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:

Mechanical metamaterials at the theoretical limit of isotropic elastic stiffness

Abstract

A wide variety of high-performance applications1 require materials for which shape control is maintained under substantial stress, and that have minimal density. Bio-inspired hexagonal and square honeycomb structures and lattice materials based on repeating unit cells composed of webs or trusses2, when made from materials of high elastic stiffness and low density3, represent some of the lightest, stiffest and strongest materials available today4. Recent advances in 3D printing and automated assembly have enabled such complicated material geometries to be fabricated at low (and declining) cost. These mechanical metamaterials have properties that are a function of their mesoscale geometry as well as their constituents3,5,6,7,8,9,10,11,12, leading to combinations of properties that are unobtainable in solid materials; however, a material geometry that achieves the theoretical upper bounds for isotropic elasticity and strain energy storage (the Hashin–Shtrikman upper bounds) has yet to be identified. Here we evaluate the manner in which strain energy distributes under load in a representative selection of material geometries, to identify the morphological features associated with high elastic performance. Using finite-element models, supported by analytical methods, and a heuristic optimization scheme, we identify a material geometry that achieves the Hashin–Shtrikman upper bounds on isotropic elastic stiffness. Previous work has focused on truss networks and anisotropic honeycombs, neither of which can achieve this theoretical limit13. We find that stiff but well distributed networks of plates are required to transfer loads efficiently between neighbouring members. The resulting low-density mechanical metamaterials have many advantageous properties: their mesoscale geometry can facilitate large crushing strains with high energy absorption2,14,15, optical bandgaps16,17,18,19 and mechanically tunable acoustic bandgaps20, high thermal insulation21, buoyancy, and fluid storage and transport. Our relatively simple design can be manufactured using origami-like sheet folding22 and bonding methods.

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

Figure 1: Material geometries.
Figure 2: Young’s, shear and bulk moduli.
Figure 3: Total stiffness.
Figure 4: The Zener anisotropy ratio.
Figure 5: Strain energy distributions.

Similar content being viewed by others

References

  1. United States National Research Council. Application of Lightweighting Technology to Military Aircraft, Vessels, and Vehicles https://doi.org/10.17226/13277 (National Academies Press, 2012)

  2. Wadley, H. N. Multifunctional periodic cellular metals. Philos. Trans. R. Soc. A 364, 31–68 (2006)

    Article  CAS  ADS  Google Scholar 

  3. Gibson, L. J., Ashby, M. F. & Harley, B. A. Cellular Materials in Nature and Medicine (Cambridge Univ. Press, 2010)

  4. Ashby, M. Hybrid materials to expand the boundaries of material–property space. J. Am. Ceram. Soc. 94, s3–s14 (2011)

    Article  CAS  Google Scholar 

  5. George, T., Deshpande, V. S. & Wadley, H. N. G. Hybrid carbon fiber composite lattice truss structures. Compos. A 65, 135–147 (2014)

    Article  CAS  Google Scholar 

  6. Ashby, M. F. et al. Metal foams: a design guide (Butterworth-Heinemann, 2000)

  7. Jang, D., Meza, L. R., Greer, F. & Greer, J. R. Fabrication and deformation of three-dimensional hollow ceramic nanostructures. Nat. Mater. 12, 893–898 (2013)

    Article  CAS  ADS  Google Scholar 

  8. Montemayor, L., Chernow, V. & Greer, J. R. Materials by design: using architecture in material design to reach new property spaces. MRS Bull. 40, 1122–1129 (2015)

    Article  Google Scholar 

  9. Zheng, X. Y. et al. Ultralight, ultrastiff mechanical metamaterials. Science 344, 1373–1377 (2014)

    Article  CAS  ADS  Google Scholar 

  10. Grenestedt, J. L. Effective elastic behavior of some models for ‘perfect’ cellular solids. Int. J. Solids Struct. 36, 1471–1501 (1999)

    Article  Google Scholar 

  11. Valdevit, L., Jacobsen, A. J., Greer, J. R. & Carter, W. B. Protocols for the optimal design of multi-functional cellular structures: from hypersonics to micro-architected materials. J. Am. Ceram. Soc. 94, s15–s34 (2011)

    Article  CAS  Google Scholar 

  12. Spadaccini, C. M. in Frontiers of Engineering: Reports on Leading-Edge Engineering from the 2015 Symposium 85–98, https://doi.org/10.17226/21825 (National Academies Press, 2016)

  13. Hashin, Z. & Shtrikman, S. A variational approach to the theory of the elastic behaviour of multiphase materials. J. Mech. Phys. Solids 11, 127–140 (1963)

    Article  MathSciNet  ADS  Google Scholar 

  14. Moongkhamklang, P. & Wadley, H. N. G. Titanium alloy lattice structures with millimeter scale cell sizes. Adv. Eng. Mater. 12, 1111–1116 (2010)

    Article  CAS  Google Scholar 

  15. Zok, F. W. et al. A protocol for characterizing the structural performance of metallic sandwich panels: application to pyramidal truss cores. Int. J. Solids Struct. 41, 6249–6271 (2004)

    Article  Google Scholar 

  16. do Rosário, J. J. et al. Self-assembled ultra high strength, ultra stiff mechanical metamaterials based on inverse opals. Adv. Eng. Mater. 17, 1420–1424 (2015)

    Article  Google Scholar 

  17. Bertoldi, K. & Boyce, M. C. Mechanically triggered transformations of phononic band gaps in periodic elastomeric structures. Phys. Rev. B 77, 052105 (2008)

    Article  ADS  Google Scholar 

  18. Babaee, S., Wang, P. & Bertoldi, K. Three-dimensional adaptive soft phononic crystals. J. Appl. Phys. 117, 244903 (2015)

    Article  ADS  Google Scholar 

  19. Mousanezhad, D. et al. Honeycomb phononic crystals with self-similar hierarchy. Phys. Rev. B 92, 104304 (2015)

    Article  ADS  Google Scholar 

  20. Bertoldi, K. & Boyce, M. C. Wave propagation and instabilities in monolithic and periodically structured elastomeric materials undergoing large deformations. Phys. Rev. B 78, 184107 (2008)

    Article  ADS  Google Scholar 

  21. Steeves, C. a. & Evans, A. G. Optimization of thermal protection systems utilizing sandwich structures with low coefficient of thermal expansion lattice hot faces. J. Am. Ceram. Soc. 94, s55–s61 (2011)

    Article  CAS  Google Scholar 

  22. Berger, J. 3-D honeycomb foam structure. US patent application 15/110,402 (2015)

  23. Gibson, L. J. & Ashby, M. F. Cellular Solids: Structure and Properties 2nd edn, 175–231 (Cambridge Univ. Press, 1997)

  24. Abaqus CAE, version 6.14-4, http://www.3ds.com/products-services/simulia/products/abaqus/abaquscae (Simulia, 2015)

  25. Deshpande, V. S., Fleck, N. A. & Ashby, M. F. Effective properties of the octet-truss lattice material. J. Mech. Phys. Solids 49, 1747–1769 (2001)

    Article  CAS  ADS  Google Scholar 

  26. Gurtner, G. & Durand, M. Stiffest elastic networks. Proc. R. Soc. Lond. A 470, 20130611 (2014)

    Article  ADS  Google Scholar 

  27. Danielsson, M., Parks, D. M. & Boyce, M. C. Three-dimensional micromechanical modeling of voided polymeric materials. J. Mech. Phys. Solids 50, 351–379 (2002)

    Article  CAS  ADS  Google Scholar 

  28. Gibson, L. J. Modelling the mechanical behavior of cellular materials. Mater. Sci. Eng. 110, 1–36 (1989)

    Article  Google Scholar 

  29. Bauer, J., Schroer, A., Schwaiger, R. & Kraft, O. Approaching theoretical strength in glassy carbon nanolattices. Nat. Mater. 15, 438–443 (2016)

    Article  CAS  ADS  Google Scholar 

  30. Vukusic, P. & Sambles, J. R. Photonic structures in biology. Nature 424, 852–855 (2003)

    Article  CAS  ADS  Google Scholar 

  31. Roberts, P. & Garboczi, E. J. Elastic moduli of model random three-dimensional closed-cell cellular solids. Acta Mater. 49, 189–197 (2001)

    Article  CAS  Google Scholar 

  32. Thomson, W. On the division of space with minimum partitional area. Acta Math. 11, 121–134 (1887)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

H.N.G.W. is grateful for support for this work by the ONR (grant number N00014-15-1-2933), managed by D. Shifler, and the DARPA MCMA programme (grant number W91CRB-10-1-005), managed by J. Goldwasser.

