0% found this document useful (0 votes)
9 views45 pages

Two-Versus Three-Dimensional Micropolar Elasticity

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
9 views45 pages

Two-Versus Three-Dimensional Micropolar Elasticity

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 45

P1: Binod

May 28, 2007 18:6 C4174 C4174˙C006

6
Two- versus Three-Dimensional
Micropolar Elasticity

Of course the continuum theory can yield, in principle, less


information about a material than does that of a correct, detailed
molecular model.
C.A. Truesdell, 1966

It was noted in Chapter 3 that lattices of beams should be modeled by


micropolar, rather than classical, elastic continua. This chapter outlines the
basic theory of such continuum models—both unrestricted and restricted
(couple-stress) ones. In particular, we first provide a formulation of basic equa-
tions in 3D, and then, in analogy to Chapter 3, we focus on planar micropolar
elasticity. Special attention is given to a generalization of the CLM result, and
its consequences. Furthermore, we discuss the problem of homogenization of
a heterogeneous Cauchy-type composite by a homogeneous micropolar-type
material. If conducted properly, one may then reduce the number of degrees
of freedom involved in, say, a finite element method, although the method
would have to account for a micropolar nature of the approximating body.
This also provides more physical insight into the so-called characteristic length,
usually an enigmatic concept appearing (and vanishing) in papers on micro-
polar theories. Although several monographs have been written on the subject
of micropolar media, most of the topics discussed in this chapter have never
been collected in a book form.

6.1 Micropolar Elastic Continua


6.1.1 Force Transfer and Degrees of Freedom
Every course on solid mechanics starts out with an introduction of the Cauchy
stress concept. This first involves identification of a finite surface area A
(= L 2 )—either in the interior of the body or on its external surface—defined
by an outer unit normal n, and a force F acting on A, Figure 6.1(a). Next,

191

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

192 Microstructural Randomness and Scaling in Mechanics of Materials

∆M

∆F

∆A

(a) (b)

FIGURE 6.1
(a) Force F and couple M acting on an internal (or external) surface area A (= L 2 ) in a
continuum; A is the area of any face of a cubic element of side L. (b) A porous medium in 2D,
viewed as a beam lattice, with each beam carrying a force and a couple locally. A unit cell of size
L is indicated with dashed lines.

one considers the ratio of F to A, and takes the limit


F(n)
lim = t(n) . (6.1)
A→0 A

It is a basic postulate of conventional solid mechanics that such a limit is well


defined, that is, that it is finite except the singularity points in the body, such
as crack tips. In the third step, following Cauchy himself, one introduces his
force-stress tensor τ as a linear mapping from n into t(n)
t(n) = τ · n. (6.2)

However, any consideration of a finite area A should involve a surface


couple M accompanying F. Thus, in analogy to (6.1), we must consider
M(n)
lim = m(n) , (6.3)
A→0 A

and, following Voigt (1887) and the brothers Eugène and François Cosserat
(1909), should introduce a couple-stress tensor µ as a linear mapping from n
into m(n)
m(n) = µ · n. (6.4)

Both t(n) and m(n) are shown acting on a face ABC of an arbitrary orientation
in Figure 6.2.
Note: The explicit consideration of µ makes τ nonsymmetric in general,
and that is the reason for using τ instead of the conventional Cauchy
stress σ .

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

Two- versus Three-Dimensional Micropolar Elasticity 193

x3 t(n)
m(n)

t(1)

C
m(1) n

x2
m(2)
B

t(2) O

A
m(3)
t(3) x1

FIGURE 6.2
Force traction and moment traction acting on face ABC with outer unit normal n of an infinitesimal
tetrahedron OABC.

If the microstructure is disregarded, we are dealing with an idealized,


homogeneous continuum in which M(n) must vanish in the limit L → 0. To
see this, take n to be aligned with n1 , and consider shear stresses τ12 and τ13 .
The torque caused by them, proportional to L 3 (τ12 − τ13 ), must disappear as
L → 0, because the cube’s volume scales as L 3 . Otherwise, we would be left
with a non-zero angular acceleration of a continuum point. This, in fact, is the
case with classical/conventional solid mechanics of Cauchy-type continua.
One then only has displacement u at a point, and assumes that m(n) = 0. But,
if the material intrinsically carries couples, we cannot disregard M(n) . Such a
situation occurs when the material has a discrete-type microstructure, such as
a beam-lattice shown in Figure 6.1(b), which simply precludes one from taking
A → 0. Here one needs to take A equal to the area of the elementary cell’s
cross-section, and the moment traction m(n) in (6.3) is defined at A finite.
For this model of force distribution in a continuous body to be fully con-
sistent with kinematics, each continuum point is endowed with six degrees of
freedom of a rigid body: three displacements ui (i = 1, . . ., 3) (or u) and three
rotations ϕi (i = 1, . . ., 3) (or ϕ), which are, in general, independent functions
of position and time. In particular, this implies that ϕ is not the same as the
macrorotation given by the gradient of u:

1
ϕi =
 e ijk uk, j (6.5)
2
Here, as before, e ijk is the Levi–Civita permutation tensor.

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

194 Microstructural Randomness and Scaling in Mechanics of Materials

When the inequality above is replaced by an equality, the continuum is


kinematically constrained—this will be discussed at various points in this
chapter. We now give some key facts in the vein of Nowacki, (1986).

6.1.2 Equations of Motion and Constitutive Equations


The equations of motion and constitutive equations of linear elasticity may
be derived from the energy conservation principle. This principle, for an adi-
abatic thermodynamic process and an anisotropic body, has the following
form:
 
d
(U + K ) = ( Xi vi + Yi wi ) dV + (ti vi + mi wi ) dS, (6.6)
dt V S

where
 
1 
U= ud V K = kd V k= ρvi vi + Iij wi w j vi = u̇i wi = ϕ̇i .
V V 2
(6.7)

In the above k is the the kinetic energy density, and u the internal energy,
both referred to a unit volume, while ρ denotes mass density and Iij is the
rotational inertia tensor. Note that the left-hand side of (6.6) represents the
rate of change of kinetic and internal energies, while the right-hand side is
the power of body forces and moments and surface forces and moments.
Let us now assume that the energy balance is invariant with respect to
rigid body motions when Xi , ti , Yi , and mi are kept fixed. Considering a
translational motion first, we substitute (with b i being an arbitrary constant
vector)

vi → vi + b i , v̇i → v̇i (6.8)

into (6.6) to get


 
 
u̇ + ρ (vi + b i ) v̇i + Iij wi ẇ j dV = [Xi (vi + b i ) + Yi wi ] dV
V V

+ [ti (vi + b i ) + mi wi ] dS. (6.9)
S

Subtracting equation (6.6) from (6.9), we obtain


 
bi ( Xi − ρ v̇i ) dV + b i ti dS = 0, (6.10)
V S

which, noting (6.2) and the Green–Gauss theorem, becomes



 
bi Xi + τji,j − ρ v̇i dV = 0. (6.11)
V

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

Two- versus Three-Dimensional Micropolar Elasticity 195

Since this has to hold for an arbitrary volume V, we obtain a local form of the
conservation of linear momentum

τji,j + Xi = ρ v̇i . (6.12)

With the above we can now simplify (6.6) to the form


  
   
u̇ + Iij wi w j dV = τji vi + Yi wi dV + [ti (vi + b i ) + mi wi ] dS,
V V S
(6.13)

which, noting (6.3) and the Green–Gauss theorem, becomes


  
   
u̇dV = τji vi, j + µji wi, j dV + µji,j + Yi − Iij w j wi dS. (6.14)
V V S

From this, the local form of conservation of energy may now be written as
 
u̇ = τji vi, j + µji wi, j + µji,j + Yi − Iij w j wi . (6.15)

Let us now postulate the energy balance to be invariant with respect to


rigid body rotations (with ωk being an arbitrary constant vector):

vi, j → vi, j − e ijk ωk wi, j → wi, j . (6.16)

Assuming u, Yi , τij , µij , and Iij to be unchanged, and proceeding in a fashion


similar as before, we arrive at a local form of the conservation of angular
momentum
..
e ijk τ jk + µji,j + Yi = Iij ϕ j . (6.17)

This equation brings about a simplification of the energy balance


 
u̇ = τji vi, j − e k ji wk + µji wi, j . (6.18)

It is now convenient to introduce two tensors describing the deformation


of the body—strain γji and torsion κij —as follows:

γji = ui, j − e k ji ϕk κji = ϕi, j . (6.19)

Evidently, γji and κij are generally nonsymmetric; κij is also called curvature
tensor, or torsion-curvature tensor. Just like in a classical continuum, we need
compatibility equations, and these are

