Ana C. Marques’s research while affiliated with University of Lausanne and other places

What is this page?


This page lists works of an author who doesn't have a ResearchGate profile or hasn't added the works to their profile yet. It is automatically generated from public (personal) data to further our legitimate goal of comprehensive and accurate scientific recordkeeping. If you are this author and want this page removed, please let us know.

Publications (108)


Correction: The long non-coding RNA Cerox1 is a post transcriptional regulator of mitochondrial complex I catalytic activity
  • Article

February 2025

·

11 Reads

eLife

Tamara M Sirey

·

·

·

[...]

·

Chris P Ponting

Fig. 5 Left: degradation rates estimated from a single sample plotted against degradation rates published in [27] (obtained using slam-seq). The red line is obtained through weighted linear regression. The weights are set as 1 − r l as indicated by the transparency of the dots. The (weighted) correlation of 55% indicates that the estimated rates are meaningful. Only genes with a mean exon TPM above 100 are taken into account. Right: Correlation between degradation rates obtained by [27] and the ones obtained our single-sample method as a function of expression level. Each line represents a biological replicate. The red dot corresponds to the data shown on the left. As expected, the correlation is higher for highly expressed genes, as the intro to exon ratios can be more reliably estimated. In this experiment, replicate 1 correlates better than the two others, indicating that it is probably of better quality
Fig. 6 Comparison of our method (SSRE) with the INSPEcT "first guess" on th same data. Top row: direct comparison (in log space) of rates obtained with our method and with the INSPEcT package on a single sample. Synthesis and processing rates are well correlated but not the degradation rate (Spearman correlation shown). The red bar indicates the diagonal. Bottom left: bars indicate the correlation of degradation rates with previously published data [27], as in Fig. 5. The INSPEcT method provides degradation rates with good correlation only for one of the three replicates (repl. 2), whereas it is the case for all three replicates using our method. The big dots indicate the slope of the regression line in log-log space (as in Fig. 5, left). Slopes obtained from SSRE estimates are closer to one, which correspond to the ideal case of a linear relationship between the (non-log) rates. Bottom center and right: Rates obtained with INSPEcT also reproduce the positive correlation between synthesis and processing rates, but they produce a negative correlation between synthesis and degradation rates, unlike our method (see Fig. 4, left) and previously published results. [27]
Estimating RNA dynamics using one time point for one sample in a single-pulse metabolic labeling experiment
  • Article
  • Full-text available

April 2022

·

72 Reads

·

2 Citations

BMC Bioinformatics

Background Over the past decade, experimental procedures such as metabolic labeling for determining RNA turnover rates at the transcriptome-wide scale have been widely adopted and are now turning to single cell measurements. Several computational methods to estimate RNA synthesis, processing and degradation rates from such experiments have been suggested, but they all require several RNA sequencing samples. Here we present a method that can estimate those three rates from a single sample. Methods Our method relies on the analytical solution to the Zeisel model of RNA dynamics. It was validated on metabolic labeling experiments performed on mouse embryonic stem cells. Resulting degradation rates were compared both to previously published rates on the same system and to a state-of-the-art method applied to the same data. Results Our method is computationally efficient and outputs rates that correlate well with previously published data sets. Using it on a single sample, we were able to reproduce the observation that dynamic biological processes tend to involve genes with higher metabolic rates, while stable processes involve genes with lower rates. This supports the hypothesis that cells control not only the mRNA steady-state abundance, but also its responsiveness, i.e., how fast steady state is reached. Moreover, degradation rates obtained with our method compare favourably with the other tested method. Conclusions In addition to saving experimental work and computational time, estimating rates for a single sample has several advantages. It does not require an error-prone normalization across samples and enables the use of replicates to estimate uncertainty and assess sample quality. Finally the method and theoretical results described here are general enough to be useful in other contexts such as nucleotide conversion methods and single cell metabolic labeling experiments.

