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.

Cell-type specific sequencing of microRNAs from complex animal tissues

Abstract

MicroRNAs (miRNAs) play an essential role in the post-transcriptional regulation of animal development and physiology. However, in vivo studies aimed at linking miRNA function to the biology of distinct cell types within complex tissues remain challenging, partly because in vivo miRNA-profiling methods lack cellular resolution. We report microRNome by methylation-dependent sequencing (mime-seq), an in vivo enzymatic small-RNA-tagging approach that enables high-throughput sequencing of tissue- and cell-type-specific miRNAs in animals. The method combines cell-type-specific 3′-terminal 2′-O-methylation of animal miRNAs by a genetically encoded, plant-specific methyltransferase (HEN1), with chemoselective small-RNA cloning and high-throughput sequencing. We show that mime-seq uncovers the miRNomes of specific cells within Caenorhabditis elegans and Drosophila at unprecedented specificity and sensitivity, enabling miRNA profiling with single-cell resolution in whole animals. Mime-seq overcomes current challenges in cell-type-specific small-RNA profiling and provides novel entry points for understanding the function of miRNAs in spatially restricted physiological settings.

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

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

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

Figure 1: The methyltransferase HEN1 from Arabidopsis thaliana methylates animal miRNAs in vivo.
Figure 2: Periodate-mediated oxidation enables selective cloning of animal miRNAs upon restrictive Ath-HEN1-directed methylation.
Figure 3: Mime-seq reproducibly recovers neuronal miRNAs in C. elegans.
Figure 4: Mime-seq unveils the miRNomes of diverse tissues in C. elegans.
Figure 5: Mime-seq has the sensitivity to retrieve miRNAs from 1/300 cells.

Similar content being viewed by others

Accession codes

Primary accessions

Gene Expression Omnibus

References

  1. Bartel, D.P. MicroRNAs: target recognition and regulatory functions. Cell 136, 215–233 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Aboobaker, A.A., Tomancak, P., Patel, N., Rubin, G.M. & Lai, E.C. Drosophila microRNAs exhibit diverse spatial expression patterns during embryonic development. Proc. Natl. Acad. Sci. USA 102, 18017–18022 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Landgraf, P. et al. A mammalian microRNA expression atlas based on small RNA library sequencing. Cell 129, 1401–1414 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Martinez, N.J. et al. Genome-scale spatiotemporal analysis of Caenorhabditis elegans microRNA promoter activity. Genome Res. 18, 2005–2015 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Park, C.Y. et al. A resource for the conditional ablation of microRNAs in the mouse. Cell Rep. 1, 385–391 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Wienholds, E. et al. MicroRNA expression in zebrafish embryonic development. Science 309, 310–311 (2005).

    Article  CAS  PubMed  Google Scholar 

  7. Chen, W. & Qin, C. General hallmarks of microRNAs in brain evolution and development. RNA Biol. 12, 701–708 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  8. Abruzzi, K., Chen, X., Nagoshi, E., Zadina, A. & Rosbash, M. RNA-seq profiling of small numbers of Drosophila neurons. Methods Enzymol. 551, 369–386 (2015).

    Article  CAS  PubMed  Google Scholar 

  9. Ristori, E. & Nicoli, S. miRNAs expression profile in zebrafish developing vessels. Methods Mol. Biol. 1214, 129–150 (2015).

    Article  PubMed  Google Scholar 

  10. Hackler, L. Jr., Masuda, T., Oliver, V.F., Merbs, S.L. & Zack, D.J. Use of laser capture microdissection for analysis of retinal mRNA/miRNA expression and DNA methylation. Methods Mol. Biol. 884, 289–304 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Herai, R.R. et al. Micro RNA detection in long-term fixed tissue of cortical glutamatergic pyramidal neurons after targeted laser-capture neuroanatomical microdissection. J. Neurosci. Methods 235, 76–82 (2014).

