Skip to main content
Log in

Anisotropy and Asymmetry of the Elastic Tensor of Lattice Materials

  • Research
  • Published:
Journal of Elasticity Aims and scope Submit manuscript

Abstract

Lattice materials formed by hinged springs or linear elastic bonds may exhibit diverse anisotropy and asymmetry features of the overall elastic behavior depending on their unit cell configuration. The recently developed singum model transfers the force-displacement relationship of the springs in the lattice to the stress-strain relationship in the continuum particle, and provides the analytical form of tangential elasticity. When a pre-stress exists in the lattice, the stiffness tensor significantly changes due to the effect of the configurational stress; existing methods like the lattice spring method, relying on a scalar energy equivalence, are insufficient in such situations. Instead, a tensorial homogenization method with the new definition of singum stress and strain, should be preferred. Different lattice structures lead to different symmetries of the stiffness tensors, which are demonstrated by five lattices. When all bonds exhibit the same length, regular hexagonal, honeycomb, and auxetic lattices demonstrate that the stiffness changes from an isotropic to anisotropic, from symmetric to asymmetric tensor. When the central symmetry of the unit cell is not satisfied, the primitive cell will contain more than one singums and the Cauchy–Born rule fails by the loss of equilibrium of the single singum. A secondary stress is induced to balance the singums. Displacement gradient \(d_{ij}=u_{j,i}\) is proposed to replace strain in the constitutive law for the general case because \(d_{12}\) and \(d_{21}\) can produce different stress states. Although the hexagonal and honeycomb lattices may exhibit isotropic behavior, for general auxetic lattices, an anisotropic and asymmetric elastic tensor is obtained with the loss of both minor and major symmetry, which is also demonstrated in a square lattice with unbalanced central symmetry and a chiral lattice. The modeling procedure and results can be generalized to three dimensions and other lattices with the anisotropic and asymmetric stiffness.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

References

  1. Berinskii, I.E.: In-plane elastic properties of auxetic multilattices. Smart Mater. Struct. 27(7), 075012 (2018)

    Article  Google Scholar 

  2. Bordiga, G., Piccolroaz, A., Bigoni, D.: A way to hypo-elastic artificial materials without a strain potential and displaying flutter instability. J. Mech. Phys. Solids 158, 104665 (2022)

    Article  MathSciNet  Google Scholar 

  3. Cabras, L., Brun, M.: A class of auxetic three-dimensional lattices. J. Mech. Phys. Solids 91, 56–72 (2016)

    Article  MathSciNet  Google Scholar 

  4. Chen, Y., Scarpa, F., Liu, Y., Leng, J.: Elasticity of anti-tetrachiral anisotropic lattices. Int. J. Solids Struct. 50(6), 996–1004 (2013)

    Article  Google Scholar 

  5. Cosserat, E., Cosserat, F.: Théorie des corps déformables. Librairie Scientifique A. Hermann et Fils (1909)

  6. Coulais, C., Sounas, D., Alu, A.: Static non-reciprocity in mechanical metamaterials. Nature 542(7642), 461–464 (2017)

    Article  Google Scholar 

  7. Delfani, M., Shodja, H.: A large-deformation thin plate theory with application to one-atom-thick layers. J. Mech. Phys. Solids 87, 65–85 (2016)

    Article  MathSciNet  Google Scholar 

  8. Delfani, M., Shodja, H., Ojaghnezhad, F.: Mechanics and morphology of single-walled carbon nanotubes: from graphene to the elastica. Philos. Mag. 93(17), 2057–2088 (2013)

    Article  Google Scholar 

  9. Ericksen, J.: On the symmetry of deformable crystals. Arch. Ration. Mech. Anal. 72, 1–13 (1979)

    Article  MathSciNet  Google Scholar 

  10. Ericksen, J.: On the Cauchy–Born rule. Math. Mech. Solids 13(3–4), 199–220 (2008)

    Article  MathSciNet  Google Scholar 

  11. Eringen, A.C.: Mechanics of micromorphic continua. In: Mechanics of Generalized Continua, pp. 18–35. Springer, Berlin (1968)

    Chapter  Google Scholar 

  12. Eringen, A.C.: Microcontinuum Field Theories: I. Foundations and Solids. Springer, Berlin (2012)

    Google Scholar 

  13. Fadda, G., Zanzotto, G.: The arithmetic symmetry of monoatomic 2-nets. Acta Crystallogr., Sect. A, Found. Crystallogr. 56(1), 36–48 (2000)

