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:

Short poly(A) tails are protected from deadenylation by the LARP1–PABP complex

Subjects

Abstract

Deadenylation generally constitutes the first and pivotal step in eukaryotic messenger RNA decay. Despite its importance in posttranscriptional regulations, the kinetics of deadenylation and its regulation remain largely unexplored. Here we identify La ribonucleoprotein 1, translational regulator (LARP1) as a general decelerator of deadenylation, which acts mainly in the 30–60-nucleotide (nt) poly(A) length window. We measured the steady-state and pulse-chased distribution of poly(A)-tail length, and found that deadenylation slows down in the 30–60-nt range. LARP1 associates preferentially with short tails and its depletion results in accelerated deadenylation specifically in the 30–60-nt range. Consistently, LARP1 knockdown leads to a global reduction of messenger RNA abundance. LARP1 interferes with the CCR4–NOT-mediated deadenylation in vitro by forming a ternary complex with poly(A)-binding protein (PABP) and poly(A). Together, our work reveals a dynamic nature of deadenylation kinetics and a role of LARP1 as a poly(A) length-specific barricade that creates a threshold for deadenylation.

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: Steady-state and dynamic poly(A) length distributions.
Fig. 2: Effective ranges of LARPs on the poly(A) tail.
Fig. 3: Loss of LARP1 accelerates deadenylation in cells.
Fig. 4: LARP1 protects mRNAs with 30–60 nt of poly(A) tails globally.
Fig. 5: LARP1 directly blocks the deadenylation by CCR4–CAF1 in vitro.
Fig. 6: Model for decelerated deadenylation by LARP1 barricade on short poly(A) tails.

Similar content being viewed by others

Data availability

Hire-PAT, BPA, SEC, double-filter binding assay and mTAIL-seq Tailseeker-processed data have been archived in Zenodo (https://doi.org/10.5281/zenodo.6462786). RNA co-IP and knockdown mTAIL-seq raw data are available in separate archives in Zenodo (https://doi.org/10.5281/zenodo.6479127 for RNA co-IP, and https://doi.org/10.5281/zenodo.5263366 for knockdown experiments). The mass spectrometry proteomics data are available via ProteomeXchange (PXD028326). Source data are provided with this paper.

Code availability

Custom codes are publicly available online (https://github.com/johapark/polya).

References

  1. Goldstrohm, A. C. & Wickens, M. Multifunctional deadenylase complexes diversify mRNA control. Nat. Rev. Mol. Cell Biol. 9, 337–344 (2008).

    Article  CAS  PubMed  Google Scholar 

  2. Roy, B. & Jacobson, A. The intimate relationships of mRNA decay and translation. Trends Genet. 29, 691–699 (2013).

    Article  CAS  PubMed  Google Scholar 

  3. Sawicki, S. G., Jelinek, W. & Darnell, J. E. 3′-Terminal addition to HeLa cell nuclear and cytoplasmic poly(A). J. Mol. Biol. 113, 219–235 (1977).

    Article  CAS  PubMed  Google Scholar 

  4. Proudfoot, N. J., Furger, A. & Dye, M. J. Integrating mRNA processing with transcription. Cell 108, 501–512 (2002).

    Article  CAS  PubMed  Google Scholar 

  5. Colgan, D. F. & Manley, J. L. Mechanism and regulation of mRNA polyadenylation. Genes Dev. 11, 2755–2766 (1997).

    Article  CAS  PubMed  Google Scholar 

  6. Kumar, A., Clerici, M., Muckenfuss, L. M., Passmore, L. A. & Jinek, M. Mechanistic insights into mRNA 3′-end processing. Curr. Opin. Struct. Biol. 59, 143–150 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Nicholson, A. L. & Pasquinelli, A. E. Tales of detailed poly(A) tails. Trends Cell Biol. 29, 191–200 (2019).

    Article  CAS  PubMed  Google Scholar 

  8. Yi, H. et al. PABP cooperates with the CCR4-NOT complex to promote mRNA deadenylation and block precocious decay. Mol. Cell 70, 1081–1088.e5 (2018).

    Article  CAS  PubMed  Google Scholar 

  9. Webster, M. W. et al. mRNA deadenylation is coupled to translation rates by the differential activities of Ccr4-Not nucleases. Mol. Cell 70, 1089–1100.e8 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Caponigro, G. & Parker, R. Multiple functions for the poly(A)-binding protein in mRNA decapping and deadenylation in yeast. Genes Dev. 9, 2421–2432 (1995).

    Article  CAS  PubMed  Google Scholar 

  11. Coller, J. M., Gray, N. K. & Wickens, M. P. mRNA stabilization by poly(A) binding protein is independent of poly(A) and requires translation. Genes Dev. 12, 3226–3235 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Sachs, A. B. & Davis, R. W. The poly(A) binding protein is required for poly(A) shortening and 60S ribosomal subunit-dependent translation initiation. Cell 58, 857–867 (1989).

    Article  CAS  PubMed  Google Scholar 

  13. Weill, L., Belloc, E., Bava, F.-A. & Méndez, R. Translational control by changes in poly(A) tail length: recycling mRNAs. Nat. Struct. Mol. Biol. 19, 577–585 (2012).

    Article  CAS  PubMed  Google Scholar 

  14. Lim, J. et al. Uridylation by TUT4 and TUT7 marks mRNA for degradation. Cell 159, 1365–1376 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Garneau, N. L., Wilusz, J. & Wilusz, C. J. The highways and byways of mRNA decay. Nat. Rev. Mol. Cell Biol. 8, 113–126 (2007).

    Article  CAS  PubMed  Google Scholar 

  16. Mugridge, J. S., Coller, J. & Gross, J. D. Structural and molecular mechanisms for the control of eukaryotic 5′–3′ mRNA decay. Nat. Struct. Mol. Biol. 25, 1077–1085 (2018).

    Article  CAS  PubMed  Google Scholar 

  17. Chen, C.-Y. A. & Shyu, A.-B. Mechanisms of deadenylation-dependent decay. Wiley Interdiscip. Rev. RNA 2, 167–183 (2011).

    Article  CAS  PubMed  Google Scholar 

  18. Yamashita, A. et al. Concerted action of poly(A) nucleases and decapping enzyme in mammalian mRNA turnover. Nat. Struct. Mol. Biol. 12, 1054–1063 (2005).

    Article  CAS  PubMed  Google Scholar 

  19. Brown, C. E. & Sachs, A. B. Poly(A) tail length control in Saccharomyces cerevisiae occurs by message-specific deadenylation. Mol. Cell. Biol. 18, 6548–6559 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Tucker, M. et al. The transcription factor associated Ccr4 and Caf1 proteins are components of the major cytoplasmic mRNA deadenylase in Saccharomyces cerevisiae. Cell 104, 377–386 (2001).

    Article  CAS  PubMed  Google Scholar 

  21. Parker, R. RNA degradation in Saccharomyces cerevisae. Genetics 191, 671–702 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Niinuma, S., Fukaya, T. & Tomari, Y. CCR4 and CAF1 deadenylases have an intrinsic activity to remove the post-poly(A) sequence. RNA 22, 1550–1559 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  23. Fabian, M. R. et al. Structural basis for the recruitment of the human CCR4-NOT deadenylase complex by tristetraprolin. Nat. Struct. Mol. Biol. 20, 735–739 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Fabian, M. R. et al. miRNA-mediated deadenylation is orchestrated by GW182 through two conserved motifs that interact with CCR4-NOT. Nat. Struct. Mol. Biol. 18, 1211–1217 (2011).