Download

Splicing of elncRNAs has evolved under purifying selection
(A) Distribution of pairwise nucleotide substitution rate between human and mouse for 1,000 randomly subsampled sets of local ancestral repeats (ARs) with matching GC-content and size as splicing related motifs (grey). The observed rate observed at splice sites (dSS, solid line arrow), ESE (dESE, dash line arrow), and U1 (dU1, dotted line arrow) sites within multi-exonic elncRNAs. Distribution of nucleotide substitution rate relative to that of randomly subsampled local ARs (dAR) for (B) splice sites, (C) ESEs, and (D) U1s within multi-exonic elncRNA (red), plncRNA (blue) and protein-coding gene (dark green). Differences between groups were tested using a two-tailed Mann-Whitney U test. *** p < 0.001. (E) Distribution of derived allele frequency (DAF) for single nucleotide variants at splicing motifs (splice sites, ESEs and U1s) within multi-exonic elncRNAs (red), plncRNAs (blue), protein-coding genes (dark green), and ARs (black). The insert illustrates variants with low derived allele frequency (DAF<0.1). Differences between groups were tested using a two-tailed Fisher’s exact test. * p < 0.05; *** p < 0.001; NS p > 0.05.
Disrupted elncRNA splicing impacts cis-gene regulation
(A) Examples of variants that can disrupt elncRNA splicing. In contrast to individuals that carry the reference genome allele at a canonical splice site, those with alternative alleles have at least one variant that disrupt the GU dinucleotide at splice donor site (denoted as HU or GV) or AG dinucleotide at splice acceptor site (denoted as BG or AH). (B) Representation of the differential splicing events between samples with different genotypes for one elncRNA, ENSG00000205786. Median differential splicing (log10 modulus fold difference in Percentage-Spliced-In (dPSI)) of each splicing event is noted next to the arrow. Decreases are represented in red and increases represented in dark blue. (C) Distribution of the log10 modulus fold difference in PSI, relative to the median of samples with reference genotype, between individuals that carry alternative or reference alleles at each corresponding elncRNA splicing event as shown in (B). (D-H) Distribution of the log10 modulus fold difference in splicing (PSI), relative to the median of samples with reference genotype, between individuals that carry alternative or reference alleles at elncRNA splice site of ENSG00000205786, of all directly affected splicing events of (D) the elncRNA and (E) all splicing events of its target protein coding genes (5 targets); as well as fold difference in expression levels (RPKM) of (F) targets, (G) non-targets and (H) the elncRNA. Differences between groups were tested using a two-tailed Mann-Whitney U test. * p < 0.05; ** p < 0.01; *** p < 0.001; NS p > 0.05.
Impact of elncRNA splicing on cis-gene regulation in the human population
(A) LINC00886 (ENSG00000240875) is a multi-exonic elncRNA whose splicing is associated with genotype of the SNP variant (rs187202716), which is also associated to the expression level of an elncRNA target, LEKR1 (ENSG00000197980). (Top panel) Distribution of the Percentage-Spliced-In (PSI) of elncRNA splicing (LINC00886, red) and target expression (RPKM) (LEKR1, green) in samples across the population that carry different alleles of the SNP variant (rs187202716). Spearman’s rho and p-values are shown. (Bottom panel) Genome browser illustrating the genomic positions of the elncRNA (LINC00886, red), its target (LEKR1, green), and their associated SNP (rs187202716, black). (B) Four models of causal inference testing that predict the relationship between joint seQTL variant (black box) associations with the splicing (sQTL) of multi-exonic elncRNAs (red boxes) and the expression level (eQTL) of their target protein-coding genes (green boxes). Schematic representation of the models of joint seQTL associations: (1) the variants are independently associated with elncRNA splicing and target expression (independent model); (2) direct association between the variant and elncRNA splicing mediates the indirect association between that and target expression (causal model); (3) direct association between the variant with target expression mediates the indirect association between that and elncRNA splicing (reactive model); and (4) the causative interaction between elncRNA splicing and target expression is more complex (undecided model). Direct associations are depicted as solid lines and indirect associations as dash lines. (C) Scatterplot depicting causal inference testing local FDR associated with each the four models (as illustrated in B). Number and proportion of joint seQTLs are provided in brackets for each model. Dotted red lines denote significance threshold at local FDR < 0.1. (D) Proportion of joint elncRNA seQTLs causally or non-casually predicted to mediate target gene expression associated with splicing junctions located at the 5´ end (red) or 3´ end (grey) ends of the transcripts. Differences between groups were tested using a two-tailed Fisher’s exact test. *** p < 0.001.
Impact of elncRNA splicing on cognate enhancer chromatin signatures in the human population
Distribution of the average log modulus fold difference in chromatin signature, relative to the median of samples with reference genotype, between individuals that carry alternative or reference alleles at elncRNA splice sites, in (A) H3K4me1 and (B) H3K27ac. (C-D) Scatterplot depicting joint elncRNA scQTLs in which elncRNA splicing is causally or non-casually predicted to mediate enhancer chromatin signatures, (C) H3K4me1 and (D) H3K27ac, respectively, using causal inference testing as illustrated using local FDR associated with the four models (as shown in Fig 4B). Dotted black lines denote significance threshold at local FDR < 0.1. Differences between groups were tested using a two-tailed Mann-Whitney U test. ** p < 0.01.
The activity of human enhancers is modulated by the splicing of their associated lncRNAs

