Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Developmental dynamics of lncRNAs across mammalian organs and species

Abstract

Although many long noncoding RNAs (lncRNAs) have been identified in human and other mammalian genomes, there has been limited systematic functional characterization of these elements. In particular, the contribution of lncRNAs to organ development remains largely unexplored. Here we analyse the expression patterns of lncRNAs across developmental time points in seven major organs, from early organogenesis to adulthood, in seven species (human, rhesus macaque, mouse, rat, rabbit, opossum and chicken). Our analyses identified approximately 15,000 to 35,000 candidate lncRNAs in each species, most of which show species specificity. We characterized the expression patterns of lncRNAs across developmental stages, and found many with dynamic expression patterns across time that show signatures of enrichment for functionality. During development, there is a transition from broadly expressed and conserved lncRNAs towards an increasing number of lineage- and organ-specific lncRNAs. Our study provides a resource of candidate lncRNAs and their patterns of expression and evolutionary conservation across mammalian organ development.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: lncRNAs expressed during mammalian organ development.
Fig. 2: Developmentally dynamic lncRNAs are enriched for functional loci.
Fig. 3: Patterns of dynamic lncRNA expression.
Fig. 4: Co-expression with adjacent protein-coding genes.

Similar content being viewed by others

Data availability

Data are available from the corresponding authors upon reasonable request.

References

  1. Cabili, M. N. et al. Integrative annotation of human large intergenic noncoding RNAs reveals global properties and specific subclasses. Genes Dev. 25, 1915–1927 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Derrien, T. et al. The GENCODE v7 catalog of human long noncoding RNAs: analysis of their gene structure, evolution, and expression. Genome Res. 22, 1775–1789 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Iyer, M. K. et al. The landscape of long noncoding RNAs in the human transcriptome. Nat. Genet. 47, 199–208 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Hon, C. C. et al. An atlas of human long non-coding RNAs with accurate 5′ ends. Nature 543, 199–204 (2017).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  5. Carninci, P. et al. The transcriptional landscape of the mammalian genome. Science 309, 1559–1563 (2005).

    ADS  CAS  PubMed  Google Scholar 

  6. Necsulea, A. et al. The evolution of lncRNA repertoires and expression patterns in tetrapods. Nature 505, 635–640 (2014).

    ADS  CAS  PubMed  Google Scholar 

  7. Washietl, S., Kellis, M. & Garber, M. Evolutionary dynamics and tissue specificity of human long noncoding RNAs in six mammals. Genome Res. 24, 616–628 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Hezroni, H. et al. Principles of long noncoding RNA evolution derived from direct comparison of transcriptomes in 17 species. Cell Rep. 11, 1110–1122 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Kopp, F. & Mendell, J. T. Functional classification and experimental dissection of long noncoding RNAs. Cell 172, 393–407 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Ponting, C. P., Oliver, P. L. & Reik, W. Evolution and functions of long noncoding RNAs. Cell 136, 629–641 (2009).

    CAS  PubMed  Google Scholar 

  11. Ulitsky, I. Evolution to the rescue: using comparative genomics to understand long non-coding RNAs. Nat. Rev. Genet. 17, 601–614 (2016).

    CAS  PubMed  Google Scholar 

  12. Necsulea, A. & Kaessmann, H. Evolutionary dynamics of coding and non-coding transcriptomes. Nat. Rev. Genet. 15, 734–748 (2014).

    CAS  PubMed  Google Scholar 

  13. Guttman, M. et al. Chromatin signature reveals over a thousand highly conserved large non-coding RNAs in mammals. Nature 458, 223–227 (2009).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  14. Ulitsky, I., Shkumatava, A., Jan, C. H., Sive, H. & Bartel, D. P. Conserved function of lincRNAs in vertebrate embryonic development despite rapid sequence evolution. Cell 147, 1537–1550 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Sauvageau, M. et al. Multiple knockout mouse models reveal lincRNAs are required for life and brain development. eLife. 2, e01749 (2013).