    Article  CAS  PubMed  Google Scholar 

  25. Van Etten, J. et al. Human Pumilio proteins recruit multiple deadenylases to efficiently repress messenger RNAs. J. Biol. Chem. 287, 36370–36383 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  26. Du, H. et al. YTHDF2 destabilizes m6A-containing RNA through direct recruitment of the CCR4-NOT deadenylase complex. Nat. Commun. 7, 12626 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Mauxion, F., Faux, C. & Séraphin, B. The BTG2 protein is a general activator of mRNA deadenylation. EMBO J. 27, 1039–1048 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Webster, M. W., Stowell, J. A. & Passmore, L. A. RNA-binding proteins distinguish between similar sequence motifs to promote targeted deadenylation by Ccr4-Not. eLife 8, e40670 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  29. Subtelny, A. O., Eichhorn, S. W., Chen, G. R., Sive, H. & Bartel, D. P. Poly(A)-tail profiling reveals an embryonic switch in translational control. Nature 508, 66–71 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Chang, H., Lim, J., Ha, M. & Kim, V. N. TAIL-seq: genome-wide determination of poly(A) tail length and 3′ end modifications. Mol. Cell 53, 1044–1052 (2014).

    Article  CAS  PubMed  Google Scholar 

  31. Lim, J., Lee, M., Son, A., Chang, H. & Kim, V. N. mTAIL-seq reveals dynamic poly(A) tail regulation in oocyte-to-embryo development. Genes Dev. 30, 1671–1682 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Legnini, I., Alles, J., Karaiskos, N., Ayoub, S. & Rajewsky, N. FLAM-seq: full-length mRNA sequencing reveals principles of poly(A) tail length control. Nat. Methods 16, 879–886 (2019).

    Article  CAS  PubMed  Google Scholar 

  33. Tudek, A. et al. Global view on the metabolism of RNA poly(A) tails in yeast Saccharomyces cerevisiae. Nat. Commun. 12, 4951 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Eisen, T. J. et al. The dynamics of cytoplasmic mRNA metabolism. Mol. Cell 77, 786–799.e10 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Lima, S. A. et al. Short poly(A) tails are a conserved feature of highly expressed genes. Nat. Struct. Mol. Biol. 24, 1057–1063 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Baer, B. W. & Kornberg, R. D. The protein responsible for the repeating structure of cytoplasmic poly(A)-ribonucleoprotein. J. Cell Biol. 96, 717–721 (1983).

    Article  CAS  PubMed  Google Scholar 

  37. Long, Y., Jia, J., Mo, W., Jin, X. & Zhai, J. FLEP-seq: simultaneous detection of RNA polymerase II position, splicing status, polyadenylation site and poly(A) tail length at genome-wide scale by single-molecule nascent RNA sequencing. Nat. Protoc. 16, 4355–4381 (2021).

    Article  CAS  PubMed  Google Scholar 

  38. Workman, R. E. et al. Nanopore native RNA sequencing of a human poly (A) transcriptome. Nat. Methods 16, 1297–1305 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Halstead, J. M. et al. Translation. An RNA biosensor for imaging the first round of translation from single cells to living animals. Science 347, 1367–1671 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Mateju, D. et al. Single-molecule imaging reveals translation of mRNAs localized to stress granules. Cell 183, 1801–1812.e13 (2020).

    Article  CAS  PubMed  Google Scholar 

  41. Bazzini, A. A., Lee, M. T. & Giraldez, A. J. Ribosome profiling shows that miR-430 reduces translation before causing mRNA decay in zebrafish. Science 336, 233–237 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Chang, H. et al. Terminal uridylyltransferases execute programmed clearance of maternal transcriptome in vertebrate embryos. Mol. Cell 70, 72–82.e7 (2018).

    Article  CAS  PubMed  Google Scholar 

  43. Xiang, K. & Bartel, D. P. The molecular basis of coupling between poly(A)-tail length and translational efficiency. eLife 10, e66493 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Blagden, S. P. et al. Drosophila Larp associates with poly(A)-binding protein and is required for male fertility and syncytial embryo development. Dev. Biol. 334, 186–197 (2009).

