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:

RETRACTED ARTICLE: Intrinsically unidirectional chemically fuelled rotary molecular motors

This article was retracted on 10 July 2023

This article has been updated

Abstract

Biological systems mainly utilize chemical energy to fuel autonomous molecular motors, enabling the system to be driven out of equilibrium1. Taking inspiration from rotary motors such as the bacterial flagellar motor2 and adenosine triphosphate synthase3, and building on the success of light-powered unidirectional rotary molecular motors4,5,6, scientists have pursued the design of synthetic molecular motors solely driven by chemical energy7,8,9,10,11,12,13. However, designing artificial rotary molecular motors operating autonomously using a chemical fuel and simultaneously featuring the intrinsic structural design elements to allow full 360° unidirectional rotary motion like adenosine triphosphate synthase remains challenging. Here we show that a homochiral biaryl Motor-3, with three distinct stereochemical elements, is a rotary motor that undergoes repetitive and unidirectional 360° rotation of the two aryl groups around a single-bond axle driven by a chemical fuel. It undergoes sequential ester cyclization, helix inversion and ring opening, and up to 99% unidirectionality is realized over the autonomous rotary cycle. The molecular rotary motor can be operated in two modes: synchronized motion with pulses of a chemical fuel and acid–base oscillations; and autonomous motion in the presence of a chemical fuel under slightly basic aqueous conditions. This rotary motor design with intrinsic control over the direction of rotation, simple chemical fuelling for autonomous motion and near-perfect unidirectionality illustrates the potential for future generations of multicomponent machines to perform mechanical functions.

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: Schematic representation of the design and concept of the continuously rotary molecular motor.
Fig. 2: Verifying the unidirectional 360° rotation of the molecular motor by desymmetrization of the lower stator.
Fig. 3: Synchronized 360° rotation with pulses of chemical fuel and acid–base oscillations.
Fig. 4: Continuous rotation of the molecular motor.

Similar content being viewed by others

Data availability

Details on the procedures, characterization and references, including spectra of new compounds and compounds made using the reported method, are available in Supplementary Information. Crystallographic data for (S,S')-Motor-1-Me, (R,S')-3-Me, Motor-1-K, Motor-1-T-MOM and Motor-Br-Me can be obtained free of charge from www.ccdc.cam.ac.uk under CCDC deposition numbers 2170186, 2170187, 2170188, 2170189 and 2170190.

Change history

  • 22 December 2022

    Editor’s Note: Readers are alerted that the authors have discovered that the motion of molecular motor reported in this paper appears to be very sensitive to pH change, which occurs during fuel consumption. A further editorial response will follow the based on the investigation of this issue.

  • 10 July 2023

    This article has been retracted. Please see the Retraction Notice for more detail: https://doi.org/10.1038/s41586-023-06380-2

References

  1. Saper, G. & Hess, H. Synthetic systems powered by biological molecular motors. Chem. Rev. 120, 288–309 (2020).

    Article  CAS  PubMed  Google Scholar 

  2. Sowa, Y. et al. Direct observation of steps in rotation of the bacterial flagellar motor. Nature 437, 916–919 (2005).

    Article  CAS  PubMed  ADS  Google Scholar 

  3. Boyer, P. D. The ATP synthase—a splendid molecular machine. Annu. Rev. Biochem. 66, 717–749 (1997).

    Article  CAS  PubMed  Google Scholar 

  4. Wang, J. & Feringa, B. L. Dynamic control of chiral space in a catalytic asymmetric reaction using a molecular motor. Science 331, 1429–1432 (2011).

    Article  CAS  PubMed  ADS  Google Scholar 

  5. Roke, D., Wezenberg, S. J. & Feringa, B. L. Molecular rotary motors: unidirectional motion around double bonds. Proc. Natl Acad. Sci. USA 115, 9423–9431 (2018).

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  6. Browne, W. R. & Feringa, B. L. Making molecular machines work. Nat. Nanotechnol. 1, 25–35 (2006).

    Article  CAS  PubMed  ADS  Google Scholar 

  7. Kelly, T. R. et al. Progress toward a rationally designed, chemically powered rotary molecular motor. J. Am. Chem. Soc. 129, 376–386 (2007).