    Article  CAS  PubMed  Google Scholar 

  12. Juzenas, S. et al. A comprehensive, cell specific microRNA catalogue of human peripheral blood. Nucleic Acids Res. 45, 9290–9301 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Kudlow, B.A., Zhang, L. & Han, M. Systematic analysis of tissue-restricted miRISCs reveals a broad role for microRNAs in suppressing basal activity of the C. elegans pathogen response. Mol. Cell 46, 530–541 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Than, M.T., Kudlow, B.A. & Han, M. Functional analysis of neuronal microRNAs in Caenorhabditis elegans dauer formation by combinational genetics and neuronal miRISC immunoprecipitation. PLoS Genet. 9, e1003592 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. He, M. et al. Cell-type-based analysis of microRNA profiles in the mouse brain. Neuron 73, 35–48 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Yu, B. et al. Methylation as a crucial step in plant microRNA biogenesis. Science 307, 932–935 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Ohara, T. et al. The 3′ termini of mouse Piwi-interacting RNAs are 2′-O-methylated. Nat. Struct. Mol. Biol. 14, 349–350 (2007).

    Article  CAS  PubMed  Google Scholar 

  18. Houwing, S. et al. A role for Piwi and piRNAs in germ cell maintenance and transposon silencing in zebrafish. Cell 129, 69–82 (2007).

    Article  CAS  PubMed  Google Scholar 

  19. Yang, Z., Ebright, Y.W., Yu, B. & Chen, X. HEN1 recognizes 21-24 nt small RNA duplexes and deposits a methyl group onto the 2′ OH of the 3′ terminal nucleotide. Nucleic Acids Res. 34, 667–675 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Ghildiyal, M. et al. Endogenous siRNAs derived from transposons and mRNAs in Drosophila somatic cells. Science 320, 1077–1081 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Alefelder, S., Patel, B.K. & Eckstein, F. Incorporation of terminal phosphorothioates into oligonucleotides. Nucleic Acids Res. 26, 4983–4988 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Horwich, M.D. et al. The Drosophila RNA methyltransferase, DmHen1, modifies germline piRNAs and single-stranded siRNAs in RISC. Curr. Biol. 17, 1265–1272 (2007).

    Article  CAS  PubMed  Google Scholar 

  23. Saito, K. et al. Pimet, the Drosophila homolog of HEN1, mediates 2′-O-methylation of Piwi- interacting RNAs at their 3′ ends. Genes Dev. 21, 1603–1608 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Billi, A.C. et al. The Caenorhabditis elegans HEN1 ortholog, HENN-1, methylates and stabilizes select subclasses of germline small RNAs. PLoS Genet. 8, e1002617 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Kamminga, L.M. et al. Differential impact of the HEN1 homolog HENN-1 on 21U and 26G RNAs in the germline of Caenorhabditis elegans. PLoS Genet. 8, e1002702 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Montgomery, T.A. et al. PIWI associated siRNAs and piRNAs specifically require the Caenorhabditis elegans HEN1 ortholog henn-1. PLoS Genet. 8, e1002616 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Park, W., Li, J., Song, R., Messing, J. & Chen, X. CARPEL FACTORY, a Dicer homolog, and HEN1, a novel protein, act in microRNA metabolism in Arabidopsisthaliana. Curr. Biol. 12, 1484–1495 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Frøkjær-Jensen, C., Davis, M.W., Ailion, M. & Jorgensen, E.M. Improved Mos1-mediated transgenesis in C. elegans. Nat. Methods 9, 117–118 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Leuschner, P.J.F., Ameres, S.L., Kueng, S. & Martinez, J. Cleavage of the siRNA passenger strand during RISC assembly in human cells. EMBO Rep. 7, 314–320 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Förstemann, K., Horwich, M.D., Wee, L., Tomari, Y. & Zamore, P.D. Drosophila microRNAs are sorted into functionally distinct argonaute complexes after production by dicer-1. Cell 130, 287–297 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Stefanakis, N., Carrera, I. & Hobert, O. Regulatory logic of pan-neuronal gene expression in C. elegans. Neuron 87, 733–750 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Tursun, B., Cochella, L., Carrera, I. & Hobert, O. A toolkit and robust pipeline for the generation of fosmid-based reporter genes in C. elegans. PLoS One 4, e4625 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Cochella, L. & Hobert, O. Embryonic priming of a miRNA locus predetermines postmitotic neuronal left/right asymmetry in C. elegans. Cell 151, 1229–1242 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Drexel, T., Mahofsky, K., Latham, R., Zimmer, M. & Cochella, L. Neuron type-specific miRNA represses two broadly expressed genes to modulate an avoidance behavior in C. elegans. Genes Dev. 30, 2042–2047 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Johnston, R.J. Jr. & Hobert, O. A microRNA controlling left/right neuronal asymmetry in Caenorhabditis elegans. Nature 426, 845–849 (2003).

