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:

Aromatic 19F-13C TROSY: a background-free approach to probe biomolecular structure, function, and dynamics

Abstract

Atomic-level information about the structure and dynamics of biomolecules is critical for an understanding of their function. Nuclear magnetic resonance (NMR) spectroscopy provides unique insights into the dynamic nature of biomolecules and their interactions, capturing transient conformers and their features. However, relaxation-induced line broadening and signal overlap make it challenging to apply NMR spectroscopy to large biological systems. Here we took advantage of the high sensitivity and broad chemical shift range of 19F nuclei and leveraged the remarkable relaxation properties of the aromatic 19F-13C spin pair to disperse 19F resonances in a two-dimensional transverse relaxation-optimized spectroscopy spectrum. We demonstrate the application of 19F-13C transverse relaxation-optimized spectroscopy to investigate proteins and nucleic acids. This experiment expands the scope of 19F NMR in the study of the structure, dynamics, and function of large and complex biological systems and provides a powerful background-free NMR probe.

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: Theoretical R2 values of 13C, 19F, and 1H resonances as a function of magnetic field strength.
Fig. 2: TROSY effect observed in a coupled 1D 13C spectrum of GB1 at multiple magnetic field strengths.
Fig. 3: 19F-13C TROSY experiments for the 42-kDa MBP with different excitation and detection schemes.
Fig. 4: The 13CF TROSY resolves all expected cross-peaks for 3-19F13C Tyr-labeled MBP at 10 °C.
Fig. 5: 13CF TROSY resonances of the 180-kDa proteasome α7 single-ring particle.
Fig. 6: TROSY selection of the narrowest component in 19F-13C correlation spectra of a 16-mer 5-fluorouracil-substituted DNA, at 5 °C.

Similar content being viewed by others

Data availability

Pulse sequences and parameter sets are available as Supplementary Software and at the laboratory website, https://artlab.dana-farber.org/19f_13c_aromatictrosy.html.

References

  1. Pervushin, K., Riek, R., Wider, G. & Wüthrich, K. Attenuated T2 relaxation by mutual cancellation of dipole-dipole coupling and chemical shift anisotropy indicates an avenue to NMR structures of very large biological macromolecules in solution. Proc. Natl Acad. Sci. USA 94, 12366–12371 (1997).

    Article  CAS  Google Scholar 

  2. Pervushin, K., Riek, R., Wider, G. & Wuthrich, K. Transverse relaxation-optimized spectroscopy (TROSY) for NMR studies of aromatic spin systems in 13C-labeled proteins. J. Am. Chem. Soc. 120, 6394–6400 (1998).

    Article  CAS  Google Scholar 

  3. Sprangers, R. & Kay, L. E. Quantitative dynamics and binding studies of the 20S proteasome by NMR. Nature 445, 618–622 (2007).

    Article  CAS  Google Scholar 

  4. Fiaux, J., Bertelsen, E. B., Horwich, A. L. & Wüthrich, K. NMR analysis of a 900K GroEL GroES complex. Nature 418, 207–211 (2002).

    Article  CAS  Google Scholar 

  5. Meyer, L. H. & Gutowsky, H. S. Electron distribution in molecules. II. Proton and fluorine magnetic resonance shifts in the halomethanes. J. Phys. Chem. 57, 481–486 (1953).

    Article  CAS  Google Scholar 

  6. Solomon, I. Relaxation processes in a system of two spins. Phys. Rev. 99, 559–565 (1955).

    Article  CAS  Google Scholar 

  7. Sykes, B. D. & Hull, W. E. in Methods in Enzymology Vol. 49 (eds. Hirs, C. H. W. & Timasheff, S. N.) 270–295 (Elsevier, 1978).

  8. Liu, J. J., Horst, R., Katritch, V., Stevens, R. C. & Wüthrich, K. Biased signaling pathways in β2-adrenergic receptor characterized by 19F-NMR. Science 335, 1106–1110 (2012).

