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.

  • Letter
  • Published:

Self-assembly of regular hollow icosahedra in salt-free catanionic solutions

Abstract

Self-assembled structures having a regular hollow icosahedral form (such as those observed for proteins of virus capsids) can occur as a result of biomineralization processes1, but are extremely rare in mineral crystallites2. Compact icosahedra made from a boron oxide have been reported3, but equivalent structures made of synthetic organic components such as surfactants have not hitherto been observed. It is, however, well known that lipids, as well as mixtures of anionic and cationic single chain surfactants, can readily form bilayers4,5 that can adopt a variety of distinct geometric forms: they can fold into soft vesicles or random bilayers (the so-called sponge phase) or form ordered stacks of flat or undulating membranes6. Here we show that in salt-free mixtures of anionic and cationic surfactants, such bilayers can self-assemble into hollow aggregates with a regular icosahedral shape. These aggregates are stabilized by the presence of pores located at the vertices of the icosahedra. The resulting structures have a size of about one micrometre and mass of about 1010 daltons, making them larger than any known icosahedral protein assembly7 or virus capsid8. We expect the combination of wall rigidity and holes at vertices of these icosahedral aggregates to be of practical value for controlled drug or DNA release.

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

Figure 1: Electron microscopy images of icosahedral aggregates.
Figure 2: Comparison of freeze-fracture (a) and cryo-TEM (b) images for two adjacent aggregates.
Figure 3: Scattering by dilute solutions of icosahedra.
Figure 4: Sketch of the aggregate structure.

Similar content being viewed by others

References

  1. Thompson, D. W. On Growth and Form (Cambridge Univ. Pres, Cambridge, 1961).

    Google Scholar 

  2. Kimoto, K. & Nishida, I. The crystal habits of small particles of aluminium and silver prepared by evaporation in clean atmosphere of argon. J. Appl. Phys. 16, 941–948 (1966).

    Article  Google Scholar 

  3. Hubert, H. H. et al. Icosahedral packing of B12 icosahedra in boron suboxide. Nature 391, 376–378 (1998).

    Article  ADS  Google Scholar 

  4. Sackmann, E. & Lipowsky, R. Handbook of Biological Physics Vol. 1B (ed. Hoff, A. J.) (North-Holland, Amsterdam, 1995).

    MATH  Google Scholar 

  5. Kaler, E. W., Murthy, A. K., Rodriguez, B. E. & Zasadzinski, J. A. N. Spontaneous vesicle formation in aqueous mixtures of single-tailed surfactants. Science 245, 1371–1374 (1989).

    Article  ADS  CAS  Google Scholar 

  6. Dubois, M. & Zemb, T. Swelling limits for bilayer microstructures: the implosion of lamellar structure versus disordered lamellae. Curr. Opin. Colloid Interf. Sci. 5, 27–37 (1997).

    Article  Google Scholar 

  7. Wals, J. et al. Tricord protease exists as an icosahedral supermolecule in vivo. Molecul. Cell. 1, 59–65 (1997).

    Article  Google Scholar 

  8. Klug, A., Franklin, R. E. & Humphrey-Owen, S. P. F. The crystal structure of Tipula iridescent virus as determined by Bragg reflection of visible light. Biochem. Biophys. Acta 32, 203–219 (1959).

    Article  CAS  Google Scholar 

  9. Dubois, M., Gulik-Kryzwicki, Th., Demé, B. & Zemb, T. Rigid organic nanodiscs of controlled size: A catanionic formulation. C. R. Acad. Sci. Paris II C 1(9), 567–575 (1998).

    Google Scholar 

  10. Scamehorn, J. F. & Hartwell, J H. in Surfactant Science Series (eds Scamehorn, J. F. & Harwell, J. H.) Vol. 33, Ch. 10 (Marcel Dekker Inc., New York, 1989).

    Google Scholar 

  11. Graciaa, A., Ghoulam, M. B., Marion, G. & Lachaise, J. Critical concentrations and compositions of mixed micelles of sodium dodecylbenzesulfonate, tetradecyltrimethylammonium bromide and polyethylenephenols. J. Phys. Chem. 93, 4167–4173 (1989).

    Article  CAS  Google Scholar 

  12. Schnur, J. Lipid tubules: a paradigm for molecularly engineered structures. Science 262, 1669–1675 (1993).

    Article  ADS  CAS  Google Scholar 

  13. Thomas, B. N., Safinya, C. R., Plano, R. J. & Clark, N. A. Lipid tubules self-assembly: length dependence on cooling rate through a first-order phase transition. Science 227, 1635–1638 (1995).

    Article  ADS  Google Scholar 

  14. Parente, R. A., Höchli, M. & Lentz, B. R. Morphology and phase behavior of two types of unilamellar vesicles prepared from synthetic phosphatidylcholines studies by freeze-fracture electron microscopy and calorimetry. Biochim. Biophys. Acta 812, 493–502 (1985).

    Article  CAS  Google Scholar 

  15. Regev, O., Marques, E. F. & Khan, A. Polymer-induced structural effects on catanionic vesicles: formation of faceted vesicles, disks, and cross-links. Langmuir 15, 642–645 (1999).

    Article  CAS  Google Scholar 

  16. Zemb, T., Dubois, M., Demé, B., Gulik-Krzywicki, T. Self-assembly of flat nanodiscs in salt-free catanionic surfactant solutions. Science 283, 816–820 (1999).

    Article  ADS  CAS  Google Scholar 

  17. Hyde, S. T. Topological transformations mediated by bilayer punctures: from sponge phases to bicontinuous monolayers and reversed sponges. Colloids Surf. A 129–130, 207–225 (1997).

