Skip to main content
Log in

On the Orientation Average Based on Central Orientation Density Functions for Polycrystalline Materials

  • Published:
Journal of Elasticity Aims and scope Submit manuscript

Abstract

The present work continues the investigation first started by Lobos et al. (J. Elast. 128(1):17–60, 2017) concerning the orientation average of tensorial quantities connected to single-crystal physical quantities distributed in polycrystals. In Lobos et al. (J. Elast. 128(1):17–60, 2017), central orientation density functions were considered in the orientation average for fourth-order tensors with certain index symmetries belonging to single-crystal quantities. The present work generalizes the results of Lobos et al. (J. Elast. 128(1):17–60, 2017) for the orientation average of tensors of arbitrary order by presenting a clear connection to the Fourier expansion of central orientation density functions and of the general orientation density function in terms of tensorial texture coefficients. The closed form of the orientation average based on a central orientation density function is represented in terms of the Fourier coefficients (referred to as texture eigenvalues) and the central orientation of the central orientation density function. The given representation requires the computation of specific isotropic tensors. A pragmatic algorithm for the automated generation of a basis of isotropic tensors is given. Applications and examples are presented to show that the representation of the orientation average offers a low-dimensional parametrization with major benefits for optimization problems in materials science. A simple implementation in Python 3 for the reproduction of all examples is offered through the GitHub repository https://github.com/mauricio-fernandez-l/centralODF-average.

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

Similar content being viewed by others

References

  1. Adams, B.L., Lyon, M., Henrie, B.: Microstructures by design: linear problems in elastic–plastic design. Int. J. Plast. 20(8–9), 1577–1602 (2004). https://doi.org/10.1016/j.ijplas.2003.11.008.

    Article  MATH  Google Scholar 

  2. Adams, B.L., Kalidindi, S.R., Fullwood, D.T.: Microstructure Sensitive Design for Performance Optimization. Butterworth-Heinemann, Waltham (2013). https://doi.org/10.1016/B978-0-12-396989-7.00001-0

    Book  Google Scholar 

  3. Andrews, D.L., Ghoul, W.A.: Eighth rank isotropic tensors and rotational averages. J. Phys. A, Math. Gen. 14(6), 1281–1290 (1999). https://doi.org/10.1088/0305-4470/14/6/008

    Article  ADS  MathSciNet  Google Scholar 

  4. Andrews, D.L., Thirunamachandran, T.: On three-dimensional rotational averages. J. Chem. Phys. 67(11), 5026 (1977). https://doi.org/10.1063/1.434725.

    Article  ADS  Google Scholar 

  5. Auffray, N.: On the algebraic structure of isotropic generalized elasticity theories. Math. Mech. Solids 20(5), 565–581 (2015). https://doi.org/10.1177/1081286513507941. http://www.scopus.com/inward/record.url?eid=2-s2.0-84930449973&partnerID=tZOtx3y1

    Article  MathSciNet  MATH  Google Scholar 

  6. Böhlke, T.: Application of the maximum entropy method in texture analysis. Comput. Mater. Sci. 32(3–4), 276–283 (2005). https://doi.org/10.1016/j.commatsci.2004.09.041

    Article  Google Scholar 

  7. Böhlke, T.: Texture simulation based on tensorial Fourier coefficients. Comput. Struct. 84(17–18), 1086–1094 (2006). https://doi.org/10.1016/j.compstruc.2006.01.006

    Article  Google Scholar 

  8. Böhlke, T., Bertram, A., Krempl, E.: Modeling of deformation induced anisotropy in free-end torsion. Int. J. Plast. 19(11 SPEC.), 1867–1884 (2003). https://doi.org/10.1016/S0749-6419(03)00043-3

    Article  MATH  Google Scholar 

  9. Böhlke, T., Risy, G., Bertram, A.: A texture component model for anisotropic polycrystal plasticity. Comput. Mater. Sci. 32(3–4), 284–293 (2005). https://doi.org/10.1016/j.commatsci.2004.09.040. http://linkinghub.elsevier.com/retrieve/pii/S092702560400206X

    Article  Google Scholar 

  10. Bunge, H.J.: In: Texture Analysis in Materials Science: Mathematical Methods, Butterworth, London (1982)

    Google Scholar 

  11. Cao, T., Cuffari, D., Bongiorno, A.: First-principles calculation of third-order elastic constants via numerical differentiation of the second Piola-Kirchhoff stress tensor. Phys. Rev. Lett. 121(21), 1 (2018). https://doi.org/10.1103/PhysRevLett.121.216001

    Article  Google Scholar 

  12. Fernández, M., GitHub repository (2019). https://github.com/mauricio-fernandez-l/centralODF-average

  13. Fullwood, D.T., Niezgoda, S.R., Adams, B.L., Kalidindi, S.R.: Microstructure sensitive design for performance optimization. Prog. Mater. Sci. 55(6), 477–562 (2010). https://doi.org/10.1016/j.pmatsci.2009.08.002

    Article  Google Scholar 

  14. Gel’fand, I.M., Minlos, R., Shapiro, Z.: Representations of the Rotation and Lorentz Groups and Their Applications. Pergamon Press, Oxford (1963). http://books.google.de/books/about/Representations_of_the_rotation_and_Lore.html?id=YbwmAAAAMAAJ&pgis=1

    MATH  Google Scholar 

  15. Glüge, R., Kalisch, J., Bertram, A.: The eigenmodes in isotropic strain gradient elasticity. In: Generalized Continua as Models for Classical and Advanced Materials, pp. 163–178 (2016)

    Chapter  Google Scholar 

  16. Guidi, M., Adams, B.L., Onat, E.T.: Tensorial representation of the orientation distribution function in cubic polycrystals. Textures Microstruct. 19(3), 147–167 (1992). https://doi.org/10.1155/TSM.19.147

    Article  Google Scholar 

  17. Hearmon, R.F.S.: ‘third-order’ elastic coefficients. Acta Crystallogr. 6(4), 331–340 (1953). https://doi.org/10.1107/s0365110x53000909

    Article  MathSciNet  MATH  Google Scholar 

  18. Jakata, K., Every, A.G.: Determination of the dispersive elastic constants of the cubic crystals Ge, Si, GaAs, and InSb. Phys. Rev. B, Condens. Matter Mater. Phys. 77(17), 1 (2008). https://doi.org/10.1103/PhysRevB.77.174301

    Article  Google Scholar 

  19. Jerphagnon, J., Chemla, D., Bonneville, R.: The description of the physical properties of condensed matter using irreducible tensors. Adv. Phys. 27(4), 609–650 (1978). https://doi.org/10.1080/00018737800101454.

    Article  ADS  Google Scholar 

  20. Kalidindi, S.R., Houskamp, J.R., Lyons, M., Adams, B.L.: Microstructure sensitive design of an orthotropic plate subjected to tensile load. Int. J. Plast. 20(8–9), 1561–1575 (2004). https://doi.org/10.1016/j.ijplas.2003.11.007.

    Article  MATH  Google Scholar 

  21. Kearsley, E.A., Fong, J.T.: Linearly independent sets of isotropic Cartesian tensors of ranks up to eight. J. Res. Natl. Bur. Stand. B, Math. Sci. 79B(1), 49 (1975). https://doi.org/10.6028/jres.079B.005. https://nvlpubs.nist.gov/nistpubs/jres/79B/jresv79Bn1-2p49_A1b.pdf

    Article  MathSciNet  MATH  Google Scholar 

  22. Korobov, A.I., Prokhorov, V.M., Mekhedov, D.M.: Second-order and third-order elastic constants of B95 aluminum alloy and B95/nanodiamond composite. Phys. Solid State 55(1), 8–11 (2013). https://doi.org/10.1134/S1063783413010216

    Article  ADS  Google Scholar 

  23. Lobos Fernández, M.: Homogenization and materials design of mechanical properties of textured materials based on zeroth-, first- and second-order bounds of linear behavior. Doctoral thesis, Karlsruhe Institute of Technology (2018). https://doi.org/10.5445/KSP/1000080683. https://publikationen.bibliothek.kit.edu/1000080683

  24. Lobos Fernández, M., Böhlke, T.: Representation of Hashin–Shtrikman bounds in terms of texture coefficients for arbitrarily anisotropic polycrystalline materials. J. Elast. 134, 1–38 (2018). https://doi.org/10.1007/s10659-018-9679-0

    Article  MathSciNet  MATH  Google Scholar 

  25. Lobos, M., Yuzbasioglu, T., Böhlke, T.: Homogenization and materials design of anisotropic multiphase linear elastic materials using central model functions. J. Elast. 128(1), 17–60 (2017). https://doi.org/10.1007/s10659-016-9615-0.

    Article  MathSciNet  MATH  Google Scholar 

  26. Mackey, J.E., Arnold, R.T.: Some combinations of third-order elastic constants for strontium titanate single crystals. J. Appl. Phys. 40(12), 4806–4811 (1969). https://doi.org/10.1063/1.1657293

    Article  ADS  Google Scholar 

  27. Man, C.-S., Huang, M.: A representation theorem for material tensors of weakly-textured polycrystals and its applications in elasticity. J. Elast. 106(1), 1–42 (2012). https://doi.org/10.1007/s10659-010-9284-3

    Article  MathSciNet  MATH  Google Scholar 

  28. Schouten, J.A.: Der Ricci-Kalkül. Springer, Berlin (1924). https://doi.org/10.1007/978-3-662-06545-7.

    Book  MATH  Google Scholar 

  29. Takahashi, S., Motegi, R.: Measurement of third-order elastic constants and applications to loaded structural materials. SpringerPlus 4(1), 1–20 (2015). https://doi.org/10.1186/s40064-015-1019-2

    Article  Google Scholar 

  30. Vilenkin, N.J.: Special Functions and the Theory of Group Representations, vol. 22. Am. Math. Soc., Providence (1968)

    Book  Google Scholar 

  31. Voigt, W.: Lehrbuch der Kristallphysik (mit Ausschluss der Kristalloptik). Teubner, Leipzig (1910)

    MATH  Google Scholar 

  32. Zheng, Q.-S., Fu, Y.-B.: Orientation distribution functions for microstructures of heterogeneous materials (II)—crystal distribution functions and irreducible tensors restricted by various material symmetries. Appl. Math. Mech. 22(8), 885–902 (2001)

    Article  MathSciNet  Google Scholar 

  33. Zheng, Q.-S., Spencer, A.J.M.: On the canonical representations for Kronecker powers of orthogonal tensors with application to material symmetry problems. Int. J. Eng. Sci. 31(4), 617–635 (1993). https://doi.org/10.1016/0020-7225(93)90054-X

    Article  MathSciNet  MATH  Google Scholar 

  34. Zou, W.-N., Zheng, Q.-S., Du, D.-X., Rychlewski, J.: Orthogonal irreducible decompositions of tensors of high order. Math. Mech. Solids 6(3), 249–267 (2001). https://doi.org/10.1177/108128650100600303

    Article  MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Mauricio Fernández.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: Generation of Bases for Isotropic Tensors

Consider the isotropic tensor

$$ \mathbb{D}_{\langle n \rangle } = \int _{\mathsf{SO(3)}}D(\omega ) \boldsymbol{Q}^{\star (n/2)} { \,\mathrm{dQ}} $$
(91)

for some Dirichlet kernel \(D(\omega )\), as considered in the present work. The isotropic tensor (91) can be computed with a given basis for the space of isotropic \(n\)-th-order tensors \(\mathsf{T}^{ \mathrm{iso}}_{n}\) with dimension \(d^{\mathrm{iso}}_{n}\), given in terms of the Motzkin sum numbers \(a_{n}\)

$$ d^{\mathrm{iso}}_{n} = a_{n} , \qquad a_{i} = \frac{i-1}{i+1}(2 a _{i-1} + 3 a_{i-2}) , \quad a_{0} = 1, \ a_{1} = 0 , $$
(92)

see, e.g., http://oeis.org/A005043. The generation of basis of \(\mathsf{T}^{\mathrm{iso}}_{n}\) is non-trivial for high \(n\), see, [4, 21] or [3]. This section offers a pragmatic computational algorithm for the automated generation of a basis, which is easily implemented in Python 3, see [12].

The fundamental isotropic tensor

$$ \mathbb{B}_{{\langle n \rangle }} = \textstyle\begin{cases} \boldsymbol{I}^{\otimes (n/2)} & n \ \text{even} \\ \boldsymbol{\epsilon }\otimes \boldsymbol{I}^{\otimes ((3-n)/2)} & n\ \text{odd} \end{cases} $$
(93)

and the list of permutations \(L_{P}\) with a total of \(n_{P}\) permutations of indices \((i_{1},i_{2},\dots ,i_{n})\)

$$ L_{P} = \{p_{1},p_{2},\dots ,p_{n_{P}}\} , \quad n_{P} = n! $$
(94)

are considered for the generation of isotropic tensors and a basis \(B_{n}\). It is shortly noted that, of course, \(n_{P}\) is much larger than \(d^{\mathrm{iso}}_{n}\), as given in Table 3.

Table 3 Numbers \(n_{P}\) and \(d^{\mathrm{iso}}_{n}\) for \(n = 4, \dots ,12\)

The application \(\rho \) of a permutation \(p \in L_{P}\) on \(\mathbb{B} _{{\langle n \rangle }}\) yields, in general, a different but still isotropic tensor

$$ \rho (\mathbb{B}_{\langle n \rangle },p) \in \mathsf{T}^{\mathrm{iso}} _{n} \quad \forall p \in L_{P} . $$
(95)

We say that two permutations \(\hat{p}_{1}\) and \(\hat{p}_{2}\) are linear independent, if the generated isotropic tensors \(\rho (\mathbb{B}_{ \langle n \rangle },\hat{p}_{1})\) and \(\rho (\mathbb{B}_{\langle n \rangle },\hat{p}_{2})\) are linear independent. Correspondingly, we say that a permutation list \(\hat{L}_{P} \subset L_{P}\) is a permutation basis, if it contains exactly \(d^{\mathrm{iso}}_{n}\) linear independent permutations.

The objective of this appendix is to generate a permutation basis computationally and pragmatically. For a straight-forward implementation with vector and matrix operations, each \(\hat{\mathbb{B}}_{{\langle n \rangle }i} = \rho (\mathbb{B}_{\langle n \rangle },\hat{p}_{i}), \hat{p}_{i} \in \hat{L}_{P}\) from a candidate list \(\hat{L}_{P}\) of length \(\hat{n} \geq d^{\mathrm{iso}}_{n}\) can be reshaped into a vector \(\hat{\underline{b}}_{i}\) delivering a matrix \(\hat{\underline{\underline{B}}}\), defined as

$$ \hat{\underline{b}}_{i} = \mathrm{vec}(\hat{\mathbb{B}}_{{\langle n \rangle }i}) \in \mathsf{R}^{3^{n}}, \qquad \hat{\underline{\underline{B}}}= (\hat{ \underline{b}}_{1} \hat{\underline{b}}_{2} \dots \hat{ \underline{b}}_{\hat{n}}) \in \mathsf{R}^{3^{n} \times \hat{n}} $$
(96)

The candidate list \(\hat{L}_{P}\) is a permutation basis if and only if \(\mathrm{rank}(\hat{\underline{\underline{B}}}) = d^{\mathrm{iso}} _{n}\). This offers the following pragmatic strategy: start with a candidate list \(\hat{L}_{P}\) and keep adding index permutations to it until \(\mathrm{rank}(\hat{\underline{\underline{B}}}) = d^{ \mathrm{iso}}_{n}\) holds, then extract the index set of linearly independent columns of \(\hat{\underline{\underline{B}}}\) (e.g., based on its QR factorization) and extract the corresponding permutation basis based on this index set. It remains to choose a generation of index permutations to be tested. Since for high \(n\), the number of permutations \(n_{P} = n!\) is large in comparison to \(d^{\mathrm{iso}} _{n}\) and ordered index permutations are highly probable to be linear dependent due to the index symmetries of \(\boldsymbol{I}\) and \(\boldsymbol{\epsilon }\), the usage of random permutations is taken into account. The straight forward idea is then to generate some sensible/pragmatic number of additional random permutations \(\hat{n}_{+} \leq d^{\mathrm{iso}}_{n}\), build the matrix \(\hat{\underline{\underline{B}}}\), compute its rank and repeat until \(\mathrm{rank}(\hat{\underline{\underline{B}}}) = d^{\mathrm{iso}} _{n}\) holds. This simple approach is given in Algorithm 1.

Algorithm 1
figure 2

Generation of a permutation basis based on random permutations

Files containing permutation bases up to \(n=12\) generated with Algorithm 1 are available at [12].

Now, we assume that a permutation basis \(\hat{L}_{P}\) has been found. Define the resulting basis of isotropic tensor as

$$ \hat{B}_{n} = \{\hat{\mathbb{B}}_{{\langle n \rangle }1},\dots \}, \qquad \hat{\mathbb{B}}_{{\langle n \rangle }i} = \rho (\mathbb{B}_{ \langle n \rangle }, \hat{p}_{i}), \quad \hat{p}_{i} \in \hat{L} _{P}. $$
(97)

The isotropic tensor (91) has a representation in terms of the basis \(\hat{B}_{n}\)

$$ \mathbb{D}_{\langle n \rangle } = \sum_{i=1}^{d^{\mathrm{iso}}_{n}} \hat{c}_{i} \hat{\mathbb{B}}_{{\langle n \rangle }i} $$
(98)

with coefficients \(\hat{c}_{i}\). These coefficients are determined through the corresponding linear system

$$ \hat{\underline{D}}= \tilde{\underline{\underline{B}}}\ \hat{\underline{c}}, \qquad \hat{D}_{i} = \mathbb{D}_{\langle n \rangle } \cdot \hat{ \mathbb{B}}_{{\langle n \rangle }i} = \int _{\mathsf{SO(3)}}D(\omega ) \bigl(\boldsymbol{Q}^{\star (n/2)} \cdot \hat{\mathbb{B}}_{{\langle n \rangle }i}\bigr) \,\mathrm{dQ}, \qquad \tilde{B}_{ij} = \hat{\mathbb{B}}_{{\langle n \rangle }i} \cdot \hat{ \mathbb{B}}_{{\langle n \rangle }j}. $$
(99)

The matrix \(\tilde{\underline{\underline{B}}}\) is regular and symmetric, such that efficient decompositions can be considered for the solution of the linear system. Alternatively, based on the permutation basis and corresponding basis isotropic tensors, a new set of orthonormal isotropic tensors (or corresponding flattened vectors) may be computed with standard Gram-Schmidt or other orthogonalization approaches to reduced the computational effort and avoid badly conditioned \(\tilde{\underline{\underline{B}}}\). Such an orthonormal basis

$$ B_{n} = \{\mathbb{B}_{{\langle n \rangle }1},\dots \} , \qquad \underline{b}_{i} = \mathrm{vec}(\mathbb{B}_{{\langle n \rangle }i}) , \qquad \mathbb{B}_{{\langle n \rangle }i} \cdot \mathbb{B}_{{\langle n \rangle }j} = \underline{b}_{i}^{ \mathrm{T}}\underline{b}_{j} = \delta _{ij} $$
(100)

would immediately deliver the results given in (49) (with \(n = 2r\)).

Appendix B: Simplification of Integrals for the Computation of \(\mathbb{D}_{{\langle 2r \rangle }\alpha }\)

As remarked after (49), the scalar

$$ q = \boldsymbol{Q}^{\star (n/2)} \cdot \mathbb{T}_{\langle n \rangle } , \quad \mathbb{T}_{\langle r \rangle } \in \mathsf{T}^{ \mathrm{iso}}_{n} $$
(101)

does not depend on the rotation axis \(\boldsymbol{n}\) of \(\boldsymbol{Q}= \boldsymbol{Q}(\boldsymbol{n},\omega )\), i.e., \(q = q(\boldsymbol{n},\omega ) = q(\tilde{\boldsymbol{n}},\omega )\) for all normalized \(\boldsymbol{n},\tilde{\boldsymbol{n}}\in T_{1}\). This can be shown by considering \(\tilde{\boldsymbol{n}}= \boldsymbol{R} \boldsymbol{n}\) with an arbitrary rotation \(\boldsymbol{R}\) and using the isotropy property of \(\mathbb{T}_{\langle n \rangle }\). First, note that

$$ \boldsymbol{Q}(\boldsymbol{R}\boldsymbol{n},\omega ) = \boldsymbol{R} \boldsymbol{Q}(\boldsymbol{n},\omega )\boldsymbol{R}^{\mathrm{T}} $$
(102)

holds. Then, keeping in mind that any transposition of an isotropic tensor is just another isotropic tensor, we represent the Rayleigh power through the tensor power and the corresponding transposition \({\mathrm{T}}_{\star }\) and compute

$$ \boldsymbol{Q}^{\star (n/2)} \cdot \mathbb{T}_{\langle n \rangle } = \bigl( \boldsymbol{Q}^{\otimes (n/2)}\bigr)^{{\mathrm{T}}_{\star }} \cdot \mathbb{T}_{\langle n \rangle } = \boldsymbol{Q}^{\otimes (n/2)} \cdot \mathbb{T}^{*}_{\langle n \rangle } , $$
(103)

where \(\mathbb{T}^{*}_{\langle n \rangle }\) denotes the corresponding transposed/permuted isotropic tensor of \(\mathbb{T}_{\langle n \rangle }\). Inserting (102) into (103) yields

$$\begin{aligned} \bigl(\boldsymbol{R}\boldsymbol{Q}(\boldsymbol{n},\omega )\boldsymbol{R} ^{\mathrm{T}}\bigr)^{\otimes (n/2)} \cdot \mathbb{T}^{*}_{\langle n \rangle } &= \boldsymbol{Q}(\boldsymbol{n},\omega )^{\otimes (n/2)} \cdot \bigl( \boldsymbol{R}^{\mathrm{T}}\star \mathbb{T}^{*}_{\langle n \rangle } \bigr) \\ &= \boldsymbol{Q}(\boldsymbol{n},\omega )^{\otimes (n/2)} \cdot \mathbb{T}^{*}_{\langle n \rangle } \quad \forall \boldsymbol{R} \in \mathsf{SO(3)}, \end{aligned}$$
(104)

meaning that \(q\), given in (101), is independent of \(\boldsymbol{n}\), such that we could just use, e.g., \(\boldsymbol{n}= \boldsymbol{e}_{1}\) for rapid computations. This, naturally, implies that, if we use the orthonormal basis (100) (with \(n=2r\)) to compute \(\mathbb{D}_{{\langle 2r \rangle }\alpha }\), then the integrals \(b_{i}^{\alpha }\) of (49) can be simplified to the one-dimensional integrals

$$ b_{i}^{\alpha }= \mathbb{D}_{{\langle 2r \rangle }\alpha }\cdot \mathbb{B}_{{\langle 2r \rangle }i} = \int _{0}^{\pi }(1+2\alpha ) D _{\alpha }( \omega ) \bigl(\boldsymbol{Q}(\boldsymbol{e}_{1},\omega )^{ \star r} \cdot \mathbb{B}_{{\langle 2r \rangle }i}\bigr) \ 4\pi \ m(\omega ) \,\mathrm{d}\omega , $$
(105)

which are easily treatable in computer algebra systems like Mathematica®12 or even with numerical modules like Numpy for Python 3, see [12].

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Fernández, M. On the Orientation Average Based on Central Orientation Density Functions for Polycrystalline Materials. J Elast 139, 331–357 (2020). https://doi.org/10.1007/s10659-019-09754-8

Download citation

  • Received:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10659-019-09754-8

Mathematics Subject Classification

Keywords

Navigation