    Article  MathSciNet  Google Scholar 

  14. François, M.L., Chen, L., Coret, M.: Elasticity and symmetry of triangular lattice materials. Int. J. Solids Struct. 129, 18–27 (2017)

    Article  Google Scholar 

  15. Kole, S., Alexander, G.P., Ramaswamy, S., Maitra, A.: Active cholesterics: odder than odd elasticity (2020). ArXiv preprint arXiv:2012.14321

  16. Lakes, R.: Experimental microelasticity of two porous solids. Int. J. Solids Struct. 22(1), 55–63 (1986)

    Article  Google Scholar 

  17. Lakes, R.: Foam structures with a negative Poisson’s ratio. Science 235, 1038–1041 (1987)

    Article  Google Scholar 

  18. Lee, C., Wei, X., Kysar, J.W., Hone, J.: Measurement of the elastic properties and intrinsic strength of monolayer graphene. Science 321(5887), 385–388 (2008)

    Article  Google Scholar 

  19. Liu, K., Paulino, G.: Nonlinear mechanics of non-rigid origami: an efficient computational approach. Proc. R. Soc. A, Math. Phys. Eng. Sci. 473(2206), 20170348 (2017)

    MathSciNet  Google Scholar 

  20. Love, A.E.H.: A Treatise on the Mathematical Theory of Elasticity. Cambridge University Press, Cambridge (1906)

    Google Scholar 

  21. Mindlin, R.: Influence of couple-stresses on stress concentrations. Tech. Rep., New York (1962)

  22. Mindlin, R.D.: Microstructure in linear elasticity. Tech. Rep., Columbia Univ. New York, Dept. of Civil Engineering and Engineering Mechanics (1963)

  23. Mindlin, R.: Stress functions for a Cosserat continuum. Int. J. Solids Struct. 1(3), 265–271 (1965)

    Article  Google Scholar 

  24. Moosavian, H., Shodja, H.: Mindlin–Eringen anisotropic micromorphic elasticity and lattice dynamics representation. Philos. Mag. 100(2), 157–193 (2020)

    Article  Google Scholar 

  25. Mura, T.: Micromechanics of Defects in Solids. Springer, Netherlands (1987). https://doi.org/10.1007/978-94-009-3489-4

    Book  Google Scholar 

  26. Nowacki, W.: Theory of Micropolar Elasticity. Springer, Berlin (1972)

    Google Scholar 

  27. Ostoja-Starzewski, M.: Lattice models in micromechanics. Appl. Mech. Rev. 55(1), 35–60 (2002)

    Article  MathSciNet  Google Scholar 

  28. Pitteri, M.: On \(\nu \)+ 1-lattices. J. Elast. 15, 3–25 (1985)

    Article  Google Scholar 

  29. Saxena, K.K., Das, R., Calius, E.P.: Three decades of auxetics research- materials with negative Poisson’s ratio: a review. Adv. Eng. Mater. 18(11), 1847–1870 (2016)

    Article  Google Scholar 

  30. Scheibner, C., Souslov, A., Banerjee, D., Surówka, P., Irvine, W., Vitelli, V.: Odd elasticity. Nat. Phys. 16(4), 475–480 (2020)

    Article  Google Scholar 

  31. Sfyris, D., Sfyris, G., Galiotis, C.: Curvature dependent surface energy for a free standing monolayer graphene: some closed form solutions of the non-linear theory. Int. J. Non-Linear Mech. 67, 186–197 (2014)

    Article  Google Scholar 

  32. Shodja, H.M., Ojaghnezhad, F., Etehadieh, A., Tabatabaei, M.: Elastic moduli tensors, ideal strength, and morphology of stanene based on an enhanced continuum model and first principles. Mech. Mater. 110, 1–15 (2017)

    Article  Google Scholar 

  33. Spadoni, A., Ruzzene, M.: Elasto-static micropolar behavior of a chiral auxetic lattice. J. Mech. Phys. Solids 60(1), 156–171 (2012)

    Article  Google Scholar 

  34. Tadmor, E.B., Miller, R.E.: Modeling Materials: Continuum, Atomistic and Multiscale Techniques. Cambridge University Press, Cambridge (2011)