January 2022

·

111 Reads

·

14 Citations

Pervasive enhancer transcription is at the origin of more than half of all long noncoding RNAs in humans. Transcription of enhancer-associated long noncoding RNAs (elncRNA) contribute to their cognate enhancer activity and gene expression regulation in cis. Recently, splicing of elncRNAs was shown to be associated with elevated enhancer activity. However, whether splicing of elncRNA transcripts is a mere consequence of accessibility at highly active enhancers or if elncRNA splicing directly impacts enhancer function, remains unanswered. We analysed genetically driven changes in elncRNA splicing, in humans, to address this outstanding question. We showed that splicing related motifs within multi-exonic elncRNAs evolved under selective constraints during human evolution, suggesting the processing of these transcripts is unlikely to have resulted from transcription across spurious splice sites. Using a genome-wide and unbiased approach, we used nucleotide variants as independent genetic factors to directly assess the causal relationship that underpin elncRNA splicing and their cognate enhancer activity. We found that the splicing of most elncRNAs is associated with changes in chromatin signatures at cognate enhancers and target mRNA expression. We provide evidence that efficient and conserved processing of enhancer-associated elncRNAs contributes to enhancer activity.


FIG. 1. Prevalence, length, and start codon sequence context of lincRNA ORFs. (A) Percentage of lincRNA transcripts with ORFs (>10 codons; red dot) and percentage of random control regions with ORFs (in introns shown as gray and in intergenic regions shown as black violins), for five species. For each lincRNA transcript, 10 length-and GþC content-matched control sequences were randomly selected from introns and intergenic regions. P values are indicated from a one-sample t-test. (B) Median length of longest ORFs in lincRNAs (red), and intronic (gray) and intergenic (black) control sequences. Error bars represent median absolute deviations. P values are indicated from Wilcoxon's rank-sum test. (C) Information content (see "Materials and Methods") for the region 612 nucleotides around AUG start codons for different ORFs (columns, indicated on top) and species (rows, indicated left). The sequence motif around mRNA start codons shows some similarity with the Kozak consensus sequence (gcc(A/G)ccAUGG). (D) Sequence similarity with the consensus mRNA sequence motif for the region 612 nucleotides around start codons (see "Materials and Methods") for mRNA coding regions (blue), lincRNA longest ORFs (red), longest ORFs in intronic and intergenic control sequences (black), and longest ORFs in 3 0 UTRs (green) and 5 0 UTRs (magenta). P values (<0.05) are indicated from Wilcoxon's rank-sum test to compare lincRNAs with control regions, 3 0 UTRs, or 5 0 UTRs. P values are marked red if the median lincRNA value is below the value for the other region.
FIG. 2. Comparison of trinucleotides ("codons") in mRNA, lincRNA, control sequences, and 3 0 UTRs. (A) First two components space from a multiple correspondence analysis performed on trinucleotide ("codon") counts (excluding start and stop codons) in mRNA coding regions (blue) and longest ORFs in lincRNAs (red), intronic and intergenic control sequences (black), and 3 0 UTRs (green) for five species. Results for 5 0 UTRs are shown in supplementary figure S2, Supplementary Material online. (B) Spearman correlation coefficients between codon frequencies for the same ORFs and species as in (A).
FIG. 4. Cell-type-specific tAIs and ribosome-binding in five human cell lines. (A) Clustering (based on Euclidean distance) of effective tRNA anticodon frequencies, estimated from smallRNA-Seq data, in five human cell lines (indicated at bottom). (B) Boxplots of cell-type-specific tAIs (zscored relative to those of intronic and intergenic control ORFs) for top-1/3 cytoplasmic expressed mRNAs (dark blue), mRNAs with cytoplasmic expression levels matching those of top-1/3 cytoplasmic lincRNAs (light blue), expressed lincRNAs (red), top-1/3 cytoplasmic expressed lincRNAs (orange), and intronic and intergenic control sequences (gray), for five human cell lines (label at bottom). The numbers of top-1/3 cytoplasmic mRNAs and lincRNAs are indicated in parentheses at the bottom. P values are calculated using Wilcoxson's rank sum test. (C) Correspondence between cell-type-specific tAIs (z-scored relative to control tAIs) and relative ribosome-binding (see "Materials and Methods") for 41 lincRNAs that are classified as cytoplasmic in all five human cell lines. Dots represent median values and error bars median absolute deviations. Pearson (r) and Spearman (q) correlation coefficients between median values are indicated.
Analysis of Eukaryotic lincRNA Sequences Indicates Signatures of Hindered Translation Linked to Selection Pressure