    Article  CAS  Google Scholar 

  9. Kim, T. H. et al. The role of ligands on the equilibria between functional states of a G protein-coupled receptor. J. Am. Chem. Soc. 135, 9465–9474 (2013).

    Article  CAS  Google Scholar 

  10. Barhate, N. B., Barhate, R. N., Cekan, P., Drobny, G. & Sigurdsson, S. T. A nonafluoro nucleoside as a sensitive 19f NMR probe of nucleic acid conformation. Org. Lett. 10, 2745–2747 (2008).

    Article  CAS  Google Scholar 

  11. Kreutz, C., Kählig, H., Konrat, R. & Micura, R. A general approach for the identification of site-specific RNA binders by 19F NMR spectroscopy: proof of concept. Angew. Chem. Int. Ed. Engl. 45, 3450–3453 (2006).

    Article  CAS  Google Scholar 

  12. Kiviniemi, A. & Virta, P. Characterization of RNA invasion by 19F NMR spectroscopy. J. Am. Chem. Soc. 132, 8560–8562 (2010).

    Article  CAS  Google Scholar 

  13. Bao, H. L. et al. Characterization of human telomere RNA G-quadruplex structures in vitro and in living cells using 19F NMR spectroscopy. Nucleic Acids Res. 45, 5501–5511 (2017).

    Article  CAS  Google Scholar 

  14. Gerig, J. T. Fluorine NMR of proteins. Prog. Nucl. Magn. Reson. Spectrosc. 26, 293–370 (1994).

    Article  CAS  Google Scholar 

  15. Brindle, K., Williams, S.-P. & Boulton, M. 19F NMR detection of a fluorine-labelled enzyme in vivo. FEBS Lett. 255, 121–124 (1989).

    Article  CAS  Google Scholar 

  16. Li, C. et al. Protein 19F NMR in Escherichia coli. J. Am. Chem. Soc. 132, 321–327 (2010).

    Article  CAS  Google Scholar 

  17. Gronenborn, A. M. et al. A novel, highly stable fold of the immunoglobulin binding domain of streptococcal protein G. Science 253, 657–661 (1991).

    Article  CAS  Google Scholar 

  18. Gardner, K. H., Zhang, X., Gehring, K. & Kay, L. E. Solution NMR studies of a 42 KDa Escherichia coli maltose binding protein/β-cyclodextrin complex: chemical shift assignments and analysis. J. Am. Chem. Soc. 120, 11738–11748 (1998).

    Article  CAS  Google Scholar 

  19. Sprangers, R. et al. TROSY-based NMR evidence for a novel class of 20S proteasome inhibitors. Biochem. 47, 6727–6734 (2008).

    Article  CAS  Google Scholar 

  20. Hogben, H. J., Krzystyniak, M., Charnock, G. T., Hore, P. J. & Kuprov, I. Spinach—a software library for simulation of spin dynamics in large spin systems. J. Magn. Reson. 208, 179–194 (2011).

    Article  CAS  Google Scholar 

  21. Kitevski-LeBlanc, J. L., Al-Abdul-Wahid, M. S. & Prosser, R. S. A mutagenesis-free approach to assignment of 19F NMR resonances in biosynthetically labeled proteins. J. Am. Chem. Soc. 131, 2054–2055 (2009).

    Article  CAS  Google Scholar 

  22. Nietlispach, D. Suppression of anti-TROSY lines in a sensitivity enhanced gradient selection TROSY scheme. J. Biomol. NMR 31, 161–166 (2005).

    Article  CAS  Google Scholar 

  23. Khan, F., Kuprov, I., Craggs, T. D., Hore, P. J. & Jackson, S. E. 19F NMR studies of the native and denatured states of green fluorescent protein. J. Am. Chem. Soc. 128, 10729–10737 (2006).

    Article  CAS  Google Scholar 

  24. Religa, T. L., Sprangers, R. & Kay, L. E. Dynamic regulation of archaeal proteasome gate opening as studied by TROSY NMR. Science 328, 98–102 (2010).