    Article  Google Scholar 

  18. Fogden, A., Daicic, J. & Kidane, A. Undulating charges in fluid membranes and their bending constants. J. Phys. (Paris) II 7, 229–248 (1997).

    CAS  Google Scholar 

  19. Fogden, A. & Daicic, J. & Kidane, A. Bending rigidity of ionic surfactant interfaces with variable charge density: the salt-free case. Colloids Surf. A 129–130, 157–165 (1997).

    Article  Google Scholar 

  20. Nelson, D. R. & Spaepen, F. Polytetrahedral order in condensed matter. Solid State Phys. 42, 1–90 (1989).

    Article  CAS  Google Scholar 

  21. Dubois, M., Belloni, L., Zemb, T., Demé, B. & Gulik-Krzywicki, T. Formation of rigid nanodiscs: Edge formation and molecular separation. Prog. Colloid Polym. Sci. 115, 238–242 (2000).

    Article  CAS  Google Scholar 

  22. Radtchenko, I. L. et al. Assembly of alternated mutlivalent ion/polyelectrolyte layers on colloidal particles. Stability of the multilayers and encapsulation of macromolecules into polyelectrolyte capsules. J. Colloid Interf. Sci. 230, 272–280 (2000).

    Article  ADS  CAS  Google Scholar 

  23. Donath, E., Sukhorukov, G. B., Caruso, F., Davis, S. A. & Möhwald, H. Novel hollow polymer shells by colloid-templated assembly of polyelectrolytes. Angew. Chem. Int. Edn 37, 2202–2205 (1998).

    Article  CAS  Google Scholar 

  24. Spector, M. S. & Zasadsinski, J. A. Topology of multivesicular liposomes, a model biliquid foam. Langmuir 12, 4704–4708 (1996).

    Article  CAS  Google Scholar 

  25. Klug, A. & Caspar, D. L. D. The structure of small viruses. Adv. Vir. Res. 7, 225–325 (1960).

    Article  CAS  Google Scholar 

  26. Crick, F. H. & Watson, J. D. The structure of small viruses. Nature 177, 473–475 (1956).

    Article  ADS  CAS  Google Scholar 

Download references

Acknowledgements

We thank G. Sukhorukov and O. Tiourina for help in confocal optical microscopy, and I. Erk for cryo-TEM. We also thank J.-L. Sikorav, P. Timmins and B. W. Ninham for critical reading of the manuscript and suggestions.

Author information

Authors and Affiliations

Authors

Supplementary information

Fig. S1 (GIF 14 KB)

Differential scanning calorimetry (DSC) of a catanionic mixture taken at a heating rate of 2°C/min at initial mole ratio r= 0.6. The chain melting transition is located at Tf= 50°C.

Fig. S2 (GIF 12 KB)

Schematic phase diagram showing the formation path of icosahedra (partial diagram for Φ 0.10). Above the chain melting temperature, stable unilamellar vesicles are formed for volume fractions in the range 0.1% to 1%. For the same composition, this sample is in a biphasic domain at room temperature. Tie lines between a lamellar phase with cristallised bilayers (Lβ) and nearly pure water (W) are shown. The "water" phase contains of the order of 1mmole of charged surfactant and the excess counter-ions, as shown by conductivity measurements. Slow cooling towards room temperature induces a liquid-liquid phase separation between a Lβ and a large volume of optically isotropic flowing solution which contains dispersed icosahedra (ICO). The surfactant volume fraction is of the order of Φ = 10-3. At rest and room temperature, icosahedra slowly flocculate after a few months following the tie-lines. These equilibrium lines connect the water corner to the Lβ phase at maximum swelling.

Fig.S3 (GIF 15 KB)

Wide angle scattering shows the in-plane distribution of deuterated chains in the Lβ phase in equilibrium with icosahedra (volume fraction 10%). (H/D) designs mixed deuterated anionic and protonated cationic surfactant at a molar ratio r = 0.65. (H/H) corresponds to a sample of same composition, but containing only hydrogenated chains. The main sharp peak corresponds to the frozen hydrocarbon chains in Lβ configuration (qf = 1.52 Å-1), whereas a broad superstructure peak appears around qf= 1.52/√3, showing the local hexagonal alternated packing. The superstructure peak is broad, showing defects in the triangular tiling, consistently with the quasi-equivalence principle.

Fig. S4 (JPG 4 KB)

Optical confocal microscopy images taken with an oil immersion objective (Leica TCS SP, excitation wavelength 525 nm), using a cationic dye (Rhodamine 6G). Left is the transmission micrograph. Two other images are confocal cuts through the same object obtained by displacing the focal plane by steps of 0.5 mm. Vertices are preferentially coloured by the positively charged dye molecule. This is an indication that excess anionic component is more pronounced around the vertices of the icosahedra. Central vertix can be seen on the top of the object, five-fold symmetry on the second cut. Distances between vertices are close to microscope resolution.

Fig. S5 (JPG 16 KB)

Electron micrographs of microstructures obtained not enough (A) or when too much (B) excess anionic surfactant is present to form icosahedra. (A) Fragments coexisting with closed objects are observed when r=0.52. Near equimolarity, there is not enough myristic acid to form the required minimal number of pores per vesicle. Some aggregates than fragment into nanodiscs upon cooling. (B) Large planar structures decorated with holes do not fold into closed objects when r=0.65. Micrographs show the coexistence of flat fragments with some facetted vesicles (bar is 1µm)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Dubois, M., Demé, B., Gulik-Krzywicki, T. et al. Self-assembly of regular hollow icosahedra in salt-free catanionic solutions. Nature 411, 672–675 (2001). https://doi.org/10.1038/35079541

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/35079541

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