    PubMed  PubMed Central  Google Scholar 

  16. Grote, P. & Herrmann, B. G. Long noncoding RNAs in organogenesis: making the difference. Trends Genet. 31, 329–335 (2015).

    CAS  PubMed  Google Scholar 

  17. Goff, L. A. et al. Spatiotemporal expression and transcriptional perturbations by long noncoding RNAs in the mouse brain. Proc. Natl Acad. Sci. USA 112, 6855–6862 (2015).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  18. Cardoso-Moreira, M. et al. Gene expression across mammalian organ development. Nature https://doi.org/10.1038/s41586-019-1338-5 (2019).

  19. Zerbino, D. R. et al. Ensembl 2018. Nucleic Acids Res. 46, D754–D761 (2018).

    CAS  PubMed  Google Scholar 

  20. Conesa, A., Nueda, M. J., Ferrer, A. & Talón, M. maSigPro: a method to identify significantly differential expression profiles in time-course microarray experiments. Bioinformatics 22, 1096–1102 (2006).

    CAS  PubMed  Google Scholar 

  21. Liu, S. J. et al. CRISPRi-based genome-scale identification of functional long noncoding RNA loci in human cells. Science 355, eaah7111 (2017).

    Google Scholar 

  22. Mukherjee, N. et al. Integrative classification of human coding and noncoding genes through RNA metabolism profiles. Nat. Struct. Mol. Biol. 24, 86–96 (2017).

    CAS  PubMed  Google Scholar 

  23. Soumillon, M. et al. Cellular source and mechanisms of high transcriptome complexity in the mammalian testis. Cell Rep. 3, 2179–2190 (2013).

    CAS  PubMed  Google Scholar 

  24. Guttman, M. & Rinn, J. L. Modular regulatory principles of large non-coding RNAs. Nature 482, 339–346 (2012).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  25. Andersson, R. et al. An atlas of active enhancers across human cell types and tissues. Nature 507, 455–461 (2014).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  26. Kutter, C. et al. Rapid turnover of long noncoding RNAs and the evolution of gene expression. PLoS Genet. 8, e1002841 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Quek, X. C. et al. lncRNAdbv2.0: expanding the reference database for functional long noncoding RNAs. Nucleic Acids Res. 43, D168–D173 (2015).

    CAS  PubMed  Google Scholar 

  28. Melé, M. et al. Chromatin environment, transcriptional regulation, and splicing distinguish lincRNAs and mRNAs. Genome Res. 27, 27–37 (2017).

    PubMed  PubMed Central  Google Scholar 

  29. Yevshin, I., Sharipov, R., Valeev, T., Kel, A. & Kolpakov, F. GTRD: a database of transcription factor binding sites identified by ChIP-seq experiments. Nucleic Acids Res. 45, D61–D67 (2017).

    CAS  PubMed  Google Scholar 

  30. Olson, E. N. Gene regulatory networks in the evolution and development of the heart. Science 313, 1922–1927 (2006).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  31. Ruf, S. et al. Large-scale analysis of the regulatory architecture of the mouse genome with a transposon-associated sensor. Nat. Genet. 43, 379–386 (2011).

    CAS  PubMed  Google Scholar 

  32. Engreitz, J. M. et al. Local regulation of gene expression by lncRNA promoters, transcription and splicing. Nature 539, 452–455 (2016).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  33. Amaral, P. P. et al. Genomic positional conservation identifies topological anchor point RNAs linked to developmental loci. Genome Biol. 19, 32 (2018).

    PubMed  PubMed Central  Google Scholar 

  34. Luo, S. et al. Divergent lncRNAs regulate gene expression and lineage differentiation in pluripotent cells. Cell Stem Cell 18, 637–652 (2016).