    Article  CAS  Google Scholar 

  25. Riek, R., Pervushin, K., Fernández, C., Kainosho, M. & Wüthrich, K. [13C,13C]- and [13C,1H]-TROSY in a triple resonance experiment for ribose-base and intrabase correlations in nucleic acids. J. Am. Chem. Soc. 123, 658–664 (2001).

    Article  CAS  Google Scholar 

  26. Seela, F. & Xu, K. DNA with stable fluorinated dA and dG substitutes: syntheses, base pairing and 19F-NMR spectra of 7-fluoro-7-deaza-2'-deoxyadenosine and 7-fluoro-7-deaza-2'-deoxyguanosine. Org. Biomol. Chem. 6, 3552–3560 (2008).

    Article  CAS  Google Scholar 

  27. García de la Torre, J., Huertas, M. L. & Carrasco, B. Hydronmr: prediction of NMR relaxation of globular proteins from atomic-level structures and hydrodynamic calculations. J. Magn. Reson. 147, 138–146 (2000).

    Article  Google Scholar 

  28. Arntson, K. E. & Pomerantz, W. C. Protein-observed fluorine NMR: a bioorthogonal approach for small molecule discovery. J. Med. Chem. 59, 5158–5171 (2016).

    Article  CAS  Google Scholar 

  29. Kitevski-LeBlanc, J. L. & Prosser, R. S. Current applications of 19F NMR to studies of protein structure and dynamics. Prog. Nucl. Magn. Reson. Spectrosc. 62, 1–33 (2012).

    Article  CAS  Google Scholar 

  30. Marsh, E. N. & Suzuki, Y. Using 19F NMR to probe biological interactions of proteins and peptides. ACS Chem. Biol. 9, 1242–1250 (2014).

    Article  CAS  Google Scholar 

  31. Danielson, M. A. & Falke, J. J. Use of 19F NMR to probe protein structure and conformational changes. Annu. Rev. Biophys. Biomol. Struct. 25, 163–195 (1996).

    Article  CAS  Google Scholar 

  32. Kitevski-LeBlanc, J. L., Evanics, F. & Scott Prosser, R. Optimizing 19F NMR protein spectroscopy by fractional biosynthetic labeling. J. Biomol. NMR 48, 113–121 (2010).

    Article  CAS  Google Scholar 

  33. Gee, C. T. et al. Protein-observed 19F-NMR for fragment screening, affinity quantification and druggability assessment. Nat. Protoc. 11, 1414–1427 (2016).

    Article  Google Scholar 

  34. Wang, M. et al. Fast magic-angle spinning 19F NMR spectroscopy of HIV-1 capsid protein assemblies. Angew. Chem. Int. Ed. Engl. 57, 16375–16379 (2018).

    Article  CAS  Google Scholar 

  35. Whittaker, M. M. & Whittaker, J. W. Construction and characterization of Pichia pastoris strains for labeling aromatic amino acids in recombinant proteins. Protein Expr. Purif. 41, 266–274 (2005).

    Article  CAS  Google Scholar 

  36. Budisa, N., Wenger, W. & Wiltschi, B. Residue-specific global fluorination of Candida antarctica lipase B in Pichia pastoris. Mol. Biosyst. 6, 1630–1639 (2010).

    Article  CAS  Google Scholar 

  37. Watts, J. K. et al. Differential stability of 2’F-ANA*RNA and ANA*RNA hybrid duplexes: roles of structure, pseudohydrogen bonding, hydration, ion uptake and flexibility. Nucleic Acids Res. 38, 2498–2511 (2010).

    Article  CAS  Google Scholar 

  38. Kitevski-Leblanc, J. L., Evanics, F. & Scott Prosser, R. Approaches to the assignment of 19F resonances from 3-fluorophenylalanine labeled calmodulin using solution state NMR. J. Biomol. NMR 47, 113–123 (2010).

    Article  CAS  Google Scholar 

  39. Kitevski-LeBlanc, J. L., Evanics, F. & Prosser, R. S. Approaches for the measurement of solvent exposure in proteins by 19F NMR. J. Biomol. NMR 45, 255–264 (2009).

