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:

Crosslinking ionic oligomers as conformable precursors to calcium carbonate

Abstract

Inorganic materials have essential roles in society, including in building construction, optical devices, mechanical engineering and as biomaterials1,2,3,4. However, the manufacture of inorganic materials is limited by classical crystallization5, which often produces powders rather than monoliths with continuous structures. Several precursors that enable non-classical crystallization—such as pre-nucleation clusters6,7,8, dense liquid droplets9,10, polymer-induced liquid precursor phases11,12,13 and nanoparticles14—have been proposed to improve the construction of inorganic materials, but the large-scale application of these precursors in monolith preparations is limited by availability and by practical considerations. Inspired by the processability of polymeric materials that can be manufactured by crosslinking monomers or oligomers15, here we demonstrate the construction of continuously structured inorganic materials by crosslinking ionic oligomers. Using calcium carbonate as a model, we obtain a large quantity of its oligomers (CaCO3)n with controllable molecular weights, in which triethylamine acts as a capping agent to stabilize the oligomers. The removal of triethylamine initiates crosslinking of the (CaCO3)n oligomers, and thus the rapid construction of pure monolithic calcium carbonate and even single crystals with a continuous internal structure. The fluid-like behaviour of the oligomer precursor enables it to be readily processed or moulded into shapes, even for materials with structural complexity and variable morphologies. The material construction strategy that we introduce here arises from a fusion of classic inorganic and polymer chemistry, and uses the same cross-linking process for the manufacture the materials.

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: Preparation and characterization of (CaCO3)n oligomers.
Fig. 2: Controllable crosslinking of (CaCO3)n oligomers.
Fig. 3: Construction of amorphous and single-crystalline-like CaCO3 bulk materials by the crosslinking of (CaCO3)n oligomers.
Fig. 4: Constructible engineering of CaCO3 single-crystalline materials by using (CaCO3)n oligomers.

Similar content being viewed by others

Data availability

The data that support the findings of this study are available from the corresponding author upon reasonable request.

References

  1. Meyers, M. A., Chen, P. Y., Lin, A. Y. M. & Seki, Y. Biological materials: structure and mechanical properties. Prog. Mater. Sci. 53, 1–206 (2008).

    CAS  Google Scholar 

  2. Riley, F. L. Silicon nitride and related materials. J. Am. Ceram. Soc. 83, 245–265 (2000).

    CAS  Google Scholar 

  3. Xu, H. H. K. et al. Calcium phosphate cements for bone engineering and their biological properties. Bone Res. 5, 17056 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Hu, C. L. & Li, Z. J. A review on the mechanical properties of cement-based materials measured by nanoindentation. Constr. Build. Mater. 90, 80–90 (2015).

    ADS  Google Scholar 

  5. Thanh, N. T. K., Maclean, N. & Mahiddine, S. Mechanisms of nucleation and growth of nanoparticles in solution. Chem. Rev. 114, 7610–7630 (2014).

    CAS  PubMed  Google Scholar 

  6. Gebauer, D., Kellermeier, M., Gale, J. D., Bergstrom, L. & Colfen, H. Pre-nucleation clusters as solute precursors in crystallisation. Chem. Soc. Rev. 43, 2348–2371 (2014).

    CAS  PubMed  Google Scholar 

  7. Gebauer, D., Volkel, A. & Colfen, H. Stable prenucleation calcium carbonate clusters. Science 322, 1819–1822 (2008).

    ADS  CAS  PubMed  Google Scholar 

  8. Kellermeier, M. et al. Colloidal stabilization of calcium carbonate prenucleation clusters with silica. Adv. Funct. Mater. 22, 4301–4311 (2012).

    CAS  Google Scholar 

  9. Wolf, S. E. et al. Carbonate-coordinated metal complexes precede the formation of liquid amorphous mineral emulsions of divalent metal carbonates. Nanoscale 3, 1158–1165 (2011).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  10. Smeets, P. J. M. et al. A classical view on nonclassical nucleation. Proc. Natl Acad. Sci. USA 114, E7882–E7890 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Gower, L. B. & Odom, D. J. Deposition of calcium carbonate films by a polymer-induced liquid-precursor (PILP) process. J. Cryst. Growth 210, 719–734 (2000).

    ADS  CAS  Google Scholar 

  12. Cheng, X. G. & Gower, L. B. Molding mineral within microporous hydrogels by a polymer-induced liquid-precursor (PILP) process. Biotechnol. Prog. 22, 141–149 (2006).

    CAS  PubMed  Google Scholar 

  13. Wolf, S. E. et al. Strong stabilization of amorphous calcium carbonate emulsion by ovalbumin: gaining insight into the mechanism of ‘polymer-induced liquid precursor’ processes. J. Am. Chem. Soc. 133, 12642–12649 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. De Yoreo, J. J. et al. Crystallization by particle attachment in synthetic, biogenic, and geologic environments. Science 349, aaa6760 (2015).

    PubMed  Google Scholar 

  15. Koltzenburg, S., Maskos, M. & Nuyken, O. (eds) Polymer Chemistry Ch. 14 (Springer, 2017).

  16. Lewis, J. A. Colloidal processing of ceramics. J. Am. Ceram. Soc. 83, 2341–2359 (2000).

    CAS  Google Scholar 

  17. Eckel, Z. C. et al. Additive manufacturing of polymer-derived ceramics. Science 351, 58–62 (2016).

    ADS  CAS  PubMed  Google Scholar 

  18. Liu, H. L., Gong, Q. H., Yue, Y. H., Guo, L. & Wang, X. Sub-1 nm nanowire based superlattice showing high strength and low modulus. J. Am. Chem. Soc. 139, 8579–8585 (2017).

    CAS  PubMed  Google Scholar 

  19. Truby, R. L. & Lewis, J. A. Printing soft matter in three dimensions. Nature 540, 371–378 (2016).

    ADS  CAS  PubMed  Google Scholar 

  20. Feinle, A., Elsaesser, M. S. & Husing, N. Sol-gel synthesis of monolithic materials with hierarchical porosity. Chem. Soc. Rev. 45, 3377–3399 (2016).

    CAS  PubMed  Google Scholar 

  21. Caminade, A. M., Hey-Hawkins, E. & Manners, I. Smart inorganic polymers. Chem. Soc. Rev. 45, 5144–5146 (2016).

    CAS  PubMed  Google Scholar 

  22. Drake, K., Mukherjee, I., Mirza, K., Ji, H. F. & Wei, Y. Phenylethynyl and phenol end-capping studies of polybiphenyloxydiphenylsilanes for cross-linking and enhanced thermal stability. Macromolecules 44, 4107–4115 (2011).

    ADS  CAS  Google Scholar 

  23. Sautter, A., Thalacker, C., Heise, B. & Wurthner, F. Hydrogen bond-directed aggregation of diazadibenzoperylene dyes in low-polarity solvents and the solid state. Proc. Natl Acad. Sci. USA 99, 4993–4996 (2002).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  24. Liu, Z. M. et al. Ionization controls for biomineralization-inspired CO2 chemical looping at constant room temperature. Phys. Chem. Chem. Phys. 17, 10080–10085 (2015).

    CAS  PubMed  Google Scholar 

  25. Franke, D. & Svergun, D. I. DAMMIF, a program for rapid ab-initio shape determination in small-angle scattering. J. Appl. Crystallogr. 42, 342–346 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Demichelis, R., Raiteri, P., Gale, J. D., Quigley, D. & Gebauer, D. Stable prenucleation mineral clusters are liquid-like ionic polymers. Nat. Commun. 2, 590 (2011).

    ADS  PubMed  Google Scholar 

  27. Michel, F. M. et al. Structural characteristics of synthetic amorphous calcium carbonate. Chem. Mater. 20, 4720–4728 (2008).

    CAS  Google Scholar 

  28. Weaver, J. C. et al. Analysis of an ultra hard magnetic biomineral in chiton radular teeth. Mater. Today 13, 42–52 (2010).

    CAS  Google Scholar 

  29. Broz, M. E., Cook, R. F. & Whitney, D. L. Microhardness, toughness, and modulus of Mohs scale minerals. Am. Mineral. 91, 135–142 (2006).

    ADS  CAS  Google Scholar 

  30. Hirano, S., Yogo, T. & Kikuta, K. Synthetic calcite single-crystals for optical-device. Prog. Cryst. Growth Charact. Mater. 23, 341–367 (1992).

    CAS  Google Scholar 

  31. Nielsen, S. S. et al. BioXTAS RAW, a software program for high-throughput automated small-angle X-ray scattering data reduction and preliminary analysis. J. Appl. Crystallogr. 42, 959–964 (2009).

    CAS  Google Scholar 

  32. Franke, D. et al. ATSAS 2.8: a comprehensive data analysis suite for small-angle scattering from macromolecular solutions. J. Appl. Crystallogr. 50, 1212–1225 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Juhás, P., Davis, T., Farrow, C. L. & Billinge, S. J. L. PDFgetX3: a rapid and highly automatable program for processing powder diffraction data into total scattering pair distribution functions. J. Appl. Crystallogr. 46, 560–566 (2013).

    Google Scholar 

  34. Van Der Spoel, D. et al. GROMACS: fast, flexible, and free. J. Comput. Chem. 26, 1701–1718 (2005).

    Google Scholar 

  35. Van der Spoel, D. et al. Gromacs User Manual version 4.5 (2010).

  36. Humphrey, W., Dalke, A. & Schulten, K. VMD: visual molecular dynamics. J. Mol. Graph. 14, 33–38 (1996).

    CAS  PubMed  Google Scholar 

  37. Berendsen, H. J., Postma, J. V., van Gunsteren, W. F., DiNola, A. & Haak, J. Molecular dynamics with coupling to an external bath. J. Chem. Phys. 81, 3684–3690 (1984).

    ADS  CAS  Google Scholar 

  38. Darden, T., York, D. & Pedersen, L. Particle mesh Ewald: an N·log(N) method for Ewald sums in large systems. J. Chem. Phys. 98, 10089–10092 (1993).

    ADS  CAS  Google Scholar 

  39. Shen, J. W., Li, C., van der Vegt, N. F. & Peter, C. Understanding the control of mineralization by polyelectrolyte additives: simulation of preferential binding to calcite surfaces. J. Phys. Chem. C 117, 6904–6913 (2013).

    CAS  Google Scholar 

  40. Raiteri, P., Gale, J. D., Quigley, D. & Rodger, P. M. Derivation of an accurate force-field for simulating the growth of calcium carbonate from aqueous solution: a new model for the calcite−water interface. J. Phys. Chem. C 114, 5997–6010 (2010).

    CAS  Google Scholar 

  41. Jorgensen, W. L. & Tiradorives, J. The Opls potential functions for proteins. Energy minimizations for crystals of cyclic peptides and crambin. J. Am. Chem. Soc. 110, 1657–1666 (1988).

    CAS  PubMed  Google Scholar 

  42. Nielsen, M. H., Aloni, S. & De Yoreo, J. J. In situ TEM imaging of CaCO3 nucleation reveals coexistence of direct and indirect pathways. Science 345, 1158–1162 (2014).

    ADS  CAS  PubMed  Google Scholar 

  43. Smeets, P. J., Cho, K. R., Kempen, R. G., Sommerdijk, N. A. & De Yoreo, J. J. Calcium carbonate nucleation driven by ion binding in a biomimetic matrix revealed by in situ electron microscopy. Nat. Mater. 14, 394–399 (2015).

    ADS  CAS  PubMed  Google Scholar 

  44. Brinker, C. J. & Scherer, G. W. Sol-gel-glass. 1. Gelation and gel structure. J. Non-Cryst. Solids 70, 301–322 (1985).

    ADS  CAS  Google Scholar 

  45. Oshida, K. et al. Analysis of pore structure of activated carbon-fibers using high-resolution transmission electron-microscopy and image-processing. J. Mater. Res. 10, 2507–2517 (1995).

    ADS  CAS  Google Scholar 

  46. Ihli, J., Kulak, A. N. & Meldrum, F. C. Freeze-drying yields stable and pure amorphous calcium carbonate (ACC). Chem. Commun. 49, 3134–3136 (2013).

    CAS  Google Scholar 

  47. Konrad, F., Gallien, F., Gerard, D. E. & Dietzel, M. Transformation of amorphous calcium carbonate in air. Cryst. Growth Des. 16, 6310–6317 (2016).

    CAS  Google Scholar 

  48. Ihli, J. et al. Dehydration and crystallization of amorphous calcium carbonate in solution and in air. Nat. Commun. 5, 3169 (2014).

    ADS  PubMed  Google Scholar 

  49. Xu, X., Han, J. T., Kim, D. H. & Cho, K. Two modes of transformation of amorphous calcium carbonate films in air. J. Phys. Chem. B 110, 2764–2770 (2006).

    CAS  PubMed  Google Scholar 

  50. Tao, J. H., Zhou, D. M., Zhang, Z. S., Xu, X. R. & Tang, R. K. Magnesium-aspartate-based crystallization switch inspired from shell molt of crustacean. Proc. Natl Acad. Sci. USA 106, 22096–22101 (2009).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  51. Kim, Y. Y. et al. An artificial biomineral formed by incorporation of copolymer micelles in calcite crystals. Nat. Mater. 10, 890–896 (2011).

    ADS  CAS  PubMed  Google Scholar 

  52. Sánchez-Iglesias, A. et al. High-yield seeded growth of monodisperse pentatwinned gold nanoparticles through thermally induced seed twinning. J. Am. Chem. Soc. 139, 107–110 (2017).

    PubMed  Google Scholar 

  53. Politi, Y., Arad, T., Klein, E., Weiner, S. & Addadi, L. Sea urchin spine calcite forms via a transient amorphous calcium carbonate phase. Science 306, 1161–1164 (2004).

    ADS  CAS  PubMed  Google Scholar 

  54. Li, L. et al. Bio-inspired enamel repair via Glu-directed assembly of apatite nanoparticles: an approach to biomaterials with optimal characteristics. Adv. Mater. 23, 4695–4701 (2011).

    CAS  PubMed  Google Scholar 

  55. Koeppe, B. et al. Polar solvent fluctuations drive proton transfer in hydrogen bonded complexes of carboxylic acid with pyridines: NMR, IR and ab initio MD study. Phys. Chem. Chem. Phys. 19, 1010–1028 (2017).

    CAS  PubMed  Google Scholar 

  56. Shawoosh, A. S. & Kutub, A. A. An investigation of the electrical, optical and DSC properties of a copper phosphate-glass composition. J. Mater. Sci. 28, 5060–5064 (1993).

    ADS  CAS  Google Scholar 

Download references

Acknowledgements

We thank J. J. De Yoreo, H. Cölfen, P. Fratzl, N. A. J. M. Sommerdijk, D. Joester and L. B. Gower for discussions; Y. Li and Y. Qiu for help with SAXS data analysis; C. Yang and B. Wang for assistance with synchrotron SAXS at the Shanghai Synchrotron Radiation Facility; C. Jin, F. Chen, W. Wang and Y. Wang for assistance with electron microscopy; J. Liu for assistance with XRD; and M. Yu and Y. Liu for assistance with NMR. This work was supported by the National Natural Science Foundation of China (21625105 and 21805241) and the China Postdoctoral Science Foundation (2017M621909 and 2018T110585). We thank W. Liu for his help and inspiration.

Author information

Authors and Affiliations

Authors

Contributions

Z.L. and R.T. initiated the project. Z.L. performed the syntheses and the FTIR, MS, NMR and XRD experiments, the calcite repair and conductivity experiments; C.S. carried out the nanoindentation and SEM experiments; B.J. performed the TEM experiments; Z.Z. performed the computer simulations; Y.Z. acquired the synchrotron SAXS and SEM data; R.T. and Z.L. supervised and supported the project; and Z.L. and X.X. analysed the data. The manuscript was written by Z.L. and R.T. All authors reviewed and approved the manuscript.

Corresponding author

Correspondence to Ruikang Tang.

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.

Peer review information Nature thanks Kazuo Onuma and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Extended data figures and tables

Extended Data Fig. 1 ESI-MS analysis of (CaCO3)n oligomers.

a, Negative-ion mode analysis of CaCl2 and TEA in ethanol solution. b, Negative-ion mode analysis of (CaCO3)n oligomers after 5 days, indicating that TEA has a long-term stabilization effect. n represents the number of Ca2+:CO32− units in one (CaCO3)n oligomer. c, d, Positive ion mode analysis of CaCl2 and TEA mixed ethanol solution (c) and (CaCO3)n oligomers (d). e, ESI tandem MS analysis of the oligomers with m/z 368, showing only one fragmentation peak at around m/z 260 with negative charges. Because no TEA could be detected in the MS–MS peak (the intensity below m/z 200 is zero), this indicates that no TEA was present in the species with m/z 368.

Extended Data Fig. 2 Synchrotron SAXS, liquid-cell TEM, dynamic light scattering and viscosity analyses of (CaCO3)n oligomers and their aggregates.

a, Scattering plots of different concentrations of (CaCO3)n measured by synchrotron SAXS. The red curve is the fitting result from DAMMIF. At a concentration of 57.0 mg l−1 an increase in the scattered intensity was detected, and this broad shoulder peak indicated the aggregation of (CaCO3)n with a large distribution of sizes. A pronounced maxima of the scattered intensity could be observed at a concentration of 285 mg l−1, demonstrating inter-oligomer correlations induced by the formation of larger (CaCO3)n aggregates. b, Size distribution function of oligomer aggregates at 57.0 mg l−1. The inset shows the simulated shape of aggregates, indicating the existence of chain-like structures, including branches at high concentrations. c, Liquid-cell TEM of the aggregates of the (CaCO3)n oligomers showing a chain-like contrast; the boundary between the background (blue) and sample (yellow) is enhanced in the false colour image on the right, and the chain-like aggregates are marked with arrows. d, Viscosity of (CaCO3)n oligomers measured at 5 °C (n ≥ 3). The change in the viscosity with concentration is better fitted with the Huggins equation (R2 = 0.98) than with the Einstein equation (R2 = 0.94), indicating that the shape of the aggregates in solution was more chain-like than spherical. e, The hydrodynamic diameter of (CaCO3)n oligomers or their aggregates at different oligomer concentrations measured by dynamic light scattering with a Ca:TEA ratio of 1:20. The increase (decrease) of the hydrodynamic diameter is reversible by condensing (diluting) the (CaCO3)n oligomer solution. f, Reversible aggregations and disaggregations of the (CaCO3)n oligomers revealed by synchrotron SAXS. g, Scheme of the reversible aggregations and disaggregations of the oligomer unit are controllable by the concentration changes. h, In situ liquid-cell TEM observation of (CaCO3)n oligomers. The chain aggregates remain in dynamical change (aggregation–disaggregation) states.

Extended Data Fig. 3 FTIR spectra of gel-like (CaCO3)n oligomers.

a, FTIR spectra showing the peak corresponding to the carbonate group in (CaCO3)n. b, The C–N bond in TEA. c, The N···H–O bond between TEA and protonated carbonate55. d, The spectra between 400 cm−1 and 4,000 cm−1 are almost the same for ethanol and gel-like oligomers, which confirms that ethanol is the major component in the gel-like (CaCO3)n oligomers. e, The peak at 3,340 cm−1 was contributed by the –OH group of ethanol, which obstructed the signal from H in N···H–O. f, The spectrum of ethyl acetate and ethyl acetate-based gel-like (CaCO3)n oligomers. The ethanol had been completely removed from the ethyl acetate-based gel-like (CaCO3)n oligomers, because the specific peaks of ethanol were not detected; by contrast, the signal at 3,000–4,000 cm−1 was detected. g, The peak at 3,555 cm−1 was contributed by H in the N···H–O bond.

Extended Data Fig. 4 Molecular dynamics simulation of the TEA-stabilized (CaCO3)n.

a, Simulated size distribution of (CaCO3)n clusters in the presence and absence of TEA. b, The branch structure in a (CaCO3)5 cluster without TEA. The circle shows the branching site. c, The branch structure in the large cluster aggregates without TEA. d, There is no branch structure in a cluster of (CaCO3)7 with TEA stabilization. e, A branch structure in (CaCO3)15 even with TEA stabilization. These results demonstrate that the stabilization effect of TEA promotes linear growth of (CaCO3)n.

Extended Data Fig. 5 (CaCO3)n oligomers in ethanol, DMSO and water.

a, d, g, Conductivity of dispersed oligomer solution with ethanol, DMSO or water as solvent. After high-speed centrifugation and redispersion of (CaCO3)n oligomers in ethanol, DMSO or water, the conductivity of the oligomer solution remained steady at 1.2 μS cm−1 in ethanol (a); the conductivity increased to 4.6 μS cm−1 in DMSO and then decreased to 3.8 μS cm−1 within 1 h in DMSO (d); the conductivity increased to 161.2 μS cm−1 but rapidly decreased to 72.0 μS cm−1 within 6 min in water (g). b, e, h, 1H NMR measurement of the ethyl group of TEA (pure or bound with oligomer) in ethanol (b), DMSO (e) and water (h). The results revealed that the chemical shift of 1H of the ethyl group of pure TEA was 2.56 ppm in ethanol and 2.67 ppm in the oligomer solution (b), and this change in chemical shift can be attributed to the interaction between the TEA and the carbonates in the oligomers. After two days, the unchanged chemical shift of 2.68 ppm corroborated the stable capping effects of the TEA in ethanol. In DMSO (e), two signals—2.57 and 2.42 ppm—were attributed to bonded and de-bonded TEA, respectively. In water (h), only the peak for de-bonded TEA in the form of [H-TEA]+ could be detected at 2.83 ppm. c, f, i, Photographs of the (CaCO3)n materials, which exhibit gel-like features in ethanol and powder-like features in DMSO and water. jl, 1H NMR measurement of the ethyl group and hydroxyl group of ethanol (pure or bound with oligomer) in ethanol (j), DMSO (k) and water (l). The negligible change in chemical shift proved that no distinct interactions occurred between ethanol and the (CaCO3)n oligomers.

Extended Data Fig. 6 Cross-linking of (CaCO3)n oligomers.

a, In situ XRD analysis during the drying of (CaCO3)n oligomers, showing the evolution from gels to an ACC monolith. b, Evolution of the RDF from gels to an ACC monolith. The peak at around 2.4 Å is attributed to Ca–O, and its intensity increased during crosslinking. c, Additional high-resolution TEM images of the crosslinking of (CaCO3)n upon the removal of TEA, showing the stepwise chain growth and branch formation. d, Statistical results of the lengths and fractal dimensions of grown (CaCO3)n oligomers. The scheme shows the anisotropic chain growth with branches after TEA removal. e, Cross-linked (CaCO3)n with a network structure over an ultra-thin carbon support film. f, EDS analysis of the crosslinked (CaCO3)n in e. g, EDS mapping analysis of the crosslinked CaCO3 oligomers.

Extended Data Fig. 7 Monolithic ACC and other inorganic materials.

a, FTIR analysis of the monolithic CaCO3, showing ACC as the only component. b, Thermal gravimetric analysis of the monolithic CaCO3, showing a CaCO3:H2O molar ratio in the ACC of 3:1. cf, Photographs and XRD patterns of multiple amorphous monoliths, including calcium phosphate (c), calcium sulfate (d), cupric phosphate (e) and manganous phosphate (f) monoliths. The subsequent crystallization manners of these samples are different. The crystallizations of amorphous cupric phosphate and amorphous manganous phosphate cannot be induced by our thermal treatments (700 °C)56. Amorphous calcium phosphate and calcium sulfate can transform to hydroxyapatite (c) and gypsum (d), respectively, by using humidity or water treatments. Although large single crystals without cracks have not been obtained, we believe that future improvements in crystallization control may provide a suitable solution. In conclusion, crosslinking oligomers provides a general strategy for the construction of amorphous monoliths with continuous structures, but an appropriate crystallization treatment is also required to extend the application of this method to the production of crystalline monoliths.

Extended Data Fig. 8 Crystallization of ACC under different conditions.

a, FTIR spectra of the monolithic CaCO3 at different time periods under 100% relative humidity at 25 °C. The peaks at 866 cm−1 belong to the amorphous phase, and those at 875 cm−1 belong to the crystalline phase. The change of the peaks with time indicates a humidity-induced crystallization process. b, Kinetics of phase transformation from ACC to calcite under different conditions: <5% relative humidity treatment at 25 °C, 100% relative humidity treatment at 25 °C, 300 °C thermal treatment, and 320 °C thermal treatment. The results show the treatment condition controls for the crystallization. It should be noted that although single crystals can be generated by humidity treatment, the size limitation is stricter than that for single crystals generated by thermal treatment. c, X-ray diffraction pattern of thermal induced calcite sample measured along <100>, <010>, <001> and <111> zone axes. The indexing rate of our sample was 79%, while that of geological single-crystal calcite was 87%.

Extended Data Fig. 9 Control of thermal-induced crystallization.

a, Scheme of thermal-induced single-crystalline calcite formation. The infrared thermal images show the thermal gradients from the proximal heating site to the distal site during the treatment, which is the key to ensure the oriented crystallization for single-crystal formation. b, POM image of a single crystalline calcite. ci, Thermal treatment at different conditions demonstrating the importance of selecting an appropriate protocol . H, T and C represent the heating rate, transformation temperature, and cooling rate, respectively. c, d, The continuous movement of the complete amorphous-crystalline crystallization frontier induced by the well-organized thermal gradient with the recommend protocol of H = 1 °C min−1, T = 320 °C and C = 5 °C min−1. e, XRD pattern of ACC with a protocol of T = 300 °C, showing no crystallization at the low heating temperature. f, A protocol of H = 1 °C min−1, T = 325 °C and C = 5 °C min−1 results in a break of the thermal gradient; a spontaneous calcite crystallization is induced at a distal site rather than the amorphous–crystalline crystallization frontier, resulting in polycrystalline formation. g, A protocol of H = 2 °C min−1, T = 320 °C and C = 5 °C min−1 also results in a thermal gradient break so that the random crystallizations occur at the amorphous–crystalline interface. h, A protocol of H = 1 °C min−1, T = 320 °C and C = 10 °C min−1 causes the formation of cracks in the resulting calcite owing to the fast cooling rate. i, Using the ACC bulk produced by the conventional particle packing, polycrystalline formation is induced after a thermal treatment of H = 1 °C min−1, T = 320 °C and C = 5 °C min−1, which is attributed to its discontinuous internal structure.

Extended Data Fig. 10 Repair of sea-urchin spine and enamel by a combination of the crosslinking oligomers.

a, POM image of natural sea-urchin spine. b, c, SEM images of natural sea-urchin spine, showing the complicated structure with a smooth surface. Inset in c is a cross-sectional view of CaCO3 on the spine. d, ATR-FTIR characterization of the sea-urchin spine, which is inorganic calcite. e, POM image of the sea-urchin spine grown using the (CaCO3)n oligomers; the crystallization is induced by a simulated sea water at 25 °C. f, g, SEM images of the repaired sea-urchin spine, which maintains the original complicated structure. A magnified image (g) shows the surface of freshly grown CaCO3. The inset of g is a cross-sectional view of the CaCO3 grown on the spine, demonstrating the continuous interfacial structure between the native sea-urchin spine and the grown calcite. h, ATR-FTIR characterization of the repaired layer on the sea-urchin spine, which is also calcite. i, SEM image of the repaired sea-urchin spine using a conventional solution crystallization method. The random precipitation of numerous calcite particles rather than the expected oriented growth is resulted, demonstrating a failure in the biomineral repair. j, Regrowth of enamel structure by using calcium phosphate oligomers and crystallization is induced by simulated oral fluid at 37 °C. Epitaxial growth of the enamel rods is observed, which is due to the continuous construction by the crosslinking oligomers.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Liu, Z., Shao, C., Jin, B. et al. Crosslinking ionic oligomers as conformable precursors to calcium carbonate. Nature 574, 394–398 (2019). https://doi.org/10.1038/s41586-019-1645-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-019-1645-x

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