    CAS  PubMed  Google Scholar 

  35. Bester, A. C. et al. An integrated genome-wide CRISPRa approach to functionalize lncRNAs in drug resistance. Cell 173, 649–664 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Jiang, W., Liu, Y., Liu, R., Zhang, K. & Zhang, Y. The lncRNA DEANR1 facilitates human endoderm differentiation by activating FOXA2 expression. Cell Rep. 11, 137–148 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Jian, X. & Felsenfeld, G. Insulin promoter in human pancreatic β cells contacts diabetes susceptibility loci and regulates genes affecting insulin metabolism. Proc. Natl Acad. Sci. USA 115, E4633–E4641 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Spigoni, G., Gedressi, C. & Mallamaci, A. Regulation of Emx2 expression by antisense transcripts in murine cortico-cerebral precursors. PLoS ONE 5, e8658 (2010).

    ADS  PubMed  PubMed Central  Google Scholar 

  39. Ramos, A. D. et al. Integration of genome-wide approaches identifies lncRNAs of adult neural stem cells and their progeny in vivo. Cell Stem Cell 12, 616–628 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Li, W., Notani, D. & Rosenfeld, M. G. Enhancers as non-coding RNA transcription units: recent insights and future perspectives. Nat. Rev. Genet. 17, 207–223 (2016).

    CAS  PubMed  Google Scholar 

  41. Liu, S. J. et al. Single-cell analysis of long non-coding RNAs in the developing human neocortex. Genome Biol. 17, 67 (2016).

    PubMed  PubMed Central  Google Scholar 

  42. Lagarde, J. et al. High-throughput annotation of full-length long noncoding RNAs with capture long-read sequencing. Nat. Genet. 49, 1731–1740 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Li, H. et al. The Sequence Alignment/Map format and SAMtools. Bioinformatics 25, 2078–2079 (2009).

    PubMed  PubMed Central  Google Scholar 

  44. Pertea, M. et al. StringTie enables improved reconstruction of a transcriptome from RNA-seq reads. Nat. Biotechnol. 33, 290–295 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Trapnell, C. et al. Differential gene and transcript expression analysis of RNA-seq experiments with TopHat and Cufflinks. Nat. Protocols 7, 562–578 (2012).

    CAS  PubMed  Google Scholar 

  46. Wang, L. et al. CPAT: Coding-Potential Assessment Tool using an alignment-free logistic regression model. Nucleic Acids Res. 41, e74 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Washietl, S. et al. RNAcode: robust discrimination of coding and noncoding regions in comparative sequence data. RNA 17, 578–594 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Altschul, S. F., Gish, W., Miller, W., Myers, E. W. & Lipman, D. J. Basic local alignment search tool. J. Mol. Biol. 215, 403–410 (1990).

    CAS  PubMed  Google Scholar 

  49. The UniProt Consortium. UniProt: the universal protein knowledgebase. Nucleic Acids Res. 45, D158–D169 (2017).

    Google Scholar 

  50. Finn, R. D. et al. The Pfam protein families database: towards a more sustainable future. Nucleic Acids Res. 44, D279–D285 (2016).

    CAS  PubMed  Google Scholar 

  51. Anders, S., Pyl, P. T. & Huber, W. HTSeq—a Python framework to work with high-throughput sequencing data. Bioinformatics 31, 166–169 (2015).

    CAS  PubMed  Google Scholar 

  52. Robinson, M. D., McCarthy, D. J. & Smyth, G. K. edgeR: a Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics 26, 139–140 (2010).

    CAS  PubMed  Google Scholar 

  53. Love, M. I., Huber, W. & Anders, S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 15, 550 (2014).

    PubMed  PubMed Central  Google Scholar 

  54. Yanai, I. et al. Genome-wide midrange transcription profiles reveal expression level relationships in human tissue specification. Bioinformatics 21, 650–659 (2005).

    CAS  PubMed  Google Scholar 

  55. Li, L., Stoeckert, C. J. J. & Roos, D. S. OrthoMCL: identification of ortholog groups for eukaryotic genomes. Genome Res. 13, 2178–2189 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Duret, L., Chureau, C., Samain, S., Weissenbach, J. & Avner, P. The Xist RNA gene evolved in eutherians by pseudogenization of a protein-coding gene. Science 312, 1653–1655 (2006).