    Article  CAS  PubMed  Google Scholar 

  8. Anelli, P. L., Spencer, N. & Stoddart, J. F. A molecular shuttle. J. Am. Chem. Soc. 113, 5131–5133 (1991).

    Article  CAS  PubMed  Google Scholar 

  9. Badjić, J. D., Balzani, V., Credi, A., Silvi, S. & Stoddart, J. F. A molecular elevator. Science 303, 1845–1849 (2004).

    Article  PubMed  ADS  Google Scholar 

  10. Cárdenas, D. J., Livoreil, A. & Sauvage, J.-P. Redox control of the ring-gliding motion in a Cu-complexed catenane: a process involving three distinct geometries. J. Am. Chem. Soc. 118, 11980–11981 (1996).

    Article  Google Scholar 

  11. Jiménez, M. C., Dietrich-Buchecker, C. & Sauvage, J.-P. Towards synthetic molecular muscles: contraction and stretching of a linear rotaxane dimer. Angew. Chem. Int. Ed. 39, 3284–3287 (2000).

    Article  ADS  Google Scholar 

  12. Silvi, S., Venturi, M. & Credi, A. Artificial molecular shuttles: from concepts to devices. J. Mater. Chem. 19, 2279–2294 (2009).

    Article  CAS  Google Scholar 

  13. Stevens, A. M. & Richards, C. J. A metallocene molecular gear. Tetrahedron Lett. 38, 7805–7808 (1997).

    Article  CAS  Google Scholar 

  14. Geeves, M. A. & Holmes, K. C. Structural mechanism of muscle contraction. Annu. Rev. Biophys. 68, 687–728 (1999).

    CAS  Google Scholar 

  15. Hua, W., Chung, J. & Gelles, J. Distinguishing inchworm and hand-over-hand processive kinesin movement by neck rotation measurements. Science 295, 844–848 (2002).

    Article  CAS  PubMed  ADS  Google Scholar 

  16. Vale, R. D. The molecular motor toolbox for intracellular transport. Cell 112, 467–480 (2003).

    Article  CAS  PubMed  Google Scholar 

  17. Erbas-Cakmak, S., Leigh, D. A., McTernan, C. T. & Nussbaumer, A. L. Artificial molecular machines. Chem. Rev. 115, 10081–10206 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. van Dijk, L. et al. Molecular machines for catalysis. Nat. Rev. Chem. 2, 0117 (2018).

    Article  Google Scholar 

  19. Aprahamian, I. The future of molecular machines. ACS Cent. Sci. 6, 347–358 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Astumian, R. D. Kinetic asymmetry allows macromolecular catalysts to drive an information ratchet. Nat. Commun. 10, 3837 (2019).

    Article  PubMed  PubMed Central  ADS  Google Scholar 

  21. Koumura, N., Zijlstra, R. W. J., van Delden, R. A., Harada, N. & Feringa, B. L. Light-driven monodirectional molecular rotor. Nature 401, 152–155 (1999).

    Article  CAS  PubMed  ADS  Google Scholar 

  22. Kelly, T. R., De Silva, H. & Silva, R. A. Unidirectional rotary motion in a molecular system. Nature 401, 150–152 (1999).

    Article  CAS  PubMed  ADS  Google Scholar 

  23. Greb, L. & Lehn, J.-M. Light-driven molecular motors: imines as four-step or two-step unidirectional rotors. J. Am. Chem. Soc. 136, 13114–13117 (2014).

    Article  CAS  PubMed  Google Scholar 

  24. Erbas-Cakmak, S. et al. Rotary and linear molecular motors driven by pulses of a chemical fuel. Science 358, 340–343 (2017).

    Article  CAS  PubMed  ADS  Google Scholar 

  25. Bach, N. N., Josef, V., Maid, H. & Dube, H. Active mechanical threading by a molecular motor. Angew. Chem. Int. Ed. 61, e202201882 (2022).

    CAS  Google Scholar 

  26. Fletcher, S. P., Dumur, F., Pollard, M. M. & Feringa, B. L. A reversible, unidirectional molecular rotary motor driven by chemical energy. Science 310, 80–82 (2005).

