Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect
ACS Publications. Most Trusted. Most Cited. Most Read
Single-Step Enantioselective Synthesis of Mechanically Planar Chiral [2]Rotaxanes Using a Chiral Leaving Group Strategy
My Activity

Figure 1Loading Img
  • Open Access
Article

Single-Step Enantioselective Synthesis of Mechanically Planar Chiral [2]Rotaxanes Using a Chiral Leaving Group Strategy
Click to copy article linkArticle link copied!

  • Chong Tian
    Chong Tian
    Department of Chemistry, University of Manchester, Oxford Road, Manchester M13 9PL, United Kingdom
    More by Chong Tian
  • Stephen D. P. Fielden
    Stephen D. P. Fielden
    Department of Chemistry, University of Manchester, Oxford Road, Manchester M13 9PL, United Kingdom
  • Borja Pérez-Saavedra
    Borja Pérez-Saavedra
    Department of Chemistry, University of Manchester, Oxford Road, Manchester M13 9PL, United Kingdom
  • Iñigo J. Vitorica-Yrezabal
    Iñigo J. Vitorica-Yrezabal
    Department of Chemistry, University of Manchester, Oxford Road, Manchester M13 9PL, United Kingdom
  • David A. Leigh*
    David A. Leigh
    Department of Chemistry, University of Manchester, Oxford Road, Manchester M13 9PL, United Kingdom
    School of Chemistry and Molecular Engineering, East China Normal University, 200062 Shanghai, China
    *E-mail: [email protected]
Open PDFSupporting Information (2)

Journal of the American Chemical Society

Cite this: J. Am. Chem. Soc. 2020, 142, 21, 9803–9808
Click to copy citationCitation copied!
https://doi.org/10.1021/jacs.0c03447
Published May 1, 2020

Copyright © 2020 American Chemical Society. This publication is licensed under CC-BY.

Abstract

Click to copy section linkSection link copied!

We report a one-step enantioselective synthesis of mechanically planar chiral [2]rotaxanes. Previous studies of such molecules have generally involved the separation of enantiomers from racemic mixtures or the preparation and separation of diastereomeric intermediates followed by post-assembly modification to remove other sources of chirality. Here, we demonstrate a simple asymmetric metal-free active template rotaxane synthesis using a primary amine, an activated ester with a chiral leaving group, and an achiral crown ether lacking rotational symmetry. Mechanically planar chiral rotaxanes are obtained directly in up to 50% enantiomeric excess. The rotaxanes were characterized by NMR spectroscopy, high-resolution mass spectrometry, chiral HPLC, single crystal X-ray diffraction, and circular dichroism. Either rotaxane enantiomer could be prepared selectively by incorporating pseudoenantiomeric cinchona alkaloids into the chiral leaving group.

Copyright © 2020 American Chemical Society

Introduction

Click to copy section linkSection link copied!

Mechanical planar chirality arises in rotaxanes with achiral components when an unsymmetrical axle is threaded through a macrocycle lacking rotational symmetry (Figure 1). (1−4) Although lacking classical elements of chirality, studies on mechanically planar chiral rotaxanes suggest their asymmetry can be well expressed for applications. (5−7) However, despite mechanically planar chiral rotaxanes being known for nearly 50 years, their enantioselective synthesis remains challenging. (1d,8) Most studies on these systems rely on the separation of enantiomers from racemic mixtures by chiral stationary phase HPLC, limiting the scale of enantioenriched material that can readily be obtained. (9)

Figure 1

Figure 1. Enantioselective synthesis of mechanically planar chiral rotaxanes through metal-free active template N-acylation using a macrocycle lacking rotational symmetry and an electrophile with a point-chiral leaving group.

Goldup et al. have addressed this synthetic problem through a chiral auxiliary approach that forms intermediate diastereomeric rotaxanes having both point chirality and mechanically planar chirality. (10,11) Separation of these diastereomeric intermediates by flash chromatography, followed by removal of the point chirality by either substitution (10) or symmetrization, (11) afforded enantioenriched mechanically planar chiral rotaxanes. The only single-step synthesis of enantioenriched mechanically planar chiral rotaxanes to date used a chiral catalyst to resolve the interconverting enantiomers of a crown ether-ammonium pseudorotaxane by capping. (12) Despite attempts to optimize this method, it produced rotaxanes in just 4% enantiomeric excess (e.e.). Here we report a simple, single-step, enantioselective synthesis of mechanically planar chiral rotaxanes that produces either enantiomer in up to 50% e.e.
Metal-free active template reactions have recently been developed in which rotaxanes (13) are spontaneously assembled under kinetic control in a single step by combining a primary amine, electrophile, and crown ether (14) in apolar solvents. Crown ethers stabilize the transition states of various nucleophilic substitution reactions through the cavity by C–H hydrogen bonding, thereby favoring the formation of rotaxanes over the unthreaded axle. Different reactions, amines, and leaving groups result in different degrees of accelerated reaction through the ring, affording different rotaxane:thread selectivities. We chose crown ether-stabilized N-acylation for the present study (Scheme 1), as this active template reaction often results in a particularly high ratio of rotaxane:thread. (14) This suggested the reaction might be tolerant of the additional functionality necessary in the macrocycle (to break rotational symmetry) and axle building blocks (to provide a chiral leaving group).

Scheme 1

Scheme 1. : (a) Achiral Rotaxane Synthesis by Active Template N-Acylation Using Rotationally Symmetrical Crown Ethers; (b) Racemic and Unoptimized Enantioselective Synthesis of a Mechanically Planar Chiral Rotaxane
An example of an active template N-acylation is the reaction of 24-crown-8 1, primary amine 2, and electrophile 3 in toluene at room temperature, producing amide [2]rotaxane 4 in 84% yield (Scheme 1a). (14b) The rate-determining step of crown ether catalyzed N-acylation reactions is the collapse of the tetrahedral intermediate formed on addition of the amine to the activated ester. (15) The nitro-phenol ester used in the reaction of 1, 2, and 3 thus provides an opportunity for a chiral directing group to be incorporated into the leaving group that could interact with a rotationally unsymmetrical macrocycle in the transition state (Figure 1). (16)

Results and Discussion

Click to copy section linkSection link copied!

Development of an Enantioselective Rotaxane Synthesis

To establish that functionalized crown ethers could take part in the active template reaction, amine 2 and activated ester 3 were treated with commercially available dibenzo-24-crown-8 (5) in toluene, yielding the corresponding [2]rotaxane, 6, in 73% yield (Scheme 1a). However, although the rotaxane axle is unsymmetrical, dibenzo-24-crown-8 (5) is D2h symmetric and so rotaxane 6 is achiral. (1) Macrocycle 7, containing two different aromatic rings, lacks rotational symmetry (it has C1h symmetry, alternatively referred to as Cs). Reaction of 7 with 2 and 3 furnished racemic mechanically planar chiral rotaxane 8 in 78% yield (Scheme 1b). The enantiomers of 8 could be separated by chiral stationary phase HPLC (see Supporting Information).
Next, we investigated the structure and location for an effective chiral leaving group in the electrophile. Preliminary screening studies identified nitrophenol ester 9, in which the chiral information stems from an O-alkylated cinchonidine unit adjacent to the nitro-group (Scheme 1b). This electrophile was reactive under the rotaxane-forming conditions despite the introduction of the deactivating electron-donating ether linkage. Combining 2, 7, and 9 in a 1:1:1 stoichiometry in toluene at room temperature afforded rotaxane 8 in 43% yield (Scheme 1b). Under similar conditions, electrophiles based on alkyl (thio)esters or with the cinchonidine unit positioned at the ortho position of the nitrophenol ring were either unreactive or generated less rotaxane (see Supporting Information). HPLC analysis of rotaxane 8 (isolated by flash chromatography) obtained from electrophile 9 revealed that the (+)-enantiomer (determined by polarimetry) had been formed in 12% e.e., confirming that a point-chiral leaving group was able to induce enantioselectivity of a mechanically planar rotaxane product.
Increasing the electronic difference between the two aromatic substituents within the macrocycle improved the enantioselectivity of the active template reaction. Macrocycle 10, with a nitro group on the catechol unit (see Supporting Information for its synthesis), afforded rotaxane (+)-11 in 23% e.e. at room temperature, which increased to 40% e.e. (55% yield) when the rotaxane-forming reaction was performed at −40 °C (Figure 2a). Lowering the reaction temperature beyond −40 °C did not result in further improvements in enantioselectivity. (17)

Figure 2

Figure 2. (a) Enantioselective synthesis of mechanically planar chiral rotaxane 11. Reaction conditions: 2 equiv of amine 2, 1 equiv each of electrophile and crown ether 10, toluene, [0.14 M], –40 °C, 24 h. (b) Partial 1H NMR spectra (600 MHz, CDCl3, 295 K) of macrocycle 10 (top), rotaxane 11 (middle) and the corresponding unthreaded axle (bottom).

The opposite enantiomer of the rotaxane, (−)-11, could be selectively accessed using electrophile 12, derived from (+)-cinchonine, a pseudoenantiomer of cinchonidine (see Supporting Information for synthesis). (18) Combining 2, 10, and 12 at −40 °C gave rotaxane (−)-11 in 50% e.e. and 51% yield (Figure 2a). The difference in enantioenrichment is a consequence of electrophiles 9 and 12 being diastereomers rather than true enantiomers.

Characterization of Rotaxanes

Comparison of the 1H NMR spectra of macrocycle 10, rotaxane 11, and the unthreaded axle (see Supporting Information for synthesis) in CDCl3 at 298 K (Figure 2b) confirmed the interlocked structure of 11. The geminal protons of the crown ether display twice the number of environments in rotaxane 11 as in unthreaded 10 due to desymmetrization of the two macrocycle faces upon rotaxane formation, while H3 and H5 of the axle (hydrogen labeling shown in Figure 2a), which are situated either side of the amide group, display significant diastereotopic splitting (Δδ = 0.39 and 0.22 ppm respectively) within the chiral environment of rotaxane 11 which, as would be expected, is absent for the corresponding achiral noninterlocked axle. Upfield shifts of H6 and H7 (Δδ = −0.32 and −0.34 ppm) in the threaded axle and HA, HB, and HC (Δδ = −0.49, −0.21, and −0.37 ppm) of the nitrocatechol unit of the threaded macrocycle result from π–π interactions involving these moieties. These intercomponent interactions may play a role in rigidifying the transition state of the collapsing tetrahedral intermediate. The large downfield shift of the amide N–H proton H4 (Δδ = +1.74 ppm) in 11 is indicative of intercomponent hydrogen bonding between the amide and the glycol chain of the macrocycle. An upshield shift of HD (Δδ = −1.29 ppm) results from hydrogen bonding with the amide oxygen atom. (14b)
Enantioenriched samples of rotaxane 11 (40% e.e. for the (+) enantiomer and 50% e.e. for the (−) enantiomer) were compared by circular dichroism (Figure 3a). The CD spectra of the mechanically planar chiral rotaxane enantiomers are symmetrical in terms of curve shape and have exciton couplings of opposite sign with maxima at 243 nm. The difference in intensity (normalized for absorption) of the spectra in Figure 3a corresponds to the difference in enantioenrichment of the samples.

Figure 3

Figure 3. (a) Circular dichroism spectra (1.0 × 10–4 M, CH2Cl2, 298 K) of (+)-11 (red) and (−)-11 (blue), baseline corrected. (b) Chemical structure of racemic rotaxane 13. (c) X-ray crystal structure of racemic rotaxane 13, side-on view showing intercomponent hydrogen bonds (in green). Hydrogen bond lengths: N47H—O16, 2.20 Å; O49—HC6, 2.63 Å. Hydrogen bond angles: N47—H–O16, 158.4°; O49—H–C6, 161.8°. (d) X-ray crystal structure of 13 viewed along the axle showing π-stacking between the macrocycle 1,2-dihydroxynaphthalene and axle bis(trifluoromethyl)phenyl rings. Centroid–centroid distance, 3.67 Å. Angle described by C40 and centroids, 97.6°. Solvate molecules and other hydrogen atoms omitted for clarity.

Although we were unable to obtain high quality single crystals of 11, single crystals of a racemic sample of 13 suitable for analysis by X-ray diffraction were grown by slow evaporation of an isopropanol/hexane solution of 13 (Figure 3b). Rotaxane 13 contains the same macrocycle as 11 and an axle derived from amine 2 and a different acyl stopper. The X-ray crystal structure of 13 (Figure 3c), containing both rotaxane enantiomers in the unit cell, shows similar intercomponent interactions to those observed by 1H NMR for 11 in solution (Figure 2b). Hydrogen bonds are present between an oxygen of the macrocycle glycol chain and the amide hydrogen atom of the axle and between the amide oxygen and a macrocycle C–H hydrogen atom (analogous to HD in 11). (14) The di(alkoxyl)naphthalene ring of the macrocycle and bis(trifluoromethyl)benzene unit of the axle π-stack (Figure 3d, closest centroid-centroid distance = 3.67 Å), (19) with the nitro-catechol moiety positioned so as to cover one face of the amide group. A similar arrangement in the transition state of the active template reaction would orient the macrocycle with respect to the axle building blocks such that one mechanically planar chiral enantiomer would be favored over the other.

Origin of Enantioselectivity

A preliminary indication of the origin of chiral transduction in these systems comes from the relative energies of the tetrahedral intermediates preceding (+)- and (−)-11, calculated at the PM6 level (20) using the Gaussian 09 software package (21) (Supporting Information and Figure 4). The collapse of similar tetrahedral intermediates has previously been shown (15a) to be the rate-determining step for the glyme catalysis of ester aminolysis. Following the Hammond postulate, the differences between the diastereomeric tetrahedral intermediates to (+)- and (−)-11 from 9 and 12 may resemble those between the transition states. The lowest energy intermediate calculated for both pseudoenantiomeric leaving groups featured an (S) stereocenter adjacent to the ammonium unit, but with the macrocycle orientation inverted for the two pseudoenantiomers (Figure 4), meaning changing between the leaving groups of 9 and 12 favors the formation of a different enantiomer of 11, as observed experimentally. The somewhat surprising indication that the two chiral leaving groups both favor an (S)-tetrahedral intermediate may reflect why the pseudoenantiomers do not generate equal and opposite e.e.’s in the active template reaction. The noncovalent interactions in the intermediate (e.g., the stacking of the electron-rich naphthalene unit with the electron-poor aryl group of the nucleophile, and the hydrogen bonding of the glycol oxygens to the H–N atoms) are reminiscent of those present in the X-ray crystal structure of rotaxane 13.

Figure 4

Figure 4. Tentative rationale for the transfer of chirality from Euclidean point-chirality (of the leaving group) to mechanical planar chirality (of the rotaxane). The lowest energy tetrahedral intermediates were modeled (see Supporting Information) using (a) electrophile 9 or (b) electrophile 12. The di(alkoxyl)naphthalene ring of the macrocycle and bis(trifluoromethyl)benzene unit originating from the nucleophile π-stack, causing the nitro-catechol ring to be positioned so as to cover one face of the tetrahedral center of the intermediate. This thermodynamically favored arrangement of components ensures the different handedness of the pseudoenantiomeric leaving groups is well-expressed in the diastereomeric transition states, resulting in enantioselectivity in the mechanically planar chiral rotaxane product. Hydrogen bonds are indicated by black dotted lines.

Also consistent with the stacking of the electron-rich naphthalene unit with the electron-poor aryl group of the nucleophile providing the driving force for organization of the transition state is the experimental evidence that decreasing the electron density of the other aromatic ring of the macrocycle increases the enantioselectivity of rotaxane formation (i.e., 12% e.e. for (+)-8; 40% e.e. for (+)-11). The less electron-rich the catechol ring is, the less it competes with the naphthalene group for π-stacking with the bis(trifluoromethyl)benzylamine and so the greater the enantiodiscrimination in the transition state.

Conclusions

Click to copy section linkSection link copied!

The examples presented demonstrate that mechanically planar chiral rotaxanes can be directly accessed in up to 50% e.e. in a single synthetic step. The chirality of the point-chiral leaving group is transferred into mechanically planar chirality in the rotaxane through metal-free active template N-acylation. Pseudoenantiomeric cinchona alkaloids allow either rotaxane enantiomer to be accessed. X-ray crystallography and molecular modeling suggest that the origin of the enantioselectivity lies in π-stacking of an electron-rich aromatic ring on the macrocycle with an electron-poor aryl group originating from the nucleophilic axle building block. This positions the second aromatic ring of the macrocycle in an orientation that blocks one face of the electrophile. Simple methods for accessing enantioenriched mechanically planar chiral rotaxanes should improve their availability for investigation in applications such as asymmetric catalysis, (7) chiral (bio)molecule sensing, (1,6,22) and novel designs (23) of molecular machinery.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.0c03447.

  • Experimental procedures, synthesis and characterization data, including circular dichroism, chiral HPLC, NMR, MS, and X-ray crystallography data (PDF)

  • Crystallographic data for 13 (CIF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
    • David A. Leigh - Department of Chemistry, University of Manchester, Oxford Road, Manchester M13 9PL, United KingdomSchool of Chemistry and Molecular Engineering, East China Normal University, 200062 Shanghai, ChinaOrcidhttp://orcid.org/0000-0002-1202-4507 Email: [email protected]
  • Authors
    • Chong Tian - Department of Chemistry, University of Manchester, Oxford Road, Manchester M13 9PL, United KingdomOrcidhttp://orcid.org/0000-0001-7264-9042
    • Stephen D. P. Fielden - Department of Chemistry, University of Manchester, Oxford Road, Manchester M13 9PL, United Kingdom
    • Borja Pérez-Saavedra - Department of Chemistry, University of Manchester, Oxford Road, Manchester M13 9PL, United Kingdom
    • Iñigo J. Vitorica-Yrezabal - Department of Chemistry, University of Manchester, Oxford Road, Manchester M13 9PL, United Kingdom
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

We thank the Engineering and Physical Sciences Research Council (EPSRC) (EP/P027067/1), the European Research Council (ERC) (Advanced Grant No. 786630), the China 1000 Talents Plan and East China Normal University for funding, the University of Manchester for a studentship (to S.D.P.F.), the Diamond Light Source (U.K.) for synchrotron beamtime on I19, the University of Manchester mass spectrometry service for high-resolution mass spectrometry, the Computational Shared Facility 3 (CSF3) at the University of Manchester for computational resources, and Jing Liu for preliminary studies. D.A.L. is a Royal Society Research Professor.

References

Click to copy section linkSection link copied!

This article references 23 other publications.

  1. 1
    (a) Bruns, C. J.; Stoddart, J. F. The Nature of the Mechanical Bond: From Molecules to Machines; John Wiley & Sons: Hoboken, NJ, 2017.
    (b) Jamieson, E. M. G.; Modicom, F.; Goldup, S. M. Chirality in rotaxanes and catenanes. Chem. Soc. Rev. 2018, 47, 52665311,  DOI: 10.1039/C8CS00097B
    (c) Evans, N. H. Chiral catenanes and rotaxanes: Fundamentals and emerging applications. Chem. - Eur. J. 2018, 24, 31013112,  DOI: 10.1002/chem.201704149
    (d) Nakazono, K.; Takata, T. Mechanical chirality of rotaxanes: Synthesis and function. Symmetry 2020, 12, 144,  DOI: 10.3390/sym12010144
  2. 2

    If a prochiral or meso thread is incorporated into a rotaxane, then chirality can arise from the position of the ring on the axle. See:

    (a) Alvarez-Pérez, M.; Goldup, S. M.; Leigh, D. A.; Slawin, A. M. Z. A chemically-driven molecular information ratchet. J. Am. Chem. Soc. 2008, 130, 18361838,  DOI: 10.1021/ja7102394
    (b) Cakmak, Y.; Erbas-Cakmak, S.; Leigh, D. A. Asymmetric catalysis with a mechanically point-chiral rotaxane. J. Am. Chem. Soc. 2016, 138, 17491751,  DOI: 10.1021/jacs.6b00303
    (c) Dommaschk, M.; Echavarren, J.; Leigh, D. A.; Marcos, V.; Singleton, T. A. Dynamic control of chiral space through local symmetry breaking in a rotaxane organocatalyst. Angew. Chem., Int. Ed. 2019, 58, 1495514958,  DOI: 10.1002/anie.201908330
  3. 3

    Mechanical planar chirality can also result from confinement of an unsymmetrical macrocycle to one side of a nonprochiral axle possessing Dnh symmetry. See:

    (a) Mochizuki, Y.; Ikeyatsu, K.; Mutoh, Y.; Hosoya, S.; Saito, S. Synthesis of mechanically planar chiral rac-[2]rotaxanes by partitioning of an achiral [2]rotaxane: Stereoinversion induced by shuttling. Org. Lett. 2017, 19, 43474350,  DOI: 10.1021/acs.orglett.7b02043
    (b) Corra, S.; de Vet, C.; Groppi, J.; La Rosa, M.; Silvi, S.; Baroncini, M.; Credi, A. Chemical on/off switching of mechanically planar chirality and chiral anion recognition in a [2]rotaxane molecular shuttle. J. Am. Chem. Soc. 2019, 141, 91299133,  DOI: 10.1021/jacs.9b00941
  4. 4

    For examples of other stereochemical consequences of threading, see:

    (a) Fuller, A.-M. L.; Leigh, D. A.; Lusby, P. J. Sequence isomerism in [3]rotaxanes. J. Am. Chem. Soc. 2010, 132, 49544959,  DOI: 10.1021/ja1006838
    (b) Talotta, C.; Gaeta, C.; Qi, Z.; Schalley, C. A.; Neri, P. Pseudorotaxanes with self-sorted sequence and stereochemical orientation. Angew. Chem., Int. Ed. 2013, 52, 74377441,  DOI: 10.1002/anie.201301570
    (c) La Manna, P.; Talotta, C.; Gaeta, C.; Soriente, A.; De Rosa, M.; Neri, P. Threading of an inherently directional calixarene wheel with oriented ammonium axles. J. Org. Chem. 2017, 82, 89738983,  DOI: 10.1021/acs.joc.7b01388
    (d) Cui, J.-S.; Ba, Q.-K.; Ke, H.; Valkonen, A.; Rissanen, K.; Jiang, W. Directional shuttling of a stimuli-responsive cone-like macrocycle on a single-state symmetric dumbbell axle. Angew. Chem., Int. Ed. 2018, 57, 78097814,  DOI: 10.1002/anie.201803349
    (e) Zheng, L.-S.; Cui, J.-S.; Jiang, W. Biomimetic synchronized motion of two interacting macrocycles in [3]rotaxane-based molecular shuttles. Angew. Chem., Int. Ed. 2019, 58, 1513615141,  DOI: 10.1002/anie.201910318
    (f) Ng, A. W. H.; Yee, C.-C.; Au-Yeung, H. Y. Radial hetero[5]catenanes: peripheral isomer sequences of the interlocked macrocycles. Angew. Chem., Int. Ed. 2019, 58, 1737517382,  DOI: 10.1002/anie.201908576
  5. 5
    Ishiwari, F.; Nakazono, K.; Koyama, Y.; Takata, T. Induction of single handed helicity of polyacetylenes using mechanically chiral rotaxanes as chiral sources. Angew. Chem., Int. Ed. 2017, 56, 1485814862,  DOI: 10.1002/anie.201707926
  6. 6
    Kameta, N.; Nagawa, Y.; Karikomi, M.; Hiratani, K. Chiral sensing for amino acid derivative based on a [2]rotaxane composed of an asymmetric rotor and an asymmetric axle. Chem. Commun. 2006, 37143716,  DOI: 10.1039/b607251h
  7. 7
    Heard, A.; Goldup, S. M. Synthesis of a mechanically planar chiral rotaxane ligand for enantioselective catalysis. Chem. 2020, 6, 9941006,  DOI: 10.1016/j.chempr.2020.02.006
  8. 8
    Glen, P. E.; O’Neill, J. A. T.; Lee, A.-L. Synthesis of a C1-symmetric Box macrocycle and studies towards active-template synthesis of mechanically planar chiral rotaxanes. Tetrahedron 2013, 69, 5768,  DOI: 10.1016/j.tet.2012.10.069
  9. 9
    (a) Yamamoto, C.; Okamoto, Y.; Schmidt, T.; Jäger, R.; Vögtle, F. Enantiomeric resolution of cycloenantiomeric rotaxane, topologically chiral catenane, and pretzel-shaped molecules: Observation of pronounced circular dichroism. J. Am. Chem. Soc. 1997, 119, 1054710548,  DOI: 10.1021/ja971764q
    (b) Schalley, C. A.; Beizai, K.; Vögtle, F. On the way to rotaxane-based molecular motors: Studies in molecular mobility and topological chirality. Acc. Chem. Res. 2001, 34, 465476,  DOI: 10.1021/ar000179i
    (c) Kameta, N.; Hiratani, K.; Nagawa, Y. A novel synthesis of chiral rotaxanes via covalent bond formation. Chem. Commun. 2004, 466467,  DOI: 10.1039/b314744d
    (d) Hirose, K.; Ukimi, M.; Ueda, S.; Onoda, C.; Kano, R.; Tsuda, K.; Hinohara, Y.; Tobe, Y. The asymmetry is derived from mechanical interlocking of achiral axle and achiral ring components – Syntheses and properties of optically pure [2]rotaxanes. Symmetry 2018, 10, 20,  DOI: 10.3390/sym10010020
    (e) Gell, C. E.; McArdle-Ismaguilov, T. A.; Evans, N. H. Modulating the expression of chirality in a mechanically chiral rotaxane. Chem. Commun. 2019, 55, 15761579,  DOI: 10.1039/C8CC10044F
    (f) Gaedke, M.; Witte, F.; Anhäuser, J.; Hupatz, H.; Schröder, H. V.; Valkonen, A.; Rissanen, K.; Lützen, A.; Paulus, B.; Schalley, C. A. Chiroptical inversion of a planar chiral redox-switchable rotaxane. Chem. Sci. 2019, 10, 1000310009,  DOI: 10.1039/C9SC03694F
  10. 10
    Bordoli, R.; Goldup, S. M. An efficient approach to mechanically planar chiral rotaxanes. J. Am. Chem. Soc. 2014, 136, 48174820,  DOI: 10.1021/ja412715m
  11. 11
    Jinks, M. A.; de Juan, A.; Denis, M.; Fletcher, C. J.; Galli, M.; Jamieson, E. M. G.; Modicom, F.; Zhang, Z.; Goldup, S. M. Stereoselective synthesis of mechanically planar chiral rotaxanes. Angew. Chem., Int. Ed. 2018, 57, 1480614810,  DOI: 10.1002/anie.201808990
  12. 12
    Makita, Y.; Kihara, N.; Nakakoji, N.; Takata, T.; Inagaki, S.; Yamamoto, C.; Okamoto, Y. Catalytic asymmetric synthesis and optical resolution of planar chiral rotaxane. Chem. Lett. 2007, 36, 162163,  DOI: 10.1246/cl.2007.162
  13. 13
    De Bo, G.; Dolphijn, G.; McTernan, C. T.; Leigh, D. A. [2]Rotaxane formation by transition state stabilization. J. Am. Chem. Soc. 2017, 139, 84558457,  DOI: 10.1021/jacs.7b05640
  14. 14
    (a) Fielden, S. D. P.; Leigh, D. A.; McTernan, C. T.; Pérez-Saavedra, B.; Vitorica-Yrezabal, I. J. Spontaneous assembly of rotaxanes from a primary amine, crown ether and electrophile. J. Am. Chem. Soc. 2018, 140, 60496052,  DOI: 10.1021/jacs.8b03394
    (b) Tian, C.; Fielden, S. D. P.; Whitehead, G. F. S.; Vitorica-Yrezabal, I. J.; Leigh, D. A. Weak functional group interactions revealed through metal-free active template rotaxane synthesis. Nat. Commun. 2020, 11, 744,  DOI: 10.1038/s41467-020-14576-7
  15. 15
    (a) Hogan, J. C.; Gandour, R. D. Structural requirements for glyme catalysis in butylaminolysis of aryl acetates in chlorobenzene. Identification of -OCH2CH2OCH2CH2OCH2CH2O- as the optimal subunit for catalysis. J. Org. Chem. 1991, 56, 28212826,  DOI: 10.1021/jo00008a044
    (b) Basilio, N.; García-Río, L.; Mejuto, J. C.; Pérez-Lorenzo, M. A. New reaction pathway in the ester aminolysis catalyzed by glymes and crown ethers. J. Org. Chem. 2006, 71, 42804285,  DOI: 10.1021/jo060389u
  16. 16

    Chiral leaving groups have previously been employed in asymmetric substitution reactions. See:

    Lepore, S. D.; Mondal, D. Recent advances in heterolytic nucleofugal leaving groups. Tetrahedron 2007, 63, 51035122,  DOI: 10.1016/j.tet.2007.03.049
  17. 17
    Heller, D.; Buschmann, H.; Scharf, H.-D. Nonlinear temperature behavior of product ratios in selection processes. Angew. Chem., Int. Ed. Engl. 1996, 35, 18521854,  DOI: 10.1002/anie.199618521
  18. 18
    Cinchona alkaloids in synthesis and catalysis: Ligands, immobilization and organocatalysis; Eui Song, C., Ed.; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, 2009.
  19. 19
    (a) Hunter, C. A.; Sanders, J. K. M. The nature of π–π interactions. J. Am. Chem. Soc. 1990, 112, 55255534,  DOI: 10.1021/ja00170a016
    (b) Riwar, L.-J.; Trapp, N.; Kuhn, B.; Diederich, F. Substituent effects in parallel displaced π–π stacking interactions: Distance matters. Angew. Chem., Int. Ed. 2017, 56, 1125211257,  DOI: 10.1002/anie.201703744
  20. 20
    Stewart, J. J. P. Optimization of parameters for semiempirical methods V. Modification of NDDO approximations and application to 70 elements. J. Mol. Model. 2007, 13, 11731213,  DOI: 10.1007/s00894-007-0233-4
  21. 21
    Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, revision D.01; Gaussian, Inc.; Wallingford, CT, 2013.
  22. 22
    (a) Lim, J. Y. C.; Marques, I.; Félix, V.; Beer, P. D. A chiral halogen-bonding [3]rotaxane for the recognition and sensing of biologically relevant dicarboxylate anions. Angew. Chem., Int. Ed. 2018, 57, 584588,  DOI: 10.1002/anie.201711176
    (b) Ba̧k, K. M.; Porfyrakis, K.; Davis, J. J.; Beer, P. D. Exploiting the mechanical bond for molecular recognition and sensing of charged species. Mater. Chem. Front. 2020, 4, 10521073,  DOI: 10.1039/C9QM00698B
  23. 23
    (a) Fu, X.; Zhang, Q.; Rao, S.-J.; Qu, D.-H.; Tian, H. One-pot synthesis of a [c2]daisy-chain-containing hetero[4]rotaxane via a self-sorting strategy. Chem. Sci. 2016, 7, 16961701,  DOI: 10.1039/C5SC04844C
    (b) Rao, S.-J.; Zhang, Q.; Mei, J.; Ye, X.-H.; Gao, C.; Wang, Q.-C.; Qu, D.-H.; Tian, H. One-pot synthesis of hetero[6]rotaxane bearing three different kinds of macrocycle through a self-sorting process. Chem. Sci. 2017, 8, 67776783,  DOI: 10.1039/C7SC03232C
    (c) Chen, S.; Wang, Y.; Nie, T.; Bao, C.; Wang, C.; Xu, T.; Lin, Q.; Qu, D.-H.; Gong, X.; Yang, Y.; Zhu, L.; Tian, H. An artificial molecular shuttle operates in lipid bilayers for ion transport. J. Am. Chem. Soc. 2018, 140, 1799217998,  DOI: 10.1021/jacs.8b09580
    (d) Zheng, X.; Zhang, Y.; Cao, N.; Li, X.; Zhang, S.; Du, R.; Wang, H.; Ye, Z.; Wang, Y.; Cao, F.; Li, H.; Hong, X.; Sue, A. C.-H.; Yang, C.; Liu, W.-G.; Li, H. Coulombic-enhanced hetero radical pairing interactions. Nat. Commun. 2018, 9, 1961,  DOI: 10.1038/s41467-018-04335-0
    (e) Li, W.-J.; Wang, W.; Wang, X.-Q.; Li, M.; Ke, Y.; Yao, R.; Wen, J.; Yin, G.-Q.; Jiang, B.; Li, X.; Yin, P.; Yang, H. B. Daisy chain dendrimers: Integrated mechanically interlocked molecules with stimuli-induced dimension modulation feature. J. Am. Chem. Soc. 2020, 142, 84738482,  DOI: 10.1021/jacs.0c02475

Cited By

Click to copy section linkSection link copied!

This article is cited by 64 publications.

  1. Stephen M. Goldup. The End of the Beginning of Mechanical Stereochemistry. Accounts of Chemical Research 2024, 57 (12) , 1696-1708. https://doi.org/10.1021/acs.accounts.4c00195
  2. Guiying Long, Yanping Deng, Wei Zhao, Guofu Zhou, Dirk J. Broer, Ben L. Feringa, Jiawen Chen. Photoresponsive Biomimetic Functions by Light-Driven Molecular Motors in Three Dimensionally Printed Liquid Crystal Elastomers. Journal of the American Chemical Society 2024, 146 (20) , 13894-13902. https://doi.org/10.1021/jacs.4c01642
  3. Peter R. Gallagher, Andrea Savoini, Abed Saady, John R. J. Maynard, Patrick W. V. Butler, Graham J. Tizzard, Stephen M. Goldup. Facial Selectivity in Mechanical Bond Formation: Axially Chiral Enantiomers and Geometric Isomers from a Simple Prochiral Macrocycle. Journal of the American Chemical Society 2024, 146 (13) , 9134-9141. https://doi.org/10.1021/jacs.3c14329
  4. Andrea Savoini, Peter R. Gallagher, Abed Saady, Stephen M. Goldup. The Final Stereogenic Unit of [2]Rotaxanes: Type 2 Geometric Isomers. Journal of the American Chemical Society 2024, 146 (12) , 8472-8479. https://doi.org/10.1021/jacs.3c14594
  5. Julio Puigcerver, Marta Marin-Luna, Javier Iglesias-Sigüenza, Mateo Alajarin, Alberto Martinez-Cuezva, Jose Berna. Mechanically Planar-to-Point Chirality Transmission in [2]Rotaxanes. Journal of the American Chemical Society 2024, 146 (5) , 2882-2887. https://doi.org/10.1021/jacs.3c11611
  6. Keisuke Wada, Kiichi Yasuzawa, Shixin Fa, Yuuya Nagata, Kenichi Kato, Shunsuke Ohtani, Tomoki Ogoshi. Diastereoselective Rotaxane Synthesis with Pillar[5]arenes via Co-crystallization and Solid-State Mechanochemical Processes. Journal of the American Chemical Society 2023, 145 (28) , 15324-15330. https://doi.org/10.1021/jacs.3c02919
  7. Martin J. Power, David T. J. Morris, Iñigo J. Vitorica-Yrezabal, David A. Leigh. Compact Rotaxane Superbases. Journal of the American Chemical Society 2023, 145 (15) , 8593-8599. https://doi.org/10.1021/jacs.3c01202
  8. Qi-Yun Feng, Yu-Qi Mao, Mei-Xiang Wang, Shuo Tong. Chiral Crown Ethers Accessed from Catalytic Enantioselective Desymmetrization Reactions. Organic Letters 2022, 24 (39) , 7107-7112. https://doi.org/10.1021/acs.orglett.2c02688
  9. Lorna Binks, Chong Tian, Stephen D. P. Fielden, Iñigo J. Vitorica-Yrezabal, David A. Leigh. Transamidation-Driven Molecular Pumps. Journal of the American Chemical Society 2022, 144 (34) , 15838-15844. https://doi.org/10.1021/jacs.2c06807
  10. Arnau Rodríguez-Rubio, Andrea Savoini, Florian Modicom, Patrick Butler, Stephen M. Goldup. A Co-conformationally “Topologically” Chiral Catenane. Journal of the American Chemical Society 2022, 144 (27) , 11927-11932. https://doi.org/10.1021/jacs.2c02029
  11. Jiaxin Hou, Guiying Long, Wei Zhao, Guofu Zhou, Danqing Liu, Dirk J. Broer, Ben L. Feringa, Jiawen Chen. Phototriggered Complex Motion by Programmable Construction of Light-Driven Molecular Motors in Liquid Crystal Networks. Journal of the American Chemical Society 2022, 144 (15) , 6851-6860. https://doi.org/10.1021/jacs.2c01060
  12. Hsien-Liang Cho, Chien-Chen Lai, Yi-Hung Liu, Hsiu-Fu Hsu, Shie-Ming Peng, Sheng-Hsien Chiu. Metal-Ion-Induced Mechanical Chirality: Achiral Rotaxane as the Only Ligand in Chiral Palladium(II)–N-Heterocyclic Carbene Complexes. Organic Letters 2022, 24 (10) , 1996-2001. https://doi.org/10.1021/acs.orglett.2c00426
  13. Shinya Tajima, Atsuya Muranaka, Masaya Naito, Noriho Taniguch, Mei Harada, Shinobu Miyagawa, Masahiro Ueda, Hikaru Takaya, Nagao Kobayashi, Masanobu Uchiyama, Yuji Tokunaga. Synthesis of a Mechanically Planar Chiral and Axially Chiral [2]Rotaxane. Organic Letters 2021, 23 (22) , 8678-8682. https://doi.org/10.1021/acs.orglett.1c02983
  14. Xiang Gao, Zheng Cui, Yue-Rong Shen, Dong Liu, Yue-Jian Lin, Guo-Xin Jin. Synthesis and Near-Infrared Photothermal Conversion of Discrete Supramolecular Topologies Featuring Half-Sandwich [Cp*Rh] Units. Journal of the American Chemical Society 2021, 143 (42) , 17833-17842. https://doi.org/10.1021/jacs.1c09333
  15. Kenji Caprice, Dávid Pál, Céline Besnard, Bartomeu Galmés, Antonio Frontera, Fabien B. L. Cougnon. Diastereoselective Amplification of a Mechanically Chiral [2]Catenane. Journal of the American Chemical Society 2021, 143 (31) , 11957-11962. https://doi.org/10.1021/jacs.1c06557
  16. Takaaki Morise, Atsuya Muranaka, Hayato Ban, Mei Harada, Masaya Naito, Kazuyuki Yoshida, Nagao Kobayashi, Masanobu Uchiyama, Yuji Tokunaga. A Chiral [3]Rotaxane Comprising Achiral Bis-macrocyclic and Dumbbell-Shaped Components. Organic Letters 2021, 23 (6) , 2120-2124. https://doi.org/10.1021/acs.orglett.1c00271
  17. Alberto Martinez-Cuezva, Adrian Saura-Sanmartin, Mateo Alajarin, Jose Berna. Mechanically Interlocked Catalysts for Asymmetric Synthesis. ACS Catalysis 2020, 10 (14) , 7719-7733. https://doi.org/10.1021/acscatal.0c02032
  18. Sean R. Barlow, Nicholas H. Evans. Synthesis of a [2]catenane by ring closing metathesis of a [2]rotaxane prepared by crown ether active templation. Organic & Biomolecular Chemistry 2024, 22 (37) , 7632-7636. https://doi.org/10.1039/D4OB01028K
  19. Alexander Ryabchun, Romain Jamagne, Javier Echavarren, Manee Patanapongpibul, Liang Zhang, Nathalie Katsonis, David A. Leigh. Macroscopic spiral rotation of microscopic objects induced by nanoscale rotaxane dynamics. Chem 2024, 10 (7) , 2196-2206. https://doi.org/10.1016/j.chempr.2024.03.013
  20. Stefan Borsley, David A. Leigh, Benjamin M. W. Roberts. Molecular Ratchets and Kinetic Asymmetry: Giving Chemistry Direction. Angewandte Chemie 2024, 136 (23) https://doi.org/10.1002/ange.202400495
  21. Stefan Borsley, David A. Leigh, Benjamin M. W. Roberts. Molecular Ratchets and Kinetic Asymmetry: Giving Chemistry Direction. Angewandte Chemie International Edition 2024, 63 (23) https://doi.org/10.1002/anie.202400495
  22. Abed Saady, Georgia K. Malcolm, Matthew P. Fitzpatrick, Noel Pairault, Graham J. Tizzard, Soran Mohammed, Ali Tavassoli, Stephen M. Goldup. A Platform Approach to Cleavable Macrocycles for the Controlled Disassembly of Mechanically Caged Molecules. Angewandte Chemie 2024, 136 (16) https://doi.org/10.1002/ange.202400344
  23. Abed Saady, Georgia K. Malcolm, Matthew P. Fitzpatrick, Noel Pairault, Graham J. Tizzard, Soran Mohammed, Ali Tavassoli, Stephen M. Goldup. A Platform Approach to Cleavable Macrocycles for the Controlled Disassembly of Mechanically Caged Molecules. Angewandte Chemie International Edition 2024, 63 (16) https://doi.org/10.1002/anie.202400344
  24. Manisha Prakashni, Suvankar Dasgupta. BP23C7: high-yield synthesis and application in constructing [3]rotaxanes and responsive pseudo[2]rotaxanes. Organic & Biomolecular Chemistry 2024, 22 (9) , 1871-1884. https://doi.org/10.1039/D3OB02094K
  25. Adrian Saura‐Sanmartin. Synthesis of ‘Impossible’ Rotaxanes. Chemistry – A European Journal 2024, https://doi.org/10.1002/chem.202304025
  26. Ryo Katoono, Kota Asai. Chiral [1]rotaxanes with a two-fold bridge based on phenylacetylene macrocycles (PAMs): two different chiroptical properties created by 7PAM. Chemistry Letters 2024, 53 (1) https://doi.org/10.1093/chemle/upad006
  27. Arnau Rodríguez-Rubio, Peter Gallagher, Stephen M. Goldup. Chiral Auxiliaries for the Synthesis of Mechanically Chiral Interlocked Molecules. 2024https://doi.org/10.1016/B978-0-32-390644-9.00139-6
  28. Romain Jamagne, Martin J. Power, Zhi-Hui Zhang, Germán Zango, Benjamin Gibber, David A. Leigh. Active template synthesis. Chemical Society Reviews 2024, 84 https://doi.org/10.1039/D4CS00430B
  29. Stephen D. P. Fielden. Crown Ether Active Template Synthesis of Rotaxanes**. ChemSystemsChem 2023, 145 https://doi.org/10.1002/syst.202300048
  30. Yan Zou, Hai‐Ning Zhang, Qiu‐Shui Mu, Li‐Long Dang, Guo‐Xin Jin. Selective Construction of Borromean Rings and Tweezer‐Like Molecular Assembly Featuring Cp*Rh/Ir Clips for Near‐Infrared Photothermal Conversion †. Chinese Journal of Chemistry 2023, 41 (23) , 3229-3237. https://doi.org/10.1002/cjoc.202300359
  31. Xiaohua Zhou, Wei Wang. Research progress in mechanically planar chiral rotaxanes. SCIENTIA SINICA Chimica 2023, 53 (12) , 2460-2475. https://doi.org/10.1360/SSC-2023-0166
  32. Mengyao Tang, Xiaoyu Yang. Catalytic Enantioselective Synthesis of Inherently Chiral Molecules: Recent Advances. European Journal of Organic Chemistry 2023, 26 (42) https://doi.org/10.1002/ejoc.202300738
  33. Qinke Cui, Xinxing Zhang. Polyrotaxane‐Based Functional Materials Enabled by Molecular Mobility and Conformational Transition †. Chinese Journal of Chemistry 2023, 41 (20) , 2715-2729. https://doi.org/10.1002/cjoc.202300239
  34. Xin Song, Xuefeng Zhu, Shengfu Wu, Wenzhuo Chen, Wei Tian, Minghua Liu. Chiroptical switching in the azobenzene‐based self‐locked [1]rotaxane by solvent and photoirradiation. Chirality 2023, 35 (10) , 692-699. https://doi.org/10.1002/chir.23567
  35. Abed Saady, Stephen M. Goldup. Triazole formation and the click concept in the synthesis of interlocked molecules. Chem 2023, 9 (8) , 2110-2127. https://doi.org/10.1016/j.chempr.2023.07.001
  36. Gen Li, Tanya K. Ronson, Roy Lavendomme, Zehuan Huang, Carles Fuertes-Espinosa, Dawei Zhang, Jonathan R. Nitschke. Enantiopure FeII4L4 cages bind steroids stereoselectively. Chem 2023, 9 (6) , 1549-1561. https://doi.org/10.1016/j.chempr.2023.03.011
  37. Shu Zhang, Arnau Rodríguez-Rubio, Abed Saady, Graham J. Tizzard, Stephen M. Goldup. A chiral macrocycle for the stereoselective synthesis of mechanically planar chiral rotaxanes and catenanes. Chem 2023, 9 (5) , 1195-1207. https://doi.org/10.1016/j.chempr.2023.01.009
  38. Rebecca L. Spicer, Curtis C. Shearman, Nicholas H. Evans. Anion Modulated Expression of Chirality in Hydrogen Bond Templated Mechanically Chiral [2]Rotaxanes. Chemistry – A European Journal 2023, 29 (21) https://doi.org/10.1002/chem.202203502
  39. Shafieq Ahmad Wagay, Lubna Khan, Rashid Ali. Recent Advancements in Ion‐Pair Receptors. Chemistry – An Asian Journal 2023, 18 (2) https://doi.org/10.1002/asia.202201080
  40. Adrian Saura‐Sanmartin. Post‐Mechanical Bond Formation Desymmetrization Approach to Obtain Mechanically Planar Chiral Rotaxanes. Asian Journal of Organic Chemistry 2023, 12 (1) https://doi.org/10.1002/ajoc.202200601
  41. Mingfeng Li, Xiu Li Chia, Chong Tian, Ye Zhu. Mechanically planar chiral rotaxanes through catalytic desymmetrization. Chem 2022, 8 (10) , 2843-2855. https://doi.org/10.1016/j.chempr.2022.08.009
  42. Carmen Lopez‐Leonardo, Adrian Saura‐Sanmartin, Marta Marin‐Luna, Mateo Alajarin, Alberto Martinez‐Cuezva, Jose Berna. Ring‐to‐Thread Chirality Transfer in [2]Rotaxanes for the Synthesis of Enantioenriched Lactams. Angewandte Chemie 2022, 134 (39) https://doi.org/10.1002/ange.202209904
  43. Carmen Lopez‐Leonardo, Adrian Saura‐Sanmartin, Marta Marin‐Luna, Mateo Alajarin, Alberto Martinez‐Cuezva, Jose Berna. Ring‐to‐Thread Chirality Transfer in [2]Rotaxanes for the Synthesis of Enantioenriched Lactams. Angewandte Chemie International Edition 2022, 61 (39) https://doi.org/10.1002/anie.202209904
  44. John R. J. Maynard, Peter Gallagher, David Lozano, Patrick Butler, Stephen M. Goldup. Mechanically axially chiral catenanes and noncanonical mechanically axially chiral rotaxanes. Nature Chemistry 2022, 14 (9) , 1038-1044. https://doi.org/10.1038/s41557-022-00973-6
  45. Jiayan Li, Donglan Huang, Qiuhui Fang, Changgui Zhao. Progress on synthesis of optically pure mechanical planar chiral rotaxanes and planar chiral macrocycle molecules. Chinese Science Bulletin 2022, 67 (20) , 2383-2392. https://doi.org/10.1360/TB-2021-1324
  46. Rosa López, Claudio Palomo. Planar Chirality: A Mine for Catalysis and Structure Discovery. Angewandte Chemie 2022, 134 (13) https://doi.org/10.1002/ange.202113504
  47. Rosa López, Claudio Palomo. Planar Chirality: A Mine for Catalysis and Structure Discovery. Angewandte Chemie International Edition 2022, 61 (13) https://doi.org/10.1002/anie.202113504
  48. Alberto de Juan, David Lozano, Andrew W. Heard, Michael A. Jinks, Jorge Meijide Suarez, Graham J. Tizzard, Stephen M. Goldup. A chiral interlocking auxiliary strategy for the synthesis of mechanically planar chiral rotaxanes. Nature Chemistry 2022, 14 (2) , 179-187. https://doi.org/10.1038/s41557-021-00825-9
  49. Ayumi Imayoshi. Discrimination of Mobile Supramolecular Chirality: Kinetic Resolution of Mechanically Planar Chiral Rotaxanes by Organocatalysis. 2022, 25-76. https://doi.org/10.1007/978-981-16-7431-0_3
  50. Yue‐Rong Shen, Xiang Gao, Zheng Cui, Guo‐Xin Jin. Rational Design and Synthesis of Interlocked [2]Catenanes Featuring Half‐Sandwich Cp*Rh/Ir Units and Pyrene‐Based Ligands †. Chinese Journal of Chemistry 2021, 39 (12) , 3303-3308. https://doi.org/10.1002/cjoc.202100546
  51. Ayumi Imayoshi, Bhatraju Vasantha Lakshmi, Yoshihiro Ueda, Tomoyuki Yoshimura, Aki Matayoshi, Takumi Furuta, Takeo Kawabata. Enantioselective preparation of mechanically planar chiral rotaxanes by kinetic resolution strategy. Nature Communications 2021, 12 (1) https://doi.org/10.1038/s41467-020-20372-0
  52. Andrew R. Puente, Adrien Bessaguet, Noël Pairault, Grégory Pieters, Jeanne Crassous, Prasad L. Polavarapu, Isabelle Opalinski, Sébastien Papot. Absolute configuration of a [1]rotaxane determined from vibrational and electronic circular dichroism spectra. Chirality 2021, 33 (11) , 773-782. https://doi.org/10.1002/chir.23365
  53. Arthur H. G. David, J. Fraser Stoddart. Chiroptical Properties of Mechanically Interlocked Molecules. Israel Journal of Chemistry 2021, 61 (9-10) , 608-621. https://doi.org/10.1002/ijch.202100066
  54. Federico Nicoli, Massimo Baroncini, Serena Silvi, Jessica Groppi, Alberto Credi. Direct synthetic routes to functionalised crown ethers. Organic Chemistry Frontiers 2021, 8 (19) , 5531-5549. https://doi.org/10.1039/D1QO00699A
  55. Eoin P. McCarney, June I. Lovitt, Thorfinnur Gunnlaugsson. Mechanically Interlocked Chiral Self‐Templated [2]Catenanes from 2,6‐Bis(1,2,3‐triazol‐4‐yl)pyridine (btp) Ligands. Chemistry – A European Journal 2021, 27 (47) , 12052-12057. https://doi.org/10.1002/chem.202101773
  56. Stefano Corra, Christiaan de Vet, Massimo Baroncini, Alberto Credi, Serena Silvi. Stereodynamics of E/Z isomerization in rotaxanes through mechanical shuttling and covalent bond rotation. Chem 2021, 7 (8) , 2137-2150. https://doi.org/10.1016/j.chempr.2021.04.010
  57. Shuntaro Amano, Stephen D. P. Fielden, David A. Leigh. A catalysis-driven artificial molecular pump. Nature 2021, 594 (7864) , 529-534. https://doi.org/10.1038/s41586-021-03575-3
  58. Liang Shen, Ming Yi, Susilo Japip, Chao Han, Lian Tian, Cher Hon Lau, Yan Wang. Breaking through permeability–selectivity trade‐off of thin‐film composite membranes assisted with crown ethers. AIChE Journal 2021, 67 (6) https://doi.org/10.1002/aic.17173
  59. Wei‐Jian Li, Qingyi Gu, Xu‐Qing Wang, Dan‐Yang Zhang, Yu‐Te Wang, Xiao He, Wei Wang, Hai‐Bo Yang. AIE‐Active Chiral [3]Rotaxanes with Switchable Circularly Polarized Luminescence. Angewandte Chemie 2021, 133 (17) , 9593-9601. https://doi.org/10.1002/ange.202100934
  60. Wei‐Jian Li, Qingyi Gu, Xu‐Qing Wang, Dan‐Yang Zhang, Yu‐Te Wang, Xiao He, Wei Wang, Hai‐Bo Yang. AIE‐Active Chiral [3]Rotaxanes with Switchable Circularly Polarized Luminescence. Angewandte Chemie International Edition 2021, 60 (17) , 9507-9515. https://doi.org/10.1002/anie.202100934
  61. Noël Pairault, Adrien Bessaguet, Romain Barat, Lucas Frédéric, Grégory Pieters, Jeanne Crassous, Isabelle Opalinski, Sébastien Papot. Diastereoselective synthesis of [1]rotaxanes via an active metal template strategy. Chemical Science 2021, 12 (7) , 2521-2526. https://doi.org/10.1039/D0SC05369D
  62. Xiang Gao, Zheng Cui, Yue-Jian Lin, Guo-Xin Jin. Construction of organometallic trefoil knots and one-dimensional chains featuring half-sandwich Cp*Rh corner units and an abnormal zwitterion ligand. Organic Chemistry Frontiers 2021, 8 (2) , 231-238. https://doi.org/10.1039/D0QO01279C
  63. Somnath Bej, Mandira Nandi, Pradyut Ghosh. A Cd( ii ) and Zn( ii ) selective naphthyl based [2]rotaxane acts as an exclusive Zn( ii ) sensor upon further functionalization with pyrene. Dalton Transactions 2021, 50 (1) , 294-303. https://doi.org/10.1039/D0DT03645E
  64. John R.J. Maynard, Stephen M. Goldup. Strategies for the Synthesis of Enantiopure Mechanically Chiral Molecules. Chem 2020, 6 (8) , 1914-1932. https://doi.org/10.1016/j.chempr.2020.07.012
Open PDF

Journal of the American Chemical Society

Cite this: J. Am. Chem. Soc. 2020, 142, 21, 9803–9808
Click to copy citationCitation copied!
https://doi.org/10.1021/jacs.0c03447
Published May 1, 2020

Copyright © 2020 American Chemical Society. This publication is licensed under CC-BY.

Article Views

9992

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. Enantioselective synthesis of mechanically planar chiral rotaxanes through metal-free active template N-acylation using a macrocycle lacking rotational symmetry and an electrophile with a point-chiral leaving group.

    Scheme 1

    Scheme 1. : (a) Achiral Rotaxane Synthesis by Active Template N-Acylation Using Rotationally Symmetrical Crown Ethers; (b) Racemic and Unoptimized Enantioselective Synthesis of a Mechanically Planar Chiral Rotaxane

    Figure 2

    Figure 2. (a) Enantioselective synthesis of mechanically planar chiral rotaxane 11. Reaction conditions: 2 equiv of amine 2, 1 equiv each of electrophile and crown ether 10, toluene, [0.14 M], –40 °C, 24 h. (b) Partial 1H NMR spectra (600 MHz, CDCl3, 295 K) of macrocycle 10 (top), rotaxane 11 (middle) and the corresponding unthreaded axle (bottom).

    Figure 3

    Figure 3. (a) Circular dichroism spectra (1.0 × 10–4 M, CH2Cl2, 298 K) of (+)-11 (red) and (−)-11 (blue), baseline corrected. (b) Chemical structure of racemic rotaxane 13. (c) X-ray crystal structure of racemic rotaxane 13, side-on view showing intercomponent hydrogen bonds (in green). Hydrogen bond lengths: N47H—O16, 2.20 Å; O49—HC6, 2.63 Å. Hydrogen bond angles: N47—H–O16, 158.4°; O49—H–C6, 161.8°. (d) X-ray crystal structure of 13 viewed along the axle showing π-stacking between the macrocycle 1,2-dihydroxynaphthalene and axle bis(trifluoromethyl)phenyl rings. Centroid–centroid distance, 3.67 Å. Angle described by C40 and centroids, 97.6°. Solvate molecules and other hydrogen atoms omitted for clarity.

    Figure 4

    Figure 4. Tentative rationale for the transfer of chirality from Euclidean point-chirality (of the leaving group) to mechanical planar chirality (of the rotaxane). The lowest energy tetrahedral intermediates were modeled (see Supporting Information) using (a) electrophile 9 or (b) electrophile 12. The di(alkoxyl)naphthalene ring of the macrocycle and bis(trifluoromethyl)benzene unit originating from the nucleophile π-stack, causing the nitro-catechol ring to be positioned so as to cover one face of the tetrahedral center of the intermediate. This thermodynamically favored arrangement of components ensures the different handedness of the pseudoenantiomeric leaving groups is well-expressed in the diastereomeric transition states, resulting in enantioselectivity in the mechanically planar chiral rotaxane product. Hydrogen bonds are indicated by black dotted lines.

  • References


    This article references 23 other publications.

    1. 1
      (a) Bruns, C. J.; Stoddart, J. F. The Nature of the Mechanical Bond: From Molecules to Machines; John Wiley & Sons: Hoboken, NJ, 2017.
      (b) Jamieson, E. M. G.; Modicom, F.; Goldup, S. M. Chirality in rotaxanes and catenanes. Chem. Soc. Rev. 2018, 47, 52665311,  DOI: 10.1039/C8CS00097B
      (c) Evans, N. H. Chiral catenanes and rotaxanes: Fundamentals and emerging applications. Chem. - Eur. J. 2018, 24, 31013112,  DOI: 10.1002/chem.201704149
      (d) Nakazono, K.; Takata, T. Mechanical chirality of rotaxanes: Synthesis and function. Symmetry 2020, 12, 144,  DOI: 10.3390/sym12010144
    2. 2

      If a prochiral or meso thread is incorporated into a rotaxane, then chirality can arise from the position of the ring on the axle. See:

      (a) Alvarez-Pérez, M.; Goldup, S. M.; Leigh, D. A.; Slawin, A. M. Z. A chemically-driven molecular information ratchet. J. Am. Chem. Soc. 2008, 130, 18361838,  DOI: 10.1021/ja7102394
      (b) Cakmak, Y.; Erbas-Cakmak, S.; Leigh, D. A. Asymmetric catalysis with a mechanically point-chiral rotaxane. J. Am. Chem. Soc. 2016, 138, 17491751,  DOI: 10.1021/jacs.6b00303
      (c) Dommaschk, M.; Echavarren, J.; Leigh, D. A.; Marcos, V.; Singleton, T. A. Dynamic control of chiral space through local symmetry breaking in a rotaxane organocatalyst. Angew. Chem., Int. Ed. 2019, 58, 1495514958,  DOI: 10.1002/anie.201908330
    3. 3

      Mechanical planar chirality can also result from confinement of an unsymmetrical macrocycle to one side of a nonprochiral axle possessing Dnh symmetry. See:

      (a) Mochizuki, Y.; Ikeyatsu, K.; Mutoh, Y.; Hosoya, S.; Saito, S. Synthesis of mechanically planar chiral rac-[2]rotaxanes by partitioning of an achiral [2]rotaxane: Stereoinversion induced by shuttling. Org. Lett. 2017, 19, 43474350,  DOI: 10.1021/acs.orglett.7b02043
      (b) Corra, S.; de Vet, C.; Groppi, J.; La Rosa, M.; Silvi, S.; Baroncini, M.; Credi, A. Chemical on/off switching of mechanically planar chirality and chiral anion recognition in a [2]rotaxane molecular shuttle. J. Am. Chem. Soc. 2019, 141, 91299133,  DOI: 10.1021/jacs.9b00941
    4. 4

      For examples of other stereochemical consequences of threading, see:

      (a) Fuller, A.-M. L.; Leigh, D. A.; Lusby, P. J. Sequence isomerism in [3]rotaxanes. J. Am. Chem. Soc. 2010, 132, 49544959,  DOI: 10.1021/ja1006838
      (b) Talotta, C.; Gaeta, C.; Qi, Z.; Schalley, C. A.; Neri, P. Pseudorotaxanes with self-sorted sequence and stereochemical orientation. Angew. Chem., Int. Ed. 2013, 52, 74377441,  DOI: 10.1002/anie.201301570
      (c) La Manna, P.; Talotta, C.; Gaeta, C.; Soriente, A.; De Rosa, M.; Neri, P. Threading of an inherently directional calixarene wheel with oriented ammonium axles. J. Org. Chem. 2017, 82, 89738983,  DOI: 10.1021/acs.joc.7b01388
      (d) Cui, J.-S.; Ba, Q.-K.; Ke, H.; Valkonen, A.; Rissanen, K.; Jiang, W. Directional shuttling of a stimuli-responsive cone-like macrocycle on a single-state symmetric dumbbell axle. Angew. Chem., Int. Ed. 2018, 57, 78097814,  DOI: 10.1002/anie.201803349
      (e) Zheng, L.-S.; Cui, J.-S.; Jiang, W. Biomimetic synchronized motion of two interacting macrocycles in [3]rotaxane-based molecular shuttles. Angew. Chem., Int. Ed. 2019, 58, 1513615141,  DOI: 10.1002/anie.201910318
      (f) Ng, A. W. H.; Yee, C.-C.; Au-Yeung, H. Y. Radial hetero[5]catenanes: peripheral isomer sequences of the interlocked macrocycles. Angew. Chem., Int. Ed. 2019, 58, 1737517382,  DOI: 10.1002/anie.201908576
    5. 5
      Ishiwari, F.; Nakazono, K.; Koyama, Y.; Takata, T. Induction of single handed helicity of polyacetylenes using mechanically chiral rotaxanes as chiral sources. Angew. Chem., Int. Ed. 2017, 56, 1485814862,  DOI: 10.1002/anie.201707926
    6. 6
      Kameta, N.; Nagawa, Y.; Karikomi, M.; Hiratani, K. Chiral sensing for amino acid derivative based on a [2]rotaxane composed of an asymmetric rotor and an asymmetric axle. Chem. Commun. 2006, 37143716,  DOI: 10.1039/b607251h
    7. 7
      Heard, A.; Goldup, S. M. Synthesis of a mechanically planar chiral rotaxane ligand for enantioselective catalysis. Chem. 2020, 6, 9941006,  DOI: 10.1016/j.chempr.2020.02.006
    8. 8
      Glen, P. E.; O’Neill, J. A. T.; Lee, A.-L. Synthesis of a C1-symmetric Box macrocycle and studies towards active-template synthesis of mechanically planar chiral rotaxanes. Tetrahedron 2013, 69, 5768,  DOI: 10.1016/j.tet.2012.10.069
    9. 9
      (a) Yamamoto, C.; Okamoto, Y.; Schmidt, T.; Jäger, R.; Vögtle, F. Enantiomeric resolution of cycloenantiomeric rotaxane, topologically chiral catenane, and pretzel-shaped molecules: Observation of pronounced circular dichroism. J. Am. Chem. Soc. 1997, 119, 1054710548,  DOI: 10.1021/ja971764q
      (b) Schalley, C. A.; Beizai, K.; Vögtle, F. On the way to rotaxane-based molecular motors: Studies in molecular mobility and topological chirality. Acc. Chem. Res. 2001, 34, 465476,  DOI: 10.1021/ar000179i
      (c) Kameta, N.; Hiratani, K.; Nagawa, Y. A novel synthesis of chiral rotaxanes via covalent bond formation. Chem. Commun. 2004, 466467,  DOI: 10.1039/b314744d
      (d) Hirose, K.; Ukimi, M.; Ueda, S.; Onoda, C.; Kano, R.; Tsuda, K.; Hinohara, Y.; Tobe, Y. The asymmetry is derived from mechanical interlocking of achiral axle and achiral ring components – Syntheses and properties of optically pure [2]rotaxanes. Symmetry 2018, 10, 20,  DOI: 10.3390/sym10010020
      (e) Gell, C. E.; McArdle-Ismaguilov, T. A.; Evans, N. H. Modulating the expression of chirality in a mechanically chiral rotaxane. Chem. Commun. 2019, 55, 15761579,  DOI: 10.1039/C8CC10044F
      (f) Gaedke, M.; Witte, F.; Anhäuser, J.; Hupatz, H.; Schröder, H. V.; Valkonen, A.; Rissanen, K.; Lützen, A.; Paulus, B.; Schalley, C. A. Chiroptical inversion of a planar chiral redox-switchable rotaxane. Chem. Sci. 2019, 10, 1000310009,  DOI: 10.1039/C9SC03694F
    10. 10
      Bordoli, R.; Goldup, S. M. An efficient approach to mechanically planar chiral rotaxanes. J. Am. Chem. Soc. 2014, 136, 48174820,  DOI: 10.1021/ja412715m
    11. 11
      Jinks, M. A.; de Juan, A.; Denis, M.; Fletcher, C. J.; Galli, M.; Jamieson, E. M. G.; Modicom, F.; Zhang, Z.; Goldup, S. M. Stereoselective synthesis of mechanically planar chiral rotaxanes. Angew. Chem., Int. Ed. 2018, 57, 1480614810,  DOI: 10.1002/anie.201808990
    12. 12
      Makita, Y.; Kihara, N.; Nakakoji, N.; Takata, T.; Inagaki, S.; Yamamoto, C.; Okamoto, Y. Catalytic asymmetric synthesis and optical resolution of planar chiral rotaxane. Chem. Lett. 2007, 36, 162163,  DOI: 10.1246/cl.2007.162
    13. 13
      De Bo, G.; Dolphijn, G.; McTernan, C. T.; Leigh, D. A. [2]Rotaxane formation by transition state stabilization. J. Am. Chem. Soc. 2017, 139, 84558457,  DOI: 10.1021/jacs.7b05640
    14. 14
      (a) Fielden, S. D. P.; Leigh, D. A.; McTernan, C. T.; Pérez-Saavedra, B.; Vitorica-Yrezabal, I. J. Spontaneous assembly of rotaxanes from a primary amine, crown ether and electrophile. J. Am. Chem. Soc. 2018, 140, 60496052,  DOI: 10.1021/jacs.8b03394
      (b) Tian, C.; Fielden, S. D. P.; Whitehead, G. F. S.; Vitorica-Yrezabal, I. J.; Leigh, D. A. Weak functional group interactions revealed through metal-free active template rotaxane synthesis. Nat. Commun. 2020, 11, 744,  DOI: 10.1038/s41467-020-14576-7
    15. 15
      (a) Hogan, J. C.; Gandour, R. D. Structural requirements for glyme catalysis in butylaminolysis of aryl acetates in chlorobenzene. Identification of -OCH2CH2OCH2CH2OCH2CH2O- as the optimal subunit for catalysis. J. Org. Chem. 1991, 56, 28212826,  DOI: 10.1021/jo00008a044
      (b) Basilio, N.; García-Río, L.; Mejuto, J. C.; Pérez-Lorenzo, M. A. New reaction pathway in the ester aminolysis catalyzed by glymes and crown ethers. J. Org. Chem. 2006, 71, 42804285,  DOI: 10.1021/jo060389u
    16. 16

      Chiral leaving groups have previously been employed in asymmetric substitution reactions. See:

      Lepore, S. D.; Mondal, D. Recent advances in heterolytic nucleofugal leaving groups. Tetrahedron 2007, 63, 51035122,  DOI: 10.1016/j.tet.2007.03.049
    17. 17
      Heller, D.; Buschmann, H.; Scharf, H.-D. Nonlinear temperature behavior of product ratios in selection processes. Angew. Chem., Int. Ed. Engl. 1996, 35, 18521854,  DOI: 10.1002/anie.199618521
    18. 18
      Cinchona alkaloids in synthesis and catalysis: Ligands, immobilization and organocatalysis; Eui Song, C., Ed.; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, 2009.
    19. 19
      (a) Hunter, C. A.; Sanders, J. K. M. The nature of π–π interactions. J. Am. Chem. Soc. 1990, 112, 55255534,  DOI: 10.1021/ja00170a016
      (b) Riwar, L.-J.; Trapp, N.; Kuhn, B.; Diederich, F. Substituent effects in parallel displaced π–π stacking interactions: Distance matters. Angew. Chem., Int. Ed. 2017, 56, 1125211257,  DOI: 10.1002/anie.201703744
    20. 20
      Stewart, J. J. P. Optimization of parameters for semiempirical methods V. Modification of NDDO approximations and application to 70 elements. J. Mol. Model. 2007, 13, 11731213,  DOI: 10.1007/s00894-007-0233-4
    21. 21
      Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, revision D.01; Gaussian, Inc.; Wallingford, CT, 2013.
    22. 22
      (a) Lim, J. Y. C.; Marques, I.; Félix, V.; Beer, P. D. A chiral halogen-bonding [3]rotaxane for the recognition and sensing of biologically relevant dicarboxylate anions. Angew. Chem., Int. Ed. 2018, 57, 584588,  DOI: 10.1002/anie.201711176
      (b) Ba̧k, K. M.; Porfyrakis, K.; Davis, J. J.; Beer, P. D. Exploiting the mechanical bond for molecular recognition and sensing of charged species. Mater. Chem. Front. 2020, 4, 10521073,  DOI: 10.1039/C9QM00698B
    23. 23
      (a) Fu, X.; Zhang, Q.; Rao, S.-J.; Qu, D.-H.; Tian, H. One-pot synthesis of a [c2]daisy-chain-containing hetero[4]rotaxane via a self-sorting strategy. Chem. Sci. 2016, 7, 16961701,  DOI: 10.1039/C5SC04844C
      (b) Rao, S.-J.; Zhang, Q.; Mei, J.; Ye, X.-H.; Gao, C.; Wang, Q.-C.; Qu, D.-H.; Tian, H. One-pot synthesis of hetero[6]rotaxane bearing three different kinds of macrocycle through a self-sorting process. Chem. Sci. 2017, 8, 67776783,  DOI: 10.1039/C7SC03232C
      (c) Chen, S.; Wang, Y.; Nie, T.; Bao, C.; Wang, C.; Xu, T.; Lin, Q.; Qu, D.-H.; Gong, X.; Yang, Y.; Zhu, L.; Tian, H. An artificial molecular shuttle operates in lipid bilayers for ion transport. J. Am. Chem. Soc. 2018, 140, 1799217998,  DOI: 10.1021/jacs.8b09580
      (d) Zheng, X.; Zhang, Y.; Cao, N.; Li, X.; Zhang, S.; Du, R.; Wang, H.; Ye, Z.; Wang, Y.; Cao, F.; Li, H.; Hong, X.; Sue, A. C.-H.; Yang, C.; Liu, W.-G.; Li, H. Coulombic-enhanced hetero radical pairing interactions. Nat. Commun. 2018, 9, 1961,  DOI: 10.1038/s41467-018-04335-0
      (e) Li, W.-J.; Wang, W.; Wang, X.-Q.; Li, M.; Ke, Y.; Yao, R.; Wen, J.; Yin, G.-Q.; Jiang, B.; Li, X.; Yin, P.; Yang, H. B. Daisy chain dendrimers: Integrated mechanically interlocked molecules with stimuli-induced dimension modulation feature. J. Am. Chem. Soc. 2020, 142, 84738482,  DOI: 10.1021/jacs.0c02475
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.0c03447.

    • Experimental procedures, synthesis and characterization data, including circular dichroism, chiral HPLC, NMR, MS, and X-ray crystallography data (PDF)

    • Crystallographic data for 13 (CIF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.