γli,h − γhi,l − e khi κik + e kli κhk = 0 κli,h = κhi,l . (6.20)

Given (6.19), we can write the energy balance (6.18) as

u̇ = τij γ̇ij + µij κ̇ij . (6.21)

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

196 Microstructural Randomness and Scaling in Mechanics of Materials

Taking u to be a state function of γij and κij , we have

∂u ∂u
u̇ = γ̇˙ ij + κ̇ij , (6.22)
∂γij ∂κij

whereby we also assume τij and µij not to be explicitly dependent on the
temporal derivatives of γij and κij . A comparison of (6.21) with (6.22) then
leads to
∂u ∂u
τij = µij = . (6.23)
∂γij ∂κij
   
Clearly, τij , γij and µij , κij are conjugate pairs. Assuming a micropolar ma-
terial of linear elastic type, its energy density is given by a scalar product

1 (1) 1 (2)
u= γij Cijkl γkl + κij Cijkl κkl , (6.24)
2 2
so that Hooke’s law is
(1) (2)
τij = Cijkl γkl µij = Cijkl κkl , (6.25)

(1) (2)
Here Cijkl and Cijkl are two micropolar stiffness tensors. Note that, due to
the existence of u, we have the basic symmetry of both stiffness (and hence,
compliance) tensors
(1) (1) (2) (2)
Cijkl = Cklij Cijkl = Cklij , (6.26)
   
but not the two other symmetries since τij , γij and µij , κij are, in general,
nonsymmetric. Indeed, this is the reason for calling this theory an asymmetric
elasticity by Nowacki (1970, 1986). The inverse of (6.25) is written as
(1) (2)
γij = Sijkl τkl κij = Sijkl µkl . (6.27)

Note: In this chapter we employ τ and γ to distinguish them, respectively,


from the symmetric stress σ and symmetric strain ε of classical continuum
theory. This convention will become very useful in homogenization of a
heterogeneous Cauchy-type composite by a homogeneous micropolar-type
material, Section 6.4.

6.1.3 Isotropic Micropolar Materials


Focusing henceforth on the centrosymmetric case, for an isotropic material,
(1) (2)
Iij = δij I , while tensors Cijkl and Cijkl of (6.25) become

(1)
Cijkl = (µ − α) δ jk δil + (µ + α) δjl δik + λδij δkl
(2)
Cijkl = (γ − ε) δ jk δil + (γ + ε) δjl δik + βδij δkl ,

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

Two- versus Three-Dimensional Micropolar Elasticity 197

where λ and µ are the Lamé constants of classical elasticity, while α, γ , ε, and
β are the micropolar constants. The free energy density is given by a scalar
product
µ+α µ−α λ
u= γji γji + γji γij + γkk γnn
2 2 2
γ +ε γ −ε β
+ κji κji + κji κij + κkk κnn , (6.28)
2 2 2
With (6.28), we can write (6.25) in two equivalent forms:

τji = (µ + α) γji + (µ − α) γij + λδij γkk τij = 2µγ(i j) + 2αγ[i j] + λδij γkk
µji = (γ + ε) κji + (γ − ε) κij + βδij κkk µij = 2γ κ(i j) + 2εκ[i j] + βδij κkk .
(6.29)

The round and square brackets indicate symmetric and antisymmetric parts
of the tensors, respectively. Of use also will be the inverse forms of this con-
stitutive law, namely,

γij = 2µ τ(i j) + 2α  τ[i j] + γ  δij τkk κij = 2γ  µ(i j) + 2ε  µ[i j] + β  δij µkk , (6.30)

in which
1 1 1 1
2µ = 2α  = 2γ  = 2ε  =
2µ 2α 2γ 2ε
(6.31)
 −λ  −β 2 2
λ = β = κ =λ+ µ  = β + γ.
6µK 6γ  3 3
Here we recognize the familiar bulk modulus κ, and its mathematically anal-
ogous micropolar quantity .
Clearly, there are six material constants: α, β, γ , ε, µ, λ. Considering the
fact that u in (6.28) is a positive definite quadratic form, one can show that
the following inequalities should hold:

3λ + 2µ > 0 µ>0 3β + 2γ > 0 γ > 0


(6.32)
µ+α >0 γ +ε >0 α>0 ε > 0.