    Book  Google Scholar 

  35. Timoshenko, S., Goodier, J.N. (eds.): Theory of Elasticity. McGraw-Hill, New York (1951)

    Google Scholar 

  36. Wallace, D.C.: Thermodynamics of Crystals. Wiley, New York (1972)

    Book  Google Scholar 

  37. Wang, M., Xu, B., Gao, C.: Recent general solutions in linear elasticity and their applications. Appl. Mech. Rev. 61(3), 030803 (2008)

    Article  Google Scholar 

  38. Wigner, E., Seitz, F.: On the constitution of metallic sodium. Phys. Rev. 43(10), 804 (1933)

    Article  Google Scholar 

  39. Willis, J.R.: Mechanics of composites. Ecole polytechnique, Département de mécanique (2002)

  40. Yin, H.: A simplified continuum particle model bridging interatomic potentials and elasticity of solids. J. Eng. Mech. 148(5), 04022017 (2022)

    Article  Google Scholar 

  41. Yin, H.: Generalization of the singum model for the elasticity prediction of lattice metamaterials and composites. J. Eng. Mech. 149(5), 04023023 (2023)

    Article  Google Scholar 

  42. Yin, H.: Improved singum model based on finite deformation of crystals with the thermodynamic equation of state. J. Eng. Mech. 149(4), 04023018 (2023)

    Article  Google Scholar 

  43. Yin, H., Sun, L., Chen, J.: Magneto-elastic modeling of composites containing chain-structured magnetostrictive particles. J. Mech. Phys. Solids 54(5), 975–1003 (2006)

    Article  Google Scholar 

  44. Yin, H., Pao, F., Zadshir, M., Lou, J., Liu, C.: Tailoring thermoelastic constants of cellular and lattice materials with pre-stress for lightweight structure. U.S. Patent App. 17/935, 155 (2022)

  45. Yin, H., Cui, J., Zadshir, M., Teka, L.: Effect of wrapping force on the effective elastic behavior of packed cylinders. J. Appl. Mech. 90(3), 031003 (2023)

    Article  Google Scholar 

  46. Zhang, W., Neville, R., Zhang, D., Scarpa, F., Wang, L., Lakes, R.: The two-dimensional elasticity of a chiral hinge lattice metamaterial. Int. J. Solids Struct. 141, 254–263 (2018)

    Google Scholar 

Download references

Acknowledgements

The author Yin thanks Professors Kefu Huang and Minzhong Wang for the fruitful discussion. The authors are also very grateful to Professor Glaucio Paulino for his encouragement and inspiration of this work and its extension to 3D origami-based lattices is underway.

Funding

This work is sponsored by the National Science Foundation IIP #1738802, IIP #1941244, CMMI #1762891, and U.S. Department of Agriculture NIFA #2021-67021-34201, whose support is gratefully acknowledged.

Author information

Authors and Affiliations

Authors

Contributions

Yin wrote the main manuscript text, and Liu implement the numerical simulation to verify the formulation.

Corresponding author

Correspondence to Huiming Yin.

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.

Appendices

Appendix A: Determination of the Force Transfer Matrix \(\textbf{R}\)

Consider a truss system of N bars, which connect to one singum node at one end with another end fixed by hinges as the cutting point. All bars are assumed to exhibit the same length and elastic constants for simplicity, and the number of the bars is large enough to make the truss system stable or indeterminate. When all the cutting points are fixed, given a force \(P_{i}\) on the singum node, the force transferred to the \(I\)th bar can be written as

$$ T_{j}^{I} = P_{i} R^{I}_{ij}, $$
(66)

which is along the bar with the direction from the singum node to the cutting point \(\textbf{n}^{I}\). Here, the conventional summation of double index notation is applied to the subscript with lower case symbols only.

Consider the force in each bar is given at \(\textbf{f}^{I} = \frac{V_{,\lambda}n_{i}}{2l_{p}^{0}}\). Without the loss of any generality, set the origin \(\textbf{0}\) at the singum node and the cutting points are fixed at \(\textbf{x}^{I}\) (\(I=1,2,\ldots , N\)) with \(l_{p}=|\textbf{x}^{I}| =\lambda l_{p}^{0}\), so \(\textbf{n}^{I}=\textbf{x}^{I}/l_{p}\).

