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:

Illuminating the dark conformational space of macrocycles using dominant rotors

Abstract

Three-dimensional conformation is the primary determinant of molecular properties. The thermal energy available at room temperature typically equilibrates the accessible conformational states. Here, we introduce a method for isolating unique and previously understudied conformations of macrocycles. The observation of unusual conformations of 16- to 22-membered rings has been made possible by controlling their interconversion using dominant rotors, which represent tunable atropisomeric constituents with relatively high rotational barriers. Density functional theory and in situ NMR measurements suggest that dominant rotor candidates for the amino-acid-based structures considered here should possess a rotational energy barrier of at least 25 kcal mol−1. Notable differences in the geometries of the macrocycle conformations were identified by NMR spectroscopy and X-ray crystallography. There is evidence that amino acid residues can be forced into rare turn motifs not observed in the corresponding linear counterparts and homodetic rings. These findings should unlock new avenues for studying the conformation–activity relationships of bioactive molecules.

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

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

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

Fig. 1: Landscape remodelling can stabilize conformations in the dark space.
Fig. 2: Dominant rotors as a means to explore complex energy landscapes of constrained chains of rotors.
Fig. 3: Structural properties of the dominant rotor-containing macrocycles.
Fig. 4: Illuminating unusual conformations and properties using dominant rotors.
Fig. 5: Observation of unusual conformations in 22-membered rings enabled by a trimethylated dominant rotor.

Similar content being viewed by others

Data availability

Crystallographic data for the structures reported in this article have been deposited at the Cambridge Crystallographic Data Centre, under deposition numbers CCDC 1956029 (11b), 1956026 (11c), 1956027 (Sa-11e), 1956031 (Ra-11e), 1956032 (Ra-12c), 1956038 (Sa-12c), 1956030 (Sa-13b) and 1956028 (Sa-13c). Copies of the data can be obtained free of charge via https://www.ccdc.cam.ac.uk/structures/. The .xyz files with outputs of all DFT and xtb calculations performed in this study are contained in the .zip file named SI_data.zip, and are also hosted on the following GitHub page: https://github.com/gabegomes/Illuminating-the-dark-conformational-space-using-dominantrotors. The dihedral angle data used to construct the Ramachandran plots of the compounds in this article are reported in Section 1.1.12 of the Supplementary Information. Correspondence and requests for materials should be addressed to A.K.Y.

References

  1. Anslyn, E. V. & Dougherty, D. A. Modern Physical Organic Chemistry. (Univ. Science Books: 2006).

  2. Hammond, G. S. A correlation of reaction rates. J. Am. Chem. Soc. 77, 334–338 (1955).

    Article  CAS  Google Scholar 

  3. Canfield, P. J. et al. A new fundamental type of conformational isomerism. Nat. Chem. 10, 615–624 (2018).

    Article  CAS  PubMed  Google Scholar 

  4. Scherer, G., Kramer, M. L., Schutkowski, M., Reimer, U. & Fischer, G. Barriers to rotation of secondary amide peptide bonds. J. Am. Chem. Soc. 120, 5568–5574 (1998).

    Article  CAS  Google Scholar 

  5. Leroux, F. Atropisomerism, biphenyls, and fluorine: a comparison of rotational barriers and twist angles. ChemBioChem 5, 644–649 (2004).

    Article  CAS  PubMed  Google Scholar 

  6. Brameld, K. A., Kuhn, B., Reuter, D. C. & Stahl, M. Small molecule conformational preferences derived from crystal structure data. A medicinal chemistry focused analysis. J. Chem. Inf. Model. 48, 1–24 (2008).

    Article  CAS  PubMed  Google Scholar 

  7. Hoffmann, R. W. Flexible molecules with defined shape—conformational design. Angew. Chem. Int. Ed. Engl. 31, 1124–1134 (1992).

    Article  Google Scholar 

  8. Golan, O., Cohen, S. & Biali, S. E. cis-syn-cis-1,2,4,5-Tetracyclohexylcyclohexane. A moderately crowded saturated hydrocarbon adopting a twist-boat conformation. J. Org. Chem. 64, 6505–6507 (1999).

    Article  CAS  Google Scholar 

  9. Englander, S. W. & Mayne, L. The nature of protein folding pathways. Proc. Natl Acad. Sci. USA 11, 15873–15880 (2014).

    Article  Google Scholar 

  10. Franco, R., Gil-Caballero, S., Ayala, I., Favier, A. & Brutscher, B. Probing conformation exchange dynamics in a short-lived protein folding intermediate by real-time relaxation–dispersion NMR. J. Am. Chem. Soc. 139, 1065–1068 (2017).

    Article  CAS  PubMed  Google Scholar 

  11. Dyson, H. J. & Wright, P. E. Unfolded proteins and protein folding studied by NMR. Chem. Rev. 104, 3607–3622 (2004).

    Article  CAS  PubMed  Google Scholar 

  12. Zhu, J., Zhu, J. & Springer, T. A. Complete integrin headpiece opening in eight steps. J. Cell Biol. 201, 1053–1068 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Brereton, A. E. & Karplus, P. A. Native proteins trap high-energy transit conformations. Sci. Adv. 1, e1501188 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  14. De Marco, R., Zhao, J., Arianna, G., Ioannone, S. & Gentilucci, L. In-peptide synthesis of imidazolidine-2-one scaffolds, equippable with proteinogenic or taggable/linkable side chains, general promoters of unusual secondary structures. J. Org. Chem. 84, 4992–5004 (2019).

    Article  PubMed  Google Scholar 

  15. Yang, D., Ng, F.-F. & Li, Z.-J. An unusual turn structure in peptides containing α-aminoxy acids. J. Am. Chem. Soc. 118, 9794–9795 (1996).

    Article  CAS  Google Scholar 

  16. Fernández-Tejada, A., Corzana, F., Busto, J. H., Avenoza, A. & Peregrina, J. M. Stabilizing unusual conformations in small peptides and glucopeptides using a hydroxylated cyclobutene amino acid. Org. Biomol. Chem. 7, 2885–2893 (2009).

    Article  PubMed  Google Scholar 

  17. Grotenbreg, G. M. et al. An unusual reverse turn structure adopted by a furanoid sugar amino acid incorporated in gramicidin S. J. Am. Chem. Soc. 126, 3444–3446 (2004).

    Article  CAS  PubMed  Google Scholar 

  18. Roesner, S. et al. Macrocyclisation of small peptides enabled by oxetane incorporation. Chem. Sci. 10, 2465–2472 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Altmayer-Henzien, A., Declerck, V., Guillot, R. & Aitken, D. J. Reactivity of 1-aminoazetidine-2-carboxylic acid during peptide forming procedures: observation of an unusual variant of the hydrazino turn. Tetrahedron Lett. 54, 802–805 (2013).

    Article  CAS  Google Scholar 

  20. Maji, S. K. et al. Peptide design using ω-amino acids: unusual turn structures nucleated by an N-terminal single γ-aminobutryic acid residue in short model peptides. J. Org. Chem. 67, 633–639 (2002).

    Article  CAS  PubMed  Google Scholar 

  21. Fitzkee, N. C. et al. Are proteins made from a limited parts list? Trends Biochem. Sci. 30, 73–80 (2005).

    Article  CAS  PubMed  Google Scholar 

  22. White, C. J. & Yudin, A. K. Contemporary strategies for peptide macrocyclization. Nat. Chem. 3, 509–524 (2011).

    Article  CAS  PubMed  Google Scholar 

  23. Malde, A. K., Hill, T. A., Iyer, A. & Fairlie, D. P. Crystal structures of protein-bound cyclic peptides. Chem. Rev. 119, 9861–9914 (2019).

    Article  CAS  PubMed  Google Scholar 

  24. Appavoo, S. D., Huh, S., Diaz, D. B. & Yudin, A. K. Conformational control of macrocycles by remote structural modification. Chem. Rev. 119, 9724–9752 (2019).

    Article  CAS  PubMed  Google Scholar 

  25. Ikuta, D. et al. Conformationally supple glucose monomers enable synthesis of the smallest cyclodextrins. Science 364, 674–677 (2019).

    Article  CAS  PubMed  Google Scholar 

  26. Wang, Y. et al. A stable silicon(0) compound with a Si=Si double bond. Science 321, 1069–1071 (2008).

    Article  CAS  PubMed  Google Scholar 

  27. Arduengo, A. J. III, Harlow, R. L. & Kline, M. A stable crystalline carbene. J. Am. Chem. Soc. 113, 361–363 (1991).

    Article  CAS  Google Scholar 

  28. Ganesamoorthy, C. et al. A silicon–carbonyl complex stable at room temperature. Nat. Chem. 12, 608–614 (2020).

    Article  CAS  PubMed  Google Scholar 

  29. Stolow, R. D. A quantitative relationship between dissociation constants and conformational equilibria. Cyclohexanecarboxylic acids. J. Am. Chem. Soc. 81, 5806–5811 (1959).

    Article  CAS  Google Scholar 

  30. Sugg, E. E., Griffin, J. F. & Portoghese, P. S. Influence of pseudoallylic strain on the conformational preference of 4-methyl-4-phenylpipecolic acid derivatives. J. Org. Chem. 50, 5032–5037 (1985).

    Article  CAS  Google Scholar 

  31. Coleman, P. J. et al. Discovery of [(2R,5R)-5-{[(5-fluoropyridin-2-yl)oxy]methyl}-2-methylpiperidin-1-yl][5-methyl-2-(pyrimidin-2-yl)phenyl]methanone (MK-6096): a dual orexin receptor antagonist with potentsleep-promoting properties. ChemMedChem 7, 415–424 (2012).

    Article  CAS  PubMed  Google Scholar 

  32. Lam, P. Y. S. et al. Rational design of potent, bioavailable, nonpeptide cyclic ureas as HIV protease inhibitors. Science 263, 380–384 (1994).

    Article  CAS  PubMed  Google Scholar 

  33. Elleraas, J. et al. Conformational studies and atropisomerism kinetics of the ALK clinical candidate lorlatinib (PF-06463922) and desmethyl congeners. Angew. Chem. Int. Ed. 55, 3590–3595 (2016).

    Article  CAS  Google Scholar 

  34. Bragg, R. A., Clayden, J., Morris, G. A. & Pink, J. H. Stereodynamics of bond rotation in tertiary aromatic amides. Chem. Eur. J. 8, 1279–1289 (2002).

    Article  CAS  PubMed  Google Scholar 

  35. Lorentzen, M. et al. Atropisomerism in tertiary biaryl 2-amides: a study of Ar–CO and Ar–Ar′ rotational barriers. J. Org. Chem. 82, 7300–7308 (2017).

    Article  CAS  PubMed  Google Scholar 

  36. Craik, D. J., Kaas, Q. & Wang, C. K. in Practical Medicinal Chemistry with Macrocycles: Design, Synthesis, and Case Studies Vol. 1 (eds Marsault, E. & Peterson, M. L.) Ch. 2 (Wiley, 2017).

  37. Henzler-Wildman, K. & Kern, D. Dynamic personalities of proteins. Nature 450, 964–972 (2007).

    Article  CAS  PubMed  Google Scholar 

  38. Grimme, S. Exploration of chemical compound, conformer, and reaction space with meta-dynamics simulations based on tight-binding quantum chemical calculations. J. Chem. Theory Comput. 15, 2847–2862 (2019).

    Article  CAS  PubMed  Google Scholar 

  39. Hosseinzadeh, P. et al. Comprehensive computational design of ordered peptide macrocycles. Science 358, 1461–1466 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Yudin, A. K. Macrocycles: lessons from the distant past, recent developments, and future directions. Chem. Sci. 6, 30–49 (2015).

    Article  CAS  PubMed  Google Scholar 

  41. Richardson, J. S. The anatomy and taxonomy of protein structure. Adv. Protein Chem. 34, 167–339 (1981).

    Article  CAS  PubMed  Google Scholar 

  42. Rubio-Martinez, J., Tomas, M. S. & Perez, J. J. Effect of the solvent on the conformational behavior of the alanine dipeptide deduced from MD simulations. J. Mol. Graph. Model. 78, 118–128 (2017).

    Article  CAS  PubMed  Google Scholar 

  43. Lovell, S. C. et al. Structure validation by Cα geometry: ϕ,ψ and Cβ deviation. Proteins 50, 437–450 (2003).

    Article  CAS  PubMed  Google Scholar 

  44. Ung, P. & Winkler, D. A. Tripeptide motifs in biology: targets for peptidomimetic design. J. Med. Chem. 54, 1111–1125 (2011).