    Article  CAS  PubMed  Google Scholar 

  37. Clark, A.M. et al. The microRNA miR-124 controls gene expression in the sensory nervous system of Caenorhabditis elegans. Nucleic Acids Res. 38, 3780–3793 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Hsieh, Y.-W., Chang, C. & Chuang, C.-F. The microRNA mir-71 inhibits calcium signaling by targeting the TIR-1/Sarm1 adaptor protein to control stochastic L/R neuronal asymmetry in C. elegans. PLoS Genet. 8, e1002864 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Simon, D.J. et al. The microRNA miR-1 regulates a MEF-2-dependent retrograde signal at neuromuscular junctions. Cell 133, 903–915 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Sokol, N.S. & Ambros, V. Mesodermally expressed DrosophilamicroRNA-1 is regulated by Twist and is required in muscles during larval growth. Genes Dev. 19, 2343–2354 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Cerro-Herreros, E., Fernandez-Costa, J.M., Sabater-Arcis, M., Llamusi, B. & Artero, R. Derepressing muscleblind expression by miRNA sponges ameliorates myotonic dystrophy-like phenotypes in Drosophila. Sci. Rep. 6, 36230 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Viswanathan, S.R., Daley, G.Q. & Gregory, R.I. Selective blockade of microRNA processing by Lin28. Science 320, 97–100 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Jannot, G., Boisvert, M.-E.L., Banville, I.H. & Simard, M.J. Two molecular features contribute to the Argonaute specificity for the microRNA and RNAi pathways in C. elegans. RNA 14, 829–835 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Steiner, F.A. et al. Structural features of small RNA precursors determine Argonaute loading in Caenorhabditis elegans. Nat. Struct. Mol. Biol. 14, 927–933 (2007).

    Article  CAS  PubMed  Google Scholar 

  45. Tomari, Y., Du, T. & Zamore, P.D. Sorting of Drosophila small silencing RNAs. Cell 130, 299–308 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Okamura, K., Liu, N. & Lai, E.C. Distinct mechanisms for microRNA strand selection by Drosophila Argonautes. Mol. Cell 36, 431–444 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Czech, B. et al. Hierarchical rules for Argonaute loading in Drosophila. Mol. Cell 36, 445–456 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Ghildiyal, M., Xu, J., Seitz, H., Weng, Z. & Zamore, P.D. Sorting of Drosophila small silencing RNAs partitions microRNA* strands into the RNA interference pathway. RNA 16, 43–56 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Corrêa, R.L., Steiner, F.A., Berezikov, E. & Ketting, R.F. MicroRNA-directed siRNA biogenesis in Caenorhabditis elegans. PLoS Genet. 6, e1000903 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Alberti, C. et al. MicroRNome by methylation-dependent sequencing (mime-seq). Protocol Exchange http://dx.doi.org/10.1038/protex.2017.171 (2017).

  51. Brenner, S. The genetics of Caenorhabditis elegans. Genetics 77, 71–94 (1974).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Sulston, J.E. & Hodgkin, J. The Nematode Caenorhabditis elegans 587–606 (Cold Spring Harbor Laboratory Press, 1988).

  53. Redemann, S. et al. Codon adaptation-based control of protein expression in C. elegans. Nat. Methods 8, 250–252 (2011).

    Article  CAS  PubMed  Google Scholar 

  54. Etchberger, J.F. et al. The molecular signature and cis-regulatory architecture of a C. elegans gustatory neuron. Genes Dev. 21, 1653–1674 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Reimão-Pinto, M.M. et al. Uridylation of RNA hairpins by tailor confines the emergence of microRNAs in Drosophila. Mol. Cell 59, 203–216 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Ameres, S.L. et al. Target RNA-directed trimming and tailing of small silencing RNAs. Science 328, 1534–1539 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Gokcezade, J., Sienski, G. & Duchek, P. Efficient CRISPR/Cas9 plasmids for rapid and versatile genome editing in Drosophila. G3 (Bethesda) 4, 2279–2282 (2014).

    Article  Google Scholar 

  58. Markstein, M., Pitsouli, C., Villalta, C., Celniker, S.E. & Perrimon, N. Exploiting position effects and the gypsy retrovirus insulator to engineer precisely expressed transgenes. Nat. Genet. 40, 476–483 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Bryantsev, A.L., Baker, P.W., Lovato, T.L., Jaramillo, M.S. & Cripps, R.M. Differential requirements for Myocyte Enhancer Factor-2 during adult myogenesis in Drosophila. Dev. Biol. 361, 191–207 (2012).

    Article  CAS  PubMed  Google Scholar 

  60. Klein, P. et al. Ret rescues mitochondrial morphology and muscle degeneration of DrosophilaPink1 mutants. EMBO J. 33, 341–355 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Bassett, A.R., Tibbit, C., Ponting, C.P. & Liu, J.-L. Mutagenesis and homologous recombination in Drosophila cell lines using CRISPR/Cas9. Biol. Open 3, 42–49 (2014).

    Article  CAS  PubMed  Google Scholar 

  62. Schindelin, J. et al. Fiji: an open-source platform for biological-image analysis. Nat. Methods 9, 676–682 (2012).

    CAS  PubMed  Google Scholar 

  63. Han, B.W., Hung, J.-H., Weng, Z., Zamore, P.D. & Ameres, S.L. The 3′-to-5′ exoribonuclease Nibbler shapes the 3′ ends of microRNAs bound to Drosophila Argonaute1. Curr. Biol. 21, 1878–1887 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Martin, M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet.journal 17, 10 (2011).

    Article  Google Scholar 

  65. 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 

  66. Di Tommaso, P. et al. Nextflow enables reproducible computational workflows. Nat. Biotechnol. 35, 316–319 (2017).

    Article  CAS  PubMed  Google Scholar 

  67. Micallef, L. & Rodgers, P. eulerAPE: drawing area-proportional 3-Venn diagrams using ellipses. PLoS One 9, e101717 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank Cochella lab members for various fosmid reporters; F. Schnorrer (IBDM, Marseille, France) for muscle-specific Gal4 driver lines; and the CGC (NIH-P40-OD010440) for worm strains. This work was supported by FP7/2007-2013 grants from the European Research Council to L.C. (ERC-StG-337161) and S.L.A. (ERC-StG-338252) and the Austrian Science Fund to L.C. (W-1207-B09 and SFB-F43-23) and S.L.A. (Y-733-B22 START, W-1207-B09 and SFB-F43-22). R.A.M. is a recipient of a DOC fellowship of the Austrian Academy of Sciences at IMBA. Basic research at IMP is supported by Boehringer Ingelheim GmbH.

Author information

Authors and Affiliations

Authors

Contributions

S.L.A. conceived the strategy. L.C. and S.L.A. designed and supervised the study. C.A. performed and analyzed all C. elegans experiments. S.L.A. and R.A.M. did in vitro and cell-culture experiments. I.S. performed all whole fly experiments. K.M. generated fosmid reporters and imaged transgenic worms. T.R.B. and J.W. performed bioinformatics analyses. L.C., C.A. and S.L.A. wrote the paper and prepared the figures.

Corresponding authors

Correspondence to Stefan L Ameres or Luisa Cochella.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Ath-HEN1 methylates animal miRNAs but does not affect miRNA steady-state abundance in C. elegans and Drosophila.

(a) Confirmation of the efficiency of discriminatory 3´ ligation between methylated and non-methylated small RNA upon oxidation. Small RNAs from Drosophila S2 cells were ligated to a pre-adenylated, 3´ sequencing adapter, either before or after oxidation with periodate, and subjected to Northern hybridization analysis using probes against the indicated small RNAs. The endogenously methylated endo-siRNA esi-2.1 served as positive control, its ligation proceeds to completion even after oxidation. The unmethylated miRNA bantam-3p served as negative control, its ligation is abolished upon oxidation. Independent repeats = 1. (b) Expression of FLAG-Ath-HEN1 and endogenous Dme-Hen1 was assessed by Western blot analysis of lysate prepared from wild-type (S2WT) S2 cells or S2 cells depleted of endogenous Dme-Hen1 by CRISPR/Cas9 genome engineering (S2-hen1KO). If indicated, wild-type (WT) or catalytic mutant (CM) Ath-HEN1 was stably expressed. Actin served as loading control. Independent repeats = 2. (c) FLAG-tagged Ath-HEN1 was immunopurified upon expression in Drosophila S2 cells and incubated with 5´ 32P radiolabeled Drosophila let-7 (21 nt) or miR-34 (24 nt) in a single- (left panel) or double-stranded (right panel) conformation. Incubation with buffer served as negative control (Contr.). MicroRNA methylation was assessed by β-elimination and high-resolution polyacrylamide gel electrophoresis. Signal corresponding to methylated (mod.) or unmethylated (unmod.) miRNAs is indicated. Independent repeats = 3. (d) Ath-HEN1 methylates endogenous miRNAs in animal cells. Ath-HEN1 driven by a constitutive Actin5C (Act5C) promoter was expressed in Drosophila S2 cells depleted of endogenous Dme-Hen1 by CRISPR-Cas9 genome editing (S2-hen1KO) and methylation status of endogenous miRNAs or the endogenous siRNA esi-2.1 was assessed by β-elimination and high-resolution Northern hybridization using probes for the indicated miRNAs. Wild-type S2 cells are shown as control. Signal corresponding to methylated (mod.) or unmethylated (unmod.) miRNAs is indicated. 2SrRNA served as control for loading and β-elimination. Independent repeats = 2. (e) Expression of FLAG-MYC-Ath-HEN1 from a UAST or UASP transgenic construct driven by Act5C-Gal4 (Act5C-Gal4; UAST-Ath-HEN1 or Act5C-Gal4; UASP-Ath-HEN1) was assessed by anti-FLAG Western blotting. Lysate was prepared from whole adult 2-5-day old female wild-type flies (w1118) or flies depleted of endogenous Dm-Hen1 by CRISPR/Cas9 genome editing (w1118; hen1m1-6) carrying the indicated transgenes. Actin served as loading control. Independent repeats = 2. (f) Ath-HEN1 methylates animal miRNAs in vivo in Drosophila. Total RNA was extracted from whole adult 2-5-day old female wild-type flies (w1118) or flies depleted of endogenous Dm-Hen1 by CRISPR/Cas9 genome editing (w1118; hen1c4-6). If indicated, flies ubiquitously express Ath-HEN1 from a UAST or UASP transgenic construct driven by Act5C-Gal4 (Act5C-Gal4; UAST-Ath-HEN1 or Act5C-Gal4; UASP-Ath-HEN1). Total RNA was subjected to b-elimination and Northern hybridization using probes against the indicated miRNAs. 2SrRNA served as control for loading and b-elimination. Independent repeats = 2. (g) Expression of MYC-Ath-HEN1 under the rps-5prom in L1-stage worms was assessed by anti-MYC Western blotting. Two independent isolates from the transgene integration procedure were tested and further used for sequencing experiments. N2 animals that do not express MYC-Ath-HEN1 served as a negative control. Gamma tubulin staining was used as loading control. Independent repeats = 2. (h) Ath-HEN1 methylates animal miRNAs in vivo in C. elegans. Total RNA extracted from wild type or transgenic rps-5prom::Ath-HEN1 L1-stage worms was subjected to b-elimination and Northern hybridization using probes against the indicated miRNAs. tRNAGly served as control for loading and b-elimination. Independent repeats = 2. (i, j) Methylation does not affect miRNA relative abundance in S2 cells or worms. Comparison of the relative abundances of all expressed miRNAs in S2-hen1KO cells stably expressing wild type (WT) or catalytic-mutant (CM) Ath-HEN1 (i) or in wt, non-transgenic L1 worms versus animals expressing rps-5prom::Ath-HEN1 (j). Spearman correlation coefficients and associated p-Values are indicated. Independent repeats = 1 for i, 2 for j. For full scans of all blots in this figure see Suppl. Fig. 9 and 10.

Supplementary Figure 2 Signal and background constraints for mime-seq.

(a) Schematic of the domain structure of plant and animal HEN1. The Arabidopsis enzyme has two dsRNA binding domains that define its specificity for short duplexes. The animal methyltransferases, despite similar enzymatic activity, do not bind or act on dsRNA and in fact do not seem to share a common ancestor with Ath-HEN1. (b) Pie charts showing the proportion of miRNAs and other small RNAs sequenced before and after oxidation of RNA obtained from wild-type, non-trangenic C. elegans L1 larvae. While miRNAs are strongly depleted, the libraries generated after oxidation still contained 5-10% of reads corresponding to miRNAs, mostly representing 26 miRNAs (Suppl. Table 6). Given their precursor structure and association with other Argonautes, and the fact that they are depleted upon oxidation in henn-1(0) background (Suppl. Tables 3, 4, 5), these are endogenous targets of Cel-HENN-1 (see discussion and Suppl. Table 6). (c) Duplex structures of two of the miRNAs that are endogenously methylated by Cel-HENN-1. (d) Comparison of the relative abundances of all expressed mature miRNAs in Drosophila S2 cells depleted of endogenous Dme-Hen1 by CRISPR-Cas9 genome editing (S2-hen1KO) and expressing Ath-HEN1, before and after oxidation treatment. Spearman correlation coefficient and associated p-Value are shown. Independent repeats = 1. (e) Depletion of endogenous RNA methyltransferase HENN-1 does not affect miRNA relative abundance in worms. Comparison of the relative abundances of all confidently detected (>10 cpm in the average of all untreated samples) mature miR strands in wt, non-transgenic L1 worms versus animals deficient for endogenous HENN-1. Spearman correlation coefficient and associated p-Value are shown. Independent repeats = 1. (f) Comparison of the relative abundances of all confidently detected (>10 cpm in the average of all untreated samples) mature miR strands in henn-1(0) worm L1s expressing rps-5prom::Ath-HEN1, before and after oxidation treatment. Spearman correlation coefficient and associated p-Value are shown. In blue are miRNAs that are significantly depleted after oxidation (p-Value <0.001) and were not detected as enriched in any other profiling experiment in this study (see text and Supplementary Table 2). Independent repeats = 2. (g) Volcano plot of the data in (f) showing the respective p-Values for all fold-changes. Independent repeats = 2.

Supplementary Figure 3 Periodate-treatment enables selective cloning of animal miR and miR* species upon restrictive Ath-HEN1-directed methylation.

(a) Volcano plots show fold-change and associated p-Values (determined by DESeq2, see methods) for the indicated number (n) of abundantly expressed fly (red) or mouse (black) miRNAs in small RNA libraries generated as described in Fig. 2a from the indicated input ratios. True-positive miRNA-recovery rate (TPR) was determined from the number of abundantly expressed fly miRNAs detected in the respective libraries above cutoff (log2 fold-change >1; p-Value<10−3) relative to the 64 fly miRNAs recovered at a read depth of >10 cpm (na=64) in untreated 1:1 input libraries. False-positive miRNA-recovery rate (FPR) was determined from the number of mouse miRNAs above cutoff (log2 fold-change >1; p-Value<10−3). (b) As described in Fig. 2a, total RNA derived from Drosophila melanogaster (dme) S2hen1 cells expressing Ath-HEN1 was mixed with total RNA from mouse (mmu) embryonic stem cells at the indicated ratios (input ratio), followed by small RNA cloning before (-ox) and after (+ox) oxidation by periodate-treatment. This experiment was conducted in triplicate. Here we focus on the fraction of miR* strands mapping to annotated mouse (mmu miR) or fly miRNAs (dme miR). The number of abundantly expressed fly miR*s (>100 cpm in untreated 1:1 input ratio libraries) still detected after the respective dilutions at a read depth of >100 cpm is indicated in parenthesis. (c) Volcano plots show fold-change and associated p-Values (determined by DESeq2, see methods) for the indicated number (n) of miR* strands of abundantly expressed fly (blue) or mouse (black) miRNAs in small RNA libraries generated as described in Fig. 2a from the indicated input ratios. True-positive miRNA-recovery rate (TPR) was determined from the number of abundantly expressed fly miR* species detected in the respective libraries above cutoff (log2fold-change >1; p-Value<10−3) relative to the expected 64 fly miR* partner strands of miRs recovered at a read depth of >10 cpm (na=64) in untreated 1:1 input libraries. False-positive miRNA-recovery rate (FPR) was determined from the number of mouse miR* species above cutoff (log2 fold-change >1; p-Value<10−3).