When all cutting points are fixed, a small variation of the singum node, \(d\textbf{x}\), will change the bond vectors into \(\textbf{x}^{I}-d\textbf{x}\), which leads to a length change

$$ dl^{I} = -n_{i}^{I} dx_{i} $$
(67)

and an orientational change

$$ dn_{i}^{I} = \frac{n^{I}_{i}n^{I}_{j}-\delta _{ij}}{\lambda l_{p}^{0}} dx_{j} $$
(68)

for the \(I\)th bonds. Note that \(d\lambda ^{I} = dl^{I}/l_{p}^{0}\).

The force variation for each bar can be obtained as

$$ df_{i}^{I} = \frac{V_{,\lambda \lambda}n_{i}^{I} d\lambda ^{I} + V_{,\lambda}dn_{i}^{I}}{2l_{p}^{0}} = \left [ \frac{-V_{,\lambda \lambda}n_{i}^{I} n_{j}^{I}}{2{l_{p}^{0}}^{2} } + \frac{V_{,\lambda}(n_{i}^{I} n_{j}^{I}-\delta _{ij})}{2\lambda{l_{p}^{0}}^{2} } \right ]dx_{j}, $$
(69)

which includes two parts: the first is caused by the length change and the second orientational change. The resultant force variation on the singum is

$$ df_{i} = \sum _{I=1}^{N} df_{i}^{I}. $$
(70)

Then the force transfer matrix \(R_{i}^{I}\) can be determined by the classical displacement method with the following procedure in 3D case, which can be reduced to 2D case straightforwardly:

  1. 1.

    Given a unit displacement \(\textbf{d}_{1}=(1,0,0)\) on the singum node, the resultant force on the singum is calculated by Eq. (70), which is a sum of N vectors in 3D, namely \(\textbf{P}_{1} = \sum _{I=1}^{N} \textbf{f}^{I1}\).

  2. 2.

    Similarly, given a unit displacement \(\textbf{d}_{2}=(0,1,0)\) or \(\textbf{d}_{3}=(0,0,1)\) on the singum node, the resultant force on the singum can also be calculated as \(\textbf{P}_{2} = \sum _{I=1}^{N} \textbf{f}^{I2}\) and \(\textbf{P}_{3}= \sum _{I=1}^{N} \textbf{f}^{I3}\).

  3. 3.

    Given any displacement \(\textbf{d}=(a,b,c)\), the resultant force will be written as \(\textbf{P}= a\textbf{P}_{1} +b\textbf{P}_{2} + c\textbf{P}_{3}\).

  4. 4.

    Solve \(a\textbf{P}_{1} +b\textbf{P}_{2} + c\textbf{P}_{3} = (-1,0,0)\) with three equations for \(a, b, c\). Using \(d \textbf{x} = (a,b,c)\), one can calculate force variation of each bar by Eq. (69), which defines \(R^{I}_{1j} = df_{j}^{I}\).

  5. 5.

    Similarly, solve \(a\textbf{P}_{1} +b\textbf{P}_{2} + c\textbf{P}_{3} = (0, -1,0)\) or \((0,0,-1)\) with three equations for \(a, b, c\). One can obtain \(R^{I}_{2j}\) or \(R^{I}_{3j}\), respectively.

Therefore, \(R^{I}_{ij}\), which shows the force for member \(I\) due to a unit force in \(x_{i}\), can be obtained.

Appendix B: A Case Study of a Honeycomb Lattice Truss System

To demonstrate the accuracy of the singum model, a case study is presented for a honeycomb lattice with harmonic potential or linear spring bonds between neighboring nodes. A MATLAB program is developed to simulate the elastic behavior of the lattice for verification of the formulation in Sect. 3, which can be straightforwardly extended to other lattices. In the program, an array of nodes are automatically generated based on the given lattice with the number of unit cells in \(X\) and \(Y\) directions. The \(X-Y\) coordinate origin is set up on the center of the lattice. When the lattice is undergone a deformation, the new coordinate \(x-y\) shares the same origin but the coordinates of the nodes change. A list of neighbors for each node is detected with the corresponding bonds and saved for the force computation step. The mid-point of each bond is also important as the potential cutting point of singum surface.