Our constitutive tensors (6.26) may alternatively be expressed in the


notation of Eringen (1966, 1999)
(1)
Cijkl = µ E δ jk δil + (µ E + α E ) δjl δik + λ E δij δkl
(2)
(6.33)
Cijkl = β E δ jk δil + γ E δjl δik + α E δij δkl ,

where, using the subscript E to denote quantities in Eringen’s notation, we


have

µE = µ − α κ E = 2α λE = λ γE = γ + ε β E = γ − ε. (6.34)

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

198 Microstructural Randomness and Scaling in Mechanics of Materials

We end this section by noting that, just like in the classical elasticity, we
can express a micropolar field problem in displacements and rotations, or in
stresses and couple-stresses. In the first case, six such equations (for ui and ϕi ,
i = 1, . . ., 3) are obtained by substituting (6.30) into the equilibrium equations
(6.12) and (6.17), and using (6.19):

..
(µ + α) ui, j j + (λ + µ − α) u j, ji + 2αe ijk ϕk + Xi = ρ ui
.. (6.35)
(γ + ε) ϕi, j j − 4αϕi + (β + γ − ε) ϕ j, ji + 2αe ijk uk + Yi = I ϕ .

This generalization of the Navier equations of classical elasticity is to be sup-


plemented by the kinematic boundary conditions on ∂ Bk and traction condi-
tions on ∂ Bt , where ∂ Bk ∪ ∂ Bt = ∂ B, that is,

τij (x, t) n j = ti (x, t) µij (x, t) n j = mi (x, t) x ∈ ∂ Bt t>0


(6.36)
ui (x, t) = f i (x, t) ui (x, t) = gi (x, t) x ∈ ∂ Bk t > 0.

Here ti , mi , f i , gi are prescribed functions.


The field equations in stresses are a generalization of the Beltrami–Michell
equations. Here we make a reference to a study of Schäfer (1967), who gen-
eralized the functions of Morrey and Maxwell and Kessel. We will return to
this topic in the 2D setting in Section 6.3 below.
As pointed out in Section 6.1.1, ϕ is an independent kinematic quantity.
However, a special model assuming the equality

1
ϕi = e ijk uk, j (6.37)
2

is sometimes used, and this is the same definition as in classical elasticity. It


is called a pseudo-continuum, a restricted, a couple-stress, or a Koiter-Mindlin
model, Koiter (1963); see also Truesdell and Toupin (1960), Grioli (1960),
Toupin (1962), Mindlin and Tiersten (1962), and Mindlin (1963). In view of
(6.37), the strains γij become symmetric and are (classically) defined as

γij = u(i, j) . (6.38)

A well-known case of such materials is the Bernoulli–Euler beam—indeed,


a 1D continuum—in which the cross-section is restricted to rotate according
to the gradient of the transverse displacement.

6.1.4 Virtual Work Principle


Let us consider fields of virtual displacement δui and rotation δϕi —both of
them continuous, consistent with boundary conditions, and infinitesimal. If
we multiply equation (6.12) by δui , and equation (6.17) by δϕi , and add the

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

Two- versus Three-Dimensional Micropolar Elasticity 199

results, we find

.. ..
[( Xi − ρ ui )δui + (Yi − I ϕ i )δϕi ]dV
V

+ [τji,j δui + (e ijk τ jk + µji,j )δϕi ]dV = 0. (6.39)
V

Upon transformation of the second of the above integrals, we obtain


 
.. ..
( Xi − ρ ui )δui dV + (Yi − I ϕ i )δϕi dV
V V
 
 
+ [ti δui + mi δϕi ] dS = τji δγji + µji δκji dV, (6.40)
V V

which expresses an equality between the virtual work of external and internal
forces; the latter are conjugate to fields of virtual strain δγji and rotation δκji .
Upon introduction of (6.29) into (6.39), we set up the variational principle

.. ..
[( Xi − ρ ui )δui + (Yi − I ϕ i )δϕi ]dV
V

+ [ti δui + (mi δϕi ]d S = δW (6.41)
S

where
 
λ
W= µγ(i j) γ(i j) + αγ[i j] γ[i j] + γkk γmm
V 2

β
+ γ κ(i j) κ(i j) + εκ[i j] κ[i j] + κkk κmm dV. (6.42)
2
This principle may be used to derive the energy conservation principle by
comparing the functions u and ϕ at a point x and time t with those quantities at
x and time t + dt. Thus, introducing δui = vi dt; δϕi = wi dt; vi = u̇i ; wi = ϕ̇i
into (6.41), we obtain
 
d
( K + W) = ( Xi vi + Yi wi ) d V + (ti vi + mi wi ) d S (6.43)
dt V S

This is the starting point for the proof of uniqueness of solutions—the proce-
dure is analogous to that in classical elasticity.

