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:

Hyperexpandable, self-healing macromolecular crystals with integrated polymer networks

An Author Correction to this article was published on 04 July 2018

This article has been updated

Abstract

The formation of condensed matter typically involves a trade-off between structural order and flexibility. As the extent and directionality of interactions between atomic or molecular components increase, materials generally become more ordered but less compliant, and vice versa. Nevertheless, high levels of structural order and flexibility are not necessarily mutually exclusive; there are many biological (such as microtubules1,2, flagella3, viruses4,5) and synthetic assemblies (for example, dynamic molecular crystals6,7,8,9 and frameworks10,11,12,13) that can undergo considerable structural transformations without losing their crystalline order and that have remarkable mechanical properties8,14,15 that are useful in diverse applications, such as selective sorption16, separation17, sensing18 and mechanoactuation19. However, the extent of structural changes and the elasticity of such flexible crystals are constrained by the necessity to maintain a continuous network of bonding interactions between the constituents of the lattice. Consequently, even the most dynamic porous materials tend to be brittle and isolated as microcrystalline powders14, whereas flexible organic or inorganic molecular crystals cannot expand without fracturing. Owing to their rigidity, crystalline materials rarely display self-healing behaviour20. Here we report that macromolecular ferritin crystals with integrated hydrogel polymers can isotropically expand to 180 per cent of their original dimensions and more than 500 per cent of their original volume while retaining periodic order and faceted Wulff morphologies. Even after the separation of neighbouring ferritin molecules by 50 ångströms upon lattice expansion, specific molecular contacts between them can be reformed upon lattice contraction, resulting in the recovery of atomic-level periodicity and the highest-resolution ferritin structure reported so far. Dynamic bonding interactions between the hydrogel network and the ferritin molecules endow the crystals with the ability to resist fragmentation and self-heal efficiently, whereas the chemical tailorability of the ferritin molecules enables the creation of chemically and mechanically differentiated domains within single crystals.

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: Packing arrangement in ferritin crystals and their expansion–contraction mediated by the infused hydrogel network.
Fig. 2: Characterization of the expansion and contraction behaviour of ferritin crystal–hydrogel hybrids.
Fig. 3: Atomic-level structural characterization of ferritin crystal–hydrogel hybrids by XRD.
Fig. 4: Self-healing behaviour and functionalization of ferritin crystal–hydrogel hybrids.

Similar content being viewed by others

Change history

  • 04 July 2018

    Change history: In this Letter, Alexander Groisman should have been listed as an author. This error has been corrected online.

References

  1. Hawkins, T., Mirigian, M., Yasar, M. S. & Ross, J. L. Mechanics of microtubules. J. Biomech. 43, 23–30 (2010).

    Article  Google Scholar 

  2. Fletcher, D. A. & Mullins, R. D. Cell mechanics and the cytoskeleton. Nature 463, 485–592 (2010).

    Article  ADS  CAS  Google Scholar 

  3. Block, S. M., Blair, D. F. & Berg, H. C. Compliance of bacterial flagella measured with optical tweezers. Nature 338, 514–518 (1989).

    Article  ADS  CAS  Google Scholar 

  4. Lewis, J. K., Bothner, B., Smith, T. J. & Siuzdak, G. Antiviral agent blocks breathing of the common cold virus. Proc. Natl Acad. Sci. USA 95, 6774–6778 (1998).

    Article  ADS  CAS  Google Scholar 

  5. Lok, S.-M. et al. Binding of a neutralizing antibody to dengue virus alters the arrangement of surface glycoproteins. Nat. Struct. Mol. Biol. 15, 312–317 (2008).

    Article  CAS  Google Scholar 

  6. Kobatake, S., Takami, S., Muto, H., Ishikawa, T. & Irie, M. Rapid and reversible shape changes of molecular crystals on photoirradiation. Nature 446, 778–781 (2007).

    Article  ADS  CAS  Google Scholar 

  7. Kim, T., Al-Muhanna, M. K., Al-Suwaidan, S. D., Al-Kaysi, R. O. & Bardeen, C. J. Photoinduced curling of organic molecular crystal nanowires. Angew. Chem. Int. Edn 52, 6889–6893 (2013).

    Article  CAS  Google Scholar 

  8. Panda, M. K. et al. Spatially resolved analysis of short-range structure perturbations in a plastically bent molecular crystal. Nat. Chem. 7, 65–72 (2015).

    Article  CAS  Google Scholar 

  9. Naumov, P., Chizhik, S., Panda, M. K., Nath, N. K. & Boldyreva, E. Mechanically responsive molecular crystals. Chem. Rev. 115, 12440–12490 (2015).

    Article  CAS  Google Scholar 

  10. Barthelet, K., Marrot, J., Riou, D. & Ferey, G. A breathing hybrid organic–inorganic solid with very large pores and high magnetic characteristics. Angew. Chem. Int. Edn Engl. 41, 281–284 (2002).

    Article  CAS  Google Scholar 

  11. Sakata, Y. et al. Shape-memory nanopores induced in coordination frameworks by crystal downsizing. Science 339, 193–196 (2013).

    Article  ADS  CAS  Google Scholar 

  12. Rabone, J. et al. An adaptable peptide-based porous material. Science 329, 1053–1057 (2010).

    Article  ADS  CAS  Google Scholar 

  13. Suzuki, Y. et al. Self-assembly of coherently dynamic, auxetic, two-dimensional protein crystals. Nature 533, 369–373 (2016).

    Article  ADS  CAS  Google Scholar 

  14. Serre, C. et al. Role of solvent-host interactions that lead to very large swelling of hybrid frameworks. Science 315, 1828–1831 (2007).

    Article  ADS  CAS  Google Scholar 

  15. Worthy, A. et al. Atomic resolution of structural changes in elastic crystals of copper(ii) acetylacetonate. Nat. Chem. 10, 65–69 (2018).

    Article  Google Scholar 

  16. Mason, J. A. et al. Methane storage in flexible metal–organic frameworks with intrinsic thermal management. Nature 527, 357–361 (2015).

    Article  ADS  CAS  Google Scholar 

  17. Couck, S. et al. An amine-functionalized MIL-53 metal–organic framework with large separation power for CO2 and CH4. J. Am. Chem. Soc. 131, 6326–6327 (2009).

    Article  CAS  Google Scholar 

  18. Chen, Q. et al. A controllable gate effect in cobalt(ii) organic frameworks by reversible structure transformations. Angew. Chem. Int. Edn 52, 11550–11553 (2013).

    Article  CAS  Google Scholar 

  19. Ghosh, S. & Reddy, C. M. Elastic and bendable caffeine cocrystals: implications for the design of flexible organic materials. Angew. Chem. Int. Edn 51, 10319–10323 (2012).

    Article  CAS  Google Scholar 

  20. Commins, P., Hara, H. & Naumov, P. Self-healing molecular crystals. Angew. Chem. Int. Edn 55, 13028–13032 (2016).

    Article  CAS  Google Scholar 

  21. Tanaka, T. et al. Phase transitions in ionic gels. Phys. Rev. Lett. 45, 1636–1639 (1980).

    Article  ADS  CAS  Google Scholar 

  22. Phadke, A. et al. Rapid self-healing hydrogels. Proc. Natl Acad. Sci. USA 109, 4383–4388 (2012).

    Article  ADS  CAS  Google Scholar 

  23. Holtz, J. H. & Asher, S. A. Polymerized colloidal crystal hydrogel films as intelligent chemical sensing materials. Nature 389, 829–832 (1997).

    Article  ADS  CAS  Google Scholar 

  24. Chen, F., Tillberg, P. W. & Boyden, E. S. Expansion microscopy. Science 347, 543–548 (2015).

    Article  ADS  CAS  Google Scholar 

  25. Elliott, J. E., Macdonald, M., Nie, J. & Bowman, C. N. Structure and swelling of poly (acrylic acid) hydrogels: effect of pH, ionic strength, and dilution on the crosslinked polymer structure. Polymer 45, 1503–1510 (2004).

    Article  CAS  Google Scholar 

  26. Theil, E. C. Ferritin: structure, gene regulation, and cellular function in animals, plants, and microorganisms. Annu. Rev. Biochem. 56, 289–315 (1987).

    Article  CAS  Google Scholar 

  27. Lawson, D. M. et al. Solving the structure of human H ferritin by genetically engineering intermolecular crystal contacts. Nature 349, 541–544 (1991).

    Article  ADS  CAS  Google Scholar 

  28. Kaya, D., Pekcan, Ö. & Yılmaz, Y. Direct test of the critical exponents at the sol-gel transition. Phys. Rev. E 69, 016117 (2004).

    Article  ADS  Google Scholar 

  29. Strandman, S. & Zhu, X. Self-healing supramolecular hydrogels based on reversible physical interactions. Gels 2, 16 (2016).

    Article  Google Scholar 

  30. Denisin, A. K. & Pruitt, B. L. Tuning the range of polyacrylamide gel stiffness for mechanobiology applications. Appl. Mater. Interfaces 8, 21893–21902 (2016).

    Article  CAS  Google Scholar 

  31. Huard, D. J., Kane, K. M. & Tezcan, F. A. Re-engineering protein interfaces yields copper-inducible ferritin cage assembly. Nat. Chem. Biol. 9, 169–176 (2013).

    Article  CAS  Google Scholar 

  32. Sontz, P. A., Bailey, J. B., Ahn, S. & Tezcan, F. A. A metal organic framework with spherical protein nodes: rational chemical design of 3D protein crystals. J. Am. Chem. Soc. 137, 11598–11601 (2015).

    Article  CAS  Google Scholar 

  33. Bradford, M. M. A rapid and sensitive method for the quantification of microgram quantities of protein utilizing the principle of protein-dye binding. Anal. Biochem. 72, 248–254 (1976).

    Article  CAS  Google Scholar 

  34. Battye, T. G. G., Kontogiannis, L., Johnson, O., Powell, H. R. & Leslie, A. G. iMOSFLM: a new graphical interface for diffraction-image processing with MOSFLM. Acta Crystallogr. D 67, 271–281 (2011).

    Article  CAS  Google Scholar 

  35. Evans, P. R. & Murshudov, G. N. How good are my data and what is the resolution? Acta Crystallogr. D 69, 1204–1214 (2013).

    Article  CAS  Google Scholar 

  36. McCoy, A. J. et al. Phaser crystallographic software. J. Appl. Cryst. 40, 658–674 (2007).

    Article  CAS  Google Scholar 

  37. Adams, P. D. et al. PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr. D 66, 213–221 (2010).

    Article  CAS  Google Scholar 

  38. Emsley, P., Lohkamp, B., Scott, W. G. & Cowtan, K. Coot: model-building tools for molecular graphics. Acta Crystallogr. D 66, 486–501 (2010).

    Article  CAS  Google Scholar 

  39. Kleywegt, G. J. & Jones, T. A. Detection, delineation, measurement and display of cavities in macromolecular structures. Acta Crystallogr. D 50, 178–185 (1994).

    Article  CAS  Google Scholar 

  40. The PyMOL Molecular Graphics System Version 1.3, https://pymol.org/2/support.html (Schrödinger LLC).

  41. Levi, S. et al. Mechanism of ferritin iron uptake: activity of the H-chain and deletion mapping of the ferro-oxidase site. J. Biol. Chem. 263, 18086–18092 (1988).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank the following colleagues for assistance: R. Alberstein for data processing; A. Rheingold, C. Moore and M. Gembicky for XRD; S. Weigand, T. Weiss and I. Rajkovic for SAXS; W.-J. Rappel for confocal microscopy; Z. Hu for performing the nanoindentation experiments. This work was primarily funded by the US Department of Energy, DOE (BES, Division of Materials Sciences, Biomolecular Materials Program, DE-SC0003844 to F.A.T.). Additional funding was provided by NSF (DMR-1602537 to F.A.T. for SAXS studies). Crystallographic data were collected at Stanford Synchrotron Radiation Lightsource (SSRL) and the Crystallography Facility of the University of California, San Diego. SAXS data were collected at SSRL and the Advanced Photon Source. SSRL and the Advanced Photon Source are supported by the DOE Office of Science, Office of Basic Energy Sciences under contracts DE-AC02-76SF00515 and DE-AC02-06CH11357, respectively.

Author information

Authors and Affiliations

Authors

Contributions

L.Z., J.B.B. and F.A.T. conceived the work and designed the experiments. L.Z. conducted the majority of the experiments and performed light microscopy analysis. L.Z. and J.B.B. performed XRD and SAXS analyses. L.Z. and R.H.S. performed confocal microscopy analysis. A.G. designed the microfluidic cells. F.A.T. wrote the manuscript with contributions from all authors.

Corresponding author

Correspondence to F. Akif Tezcan.

Ethics declarations

Competing interests

L.Z., J.B.B. and F.A.T. have submitted a patent application based on the work described here.

Additional information

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

Extended data figures and tables

Extended Data Fig. 1 Distribution of electrostatic charge on the surface of ferritin and size distribution of ferritin in solution.

a, pH-dependent zeta potentials of ferritin, determined by dynamic-light-scattering measurements. b, Dynamic-light-scattering profile of ferritin (200 μM) in a solution of 50 mM HEPES (pH 7.0). c, Representation of the electrostatic charge distribution on the ferritin surface, as viewed along the two-, three- and four-fold symmetry axes. Positive (+5kBT/e) and negative (−5kBT/e) charges are shown in blue and red, respectively. kB, Boltzmann constant; e, electron charge.

Extended Data Fig. 2 Molecular diffusion and polymerization in ferritin crystals, monitored using confocal microscopy.

a, Diffusion of rhodamine B into a ferritin crystal over 15 min. b, c, In crystallo polymerization of the hydrogel network, monitored through the decrease of integrated pyranine fluorescence (green fluorescence channel). The corresponding bright-field (DIC) images show the diffusion of the aqueous NaCl solution into the crystal. The ring-shaped diffusion front becomes evident at time t = 108 s and disappears by t = 216 s. The crystal expands by approximately 5% (edge length) during polymerization. Scale bars in a and b correspond to 100 μm. d, Scanning electron microscopy images of native ferritin crystals (top) and crystal–hydrogel hybrids (bottom).

Extended Data Fig. 3 Quantification of an acrylic acid analogue using 19F NMR.

a, 19F-NMR spectrum, showing peak assignments for the trifluoroacetic acid standard, free 2-(trifluoromethyl)acrylic acid, and 2-(trifluoromethyl)acrylic acid incorporated into the polymer. b, Diagram illustrating the experimental protocol for the quantification of 2-(trifluoromethyl)acrylic acid uptake into the crystal lattice. The concentration of the 2-(trifluoromethyl)acrylic acid in the crystal lattice (155.6 mM) is approximately the same as its concentration in the soaking solution (see Methods for details).

Extended Data Fig. 4 Isotropic hyperexpansion of ferritin crystal–hydrogel hybrids.

a, b, Continuous expansion of two different crystal–hydrogel hybrids in deionized water, monitored using confocal microscopy. Crystal facets are still discernible after expansion for more than 2 h. Scale bars correspond to 100 μm. c, Ferritin release into the solution from expanding crystal–hydrogel hybrids (n > 10,000) over about 4 h. Negligible ferritin release is observed until about 1 h. Protein concentrations were determined using the Bradford assay. d, Confocal microscopy images of highly expanded crystal–hydrogel hybrids, showing the structural deterioration of the facets and the edges.

Extended Data Fig. 5 Expansion and contraction behaviour of crystal–hydrogel hybrids in the presence of different metal ions.

a, Light micrographs of the crystal–hydrogel hybrids at different stages of expansion and contraction in response to different metal ions. b, XRD patterns (T = 273 K) of expanded crystal–hydrogel hybrids, acquired upon contraction with different metal ions. Contraction with divalent cations (Ca, Mg, Cd, Zn, Ni and Co) reproducibly leads to the recovery of the full atomic-level order, whereas contraction with monovalent cations (Li, Na and K) only reinstates low-order diffraction peaks.

Extended Data Fig. 6 Successive expansion–contraction cycles for a single ferritin crystal–hydrogel hybrid.

Light micrographs of a hybrid crystal at pre- and post-expansion stages in each cycle are shown on the left, and the corresponding changes in edge length upon expansion–contraction are shown on the right. The separation between the major ticks of the ruler is 100 μm. The crystal expands to a lesser extent during the first expansion cycle, which we ascribe to residual CaCl2 (which forms strong polymer–polymer and protein–protein interactions) remaining in the solution that is transferred on the loop along with the crystal. The subsequent variability in the rate and extent of expansion is attributed to the different amounts of residual NaCl transferred in each cycle.

Extended Data Fig. 7 Alternative hydrogel formulations.

a, Alternative monomer combinations that yield successful in crystallo polymerization and crystal expansion. b, Monomer combinations that lead to crystal dissolution during polymerization. c, A crystal soaked in a solution containing polyacrylate (molecular weight, Mw = 2,100 Da) dissolves upon being transferred into water. The separation between the major ticks of the ruler is 100 μm. MBAm, N,N′-methylenebis(acrylamide).

Extended Data Fig. 8 SAXS imaging of a single crystal–hydrogel hybrid in a microfluidic chip.

a, Schematic diagram of the microfluidic chip. b, Side-view representations of the microfluidic chip. c, Photograph of the microfluidic chip, mounted on beamline 4-2 at SSRL. d, Single-crystal SAXS diffraction patterns observed at different stages of crystal expansion and contraction. The Miller indices of each visible spot are indicated. Reflections with the highest signal-to-noise ratio (II) are circled in red. e, Spot profiles of the highest-II reflections indicated in d.

Extended Data Fig. 9 Mechanical and thermal properties of ferritin crystal–hydrogel hybrids.

a, Light-microscopy images showing the fragmentation of a native ferritin crystal and of a crystal–hydrogel hybrid upon application of external force with a needle at the location indicated with the arrow. The separation between the major ticks of the ruler is 100 μm. b, Temperature dependence of the SAXS profiles of native ferritin crystals and crystal–hydrogel hybrids. The small-angle reflections (that is, periodic order) in both samples are maintained at 80 °C (the maximal temperature experimentally attainable). c, Determination of the hardness and reduced modulus of native ferritin crystals and crystal–hydrogel hybrids using atomic force microscopy nanoindentation measurements. d, Light-microscopy images showing the expansion and contraction of a crystal–hydrogel hybrid containing Fe-loaded ferritin molecules. The separation between the major ticks of the ruler is 100 μm.

Extended Data Table 1 X-ray data collection and refinement statistics

Supplementary information

Video 1: In-crystallo formation of hydrogel network monitored by pyranine quenching (see Extended Data Fig. 2 for static snapshots)

To monitor the process of in-crystallo polymerization, a ferritin crystal infused with polymer precursors pyranine is imaged by confocal microscopy, using both fluorescent (green) and differential interference contrast (DIC) channels. 10 µL polymerization solution is added onto the crystal to initiate in-crystallo polymerization. Fluorescence from pyranine rapidly decreases in the first minute, indicating that the polymerization process was complete in <2 min. Polymerization is followed by the intrusion of aqueous NaCl solution into the lattice; the movement of the solvent front into the center of the crystal is observed as a ring that progressively becomes smaller.

Video 2: Structural evolution of a ferritin crystal-hydrogel hybrid during polymerization, expansion and contraction (see Main Text Fig. 2 for static snapshots)

A ferritin crystal infused with polymer precursors was placed on a microscopic ruler and imaged by light microscopy. 10 µL of the polymerization solution is added to initiate in-crystallo polymerization. The diffusion ring of the aqueous NaCl is clearly visible. After polymerization, the crystal is placed onto a clean ruler, deionized water is added and crystal-hydrogel hybrid expansion proceeds for ca. 10 min, whereby the crystal grows to ca. 140% of its original dimensions. Contraction is then mediated by NaCl addition, and a further incremental amount of contraction is achieved by CaCl2 addition.

Video 3: Examples of self-healing behavior in ferritin crystal-hydrogel hybrids (see Main Text Figs. 4a and 4b for static snapshots)

This video depicts three crystals (Crystals 1, 2 and 3) that undergo self-healing. Crystal 1 and Crystal 2 are crystal-hydrogel hybrids that develop cracks during contraction with CaCl2, and then undergo self-healing. In the case of Crystal 3, cracks appear during polymerization but are healed spontaneously. These cracks continually reappear and disappear during crystal expansion, and are aggravated for a brief period during the rapid, NaCl-induced contraction, but they eventually self-heal almost completely.

Video 4: Expansion and contraction behavior of functionalized ferritin crystals (see Extended Data Fig. 9d, and Main Text Figs. 4d and 4e for static snapshots)

Part 1: A crystal-hydrogel hybrid containing Fe-loaded ferritin expands in water and contracts upon addition of NaCl. Part 2: An expandable core/expandable shell crystal-hydrogel hybrid (core is labeled with rhodamine) expands in water and then contracts upon addition of NaCl. Part 3: An non-expandable core/expandable shell crystal-hydrogel hybrid (core is labeled with rhodamine and crosslinked with glutaraldehyde) undergoes shell-fragmentation upon transfer into water.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhang, L., Bailey, J.B., Subramanian, R.H. et al. Hyperexpandable, self-healing macromolecular crystals with integrated polymer networks. Nature 557, 86–91 (2018). https://doi.org/10.1038/s41586-018-0057-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-018-0057-7

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