Figure 6 schematically illustrates the honeycomb lattices. The boundary points can be selected by two cases. Figure 6(a) uses the mid-points of the bonds for the boundary; whereas Fig. 6(b) exhibits the boundary on the nodes. If the lattice approaches the infinite domain, the boundary selection produces negligible effects to the effective mechanical behavior. However, when finite unit cells are used, they may produce big differences with different convergence rates to the solution. Due to the periodicity of the singum, Fig. 6(a) provides a better performance and will be used in the study. The algorithm is structured and implemented as follows:

  1. 1.

    Initialize the simulation box by periodically extending the singum along \(x\) and \(y\) direction with \(N_{x}\) and \(N_{y}\) replications. The 4 sides of the box are made of loading bars as a boundary layer. The node on the boundary is connected to the loading bars by hinges. The initial bond length and force are at \(r=2l_{p}^{0}\) and \(\textbf{F}^{I}=0\), respectively, so \(\lambda =1\).

  2. 2.

    Given a testing mode, such as tension or shearing, apply the corresponding uniform singum strain variation \(\delta d_{ij}=10^{-6}\) to the simulation box according to the Cauchy–Born (CB) rule. The new positions of all mid-points are updated through an affine transformation with \(\delta d_{ij}\).

  3. 3.

    Calculate the coordinate of each node from the mid-points by the equilibrium of the node. The length change of each fiber and \(\lambda \), and use the potential function \(V(\lambda )\) to calculate the equilibrium bond forces. For each loading bar, collect all bond forces and calculate the effective stress vector on each bar in the deformed configuration. Using the effective stress vector on the 4 edges, one can obtain the stress variation caused by \(\delta d_{ij}=10^{-6}\) at \(\lambda =1\), and thus calculate the elastic tensor.

  4. 4.

    For any normalized fiber length or stretch ratio, namely \(\lambda ^{i}\), the coordinate of each node and the force in each bond can be calculated with the harmonic potential \(V\) in Eq. (5). Repeat Steps 2 and 3 to calculate the elastic tensor at \(\lambda ^{i}\). Therefore, the relation of elastic tensor and \(\lambda \) can be calculated.

Fig. 6
figure 6

The honeycomb lattice truss system with boundary: (a) at the mid-points of the bonds and (b) at the nodes

Although the above calculation can guarantee the equilibrium of the internal nodes immediately with the periodic microstructure during the deformation, the result of the elastic tensor is affected by the cut-off for the boundary length but will quickly converge with the number of unit cell \(N_{x}\) and \(N_{y}\).

On the contrary, if Fig. 6(b) is used directly to generate the new positions of nodes following the C-B rule, the equilibrium positions of the internal nodes will not be periodic anymore, and a very large lattice is required to reach the convergent solution, which is demonstrated in Fig. 7(a).

Fig. 7
figure 7

Numerical simulation results of elastic tensors changing with different simulation scales and different boundary construction types(a) the comparison of the numerical simulation and the prediction of elastic tensors with different \(\lambda \) (b)

As a numerical example, consider a lattice with \(k=1000N/m\), \(l_{p}^{0}=1mm\), and \(\lambda =1\). Equation (33) provides \(C_{1111}=C_{2211}=288.6751\)N/m and \(C_{1212}=0\)N/m respectively. 7(a) shows the prediction of \(C_{1111}\) and \(C_{1122}\) changing with the increase of \(N_{x}\) or \(N_{y}\) (here we use \(N_{x}=N_{y}\)). Obviously, for type A lattice in Fig. 6(a), the simulation converges quickly with a few nodes. For instance, with only 184 nodes in total, the simulation results gave \(C_{1111}=288.6520\)N/m, compared to the prediction value from the singum model, the difference was only about 0.008%. However, for type B lattice in Fig. 6(b), the simulation converges much slower and requires more nodes to dilute the boundary effects as the displacement of nodes on top or bottom boundary does not follow the periodic distribution.

Using the type A lattice, we change the stretch ratio \(\lambda \) between 1.0 to 1.2, and show the comparison of the numerical simulation and the prediction of Eq. (33) in 7(b), which exhibit excellent agreement. Indeed, the singum model provides the exact solution for the lattice with the potential between short-range bonds.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yin, H., Liu, C. Anisotropy and Asymmetry of the Elastic Tensor of Lattice Materials. J Elast 154, 659–691 (2023). https://doi.org/10.1007/s10659-023-10028-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10659-023-10028-7

Keywords

Mathematics Subject Classification

Navigation