    ADS  CAS  PubMed  Google Scholar 

  57. Hezroni, H. et al. A subset of conserved mammalian long non-coding RNAs are fossils of ancestral protein-coding genes. Genome Biol. 18, 162 (2017).

    PubMed  PubMed Central  Google Scholar 

  58. Smit, A. F. A., Hubley, R. & Green, P. RepeatMasker Open-4.0. v.4.0.6 http://www.repeatmasker.org (2013–2015).

  59. Chen, J. et al. Evolutionary analysis across mammals reveals distinct classes of long non-coding RNAs. Genome Biol. 17, 19 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Quinlan, A. R. & Hall, I. M. BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics 26, 841–842 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Hinrichs, A. S. et al. The UCSC Genome Browser Database: update 2006. Nucleic Acids Res. 34, D590–D598 (2006).

    CAS  PubMed  Google Scholar 

  62. Wucher, V. et al. FEELnc: a tool for long non-coding RNA annotation and its application to the dog transcriptome. Nucleic Acids Res. 45, e57 (2017).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  63. Kolde, R. pheatmap: Pretty Heatmaps. v.1.0.10 (2015).

  64. Hensman, J., Rattray, M. & Lawrence, N. D. Fast variational inference in the conjugate exponential family. In Advances in Neural Information Processing Systems 25 1–9 (2012).

  65. Hensman, J., Lawrence, N. D. & Rattray, M. Hierarchical Bayesian modelling of gene expression time series across irregularly sampled replicates and clusters. BMC Bioinformatics 14, 252 (2013).

    PubMed  PubMed Central  Google Scholar 

  66. Hensman, J., Rattray, M. & Lawrence, N. D. Fast nonparametric clustering of structured time-series. IEEE Trans. Pattern Anal. Mach. Intell. 37, 383–393 (2015).

    PubMed  Google Scholar 

  67. Wang, J., Vasaikar, S., Shi, Z., Greer, M. & Zhang, B. WebGestalt 2017: a more comprehensive, powerful, flexible and interactive gene set enrichment analysis toolkit. Nucleic Acids Res. 45 (W1), W130–W137 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  68. Kaessmann, H. Origins, evolution, and phenotypic impact of new genes. Genome Res. 20, 1313–1326 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  69. Revelle, W. psych: Procedures for Personality and Psychological Research. R package v.1.8.4 https://cran.r-project.org/web/packages/psych/index.html (2017).

  70. Ebisuya, M., Yamamoto, T., Nakajima, M. & Nishida, E. Ripples from neighbouring transcription. Nat. Cell Biol. 10, 1106–1113 (2008).

    CAS  PubMed  Google Scholar 

  71. R Core Team. R: A Language and Environment for Statistical Computing https://www.r-project.org/ (R Foundation for Statistical Computing, 2008).

  72. Wickham, H., Romain, F., Henry, L. & Müller, K. dplyr: A Grammar of Data Manipulation. v.0.7.6 https://cran.r-project.org/web/packages/dplyr/index.html (2017).

  73. Wickham, H. tidyr: Easily Tidy Data with ‘spread()’ and ‘gather()’ Functions. v.0.8.1 https://tidyr.tidyverse.org/ (2018).

  74. Wickham, H. stringr: Simple, Consistent Wrappers for Common String Operations. v.1.3.1 https://stringr.tidyverse.org/ (2018).

  75. Dowle, M. & Srinivasan, A. data.table: Extension of ‘data.frame’. v.1.11.4 https://cran.r-project.org/web/packages/data.table/index.html (2017).

