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:

Competition between chiral solvents and chiral monomers in the helical bias of supramolecular polymers

Abstract

Solute–solvent interactions are key for the assembly and proper functioning of biomacromolecules and play important roles in many fields of organic and polymer chemistry. Despite numerous reports describing the effects of (chiral) solvents on helical conformations of (supramolecular) polymers, the combination of chiral solvents and chiral monomers is unexplored. Here we report diastereomeric differences in the supramolecular polymerization of enantiomers of chiral triphenylene-2,6,10-tricarboxamides in chiral chlorinated solvents. Competition between the preferences induced by the stereocentres of the assembled monomers and those present in the solvent molecules results in unforeseen temperature-dependent solvation effects. By combining experiments and mathematical modelling, we show that the observed differences between enantiomers originate from the combined additive entropic effects of stereocentres present in the monomer and in the solvent. Remarkably, copolymerizations show that the chiral solvent can bias the copolymer helicity and thereby overrule the helical preference of the monomers. Our results highlight the importance of cumulative solvation effects in supramolecular polymerizations.

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: TTAs form supramolecular polymers of left- and right-handed helicities in chlorinated solvents.
Fig. 2: Synergistic effects of solvent in deracemization of supramolecular polymers.
Fig. 3: Chiral solvation is strong enough to break the mirror symmetry between (R)-1 and (S)-1.
Fig. 4: Optically active solvents dictate the helicity in the copolymerizations: majority-rules and ‘sergeants-and-soldiers’ experiments.
Fig. 5: AFM shows a difference in morphology of fibres spin-coated from chiral solvent.

Similar content being viewed by others

Data availability

All data supporting the findings are included in the manuscript and Supplementary Information. Individual data files are also available upon request from the corresponding authors.

References

  1. Kulkarni, M. & Mukherjee, A. Understanding B-DNA to A-DNA transition in the right-handed DNA helix: perspective from a local to global transition. Prog. Biophys. Mol. Biol. 128, 63–73 (2017).

    Article  CAS  PubMed  Google Scholar 

  2. Fuller, W., Forsyth, T. & Mahendrasingam, A. Water–DNA interactions as studied by X-ray and neutron fibre diffraction. Philos. Trans. R. Soc. B Biol. Sci. 359, 1237–1248 (2004).

    Article  CAS  Google Scholar 

  3. Siebert, T., Guchhait, B., Liu, Y., Fingerhut, B. P. & Elsaesser, T. Range, magnitude, and ultrafast dynamics of electric fields at the hydrated DNA surface. J. Phys. Chem. Lett. 7, 3131–3136 (2016).

    Article  CAS  PubMed  Google Scholar 

  4. Chong, S. H. & Ham, S. Anomalous dynamics of water confined in protein–protein and protein–DNA interfaces. J. Phys. Chem. Lett. 7, 3967–3972 (2016).

    Article  CAS  PubMed  Google Scholar 

  5. Chong, S. H. & Ham, S. Dynamics of hydration water plays a key role in determining the binding thermodynamics of protein complexes. Sci. Rep. 7, 8744 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  6. Oshima, H., Hayashi, T. & Kinoshita, M. Statistical thermodynamics for actin–myosin binding: the crucial importance of hydration effects. Biophys. J. 110, 2496–2506 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Hunter, C. A. Quantifying intermolecular interactions: guidelines for the molecular recognition toolbox. Angew. Chem. Int. Ed. 43, 5310–5324 (2004).

    Article  CAS  Google Scholar 

  8. Kanagaraj, K., Alagesan, M., Inoue, Y. & Yang, C. in Comprehensive Supramolecular Chemistry Vol. II (eds Atwood, J., Gokel, G. W. & Barbour, L.) 11–60 (Elsevier, 2017).

  9. Chandler, D. Interfaces and the driving force of hydrophobic assembly. Nature 437, 640–647 (2005).

    Article  CAS  PubMed  Google Scholar 

  10. Cabot, R. & Hunter, C. A. Molecular probes of solvation phenomena. Chem. Soc. Rev. 41, 3485–3492 (2012).

    Article  CAS  PubMed  Google Scholar 

  11. Otto, S. The role of solvent cohesion in nonpolar solvation. Chem. Sci. 4, 2953–2959 (2013).

    Article  CAS  Google Scholar 

  12. Van Zee, N. J. et al. Potential enthalpic energy of water in oils exploited to control supramolecular structure. Nature 558, 100–103 (2018).

    Article  PubMed  CAS  Google Scholar 

  13. Dong, S. et al. Structural water as an essential comonomer in supramolecular polymerization. Sci. Adv. 3, eaao0900 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  14. Yashima, E., Matsushima, T. & Okamoto, Y. Chirality assignment of amines and amino alcohols based on circular dichroism induced by helix formation of a stereoregular poly((4-carboxyphenyl)acetylene) through acid–base complexation. J. Am. Chem. Soc. 119, 6345–6359 (1997).

    Article  CAS  Google Scholar 

  15. Rao, K. V., Miyajima, D., Nihonyanagi, A. & Aida, T. Thermally bisignate supramolecular polymerization. Nat. Chem. 9, 1133–1139 (2017).

    Article  CAS  Google Scholar 

  16. Biot, M. & Rapport de, M. Biot sur ce mémoire. C. R. Acad. Sci. 29, 433–447 (1849).

    Google Scholar 

  17. Pasteur, L. Recherches sur les relations qui peuvent exister entre la forme crystalline, la composition chimique et le sens de la polarisation rotatoire. Ann. Chim. Phys. 24, 442–459 (1848).

    Google Scholar 

  18. Cantekin, S. et al. A stereoselectively deuterated supramolecular motif to probe the role of solvent during self-assembly processes. Chem. Commun. 48, 3803–3805 (2012).

    Article  CAS  Google Scholar 

  19. Jacques, J., Collet, A. & Wilen, S. H. Enantiomers, Racemates and Resolutions (Krieger, 1991).

  20. Groen, M. B., Schadenberg, H. & Wynberg, H. Synthesis and resolution of some heterohelicenes. J. Org. Chem. 36, 2797–2809 (1971).

    Article  Google Scholar 

  21. Lüttringhaus, A. & Berrer, D. Zur Struktur der Lösungen—III. Racemat-spaltung durch ein optisch aktives Lösungsmittel. Tetrahedron Lett. 10, 10–12 (1959).

    Article  Google Scholar 

  22. Tulashie, S. K. The Potential of Chiral Solvents in Enantioselective Crystallization (Otto-von-Guericke Universität Magdeburg, 2010).

  23. Yang, L., Adam, C., Nichol, G. S. & Cockroft, S. L. How much do van der Waals dispersion forces contribute to molecular recognition in solution? Nat. Chem. 5, 1006–1010 (2013).

    Article  CAS  PubMed  Google Scholar 

  24. Fujiki, M. Supramolecular chirality: solvent chirality transfer in molecular chemistry and polymer chemistry. Symmetry 6, 677–703 (2014).

    Article  CAS  Google Scholar 

  25. Green, M. M., Khatri, C. & Peterson, N. C. A macromolecular conformational change driven by a minute chiral solvation energy. J. Am. Chem. Soc. 115, 4941–4942 (1993).

    Article  CAS  Google Scholar 

  26. Palmans, A. R. A., Vekemans, J. A. J. M., Havinga, E. E. & Meijer, E. W. Sergeants-and-soldiers principle in chiral columnar stacks of disc-shaped molecules with C3 symmetry. Angew. Chem. Int. Ed. 36, 2648–2651 (1997).

    Article  CAS  Google Scholar 

  27. Kawagoe, Y., Fujiki, M. & Nakano, Y. Limonene magic: noncovalent molecular chirality transfer leading to ambidextrous circularly polarised luminescent π-conjugated polymers. New J. Chem. 34, 637–647 (2010).

    Article  CAS  Google Scholar 

  28. Nakashima, H., Koe, J. R., Torimitsu, K. & Fujiki, M. Transfer and amplification of chiral molecular information to polysilylene aggregates. J. Am. Chem. Soc. 123, 4847–4848 (2001).

    Article  CAS  PubMed  Google Scholar 

  29. Isare, B. et al. Chirality in dynamic supramolecular nanotubes induced by a chiral solvent. Chem. Eur. J. 16, 173–177 (2010).

    Article  CAS  PubMed  Google Scholar 

  30. Stepanenko, V., Li, X. Q., Gershberg, J. & Würthner, F. Evidence for kinetic nucleation in helical nanofiber formation directed by chiral solvent for a perylene bisimide organogelator. Chem. Eur. J. 19, 4176–4183 (2013).

    Article  CAS  PubMed  Google Scholar 

  31. Nagata, Y., Takeda, R. & Suginome, M. Asymmetric catalysis in chiral solvents: chirality transfer with amplification of homochirality through a helical macromolecular scaffold. ACS Cent. Sci. 5, 1235–1240 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Green, M. M., Andreola, C., Muñoz, B., Reidy, M. P. & Zero, K. Macromolecular stereochemistry: a cooperative deuterium isotope effect leading to a large optical rotation. J. Am. Chem. Soc. 110, 4063–4065 (1988).

    Article  CAS  Google Scholar 

  33. Nagata, Y. et al. Elucidating the solvent effect on the switch of the helicity of poly(quinoxaline-2,3-diyl)s: a conformational analysis by small-angle neutron scattering. J. Am. Chem. Soc. 140, 2722–2726 (2018).

    Article  CAS  PubMed  Google Scholar 

  34. van’t Hoff, J. H. Die Lagerung Der Atome in Raume (FB&C, 2018).

  35. Green, M. M. et al. Majority rules in the copolymerization of mirror image isomers. J. Am. Chem. Soc. 117, 4181–4182 (1995).

    Article  CAS  Google Scholar 

  36. Green, M. M. et al. Macromolecular stereochemistry: the out-of-proportion influence of optically active comonomers on the conformational characteristics of polyisocyanates. The sergeants and soldiers experiment. J. Am. Chem. Soc. 111, 6452–6454 (1989).

    Article  Google Scholar 

  37. Cantekin, S., de Greef, T. F. A. & Palmans, A. R. A. Benzene-1,3,5-tricarboxamide: a versatile ordering moiety for supramolecular chemistry. Chem. Soc. Rev. 41, 6125–6137 (2012).

    Article  CAS  PubMed  Google Scholar 

  38. Markvoort, A. J., ten Eikelder, H. M. M., Hilbers, P. A. J., de Greef, T. F. A. & Meijer, E. W. Theoretical models of nonlinear effects in two-component cooperative supramolecular copolymerizations. Nat. Commun. 2, 509 (2011).

    Article  PubMed  CAS  Google Scholar 

  39. Roman, M. et al. Supramolecular balance: using cooperativity to amplify weak interactions. J. Am. Chem. Soc. 132, 16818–16824 (2010).

    Article  CAS  PubMed  Google Scholar 

  40. Lv, Z., Chen, Z., Shao, K., Qing, G. & Sun, T. Stimuli-directed helical chirality inversion and bio-applications. Polymers 8, 310 (2016).

    Article  PubMed Central  CAS  Google Scholar 

  41. Go, M. et al. Temperature-controlled helical inversion of asymmetric triphenylamine-based supramolecular polymers; difference of handedness at the micro- and macroscopic levels. Org. Chem. Front. 6, 1100–1108 (2019).

    Article  CAS  Google Scholar 

  42. Harada, N., Nakanishi, K. & Berova, N. in Comprehensive Chirooptical Spectroscopy Vol. 2 (eds Berova, N., Polavarapu, P. L., Nakanashi, K. & Woody, R. W.) 115–166 (John Wiley, 2012).

  43. Kang, J. et al. C5-Symmetric chiral corannulenes: desymmetrization of bowl inversion equilibrium via ‘Intramolecular’ hydrogen-bonding network. J. Am. Chem. Soc. 136, 10640–10644 (2014).

    Article  CAS  PubMed  Google Scholar 

  44. Garza, A. J. Solvation entropy made simple. J. Chem. Theory Comput. 15, 3204–3214 (2019).

    Article  CAS  PubMed  Google Scholar 

  45. Green, M. M. et al. A helical polymer with a cooperative response to chiral information. Science 268, 1860–1866 (2006).

    Article  Google Scholar 

  46. Meyerhoffer, W. & van’t Hoff, J. H. Gleichgewichte der Stereomeren (B.G. Teubner, 1906).

  47. De Windt, L. N. J. et al. Detailed approach to investigate thermodynamically controlled supramolecular copolymerizations. Macromolecules 52, 7430–7438 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  48. Ślęczkowski, M. L., Meijer, E. W. & Palmans, A. R. A. Cooperative folding of linear poly(dimethyl siloxane)s via supramolecular interactions. Macromol. Rapid Commun. 38, 1700566 (2017).

    Article  CAS  Google Scholar 

  49. Stals, P. J. M., Smulders, M. M. J., Martín-Rapún, R., Palmans, A. R. A. & Meijer, E. W. Asymmetrically substituted benzene-1,3,5-tricarboxamides: self-assembly and odd–even effects in the solid state and in dilute solution. Chem. Eur. J. 15, 2071–2080 (2009).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank L. de Windt for providing BTA and SBTA monomers, C. Kulkarni for fruitful discussions about solvation, and B. Markvoort and H. ten Eikelder for discussion on the calculations. This work was partially supported by the European Union’s Horizon 2020 research and innovation programme (Marie Skłodowska‐Curie grant agreement number 642083), the National Science Center, Poland (MINIATURA 3, grant number 2019/03/X/ST3/02148) and NWO (TOP-PUNT grant 10018944). Part of this work was carried out on the Dutch national e-infrastructure with the support of the SURF Cooperative.

Author information

Authors and Affiliations

Authors

Contributions

M.L.Ś. initiated the project, synthesized all materials and performed the spectroscopic, spectrometric and calorimetric analyses. M.F.J.M. developed the mathematical models and performed the fits to the experimental data. P.Ś. performed AFM measurements. M.L.Ś., M.F.J.M., A.R.A.P. and E.W.M. wrote the manuscript with input from all authors.

Corresponding authors

Correspondence to Anja R. A. Palmans or E. W. Meijer.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

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

Extended data

Extended Data Fig. 1 Spectroscopic characterization of TTAs 1 in bulk and concentrated CHCl3 solutions.

a) FT-IR spectra of TTAs 1 in bulk; b) FT-IR spectra of (S)-1 in CHCl3 at different concentrations; c) CD-spectra of (S)-1 at 512 μM at 20 °C and −5 °C; d) UV-Vis spectra of (S)-1 at 20 °C and −5 °C; AFM image of 100 μM of e) (S)-1 and f) (R)-1 spin-coated from 1-chlorooctane. FT-IR spectra in bulk (a) show absorption bands that are characteristic for helical supramolecular polymers based on three-fold hydrogen bond in previously studied, analogous benzene-1,3,5-tricarboxamides49. This absorption band pattern is maintained in concentrated solutions of chloroform (b). CD spectra of the same solutions at 512 μM (c) show a positive CD effect which saturates at −5 °C, thereby confirming full aggregation, which is also supported by UV-Vis spectra (d), which show hypsochromic shift upon cooling from molecularly dissolved state. Therefore, based on these multiple spectroscopic methods, we conclude that all experiments discussed below, which show the same CD pattern, are indicative of the formation of helically arranged monomers, stabilized by threefold hydrogen bonds within the supramolecular polymers. In addition, the AFM images of solutions of (S)-1 and (R)-1 spin-coated from 1-chlorooctane show long, fibrous structures obtained from 1-chlorooctane.

Extended Data Fig. 2 Cooling rate and concentration dependency on supramolecular polymerization of (R)-1.

a) VT-CD of (R)-1 in 1-chlorooctane at 300 μM at different cooling rates; b) VT-CD of (R)-1 in 1-chlorooctane at different concentrations at a cooling rate of 0.5 K min−1; c) VT-CD of (R)-1 in (S)-ClMeBu at different concentrations at a cooling rate of 0.5 K min−1; d) VT-CD of (R)-1 in (S)-CldMeOct at different concentrations at a cooling rate of 0.1 K min−1. Dashed lines in (b), (c) and (d) signify helix inversion temperature. Superimposable VT-CD of (R)-1 at 300 μM at variable cooling rate (a) show that the two competing helical states P and M are in thermodynamic equilibrium. When the concentration of the monomer is lowered (b), the stable aggregates are formed below the transition temperature, thus only M-helical aggregates for (R)-1 are observed. When the same experiments is carried in other solvents (c and d), only the transition temperature is shifted, while the competition between the states PR and MR remains monomer-concentration independent, supporting our thesis that it is the differential solvation that drives the thermodynamic equilibrium towards one of the helical states.

Extended Data Fig. 3 Influence of chemical structure and e.e. of the chiral solvent on the polymerization of n-1.

a) CD spectra of n-1 in different chiral, non-racemic solvents; b) VT-CD of n-1 of the samples from experiments showed on a); c) VT-CD of n-1 in 1-chloro-2-methylbutane at variable enantiomeric excess (e.e.) of the solvent. Optical path length: 0.1 cm. Complete mirror-symmetry breaking is achieved with use of the optically active solvents as confirmed by the values of the calculated molar circular dichroism from CD spectra in chiral solvents (a), Δε, of n-1, which are identical to the ones calculated for (S)-1 and (R)-1. The values were calculated using equation: ∆ε = (CD effect)/(32980 ∙ c ∙ l). Wherein c is the concentration of TTA in mol L-1 and l is the optical path length in cm. The determined values are: Δεn-1 = 182 L mol−1 cm-1 (in (S)-CldMeOct) and Δε(R)-1 = 187 L mol−1 cm−1 (in 1-chlorooctane). Optically active solvents induce only a single helical preference in the regime measured, as no transition between helical states is observed on the VT-CD (b). VT-CD of n-1 in ClMeBu at various e.e. shows non-linear dependency of the helical bias on the e.e. of the solvent (c). The results show an interesting insight into the solvation of n-1. It is clearly visible that at higher temperatures the helical bias is lower for all solutions below 25% of the solvent e.e. The increase of the bias is visible only at lower temperatures, where solvation is higher.

Extended Data Fig. 4 Influence of chemical structure and e.e. of the chiral solvent on the polymerization of (R)-1 and (S)-1.

a) VT-CD of (R)-1 in chiral solvents at rate 0.5 K min−1; b) VT-CD of (S)-1 and (R)-1 in (S)-CldMeOct (cooling rate 0.1 K min−1, 30 µM); c) VT-CD of (S)-1 in chiral solvents; d) VT-CD at 266 nm of (R)-1 (25 µM) and (S)-1 (20 µM) in (R)-CldMeOct (cooling rate 0.1 K min−1); e) VT-CD of (R)-1 in ClMeBu with varying e.e. of the solvent. VT-CD experiments of (R)-1 and (S)-1 in chiral solvents (a, c) show different profiles. Mirror-symmetry of the interactions is depicted by VT-CD of (S)-1 and (R)-1 in (S)-CldMeOct (b) and (R)-CldMeOct (d). Due to mirror-symmetry in interactions between (S)-1 / (S)-CldMeOct and (R)-1/ (R)-CldMeOct and vice versa, the helix inversion temperatures are identical. The sensitivity of these systems towards chiral solvation is further corroborated with the VT-CD of (R)-1in ClMeBu at varied e.e of the solvent (e). As the e.e is varied from racemic to enantiopure (S)-ClMeBu, the relative stabilities of the P and M helical aggregates are shifted and the transition temperature between the two helical states shifts to a lower temperature. Please note that concentrations in Figures b and d are different due to measurement error as well as concentrations of (S)-1 and (R)-1 in Figure d are also not equal. Since we assumed concentration-independence, we did not correct the experiments.

Extended Data Fig. 5 Calorimetric characterization of the helix inversion of (S)-1.

a) Micro-DSC curves of (S)-1 in (−)-MntCl (500 μM, 1 K min−1); b) VT-CD of (S)-1 solutions used for Micro-DSC experiments. Micro-DSC of (S)-1 in (−)-MntCl at 500 µM reveals a very small but reversible enthalpy change (a) which is observed around the temperature where the P-M transition occurs in the VT-CD experiments of the same samples (b). The exothermic transition upon cooling suggests that higher solvation is observed at low temperatures, which is a common phenomenon in crystals. The thermal hysteresis between heating and cooling, which is visible on the VT-CD curves is caused by high concentration which can alter the kinetics of the process due to high viscosity of the solution, but does not affect the thermodynamics.

Extended Data Fig. 6 Thermodynamic parameters obtained from the global fits of the homopolymerization of 1.

The thermodynamic parameters of (R)-1 and (S)-1 in (S)-CldMeOct obtained from the global fits of the thermodynamic mass-balance model to the experimental VT-CD data of all concentrations reported in Fig. 3 in the main text. aΔHe, X and ΔSe, X indicate the enthalpy and entropy of elongation of the polymers with X helicity by addition of a monomer. bNPX indicates the enthalpic nucleation penalty, which relates the formation of X-helical nuclei to the elongation of polymers with that helicity. cR is the gas constant, Δν the difference in the solvent interaction term between the polymer present above and below the inversion temperature and [solv] indicates the concentration of the solvent. In the fits from which those values were obtained ΔSlow T, Δν, NPhigh T and ΔNP were used as fit parameters and the other parameters were derived from these values. See pages 11–13 in the Supplementary Information for details.

Extended Data Fig. 7 Influence of the addition of decalin on the polymerization of (R)-1 and (S)-1 in (S)-CldMeOct.

VT-CD experiments in (S)-CldMeOct / decalin solutions of a) (R)-1; b) (S)-1 at 20 μM (cooling rate: 0.1 K min-1, l = 1 cm). Dilution of (S)-CldMeOct with decalin causes stabilization of the higher solvation state and a gradual shift of the P-M transition temperature (a and b). To explain this trend, we must set the opposite reference than it is shown on the Figure. The reference point is at 0% of (S)-CldMeOct (so at 100% decalin). At this composition the (S)-1 and (R)-1 aggregates are enantiomeric, and the VT-CD curves are mirror images. Titration of the decalin solutions with (S)-CldMeOct causes a non-symmetric response due to the arising diastereomeric relationship between (S)-1 and (R)-1 upon addition of the chiral, non-racemic solvent. Upon addition of (S)-CldMeOct, the M helices of (R)-1 as well as of (S)-1 are destabilized with respect to the P helices. Since M helices of (R)-1 are stable at lower temperatures, the P-M transition temperature of (R)-1 is moved faster towards low temperatures than the P-M transition temperature of (S)-1, which favors P helices at low temperatures. The preference of the solvent towards P helices is clearly visible on the copolymerization experiments (see Extended Data Fig. 8).

Extended Data Fig. 8 Additional plots and thermodynamic characterization of the copolymerization of (S)-1 and (R)-1 in (S)-CldMeOct.

a) VT-CD traces of the copolymerization of (S)-1 and (R)-1 in (S)-CldMeOct at different enantiomeric excesses of the monomer (c = 30 μM, cooling rate = 0.1 K min−1, optical path length = 0.1 cm). The vertical dashed line shows the temperature of intersection of all curves; b) Complete set of selected experimental (squares) and fitted (solid line) data of supramolecular copolymerization of (R)-1 and (S)-1 obtained from global fit; c) extracted CD effects at 266 nm of various compositions at temperatures of 20 °C and 81 °C; d) extracted P-M transition temperatures (Tinv) as function of the monomer composition. The copolymerization of (S)-1 with (R)-1 in (S)-CldMeOct reveals the solvent and the temperature effect on the preferred helicity of the copolymers. The chiral solvent generates an energetic difference between the P and M helices, thereby making them diastereomers. Clear stabilization of the P helix is visible since remarkably, at a temperature of 69 °C, no M helix can be present regardless of the monomer composition (a, dashed line). Moreover, plotting the CD effect at 20 and 81 °C always shows an excess of the P helices (positive CD effect) over M-helices (negative CD effect) with respect to the monomer e.e. (c). Interestingly, the P-M transition temperature with respect to the monomer e.e. shows a quasi-hyperbolic dependency (d), which is not symmetric due to the diastereomeric relationship. This means that addition of (S)-1 to (R)-1 at high (R)-1 content has an opposite effect on relative stability of P and M helices than addition of (R)-1 to (S)-1 at high (S)-1 content.

Supplementary information

Supplementary Information

Synthetic procedures, Supplementary Fig. 1 and Table 1.

Matlab scripts

Matlab files for the fitting of the solvent-dependent homopolymerizations and copolymerizations.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ślęczkowski, M.L., Mabesoone, M.F.J., Ślęczkowski, P. et al. Competition between chiral solvents and chiral monomers in the helical bias of supramolecular polymers. Nat. Chem. 13, 200–207 (2021). https://doi.org/10.1038/s41557-020-00583-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41557-020-00583-0

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