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:

Accurate design of translational output by a neural network model of ribosome distribution

Subjects

This article has been updated

Abstract

Synonymous codon choice can have dramatic effects on ribosome speed and protein expression. Ribosome profiling experiments have underscored that ribosomes do not move uniformly along mRNAs. Here, we have modeled this variation in translation elongation by using a feed-forward neural network to predict the ribosome density at each codon as a function of its sequence neighborhood. Our approach revealed sequence features affecting translation elongation and characterized large technical biases in ribosome profiling. We applied our model to design synonymous variants of a fluorescent protein spanning the range of translation speeds predicted with our model. Levels of the fluorescent protein in budding yeast closely tracked the predicted translation speeds across their full range. We therefore demonstrate that our model captures information determining translation dynamics in vivo; that this information can be harnessed to design coding sequences; and that control of translation elongation alone is sufficient to produce large quantitative differences in protein output.

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: Design and performance of a neural network model of translation elongation.
Fig. 2: Performance comparisons on low-coverage genes and with competing models.
Fig. 3: Interpretation of models of translation-elongation rates.
Fig. 4: Design of synonymous sequences showing that elongation rate affects translation output.

Similar content being viewed by others

Change history

  • 09 July 2018

    In the version of this article initially published, Supplementary Notes 1—3 and Supplementary Tables 1 and 2 were omitted from the Supplementary Text and Figures file. The error has been corrected.

References

  1. Ishimura, R. et al. Ribosome stalling induced by mutation of a CNS-specific tRNA causes neurodegeneration. Science 345, 455–459 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  2. Goodarzi, H. et al. Modulated expression of specific tRNAs drives gene expression and cancer progression. Cell 165, 1416–1427 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  3. Kirchner, S. et al. Alteration of protein function by a silent polymorphism linked to tRNA abundance. PLoS Biol. 15, e2000779 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  4. Zhao, F., Yu, C.-H. & Liu, Y. Codon usage regulates protein structure and function by affecting translation elongation speed in Drosophila cells. Nucleic Acids Res. 45, 8484–8492 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  5. Shah, P., Ding, Y., Niemczyk, M., Kudla, G. & Plotkin, J. B. Rate-limiting steps in yeast protein translation. Cell 153, 1589–1601 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  6. Presnyak, V. et al. Codon optimality is a major determinant of mRNA stability. Cell 160, 1111–1124 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  7. Bazzini, A. A. et al. Codon identity regulates mRNA stability and translation efficiency during the maternal-to-zygotic transition. EMBO J. 35, 2087–2103 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  8. Ingolia, N. T., Ghaemmaghami, S., Newman, J. R. S. & Weissman, J. S. Genome-wide analysis in vivo of translation with nucleotide resolution using ribosome profiling. Science 324, 218–223 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  9. Stadler, M. & Fire, A. Wobble base-pairing slows in vivo translation elongation in metazoans. RNA 17, 2063–2073 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  10. Dana, A. & Tuller, T. Determinants of translation elongation speed and ribosomal profiling biases in mouse embryonic stem cells. PLOS Comput. Biol. 8, e1002755 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  11. Charneski, C. A. & Hurst, L. D. Positively charged residues are the major determinants of ribosomal velocity. PLoS Biol. 11, e1001508 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  12. Gardin, J. et al. Measurement of average decoding rates of the 61 sense codons in vivo. eLife 3, e03735 (2014).

    Article  PubMed Central  Google Scholar 

  13. Lareau, L. F., Hite, D. H., Hogan, G. J. & Brown, P. O. Distinct stages of the translation elongation cycle revealed by sequencing ribosome-protected mRNA fragments. eLife 3, e01257 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  14. Pop, C. et al. Causal signals between codon bias, mRNA structure, and the efficiency of translation and elongation. Mol. Syst. Biol. 10, 770 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  15. Liu, T.-Y. & Song, Y. S. Prediction of ribosome footprint profile shapes from transcript sequences. Bioinformatics 32, i183–i191 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  16. Zhang, S. et al. Analysis of ribosome stalling and translation elongation dynamics by deep learning. Cell Syst. 5, 212–220.e6 (2017).