Supplementary Figure 4 Ath-HEN1-directed methylation followed by oxidation-dependent cloning preserves the relative quantity of miR and miR* species.

Relative abundance of the indicated number of fly miR (red) and miR* (blue) species in the indicated small RNA libraries generated as described in Fig. 2a. Only miR and miR* strands that were significantly enriched (log2 fold-change > 1; p-Value<10−3) upon oxidation in the respective input dilution, as determined in Fig. 2b and Suppl. Fig. 3a and c, are shown. Spearman correlation coefficient (rs) and associated p-Value were determined separately for miR (red) and miR* (blue) species. Mean±stdev of three independent experiments are shown.

Supplementary Figure 5 Pan-neuronal mime-seq is highly reproducible and robust to Ath-HEN1 dose.

(a) Western blot analysis of the expression of MYC-Ath-HEN1 driven under three different pan-neuronal promoters in C. elegans (biological duplicates are shown). Tubulin detection was used for loading control. Non-transgenic, N2 animals were used for negative control. For full scans of the blots see Suppl. Fig. 10. Independent repeats = 2. (b, d, f) Comparison of the relative abundances of all confidently detected (>10 cpm in the average of all untreated samples) mature miR strands in henn-1(0) worm L1s expressing Ath-HEN1 under three different pan-neuronal drivers, before and after oxidation treatment, shows that some miRNAs are enriched while others are depleted by the oxidation treatment. (c, e, g) Comparison of the relative abundances of all confidently detected (>10 cpm in the average of all untreated samples) miRNAs for each pair of biological replicates, for both treated and untreated samples, in all three pan-neuronal experiments. Spearman correlation coefficients and their significance are shown. All experiments in this figure were performed on two biologically independent samples.

Supplementary Figure 6 Methylated miR* enrichment supports neuronal miRNA assignments.

For each pan-neuronal mime-seq experiment, we analyzed the change in relative abundance after periodate-mediated oxidation of all miR strands detected above an average of 10 cpm in all the untreated samples, as well as of all miR* strands corresponding to the detected miRs, regardless of their abundance. The correlations between the enrichment of both species is plotted for each experiment and the Pearson correlation coefficients (rp) and associated two-tailed p-Values (calculated in Prism7) are shown. The Venn diagrams display how many miR strands were significantly enriched (log2 fold-change > 1; p-Value<10−3) upon oxidation in the respective experiment and from their respective miR* strands, how many were also enriched (log2 fold-change > 1; p-Value<0.0013) and expressed above 10 cpm. The majority of miR* strands that do not support miR enrichment are expressed below 10 cpm in the unoxidized samples. Experiments in this figure were performed on two biologically independent samples.

Supplementary Figure 7 Mime-seq reveals the miRNomes of C. elegans tissues of diverse composition and size.

(a) Western blot analysis of the expression of MYC-Ath-HEN1 driven under different tissue-specific promoters in C. elegans (biological duplicates are shown). Tubulin detection was used for loading control. Non-transgenic, N2 animals were used for negative control. For full scans of the blots see Suppl. Fig. 10. Independent repeats = 2 for elt-2 and myo-2, 1 for myo-3. (b, d, f) Comparison of the relative abundances of all confidently detected (>10 cpm in the average of all untreated samples) mature miR strands in henn-1(0) worm L1s expressing Ath-HEN1 under three different tissue-specific drivers, before and after oxidation treatment, shows that some miRNAs are enriched while others are depleted by the oxidation treatment. (c, e, g) Comparison of the relative abundances of all confidently detected (>10 cpm in the average of all untreated samples) mature miR strands for each pair of biological replicates, for both treated and untreated samples, in all three tissue-specific experiments. Spearman correlation coefficients and their significance are shown. All experiments in this figure were performed on two biologically independent samples.

