P, Mardinoglu A, et al. Proteomics. Tissue-based map of the human proteome. Science. 2015;347(6220):1260419. pmid:25613900.
  73. 73. Slabicki M, Theis M, Krastev DB, Samsonov S, Mundwiller E, Junqueira M, et al. A genome-scale DNA repair RNAi screen identifies SPG48 as a novel gene associated with hereditary spastic paraplegia. PLoS Biol. 2010;8(6):e1000408. pmid:20613862; PubMed Central PMCID: PMCPMC2893954.
  74. 74. Ait Yahya-Graison E, Aubert J, Dauphinot L, Rivals I, Prieur M, Golfier G, et al. Classification of human chromosome 21 gene-expression variations in Down syndrome: impact on disease phenotypes. Am J Hum Genet. 2007;81(3):475–91. pmid:17701894; PubMed Central PMCID: PMCPMC1950826.
  75. 75. Rodriguez JM, Maietta P, Ezkurdia I, Pietrelli A, Wesselink JJ, Lopez G, et al. APPRIS: annotation of principal and alternative splice isoforms. Nucleic Acids Res. 2013;41(Database issue):D110–7. pmid:23161672; PubMed Central PMCID: PMCPMC3531113.
  76. 76. Cunningham F, Amode MR, Barrell D, Beal K, Billis K, Brent S, et al. Ensembl 2015. Nucleic Acids Res. 2015;43(Database issue):D662–9. pmid:25352552; PubMed Central PMCID: PMCPMC4383879.