December 2021

·

34 Reads

Molecular Biology and Evolution

Long intergenic non-coding RNAs (lincRNAs) represent a large fraction of transcribed loci in eukaryotic genomes. Although classified as non-coding, most lincRNAs contain open reading frames (ORFs), and it remains unclear why cytoplasmic lincRNAs are not or very inefficiently translated. Here, we analysed signatures of hindered translation in lincRNA sequences from five eukaryotes, covering a range of natural selection pressures. In fission yeast and C elegans, i.e. species under strong selection, we detected significantly shorter ORFs, a suboptimal sequence context around start codons for translation initiation, and trinucleotides (“codons”) corresponding to less abundant tRNAs than for neutrally evolving control sequences, likely impeding translation elongation. For human, we detected signatures for cell type-specific hindrance of lincRNA translation, in particular codons in abundant cytoplasmic lincRNAs corresponding to lower expressed tRNAs than control codons, in three out of five human cell lines. We verified that varying tRNA expression levels between cell lines are reflected in the amount of ribosomes bound to cytoplasmic lincRNAs in each cell line. We propose that codons at ORF starts are particularly important for reducing ribosome-binding to cytoplasmic lincRNA ORFs. Altogether, our analyses indicate that in species under stronger selection lincRNAs evolved sequence features generally hindering translation and support cell type-specific hindrance of translation efficiency in human lincRNAs. The sequence signatures we have identified may improve predicting peptide- and genuine non-coding lincRNAs in a cell type.


Figure 3. Micropeptide-encoding transcript expression is post-transcriptionally regulated by miRNAs.
Translation is required for miRNA-dependent decay of endogenous transcripts

December 2020

·

94 Reads

·

30 Citations

The EMBO Journal

Post-transcriptional repression of gene expression by miRNAs occurs through transcript destabilization or translation inhibition. mRNA decay is known to account for most miRNA-dependent repression. However, because transcript decay occurs co-translationally, whether target translation is a requirement for miRNA-dependent transcript destabilization remains unknown. To decouple these two molecular processes, we used cytosolic long noncoding RNAs (lncRNAs) as models for endogenous transcripts that are not translated. We show that, despite interacting with the miRNA-loaded RNA-induced silencing complex, the steady-state abundance and decay rates of these transcripts are minimally affected by miRNA loss. To further validate the apparent requirement of translation for miRNA-dependent decay, we fused two lncRNA candidates to the 3'-end of a protein-coding gene reporter and found this results in their miRNA-dependent destabilization. Further analysis revealed that the few natural lncRNAs whose levels are regulated by miRNAs in mESCs tend to associate with translating ribosomes, and possibly represent misannotated micropeptides, further substantiating the necessity of target translation for miRNA-dependent transcript decay. In summary, our analyses suggest that translation is required for miRNA-dependent transcript destabilization, and demonstrate that the levels of coding and noncoding transcripts are differently affected by miRNAs.