    Article  PubMed  CAS  Google Scholar 

  17. Dao Duc, K. & Song, Y. S. The impact of ribosomal interference, codon usage, and exit tunnel interactions on translation elongation rate variation. PLoS Genet. 14, e1007166 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  18. Fang, H. et al. Scikit-ribo enables accurate estimation and robust modeling of translation dynamics at codon resolution. Cell Syst. 6, 180–191.e4 (2018).

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  19. O’Connor, P. B. F., Andreev, D. E. & Baranov, P. V. Comparative survey of the relative impact of mRNA features on local ribosome profiling read density. Nat. Commun. 7, 12915 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  20. Artieri, C. G. & Fraser, H. B. Accounting for biases in riboprofiling data indicates a major role for proline in stalling translation. Genome Res. 24, 2011–2021 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  21. Hussmann, J. A., Patchett, S., Johnson, A., Sawyer, S. & Press, W. H. Understanding biases in ribosome profiling experiments reveals signatures of translation dynamics in yeast. PLoS Genet. 11, e1005732 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  22. Weinberg, D. E. et al. Improved ribosome-footprint and mRNA measurements provide insights into dynamics and regulation of yeast translation. Cell Rep. 14, 1787–1799 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  23. Sharp, P. M., Tuohy, T. M. & Mosurski, K. R. Codon usage in yeast: cluster analysis clearly differentiates highly and lowly expressed genes. Nucleic Acids Res 14, 5125–5143 (1986).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  24. Plotkin, J. B. & Kudla, G. Synonymous but not the same: the causes and consequences of codon bias. Nat. Rev. Genet. 12, 32–42 (2011).

    Article  PubMed  CAS  Google Scholar 

  25. dos Reis, M., Savva, R. & Wernisch, L. Solving the riddle of codon usage preferences: a test for translational selection. Nucleic Acids Res. 32, 5036–5044 (2004).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  26. Letzring, D. P., Dean, K. M. & Grayhack, E. J. Control of translation efficiency in yeast by codon-anticodon interactions. RNA 16, 2516–2528 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  27. Gamble, C. E., Brule, C. E., Dean, K. M., Fields, S. & Grayhack, E. J. Adjacent codons act in concert to modulate translation efficiency in yeast. Cell 166, 679–690 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  28. Murphy, F. V. IV & Ramakrishnan, V. Structure of a purine-purine wobble base pair in the decoding center of the ribosome. Nat. Struct. Mol. Biol. 11, 1251–1252 (2004).

    Article  PubMed  CAS  Google Scholar 

  29. McGlincy, N. J. & Ingolia, N. T. Transcriptome-wide measurement of translation by ribosome profiling. Methods 126, 112–129 (2017).

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  30. Iwasaki, S., Floor, S. N. & Ingolia, N. T. Rocaglates convert DEAD-box protein eIF4A into a sequence-selective translational repressor. Nature 534, 558–561 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  31. Schuller, A. P., Wu, C. C.-C., Dever, T. E., Buskirk, A. R. & Green, R. eIF5A functions globally in translation elongation and termination. Mol. Cell 66, 194–205.e5 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  32. Sheff, M. A. & Thorn, K. S. Optimized cassettes for fluorescent protein tagging in Saccharomyces cerevisiae. Yeast 21, 661–670 (2004).

    Article  PubMed  CAS  Google Scholar 

  33. Chu, D. et al. Translation elongation can control translation initiation on eukaryotic mRNAs. EMBO J. 33, 21–34 (2014).

    Article  PubMed  CAS  Google Scholar 

  34. Qian, W., Yang, J.-R., Pearson, N. M., Maclean, C. & Zhang, J. Balanced codon usage optimizes eukaryotic translational efficiency. PLoS Genet. 8, e1002603 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  35. Langmead, B., Trapnell, C., Pop, M. & Salzberg, S. L. Ultrafast and memory-efficient alignment of short DNA sequences to the human genome. Genome Biol. 10, R25 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  36. Li, B. & Dewey, C. N. RSEM: accurate transcript quantification from RNA-Seq data with or without a reference genome. BMC Bioinformatics 12, 323 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  37. Battenberg, E. et al. Lasagne: first release. https://doi.org/10.5281/zenodo.27878 (2015).