    Article  CAS  PubMed  Google Scholar 

  45. Burrows, C. et al. The RNA binding protein Larp1 regulates cell division, apoptosis and cell migration. Nucleic Acids Res. 38, 5542–5553 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Yang, R. et al. La-related protein 4 binds poly(A), interacts with the poly(A)-binding protein MLLE domain via a variant PAM2w motif, and can promote mRNA stability. Mol. Cell. Biol. 31, 542–556 (2011).

    Article  CAS  PubMed  Google Scholar 

  47. Miroci, H. et al. Makorin ring zinc finger protein 1 (MKRN1), a novel poly(A)-binding protein-interacting protein, stimulates translation in nerve cells. J. Biol. Chem. 287, 1322–1334 (2012).

    Article  CAS  PubMed  Google Scholar 

  48. Hildebrandt, A. et al. The RNA-binding ubiquitin ligase MKRN1 functions in ribosome-associated quality control of poly(A) translation. Genome Biol. 20, 216 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  49. Aoki, K. et al. LARP1 specifically recognizes the 3′ terminus of poly(A) mRNA. FEBS Lett. 587, 2173–2178 (2013).

    Article  CAS  PubMed  Google Scholar 

  50. Mattijssen, S. et al. mRNA codon-tRNA match contributes to LARP4 activity for ribosomal protein mRNA poly(A) tail length protection. eLife 6, e28889 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  51. Cruz-Gallardo, I. et al. LARP4A recognizes polyA RNA via a novel binding mechanism mediated by disordered regions and involving the PAM2w motif, revealing interplay between PABP, LARP4A and mRNA. Nucleic Acids Res. 47, 4272–4291 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Merret, R. et al. The association of a La module with the PABP-interacting motif PAM2 is a recurrent evolutionary process that led to the neofunctionalization of La-related proteins. RNA 19, 36–50 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Xie, J., Kozlov, G. & Gehring, K. The ‘tale’ of poly(A) binding protein: the MLLE domain and PAM2-containing proteins. Biochim. Biophys. Acta 1839, 1062–1068 (2014).

    Article  CAS  PubMed  Google Scholar 

  54. Maraia, R. J., Mattijssen, S., Cruz-Gallardo, I. & Conte, M. R. The La and related RNA-binding proteins (LARPs): structures, functions, and evolving perspectives. Wiley Interdiscip. Rev. RNA https://doi.org/10.1002/wrna.1430 (2017).

  55. Mattijssen, S., Iben, J. R., Li, T., Coon, S. L. & Maraia, R. J. Single molecule poly(A) tail-seq shows LARP4 opposes deadenylation throughout mRNA lifespan with most impact on short tails. eLife 9, e59186 (2020).

  56. Mattijssen, S. et al. The isolated La-module of LARP1 mediates 3′ poly(A) protection and mRNA stabilization, dependent on its intrinsic PAM2 binding to PABPC1. RNA Biol. https://doi.org/10.1080/15476286.2020.1860376 (2020).

  57. Fonseca, B. D. et al. La-related protein 1 (LARP1) represses terminal oligopyrimidine (TOP) mRNA translation downstream of mTOR complex 1 (mTORC1). J. Biol. Chem. 290, 15996–16020 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Gentilella, A. et al. Autogenous control of 5′TOP mRNA stability by 40S ribosomes. Mol. Cell 67, 55–70.e4 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Lahr, R. M. et al. La-related protein 1 (LARP1) binds the mRNA cap, blocking eIF4F assembly on TOP mRNAs. eLife 6, e24146 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  60. Hong, S. et al. LARP1 functions as a molecular switch for mTORC1-mediated translation of an essential class of mRNAs. eLife 6, e25237 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  61. Ogami, K. et al. mTOR- and LARP1-dependent regulation of TOP mRNA poly(A) tail and ribosome loading. Cell Rep. 41, 111548 (2022)

  62. Lahr, R. M. et al. The La-related protein 1-specific domain repurposes HEAT-like repeats to directly bind a 5′TOP sequence. Nucleic Acids Res. 43, 8077–8088 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Lim, J. et al. Mixed tailing by TENT4A and TENT4B shields mRNA from rapid deadenylation. Science 361, 701–704 (2018).

    Article  CAS  PubMed  Google Scholar 

  64. Al-Ashtal, H. A., Rubottom, C. M., Leeper, T. C. & Berman, A. J. The LARP1 La-module recognizes both ends of TOP mRNAs. RNA Biol. 18, 248–258 (2021).

  65. Wilbertz, J. H. et al. Single-molecule imaging of mRNA localization and regulation during the integrated stress response. Mol. Cell 73, 946–958.e7 (2019).

    Article  CAS  PubMed  Google Scholar 

  66. Weidenfeld, I. et al. Inducible expression of coding and inhibitory RNAs from retargetable genomic loci. Nucleic Acids Res. 37, e50 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  67. Schaft, J., Ashery-Padan, R., van der Hoeven, F., Gruss, P. & Francis Stewart, A. Efficient FLP recombination in mouse ES cells and oocytes. Genesis 31, 6–10 (2001).

    Article  CAS  PubMed  Google Scholar 

  68. Perez-Riverol, Y. et al. The PRIDE database and related tools and resources in 2019: improving support for quantification data. Nucleic Acids Res. 47, D442–D450 (2019).

    Article  CAS  PubMed  Google Scholar 

  69. Nguyen, T. A. et al. Functional anatomy of the human microprocessor. Cell 161, 1374–1387 (2015).

    Article  CAS  PubMed  Google Scholar 

  70. Wee, L. M., Flores-Jasso, C. F., Salomon, W. E. & Zamore, P. D. Argonaute divides its RNA guide into domains with distinct functions and RNA-binding properties. Cell 151, 1055–1067 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Iwakawa, H.-O. & Tomari, Y. Molecular insights into microRNA-mediated translational repression in plants. Mol. Cell 52, 591–601 (2013).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank our laboratory members for their support, especially S. Lee for Tet-On inducible TOP TRICK reporter cell line construction and Y. Jung for insights into the inference of deadenylation rates. We also thank J. Kim and J.-S. Kim for mass spectrometry and K. Schönig and J. A. Chao for kind donation of the master cell line and materials used in our reporter cell line construction. This research was supported by the Institute for Basic Science from the Ministry of Science, ICT and Future Planning of Korea (grant no. IBS-R008-D01 to J.P., M.K., H.Y., K.B., Y.C., Y.-s.L., J.L. and V.N.K.); BK21 research fellowships from the Ministry of Education of Korea (to M.K. and Y.C.); and a grant funded by the Kwanjeong Educational Foundation (to M.K.).

Author information

Authors and Affiliations

Authors

Contributions

J.P., H.Y., K.B. and V.N.K. conceived the project. J.P., M.K., H.Y. and Y.C. conducted the cell- and sequencing-related experiments. J.P. performed the bioinformatics analysis. K.B. purified the recombinant proteins and performed the in vitro assay. Y.-s.L. analyzed the PAL-seq dataset. J.L. generated the mTAIL-seq library for the decay machinery knockdown samples. J.P., M.K., H.Y., K.B. and V.N.K. wrote the paper.

Corresponding author

Correspondence to V. Narry Kim.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Structural & Molecular Biology thanks Jixian Zhai and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Peer reviewer reports are available. Primary Handling Editor: Sara Osman, in collaboration with the Nature Structural & Molecular Biology team.

Additional information

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

Extended data

Extended Data Fig. 1 Simulation to test the relationship between the relative deadenylation rate and the local poly(A) density at a steady state.

a. The production (π) of mRNAs and their poly(A) tails, which includes transcription and 3′ end processing, was modeled with a Gaussian distribution. Deadenylation (δ) and decay (φ) were modeled with logistic functions of the poly(A) length. Deadenylation rates were varied with three different parameters (i)–(iii): uniform, decreasing, and increasing along the poly(A) length (long-to-short), respectively. The number of transcripts over the cycles is shown on the third column to ensure that the system reaches a steady state. The light green-to-black lines in the plots in the right-most column represent the poly(A) length distribution over the cycles. b. Another computer simulation with a different production model. The initial distribution of poly(A) length in the production (π) step is modeled with the negative binomial distribution to account for a more realistic scenario based on the experimental observation34.

Extended Data Fig. 2 Validation experiments after knockdown of decay machinery.

a. Schematic of Tet-On inducible TOP TRICK reporter. Reporter mRNAs are only transcribed in the presence of doxycycline. SV40 poly(A) signal on its 3′ UTR was targeted for measuring poly(A)-tail length by Hire-PAT assay. b. Global poly(A) length distributions measured by mTAIL-seq upon depletion of decay factors. c. Poly(A) length distributions of pulse-expressed reporter mRNAs measured by Hire-PAT assay upon depletion of decay factors at various time points post-wash. d. Steady-state poly(A) length distribution of endogenous mRNAs measured by Hire-PAT assay upon depletion of decay factors.

Extended Data Fig. 3 LARPs as PABPC1 interacting partners and validation experiments evaluating LARPs.

a. Log2 fold of LFQ intensities for anti-PABPC1 over anti-Myc co-immunoprecipitates. The proteins with > 4 folds in both replicates are highlighted and shown in a separate panel. The direct target protein PABPC1 is colored in red, and the co-purified cytosolic RNA-binding proteins are colored in blue. b. The domain structures of LARP1, 4, and 4B. PABP-interacting domains are colored in red. c. Western blotting of cytosol RIP-BPA samples (Exp. 3) in Fig. 2a, b. d. Western blotting of knockdown-BPA samples (Exp. 2) in Fig. 2c, d. e. Poly(A) length distribution of the LARP4, LARP4B single knockdowns and LARP4/4B double knockdown RNAs measured by BPA. f. Western blotting of translational machinery including PABPC1 upon LARP1 knockdown (n = 4).

Source data

Extended Data Fig. 4 Validation experiments for LARP1’s effect on TOP and non-TOP mRNAs.

a. Schematics of the TOP and non-TOP reporter mRNAs. Five cytosines including cytidine on +1 position were replaced with purines58. CTE on their 3′ UTR was targeted for measuring their expression level by RT-qPCR. b. Comparison of LARP1 association with TOP or non-TOP reporter mRNAs. Left: the experimental scheme. Doxycycline was treated for 72 hrs to induce expressions of TOP or non-TOP reporters and reach a steady-state. Center: RT-qPCR of LARP1 co-immunoprecipitated RNAs relative to U1 snRNA (n = 1). The TOP and non-TOP reporter mRNAs were targeted by CTE in their 3′ UTR. GAPDH mRNA was used as a negative control for LARP1 association. Right: Western blotting of the input and LARP1-immunoprecipitated samples. Normal rabbit IgG (NRG) was used as a negative control for immunoprecipitation. c. Poly(A) length distributions of pulse-expressed TOP (left) and non-TOP (right) reporter mRNAs measured by Hire-PAT assay upon depletion of LARP1 using another siRNA (siLARP1 #2). The time indicates the minutes post-wash. d. Steady-state poly(A) length distribution of endogenous mRNAs measured by Hire-PAT assay upon depletion of LARP1 using another siRNA (siLARP1 #2). e. Steady-state poly(A) length distribution of endogenous mRNAs measured by Hire-PAT assay in LARP1-depleted HEK293T and HCT116 cells. f. Abundance changes of the same endogenous mRNAs in (e) relative to U1 snRNA measured by RT-qPCR in LARP1-depleted HEK293T and HCT116 cells (n = 5). Data are represented as mean ± SEM. P-values (*p < 0.05, **p < 0.01, ***p < 0.001) were calculated from the one-sided Student’s t-test.

Source data

Extended Data Fig. 5 Steady-state poly(A) length distribution and mRNA abundance changes upon LARPs knockdown.

a. Global poly(A) length distributions of TOP (left) and non-TOP (right) genes in the LARP1 RNA co-IP sample determined by mTAIL-seq. mt-mRNAs were excluded. b.Global poly(A) length distributions of TOP (left) and non-TOP (right) genes in the LARP1 knockdown sample determined by mTAIL-seq. mt-mRNAs were excluded. c. Poly(A) length distributions of mt-mRNAs determined by mTAIL-seq upon depletion of LARP1 or LARP4/4B. d. Global poly(A) length distributions determined by mTAIL-seq upon depletion of LARP4/4B. mt-mRNAs were excluded. The right plot is scaled by the total abundance of mRNAs relative to mt-mRNAs. e. RT-qPCR of the steadily-expressed TOP reporter relative to U1 snRNA upon LARP1 or LARP4/4B depletion (n = 3). The reporter mRNA was targeted by the common constitutive transport element (CTE) in its 3′ UTR. Data are represented as mean ± SEM. P-values (*p < 0.05) were calculated from one-sided Student’s t-test. f. RT-qPCR of the endogenous mRNAs relative to U1 snRNA upon depletion of LARP4/4B measured by qRT-PCR (n = 4). Data are represented as mean ± SEM. P-values (*p < 0.05) were calculated from the one-sided Student’s t-test.

Source data

Extended Data Fig. 6 In vitro experiments with LARP1, PABPC1 and the CCR-CAF1 subcomplex.

a. InstantBlue staining of purified recombinant proteins (n = 1). b. Double-filter binding assay with PABPC1 and RNAs bearing various lengths of poly(A) tail (left) or LARP1 and the RNAs (right). The mean values ± SE from three independent experiments are shown in the graphs. ‘Kd’ and ‘n.d’. denote ‘dissociation constant’ and ‘not determined’, respectively. c. Analytical size-exclusion chromatography with the purified CCR4-CAF1 subcomplex (C) and LARP1-PABPC1-A25 RNA (LPR). Absorbance at 260 nm and 280 nm were drawn in dashed (A260) or solid (A280) lines. d. In vitro deadenylation assay in the presence or absence of PABPC1 and an increasing amount of LARP1. The A50 substrates were pre-incubated with or without PABPC1 and LARP1 prior to adding the CCR4-CAF1 subcomplex (n = 1). e. Time course in vitro deadenylation assay in the presence or absence of recombinant LARP1 using longer poly(A)-tail substrates (A54, A110). The RNA substrates were pre-incubated with PABPC1 and LARP1 prior to adding the CCR4-CAF1 subcomplex (n = 1).

Source data

Supplementary information

Supplementary Information

Supplementary Table 1: Oligonucleotides used in this study.

Reporting Summary

Peer Review File

Source data

Source Data Fig. 2

Raw gel images from bulk poly(A)-tail assay.

Source Data Fig. 3

RT–qPCR data.

Source Data Fig. 5

Uncropped gel images for Fig. 5a,b.

Source Data Fig. 5

Size-exclusion chromatography (SEC) data.

Source Data Extended Data Fig. 3

Raw gel images from western blotting and bulk poly(A)-tail assay.

Source Data Extended Data Fig. 4

RT–qPCR data.

Source Data Extended Data Fig. 4

Raw gel images from western blotting.

Source Data Extended Data Fig. 5

RT–qPCR data.

Source Data Extended Data Fig. 6

Raw data from double-filter binding assay and SEC.

Source Data Extended Data Fig. 6

Raw gel/blot images.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Park, J., Kim, M., Yi, H. et al. Short poly(A) tails are protected from deadenylation by the LARP1–PABP complex. Nat Struct Mol Biol 30, 330–338 (2023). https://doi.org/10.1038/s41594-023-00930-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41594-023-00930-y

This article is cited by

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