BAZ2A safeguards genome architecture of ground-state pluripotent stem cells

December 2020

·

44 Reads

·

12 Citations

Chromosomes have an intrinsic tendency to segregate into compartments, forming long-distance contacts between loci of similar chromatin states. How genome compartmentalization is regulated remains elusive. Here, comparison of mouse ground-state embryonic stem cells (ESCs) characterized by open and active chromatin, and advanced serum ESCs with a more closed and repressed genome, reveals distinct regulation of their genome organization due to differential dependency on BAZ2A/TIP5, a component of the chromatin remodeling complex NoRC. On ESC chromatin, BAZ2A interacts with SNF2H, DNA topoisomerase 2A (TOP2A) and cohesin. BAZ2A associates with chromatin sub-domains within the active A compartment, which intersect through long-range contacts. We found that ground-state chromatin selectively requires BAZ2A to limit the invasion of active domains into repressive compartments. BAZ2A depletion increases chromatin accessibility at B compartments. Furthermore, BAZ2A regulates H3K27me3 genome occupancy in a TOP2A-dependent manner. Finally, ground-state ESCs require BAZ2A for growth, differentiation, and correct expression of developmental genes. Our results uncover the propensity of open chromatin domains to invade repressive domains, which is counteracted by chromatin remodeling to establish genome partitioning and preserve cell identity.


A circular RNA generated from an intron of the insulin gene controls insulin secretion

November 2020

·

263 Reads

·

75 Citations

Fine-tuning of insulin release from pancreatic β-cells is essential to maintain blood glucose homeostasis. Here, we report that insulin secretion is regulated by a circular RNA containing the lariat sequence of the second intron of the insulin gene. Silencing of this intronic circular RNA in pancreatic islets leads to a decrease in the expression of key components of the secretory machinery of β-cells, resulting in impaired glucose- or KCl-induced insulin release and calcium signaling. The effect of the circular RNA is exerted at the transcriptional level and involves an interaction with the RNA-binding protein TAR DNA-binding protein 43 kDa (TDP-43). The level of this circularized intron is reduced in the islets of rodent diabetes models and of type 2 diabetic patients, possibly explaining their impaired secretory capacity. The study of this and other circular RNAs helps understanding β-cell dysfunction under diabetes conditions, and the etiology of this common metabolic disorder.