  76. Wickham, H. ggplot2: Elegant Graphics for Data Analysis 2nd edn (Springer, 2016).

  77. Auguie, B. gridExtra: Miscellaneous Functions for ‘Grid’ Graphics. v.2.3 https://rdrr.io/cran/gridExtra/ (2017).

  78. Wickham, H. Reshaping data with the reshape package. J. Stat. Softw. 21, 1–20 (2007).

    Google Scholar 

  79. Wickham, H. The split-apply-combine strategy for data analysis. J. Stat. Softw. 40, 1–29 (2011).

    Google Scholar 

  80. Lê, S., Josse, J. & Husson, F. FactoMineR: an R Package for multivariate analysis. J. Stat. Softw. 25, 1–18 (2008).

    Google Scholar 

Download references

Acknowledgements

We thank S. Anders, M. Sepp, E. Leushkin and members of the Kaessmann group for discussions, M. Sanchez-Delgado and N. Trost for assistance in figure design, and I. Moreira for help in the development of the interactive tool. We acknowledge support by the state of Baden-Württemberg through bwHPC and the German Research Foundation (DFG) through grant INST 35/1134-1 FUGG. This research was supported by grants from the European Research Council (615253, OntoTransEvol) and Swiss National Science Foundation (146474) to H.K., by the Marie Curie FP7-PEOPLE-2012-IIF to M.C.-M. (329902) and by a scholarship for MSc studies by the Alexander S. Onassis Public Benefit Foundation (F ZL 084-1/2015-2016) to I.S.

Peer review information

Nature thanks Camille Berthelot, Igor Ulitsky and the other anonymous reviewer(s) for their contribution to the peer review of this work.

Author information

Authors and Affiliations

Authors

Contributions

M.C.-M. and H.K. conceived and organized the study based on an original design by H.K. R.M. performed the lncRNA annotation and orthology assignment. I.S. performed all other analyses, under the supervision of M.C.-M. and H.K. I.S., M.C.-M. and H.K. wrote the manuscript, with input from R.M.

Corresponding authors

Correspondence to Margarida Cardoso-Moreira or Henrik Kaessmann.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data figures and tables

Extended Data Fig. 1 Annotation and orthology assignment of lncRNAs.

a, Schematic representation of the lncRNA annotation pipeline. b, Schematic representation of the pipeline for the detection of 1:1 lncRNA families.

Extended Data Fig. 2 Genomic classification and expression patterns of lncRNAs.

a, Distribution of lncRNAs among genomic classes in each species. b, Comparison of genomic classes (left), evolutionary age (middle) and organ of maximum expression (right) for known (Ensembl19) and newly annotated (novel) human lncRNAs. c, Number of species with a detected lncRNA member for human families of various evolutionary ages. d, Comparison of the fraction of species with a detected lncRNA member for human families conserved across mammals (180 Ma) and amniotes (300 Ma) with a previous study8. e, Fraction of lncRNAs and protein-coding gene orthologues found in conserved synteny with at least one protein-coding gene neighbour for increasing evolutionary distances. f, Organ of maximum expression for expressed lncRNAs (≥1 RPKM) in each species. g, Number of lncRNAs expressed (≥1 RPKM) in each species during the development of each organ (in logarithmic scale).

Extended Data Fig. 3 Features of developmentally dynamic lncRNA expression.

a, Representative examples of human developmentally dynamic (n = 5,887) and non-dynamic (n = 25,791) lncRNA expression profiles (mean expression; vertical bars represent the minimum and maximum values across replicates) for varying levels of maximum expression, replicate reproducibility and expression windows. The vertical dashed line represents birth; the horizontal dashed line marks 1 RPKM. b, Summary statistics for the lncRNAs and protein-coding genes in this study. c, Number of organs with developmentally dynamic expression for dynamic lncRNAs and protein-coding genes in each species. d, e, Tissue-specificity (d) and median time-specificity (e) of non-dynamic and dynamic lncRNAs, and protein-coding genes, across species. Tissue- and time-specificity indexes range from 0 (broad expression) to 1 (specific expression). All comparisons between non-dynamic and dynamic lncRNAs, and protein-coding genes are significant (P = 2.2 × 10−16, two-sided Mann–Whitney U-test). f, Maximum expression levels (log10(RPKM)) for developmentally dynamic and non-dynamic lncRNAs across species (excluding samples from the sexually mature testis). Developmentally dynamic lncRNAs are more highly expressed in all species (P = 2.2 × 10−16, two-sided Mann–Whitney U-test). Box plots are as in Fig. 2.