Supplementary Figure 8 Mime-seq can be performed in wild-type background in worms and flies.

(a) Expression of Ath-HEN1 in the ASE neurons was monitored by joint expression of mCherry-H2B using a T2A peptide. At least 10 animals were analyzed from each of two independent transgenic lines. Scale bar = 10 μm. (b) Comparison of the relative abundances of all confidently detected (>10 cpm in the average of all untreated samples) mature miR strands in henn-1(0) worm L1s expressing Ath-HEN1 under the ASE-specific promoter, before and after oxidation treatment. Reproducibility of the two biological replicates, for both treated and untreated conditions, are shown. Spearman correlation coefficients and their significance are shown. (c) Same as in (b) but with animals carrying wild-type (wt) endogenous henn-1. Experiments in this figure were performed on two biologically independent samples. (d) Venn diagram shows the overlap between the significantly enriched miRNAs (log2 Fold Change>1, p-Value<0.001 for henn-1(tm4477) or <0.01 for wt) in each of the two backgrounds. (e) Muscle mime-seq results from C. elegans from both henn-1(0) and wild-type backgrounds are shown as volcano plots. These show miRNAs enriched or depleted following oxidation treatment, and their associated p-Values. The wt enrichments shown are after subtraction of the background enrichment in non-transgenic, wt animals to account for endogenous Cel-HENN-1-mediated methylation. The Venn diagram shows the overlap between the significantly enriched miRNAs (log2 fold-change>1, p-Value<0.001 for henn-1(0) or <0.01 for wt) in each of the two backgrounds. Experiments in this figure were performed on two biologically independent samples. (f) Muscle mime-seq results from Drosophila using all muscle or flight-muscle-specific drivers are shown as volcano plots. These show miRNAs enriched or depleted following oxidation treatment, and their associated p-Values. The Venn diagram shows the overlap between the miRNAs enriched >2-fold in each of the two experiments. Experiments in this figure were performed in single replicates.

Supplementary Figure 9 Source images for Drosophila experiments.

Full scans of Northern hybridization and Western blots associated with (a) Suppl. Fig. 1c, (b) Suppl. Fig. 1a, (c) Suppl. Fig. 1b, (d) Suppl. Fig. 1d (black and red dotted box), (e) Suppl. Fig. 1e, (f) Fig. 1b (black dotted box) and Suppl. Fig. 1f (black and red dotted box). Marker lane is indicated (M).

Supplementary Figure 10 Source images for C. elegans experiments.

Full scans of Northern hybridization and Western blots associated with (a) Fig. 1c (black dotted box) and Suppl. Fig. 1h (red dotted box), (b) Suppl. Fig. 5a (red dotted box), (c) Suppl. Fig. 7a (long red dotted box) and Suppl. Fig. 1g (short red dotted box). Marker lane is indicated (M).

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–10 and Supplementary Tables 7–8 (PDF 2554 kb)

Life Sciences Reporting Summary (PDF 166 kb)

Supplementary Protocol

MicroRNome by methylation-dependent sequencing (mime-seq) (PDF 6917 kb)

Supplementary Table 1

Synthetic dilution experiment to determine thesensitivity of mime-seq (XLSX 1373 kb)

Supplementary Table 2

A few miRNAs are not efficiently methylated by AthHEN1 (XLSX 58 kb)

Supplementary Table 3

Pan-neuronal miRNAs from C. elegans revealed by mime-seq and comparison with transcriptional reporters (XLSX 111 kb)

Supplementary Table 4

Tissue-specific miRNAs from C. elegans revealed by mime-seq (XLSX 112 kb)

Supplementary Table 5

The miRNome of a single neuron class in C. elegans revealed by mime-seq in wild-type and henn-1(0) backgrounds (XLSX 164 kb)

Supplementary Table 6

miRNAs endogenously methylated by Cel-HENN-1 (XLSX 68 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Alberti, C., Manzenreither, R., Sowemimo, I. et al. Cell-type specific sequencing of microRNAs from complex animal tissues. Nat Methods 15, 283–289 (2018). https://doi.org/10.1038/nmeth.4610

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nmeth.4610

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