    Article  CAS  Google Scholar 

  40. Kovacs, H. & Kupče, Ē. Parallel MR spectroscopy with simultaneous detection of 1H and 19F nuclei. Magn. Reson. Chem. 54, 544–560 (2016).

    Article  CAS  Google Scholar 

  41. Tang, P., Furuya, T. & Ritter, T. Silver-catalyzed late-stage fluorination. J. Am. Chem. Soc. 132, 12150–12154 (2010).

    Article  CAS  Google Scholar 

  42. Frieden, C., Hoeltzli, S. D. & Bann, J. G. The preparation of 19F-labeled proteins for NMR studies. Methods Enzymol. 380, 400–415 (2004).

    Article  CAS  Google Scholar 

  43. Salwiczek, M., Nyakatura, E. K., Gerling, U. I., Ye, S. & Koksch, B. Fluorinated amino acids: compatibility with native protein structures and effects on protein-protein interactions. Chem. Soc. Rev. 41, 2135–2171 (2012).

    Article  CAS  Google Scholar 

  44. Konas, D. W., Seci, D. & Tamimi, S. Synthesis of (L)-4-fluorotryptophan. Synth. Commun. 42, 144–152 (2012).

    Article  CAS  Google Scholar 

  45. Takeuchi, Y., Tarui, T. & Shibata, N. A novel and efficient synthesis of 3-fluorooxindoles from indoles mediated by Selectfluor. Org. Lett. 2, 639–642 (2000).

    Article  CAS  Google Scholar 

  46. Kim, T. H. et al. The role of dimer asymmetry and protomer dynamics in enzyme catalysis. Science 355, eaag2355 (2017).

    Article  Google Scholar 

  47. Hoang, J. & Prosser, R. S. Conformational selection and functional dynamics of calmodulin: a 19F nuclear magnetic resonance study. Biochem. 53, 5727–5736 (2014).

    Article  CAS  Google Scholar 

  48. Schwieters, C. D. et al. Solution structure of the 128 kDa enzyme I dimer from Escherichia coli and its 146 kDa complex with HPr using residual dipolar couplings and small- and wide-angle X-ray scattering. J. Am. Chem. Soc. 132, 13026–13045 (2010).

    Article  CAS  Google Scholar 

  49. Depetris, I. et al. Fluoropyrimidine-induced cardiotoxicity. Crit. Rev. Oncol. Hematol. 124, 1–10 (2018).

    Article  Google Scholar 

  50. Voeks, D. et al. Gene therapy for prostate cancer delivered by ovine adenovirus and mediated by purine nucleoside phosphorylase and fludarabine in mouse models. Gene Ther. 9, 759–768 (2002).

    Article  CAS  Google Scholar 

  51. Delaglio, F. et al. NMRPipe: a multidimensional spectral processing system based on UNIX pipes. J. Biomol. NMR 6, 277–293 (1995).

    Article  CAS  Google Scholar 

  52. Fogh, R. et al. The CCPN project: an interim report on a data model for the NMR community. Nat. Struct. Biol. 9, 416–418 (2002).

    Article  CAS  Google Scholar 

  53. Shaka, A. J., Keeler, J., Frenkiel, T. & Freeman, R. An improved sequence for broadband decoupling: WALTZ-16. J. Magn. Reson. 52, 335–338 (1983).

    CAS  Google Scholar 

Download references

Acknowledgements

We thank K.E. Leigh and M. Kostic for help in proofreading of the manuscript presented here and discussions regarding the work, and D. Baldisseri for help with the NMR experiments. We are especially grateful to L.E. Kay (Departments of Molecular Genetics, Biochemistry, and Chemistry, University of Toronto, Toronto, Ontario, Canada) for sharing the plasmid carrying the single-ring α7 proteasome particle with us. This research was supported by the NIH (grant nos. GM047467, GM129026 and AI037581 to G.W.), the Claudia Adams Barr Program for Innovative Cancer Research (H.A.), the Austrian Science Fund FWF (Schrödinger Fellowship no. J3872-B21 to A.B.), and the National Health and Medical Research Council Australia (C. J. Martin Fellowship no. APP1090444 to S.C.). Maintenance of the NMR instruments used for this research was supported by NIH grant no. EB002026.