    Article  CAS  PubMed  ADS  Google Scholar 

  27. Collins, B. S. L., Kistemaker, J. C. M., Otten, E. & Feringa, B. L. A chemically powered unidirectional rotary molecular motor based on a palladium redox cycle. Nat. Chem. 8, 860–866 (2016).

    Article  CAS  Google Scholar 

  28. Zhang, Y. et al. A chemically driven rotary molecular motor based on reversible lactone formation with perfect unidirectionality. Chem 6, 2420–2429 (2020).

    Article  CAS  Google Scholar 

  29. Corra, S. et al. Artificial supramolecular pumps powered by light. Chem. Eur. J. 27, 11076–11083 (2021).

    Article  CAS  PubMed  Google Scholar 

  30. Guentner, M. et al. Sunlight-powered kHz rotation of a hemithioindigo-based molecular motor. Nat. Commun. 6, 8406 (2015).

    Article  CAS  PubMed  ADS  Google Scholar 

  31. Koumura, N., Geertsema, E. M., van Gelder, M. B., Meetsma, A. & Feringa, B. L. Second generation light-driven molecular motors. Unidirectional rotation controlled by a single stereogenic center with near-perfect photoequilibria and acceleration of the speed of rotation by structural modification. J. Am. Chem. Soc. 124, 5037–5051 (2002).

    Article  CAS  PubMed  Google Scholar 

  32. Lubbe, A. S. et al. Photoswitching of DNA hybridization using a molecular motor. J. Am. Chem. Soc. 140, 5069–5076 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Pooler, D. R. S., Lubbe, A. S., Crespi, S. & Feringa, B. L. Designing light-driven rotary molecular motors. Chem. Sci. 12, 14964–14986 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Štacko, P. et al. Locked synchronous rotor motion in a molecular motor. Science 356, 964–968 (2017).

    Article  PubMed  ADS  Google Scholar 

  35. Zhao, D., van Leeuwen, T., Cheng, J. & Feringa, B. L. Dynamic control of chirality and self-assembly of double-stranded helicates with light. Nat. Chem. 9, 250–256 (2016).

    Article  PubMed  Google Scholar 

  36. Wilson, M. R. et al. An autonomous chemically fuelled small-molecule motor. Nature 534, 235–240 (2016).

    Article  CAS  PubMed  ADS  Google Scholar 

  37. Borsley, S., Kreidt, E., Leigh, D. A. & Roberts, B. M. W. Autonomous fuelled directional rotation about a covalent single bond. Nature 604, 80–85 (2022).

    Article  CAS  PubMed  ADS  Google Scholar 

  38. Bringmann, G. et al. Atroposelective synthesis of axially chiral biaryl compounds. Angew. Chem. Int. Ed. 44, 5384–5427 (2005).

    Article  CAS  Google Scholar 

  39. Seeman, J. I. The Curtin–Hammett principle and the Winstein–Holness equation: new definition and recent extensions to classical concepts. J. Chem. Educ. 63, 42–48 (1986).

    Article  CAS  Google Scholar 

  40. Koumura, N., Geertsema, E. M., Meetsma, A. & Feringa, B. L. Light-driven molecular rotor: unidirectional rotation controlled by a single stereogenic center. J. Am. Chem. Soc. 122, 12005–12006 (2000).

    Article  CAS  Google Scholar 

  41. Aikawa, K., Miyazaki, Y. & Mikami, K. Stable axial chirality in metal complexes bearing 4,4′-substituted BIPHEPs: application to catalytic asymmetric carbon–carbon bond-forming reactions. Bull. Chem. Soc. Jpn 85, 201–208 (2012).

    Article  CAS  Google Scholar 

  42. Mori, K. et al. Enantioselective synthesis of multisubstituted biaryl skeleton by chiral phosphoric acid catalyzed desymmetrization/kinetic resolution sequence. J. Am. Chem. Soc. 135, 3964–3970 (2013).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We are grateful for the support of this work by the National Natural Science Foundation of China (21971267), the Fundamental Research Funds for the Central Universities, Sun Yat-sen University (22lgqb34) and the programme for Guangdong Introducing Innovative and Entrepreneurial Teams (2017ZT07C069). B.L.F. acknowledges financial support from the European Research Council (ERC; advanced grant number 694345 to B.L.F.) and the Dutch Ministry of Education, Culture and Science (Gravitation programme number 024.001.035).