Figure 2. (A) Number of differentially expressed genes (DEGs) (messenger RNAs (mRNAs)) per comparison. The effect of iodide on wild-type (WT) mice (WTI/WTC) and on Nrf2 knockout (KO) mice (KOI/KO), as well as the effect of genotype (KOC/WTC) on gene expression, were assessed by the DESeq2 method, with a false discovery rate (FDR) set at 0.05 and the minimum gene expression fold change set at 1.5. Table S2 (Supplementary Materials) contains full lists of DEGs. (B) Venn diagrams with upregulated and downregulated genes (mRNAs) after iodide in WT and Nrf2 KO mice. (C) Volcano plots of genes that are differentially expressed (FDR < 0.05, fold change ≥ 1.5) in WT mice after iodide (WTI/WTC). (D) Volcano plots of genes that are differentially expressed (FDR < 0.05, fold change ≥ 1.5) in Nrf2 KO mice versus WT mice (KOC/WTC) without excess iodide. Some gene names are indicatively noted.
Figure 4. (A) Top canonical pathways enriched among DEGs (mRNAs) in Nrf2 KO mice compared to WT mice (KOC/WTC) ranked by p-value, as assessed by Ingenuity Pathway Analysis (IPA). The absolute z-score was set at ≥1.5. White bars indicate downregulation of the respective pathway (negative z-score), while black bars indicate upregulation of the respective pathway (positive z-score). The asterisk (*) before the pathway name indicates that it is relevant to Nrf2 signaling. (B) Top canonical pathways enriched in WT mice after iodide (WTI/WTC) ranked by p-value (as assessed by
Figure 5. Pathway analysis of genes (mRNAs) that respond differentially to iodide among Nrf2 KO and WT mice using Parametric Gene Set Enrichment Analysis (PGSEA). The heatmap depicts the trend for each Gene Ontology (GO) biological process for each sample. Red squares indicate upregulation (positive z-score), while blue squares indicate downregulation (negative z-score). GO terms marked with an asterisk (*) refer to processes relevant to inflammation/autoimmunity. The top 30 pathways are shown, with FDR set at <0.05. Table S5 (Supplementary Materials) shows the z-score values for each pathway per sample and the respective p-values for each pathway.
Figure 6. (A) Heatmap of genes (mRNAs) that participate in the "activation of T-lymphocytes" process, which is enriched in Nrf2 KO mice after iodide as compared to WT mice after iodide. This process is upregulated in Nrf2 KO mice after iodide (z-score = 3.466, IPA algorithm), while, in WT mice, it is not enriched at all after iodide. Red color indicates upregulation by ≥1.5-fold, while blue color indicates downregulation by ≥1.5-fold (p < 0.05), and white color indicates gene expression change <1.5-fold and/or p ≥ 0.05. (B) Heatmap of genes (mRNAs) that participate in the "Nrf2-mediated oxidative stress response", which is enriched in Nrf2 KO mice treated after iodide as compared to WT mice after iodide. This pathway is upregulated in WT mice after iodide (z-score = 2.333, IPA algorithm), while, in Nrf2 KO mice, it is not enriched at all after iodide. Color-coding in the heatmap is similar to that in panel A.
Figure 7. (A) Comparison of predicted upstream regulators of mRNAs differentially expressed after iodide treatment in WT mice as compared to Nrf2 KO mice (WTI/WTC vs. KOI/KOC, z-score ≥4 and p < 0.01). Red color indicates upregulation (positive z-score), while white color indicates absence of enrichment of the upstream regulator. The graph was generated by the comparison analysis of IPA. (B) Gene expression heatmap of genes that participate in the fibrosis pathway, which is predicted to be upregulated by the enriched upstream regulator TGFβ after iodide. Red color indicates ≥1.5-fold upregulation (p < 0.05), while white color indicates gene expression change <1.5-fold and/or p ≥ 0.05. The right column continues after the left one. Genes are presented from top to bottom from lower to higher p-value in either comparison (WTI/WTC, KOI/KOC).
The Transcriptomic Response of the Murine Thyroid Gland to Iodide Overload and the Role of the Nrf2 Antioxidant System

September 2020

·

142 Reads

·

11 Citations

Background: Thyroid follicular cells have physiologically high levels of reactive oxygen species because oxidation of iodide is essential for the iodination of thyroglobulin (Tg) during thyroid hormone synthesis. Thyroid follicles (the functional units of the thyroid) also utilize incompletely understood autoregulatory mechanisms to defend against exposure to excess iodide. To date, no transcriptomic studies have investigated these phenomena in vivo. Nuclear erythroid factor 2 like 2 (Nrf2 or Nfe2l2) is a transcription factor that regulates the expression of numerous antioxidant and other cytoprotective genes. We showed previously that the Nrf2 pathway regulates the antioxidant defense of follicular cells, as well as Tg transcription and Tg iodination. We, thus, hypothesized that Nrf2 might be involved in the transcriptional response to iodide overload. Methods: C57BL6/J wild-type (WT) or Nrf2 knockout (KO) male mice were administered regular water or water supplemented with 0.05% sodium iodide for seven days. RNA from their thyroids was prepared for next-generation RNA sequencing (RNA-Seq). Gene expression changes were assessed and pathway analyses were performed on the sets of differentially expressed genes. Results: Analysis of differentially expressed messenger RNAs (mRNAs) indicated that iodide overload upregulates inflammatory-, immune-, fibrosis- and oxidative stress-related pathways, including the Nrf2 pathway. Nrf2 KO mice showed a more pronounced inflammatory-autoimmune transcriptional response to iodide than WT mice. Compared to previously published datasets, the response patterns observed in WT mice had strong similarities with the patterns typical of Graves' disease and papillary thyroid carcinoma (PTC). Long non-coding RNAs (lncRNAs) and microRNAs (miRNAs) also responded to iodide overload, with the latter targeting mRNAs that participate mainly in inflammation pathways. Conclusions: Iodide overload induces the Nrf2 cytoprotective response and upregulates inflammatory, immune, and fibrosis pathways similar to autoimmune hyperthyroidism (Graves' disease) and PTC.