Author information

Authors and Affiliations

Authors

Contributions

J.B.B. created the ideas, conceived and designed the new material geometries, and performed the structural analysis. R.M.M. developed the analytical models for strain energy and moduli, and, with H.N.G.W., contributed to refining the concepts, contextualizing the results, and providing critiques and assessments.

Corresponding author

Correspondence to J. B. Berger.

Ethics declarations

Competing interests

The material geometry identified in this work to achieve the theoretical bounds in performance has been included in a Patent Cooperation Treaty (PCT/US2015/010458) by Nama Development, LLC (DE), which is majority-owned by J.B.B.

Extended data figures and tables

Extended Data Figure 1 Property space of isotropic and nearly isotropic materials.

Metamaterial geometries with suboptimal performance have been omitted. A theoretical bound (Ω = 1) limits the performance of all material systems and is defined by the highest performance possible for a two-phase system, which is achieved by the single-crystal diamond and void system. The parameter Ω is defined in the text. The cubic + octet material can bound property space when composed of materials with maximal properties, such as diamond, beryllium, boron carbide, fibre composites and lightweight alloys. Fabrication techniques limit our ability to achieve a wide and otherwise unoccupied region of property space (labelled ‘Metamaterials’).

Extended Data Figure 2 Strain energy distributions from axial stress.

af, Strain energy distributions in the geometries from Fig. 1, when subject to uniaxial stress; Uε is the local strain energy density and is the average solid-phase strain energy density. Macroscopic loads are transmitted through stiff networks of members aligned with the principal stress direction. Strains are small, but scaled to reveal the nature of the deformations. The two-dimensional connectedness of material in closed-cell geometries enables effective transmission of loads between neighbouring members, facilitating materials that can achieve the theoretical bounds (a). Open-cell and stochastic materials (df) have large strain energy concentrations.

Extended Data Figure 3 Strain energy distributions from shear loading.

af, Strains are scaled to highlight the nature of the deformations. The displacements in stiff closed-cell materials (ac) are largely affine and absent of bending. Despite the identical alignment of material in the octet truss (f) and octet foam (c), the absence of membrane stress enables substantial bending to take place in the former, open-cell configuration. All geometries have .

Extended Data Figure 4 Strain energy distributions from hydrostatic loading.

af, Strains are scaled to highlight the nature of the deformations. In maximally stiff materials, the deformations are limited to the filling of void space through member swelling. The displacements are primarily affine and strain energy distributions are nearly uniform. Poor alignment of neighbouring cell walls in the quasi-random material allows some bending to occur (d). All geometries have .

Extended Data Figure 5 Moduli of quasi-random and stochastic foams.

a, The normalized average Young’s, shear and bulk moduli of quasi-random foam, determined using finite-element models, are plotted against relative density. The coloured data (from this study) are fitted to third-order polynomials that are forced to go through the origin (0, 0) and the point (1, 1), corresponding to empty space and a dense solid, respectively. Data are also fitted to the model of ref. 23 (‘G-A’) for the stiffness of isotropic cellular materials, using ϕ = 2/3 (dashed line). (ϕ is the fraction of material subject to bending, and (1−ϕ) is the fraction of material subject to stretching.) Experimental data for Young’s modulus (open circles) are taken from ref. 23. A similar, but more extensive, finite-element study31 (open squares) produced similar results. b, Close-up of the grey shaded region in a.

Source data

Extended Data Table 1 Material properties used to populate property space

Supplementary information

Supplementary Information

This file contains Supplementary Equations and Methods. It contains derivations for elastic strain energy and stiffness of octet-foam, cubic-foam, and octet-foam and cubic-foam combined, and methods for finite element verification. (PDF 1296 kb)

PowerPoint slides

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Berger, J., Wadley, H. & McMeeking, R. Mechanical metamaterials at the theoretical limit of isotropic elastic stiffness. Nature 543, 533–537 (2017). https://doi.org/10.1038/nature21075

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature21075

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