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:

Distinct chemical environments in biomolecular condensates

An Author Correction to this article was published on 25 October 2023

This article has been updated

Abstract

Diverse mechanisms have been described for selective enrichment of biomolecules in membrane-bound organelles, but less is known about mechanisms by which molecules are selectively incorporated into biomolecular assemblies such as condensates that lack surrounding membranes. The chemical environments within condensates may differ from those outside these bodies, and if these differed among various types of condensate, then the different solvation environments would provide a mechanism for selective distribution among these intracellular bodies. Here we use small molecule probes to show that different condensates have distinct chemical solvating properties and that selective partitioning of probes in condensates can be predicted with deep learning approaches. Our results demonstrate that different condensates harbor distinct chemical environments that influence the distribution of molecules, show that clues to condensate chemical grammar can be ascertained by machine learning and suggest approaches to facilitate development of small molecule therapeutics with optimal subcellular distribution and therapeutic benefit.

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: Therapeutic small molecules concentrate in distinct intracellular environments.
Fig. 2: Selective partitioning of small molecules in simple condensates.
Fig. 3: Deep learning discovers compounds with selective partitioning behaviors.
Fig. 4: Live cell partitioning predicted by deep learning classifiers.
Fig. 5: Small molecule–protein interactions in condensates.

Similar content being viewed by others

Data availability

Source and probe partitioning screen data are available at https://doi.org/10.6084/m9.figshare.23736693, https://doi.org/10.6084/m9.figshare.23535258.

Code availability

Code is available on GitHub: machine learning tools, https://github.com/pgmikhael/ChemicalGrammar; in vitro droplet assays, https://github.com/jehenninger/in_vitro_droplet_assay; RDKit Calculations, https://github.com/hrkilgore/rdkit_scripts and PLSR Model, https://github.com/uberholzer/partitioning_PLSR.

Change history

References

  1. Banani, S. F., Lee, H. O., Hyman, A. A. & Rosen, M. K. Biomolecular condensates: organizers of cellular biochemistry. Nat. Rev. Mol. Cell Biol. 18, 285–285 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Alberts, B. et al. Molecular Biology of the Cell 7th edn (W. W. Norton & Company, 2022).

  3. Klein, I. A. et al. Partitioning of cancer therapeutics in nuclear condensates. Science 368, 1386–1392 (2020).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  4. Kilgore, H. R. & Young, R. A. Learning the chemical grammar of biomolecular condensates. Nat. Chem. Biol. 18, 1298–1306 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Shin, Y. & Brangwynne, C. P. Liquid phase condensation in cell physiology and disease. Science 357, eaaf4382 (2017).

    Article  PubMed  Google Scholar 

  6. Langdon, E. M. & Gladfelter, A. S. A new lens for RNA localization: liquid-liquid phase separation. Annu. Rev. Microbiol. 72, 255–271 (2018).

    Article  CAS  PubMed  Google Scholar 

  7. Lyon, A. S., Peeples, W. B. & Rosen, M. K. A framework for understanding the functions of biomolecular condensates across scales. Nat. Rev. Mol. Cell Biol. 22, 215–235 (2021).

    Article  CAS  PubMed  Google Scholar 

  8. Alberti, S. & Hyman, A. A. Biomolecular condensates at the nexus of cellular stress, protein aggregation disease and ageing. Nat. Rev. Mol. Cell Biol. 22, 196–213 (2021).

    Article  CAS  PubMed  Google Scholar 

  9. Feric, M. & Misteli, T. Phase separation in genome organization across evolution. Trends Cell Biol. 31, 671–685 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Forman-Kay, J. D., Ditlev, J. A., Nosella, M. L. & Lee, H. O. What are the distinguishing features and size requirements of biomolecular condensates and their implications for RNA-containing condensates? RNA 28, 36–47 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Du, M. & Zhijian, Chen J. DNA-induced liquid phase condensation of cGAS activates innate immune signaling. Science 361, 704–709 (2018).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  12. Howard, T. P. & Roberts, C. W. M. Partitioning of chemotherapeutics into nuclear condensates-opening the door to new approaches for drug development. Mol. Cell 79, 544–545 (2020).