  38. The Theano Development Team et al. Theano: a Python framework for fast computation of mathematical expressions. https://arxiv.org/abs/1605.02688 (2016).

  39. Ritz, C. & Spiess, A.-N. qpcR: an R package for sigmoidal model selection in quantitative real-time polymerase chain reaction analysis. Bioinformatics 24, 1549–1551 (2008).

    Article  PubMed  CAS  Google Scholar 

  40. Burdukiewicz, M. et al. Methods for comparing multiple digital PCR experiments. Biomol. Detect. Quantif. 9, 14–19 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  41. Gibson, D. G. et al. Enzymatic assembly of DNA molecules up to several hundred kilobases. Nat. Methods 6, 343–345 (2009).

    Article  PubMed  CAS  Google Scholar 

  42. Gueldener, U., Heinisch, J., Koehler, G. J., Voss, D. & Hegemann, J. H. A second set of loxP marker cassettes for Cre-mediated multiple gene knockouts in budding yeast. Nucleic Acids Res. 30, e23 (2002).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  43. Daniel Gietz, R. & Woods, R. A. Transformation of yeast by lithium acetate/single-stranded carrier DNA/polyethylene glycol method. Methods Enzymol. 350, 87–96 (2002).

    Article  PubMed  Google Scholar 

  44. Hahne, F. et al. flowCore: a Bioconductor package for high throughput flow cytometry. BMC Bioinformatics 10, 106 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  45. Sarkar, D., Le Meur, N. & Gentleman, R. Using flowViz to visualize flow cytometry data. Bioinformatics 24, 878–879 (2008).

    Article  PubMed  CAS  Google Scholar 

  46. Ares, M. Isolation of total RNA from yeast cell cultures. Cold Spring Harb. Protoc. 2012, 1082–1086 (2012).

    PubMed  Google Scholar 

Download references

Acknowledgements

We are grateful to N. Ingolia and S. McCurdy for discussion. This work was supported by the National Cancer Institute of the National Institutes of Health, under award R21CA202960 to L.F.L., and by the National Institute of General Medical Sciences of the National Institutes of Health, under award P50GM102706 to the Berkeley Center for RNA Systems Biology. R.T. was supported by the Department of Defense through the National Defense Science & Engineering Graduate Fellowship (NDSEG) Program. This work made use of the Vincent J. Coates Genomics Sequencing Laboratory at the University of California, Berkeley, supported by National Institutes of Health S10 Instrumentation grant OD018174, and the UC Berkeley flow cytometry core facilities.

Author information

Authors and Affiliations

Authors

Contributions

L.F.L., R.T., and N.J.M. designed the study, with input from L.P. R.T. developed the software and performed modeling, and R.T., L.P., and L.F.L. analyzed and interpreted the computational results. N.J.M. designed and created the yeast strains and performed expression experiments, with assistance from M.E.G. and N.N. M.E.G. performed yeast ribosome profiling. N.J.M. and L.F.L. analyzed and interpreted the experimental data. R.T. and L.F.L. wrote the manuscript, with input from all authors.

Corresponding author

Correspondence to Liana F. Lareau.

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.

Integrated supplementary information

Supplementary Figure 1 Neural network model structure.

(A) Counts are predicted at the A site codon. In the model shown, a sequence neighborhood spanning from 5 codons upstream of the A site (codon -5) to 4 codons downstream of the A site (codon + 4) is used as the predictive region. This neighborhood is divided into codons and encoded via one-hot encoding (purple) for input into a regression model. We also encode the same region as nucleotide features (green) and include these features in the model. Finally, we compute RNA structure scores on three 30 nt sliding structure windows that span the width of a typical 28 nt footprint. These windows start 17, 16, and 15 nucleotides before the start of the A site. (B) These features are concatenated in a vector, which is used as the input to a fully connected feedforward neural network model. Each model in this paper contains one hidden layer with 200 hidden units, and a tanh activation function on the hidden units. The output layer contains one unit with a ReLU activation function to enforce nonnegativity of predicted scaled counts.

Supplementary Figure 2 Change in MSE upon including mRNA structure.