Extended Data Fig. 4 Functionality signature enrichments of developmentally dynamic lncRNAs.

a, Fraction of developmentally dynamic human lncRNAs (n = 5,887) for different genomic classes. Overrepresented classes were determined by comparing the fraction of dynamic lncRNAs in each class against all other classes. b, Normalized density distribution of the distance to the nearest protein-coding gene for dynamic (n = 5,887) and non-dynamic (n = 25,791) human lncRNAs. c, Generation of expression-matched dynamic (n = 2,906) and non-dynamic (n = 3,098) lncRNAs and their distribution among genomic classes. d, Fraction of developmentally dynamic human lncRNAs among isoforms with an increasing number of exons. The number of exons is significantly higher for developmentally dynamic lncRNAs (P = 2.2 × 10−16, two-sided Mann–Whitney U-test). e, Fraction of human lncRNAs that are intergenic, developmentally dynamic and that do not overlap enhancers25 (n = 16,481) among different age groups. f, Fraction of developmentally dynamic genes across expression-matched (n = 6,004) human lncRNAs of different age groups (top) and functionally characterized lncRNAs27 (bottom). g, Generation of expression-matched, lowly expressed (0.25–0.75 RPKM) dynamic (n = 798) and non-dynamic (n = 717) human lncRNAs and their distribution across different age groups. h, Fraction of developmentally dynamic human lncRNAs (n = 5,887) with or without a mouse (dynamic or not) orthologue (P = 2.2 × 10−16, Fisher’s exact test). i, Similarity of spatiotemporal expression (Spearman correlation coefficient between human and mouse organs/developmental stages) for 1:1 orthologues. j, Expression similarity across matched organs and developmental stages for mouse and rat 1:1 orthologous lncRNAs that are dynamic in both species, for different evolutionary ages. k, Fraction of lncRNAs present in the CRISPRi screen library21 resulting in a significant growth phenotype (hits) in at least one cell line for lncRNAs present (n = 2,364) or absent (n = 14,037) in our annotation and dynamic (n = 1,093) or non-dynamic (n = 1,277). l, Fraction of lncRNAs present in the CRISPRi screen library21 resulting in a significant growth phenotype (hits) in expression-matched dynamic (n = 2,906) and non-dynamic lncRNAs (n = 3,098). Box plots are as in Fig. 2. In al, statistical tests are two-sided.

Extended Data Fig. 5 Transcriptional regulation of dynamic lncRNAs in mouse.

a, Fraction of promoters of protein-coding genes, dynamic and non-dynamic lncRNAs, and size-matched random intergenic regions that overlap with binding sites for TFs. Each data point corresponds to a TF (n = 355). Box plots are as in Fig. 2. b, Selection of the 50 TFs with the highest binding variability across promoters of lncRNAs that were dynamic in different organs (in blue). TFs with maximum binding frequency ≤ 0.05 (red line) were not considered, as their high variability is probably associated with a low binding frequency. c, Spatiotemporal expression patterns of the 50 most variable TFs in mouse. The heat map is clustered by rows and shows expression levels in counts (after variance-stabilizing transformation).

Extended Data Fig. 6 Patterns of lncRNA expression in mammalian development.

a, Number of differentially expressed protein-coding genes and dynamic lncRNAs between adjacent stages of organ development in human, rat, rabbit, opossum and chicken. b, Number of differentially expressed ‘isolated intergenic’ (more than 100 kb from the closest protein-coding-gene) dynamic lncRNAs between adjacent stages during mouse development.

Extended Data Fig. 7 Clustering of dynamic lncRNAs based on developmental trajectories.

Clusters of developmentally dynamic lncRNAs and protein-coding genes across mouse organs (brain = 14,629 genes; cerebellum = 13,166; heart = 12,382; kidney = 14,634; liver = 13,888; ovary = 12,694; testis = 13,749). Grey lines represent individual gene trajectories and solid lines posterior mean trajectories for each cluster. Clusters are arranged by decreasing fraction of lncRNAs. Enriched representative biological processes (Benjamini–Hochberg adjusted P < 0.05, hypergeometric test) are shown for each cluster.