Author information

Authors and Affiliations

Authors

Contributions

D.Z., K.M., Y.Z. and B.L.F. conceived the project. K.M., Y.Z., Z.D. and X.M. carried out the synthesis and characterized the motion of the molecular motor. K.M. and Y.Y. performed all XRD measurements and structural analysis. D.Z. and B.L.F. guided the research; K.M., Y.Z., D.Z. and B.L.F wrote the manuscript. All authors discussed the results and commented on the manuscript.

Corresponding authors

Correspondence to Ben L. Feringa or Depeng Zhao.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature thanks He Tian and the other, anonymous, reviewer(s) 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.

This article has been retracted. Please see the retraction notice for more detail: https://doi.org/10.1038/s41586-023-06380-2

Extended data figures and tables

Extended Data Fig. 1 Single-crystal structures of 3-Me, Motor-1-Me, Motor-1-K and MOM-derived thermodynamic product (Motor-1-T-MOM).

(a) The absolute configuration of molecular motors was unambiguously established by single-crystal X-ray diffraction as well (Here, the S or R denotes the configuration of the chiral centre at the α position of carboxylic acid group, S’ or R’ denotes the configuration of the chiral centre of the benzyl position). Single crystal X-ray diffraction experiments of the corresponding methyl esters 3-Me and Motor-1-Me were performed to gain insight into the relationship between configurations and cyclization selectivity. The results indicated the orientation of the carboxylic group is key to realize high selectivity of cyclization. Viewed along the biaryl C–C bond (the axle of rotation), the carboxylic group in (S,S’)-syn-isomer−1 favours bridge formation with the hydroxyl group on the right of the stator aryl to generate exclusively the kinetic isomer. However, the methyl group at the α-position of the carboxylic group in (R,S’)-anti-isomer 3 blocked the carboxylic group from reacting with hydroxyl group on the right side of the stator aryl and therefore cyclized with the left hydroxyl group to give dominantly a thermodynamic isomer. (b) Single-crystal X-Ray diffraction of Motor-1-K and MOM-protected Motor-1-T-MOM confirmed that the rotational direction of Motor-1 was clockwise. It is interesting to note the favoured tub-shaped conformation of the eight-membered ring in Motor-1-T and Motor-1-K. In Motor-1-K, the methyl group adopts a pseudoaxial orientation while the methoxy group is located in a pseudoequatorial position. However, in the case of Motor-1-T, the methoxy group is in a pseudoaxial orientation and the methyl group adopts a pseudoequatorial position. The eight-membered ring flipping can be explained by Winstein–Holness A value which is used to describe the conformational preference of an equatorial compared to an axial substituent in a substituted cyclohexane. As the A value of a methyl group (1.7 kcal/mol) is larger than the A value of a methoxy group (0.6 kcal/mol), the methyl group has a strong tendency for the equatorial position which also drives the helix inversion of the biaryl and the conversion of Motor-1-K to Motor-1-T (P to M).

Extended Data Fig. 2 1H NMR spectra of three stations of Motor-1.

1H NMR spectra of three states (Motor-1, Motor-1-K, Motor-1-T) during the 180° unidirectional rotation. In the first step, Motor-1 was cyclized to give a high-energy isomer Motor-1-K in the presence of the fuel EDC in DCM at 0 °C and this is the only energy input step in the half cycle of rotation. Motor-1-K unstable isomer then underwent an energetic downhill helix inversion to yield the Motor-1-T with a final ratio of T/K = 4/1. Analysis of the kinetic data provided the Gibbs free energy of activation ΔG° = 103.6 kJ/mol (t1/2 = 45 h at 298.15 K) (Supplementary Fig. 9).

Supplementary information

Supplementary Information

This Supplementary Information file contains 17 sections and includes 40 figures and 5 tables.

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.

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mo, K., Zhang, Y., Dong, Z. et al. RETRACTED ARTICLE: Intrinsically unidirectional chemically fuelled rotary molecular motors. Nature 609, 293–298 (2022). https://doi.org/10.1038/s41586-022-05033-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-022-05033-0

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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