    Article  CAS  PubMed  Google Scholar 

  13. Lyons, H. et al. Functional partitioning of transcriptional regulators by patterned charge blocks. Cell 186, 327–345.e328 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Sabari, B. R. et al. Coactivator condensation at super-enhancers links phase separation and gene control. Science 361, eaar3958 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  15. Larson, A. G. et al. Liquid droplet formation by HP1α suggests a role for phase separation in heterochromatin. Nature 547, 236–240 (2017).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  16. Patel, A. et al. A liquid-to-solid phase transition of the ALS protein FUS accelerated by disease mutation. Cell 162, 1066–1077 (2015).

    Article  CAS  PubMed  Google Scholar 

  17. Rogers, D. & Hahn, M. Extended-connectivity fingerprints. J. Chem. Inf. Model. 50, 742–754 (2010).

    Article  CAS  PubMed  Google Scholar 

  18. Mater, A. C. & Coote, M. L. Deep learning in chemistry. J. Chem. Inf. Model. 59, 2545–2559 (2019).

    Article  CAS  PubMed  Google Scholar 

  19. Walters, W. P. & Barzilay, R. Applications of deep learning in molecule generation and molecular property prediction. Acc. Chem. Res. 54, 263–270 (2021).

    Article  CAS  PubMed  Google Scholar 

  20. Hunter, C. A. & Sanders, J. K. M. The nature of π-π interactions. J. Am. Chem. Soc. 112, 5525–5534 (1990).

    Article  CAS  Google Scholar 

  21. Thakuria, R., Nath, N. K. & Saha, B. K. The nature and applications of π–π interactions: a perspective. Cryst. Growth Des. 19, 523–528 (2019).

    Article  CAS  Google Scholar 

  22. Dougherty, D. A. The cation−π interaction. Acc. Chem. Res. 46, 885–893 (2013).

    Article  CAS  PubMed  Google Scholar 

  23. Stewart, J. A., McCormack, J. J. & Krakoff, I. H. Clinical and clinical pharmacologic studies of mitoxantrone. Cancer Treat. Rep. 66, 1327–1331 (1982).

    CAS  PubMed  Google Scholar 

  24. Kaur, R., Manjal, S. K., Rawal, R. K. & Kumar, K. Recent synthetic and medicinal perspectives of tryptanthrin. Bioorg. Med. Chem. 25, 4533–4552 (2017).

    Article  CAS  PubMed  Google Scholar 

  25. Wang, J. et al. A molecular grammar governing the driving forces for phase separation of prion-like RNA binding proteins. Cell 174, 688–699.e616 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Martin, E. W. et al. Valence and patterning of aromatic residues determine the phase behavior of prion-like domains. Science 367, 694–699 (2020).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  27. Choi, J.-M., Holehouse, A. S. & Pappu, R. V. Physical principles underlying the complex biology of intracellular phase transitions. Annu. Rev. Biophys. 49, 107–133 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Landrum, G. et al. RDKit: Open-Source chemoinformatics. Zenodo https://doi.org/10.5281/zenodo.4750957 (2021).

  29. Pedregosa, F. et al. Scikit-learn: machine learning in Python. J. Mach. Learn. Res. 12, 2825–2830 (2011).

    MathSciNet  Google Scholar 

  30. Mevik, B.-H. & Wehrens, R. The pls package: principal component and partial least squares regression in R. J. Stat. Softw. 18, 1–23 (2007).

    Article  Google Scholar 

  31. R Core Team. R: A Language and Environment for Statistical Computing (R Foundation for Statistical Computing, 2022); https://www.R-project.org/

  32. Mehmood, T., Liland, K. H., Snipen, L. & Sæbø, S. A review of variable selection methods in partial least squares regression. Chemom. Intell. Lab. Syst. 118, 62–69 (2012).

    Article  CAS  Google Scholar 

  33. Yang, K. et al. Analyzing learned molecular representations for property prediction. J. Chem. Inf. Mod. 59, 3370–3388 (2019).

    Article  CAS  Google Scholar 

  34. Jin, W., Barzilay, R. & Jaakkola, T. Multi-objective molecule generation using interpretable substructures. In Proc. 37th Int. Conf. Mach. Learn. (PMLR, 2020).

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

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank J. Platt, J. H. Cheah, C. Soule, W. Salmon and C. Rogers for comments and discussion and A. Tubelli for graphic art. Support was provided by NIH grant nos. GM144283 (R.A.Y.) and CA155258 (R.A.Y.), NSF grant no. PHY2044895 (R.A.Y.), Damon Runyon Cancer Research Foundation Fellowship grant no. 2458-22 (H.R.K.), NSF Graduate Research Fellowship grant no. 1745302 (K.J.O.), the MIT Jameel Clinic for Machine Learning in Health (R.B., P.G.M.) and Eric and Wendy Schmidt Center, Broad Institute (P.G.M.), Basic Science Research Institute Fund grant no. 2021R1A6A1A10042944 (Y.-T.C.) and a National Research Foundation of Korea grant funded by the Korean government (MSIT) (no. 2023R1A2C300453411 to Y.-T.C.).

Author information

Authors and Affiliations

Authors

Contributions

H.R.K., R.B., A.B. and R.A.Y. conceived the idea. Y.-T.C. provided the fluorescent probes. H.R.K., K.J.O., A.B., N.M.H. and C.V.D. performed the experiments. H.R.K., P.G.M. and K.J.O. analyzed the data. H.R.K., P.G.M., K.J.O, T.I.L. and R.A.Y. prepared the manuscript.

Corresponding authors

Correspondence to Henry R. Kilgore or Richard A. Young.

Ethics declarations

Competing interests

R.A.Y. is a founder and shareholder of Syros Pharmaceuticals, Camp4 Therapeutics, Omega Therapeutics, Dewpoint Therapeutics and Paratus Sciences, and has consulting or advisory roles at Precede Biosciences and Novo Nordisk. R.B. has consulting or advisory roles at Dewpoint Therapeutics, J&J, Amgen, Outcomes4Me, Immunai and Firmenich. H.R.K. is a consultant of Dewpoint Therapeutics. The remaining authors declare no competing interests.

Peer review

Peer review information

Nature Chemical Biology thanks the anonymous reviewers for their contribution to the peer review of this work.

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 Live cell confocal and two-photon imaging of endogenously fluorescent drugs.

HCT-116 cells were incubated with a drug or natural product at 50 μM for 1 hour and then imaged with a confocal or two-photon microscope. Image for sunitinib is also shown in Fig. 1. Scale: 10 μm.

Extended Data Fig. 2 Subcellular distribution of small molecules in a breast cancer cell line.

Live MCF7 cells were incubated with a drug or natural product at 50 μM for 1 hour prior to confocal imaging. Scale: 10 μm R = CH2CH2OCH2CH2NH2, R2 = CH2CH2N(CH2CH3)2.

Extended Data Fig. 3 Subcellular distribution of small molecules in a prostate cancer cell line.

Live PC3 cells were incubated with a drug or natural product at 50 μM for 1 hour prior to confocal imaging. Scale: 10 μm. R = CH2CH2OCH2CH2NH2, R2 = CH2CH2N(CH2CH3)2.

Extended Data Fig. 4 Additional analysis of condensate selectivity in fluorescent probe partitioning.

(a-c) Dot plots comparing the partition ratios of probes above the 90th percentile in, a, MED1, b, NPM1, and c, HP1α against their partition ratios in other condensates. P values computed with a two-sided unpaired t-test, n = 50 and d.f.=98 for all groups, with t-statistics and effect sizes (η2) as follows: MED1-NPM1 (t = 22.87, η2 = 0.84), MED1- HP1α (t = 19.88, η2 = 0.80), NPM1-MED1 (t = 3.43, η2 = 0.11), NPM1- HP1α (t = 8.24, η2 = 0.41), HP1α -MED1 (t = 3.15, η2 = 0.09), HP1α -NPM1 (t = 4.09, η2 = 0.15). (d-f) Dot plots comparing the partition ratios of probes below the 10th percentile in d, MED1, e, NPM1, and f, HP1α against their partition ratios in other condensates. P values computed with a two-sided unpaired t-test, n = 50 and d.f.=98 for all groups, with test statistics as follows: MED1-NPM1 (t = 3.79, η2 = 0.13), MED1- HP1α (t = 4.65, η2 = 0.18), NPM1-MED1 (t = 3.12, η2 = 0.09), NPM1- HP1α (t = 2.60, η2 = 0.07), HP1α -MED1 (t = 3.76, η2 = 0.13), HP1α -NPM1 (t = 4.20, η2 = 0.15). (g-i) Dot plots comparing the percentile ranks of probes below the 10th percentile in g, MED1, h, NPM1, and i, HP1α against their percentiles in other condensates. P values computed with a two-sided Wilcoxon matched-pairs signed rank test, n = 50 for all groups, test statistic |W | (left to right): 915, 1003, 928, 1137, 1099, 681. For panels a-i, centerline and error bars represent mean ± s.d., P values were not adjusted for multiple comparisons, sample size n = 100 probes.

Extended Data Fig. 5 Probe features suggest a chemical grammar in condensates.

a, Cartoon depicting how similar molecules (here, sharing color) might interact with the same chemical environment. b, Schematic showing calculation of Tanimoto similarity matrices comparing fluorescent probes by their Morgan Fingerprints. c, Schematic and d, dot plots showing calculation of mean Tanimoto similarities from matrices of fluorescent probes compared against each other in high-to-high (H-H), high-to-low (H-L) and low-to-low (L-L) partitioning regions. e, Graphic and f, dot plots show the comparison of high partitioning probes between condensates through quantification of matrices, significance between groups was not assessed. Centerline and error bars represent mean ± s.d. Panel d, all comparisons were statistically significant with P value, P < 0.0001 (asterisks do not appear in figure), sample size MED1 n = 120. NPM1 n = 100. HP1α n = 100, without adjustment for multiple comparisons. Unpaired two-sided t-test statistic and degrees of freedom: MED1 H-H, t = 9.5, df = 238. MED1 H-L, t = 12.7, df = 238. MED1 L-L, t = 7.3, df = 238. NPM1 H-H, t = 12.17, df = 198. NPM1 H-L, t = 7.4, df = 198. NPM1 L-L, t = 9.4, df = 198. HP1α H-H, t = 4.8, df = 198. HP1α H-L, t = 10.7, df = 198. HP1α L-L, t = 8.3, df = 198.

Extended Data Fig. 6 Comparison of deep learning and Tanimoto similarity approaches toward predicting partitioning behaviors.

Receiver operating characteristic curves quantifying the success of deep learning and Tanimoto similarity approaches for determining partitioning behavior of small molecules into condensates. a, Deep learning-based approach to probe classification (compounds were considered to have partitioned if probes concentrated above 2.7 for MED1, 2.7 for NPM1, and 2.0 for HP1α). Tanimoto similarity approach to probe classification in b, MED1, c, NPM1, and d, HP1α condensates across different thresholds of Tanimoto similarity (compounds were considered to have concentrated into a condensate if probe partition ratio was K > 2.0 for MED1, NPM1, and HP1α).

Extended Data Fig. 7 Live cell two-photon imaging of small molecules in mouse embryonic stem cells.

Live mouse embryonic stem cells were incubated with a drug or natural product and assayed with two-photon imaging. Drugs and natural products are listed in Supplementary Table 1 and their predicted subcellular distribution from machine learning is given in Supplementary Table 2. Scale: 50 μm.

Extended Data Fig. 8 Receiver operating characteristic curves comparing performance of Tanimoto similarity and deep learning classifiers on in vivo compounds.

a, Performance of NPM1 deep learning model (AUC-ROC = 0.62) and Tanimoto similarity (AUC-ROC = 0.52) at identifying drugs and natural products that concentrate in the nucleolus. b, Performance of HP1α deep learning model (AUC-ROC = 0.59) and Tanimoto similarity (AUC-ROC = 0.52) at identifying drugs and natural products that concentrate in chromocenters.

Extended Data Fig. 9 Live cell images showing the partitioning of small molecules into nuclear compartments.

a, Micrograph and line plot showing the signal intensity from mitoxantrone along that indicated gray arrow in HCT-116 cells. b, Mouse embryonic stem cells stained with the DNA dye Hoechst. (a) chromocenter, (b) perinuclear heterochromatin, (c) nucleolus. Zoom (2x). c, Micrograph and line plot showing the signal intensity from tryptanthrin along the indicated gray arrow in mouse embryonic stem cells. Scale: 10 μm. Images were recorded after 1 hour of incubation for tryptanthrin and mitoxantrone, and 10 minutes for Hoechst.

Supplementary information

Supplementary Information

Supplementary Figs. 1–12 and Tables 1–3.

Reporting Summary

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

Kilgore, H.R., Mikhael, P.G., Overholt, K.J. et al. Distinct chemical environments in biomolecular condensates. Nat Chem Biol 20, 291–301 (2024). https://doi.org/10.1038/s41589-023-01432-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41589-023-01432-0

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