6.1.5 Hamilton’s Principle


Consider now a micropolar elastic body undergoing some motion between
times t = t1 and t = t2 . We now compare the actual displacements u (x, t) and
rotations ϕ (x, t) with the virtual u (x, t) + δu and ϕ (x, t) + δϕ, whereby the
latter are chosen so as to satisfy the conditions

δu (x, t1 ) = δu (x, t2 ) = 0 δϕ (x, t1 ) = δϕ (x, t2 ) = 0. (6.44)

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

200 Microstructural Randomness and Scaling in Mechanics of Materials

The virtual work principle is now written as



.. ..
δL − (ρ ui δui + I ϕ i δϕi )dV = δW. (6.45)
V

where
 
δL = ( Xi δui + Yi δϕi ) dV + (ti δui + mi δwi ) dS. (6.46)
V S

Integrating (6.45) over the interval t1 ≤ t ≤ t2 ,


 t2  t2  t2 
.. ..
δ Wdt = δLdt − dt (ρ ui δui + I ϕ i δϕi )dV, (6.47)
t1 t1 t1 V

while introducing the variation of kinetic energy,


   
∂ .. ∂ ..
δK = ρ ( u̇i δui ) dV − ρ ui δui dV + I ( ϕ̇i δϕi ) dV − I ϕ i δϕi dV,
V ∂t V V ∂t V
(6.48)

and integrating it also over t1 ≤ t ≤ t2 , and taking note of (6.44), we find


 t2  t2   t2 
.. ..
δ K dt = −ρ dt ui δui dV − I dt ϕ i δϕi dV. (6.49)
t1 t1 V t1 V

In view of (6.47) and (6.49), we finally obtain Hamilton’s principle generalized


to a micropolar medium
 t2  t2
δ (W − K ) dt = δ Ldt. (6.50)
t1 t1

Variation and integration on the right-hand side of (6.50) commute when


the external forces are conservative and derivable from a potential V. In that
case,
 
∂V ∂V ∂V ∂V
δL = − δui + δϕi = −δ ui + ϕi , (6.51)
∂ui ∂ϕi ∂ui ∂ϕi

and (6.50) becomes


 t2
δ ( − K ) dt = 0 =W+V (6.52)
t1

in which  denotes the total potential energy.

6.1.6 Reciprocity Relation


Consider an isotropic body of volume V and bounding surface ∂ B loaded by
a system {X, Y, t, m}, which then results in {u, ϕ}. The initial conditions are

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

Two- versus Three-Dimensional Micropolar Elasticity 201

homogeneous:
ui (x, 0) = 0 u̇i (x, 0) = 0 ϕi (x, 0) = 0 ϕ̇i (x, 0) = 0 (6.53)
Also, consider another loading system X , Y , t , m acting on the same body,
resulting in u , ϕ  , both causes and effects being now denoted by primes.
This is subject to analogous initial conditions as in the first case.
We now apply Laplace transformation to the constitutive equations (6.29)
to get
τji = (µ + α) γji + (µ − α) γij + λδij γkk
(6.54)
µji = (γ + ε) κji + (γ − ε) κij + βδij κkk ,
where
 ∞
τji (x, p) = τji (x, t) e − pt dt, etc. (6.55)
0
Proceeding in a similar fashion with the primed quantities τji (x, p), etc., a
statement entirely similar to (6.54) above is obtained.
It is easy to verify that the following is true:

τji γji + µji κji = τji γji + µji κji ; (6.56)


and this, upon a volume integration, becomes
     
τji γji + µji κji dV = τji γji + µji κji dV. (6.57)
V V

Next, carry out the Laplace transformation on the equations of motion (6.12)
and (6.17) corresponding to the first loading system to get

τji,j + Xi = p 2 ρ ui 
e ijk τ jk + µ ji,j + Yi = p I ϕi .
2
(6.58)

Noting a completely analogous relation corresponding to the second loading


system, (6.57) can be converted to
 
  
   
Xi ui + Yi ϕi − p ρ ui − p I ϕi dV +
2 2
ti ui + mi ϕi dS
V S
 
   
= Xi ui + Yi ϕi dV + ti ui + mi ϕi dS. (6.59)
V S

Upon carrying out the inverse Laplace transformation, we arrive at the the-
orem of reciprocity of work of causes and effects in both loading systems
(Sandru, 1966):
 
   
Xi ∗ ui + Yi ∗ ϕi dV + ti ∗ ui + mi ∗ ϕi dS
V S
 
   
= Xi ∗ ui + Yi ∗ ϕi dV + ti ∗ ui + mi ∗ ϕi dS, (6.60)
V S

where ∗ denotes a convolution operation between Xi and ui .