Author information

Authors and Affiliations

Authors

Contributions

A.B., S.C., G.W., V.M.G., I.K., K.T., and H.A. designed the research. A.B., S.C., A.D., D.L.R., N.T.B., C.D.C., O.I.P., V.M.G., C.A., M.Z., H.K., I.K., K.T., and H.A. performed experiments. A.B., S.C., A.D., V.M.G., H.K., I.K., and H.A. analyzed the data. A.B., S.C., A.D., G.W., I.K., D.L.R., V.M.G., K.T., and H.A. wrote the paper.

Corresponding authors

Correspondence to Koh Takeuchi or Haribabu Arthanari.

Ethics declarations

Competing interests

V.M.G. is the founder of FB Reagents Ltd, a company that provides isotopically enriched NMR reagents. C.A. and H.K. work for Bruker Biospin Corporation, which is a manufacturer of equipment used in this work.

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 Calculated magnetic field dependence of the transverse relaxation rates for 13CF nuclei and 19FC nuclei in aromatic amino acids with various 19F substitutions.

Blue triangles represent the relaxation rates of decoupled resonances, and red squares represent those of TROSY resonances. The points are connected by a dashed line to provide visual clarity. The relaxation rates are calculated for the respective 19F-bearing amino acids incorporated into a protein with a rotational correlation time (τc) of 25 ns.

Supplementary Figure 2 Selective detection of the TROSY component.

a,b, 13C-19F HSQCs of 3-19F13C Tyr-labeled MBP recorded at 600 MHz: (a) decoupled and (b) coupled. The expected narrowing for the TROSY component is observed in b. A dashed rectangle highlights the four components of one decoupled spin-pair (red). The top right component is clearly more intense than the top left component and the bottom components, which are anti-TROSY with respect to 13CF are substantially broadened. In addition, two intense (more mobile) 13CF TROSY peaks are highlighted and a 1D trace identifies the component to the right (top right) as the TROSY component. c, The TROSY component was selectively observed by the ST2PT scheme. The negative signal (purple) originates from imperfect cancelation of the anti-TROSY component.

Supplementary Figure 3 Sensitivity-enhanced (SE) and non-SE out-and-back 13C-19F TROSYs exhibit comparable sensitivity.

a,b, Pulse schemes for (a) SE and (b) non-SE TROSY versions. c,d, The TROSY spectra of 3-19F13C Tyr-labeled MBP recorded at 600 MHz with SE (c) and non-SE (d) pulse sequences show no substantial difference in sensitivity. Owing to the fast-relaxing 19F magnetization spending more time in the transverse plane in the SE version of the pulse sequence, the typical sensitivity gains associated with the SE experiment were offset. All narrow rectangles denote 90° pulses, and broad rectangles represent 180° pulses. 1H refocusing and decoupling were not possible because of the architecture of the probe, but could in principle be implemented and hence are represented in the pulse sequence. The duration of the delay T was set to 2.08 ms (= 1/2 JCF). Δ is a delay to refocus any chemical shift evolution occurring during evolution of the first point in the indirect dimension, whereas Δ1 refocuses chemical shift evolution occurring during the last gradient and the gradient recovery period. All pulses, unless otherwise noted, are applied along the x-axis. The phase cycles used are indicated under the respective pulse programs. The quadrature detection in the indirect dimension for a was achieved using the Echo-Anti-echo detection scheme, and for b it was done using the States-TPPI approach by incrementing the phase ϕ1. The duration and strengths of the gradients are g0 = (the length of the indirect encoding period, 1.65 Gauss/cm); g1 = (750 μs, 16.5 Gauss/cm); g2 = (500 μs, 44 Gauss/cm); g3 = (750 μs, 24.75 Gauss/cm); g4 = (750 μs, 27.5 Gauss/cm) and g5 = (500 μs, 23.51 Gauss/cm) in the SE version of the experiment (a). The duration and strengths of the gradients in the non-SE TROSY experiments (b) were g0 = (the length of the indirect encoding period, 1.65 Gauss/cm); g1 = (700 ms, 12.9 Gauss/cm); g2 = (500 ms, –33.0 Gauss/cm); g3 = (700 ms, 22 Gauss/cm); g4 = (700 ms, 31.68 Gauss/cm) and g5 = (500 ms, 23.52 Gauss/cm).