    Article  CAS  PubMed  Google Scholar 

  45. Niggli, D. A., Ebert, M.-O., Lin, Z., Seebach, D. & van Gunsteren, W. F. Helical content of a β3-octapeptide in methanol: molecular dynamics simulations explain a seeming discrepancy between conclusions derived from CD and NMR data. Chem. Eur. J. 18, 586–593 (2012).

    Article  CAS  PubMed  Google Scholar 

  46. Pavone, V. et al. Discovering protein secondary structures: classification and description of isolated α-turns. Biopolymers 38, 705–721 (1996).

    Article  CAS  PubMed  Google Scholar 

  47. Chou, K.-C. Prediction and classification of α-turn types. Biopolymers 42, 837–853 (1997).

    Article  CAS  PubMed  Google Scholar 

  48. Weinhold, F., Landis, C. R. & Glendening, E. D. What is NBO analysis and how is it useful? Int. Rev. Phys. Chem. 35, 399–440 (2016).

    Article  CAS  Google Scholar 

  49. Newberry, R. W. & Raines, R. T. The n→π* interaction. Acc. Chem. Res. 50, 1838–1846 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Wenzell, N. A. et al. Electronic and steric control of n→π* interactions: stabilization of the α-helix conformation without a hydrogen bond. ChemBioChem 20, 963–967 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Fernández, A. Keeping dry and crossing membranes. Nat. Biotechnol. 22, 1081–1084 (2004).

    Article  PubMed  Google Scholar 

  52. Rajashankar, K. R. & Ramakumar, S. π-Turns in proteins and peptides: classification, occurrence, hydration and sequence. Protein Sci. 5, 932–946 (1996).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Dasgupta, B. & Chakrabarti, P. pi-Turns: types, systematics and the context of their occurrence in protein structures. BMC Struct. Biol. 8, 39 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  54. VnmrJ 3.2 (Agilent Technologies, 2011).

  55. Frost, J. R., Scully, C. C. G. & Yudin, A. K. Oxadiazole grafts in peptide macrocycles. Nat. Chem. 8, 1105–1111 (2016).

    Article  CAS  PubMed  Google Scholar 

  56. Frisch, J. M. et al. Gaussian 16, Revision C.01 (Gaussian, 2016).

  57. O’Boyle, N. M. et al. Open Babel: an open chemical toolbox. J. Cheminform. 3, 33 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  58. Open Babel, Version 2.3.1 (2011); http://openbabel.org/

  59. Bannwarth, C., Ehlert, S. & Grimme, S. GFN2-xTB—an accurate and broadly parametrized self-consistent tight-binding quantum chemical method with multipole electrostatics and density-dependent dispersion contributions. J. Chem. Theory Comput. 15, 1652–1671 (2019).

    Article  CAS  PubMed  Google Scholar 

  60. Scalmani, G. & Frisch, M. J. Continuous surface charge polarizable continuum models of solvation. I. General formalism. J. Chem. Phys. 132, 114110 (2010).

    Article  PubMed  Google Scholar 

  61. Becke, A. D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 98, 5648–5652 (1993).

    Article  CAS  Google Scholar 

  62. Lee, C., Yang, W. & Parr, R. G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 37, 785–789 (1988).

    Article  CAS  Google Scholar 

  63. Grimme, S., Antony, J., Ehrlich, S. & Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 132, 154104 (2010).

    Article  PubMed  Google Scholar 

  64. Grimme, S., Ehrlich, S. & Goerigk, L. Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem 32, 1456–1465 (2011).

    Article  CAS  PubMed  Google Scholar 

  65. Luchini, G., Alegre-Requena, J. V., Funes-Ardoiz, I. & Paton, R. S. GoodVibes: automated thermochemistry for heterogeneous computational chemistry data [version 1; peer review: 2 approved with reservations]. F1000Res. 9, 291 (2020).

    Article  Google Scholar 

  66. Glendening, E. D., Landis, C. R. & Weinhold, F. NBO 6.0: natural bond orbital analysis program. J. Comput. Chem. 34, 1429–1437 (2013).

    Article  CAS  PubMed  Google Scholar 

  67. Chemcraft, Version 1.8 (2019); https://www.chemcraftprog.com

  68. Legault, C. Y., CYLview, 1.0b (Université de Sherbrooke, 2009); http://www.cylview.org

  69. The PyMOL Molecular Graphics System, Version 2.3.4 (Schrödinger, 2019).

  70. Pedregal, J. R.-G., Gómez-Orellana, P. & Maréchal, J.-D. ESIgen: electronic supporting information generator for computational chemistry publications. J. Chem. Inf. Model. 58, 561–564 (2018).

    Article  Google Scholar 

  71. Clayden, J., Moran, W. J., Edwards, P. J. & LaPlante, S. R. The challenge of atropisomerism in drug discovery. Angew. Chem. Int. Ed. 48, 6398–6401 (2009).

    Article  CAS  Google Scholar 

  72. LaPlante, S. R. et al. Assessing atropisomer axial chirality in drug discovery and development. J. Med. Chem. 54, 7005–7022 (2011).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank the Natural Sciences and Engineering Research Council (NSERC) for financial support. D.B.D. and S.D.A. thank the NSERC for PGS-D funding. G.P.G. thanks the NSERC for a Banting Postdoctoral Fellowship. We thank A. J. Lough for X-ray structure determination and D. C. Burns of the CSICOMP NMR facility for assistance with spectroscopic experiments. We thank Compute Canada for computational resources. DFT and NBO computations were performed on the Niagara supercomputer at the SciNet HPC Consortium. SciNet is funded by the Canada Foundation for Innovation, the Government of Ontario, Ontario Research Fund - Research Excellence and the University of Toronto. Helpful discussions with R. Mendoza-Sanchez, C. C. G. Scully, A. Holownia, S. K. Liew, H. S. Soor, C. N. Apte and A. L. Roughton are greatly appreciated. We also thank A. Aspuru-Guzik, K. Z. Demmans, S. C. Khojasteh, L. A. Dutra, F. Sprang and B. Hamzaev for their help on ongoing projects related to the use of the dominant rotors described here.

Author information

Authors and Affiliations

Authors

Contributions

A.K.Y. and D.B.D. conceived the study and designed the experiments; D.B.D., A.F.B., Y.L. and T.J.M. carried out macrocycle synthesis, purification and characterization; D.B.D. obtained and analysed the kinetic and thermodynamic data; S.D.A. and D.B.D. computed the NMR-based structures and performed unrestrained MD simulations; Y.L. and D.B.D. obtained X-ray quality crystals of the macrocycles; G.P.G. performed DFT and NBO calculations; D.B.D. and A.K.Y. wrote the paper with help from all the authors.

Corresponding author

Correspondence to Andrei K. Yudin.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Chemistry thanks the anonymous reviewers for their contribution to the peer review of this work.

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

Extended data

Extended Data Fig. 1 Characterization of dominant rotor-containing macrocyclic peptides.

a, 2D Exchange spectroscopy of 11b (top) and 11c (bottom) in DMSO-d6 at 25 °C. The negative cross-peak in the spectrum of 11b suggests an exchange between two conformers on the NMR time scale due to atropisomerism at the biaryl rotor71,72. b, LC-MS chromatograms of linear starting material Fmoc-Trz-AGF (top), Ra-12c (middle) and Sa-12c (bottom) atropdiastereomeric macrocycles.

Extended Data Fig. 2 Hydrogen-deuterium exchange of the backbone amide groups of Sa- and Ra-11e.

Differences in amide NH hydrogen-deuterium exchange rates and chemical shift/temperature coefficients show that dominant rotors can stabilize distinct solution conformations.

Extended Data Fig. 3 Contrasting the structural features of dominant rotor peptides and homodetic counterparts.

a, Hydrogen-bonding profile and temperature-chemical shift coefficients (Tcoeff) of homodetic peptide 15 and dominant rotor 12c macrocycles, based on the AGF sequence, as measured by VT 1H NMR spectroscopy. Hydrogen atoms engaged in intramolecular hydrogen bonding and their corresponding NMR signals are highlighted here by grey rectangles. b, Frequency histograms of the AGF conformation over the unrestrained MD trajectory revealed a large population shift in 15 relative to 12c. The macrocycle flexibility in the dominant rotor peptides was assessed computationally and deduced to be minimal, with average backbone atom deviations below 0.2 Å. Both Ra- and Sa-12c structures were stabilized by 13-membered hydrogen bonds with lengths (Å) of 2.3 ± 0.7 and 2.4 ± 1.0, respectively, over the course of the 100ns MD trajectories. An overlay of the backbone atoms for the 10 lowest energy clusters in each peptide shows that dominant rotor peptide conformations are closely matched and are well-behaved in solution. Side chains are omitted for clarity. c, An overlay of the X-ray crystal structure (grey) and NMR solution structure (blue) of compound Ra-12c show a close overlap between the solid- and solution-phase structures. d, Energetic contribution of hydrogen bonds to conformational stabilization in Ra-12c. Hydrogen bond distances (Å): α, 1.85; β, 2.71. Second-order stabilization energies (in kcal mol−1): α, 9.1; β, < 0.5. NBO analysis of nOσ*NH interactions between the dominant rotor carbonyl and transannular amides. e, The NBO interaction of the nOπ*CO between the dominant rotor and Ala1 was calculated to be 2.8 kcal mol−1 with an O•••C distance of 2.65 Å and OCO angle of 103.4º.