We compared the performance of a model using a sequence neighborhood from codons -7 to + 5 against models that each also include the folding energy of a single 30 nt window. Each point shows the change in MSE upon including a particular window. The greatest improvement in MSE is achieved by including a window starting at nucleotide position -17 and ending at position 12. This is roughly coterminal with a typical 28 nt ribosome footprint (nucleotide positions -15 to + 12).

Supplementary Figure 3 Relative contributions of A-site codons and context.

(A) Pearson correlation of observed vs. predicted scaled counts per codon, for a model using codons -3 to + 2 and associated nucleotides (x-axis) and a model using the same region but without the A site (y-axis). Codons whose inclusion in the model leads to significantly better prediction (higher correlation between the observed and predicted scaled counts), per a t-test of the Fisher transformation of correlations with an FDR of 5%, are shown in red. (B) We extracted the sequence context (codons -3 to + 2) for all positions where the squared error was higher for an A-site-only model than for a model with codons -3 to + 2 but no A site. The proportion of codons at each position in this set was compared to the overall distribution of codons in the test set with a two-sided proportion test, using an FDR of 5%, and codons with adjusted p < 0.05 are shown in the plot.

Supplementary Figure 4 Predictive values of codon positions.

Predictive value of codon positions in (A) a human ribosome profiling data set using Circligase II (Iwasaki, S. et al., Nature. 534(7608), 558-61, 2016), and (B) a yeast ribosome profiling dataset using Circligase I (Schuller, A. et al., Mol. Cell. 66, 194–205.e5, 2017). As in Fig. 3a, we trained a reference model on codons -7 to + 5 (with nucleotide features over the same neighborhood), and then a series of leave-one-out models each excluding exactly one codon in the sequence neighborhood, along with the corresponding nucleotides. For each model, we compute Pearson correlations between the true and predicted scaled counts over all codons in the test set. Shown is the difference in Pearson correlations between the reference model and the leave-one-out models. Higher Δr indicates increased importance of that codon position to model predictions.

Supplementary Figure 5 Circligase I and II preferences.

(A) Mean contributions to scaled counts at the 5´ end of a ribosome footprint, for yeast data sets generated with Circligase II (our data) and Circligase I (Schuller, A. et al., Mol. Cell. 66, 194–205.e5, 2017). Scores are from the -5 codon position. To generate these scores, we trained models only on 28 nt footprints with their 5´ end aligning with the beginning of the -5 codon. (B) Ligation efficiency of CircLigase I enzyme, as in Fig. 3g.

Supplementary Figure 6 Flow cytometry gating strategy.

Scatter plot of forward scatter area (FSC, arbitrary units) against side scatter area (SSC, arbitrary units) for each of the 50000 events collected for a representative flow cytometry sample of diploid yeast expressing mCherry and a differentially optimized eCitrine. Events are colored by their density on the plot, low density points being colored blue moving to high density events being colored dark red. Events outside the plotted area are denoted by grey lines at the edge of the plot. Annotated by red regions are seven areas of high local density defined by the curv2filter method, each with the number of events they contain. For each sample, events within the most populous region were taken forward for further analysis; in this representative sample, these would be the events within Area 3.

Supplementary Figure 7 yECitrine mRNA level and fluorescence.

(A) eCitrine:mCherry fluorescence ratio as in Fig. 3c, including this ratio for the yECitrine sequence (magenta). (B) eCitrine:mCherry mRNA ratio measured by qPCR in biological replicates of four strains (colors as in Fig. 4). Each data point represents the ratio of medians of three technical replicates, normalized to the median ratio of the highest expression strain.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–7, Supplementary Notes 1–3 and Supplementary Tables 1 and 2

Reporting Summary

Supplementary Table 3

Codon contributions. Weights learned as described in the methods for each codon in positions -5 to +4, used to create figure 3B

Supplementary Dataset 1

Annotated sequences of plasmids for eCitrine strain creation, in genbank format

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Tunney, R., McGlincy, N.J., Graham, M.E. et al. Accurate design of translational output by a neural network model of ribosome distribution. Nat Struct Mol Biol 25, 577–582 (2018). https://doi.org/10.1038/s41594-018-0080-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41594-018-0080-2

This article is cited by

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research