Supplementary Figure 4 Contribution of DD and CSA mechanisms to the relaxation rates of various product states in the 13C-19F TROSY experiment.

Relaxation rates of 13C-19F product states are shown, with contributions from 19F-13C DD (green), 19F CSA (red), and 13C CSA (blue). The theoretical calculations were performed for 3-19F13C Tyr incorporated into a protein with a rotational correlation time (τc) of 25 ns at a magnetic field strength of 600 MHz.

Supplementary Figure 5 Pulse schemes for 13C-start, 19F-detected TROSY and 19F-start, 13C-detected TROSY.

Narrow rectangular bars denote flip angles of 90°, and wide bars denote flip angles of 180°. 1H refocusing and decoupling during acquisition were not possible because of the architecture of the probe, but could in principle be implemented and hence are represented in the pulse sequence. The duration of the delay T was set to 2.08 ms (= 1/2 JCF). All pulses, unless otherwise noted, are applied along the x-axis. The phase cycles used are indicated under the respective pulse programs. The indirect dimension is acquired using the Echo-Anti-echo detection scheme by changing ϕ1 for quadrature detection. For both the 13C-start, 19F-detected TROSY and 19F-start, 13C-detected TROSY, ϕ2 and ϕ3 are set to y and –y, respectively. The duration and strengths of the gradients are g3 = (750 μs, 24.75 Gauss/cm) and g4 = (750 μs, 27.5 Gauss/cm).

Supplementary Figure 6 Illustration of spin-state selection by ST2PT scheme.

Overlay of four out-and-stay 19F-13C correlation spectra where the various components of a coupled 19F-13C spin system were recorded with spin-state selection on an MBP sample at 600 MHz, labeled with 3-19F13C Tyr. The four components corresponding to a single 19F-13C correlation (a representative 3-19F13C Tyr residue in MBP) are highlighted in a dashed rectangular box. The top-right component (brown) is the TROSY resonance.

Supplementary Figure 7 Longitudinal relaxation times (T1) for 13CF and 19FC nuclei in 3-19FC Tyr-labeled GB1 at 500 MHz.

a,b, 1D 13CF (a) and 1D 19FC (b) spectra depicting the region with peaks from 3-19F13C Tyr. c,d, Longitudinal relaxation curves (blue) and relaxation times (T1) fitted to experimentally measured values (red points) for peaks 1–6 in a (c) and b (d).

Supplementary Figure 8 Calculated transverse relaxation rates (R2) as a function of rotational correlation time.

ad, R2 for (a) 13CF and (c) 19FC in 3-19F13C Tyr and (b) 13CH and (d) 1HC in Tyr as a function of rotational correlation times (τc) ranging from 5 to 95 ns at a field strength of 600 MHz. Blue triangles represent the R2 of decoupled resonances, and red squares represent those of the narrow TROSY resonances. The points are connected by a dashed line to provide visual clarity.

Supplementary Figure 9 19F-13C TROSY enables 13C-detection with high resolution for large-molecular-weight systems: comparing the TROSY and anti-TROSY components.

a,b, The TROSY cross-peaks in a at 25 °C (e.g., orange peak) are considerably sharper than their anti-TROSY counterparts (e.g., blue peak) in b. The bottom panels show spectra recorded at 10 °C, where the temperature was decreased to mimic the relaxation of a larger-molecular-weight system. c,d, TROSY cross-peaks in c remain sharp and intense (e.g., red peak), whereas in d the corresponding anti-TROSY cross-peaks broaden considerably and are barely visible (e.g., dark blue peak).