Figure 3.
Noncanonical targeting contributes significantly to miRNA-mediated regulation

July 2020

·

69 Reads

·

1 Citation

Determining which genes are targeted by miRNAs is crucial to elucidate their contributions to diverse biological processes in health and disease. Most miRNA target prediction tools rely on the identification of complementary regions between transcripts and miRNAs. Whereas important for target recognition, the presence of complementary sites is not sufficient to identify transcripts targeted by miRNAs. Here, we describe an unbiased statistical genomics approach that explores genetically driven changes in gene expression between human individuals. Using this approach, we identified transcripts that respond to physiological changes in miRNA levels. We found that a much smaller fraction of mRNAs expressed in lymphoblastoid cell lines (LCLs) than what is predicted by other tools is targeted by miRNAs. We estimate that each miRNA has a relatively small number of targets. The transcripts we predict to be miRNA targets are enriched in AGO-binding and previously validated miRNAs target interactions, supporting the reliability of our predictions. Consistent with previous analysis, these targets are also enriched among dosage sensitive and highly controlled genes. Almost a third of genes we predict to be miRNA targets lack sequence complementarity to the miRNA seed region (noncanonical targets). These noncanonical targets have higher complementary with the miRNA 3' end. The impact of miRNAs on the levels of their canonical or noncanonical targets is identical supporting the relevance of this poorly explored mechanism of targeting.


Citations (36)


... Recent developments in in vivo uorescence and superresolution microscopy enabled the visualization of transcription dynamics in living cells (5,41,42). scRNA-seq and lineage tracing have been coupled to combine clonal information with cell transcriptomes (43,44), and metabolic labeling has been used to track newly synthesized RNA to allow the study of transcriptional dynamics to investigate how perturbations impact gene expression (45)(46)(47). However, understanding how the initial state of a cell's transcriptome in uences its response to a perturbation remains challenging due to the assumptions required to infer a cell's initial state when using methodologies that require cell lysis (10). ...

Reference:

Single-cell nanobiopsy enables multigenerational longitudinal transcriptomics of cancer cells
Estimating RNA dynamics using one time point for one sample in a single-pulse metabolic labeling experiment

BMC Bioinformatics

... Rights reserved. of nearby genes [15]. LncRNAs transcribed near enhancer regions modulate chromatin looping and splicing [16] while also playing key roles in nuclear organization, sorting, and transport [17]. The functional versatility of lncRNAs is driven by the sequence motifs and secondary structures they contain. ...

The activity of human enhancers is modulated by the splicing of their associated lncRNAs

... In prostate cancer BAZ2A is overexpressed and is involved in maintaining prostate cancer cell growth 45,46 . Interestingly, BAZ2A overexpression is also tightly associated with a molecular subtype displaying a CpG island methylator phenotype (CIMP) 47,72,73 , which could suggest an elevated silencing through repressive histone marks and DNA methylation of p14/p16 and p15 loci in pancreatic cancer. Collectively, our research discoveries will make it possible to use novel epigenetic machineries as therapeutic targets for speci cally killing the harmful cancer-forming cells in patients. ...