Extended Data Fig. 8 Characteristics of dynamic lncRNAs expressed in different developmental stages.

a, Expression similarity between human and mouse 1:1 orthologous protein-coding genes (n = 16,078), developmentally dynamic (n = 281) and non-dynamic (n = 1,386) lncRNAs across organs/developmental stages. Each point corresponds to the Spearman correlation coefficient of expression between human and mouse orthologues for matching samples. Lines and the 95% confidence interval (shaded regions) correspond to linear model predictions. Spearman correlation coefficients between expression similarity and developmental stage are given for each comparison. b, Expression similarity between dynamic human and mouse orthologous lncRNAs from a, summarized by organ. c, Fraction of conserved (≥80 Ma) dynamic lncRNAs expressed in each mouse organ during development. The colour signifies the focal organ for each comparison. d, Tissue-specificity for mouse lncRNAs with different developmental trajectories. e, Fraction of human lncRNAs with different developmental trajectories among functionally characterized lncRNAs27 (n = 59). f, CRISPRi growth screen hits21 (n = 98). g, Fraction of late-expressed dynamic (n = 2,956) and non-dynamic (n = 25,791) lncRNAs for different age groups and functionally characterized27 human lncRNAs. Box plots are as in Fig. 2. *P < 0.05, **P < 0.01, ***P < 0.001, two-sided Mann–Whitney U-test (bd) or Fisher’s exact test (eg).

Extended Data Fig. 9 Co-expression of dynamic lncRNAs with adjacent protein-coding genes.

a, Normalized density distribution of Pearson correlation coefficients (r) of spatiotemporal gene expression between adjacent paralogous (human = 267; mouse = 263) and non-paralogous (human = 3,359; mouse = 3,382) mRNA–mRNA pairs. b, Number of paralogous (human = 267; mouse = 263) and non-paralogous (human = 3,359; mouse = 3,382) adjacent mRNA–mRNA pairs detected as co-expressed above a range of Pearson’s r cut-offs. c, Relationship between distance and Pearson correlation of expression for lncRNA–mRNA (human = 4,881; mouse = 4,722) and mRNA–mRNA (human = 3,359; mouse = 3,382) pairs. Lines were estimated through LOESS regression and the 95% confidence interval is shown in grey. d, Distribution of Pearson’s r for lncRNA–mRNA and mRNA–mRNA pairs across different distance intervals. Box plots are as in Fig. 2. e, Density distributions of Pearson’s r between a protein-coding gene and its nearest dynamic lncRNA (human = 2,440; mouse = 2,549) and protein-coding gene (human = 1,606; mouse = 1,777) after excluding antisense and divergently transcribed lncRNAs. f, Enriched biological processes among human protein-coding genes with significantly higher expression correlations with their adjacent dynamic lncRNA than with the control protein-coding gene (n = 358; Benjamini–Hochberg adjusted P < 0.01, hypergeometric test; data for mouse are shown in Fig. 4b). In ae, statistical tests are two-sided.

Supplementary information

Reporting Summary

Supplementary Information

This file contains legends for Supplementary Tables 1-16.

Supplementary Tables 1-16

Supplementary Tables 1-16.

Supplementary Data 1

This file contains the lncRNA annotations used in this study in gtf format. Coordinates correspond to the following genome assemblies (human: hg19; rhesus macaque: rheMac3; mouse: mm10; rat: Rnor_5.0; rabbit: OryCun2.0; opossum: monDom5; chicken: Galgal4).

Supplementary Data 2

This file contains expression tables (in RPKM) for lncRNAs, putative new coding genes (denoted with the suffix “.coding”) and Ensembl-annotated transcribed regions that don’t overlap our lncRNAs in the same strand (v75 for human, v77 for all other species).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sarropoulos, I., Marin, R., Cardoso-Moreira, M. et al. Developmental dynamics of lncRNAs across mammalian organs and species. Nature 571, 510–514 (2019). https://doi.org/10.1038/s41586-019-1341-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-019-1341-x

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing