scientific commentaries\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoFOUNDATIONS
ADVANCES
ISSN: 2053-2733

On symmetries of higher-order elastic constants

crossmark logo

aTerminal Effects Division, DEVCOM ARL, Aberdeen Proving Ground, MD, 21005-5066, USA
*Correspondence e-mail: john.d.clayton1.civ@army.mil

In elastic crystals, a hyperelastic description is conventionally assumed, and the strain energy potential is idealized as a Taylor-series expansion in strain about an unstrained reference state. Coefficients of quadratic terms are second-order or linear elastic constants. Coefficients of higher-order terms are elastic constants of third order, fourth order, and so on. Recently published work by Telyatnik [Acta Cryst. (2024), A80, 394–404] extends prior knowledge of symmetry properties for anisotropic elastic constants of single crystals, as well as transversely isotropic and isotropic solids, to terms up to sixth order. Effective elastic constants for polycrystalline aggregates, with possible anisotropy, were reported by Telyatnik, in the same article, to the same order. A terse summary of nonlinear crystal elasticity and independent elastic constants of orders two and three are given in this commentary for context. Methods and results of Telyatnik, anticipated to be of great utility to crystal elasticity research, are then highlighted.

1. Nonlinear elasticity of crystals

Linear elasticity accurately describes the mechanical response of elastic solids when deformations are small and the stress–strain response is linear, meaning Hooke's law applies. Linear elasticity theory (Voigt, 1910[Voigt, W. (1910). Lehrbuch der Kristallphysik. Leipzig: Teubner.]; Love, 1927[Love, A. (1927). A Treatise on the Mathematical Theory of Elasticity, 4th ed. Cambridge University Press.]; Hearmon, 1946[Hearmon, R. (1946). Rev. Mod. Phys. 18, 409-440.]) can be successfully applied to many, if not most, problems in structural mechanics. A nonlinear theory, on the other hand, is needed for accurate descriptions of mechanics of solids when deformations are large or when linearity breaks down.

All known solids are ultimately nonlinear. No finite volume of any real material can be compressed to a point of infinitesimal size: its bulk modulus must eventually increase with decreasing volume to prevent this. Nonlinear elasticity can describe wave propagation in pre-stressed crystals (Thurston & Brugger, 1964[Thurston, R. & Brugger, K. (1964). Phys. Rev. 133, A1604-A1610.]; Thurston et al., 1966[Thurston, R., McSkimin, H. & Andreatch, P. (1966). J. Appl. Phys. 37, 267-275.]; Thurston, 1974[Thurston, R. (1974). Handbuch der Physik, edited by C. Truesdell, vol. VI, pp. 109-308. Berlin: Springer.]) and short-range core effects from lattice defects (Teodosiu, 1982[Teodosiu, C. (1982). Elastic Models of Crystal Defects. Berlin: Springer.]). At regimes departing more from linearity (Chang & Barsch, 1967[Chang, Z. & Barsch, G. (1967). Phys. Rev. Lett. 19, 1381-1382.]), or in shock compression, elastic constants up to order four have been measured (Fowles, 1967[Fowles, R. (1967). J. Geophys. Res. 72, 5729-5742.]; Graham, 1972[Graham, R. (1972). J. Acoust. Soc. Am. 51, 1576-1581.]). In strong crystals such as quartz, sapphire and diamond, large uniaxial compressive deformations can be reached before inelasticity from dislocation motion, deformation twinning or fracture ensues (Clayton, 2019[Clayton, J. (2019). Nonlinear Elastic and Inelastic Models for Shock Compression of Crystalline Solids. Cham: Springer.]). Chen et al. (2020[Chen, H., Zarkevich, N., Levitas, V., Johnson, D. & Zhang, X. (2020). npj Comput. Mater. 6, 115.]) showed the importance of anisotropic constants up to fifth order on the stability of silicon.

Symmetries of second-order constants for anisotropic elasticity of crystals have been known for over a century (Voigt, 1910[Voigt, W. (1910). Lehrbuch der Kristallphysik. Leipzig: Teubner.]; Hearmon, 1946[Hearmon, R. (1946). Rev. Mod. Phys. 18, 409-440.]). Independent third-order constants have been known for all crystal classes since works by Fumi (1951[Fumi, F. (1951). Phys. Rev. 83, 1274-1275.]) and Hearmon (1953[Hearmon, R. F. S. (1953). Acta Cryst. 6, 331-340.]), and those of fourth-order since the work of Brendel (1979[Brendel, R. (1979). Acta Cryst. A35, 525-533.]). Measurements of third-order constants often rely on sound speeds in pre-stressed crystals (Thurston & Brugger, 1964[Thurston, R. & Brugger, K. (1964). Phys. Rev. 133, A1604-A1610.]; Brugger, 1965[Brugger, K. (1965). J. Appl. Phys. 36, 759-768.]; Thurston, 1974[Thurston, R. (1974). Handbuch der Physik, edited by C. Truesdell, vol. VI, pp. 109-308. Berlin: Springer.]). Contemporary first-principles atomic simulations involving density functional theory (DFT) have been used to predict constants up to orders four and five (Chen et al., 2020[Chen, H., Zarkevich, N., Levitas, V., Johnson, D. & Zhang, X. (2020). npj Comput. Mater. 6, 115.]; Pandit & Bongiorno, 2023[Pandit, A. & Bongiorno, A. (2023). Comput. Phys. Commun. 288, 108751.]). Since elastic constants are ultimately related to interatomic forces (Born & Huang, 1954[Born, M. & Huang, K. (1954). Dynamical Theory of Crystal Lattices. London: Oxford University Press.]; Wallace, 1972[Wallace, D. (1972). Thermodynamics of Crystals. New York: John Wiley and Sons.]), characteristics of elastic constants give insight into atomic-scale physics, and vice versa. Higher-order constants are associated with anharmonicity (Hiki & Granato, 1966[Hiki, Y. & Granato, A. (1966). Phys. Rev. 144, 411-419.]; Hiki, 1981[Hiki, Y. (1981). Annu. Rev. Mater. Sci. 11, 51-73.]).

A brief primer is given here; theoretical presentations of nonlinear elasticity for anisotropic crystals are available in books on the subject (Thurston, 1974[Thurston, R. (1974). Handbuch der Physik, edited by C. Truesdell, vol. VI, pp. 109-308. Berlin: Springer.]; Teodosiu, 1982[Teodosiu, C. (1982). Elastic Models of Crystal Defects. Berlin: Springer.]; Clayton, 2011[Clayton, J. (2011). Nonlinear Mechanics of Crystals. Dordrecht: Springer.]). Let [{\bf{x}} = {\bf{x}}({\bf{X}},t)] be the spatial position vector of a material particle that occupied a reference position [{\bf{X}}] at some initial time [t = t_{0}]. Cartesian coordinates (xk,XK) with [k,K = 1,2,3] are used for Euclidean 3-space, and repeated indices are summed. The deformation gradient [{\bf{F}}({\bf{X}},t)] is the two-point tensor

[{\bf{F}} = {\partial{\bf{x}}}/{\partial{\bf{X}}}\,\leftrightarrow\,F_{iJ} = { \partial x_{i}}/{\partial X_{J}},\quad\det{\bf{F}}\,\gt\,0.\eqno(1)]

Per the Cauchy–Born rule, primitive Bravais lattice vectors of a crystalline material at point [({\bf X},t)] deform affinely with [{\bf{F}}({\bf{X}},t)] (Born & Huang, 1954[Born, M. & Huang, K. (1954). Dynamical Theory of Crystal Lattices. London: Oxford University Press.]; Clayton, 2011[Clayton, J. (2011). Nonlinear Mechanics of Crystals. Dordrecht: Springer.]). The classical strain measure for nonlinear crystal elasticity is the Green–Lagrange strain (Wallace, 1972[Wallace, D. (1972). Thermodynamics of Crystals. New York: John Wiley and Sons.]; Thurston, 1974[Thurston, R. (1974). Handbuch der Physik, edited by C. Truesdell, vol. VI, pp. 109-308. Berlin: Springer.]; Clayton, 2011[Clayton, J. (2011). Nonlinear Mechanics of Crystals. Dordrecht: Springer.]):

[{\bf{E}} = {\textstyle{{{1} \over {2}}}}({\bf{F}}^{\rm T}{\bf{F}}-{\bf 1}) = {\bf {E}}^{\rm T}\leftrightarrow E_{IJ} = {\textstyle{{{1} \over {2}}}}(F_{kI}F_{kJ} -\delta_{IJ}) = E_{JI}.\eqno(2)]

Strain energy per unit initial volume is [W({\bf E}({\bf F})) = {\bar {\bf W}}({\bf{F}})]. Isentropic or isothermal conditions are implied; elastic constants are isentropic or isothermal values. Cauchy (true) stress is

[{\boldsigma} = {{1} \over {\det{\bf{F}}}}{\bf{F}}{{\partial W} \over {\partial{\bf{E}} }}{\bf{F}}^{{\rm T}}\leftrightarrow\sigma_{ij} = {{1} \over {\det F_{kK}}}F_{iL} {{\partial W} \over {\partial E_{LM}}}F_{jM}.\eqno(3)]

Using Greek indices [\alpha,\beta,\gamma,\ldots = 1,2,\ldots,6] for Voigt notation (Thurston, 1974[Thurston, R. (1974). Handbuch der Physik, edited by C. Truesdell, vol. VI, pp. 109-308. Berlin: Springer.]; Clayton, 2011[Clayton, J. (2011). Nonlinear Mechanics of Crystals. Dordrecht: Springer.]), series expansion of W gives

[\eqalignno{ W& = W_{0}+C_{\alpha}E_{\alpha}+{\textstyle{{{1} \over { 2}}}}C_{\alpha\beta}E_{\alpha}E_{\beta}+{\textstyle{{{1} \over {6}}}}C_{\alpha\beta \gamma}E_{\alpha}E_{\beta}E_{\gamma}&\cr &+{\textstyle{{{1} \over {24}}}}C_{\alpha\beta\gamma\delta}E_{\alpha} E_{\beta}E_{\gamma}E_{\delta}+{\textstyle{{{1} \over {120}}}}C_{\alpha\beta\gamma \delta\epsilon}E_{\alpha}E_{\beta}E_{\gamma}E_{\delta}E_{\epsilon}+\ldots.&\cr &&(4)}]

The constant datum energy is W0, first-order [C_{\alpha} = 0] for a stress-free reference state, second-order constants are [C_{\alpha\beta}], third-order are [C_{\alpha\beta\gamma}], and so on. Constants have implied symmetries:

[C_{\alpha\beta} = C_{\beta\alpha},\quad C_{\alpha\beta\gamma} = C_{\beta\alpha \gamma} = C_{\alpha\gamma\beta} = C_{\gamma\beta\alpha},\quad\ldots.\eqno(5)]

For materials (i.e. crystal classes) of lowest (i.e. triclinic) symmetry, [C_{\alpha\beta}] and [C_{\alpha\beta\gamma}] have 21 and 56 independent components, respectively.

Denote by [{\bf{R}}] an orthogonal matrix ([{\bf{R}}^{{\rm T}}{\bf{R}} = {\bf 1}]) that belongs to the symmetry group of transformations for a given material (e.g. its Laue group or crystal class). Then [{\bar W}({\bf{F}}{\bf{R}}) = {\bar W}({\bf{F}})] yields constraints among the elastic constants of each order dictated by that symmetry group. The greater the intrinsic symmetry, the more expansive the symmetry group, and heuristically the fewer independent elastic constants of a given order. For example, if a material is isotropic, [{\bf{R}}] can be any rotation, limiting the number of independent [C_{\alpha\beta}] to two and [C_{\alpha\beta\gamma}] to three.

For crystal structures, standard conventions are used to relate coordinate axes to directions in the lattice (Brainerd et al., 1949[Brainerd, J. et al. (1949). Proc. Inst. Radio Eng. 37, 1378-1395.]); an example is shown in Fig. 1[link] for a triclinic crystal. Independent second- and third-order elastic constants (Fumi, 1951[Fumi, F. (1951). Phys. Rev. 83, 1274-1275.]; Hearmon, 1953[Hearmon, R. F. S. (1953). Acta Cryst. 6, 331-340.]) for all 11 Laue groups and isotropic solids are reported for easy reference in Tables 1[link] and 2[link], following Brugger (1965[Brugger, K. (1965). J. Appl. Phys. 36, 759-768.]), Thurston (1974[Thurston, R. (1974). Handbuch der Physik, edited by C. Truesdell, vol. VI, pp. 109-308. Berlin: Springer.]) and Clayton (2011[Clayton, J. (2011). Nonlinear Mechanics of Crystals. Dordrecht: Springer.]).

Table 1
Second-order elastic constants

N, triclinic; M, monoclinic; O, orthorhombic; R, rhombohedral; T, tetragonal; H, hexagonal; C, cubic; iso, isotropic; I and II, classes of respective higher and lower symmetry. [{\rm A}: = {\textstyle{{{1} \over {2}}}}(C_{11}-C_{12})]. Bottom row: No. of independent constants. See Thurston (1974[Thurston, R. (1974). Handbuch der Physik, edited by C. Truesdell, vol. VI, pp. 109-308. Berlin: Springer.]), Clayton (2011[Clayton, J. (2011). Nonlinear Mechanics of Crystals. Dordrecht: Springer.]).

N M O TII TI RII RI HII HI CII CI iso
11 11 11 11 11 11 11 11 11 11 11 11
12 12 12 12 12 12 12 12 12 12 12 12
13 13 13 13 13 13 13 13 13 12 12 12
14 0 0 0 0 14 14 0 0 0 0 0
15 15 0 0 0 15 0 0 0 0 0 0
16 0 0 16 0 0 0 0 0 0 0 0
22 22 22 11 11 11 11 11 11 11 11 11
23 23 23 13 13 13 13 13 13 12 12 12
24 0 0 0 0 −14 −14 0 0 0 0 0
25 25 0 0 0 −15 0 0 0 0 0 0
26 0 0 −16 0 0 0 0 0 0 0 0
33 33 33 33 33 33 33 33 33 11 11 11
34 0 0 0 0 0 0 0 0 0 0 0
35 35 0 0 0 0 0 0 0 0 0 0
36 0 0 0 0 0 0 0 0 0 0 0
44 44 44 44 44 44 44 44 44 44 44 A
45 0 0 0 0 0 0 0 0 0 0 0
46 46 0 0 0 −15 0 0 0 0 0 0
55 55 55 44 44 44 44 44 44 44 44 A
56 0 0 0 0 14 14 0 0 0 0 0
66 66 66 66 66 A A A A 44 44 A
                       
21 13 9 7 6 7 6 5 5 3 3 2

Table 2
Third-order elastic constants

Notation in first row follows Table 1[link]. [{\rm A}: = C_{111}+C_{112}-C_{222}]; [{\rm B}: = -{\textstyle{{{1} \over {2}}}}(C_{115}+3C_{125})]; [{\rm C}: = {\textstyle{{{1} \over {2}}}}(C_{114}+3C_{124})]; [{\rm D}: = -{\textstyle{{{1} \over {4}}}}(2C_{111}+C_{112}-3C_{222})]; [{\rm E}: = -C_{114}-2C_{124}]; [{\rm F}: = -C_{115}-2C_{125}]; [{\rm G}: = -{\textstyle{{{1} \over {2}}}}(C_{115}-C_{125})]; [{\rm H}: = {\textstyle{{{1} \over {2}}}}(C_{114}-C_{124})]; [{\rm I}: = {\textstyle{{{1} \over {4}}}}(2C_{111}-C_{112}+C_{222})]; [{\rm J}: = {\textstyle{{{1} \over {2}}}}(C_{113}-C_{123})]; [{\rm K}: = -{\textstyle{{{1} \over {2}}}}(C_{144}-C_{155})]; [{\rm L}: = {\textstyle{{{1} \over {2}}}}(C_{112}-C_{123})]; [{\rm M}: = {\textstyle{{{1} \over {4}}}}(C_{111}-C_{112})]; [{\rm N}: = {\textstyle{{{1} \over {8}}}}(C_{111}-3C_{112}+2C_{123})]. Bottom row: No. of independent constants. See Thurston (1974[Thurston, R. (1974). Handbuch der Physik, edited by C. Truesdell, vol. VI, pp. 109-308. Berlin: Springer.]), Clayton (2011[Clayton, J. (2011). Nonlinear Mechanics of Crystals. Dordrecht: Springer.]).

N M O TII TI RII RI HII HI CII CI iso
111 111 111 111 111 111 111 111 111 111 111 111
112 112 112 112 112 112 112 112 112 112 112 112
113 113 113 113 113 113 113 113 113 113 112 112
114 0 0 0 0 114 114 0 0 0 0 0
115 115 0 0 0 115 0 0 0 0 0 0
116 0 0 116 0 116 0 116 0 0 0 0
122 122 122 112 112 A A A A 113 112 112
123 123 123 123 123 123 123 123 123 123 123 123
124 0 0 0 0 124 124 0 0 0 0 0
125 125 0 0 0 125 0 0 0 0 0 0
126 0 0 0 0 −116 0 −116 0 0 0 0
133 133 133 133 133 133 133 133 133 112 112 112
134 0 0 0 0 134 134 0 0 0 0 0
135 135 0 0 0 135 0 0 0 0 0 0
136 0 0 136 0 0 0 0 0 0 0 0
144 144 144 144 144 144 144 144 144 144 144 L
145 0 0 145 0 145 0 145 0 0 0 0
146 146 0 0 0 B 0 0 0 0 0 0
155 155 155 155 155 155 155 155 155 155 155 M
156 0 0 0 0 C C 0 0 0 0 0
166 166 166 166 166 D D D D 166 155 M
222 222 222 111 111 222 222 222 222 111 111 111
223 223 223 113 113 113 113 113 113 112 112 112
224 0 0 0 0 E E 0 0 0 0 0
225 225 0 0 0 F 0 0 0 0 0 0
226 0 0 −116 0 116 0 116 0 0 0 0
233 233 233 133 133 133 133 133 133 113 112 112
234 0 0 0 0 −134 −134 0 0 0 0 0
235 235 0 0 0 −135 0 0 0 0 0 0
236 0 0 −136 0 0 0 0 0 0 0 0
244 244 244 155 155 155 155 155 155 166 155 M
245 0 0 −145 0 −145 0 −145 0 0 0 0
246 246 0 0 0 G 0 0 0 0 0 0
255 255 255 144 144 144 144 144 144 144 144 L
256 0 0 0 0 H H 0 0 0 0 0
266 266 266 166 166 I I I I 155 155 M
333 333 333 333 333 333 333 333 333 111 111 111
334 0 0 0 0 0 0 0 0 0 0 0
335 335 0 0 0 0 0 0 0 0 0 0
336 0 0 0 0 0 0 0 0 0 0 0
344 344 344 344 344 344 344 344 344 155 155 M
345 0 0 0 0 0 0 0 0 0 0 0
346 346 0 0 0 −135 0 0 0 0 0 0
355 355 355 344 344 344 344 344 344 166 155 M
356 0 0 0 0 134 134 0 0 0 0 0
366 366 366 366 366 J J J J 144 144 L
444 0 0 0 0 444 444 0 0 0 0 0
445 445 0 0 0 445 0 0 0 0 0 0
446 0 0 446 0 145 0 145 0 0 0 0
455 0 0 0 0 −444 −444 0 0 0 0 0
456 456 456 456 456 K K K K 456 456 N
466 0 0 0 0 124 124 0 0 0 0 0
555 555 0 0 0 −445 0 0 0 0 0 0
556 0 0 −446 0 −145 0 −145 0 0 0 0
566 566 0 0 0 125 0 0 0 0 0 0
666 0 0 0 0 −116 0 −116 0 0 0 0
                       
56 32 20 16 12 20 14 12 10 8 6 3
[Figure 1]
Figure 1
Standard coordinate axes [\{X_{1},X_{2},X_{3}\}\rightarrow\{X,Y,Z\}] and natural lattice directions (a, b, c) for a triclinic class I crystal. Based on Brainerd et al. (1949[Authier, A. & Zarembowitch, A. (2003). International Tables for Crystallography Vol. D, Physical Properties of Crystals, edited by A. Authier, pp. 72-98. Dordrecht: Springer.]) (redrawn by the author).

Also of theoretical interest are Cauchy symmetries arising if all interatomic forces are central, as from a pair potential (Love, 1927[Love, A. (1927). A Treatise on the Mathematical Theory of Elasticity, 4th ed. Cambridge University Press.]; Born & Huang, 1954[Born, M. & Huang, K. (1954). Dynamical Theory of Crystal Lattices. London: Oxford University Press.]). In tensor form, second-order constants then obey [C_{IJKL} = C_{IKJL} = C_{ILKJ}]; similar constraints arise at higher orders. In cubic classes of greatest symmetry, [C_{12} = C_{44}], [C_{112} = C_{155}] and [C_{123} = C_{144} = C_{456}] in Voigt notation. In a Cauchy isotropic solid, [C_{\alpha\beta}] and [C_{\alpha\beta\gamma}] each contain but one independent constant, and Poisson's ratio [\nu = {{1} \over {4}}].

Symmetric second-order strain tensors differing from [{\bf{E}}] of (2[link]) have been used in nonlinear elastic potentials akin to (4[link]), with noted advantages for describing DFT (Nielsen, 1986[Nielsen, O. (1986). Phys. Rev. B, 34, 5808-5819.]) and shock (Clayton, 2019[Clayton, J. (2019). Nonlinear Elastic and Inelastic Models for Shock Compression of Crystalline Solids. Cham: Springer.]) data. Symmetries of elastic constant tensors of all orders are unchanged (e.g. Tables 1[link] and 2[link] remain valid) so long as the strain tensor has components referred to {XK}; any such strain transforms the same under the action of [{\bf R}]. Values of [C_{\alpha\beta}] are identical for all such strain measures, but values of [C_{\alpha\beta\gamma}] and constants of successively higher orders generally differ, and (3[link]) is transformed for strain different from [{\bf E}]. Another example is linear elasticity, for which the small strain tensor is [{\boldepsilon} = {{1} \over {2}}[\partial{\bf{u}}/\partial{\bf{X}}+(\partial{\bf{u} }/\partial{\bf{X}})^{{\rm T}}]] with displacement [{\bf{u}} = {\bf{x}}-{\bf{X}}]. To first order in [\partial{\bf{u}}/\partial{\bf{X}}], [{\boldepsilon}\approx{\bf{E}}]. Informally, when [|{\bf{F}}-{\bf 1}|\ll 1], [W_{0} = 0] and W is truncated at order two, (3[link]) and (4[link]) give Hooke's law:

[\sigma_{\alpha} = {{\partial W} \over {\partial\epsilon_{\alpha}}} = C_{\alpha\beta} \epsilon_{\beta},\quad W = {\textstyle{{{1} \over {2}}}}C_{\alpha\beta}\epsilon_{ \alpha}\epsilon_{\beta}.\eqno(6)]

The effective elastic constants of heterogeneous solids (e.g. polycrystalline aggregates) depend on the properties and orientations of constituents (Gnaupel-Herold, 2023[Gnäupel-Herold, T. (2023). J. Appl. Cryst. 56, 1658-1673.]). If an aggregate has a certain target symmetry, estimates or bounds on its effective `averaged' constants can be obtained from theoretical averaging schemes. For second-order constants, well known estimates include upper and lower bounds of Voigt and Reuss, respectively, Hill's proposition (Hill, 1952[Hill, R. (1952). Proc. Phys. Soc. A, 65, 349-354.]) that interpolates between the two, and `self-consistent' models (Kroner, 1958[Kroner, E. (1958). Z. Phys. 151, 504-518.]; De Wit, 1997[De Wit, R. (1997). J. Appl. Cryst. 30, 510-511.]). Such methods were extended to third-order constants for isotropic (Barsch, 1968[Barsch, G. (1968). J. Appl. Phys. 39, 3780-3793.]; Lubarda, 1997[Lubarda, V. (1997). J. Mech. Phys. Solids, 45, 471-490.]) and textured (Johnson, 1985[Johnson, G. (1985). J. Appl. Mech. 52, 659-663.]; Kube & Turner, 2016[Kube, C. & Turner, J. (2016). J. Elast. 122, 157-177.]) polycrystals. Isotropic Voigt averages of fourth-order constants for aggregates of cubic crystals were derived by Krasilnikov & Vekilov (2019[Krasilnikov, O. & Vekilov, Y. (2019). Phys. Rev. B, 100, 134107.]).

2. Nonlinear crystal elasticity to sixth order

Telyatnik (2024[Telyatnik, R. S. (2024). Acta Cryst. A80, 394-404.]) recently developed numerically efficient algorithms for symbolic computations of effective elastic constants of orders two through six for polycrystalline aggregates having overall target symmetries of any crystal class, transverse isotropy or full isotropy. Constituent crystallites can have any anisotropy. Effective constants are defined as arithmetic averages over the minimal set of symmetry operations generating a given target symmetry. In cases of targeted transverse or full isotropy, continuous integrals replace discrete averages. For efficiency, nested calculations for higher symmetries apply precomputed averages from lower symmetries. Gauss–Jordan elimination is used to find algebraic relationships among all elastic constants of a given order, for each target symmetry. Computations exceed capabilities of existing tools supplementing Vol. D of the International Tables for Crystallography (Authier & Zarembowitch, 2003[Authier, A. & Zarembowitch, A. (2003). International Tables for Crystallography Vol. D, Physical Properties of Crystals, edited by A. Authier, pp. 72-98. Dordrecht: Springer.]).

Independent elastic constants and symmetry relationships for all crystal classes, transverse isotropy and full isotropy, for orders two through six, are provided in Appendix A of Telyatnik (2024[Telyatnik, R. S. (2024). Acta Cryst. A80, 394-404.]) and Telyatnik (2021[Telyatnik, R. (2021). Mendeley Data, V2, https://data.mendeley.com/datasets/mf8rbjzwmw/2.]). Previously, such information (e.g. as in Tables 1[link] and 2[link]) was available only for constants of all crystal classes and isotropy, to order four (Brendel, 1979[Brendel, R. (1979). Acta Cryst. A35, 525-533.]). Averages for independent elastic constants, again up to sixth order, are given in Appendix B of Telyatnik (2024[Telyatnik, R. S. (2024). Acta Cryst. A80, 394-404.]) and Telyatnik (2021[Telyatnik, R. (2021). Mendeley Data, V2, https://data.mendeley.com/datasets/mf8rbjzwmw/2.]). These include the target symmetries of all crystal classes, transverse isotropy and isotropy. Other appendices (Telyatnik, 2021[Telyatnik, R. (2021). Mendeley Data, V2, https://data.mendeley.com/datasets/mf8rbjzwmw/2.]; Telyatnik, 2024[Telyatnik, R. S. (2024). Acta Cryst. A80, 394-404.]) list all components and rotation matrices. For anisotropic target symmetries, averages are defined as described in the preceding paragraph. These averages do not incorporate data on local crystal orientations (i.e. distribution functions) included in some prior definitions of effective third-order constants (Johnson, 1985[Johnson, G. (1985). J. Appl. Mech. 52, 659-663.]; Kube & Turner, 2016[Kube, C. & Turner, J. (2016). J. Elast. 122, 157-177.]). Therefore, anisotropic aggregate constants of Telyatnik (2024[Telyatnik, R. S. (2024). Acta Cryst. A80, 394-404.]) cannot be expected to reproduce effective constants of textured polycrystalline metals, for example. However, these anisotropic aggregate values can serve as higher-symmetry approximations for constants of classes of lower true symmetry (Telyatnik, 2021[Telyatnik, R. (2021). Mendeley Data, V2, https://data.mendeley.com/datasets/mf8rbjzwmw/2.]).

For isotropic target symmetry, averaged constants (Telyatnik, 2024[Telyatnik, R. S. (2024). Acta Cryst. A80, 394-404.]) are consistent with Voigt's postulate. Previous isotropic averages were usually limited to second and third orders, with the latter for constituents having cubic or hexagonal symmetry (Barsch, 1968[Barsch, G. (1968). J. Appl. Phys. 39, 3780-3793.]; Lubarda, 1997[Lubarda, V. (1997). J. Mech. Phys. Solids, 45, 471-490.]; Kube & Turner, 2016[Kube, C. & Turner, J. (2016). J. Elast. 122, 157-177.]). Previously, the highest known order of derived, isotropic Voigt-averaged constants was four, as reported in Appendix A of Krasilnikov & Vekilov (2019[Krasilnikov, O. & Vekilov, Y. (2019). Phys. Rev. B, 100, 134107.]) for crystallites of general anisotropy (e.g. triclinic symmetry) and verified independently by Telyatnik (2024[Telyatnik, R. S. (2024). Acta Cryst. A80, 394-404.]). Therefore, the isotropic Voigt-type averages for five independent fifth-order constants and seven independent sixth-order constants derived by Telyatnik (2024[Telyatnik, R. S. (2024). Acta Cryst. A80, 394-404.]) appear to be a new and valuable contribution to nonlinear elasticity theory for solid crystals.

Finally, note that for some solids under extreme strain [e.g. rubbery polymers (Ogden, 1984[Ogden, R. (1984). Non-Linear Elastic Deformations. Chichester: Ellis-Horwood.]) and biological tissues (Fung, 1993[Fung, Y.-C. (1993). Biomechanics: Mechanical Properties of Living Tissues, 2nd ed. New York: Springer.])] Taylor polynomials like (4[link]) can be cumbersome so are often replaced with other functional forms (e.g. exponentials) needing fewer constants to fit data. These solids are often idealized as incompressible, for which (3[link]) and (4[link]) are inappropriate.

References

First citationAuthier, A. & Zarembowitch, A. (2003). International Tables for Crystallography Vol. D, Physical Properties of Crystals, edited by A. Authier, pp. 72–98. Dordrecht: Springer.  Google Scholar
First citationBarsch, G. (1968). J. Appl. Phys. 39, 3780–3793.  CrossRef CAS Web of Science Google Scholar
First citationBorn, M. & Huang, K. (1954). Dynamical Theory of Crystal Lattices. London: Oxford University Press.  Google Scholar
First citationBrainerd, J. et al. (1949). Proc. Inst. Radio Eng. 37, 1378–1395.  Google Scholar
First citationBrendel, R. (1979). Acta Cryst. A35, 525–533.  CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationBrugger, K. (1965). J. Appl. Phys. 36, 759–768.  CrossRef Web of Science Google Scholar
First citationChang, Z. & Barsch, G. (1967). Phys. Rev. Lett. 19, 1381–1382.  CrossRef CAS Web of Science Google Scholar
First citationChen, H., Zarkevich, N., Levitas, V., Johnson, D. & Zhang, X. (2020). npj Comput. Mater. 6, 115.  Web of Science CrossRef Google Scholar
First citationClayton, J. (2011). Nonlinear Mechanics of Crystals. Dordrecht: Springer.  Google Scholar
First citationClayton, J. (2019). Nonlinear Elastic and Inelastic Models for Shock Compression of Crystalline Solids. Cham: Springer.  Google Scholar
First citationDe Wit, R. (1997). J. Appl. Cryst. 30, 510–511.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationFowles, R. (1967). J. Geophys. Res. 72, 5729–5742.  CrossRef CAS Web of Science Google Scholar
First citationFumi, F. (1951). Phys. Rev. 83, 1274–1275.  CrossRef CAS Web of Science Google Scholar
First citationFung, Y.-C. (1993). Biomechanics: Mechanical Properties of Living Tissues, 2nd ed. New York: Springer.  Google Scholar
First citationGnäupel-Herold, T. (2023). J. Appl. Cryst. 56, 1658–1673.  Web of Science CrossRef IUCr Journals Google Scholar
First citationGraham, R. (1972). J. Acoust. Soc. Am. 51, 1576–1581.  CrossRef CAS Web of Science Google Scholar
First citationHearmon, R. (1946). Rev. Mod. Phys. 18, 409–440.  CrossRef CAS Web of Science Google Scholar
First citationHearmon, R. F. S. (1953). Acta Cryst. 6, 331–340.  CrossRef IUCr Journals Web of Science Google Scholar
First citationHiki, Y. (1981). Annu. Rev. Mater. Sci. 11, 51–73.  CrossRef CAS Web of Science Google Scholar
First citationHiki, Y. & Granato, A. (1966). Phys. Rev. 144, 411–419.  CrossRef CAS Web of Science Google Scholar
First citationHill, R. (1952). Proc. Phys. Soc. A, 65, 349–354.  CrossRef Web of Science Google Scholar
First citationJohnson, G. (1985). J. Appl. Mech. 52, 659–663.  CrossRef Web of Science Google Scholar
First citationKrasilnikov, O. & Vekilov, Y. (2019). Phys. Rev. B, 100, 134107.  Web of Science CrossRef Google Scholar
First citationKroner, E. (1958). Z. Phys. 151, 504–518.  CAS Google Scholar
First citationKube, C. & Turner, J. (2016). J. Elast. 122, 157–177.  Web of Science CrossRef Google Scholar
First citationLove, A. (1927). A Treatise on the Mathematical Theory of Elasticity, 4th ed. Cambridge University Press.  Google Scholar
First citationLubarda, V. (1997). J. Mech. Phys. Solids, 45, 471–490.  CrossRef CAS Web of Science Google Scholar
First citationNielsen, O. (1986). Phys. Rev. B, 34, 5808–5819.  CrossRef CAS Web of Science Google Scholar
First citationOgden, R. (1984). Non-Linear Elastic Deformations. Chichester: Ellis-Horwood.  Google Scholar
First citationPandit, A. & Bongiorno, A. (2023). Comput. Phys. Commun. 288, 108751.  Web of Science CrossRef Google Scholar
First citationTelyatnik, R. (2021). Mendeley Data, V2, https://data.mendeley.com/datasets/mf8rbjzwmw/2Google Scholar
First citationTelyatnik, R. S. (2024). Acta Cryst. A80, 394–404.  Web of Science CrossRef IUCr Journals Google Scholar
First citationTeodosiu, C. (1982). Elastic Models of Crystal Defects. Berlin: Springer.  Google Scholar
First citationThurston, R. (1974). Handbuch der Physik, edited by C. Truesdell, vol. VI, pp. 109–308. Berlin: Springer.  Google Scholar
First citationThurston, R. & Brugger, K. (1964). Phys. Rev. 133, A1604–A1610.  CrossRef Google Scholar
First citationThurston, R., McSkimin, H. & Andreatch, P. (1966). J. Appl. Phys. 37, 267–275.  CrossRef CAS Web of Science Google Scholar
First citationVoigt, W. (1910). Lehrbuch der Kristallphysik. Leipzig: Teubner.  Google Scholar
First citationWallace, D. (1972). Thermodynamics of Crystals. New York: John Wiley and Sons.  Google Scholar

This article is published by the International Union of Crystallography. Prior permission is not required to reproduce short quotations, tables and figures from this article, provided the original authors and source are cited. For more information, click here.

Journal logoFOUNDATIONS
ADVANCES
ISSN: 2053-2733
Follow Acta Cryst. A
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds