21.05.2019 Views

Kroner

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Derivation and role of <strong>Kroner</strong>’s<br />

cubic in the calculation of shear<br />

moduli of an aggregate of cubic<br />

crystals<br />

Armand Attia<br />

April 12, 2019<br />

This work is licensed under the Creative Commons Attribution 4.0<br />

International License. To view a copy of this license, visit<br />

http://creativecommons.org/licenses/by/4.0/ or send a letter to Creative<br />

Commons, PO Box 1866, Mountain View, CA 94042, USA.<br />

Contents<br />

1 Abstract 3<br />

2 Introduction 4<br />

3 Properties of fourth-order tensors 4<br />

4 Equilibrium 6<br />

5 Integral equation for strains 6<br />

6 Role of Eshelby’s tensor in integral equation for strains 10<br />

7 Symmetry properties and model of tensor D 11<br />

8 Derivation of tensor A for isotropic matrix and isotropic inhomogeneity<br />

12<br />

9 Self consistent scheme for cubic polycrystals 13<br />

10 Elastic moduli of a cubic polycrystal 15<br />

11 Rotational averages 17<br />

12 Derivation of tensor A for anisotropic inhomogeneity and <strong>Kroner</strong>’s<br />

cubic 19<br />

13 Comparison with experimental data 22<br />

14 Major symmetries of tensor D 25


15 Iterative Newton’s method solution of <strong>Kroner</strong>’s cubic 28


1 Abstract<br />

The pioneering work of Hershey [1], <strong>Kroner</strong> [2], and Eshelby [3] has laid the<br />

foundations for modeling polycrystals. The focus here is on <strong>Kroner</strong>’s cubic<br />

which provides a remarkable estimate of the shear modulus of an aggregate of<br />

cubic crystals, arranged in such random orientations that a rotational average<br />

provides a good isotropic estimate. A detailed derivation of <strong>Kroner</strong>’s cubic is<br />

provided beginning with the Green tensor calculation by Weinberger [4], leading<br />

to an integral equation for strains by Gubernatis [5], from which Markov [6]<br />

derives <strong>Kroner</strong>’s cubic. Comparison with experimental data shows good agreement<br />

except for the Carbon family. Markov’s approach opens the potential for<br />

modeling non-cubic crystals.


2 Introduction<br />

This is a detailed review of the derivation of <strong>Kroner</strong>’s cubic [2], which is used<br />

to calculate the macroscopic shear modulus of an aggregate of cubic crystals.<br />

The macroscopic modulus is assumed to apply to an isotropic ensemble of an<br />

aggregate of cubic crystals in perfectly randomized orientations.<br />

The first major part of this derivation is based on the treatment of equilibrium<br />

by Gubernatis [5] and the calculation of the Green tensor function as<br />

derived by Weinberger et al [4], all leading to an integral equation for strains<br />

by Gubernatis [5].<br />

The second major part is based on Markov’s [6] treatment for deriving <strong>Kroner</strong>’s<br />

cubic from this integral equation. The third major part compares measured<br />

shear moduli with the values predicted by <strong>Kroner</strong>’s cubic with Hashin’s bounds.<br />

But first, a few words about fourth-order tensors.<br />

3 Properties of fourth-order tensors<br />

Fourth-order tensors are designated with such script (C), while second-order<br />

tensors are designated without such script (C). Markov [6] defines the composition<br />

of two fourth-order tensors C : B as:<br />

(C : B) ijkl = C ijrs B srkl (1)<br />

For a fourth-order tensor B with only minor symmetry:<br />

the identity tensor I<br />

B ijkl = B jikl (2)<br />

B ijkl = B jilk (3)<br />

has the property:<br />

Since:<br />

I ijkl = 1 2 (δ ikδ jl + δ il δ jk ) (4)<br />

B : I = I : B = B (5)<br />

B ijrs I srkl = 1 2 B ijrs(δ sk δ rl + δ sl δ rk ) = 1 2 (B ijlk + B ijkl ) = B ijkl (6)<br />

I ijrs B srkl = 1 2 (δ irδ js + δ is δ jr )B srkl = 1 2 (B jilk + B ijkl ) = B ijkl (7)<br />

Consider isotropic fourth-order tensors B with both major and minor symmetries:<br />

B ijkl = B jikl = B ijlk = B klij (8)


Such a tensor can be decomposed as a linear combination of the following<br />

orthogonal set of fourth-order tensors I ′ and I ′′<br />

where<br />

with the properties:<br />

so that:<br />

B = b ′ I ′ + b ′′ I ′′ (9)<br />

I ′ ijkl = 1 3 δ ijδ kl (10)<br />

I” ijkl = I ijkl − I ′ ijkl (11)<br />

I ′ : I ′ = I ′ (12)<br />

I ′′ : I ′′ = I ′′ (13)<br />

I ′ : I ′′ = I ′ : (I − I ′ ) = I ′ : I − I ′ : I ′ = 0 (14)<br />

Note, that because of the above properties of I ′ and I ′′ , we find that:<br />

B : I ′ = b ′ I ′ (15)<br />

B : I ′′ = b ′′ I ′′ (16)<br />

so that it is clear that b ′ and b ′′ are eigenvalues of B with corresponding eigentensors<br />

I ′ and I ′′ and that with respect to these eigentensors, B is in diagonal<br />

form. Therefore, its inverse is a tensor A given in terms of the reciprocals of<br />

the eigenvalues (if they do not vanish):<br />

A = B −1 = 1 b ′ I′ + 1<br />

b ′′ I′′ (17)<br />

The scalar product of two fourth-order tensors is defined as:<br />

A :: B = A ijkl B ijkl (18)<br />

Notice that while the tensors I ′ and I ′′ are orthogonal, only I ′ is normalized,<br />

i.e.:<br />

I ′ :: I ′ = 1 (19)<br />

I ′′ :: I ′′ = 5 (20)<br />

Therefore, given an isotropic fourth-order tensor B with major and minor symmetries,<br />

if we write<br />

Then to solve for b ′ and b ′′ , we have:<br />

since I ′ is normalized, but:<br />

B = b ′ I ′ + b ′′ I ′′ (21)<br />

b ′ = B :: I ′ (22)<br />

since I ′′ is not normalized.<br />

b ′′ =<br />

B :: I′′<br />

I ′′ :: I ′′ (23)


4 Equilibrium<br />

Consider a linear elastic solid. At equilibrium with no body forces, the stress<br />

satisfies:<br />

∂σ ij (x)<br />

∂x j<br />

= ∂<br />

∂x j<br />

(C ijkl (x)ɛ kl (x)) = 0 (24)<br />

where the strain is related to the displacement via:<br />

and following Gubernatis [5]:<br />

(25)<br />

ɛ kl = 1 2 (u k,l + u l,k ) (26)<br />

C(x) = C 0 + [C](x) (27)<br />

[C](x) = C 2 (x) − C 1 (x) (28)<br />

in which the elastic modulus C 0 is constant, and the elastic moduli C 1 (x) and<br />

C 2 (x) may vary over an infinite region of volume V but differ only over an<br />

inclusion of finite volume W . C 1 is associated with the ”matrix” region, and<br />

C 2 is associated with an inhomogeneity in the inclusion. We will focus on a<br />

spherical inclusion and on constant moduli C 1 and C 2 .<br />

Therefore, the equilibrium requirement Eq. 24 may be written as:<br />

−C 0 klmnɛ mn,l = ([C] klmn ɛ mn ) ,l<br />

(29)<br />

5 Integral equation for strains<br />

This concerns the determination of displacement fields in a solid matrix (medium<br />

1) containing an inhomogeneity (medium 2) whose elastic properties differ from<br />

those of the surrounding matrix. The goal is to relate the strain in the inhomogeneity<br />

to the known strain when there is no inhomogeneity. Only the<br />

matrix is assumed to be isotropic. Since the equilibrium equation is based on<br />

linear elasticity, Green’s (tensor) function can be used to construct the general<br />

solution for the displacement. The construction of the Green tensor function<br />

follows Weinberger et al [4]. The Green tensor function G mr (x, x ′ ) is defined as<br />

the displacement at x in the m th direction in response to a concentrated unit<br />

force applied in the r th direction at x ′ . In an infinite homogeneous body, the<br />

Green’s function depends only on the relative displacement between the points<br />

because the underlying differential operator here has constant coefficients and<br />

thus is invariant under translation (See Stakgold [7] p. 49). Therefore, Green’s<br />

function may be written as:<br />

G mr (x, x ′ ) = G mr (x − x ′ ) (30)


For a constant point force F acting at x ′ , the displacement at x is given by:<br />

u m (x) = G mr (x − x ′ )F r (x ′ ) (31)<br />

So that the displacement gradient and the strain take the form:<br />

u m,n = G mr,n (x − x ′ )F r (32)<br />

ɛ mn = 1 2 (u m,n + u n,m ) (33)<br />

and from the constitutive relation, the stress is given by:<br />

σ kl = C 0 klmnɛ mn (34)<br />

because of the symmetry of the linear elastic coefficients:<br />

σ kl = C 0 klmnu m,n (35)<br />

σ kl = C 0 klmnG mr,n (x − x ′ )F r (36)<br />

Now, following Weinberger et al [4], for a given volume V , from the definition<br />

of the Dirac delta function, the k component of the force F acting at x ′ can be<br />

described by:<br />

∫<br />

F k δ(x − x ′ )dV (37)<br />

V<br />

Equilibrium requires that the force F be balanced by tractions acting on the<br />

surface S enclosing this volume V , so that:<br />

∫<br />

∫<br />

F k δ(x − x ′ )dV + σ kl n l dS = 0 (38)<br />

V<br />

S<br />

∫<br />

∫<br />

F k δ(x − x ′ )dV + CklmnG 0 mr,n (x − x ′ )n l F r dS = 0 (39)<br />

V<br />

S<br />

Applying Gauss’s theorem:<br />

∫<br />

∫<br />

F k δ(x − x ′ )dV + CklmnG 0 mr,nl (x − x ′ )F r dV = 0 (40)<br />

V<br />

V<br />

∫<br />

[F r δ kr δ(x − x ′ ) + CklmnG 0 mr,nl (x − x ′ )F r ]dV = 0 (41)<br />

from which:<br />

V<br />

−C 0 klmnG mr,nl (x − x ′ ) = δ kr δ(x − x ′ ) (42)<br />

For an unbounded isotropic matrix, Weinberger et al [4] derive that Green<br />

tensor function for x ′ = 0 in their Eq. (1.79) as:<br />

G mr (x) =<br />

1<br />

(<br />

(3 − 4ν)δ mr + x )<br />

mx r<br />

16πµ(1 − ν)x<br />

x 2<br />

(43)


which clearly shows that G mr vanishes as x = |x| approaches infinity, and where<br />

µ, ν are the shear modulus and Poisson’s ratio of the matrix material. Following<br />

Gubernatis [5], the solution for the displacement is decomposed into:<br />

u i (x) = u 0 i (x) + u 1 i (x) (44)<br />

where u 0 i (x) is any solution of the homogeneous equation:<br />

C 0 klmnɛ mn,l = 0 (45)<br />

and u 1 i (x) satisfies the inhomogeous equation Eq. 29. The solution for the<br />

displacement u 1 i (x) is given by applying Stakgold [7] Eq. (5.100) to Eqs. 29<br />

and 42 to find the displacement response at x due to the equilibrating ”force”<br />

at x ′ given by the right hand side of Eq. 29:<br />

∫<br />

u 1 i (x) = dx ′ G ik (x − x ′ )σkl,l ∗ ′(x′ ) (46)<br />

or,<br />

∫<br />

u 1 i (x) =<br />

σ ∗ kl,l ′(x′ ) =<br />

V<br />

V<br />

∂<br />

∂x ′ l<br />

([C] klmn (x ′ )ɛ mn (x ′ )) (47)<br />

dx ′ G ik (x − x ′ ) ∂<br />

∂x ′ ([C] klmn (x ′ )ɛ mn (x ′ )) (48)<br />

l<br />

so that the total displacement is given by:<br />

∫<br />

u i (x) = u 0 i (x) + dx ′ G ik (x − x ′ ) ∂ (<br />

)<br />

V<br />

∂x ′ [C] klmn (x ′ )ɛ mn (x ′ )<br />

l<br />

(49)<br />

Integrating by parts:<br />

∫<br />

u i (x) = u 0 i (x) + dx ′ ∂<br />

(<br />

)<br />

V ∂x ′ G ik (x − x ′ )[C] klmn (x ′ )ɛ mn (x ′ ) (50)<br />

l<br />

∫<br />

− dx ′ ∂ (<br />

Gik (x − x ′ ) ) [C] klmn (x ′ )ɛ mn (x ′ ) (51)<br />

V<br />

∂x ′ l<br />

Applying the divergence theorem and using the fact that the Green function<br />

vanishes on the boundary at infinity:<br />

∫<br />

u i (x) = u 0 i (x) − dx ′ ∂ (<br />

Gik (x − x ′ ) ) [C] klmn (x ′ )ɛ mn (x ′ ) (52)<br />

Since:<br />

∂<br />

∂x ′ l<br />

V<br />

∂x ′ l<br />

G ik (x − x ′ ) = − ∂<br />

∂x l<br />

G ik (x − x ′ ) (53)<br />

The displacement solution becomes:<br />

∫<br />

u i (x) = u 0 i (x) + dx ′ ∂ (<br />

Gik (x − x ′ ) ) [C] klmn (x ′ )ɛ mn (x ′ ) (54)<br />

∂x l<br />

V


or, defining:<br />

we can write:<br />

∫<br />

u i (x) = u 0 i (x) +<br />

Now calculate the strains.<br />

G ik,l (x − x ′ ) =<br />

V<br />

∂<br />

∂x l<br />

(<br />

Gik (x − x ′ ) ) (55)<br />

dx ′( G ik,l (x − x ′ ) ) [C] klmn (x ′ )ɛ mn (x ′ ) (56)<br />

ɛ ij = 1 2 (u i,j + u j,i ) (57)<br />

Since only the Green tensor function depends on x:<br />

∫<br />

u i,j = u 0 i,j + dx ′( G ik,lj (x − x ′ ) ) [C] klmn (x ′ )ɛ mn (x ′ ) (58)<br />

V<br />

∫<br />

u j,i = u 0 j,i + dx ′( G jk,li (x − x ′ ) ) [C] klmn (x ′ )ɛ mn (x ′ ) (59)<br />

V<br />

∫<br />

ɛ ij (x) = ɛ 0 ij(x) + dx ′ G ijkl (x − x ′ )[C] lkmn (x ′ )ɛ mn (x ′ ) (60)<br />

where:<br />

V<br />

G ijkl (x) = 1 2 (G ik,lj(x) + G jk,li (x)) (61)<br />

Since the moduli differ only over the inclusion W , Eq. 60 reduces to:<br />

∫<br />

ɛ ij (x) = ɛ 0 ij(x) + dx ′ G ijkl (x − x ′ )[C] lkmn (x ′ )ɛ mn (x ′ ) (62)<br />

W<br />

Markov’s [6] treatment of this integral equation begins here. Markov proceeds<br />

as follows. ”If ɛ 0 is constant, then from Eq. 62 the strain within the inhomogeneity<br />

will also be constant provided that the moduli differ but by constant<br />

values and the following tensor field is constant within the inhomogeneity”:<br />

∫<br />

P ijkl (x) = − G ijkl (x − x ′ )dx ′ (63)<br />

So that:<br />

But:<br />

so that:<br />

W<br />

ɛ ij = ɛ 0 ij − P ijkl [C] lkmn ɛ mn (64)<br />

I ijmn ɛ mn = ɛ ij (65)<br />

(I ijmn + P ijkl [C] lkmn )ɛ mn = ɛ 0 ij (66)


Therefore, letting:<br />

we have:<br />

Or, with:<br />

B ijmn = I ijmn + P ijkl [C] lkmn (67)<br />

B ijmn ɛ mn = ɛ 0 ij (68)<br />

B = I + P : [C] (69)<br />

Markov [6] identifies a fourth-order tensor A given by:<br />

A = (I + P : [C]) −1 = B −1 (70)<br />

Finally, Markov [6] connects the constant strain ɛ 0 with the constant strain in<br />

the ellipsoidal cavity:<br />

ɛ = A : ɛ 0 (71)<br />

In the next two sections, it will be shown that the tensor P is constant, which<br />

will allow us to derive the tensor A.<br />

6 Role of Eshelby’s tensor in integral equation<br />

for strains<br />

In order to make further progress from the integral equation above toward <strong>Kroner</strong>’s<br />

cubic, we need to derive an important connection between the tensor P<br />

and another tensor T related to Eshelby’s tensor, as described below.<br />

For this purpose, we will follow Weinberger et al [4]. They introduce the<br />

Eshelby tensor S (See [3] Eq. (3.3) p. 104) that relates the ”constrained strain<br />

inside the inclusion to the (eigen)strain in the inclusion in response to zero<br />

stress”. The Eshelby tensor can be written as:<br />

Weinberger et al [4] Eq. (2.13) write:<br />

where by letting:<br />

S = T : C (72)<br />

S ijmn = − 1 2 C lkmn(D iklj + D jkli ) (73)<br />

T ijkl = − 1 2 (D iklj + D jkli ) (74)<br />

we can relate T to P, and where, for an ellipsoidal inclusion, Weinberger et al<br />

[4] Eq. (2.16) give:<br />

∫<br />

D ijkl (x) = G ij,kl (x − x ′ )dV (x ′ ) (75)<br />

V 0<br />

From Eqs. 61, 63, 73, 74, and 75, it is clear that:<br />

T = P (76)


7 Symmetry properties and model of tensor D<br />

In order to show that tensor T (and therefore that P) is a constant, we need<br />

to establish some symmetry properties of tensor D. Weinberger et al [4] Eq.<br />

(2.61) calculate the D ijkl tensor when x lies inside the ellipsoidal inclusion as:<br />

D ijkl = − abc<br />

4π<br />

where, from [4] Eq. (1.70):<br />

∫ π ∫ 2π<br />

0<br />

0<br />

(zz) −1<br />

ij z sinΦ<br />

kz l dΘdΦ (77)<br />

β3 (zz) −1<br />

ij = 1 (<br />

δ ij − λ )<br />

1 + µ 1<br />

z i z j<br />

µ 1 λ 1 + 2µ 1<br />

and from [4] Eqs. (2.17), (2.42), (2.43), (2.44):<br />

For a unit sphere, from [4] Eq. (2.46):<br />

D ijkl = − 1<br />

4πµ 1<br />

∫ π<br />

0<br />

(78)<br />

z 1 = sinΦcosΘ (79)<br />

z 2 = sinΦsinΘ (80)<br />

∫ 2π<br />

0<br />

z 3 = cosΦ (81)<br />

(δ ij + 2κ 1 z i z j ) z k z l sinΦdΘdΦ (82)<br />

where, given Lame’s constant λ 1 , shear modulus µ 1 , Poisson’s ratio ν 1 , and bulk<br />

modulus k 1 for the isotropic matrix:<br />

κ 1 = − 1 1<br />

= − 1 λ 1 + µ 1<br />

(83)<br />

4 1 − ν 1 2 λ 1 + 2µ 1<br />

ν 1 = 3k 1 − 2µ 1<br />

2(3k 1 + µ 1 )<br />

κ 1 = (− 1 2 ) (3k 1 + µ 1 )<br />

(3k 1 + 4µ 1 )<br />

From Eq. 82, it is clear that D satisfies the following minor symmetries:<br />

(84)<br />

(85)<br />

D ijkl = D jikl = D ijlk (86)<br />

However, it is also possible to show that D satisfies the major symmetry:<br />

This is shown in Section 14 so that we can write:<br />

D ijkl = − 1 (1 + 2 3µ 1 5 κ 1)δ ij δ kl − 2<br />

15<br />

D ijkl = D klij (87)<br />

or, in terms of the isotropic basis tensors:<br />

D = − 1 µ 1<br />

[<br />

(1 + 2 3 κ 1)I ′ + 4<br />

15 κ 1I ′′ ]<br />

κ 1<br />

µ 1<br />

(δ ik δ jl + δ il δ jk ) (88)<br />

(89)


We now return to the tensor T defined in Eq. 74:<br />

D iklj = γδ ik δ lj + ω(δ il δ kj + δ ij δ kl ) (90)<br />

D jkli = γδ jk δ li + ω(δ jl δ ki + δ ij δ kl ) (91)<br />

T ijkl = − 1 2 (D iklj + D jkli ) (92)<br />

T ijkl = 1 ( 2<br />

3µ 1 5 κ 1δ ij δ kl + 1 2 (1 + 4 )<br />

5 κ 1)(δ ik δ lj + δ il δ kj )<br />

(93)<br />

or, in terms of the isotropic basis tensors:<br />

Using Eq. 85:<br />

T = t ′ I ′ + t ′′ I ′′ (94)<br />

t ′ = 1 + 2κ 1<br />

3µ 1<br />

(95)<br />

t ′′ = 1 + 4 5 κ 1<br />

3µ 1<br />

(96)<br />

t ′ =<br />

1<br />

3k 1 + 4µ 1<br />

(97)<br />

t ′′ = 3<br />

5µ 1<br />

k 1 + 2µ 1<br />

3k 1 + 4µ 1<br />

(98)<br />

Thus, the tensor T and therefore tensor P is a constant.<br />

8 Derivation of tensor A for isotropic matrix<br />

and isotropic inhomogeneity<br />

Since P = T , we have from the above:<br />

P = p 1 I ′ + p 2 I ′′ (99)<br />

p 1 = t ′ (100)<br />

p 2 = t ′′ (101)<br />

For the special case where both the matrix and the inhomogeneity are isotropic,<br />

with subscript ”1” for the isotropic matrix and subscript ”2” for the isotropic<br />

inhomogeneity, the corresponding elastic moduli are written as:<br />

so that:<br />

C 1 = 3k 1 I ′ + 2µ 1 I ′′ (102)<br />

C 2 = 3k 2 I ′ + 2µ 2 I ′′ (103)<br />

P : [C] = 3[k]p 1 I ′ + 2[µ]p 2 I ′′ (104)


where:<br />

[k] = k 2 − k 1 (105)<br />

[µ] = µ 2 − µ 1 (106)<br />

We now proceed to derive the tensor A in Eq. 70. Putting Eq. 104 into Eq. 70<br />

and using Eq. 11 we have:<br />

I + P : [C] ≡ B = b ′ I ′ + b ′′ I ′′ (107)<br />

b ′ = 1 + 3[k]p 1 (108)<br />

b ′′ = 1 + 2[µ]p 2 (109)<br />

Note, that because of the properties of fourth-order tensors described in Section<br />

3:<br />

which can be written as:<br />

A = B −1 = 1 b ′ I′ + 1<br />

b ′′ I′′ (110)<br />

A = a ′ I ′ + a ′′ I ′′ (111)<br />

a ′ = 1 b ′ = k 1<br />

k 1 + 3[k]k 1 p 1<br />

(112)<br />

a ′′ = 1<br />

b ′′ = µ 1<br />

µ 1 + 2[µ]µ 1 p 2<br />

(113)<br />

Comparing Eqs. 112 and 113 with Markov’s [6] Eq. (4.58), we find that:<br />

3k 1 p 1 = α 1 (114)<br />

2µ 1 p 2 = β 1 (115)<br />

Note that Markov has an error in his Eq. (4.58) for the basis tensor I ′′<br />

ijkl which<br />

he writes as:<br />

which should be corrected to read:<br />

1<br />

2 (δ ijδ kl + δ il δ jk − 2 3 δ ijδ kl ) (116)<br />

I ′′<br />

ijkl = 1 2 (δ ikδ jl + δ il δ jk − 2 3 δ ijδ kl ) (117)<br />

This concludes the derivation of tensor A for isotropic matrix and isotropic<br />

inhomogeneity.<br />

9 Self consistent scheme for cubic polycrystals<br />

The goal is to derive an equation (<strong>Kroner</strong>’s cubic) that relates the shear modulus<br />

of the aggregate of cubic crystals to the elastic coefficients of a cubic crystal.


Markov [6] assumes that ”...this aggregate (polycrystal) is an assembly of<br />

monocrystals, homogeneous grains with one and the same elastic properties,<br />

defined by the elastic tensor C. The crystallographic axes of each grain vary.<br />

Hence the tensor C is rotated in a complicated manner when one moves across<br />

the solid and exactly this is what makes the polycrystal heterogeneous”. To<br />

keep things simple, Markov assumes that ”there exist no preferable orientations<br />

of grains, i.e. of the crystallographic axes. Hence there is no texture presented<br />

so that the polycrystal is macroscopically isotropic, with a tensor of effective<br />

moduli”:<br />

C ∗ = 3k ∗ I ′ + 2µ ∗ I ′′ (118)<br />

where I ′ and I ′′ are given by Eqs. 11, 4, and 10. Markov continues. ”Our aim<br />

is to develop a certain simple approximate scheme of self-consistent type for<br />

evaluating C ∗ by means of the given elastic tensor C for a single grain. Imagine<br />

each grain is a sphere, immersed into an unbounded matrix with the effective,<br />

but yet unknown, [assumed isotropic] properties C ∗ . Fix one of the grains, W .<br />

According to Eq. 71 the strain within such a grain is constant”:<br />

ɛ gr = A(C, C ∗ ) : ¯ɛ (119)<br />

where ¯ɛ is the ”prescribed macrostrain tensor, applied to the polycrytalline representative<br />

volume element (RVE)”, interpreted here as the strain ɛ 0 resulting<br />

when the moduli C 1 and C 2 match. Moreover, in virtue of Eqs. 70 and 99,<br />

Markov claims that given:<br />

so that:<br />

A(C, C ∗ ) ≡ A ∗ (120)<br />

A ∗ = B ∗(−1) (121)<br />

B ∗ = [I + P ∗ : (C − C ∗ )] (122)<br />

P ∗ = p ∗ 1I ′ + p ∗ 2I ′′ (123)<br />

p ∗ 1<br />

1 =<br />

3k ∗ + 4µ ∗ (124)<br />

p ∗ 2 = 3<br />

5µ ∗ k ∗ + 2µ ∗<br />

3k ∗ + 4µ ∗ (125)<br />

It appears that Markov’s argument is that Eq. 99 applies to both the isotropic<br />

matrix and the approximately isotropic aggregate of cubic (anisotropic) crystals<br />

in perfectly random orientations. The difference between these two configurations<br />

is that in the case of the isotropic aggregate, the elastic modulus C ∗ is<br />

unknown.<br />

Markov is expanding the role of tensor A to apply to the case where the<br />

inhomogeneity is not isotropic. Note that the tensor P was derived from the<br />

Green’s function for the isotropic matrix alone. The constitutive tensor C is now<br />

allowed to be anisotropic. This is consistent with the previous development.


Markov now describes the critical condition to be satisfied for determining<br />

this aggregate effective elastic modulus: ”Let us now average Eq. 119 with<br />

respect to all possible crystallographic orientations of the axes of the grain.<br />

This operation will be denoted by 〈·〉 Ω<br />

”. Then:<br />

”The key observation now is the identity”:<br />

〈ɛ gr 〉 Ω<br />

= 〈A ∗ 〉 Ω<br />

: ¯ɛ (126)<br />

〈ɛ gr 〉 Ω<br />

= ¯ɛ (127)<br />

”i.e. the strain averaged over all crystallographic orientations in a grain equals<br />

the macrostrain. Inserting Eq. 126 into Eq. 127:<br />

〈A ∗ 〉 Ω<br />

= I (128)<br />

through which the unknown effective tensor C ∗ can be found”.<br />

10 Elastic moduli of a cubic polycrystal<br />

For a cubic polycrystal the calculation of the unknown effective tensor leads to<br />

<strong>Kroner</strong>’s cubic for the effective shear modulus.<br />

The tensor of the elastic cubic crystal moduli C in this case is given by<br />

Markov as:<br />

C = 3kI ′ + 2µ 2 I ′′ + 2(µ 1 − µ 2 )O h (129)<br />

O h = e 4 1 + e 4 2 + e 4 3 (130)<br />

Markov does not define e 4 i , saying only that O h is the ”basic fourth-rank tensor<br />

with cubic symmetry (whose axes are along the orthonormal crystallographic<br />

basis e i , i = 1, 2, 3)”. Turning to [8], Bertram defines e 4 i as (no sum on i):<br />

e 4 i = e i ⊗ e i ⊗ e i ⊗ e i (131)<br />

To simplify the calculation of the shear modulus, Markov introduces ”three<br />

basic fourth-rank tensors with cubic symmetry”:<br />

Σ 1 = I ′ (132)<br />

Σ 2 = I ′ + I ′′ − O h (133)<br />

Σ 3 = O h − I ′ (134)<br />

Markov claims that these three tensors are orthogonal, which is true after applying<br />

the correction in Eq. 117; that is, for each i, j = 1, 2, 3, i ≠ j:<br />

This correction also allows the relation:<br />

Σ i : Σ j = 0 (135)<br />

Σ 1 + Σ 2 + Σ 3 = I (136)


Finally, this correction also confirms that for each i = 1, 2, 3, with no implied<br />

sum on i:<br />

Σ i : Σ i = Σ i (137)<br />

With respect to this basis, Markov’s elastic coefficients take the form:<br />

where:<br />

C = 3K ′ Σ 1 + 2µ 2 Σ 2 + 2µ 1 Σ 3 (138)<br />

K ′ = k + 2 3 (µ 1 − µ 2 ) (139)<br />

To interpret Markov’s formulation for the elastic coefficients of the cubic crystal<br />

in Eq. 129, we introduce the elastic coefficients of the cubic crystal in the form<br />

described by Lubarda [9] in Eq. (3.1) (Note Lubarda’s sign error: +2c 44 should<br />

be −2c 44 ):<br />

C ijkl = c 12 δ ij δ kl + 2c 44 I ijkl + (c 11 − c 12 − 2c 44 )Âijkl (140)<br />

which is equivalent in Markov’s notation to:<br />

or,<br />

C = 3c 12 I ′ + 2c 44 (I ′ + I ′′ ) + (c 11 − c 12 − 2c 44 )O h (141)<br />

C = (3c 12 + 2c 44 )I ′ + 2c 44 (I ′′ ) + (c 11 − c 12 − 2c 44 )O h (142)<br />

but, in terms of Markov’s basis:<br />

C = (c 11 + 2c 12 )Σ 1 + 2c 44 Σ 2 + (c 11 − c 12 )Σ 3 (143)<br />

Comparing Eqs. 138 with 143 shows that:<br />

In the sequel, the symbolic notation:<br />

will be useful.<br />

K ′ = 1 3 (c 11 + 2c 12 ) (144)<br />

µ 1 = 1 2 (c 11 − c 12 ) (145)<br />

µ 2 = c 44 (146)<br />

C = (α, β, γ) ⇔ C = αΣ 1 + βΣ 2 + γΣ 3 (147)


11 Rotational averages<br />

Before proceeding with the averaging process of the constitutive coefficients<br />

and, in particular, of the tensor A, it will be useful to describe the details of<br />

calculating rotational averages.<br />

According to Eqs. (2) and (3) in Andrews [10], the rotational average of a<br />

fourth order tensor C is given by:<br />

〈C〉 i1i 2i 3i 4<br />

= Λ i1i 2i 3i 4;λ 1λ 2λ 3λ 4<br />

C λ1λ 2λ 3λ 4<br />

(148)<br />

and where Λ is defined by Eq. (19) in Andrews [10] as:<br />

⎛<br />

Λ i1i 2i 3i 4;λ 1λ 2λ 3λ 4<br />

= 1 ( )<br />

δi1i<br />

30<br />

2<br />

δ i3i 4<br />

δ i1i 3<br />

δ i2i 4<br />

δ i1i 4<br />

δ i2i 3<br />

⎝<br />

4 −1 −1<br />

−1 4 −1<br />

−1 −1 4<br />

⎞ ⎛<br />

⎠ ⎝<br />

Let’s begin with finding the rotational average of tensor I ′ .<br />

In the following, the subscripts h and Ω are understood. Following Andrews,<br />

we have:<br />

where we must have the correspondence:<br />

and where<br />

〈I ′ 〉 ijkl<br />

= Λ ijkl;αβγω I ′ αβγω (150)<br />

(i 1 , i 2 , i 3 , i 4 ) ⇔ (i, j, k, l) (151)<br />

(λ 1 , λ 2 , λ 3 , λ 4 ) ⇔ (α, β, γ, ω) (152)<br />

Carrying out the above calculations results in:<br />

Similarly, with the correspondence:<br />

we find:<br />

and, with the correspondence:<br />

I ′ αβγω = 1 3 δ αβδ γω (153)<br />

〈I ′ 〉 = I ′ (154)<br />

(i 1 , i 2 , i 3 , i 4 ) ⇔ (i, k, j, l) (155)<br />

(λ 1 , λ 2 , λ 3 , λ 4 ) ⇔ (α, γ, β, ω) (156)<br />

〈δ ik δ jl 〉 = δ ik δ jl (157)<br />

(i 1 , i 2 , i 3 , i 4 ) ⇔ (i, l, j, k) (158)<br />

(λ 1 , λ 2 , λ 3 , λ 4 ) ⇔ (α, ω, β, γ) (159)<br />

δ λ1λ 2<br />

δ λ3λ 4<br />

δ λ1λ 3<br />

δ λ2λ 4<br />

δ λ1λ 4<br />

δ λ2λ 3<br />

⎞<br />

⎠(149)


we find:<br />

〈δ il δ jk 〉 = δ il δ jk (160)<br />

so that:<br />

〈I〉 = I (161)<br />

and, therefore:<br />

〈I ′′ 〉 = I ′′ (162)<br />

Let’s now find the rotational average of tensor O h . Following Andrews, we have:<br />

From Eqs. 130 and 131:<br />

So that:<br />

〈O〉 ijkl<br />

= Λ ijkl;αβγω O αβγω (163)<br />

e 4 i = δ αi e α ⊗ δ βi e β ⊗ δ γi e γ ⊗ δ ωi e ω (164)<br />

O αβγω =<br />

3∑<br />

(δ αi δ βi δ γi δ ωi ) (165)<br />

i=1<br />

and from Eqs. 149, 163, and 165 we obtain:<br />

Now:<br />

so that:<br />

〈O ijkl 〉 = 1<br />

30 (δ αβδ γω A ijkl + δ αγ δ βω B ijkl + δ αω δ βγ C ijkl )O αβγω (166)<br />

A ijkl = 4δ ij δ kl − δ ik δ jl − δ il δ jk (167)<br />

B ijkl = −δ ij δ kl + 4δ ik δ jl − δ il δ jk (168)<br />

C ijkl = −δ ij δ kl − δ ik δ jl + 4δ il δ jk (169)<br />

δ αβ δ γω O αβγω = 3 (170)<br />

δ αγ δ βω O αβγω = 3 (171)<br />

δ αω δ βγ O αβγω = 3 (172)<br />

〈O ijkl 〉 = 1<br />

10 (A ijkl + B ijkl + C ijkl ) = 1 5 (δ ijδ kl + δ ik δ jl + δ il δ jk ) (173)<br />

We now define H as:<br />

so that:<br />

H ijkl = δ ij δ kl + δ ik δ jl + δ il δ jk (174)<br />

〈O〉 = 1 5 H (175)


Note that from Eqs. 10, 11, and 174:<br />

H = 5I ′ + 2I ′′ (176)<br />

We now prove the following identities which we will need in the sequel:<br />

From 133, 154, 162,173, and 176:<br />

〈Σ 2 〉 = 3 5 I′′ (177)<br />

〈Σ 3 〉 = 2 5 I′′ (178)<br />

〈Σ 2 〉 = 〈I ′ 〉 + 〈I”〉 − 〈O〉 (179)<br />

〈Σ 2 〉 = I ′ + I” − 1 5 H (180)<br />

〈Σ 2 〉 = I ′ + I” − 1 5 (5I′ + 2I”) (181)<br />

〈Σ 2 〉 = 3 I” (182)<br />

5<br />

Eq. 178 is shown is a similar fashion. From Eqs. 133 and 134:<br />

so that:<br />

Similarly, from Eq. 134:<br />

Σ 2 + Σ 3 = I ′′ (183)<br />

〈Σ 1 〉 = 〈I ′ 〉 = Σ 1 (184)<br />

〈Σ 2 〉 = 3 5 I” = 3 5 (Σ 2 + Σ 3 ) (185)<br />

〈Σ 3 〉 = 〈O〉 − 〈I ′ 〉 (186)<br />

〈Σ 3 〉 = 1 5 H − I′ (187)<br />

〈Σ 3 〉 = 1 5 (5I′ + 2I ′′ ) − I ′ (188)<br />

〈Σ 3 〉 = 2 5 I′′ (189)<br />

〈Σ 3 〉 = 2 5 (Σ 2 + Σ 3 ) (190)<br />

12 Derivation of tensor A for anisotropic inhomogeneity<br />

and <strong>Kroner</strong>’s cubic<br />

For the tensor C of a single grain, see Eq. 138, we have:<br />

C = (α, β, γ), α = 3K ′ , β = 2µ 2 , γ = 2µ 1 (191)


Due to the orthogonal properties of the Σ ′ is, inversion and multiplication take<br />

the forms:<br />

C −1 = ( 1 α , 1 β , 1 γ ) (192)<br />

With C ′ = (α ′ , β ′ , γ ′ ) and C ′′ = (α ′′ , β ′′ , γ ′′ ):<br />

C ′ : C ′′ = (α ′ α ′′ , β ′ β ′′ , γ ′ γ ′′ ) (193)<br />

The most important formula, however, concerns the tensors of the form given<br />

by Eq. 147, averaged over all orientations, using Eqs. 184, 185, 190:<br />

From Eqs. 122 and 191 the quantity:<br />

has the form:<br />

〈C〉 Ω<br />

= 〈(α, β, γ)〉 Ω<br />

= (α, ¯β, ¯β) (194)<br />

¯β = 1 (3β + 2γ) (195)<br />

5<br />

B ∗ = I + P ∗ : (C − C ∗ ) (196)<br />

B ∗ = (1, 1, 1) + (p ∗ 1, p ∗ 2, p ∗ 2) · [(3K ′ , 2µ 2 , 2µ 1 ) − (3k ∗ , 2µ ∗ , 2µ ∗ )] (197)<br />

which reduces to:<br />

B ∗ = (1 + 3p ∗ 1(K ′ − k ∗ ), 1 + 2p ∗ 2(µ 2 − µ ∗ ), 1 + 2p ∗ 2(µ 1 − µ ∗ )) (198)<br />

But Eq. 198 is the inverse of Eq. 120, so that:<br />

A ∗ 1<br />

= (<br />

1 + 3p ∗ 1 (K′ − k ∗ ) , 1<br />

1 + 2p ∗ 2 (µ 2 − µ ∗ ) , 1<br />

1 + 2p ∗ 2 (µ 1 − µ ∗ ) ) (199)<br />

whose average follows Eqs. 194 and 154:<br />

〈A ∗ 〉 Ω<br />

= (α ∗ , β ∗ , β ∗ ) (200)<br />

α ∗ 1<br />

=<br />

1 + 3p ∗ 1 (K′ − k ∗ )<br />

(201)<br />

β ∗ = 3 1<br />

5 1 + 2p ∗ 2 (µ 2 − µ ∗ ) + 2 1<br />

5 1 + 2p ∗ 2 (µ 1 − µ ∗ )<br />

(202)<br />

Finally, the average of A ∗ must satisfy Eq. 128 subject to Eq. 136, so that:<br />

α ∗ Σ 1 + β ∗ (Σ 2 + Σ 3 ) = I = I ′ + I ′′ (203)<br />

Σ 2 + Σ 3 = I ′′ (204)<br />

α ∗ I ′ + β ∗ I ′′ = I ′ + I ′′ (205)<br />

(α ∗ − 1)I ′ + (β ∗ − 1)I ′′ = 0 (206)<br />

(207)


Since I ′ and I ′′ are orthogonal:<br />

α ∗ = 1 (208)<br />

β ∗ = 1 (209)<br />

from which:<br />

k ∗ = K ′ (210)<br />

and substituting Eq. 125 into Eq. 202 and applying Eq. 209 gives a cubic<br />

equation for the aggregate shear modulus µ ∗ in the form:<br />

(µ ∗ ) 3 + a(µ ∗ ) 2 − bµ ∗ − c = 0 (211)<br />

a = 9k∗ + 4µ 1<br />

8<br />

b = 3µ 2(k ∗ + 4µ 1 )<br />

8<br />

c = 3k∗ µ 1 µ 2<br />

4<br />

Using Eq. 210, 144, 145, and 146, the above coefficients of the cubic become:<br />

(212)<br />

(213)<br />

(214)<br />

a = 1 8 (5c 11 + 4c 12 ) (215)<br />

b = 1 8 c 44(7c 11 − 4c 12 ) (216)<br />

c = 1 8 c 44(c 11 + 2c 12 )(c 11 − c 12 ) (217)<br />

<strong>Kroner</strong>’s [2] Eq. (22) shows the cubic for the aggregate shear modulus G as:<br />

Converting to Voigt notation:<br />

G 3 + ᾱG 2 + ¯βG + ¯γ = 0 (218)<br />

ᾱ = (3 ¯K + 4¯ν)<br />

8<br />

¯β = −( ¯K + 12¯ν) ¯µ 8<br />

(219)<br />

(220)<br />

¯γ = − ¯K ¯µ ¯ν 4<br />

(221)<br />

¯K = c 1111 + 2c 1122 (222)<br />

¯µ = c 1212 (223)<br />

¯ν = (c 1111 − c 1122 )<br />

2<br />

(224)<br />

¯K = c 11 + 2c 12 (225)<br />

¯µ = c 66 (226)<br />

¯ν = (c 11 − c 12 )<br />

2<br />

(227)


But because of cubic symmetry, c 66 = c 44 , so that <strong>Kroner</strong>’s coefficients ᾱ, ¯β, ¯γ<br />

are consistent with the above coefficients a, b, c, namely:<br />

ᾱ = 1 8 (5c 11 + 4c 12 ) = a (228)<br />

¯β = − c 44<br />

8 (7c 11 − 4c 12 ) = −b (229)<br />

¯γ = − c 44<br />

8 (c 11 + 2c 12 )(c 11 − c 12 ) = −c (230)<br />

as required, thus completing the derivation of <strong>Kroner</strong>’s cubic.<br />

13 Comparison with experimental data<br />

Hashin and Shtrikman [11], [12] have derived bounds for the isotropic shear<br />

modulus of aggregates of cubic crystals. Gschneider [13] provides experimental<br />

data, and Simmons [14] provides cubic crystal elastic coefficients in Mbar. Sisodia<br />

[15] compares some experimental data with calculated shear modulus and<br />

reports the formulation of the Hashin-Shtrikman bounds, G ∗ 1 and G ∗ 2 in terms<br />

of the single crystal cubic coefficients c ij as:<br />

where<br />

(<br />

) −1<br />

G ∗ 5<br />

1 = G 1 + 3 − 4β 1 (231)<br />

G 2 − G 1<br />

(<br />

G ∗ 2 = G 2 + 2<br />

5<br />

G 1 − G 2<br />

− 6β 2<br />

) −1<br />

(232)<br />

β 1 = − 3 K + 2G 1<br />

5 G 1 (3K + 4G 1 )<br />

β 2 = − 3 K + 2G 2<br />

5 G 2 (3K + 4G 2 )<br />

An anisotropy index à is defined by:<br />

K = c 11 + 2c 12<br />

3<br />

(233)<br />

(234)<br />

(235)<br />

G 1 = c 11 − c 12<br />

2<br />

(236)<br />

G 2 = c 44 (237)<br />

à = G 2<br />

G 1<br />

= 2c 44<br />

c 11 − c 12<br />

(238)<br />

which reduces to unity for isotropic material. Table 1 shows the cubic coefficients<br />

and the anisotropy index for several materials. Notice that Tungsten (W) and<br />

Molybdenum (Mo) come closest to being isotropic.<br />

To compare the shear modulus determined by <strong>Kroner</strong>’s cubic with experimental<br />

values, we use a simple iterative Newton’s method, beginning with a trial


value at the arithmetic average of the Hashin-Shtrikman’s bounds, as described<br />

in Section 15. Table 1 shows these bounds and compares this calculated value<br />

with the experimental value. Notice that for elements which are not in Group<br />

IV-A, the error is under 20%, while for Group IV-A, the error varies from 40%<br />

to 74%.<br />

Sisodia [15] mentions that if the anisotropic index à > 1, then G∗ 2 > G ∗ 1<br />

and vice versa. Notice also from Table 1, for those (most) elements where the<br />

anisotropy index exceeds unity, it is the case that G ∗ 2 > G ∗ 1. For Molybdenum<br />

and Tungsten which are practically isotropic, the bounds coincide. Finally,<br />

Niobium is the only element with à < 1, and there G∗ 2 < G ∗ 1.<br />

Units:<br />

kg f<br />

cm 2<br />

Gschneider’s experimental values are given in units of × 10 6 . This is<br />

converted to GPa by multiplying by gravitational acceleration as 9.80665 m s<br />

, so 2<br />

that Gschneider’s units are converted as follows:<br />

kg f<br />

cm 2 × 106 × 10 4 cm2<br />

m 2 × 9.80665 m s 2 × N<br />

kg m s 2<br />

× P a<br />

N<br />

m 2<br />

× GP a<br />

10 9 P a<br />

(239)<br />

and 1 kg f<br />

cm<br />

×10 6 then corresponds to 98.0665 GPa. For example, with Aluminum,<br />

2<br />

Schneider’s value of 0.271 kg f<br />

cm<br />

× 10 6 corresponds to 26.6 GPa.<br />

2


Table 1: Calculated (Gk) vs. Experimental (Gx) Shear Modulus (GPa)<br />

Element Group c11 c12 c44 G1* G2* Gk Gx Error Anisotropy<br />

Ag11650 I-B 124 93.4 46.1 29 30.9 30.2 28.6 0.06 3.01<br />

Ag11649 I-B 120 89.7 43.7 28 29.6 29 28.6 0.01 2.88<br />

Ag11651 I-B 121 89.9 42.6 28 29.4 28.9 28.6 0.01 2.74<br />

Ag11678 I-B 122 90.7 45.4 29.2 30.8 30.2 28.6 0.06 2.90<br />

Au10555 I-B 187 155 42.9 28.6 29.9 29.4 27.6 0.07 2.68<br />

Cu10380 I-B 170 115 61 44.1 45.2 44.8 45.1 -0.01 2.22<br />

Cu10381 I-B 176 128 75.2 46 49.4 48.2 45.1 0.07 3.13<br />

Cu10382 I-B 170 124 64.5 41.9 44.2 43.3 45.1 -0.04 2.80<br />

Cu10383 I-B 168 121 75.4 46 49.4 48.2 45.1 0.07 3.21<br />

Cu10384 I-B 170 123 75.3 46 49.5 48.2 45.1 0.07 3.20<br />

Cu10385 I-B 171 124 75.5 46 49.5 48.2 45.1 0.07 3.21<br />

Cu10423 I-B 168 121 75.7 46.2 49.6 48.4 45.1 0.07 3.22<br />

Al10037 III-A 106 59 28.5 26.3 26.4 26.3 26.6 -0.01 1.21<br />

Ge10531 IV-A 132 50.9 66.9 54.5 54.8 54.7 39.2 0.40 1.65<br />

Ge10532 IV-A 129 47.9 67 54.7 55 54.9 39.2 0.40 1.65<br />

Ge10538 IV-A 129 48.3 67.1 54.5 54.9 54.8 39.2 0.40 1.66<br />

Pb10681 IV-A 47.7 40.3 14.4 8 9 8.66 5.39 0.61 3.89<br />

Si11644 IV-A 168 66 84 68.6 69 68.9 39.7 0.74 1.65<br />

Si11645 IV-A 166 63.9 79.6 66.5 66.7 66.6 39.7 0.68 1.56<br />

Nb11140 V-B 246 139 29.3 37.7 37.2 37.4 37.5 0.00 0.55<br />

Ta11928 V-B 260 154 82.6 69 69.3 69.2 68.6 0.01 1.56<br />

V12078 V-B 196 133 67 49.3 50.3 49.9 46.5 0.07 2.13<br />

W12020 VI-B 513 206 153 153 153 153 153 0.00 1.00<br />

Mo10974 VI-B 470 168 107 123 123 123 116 0.06 0.71<br />

Mo10995 VI-B 441 172 122 127 127 127 116 0.09 0.91<br />

Fe10650 VIII 231 135 116 80.8 83.4 82.4 81.5 0.01 2.42<br />

Ni11110 VIII 248 155 124 82.4 86.1 84.7 75 0.13 2.67<br />

Ni11111 VIII 247 144 124 86.1 88.9 87.8 75 0.17 2.41<br />

Pd11222 VIII 227 176 71.7 46.7 49.2 48.3 51.1 -0.05 2.81<br />

Pt11224 VIII 347 251 76.5 63.5 63.8 63.7 61 0.04 1.59<br />

Th12001 actinide 75.3 48.9 47.8 27.3 30 29 27.9 0.04 3.62


14 Major symmetries of tensor D<br />

Write:<br />

D ijkl = − 1<br />

4πµ (M ijkl + 2κ 1 N ijkl ) (240)<br />

M ijkl =<br />

N ijkl =<br />

∫ π ∫ 2π<br />

0 0<br />

∫ π ∫ 2π<br />

0<br />

0<br />

(δ ij z k z l sinΦ) dΘdΦ (241)<br />

(z i z j z k z l sinΦ) dΘdΦ (242)<br />

It is clear that N satisfies both major and minor symmetries. It remains to<br />

show that M ijkl satisfies major symmetry. Clearly, if i ≠ j, then M ijkl = 0. It<br />

can also be shown that if k ≠ l, then M ijkl = 0. For this purpose, since M ijkl<br />

satisfies minor symmetries, it suffices to show that:<br />

M ij12 = M ij13 = M ij23 = 0 (243)<br />

Putting Eqs. 79,80,81 into Eq. 241 as needed to form M ij12 , M ij13 , M ij23 , we<br />

find that in each case the Θ-integral vanishes.<br />

For M ij12 , the Θ-integral involves:<br />

∫ 2π<br />

0<br />

(cosΘsinΘ)dΘ = 1 4<br />

∫ 2π<br />

For M ij13 , the Θ-integral involves:<br />

∫ 2π<br />

For M ij23 , the Θ-integral involves:<br />

0<br />

∫ 2π<br />

0<br />

0<br />

(sin2Θ)d(2Θ) = (− 1 4 )[cos2Θ]2π 0 = 0 (244)<br />

cosΘdΘ = [sinΘ] 2π<br />

0 = 0 (245)<br />

sinΘdΘ = −[cosΘ] 2π<br />

0 = 0 (246)<br />

From the above, it is clear that the only non-zero values of M ijkl are obtained<br />

only if i = j and k = l. With i = j, M ijkl reduces to the second order quantity:<br />

M kl =<br />

∫ π ∫ 2π<br />

0<br />

0<br />

(z k z l sinΦ) dΘdΦ (247)<br />

With k = l = 1, 2, 3 and putting Eqs. 79,80,81 into Eq. 247 as needed, we find<br />

that:<br />

M 11 =<br />

∫ π ∫ 2π<br />

0<br />

0<br />

sin 2 Φcos 2 ΘsinΦdΘdΦ =<br />

[ Θ<br />

=<br />

2 + 1 ] 2π<br />

4 sin2Θ<br />

0<br />

∫ 2π<br />

0<br />

cos 2 ΘdΘ<br />

∫ π<br />

[<br />

− 1 3 cosΦ(sin2 Φ + 2)<br />

0<br />

sin 3 ΦdΦ (248)<br />

] π<br />

0<br />

= 4 3 π (249)


Similarly:<br />

M 33 =<br />

Thus, for:<br />

∫ π ∫ 2π<br />

M 22 =<br />

∫ 2π<br />

sin 2 ΘdΘ<br />

∫ π<br />

0<br />

0<br />

cos 2 ΦsinΦdΘdΦ = (−1)2π<br />

0 0<br />

0<br />

sin 3 ΦdΦ = 4 3 π (250)<br />

∫ π<br />

cos 2 Φd(cosΦ) = 4 3 π (251)<br />

ijkl = 1111, 1122, 1133, 2211, 2222, 2233, 3311, 3322, 3333 (252)<br />

M ijkl = 4 3 π (253)<br />

These results clearly show that, given all the non-zero terms, M ijkl does have<br />

major symmetry as well as minor symmetry. Now that we have shown that<br />

both N ijkl and M ijkl have both major and minor symmetry, we can construct<br />

the following model for D ijkl which assumes both major and minor symmetry:<br />

D ijkl = γδ ij δ kl + ω(δ ik δ jl + δ il δ jk ) (254)<br />

where γ is given by any member of the list in Eq. 252, say 1122 for which:<br />

=<br />

∫ π ∫ 2π<br />

0<br />

0<br />

N 1122 =<br />

∫ π ∫ 2π<br />

0<br />

0<br />

(<br />

z<br />

2<br />

1 z 2 2sinΦ ) dΘdΦ (255)<br />

(<br />

sin 2 Φcos 2 Θsin 2 Φsin 2 ΘsinΦ ) dΘdΦ (256)<br />

=<br />

∫ 2π<br />

0<br />

cos 2 Θsin 2 ΘdΘ<br />

∫ π<br />

0<br />

∫ π<br />

( ∫ 1 2π<br />

) ( )<br />

4<br />

= sin 2 (2Θ)d(2Θ) sin 3 ΦdΦ<br />

8 0<br />

5 0<br />

( [ ] ) 2π (<br />

1 1<br />

=<br />

8 2 (2Θ) ( 4 [<br />

0<br />

5 ) − 1 ] π )<br />

3 (cosΦ)(2) = 4π<br />

0<br />

15<br />

sin 5 ΦdΦ (257)<br />

(258)<br />

(259)<br />

so that:<br />

D 1122 = − 1 (M 1122 + 2κ 1 N 1122 )<br />

4πµ 1<br />

(260)<br />

= − 1 ( 4 4πµ 1 3 π + 2κ 4π<br />

1<br />

15 ) (261)<br />

= − 1<br />

3µ 1<br />

(1 + 2 5 κ 1) (262)<br />

γ = D 1122 = − 1<br />

3µ 1<br />

(1 + 2 5 κ 1) (263)


Finally, ω represents any D ijkl term for which M ijkl = 0 but N ijkl ≠ 0, that is,<br />

for the following list of (ijkl) indices for which (i ≠ j and k ≠ l) but for which<br />

(i = k and j = l) or (i = l and j = k):<br />

1212, 2121, 1313, 3131, 2323, 3232, 1221, 2112, 1331, 3113, 2332, 3223 (264)<br />

For D 1313 , M 1313 = 0 and:<br />

=<br />

∫ 2π<br />

0<br />

=<br />

=<br />

∫ 2π<br />

0<br />

∫ π ∫ 2π<br />

0<br />

N 1313 =<br />

0<br />

∫ 2π<br />

=<br />

0<br />

∫ π ∫ 2π<br />

0<br />

0<br />

(<br />

z<br />

2<br />

1 z 2 3sinΦ ) dΘdΦ (265)<br />

(<br />

sin 2 Φcos 2 Θcos 2 ΦsinΦ ) dΘdΦ (266)<br />

cos 2 ΘdΘ<br />

∫ π<br />

0<br />

sin 3 Φcos 2 ΦdΦ (267)<br />

(∫ π<br />

cos 2 ΘdΘ sin 3 Φ ( 1 − sin 2 Φ ) )<br />

dΦ<br />

(∫ π<br />

cos 2 ΘdΘ sin 3 ΦdΦ −<br />

0<br />

0<br />

∫ π<br />

0<br />

(268)<br />

)<br />

sin 5 ΦdΦ = 4<br />

15 π (269)<br />

so that:<br />

ω = D 1313 = − 2κ 1<br />

N 1313 = − 2κ 1 4π<br />

4πµ 1 4πµ 1 15 = − 2 15<br />

κ 1<br />

µ 1<br />

(270)<br />

giving our model as:<br />

D ijkl = − 1 (1 + 2 3µ 1 5 κ 1)δ ij δ kl − 2 κ 1<br />

(δ ik δ jl + δ il δ jk ) (271)<br />

15 µ 1


15 Iterative Newton’s method solution of <strong>Kroner</strong>’s<br />

cubic<br />

Rewrite Eq. 211 as:<br />

x 3 + ax 2 − bx − c = 0 (272)<br />

where x = µ ∗ . It turns out that it is easier to use Newton’s method rather than<br />

the formulas for solving a cubic equation. To this end, we can write an exact<br />

Taylor’s expansion in the form:<br />

f(x + ∆x) = f(x) + f ′ (x)∆x + 1 2 f ′′ (x)(∆x) 2 + 1 6 f ′′′ (x)(∆x) 3 (273)<br />

since all derivatives greater than third order vanish. Now apply Newton’s<br />

method to first, second, or third order, as follows. Write:<br />

f(x n+1 ) = f(x n ) + f ′ (x n )∆x + 1 2 f ′′ (x n )(∆x) 2 + 1 6 f ′′′ (x n )(∆x) 3 (274)<br />

x n+1 = x n + ∆x (275)<br />

Newton’s method is usually based only on a first order expansion, solving for<br />

∆x assuming that the first derivative does not vanish:<br />

f(x n+1 ) = 0 ≈ f(x n ) + f ′ (x n )∆x (276)<br />

∆x = −f(x n)<br />

f ′ (x n )<br />

(277)<br />

We consider the case where the first derivative is too small by going to a second<br />

order approximate solution, and if both first and second derivatives are too<br />

small, by going to a third order solution. Details follow. ”Too small” means that<br />

the normalized value of the derivative is less than 10 −10 . If the first derivative<br />

is too small, then we check the second derivative. If it is also too small, then<br />

we use the third order approximation, and the Taylor’s expansion reduces to:<br />

f n+1 = f(x n+1 ) = 0 = f n + 1 6 f ′′′ (x n )(∆x) 3 (278)<br />

But from Eq.272, f ′′′ (x n ) = 6, so that we can solve for ∆x:<br />

f n = f(x n ) (279)<br />

∆x = (−f n ) 1 3 (280)<br />

If the first derivative is too small, but the second derivative is not too small,<br />

then from the second order approximation, we can solve for ∆x provided that<br />

< 0, so that:<br />

f n<br />

f ′′ n<br />

∆x =<br />

√<br />

−2 f n<br />

f ′′ n<br />

(281)


otherwise there is no solution (More below). If the first derivative is sufficiently<br />

large, then we can consider again the second order approximation:<br />

f(x n+1 ) = 0 ≈ f(x n ) + f ′ (x n )∆x + 1 2 f ′′ (x n )(∆x) 2 (282)<br />

Checking the discriminant D:<br />

D = (f ′ n) 2 − 2f ′′<br />

nf n (283)<br />

If this discriminant is negative, then we revert to the first order solution Eq.<br />

277. If the discriminant is positive but the second derivative is too small, the<br />

we also revert to the first order solution Eq. 277. Finally, if the discriminant is<br />

positive and the second derivative is sufficiently large, then we get the second<br />

order solution for ∆x:<br />

∆x = −f ′ n + √ D<br />

f ′′ n<br />

(284)<br />

We choose the root with the ”+” sign because as f n ⇒ 0, ∆x ⇒ 0. Note that<br />

all cases reported here fell in this second order solution, where both first and<br />

second derivatives were large and the discriminant was strictly positive.<br />

The case above without solution is interesting. This is the case where f n ′ is<br />

have the same sign. If they are both positive,<br />

this means that locally (about x n ), f(x) opens upward but does not cross the<br />

x-axis. Similarly, if they are both negative, f(x) opens downward but does not<br />

cross the x-axis. In both of these cases, there is clearly no solution.<br />

too small and both f n and f ′′<br />

n


References<br />

[1] A. V. Hershey. The elasticity of an isotropic aggregate of anisotropic cubic<br />

crystals. Journal of Applied Mechanics, 21:236–240, September 1954.<br />

[2] E. <strong>Kroner</strong>. Berechnung der elastischen konstanten des vielkristalls aus den<br />

konstanten des einkristalls. Zeitschrift fur physik, 151:504–518, 1958.<br />

[3] J. D. Eshelby. Elastic inclusions and inhomogeneities. In I.N. Sneddon<br />

and R. Hill, editors, Progress in solid mechanics, volume II, pages 88–140.<br />

North-Holland, 1961.<br />

[4] C. Weinberger, W. Cai, and D.M. Barnett. Lecture notes - elasticity of<br />

microscopic structures. http://micro.stanford.edu/~caiwei/me340b/,<br />

September 2005.<br />

[5] J.E. Gubernatis and J.A. Krumhansl. Macroscopic engineering properties<br />

of polycrystalline materials: Elastic properties. Journal of Applied Physics,<br />

46(5):1875–1883, May 1975.<br />

[6] K. Markov. Heterogeneous Media: micromechanics modeling methods and<br />

simulations, chapter ”Elementary micromechanics of heteregeneous media”,<br />

pages 1–146. Birkhauser, 2000.<br />

[7] I. Stakgold. Boundary value problems of mathematical physics, volume II.<br />

Macmillan, 1968.<br />

[8] A. Bertram and J. Olschewski. Formulation of anisotropic linear viscoelastic<br />

constitutive laws by a projection method. In A. D. Freed, editor,<br />

High temperature constitutive modeling-theory and application, volume<br />

AMD 121, pages 129–137. American Society of Mechanical Engineers, 1991.<br />

[9] V.A. Lubarda. New estimates of the third-order elastic constants for<br />

isotropic aggregates of cubic crystals. J. Mech. Phys. Solids, 45(4):471–<br />

490, 1997.<br />

[10] D. L. Andrews and T. Thirunamachandran. On three-dimensional rotational<br />

averages. Journal of chemical physics, 67(11):5026–5033, 1977.<br />

[11] Z. Hashin and S. Shtrikman. On some variational principles in anisotropic<br />

and nonhomogeneous elasticity. J. Mech. Phys. Solids, 10:335–342, 1962.<br />

[12] Z. Hashin and S. Shtrikman. A variational approach to the theory of the<br />

elastic behaviour of polycrystals. J. Mech. Phys. Solids, 10:343–352, 1962.<br />

[13] Gschneider. Solid State Physics, chapter Physical properties and interrelationships<br />

of metallic and semimetallic elements, pages 275–297. Academic<br />

Press, 1964.<br />

[14] G. Simmons and H. Wang. Single crystal elastic constants and calculated<br />

aggregate properties: a handbook. MIT Press, 1971.


[15] P. Sisodia and M.P. Verma. Shear moduli of polycrystalline cubic elements.<br />

J. Phys. Chem. Solids, 50(2):223–224, 1989.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!