Supplementary Figure 10 High-resolution 19F-13C TROSY facilitates the detection of minor conformations.

Low-population cross-peaks in the 19F-13C TROSY spectrum of MBP are shown in green circles. These resonances are likely to have originated from rotamers of 3-19F13C Tyr due to (1) rotation along the Cγ–Cβ bond or (2) rotation along the phenolic Cζ–O bond, which determines the proximity of the hydroxyl hydrogen atom to the fluorine atom. Similarly, slow flipping of the hydroxyl group around the phenolic Cζ–O bond between an intra-residue O-H-F and an inter-residue H-bonding configuration can produce two sets of resonances of unequal intensity. Such rotamers of 3-19F Tyr have been previously observed by X-ray crystallography (J. Mol. Biol. 281, 323–339; 1998). In a few instances, for the very strong cross-peaks, we observed a minor conformation that corresponds to the non-TROSY component along the 13C dimension; this is marked by a red solid rectangle. The non-TROSY component can be recognized by its 19F resonance frequency, which is identical to that of its TROSY counterpart with a 13C frequency separation of ~240 Hz, corresponding to the 1JFC coupling constant.

Supplementary Figure 11 Calculated magnetic field dependence of R2 for 13CF and 19FC nuclei in various 19F-substituted nucleobases.

Blue triangles represent the R2 of decoupled resonances, and red squares represent those of TROSY resonances. The relaxation rates are calculated for the respective 19FC-bearing nucleobase assumed to be incorporated into a nucleic acid with a rotational correlation time (τc) of 25 ns. The points are connected by a dashed line to provide visual clarity.

Supplementary Figure 12 R2 calculated for 13CF, 13CH, 19FC, and 1HC nuclei in 5-19F uracil or uracil as a function of magnetic field strength.

a,b, R2 of 13CF as part of 5-19F uracil (a) and R2 of 5-13CH in uracil (b). c,d, R2 of the uracil 5-19FC uracil (c) and R2 of 5-1HC (d). Here uracil and 5-19F uracil relaxation times are calculated with a rotational correlation time (τc) of 25 ns. Red squares represent the R2 of the TROSY components, and blue triangles represent the R2 of the decoupled resonances. The points are connected by a dashed line to provide visual clarity.

Supplementary Figure 13 1H decoupling improves resolution and sensitivity of the 19F-13C TROSY experiment.

a, 13C-19F TROSY-SE of 3-19F13C Tyr-labeled GB1 without 1H decoupling. Here, 3JCH coupled peaks are further split by the 2JCC into a doublet of doublets along the 13C dimension. This is further emphasized by the vertical 1D trace at –137 p.p.m. on the 19F axis. b, 13C-19F TROSY-SE spectrum of GB1 with 1H decoupling. The 1H coupling is removed, and the peaks are pure doublets in the 13C dimension, owing to the ~7-Hz 2JCC splitting.

Supplementary Figure 14 The 13CF TROSY resonance is the slowest relaxing nucleus in biological NMR.

a,b, Theoretical transverse relaxation rates of 13CF, 15NH and 13CH at the magnetic field strength where the TROSY effect for any given spin-pair is maximized, calculated for (a) a 42-kDa protein (τc at 25 °C = 25 ns) and (b) a 110-kDa protein (τc at 25 °C = 65 ns). Simulated free induction decays (top) compare and visualize the influence of transverse relaxation on the height and line width of the TROSY resonance of the aromatic 13CF, aromatic 13CH and backbone 15NH spin states.

Supplementary information

Supplementary Text and Figures

Supplementary Figs. 1–14, Supplementary Table 1, and Supplementary Notes 1–3

Reporting Summary

Supplementary Software

Pulse programs and parameter sets.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Boeszoermenyi, A., Chhabra, S., Dubey, A. et al. Aromatic 19F-13C TROSY: a background-free approach to probe biomolecular structure, function, and dynamics. Nat Methods 16, 333–340 (2019). https://doi.org/10.1038/s41592-019-0334-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41592-019-0334-x

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