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

202 Microstructural Randomness and Scaling in Mechanics of Materials

This is one of the most interesting theorems of micropolar elasticity. Its


generality offers a possibility of integration of the equations of elastodynamics
using Green’s function.

6.1.7 Elements of Micropolar Elastodynamics


6.1.7.1 Basic Equations
We begin by noting that equations (6.35) can also be written in a vector form

2u + (λ + µ − α) grad divu + 2α rotϕ + X = 0


(6.61)
4ϕ + (β + γ − ε) grad divϕ + 2α rotu + Y = 0,

where

2 = (µ + α) ∇ 2 − ρ ∂2t 4 = (γ + ε) ∇ 2 − 4α + I ∂2t , (6.62)

are the d’Alembert and Klein–Gordon operators, respectively, and ∂2t indicates
the second derivative with respect to time.
The physics represented by this coupled system of hyperbolic differential
equations can be understood by operating either with divergence or rotation
upon it. In the first case, we find

1 divu + divX = 0
(6.63)
3 divϕ + divY = 0,

where again we introduced two partial differential operators

1 = (λ + 2µ) ∇ 2 − ρ ∂2t 3 = (β + 2γ ) ∇ 2 − 4α + I ∂2t . (6.64)

On the other hand, upon carrying out a rotation on (6.61), we find

3 r otu + 2α rot rotϕ + r otX = 0


(6.65)
3 r otϕ + 2α rot rotu + r otY = 0.

If we now operate with 1 4 on (6.61)1 and employ (6.63)1 and (6.65)1 ,


and similarly we operate with 3 4 on (6.62)2 and employ (6.63)2 and (6.65)2 ,
we shall find
   
2 4 + 4α ∇ u = − 1 4 − grad div X + 2α rot 1 Y
2 2
1
    (6.66)
2 4 + 4α ∇ ϕ = − 2 3 − grad div Y + 2α rot 1 X,
2 2
3

where

 = (λ + µ − α) 4 − 4α 2  = (β + γ − ε) 2 − 4α 2 . (6.67)

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

Two- versus Three-Dimensional Micropolar Elasticity 203

Now, let us introduce a representation for displacements and rotations,


employing two vector functions F and G:
 
u= 1 4 − grad div F − 2α rot 3 G
  (6.68)
ϕ= 2 3 − grad div G − 2α rot 1 F.

With the substitution of (6.68) into (6.66) we find two equations governing F
and G:
 
1 2 4 + 4α 2 ∇ 2 F + X = 0
  (6.69)
4 + 4α ∇ G + Y = 0.
2 2
3 2

More insight into what is represented by the equations of motion (6.61)


may be gained by using Helmholtz’s theorem for u and ϕ:

u = grad  + r ot , div  = 0
(6.70)
ϕ = grad  + r ot H, div H = 0.

If we also introduce the same type of decomposition for body forces and
body couples, that is,
 
X = ρ grad ϑ + r ot χ , div χ = 0
  (6.71)
Y = I grad σ + r ot η , div η = 0.

we find that (6.61) reduces to four wave equations

1 + ρϑ = 0 3 + Iσ = 0
(6.72)
2 + 2α rotH + ρχ = 0 4H + 2α rot + I η = 0.

The first of these equations represents a longitudinal wave motion, identi-


cal with what is known from classical elastodynamics. The second equation
represents a longitudinal wave of (micro)rotation. The third and fourth equa-
tions represent propagation of two transverse waves—in displacements and
in rotations—which may be cast in the form
 
2 4 + 4α 2 ∇ 2  = 2α I rot η − ρ 4χ
  (6.73)
4 + 4α ∇ H = 2αρ rot χ − I 2 η.
2 2
2

These waves were investigated by Ignaczak (1970).