Extended Data Fig. 4 Generality of hydrogen-bonding patterns across dominant rotor-containing macrocycles.

a, Structural scope of dominant rotor peptides with Ra-(red) and Sa-(grey) average backbone solution structures superimposed on the three consecutive natural amino acid residues. Root mean-squared deviation (RMSD) shown for macrocycle backbone atoms. The type II αRU-turn was detected in every Ra-well except for 12h, which exhibited a left-handed helical turn. Apart from 12g, the Sa-wells are variable and access a variety of different turn types and conformations containing non-hydrogen bonded amides. Dominant rotor omitted for clarity. Gibbs free energy differences in kcal mol−1 are shown in brackets.

Extended Data Fig. 5 Choosing bis-methylated dominant rotors can result in additional differentiation of the observed conformational states.

Replacement of the mono-methyl dominant rotor in 12c with a bis-methylated rotor disrupts α-turn formation and stabilizes a non-classical β-turn type in Ra-12d (grey). This conformational change in the Ra-well is supported by a decrease in the coupling constant (3JHNCH) of the alanine NH residue from 5.3 to 2.3 Hz. The increase from 4.7 to 7.7 Hz in the Sa-well also indicates a change in the alanine ϕ dihedral angle of Sa-12d (green). Ramachandran plots of the three α-carbon atoms in the AGF sequence of the 10 lowest energy clusters for each dominant rotor peptide show that the Ra-12d conformer is less scattered than Sa-12d, and therefore more rigid.

Extended Data Fig. 6 Observation of unusual conformations in 22-membered rings.

Arrow diagram of the PGLGF (proline-glycine-leucine-glycine-phenylalanine) sequence shows that Gly2 and Phe5 adopt vastly different backbone conformations in two-well system 14c. The NMR-solution structure of Ra-14c (cyan) shows that the dominant rotor carbonyl engages the Gly2 amide NH in an inverse γ-turn, which enforces a type I-αLS turn for the LGF segment. Sa-14c (green) adopts a novel π-turn with an internal β-turn that is stabilized by a hydrogen bond between the carbonyl oxygen and amide NH of the dominant rotor. VT NMR studies are in agreement with these unique hydrogen-bond patterns.

Supplementary information

Supplementary Information

Supplementary Methods, Discussion, Figs. 1–13, Tables 1 and 2, Appendix 1: NMR spectra, Appendix 2: X-ray coordinate data, Appendix 3: torsion scan coordinate data, Appendix 4: NBO coordinate data.

Supplementary Data 1

.xyz files of the geometries obtained from DFT calculations.

Supplementary Data 2

Crystallographic data for compound 11c. CCDC reference 1956026.

Supplementary Data 3

Crystallographic data for compound Sa-11e. CCDC reference 1956027.

Supplementary Data 4

Crystallographic data for compound Sa-13c. CCDC reference 1956028.

Supplementary Data 5

Crystallographic data for compound 11b. CCDC reference 1956029.

Supplementary Data 6

Crystallographic data for compound Sa-13b. CCDC reference 1956030.

Supplementary Data 7

Crystallographic data for compound Ra-11e. CCDC reference 1956031.

Supplementary Data 8

Crystallographic data for compound Ra-12c. CCDC reference 1956032.

Supplementary Data 9

Crystallographic data for compound Sa-12c. CCDC reference 1956038.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Diaz, D.B., Appavoo, S.D., Bogdanchikova, A.F. et al. Illuminating the dark conformational space of macrocycles using dominant rotors. Nat. Chem. 13, 218–225 (2021). https://doi.org/10.1038/s41557-020-00620-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41557-020-00620-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