BAZ2A safeguards genome architecture of ground-state pluripotent stem cells

... Apart from the relatively short, unstable, and unspliced bidirectional eRNAs (De Santa et al., 2010;Kim et al., 2010;Djebali et al., 2012), a subset of enhancers transcribe unidirectional lncRNAs showing different characteristics in terms of splicing and stability Marques et al., 2013;Hon et al., 2017;Gil and Ulitsky, 2018;Tan et al., 2018). In a previous study, using solely the H3K4me1 histone modification mark and the correlation of the H3K4me1 signal with cell type-specific expression of putative mRNA targets to predict enhancers (Corradin et al., 2014), it was estimated that about one third of Gencode-annotated lincRNAs overlap cell type-specific enhancers (Vucicevic et al., 2015). ...

An unexpected contribution of lincRNA splicing to enhancer function

... Although much progress has been made, target identification remains a challenge because of the limited understanding of the molecular basis of miRNA-target coupling, but also due to the context-dependence of post-transcriptional regulation and miRNA mode of action [5][6][7]. Generally, miRNAs downregulate proteins through a combination of translational inhibition and promotion of mRNA decay [8][9][10], even though which mechanisms of action of microRNAs are the most dominant remains a matter of debate. The emergence of high-throughput methods in the past decades has allowed researchers to address the question of miRNA action on a global scale. ...

Translation is required for miRNA-dependent decay of endogenous transcripts

The EMBO Journal

... The islets of rats and individuals with type 2 diabetes have decreased levels of ci-Ins2/ci-INS. 45 CircHIPK3 plays a role in insulin secretion and beta-cell proliferation, while CircCAMSAP1 has implications for insulin resistance and glucose homeostasis. 46 Shan et al. 47 reported that retinal vascular dysfunction in diabetes mellitus is mediated by circHIPK3. ...

A circular RNA generated from an intron of the insulin gene controls insulin secretion

... The hypertranscription and active chromatin state of the ESC genome are mirrored in the nucleolus of ESCs due to the lack of heterochromatic and silent rRNA genes and the higher rRNA transcription levels relative to differentiated cells (Savi c et al., 2014;Schlesinger et al., 2009) (Figure 1). DNA hypomethylation of rRNA genes and their lack of silencing was reported in both ground-state pluripotent and developmentally primed mESCs (Dalcher et al., 2019;Savi c et al., 2014;Schlesinger et al., 2009). These ESC types are known to have distinct epigenetic signatures, such as a low DNA methylation content in ground-state pluripotent cells and high CpG methylation in primed mESCs (Ficz et al., 2013;Habibi et al., 2013;Leitch et al., 2013;Marks et al., 2012). ...

TIP5 safeguards genome architecture of ground-state pluripotent stem cells
  • Citing Preprint
  • December 2019

... Transcriptome analyses of the thyroid gland focusing on thyroid dysfunction have been limited to reports on a murine model of Graves' disease 40 and a murine model of iodine treatment. 41 To the best of our knowledge, there have been no clinical reports about the thyroid transcriptome of patients with hyperthyroidism. In the present study, our hyperthyroid mice provided valuable information on the molecular signature of hyperthyroidism. ...

The Transcriptomic Response of the Murine Thyroid Gland to Iodide Overload and the Role of the Nrf2 Antioxidant System

... Thus, multi-OMICs integration facilitated the identification of 759 candidate miRNA targets corresponding to 7% of expressed genes. Interestingly, Tan et al (preprint: 2020) estimated that 6% of expressed genes were regulated by miRNAs in lymphoblastoid cell lines (preprint: Tan et al, 2020), suggesting that this percentage might reflect functionally relevant interaction number magnitudes in various contexts. ...

Noncanonical targeting contributes significantly to miRNA-mediated regulation

... poly-Gly/Ala tail. 101 Aberrant elongation of CAG/CTG repeat sequences is associated with the onset of Huntington's disease, spinocerebellar ataxia, and myotonic dystrophy. 102,103 Research efforts have focused on developing therapeutic agents that target and bind to CAG/CTG repeat sequences to treat these neurological disorders. ...

Three-dimensional chromatin interactions remain stable upon CAG/CTG repeat expansion

Science Advances