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

204 Microstructural Randomness and Scaling in Mechanics of Materials

6.1.7.2 Plane Monochromatic Waves


Consider a plane monochromatic wave propagating in an infinite body, and
harmonic in time. Let the wavefront at time t = 0 be in the plane p = xi ni ,
where n is the unit vector normal to the plane. The displacements and rota-
tions are now of the form

u j = Aj exp [−ik (t − nk xk )] ϕ j = B j exp [−ik (t − nk xk )] , (6.74)

where k = ω/ = 2π/l is the phase velocity, ω is the angular velocity, and
l is the wave length. Introducing (6.74) into (6.35), we arrive at a system of
algebraic equations
  2αi
µ + α − ρ2 Aj + (λ + µ − α) n j xk Ak + e jkl nl Bk = 0,
k
 (6.75)
4α 2αi
γ + ε + 2 − I 2 B j + (β + γ − ε) n j xk Bk + e jkl nl Ak = 0.
k k
Setting this system’s determinant to zero leads to

  4α
λ + 2µ − ρ 2
β + 2γ + 2 − I 2
k
   (6.76)
  4α 4α 2
µ + α − ρ2 γ + ε + 2 − I 2 − 2 = 0,
k k
from which we determine phase velocities of various plane waves. The first
term in (6.76) yields
 1/2
λ + 2µ
= (6.77)
ρ
The second term in (6.76) yields dispersive (i.e., ω-dependent) wave propa-
gation
 −1/2  1/2
ω2 β + 2γ 4α
 = 3 1 − 02 , 3 = , ω02 = , (6.78)
ω I I
which has physical meaning only for ω > ω0 , because this condition ensures
real values of . The third term in (6.76) yields a quartic equation

  4α 4α 2
µ + α − ρ2 γ + ε + 2 − I 2 − 2 = 0. (6.79)
k k

6.1.8 Noncentrosymmetric Micropolar Elasticity


Recall that the constitutive equations relate force stresses with strains on one
hand, and moment stresses with curvatures on the other. In general, however,
there is a possibility of a direct coupling between force-type and couple-type

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

Two- versus Three-Dimensional Micropolar Elasticity 205

effects in the constitutive model, whereby τij would also be a function of κkl ,
and µij a function of γkl . A simple 1D case of chirality is the helix already
discussed in Section 3.5 of Chapter 3. Thus, generalizing (6.25), we write
(1) (3)
τij = Cijkl γkl + Cijkl κkl ,
(4) (2)
(6.80)
µij = Cijkl γkl + Cijkl κkl .

This is called either non-centrosymmetry or hemitropy (Aero and Kuvshinskii


1964, 1969; Nowacki, 1986), or chirality (Lakes and Benedict, 1982; Lakes, 2001).
Note that all the developments of this chapter preceding this equation per-
tained to centrosymmetric materials. Note that, by the argument of reciprocity,
(3) (4)
Cijkl = Cijkl . (6.81)

As Lakes points out, because the coordinate changes in the centrosymmetric


material result in the transformation matrix

a im = −δjm , (6.82)
(1)
the stiffness tensor Cijkl satisfies

(1) (1) (1)


Cijkl = a im a jn a ko a lp Cmnop
(1)
= (−1) 4 Cijkl = Cijkl . (6.83)

6.2 Classical vis-à-vis Nonclassical (Elasticity) Models


6.2.1 A Brief History
Following its birth, the theory of the Cosserat brothers (1896, 1909) remained
dormant for half a century, apparently the only exceptions being the works
of Somigliana (1910) and Sudria (1935); see also Ball and James (2002). This
hibernation was likely due to the theory’s generality (as a nonlinear theory
with finite motions and inelastic interactions) and its presentation as a uni-
fied theory incorporating mechanics, optics, magnetism and electrodynam-
ics. The dynamic growth of continuum mechanics and thermodynamics (e.g.,
Ericksen and Truesdell, 1958; Truesdell and Toupin, 1960) begun in the fifties
and sixties brought the work of the Cosserat brothers back into focus. Funda-
mentals of a general linear Cosserat continuum were given by Günther (1958),
who discussed in detail the 1-, 2-, and 3D Cosserat models, as well as their
significance in the dislocation theory, and Schäfer (1962), who focused on the
planar case. From that period one should also mention several other works.
Thus, Grioli (1960) established the constitutive relations for finite deforma-
tions of perfectly elastic solids. Aero and Kuvshinskii (1960) independently
derived the equilibrium equations and constitutive relations for anisotropic
solids in the linearized theory. Mindlin and Tiersten (1962) established the

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

206 Microstructural Randomness and Scaling in Mechanics of Materials

boundary conditions; see also Kröner (1963), Koiter (1963), and Eringen (1968).
An expression of the growing interest in Cosserat theory was soon found in
symposia (e.g., Kröner 1968) and monographs on the subject (e.g., Nowacki,
1970, 1986; Stojanovic, 1970; Brulin and Hsieh, 1982).
Building on the shoulders of the Cosserats, and to account for increasing
levels of complexity, other, more general theories accounting for higher-order
interactions such as monopolar, multipolar, and strain-gradient were intro-
duced (see e.g., Green and Rivlin, 1964; Toupin, 1962, 1964; Jaunzemis, 1967;
Tiersten and Bleustein, 1974). There are also “micropolar,” “microstretch,”
and most generally “micromorphic” continua (Eringen, 1999, 2001; Mariano,
2001). To clarify the key concepts here, following Goddard (2006), let us con-
sider a series expansion of the velocity (or infinitesimal displacement) field v

v (x) = v0 + L1 · r + L2 · r2 + · · ·, (6.84)

where

r = x − x0 , Ln = 1
n! (∇ ⊗ v) 0T , (6.85)

and rn denotes the n-fold symmetric tensor product ⊗n r. This allows an


expansion for the global stress-power density in a simple continuum
 
1
ẇ = σ : LdV = ẇn , (6.86)
V V n

with

ẇn = σn : Ln , σn := V σ : rn dV, (6.87)

where L = (∇ ⊗ v) T is is the first velocity gradient, its dual being the Cauchy
stress σ . Furthermore, while we easily see that the higher-order kinematic
quantities Ln are conjugate to the stress moments σn , there are two ways to
interpret the Ln :
1. Multipolar: the Ln are identical with the higher gradients of the
velocity field. This viewpoint was advanced by Green and Rivlin
(1964a,b) and Mindlin (1963).
2. Micromorphic: the Ln are intrinsic particulate fields (i.e., pertaining
to generally deformable particles making up the macro-continuum),
which require their own constitutive equations. This approach dates
back to the Cosserat brothers, and was then further pursued by
Eringen (1999) leading to microstretch and micromorphic theories.
The microscopic treatment dictates that, in a multipolar continuum, the
stress moments should satisfy a hierarchy of balances
∇ · σn+1
T
+ σn = Gn+1 for n = 0, 1, . . ., (6.88)

where σ0 := 0, σ1 := σ , and the Gs represent extrinsic body moments accu-


mulation of intrinsic multipolar momenta, which effectively vanish in the

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

Two- versus Three-Dimensional Micropolar Elasticity 207

quasi-static limit. Equation (6.88) leads to the following balance:


   
∂V σn + x ⊗ σ n−1 + · · · + x ⊗σ · dS = V Gn dV for n = 2, 3, . . .,
n−1
(6.89)

In a certain sense, all of these theories can be considered as simpler cases of


“nonlocal continuum theories” (Eringen and Hanson, 2002), which, according
to these authors, “are concerned with material bodies whose behavior at any
interior point depends on the state of all other points in the body—rather than
only on an effective field resulting from these points—in addition to its own
state and the state of some calculable external field.” Focusing henceforth on
micropolar theories, we would like to note their extensions beyond purely
elastic material behaviors. An extension pertaining to thermoelasticity was
already given by Nowacki (1966) and Tauchert et al. (1968); see Dhaliwal
and Singh (1987) for a review. A micropolar generalization of viscoelasticity,
with a focus on waves, was presented by Maugin (1974). Beginning with
Green and Naghdi (1965), Mişicu (1964), and Sawczuk (1967), there has also
been research on (elastic-)plastic continua with microstructure, e.g., Fleck
et al. (1994) and Hutchinson (2000). This has then led to strain-gradient models
(Aifantis, 1987; Zbib and Aifantis, 1989; Fleck and Hutchinson, 1997; Pamin,
2005). Extensive research has also been done on micropolar fluid mechanics
(e.g., Cowin, 1974; Eringen, 2001).
Although the nonclassical theories have become very advanced mathe-
matically and explained effects that could not be brought out by classical
theories, they usually lacked the input of physically based constitutive coef-
ficients. Besides the beam lattices discussed in Chapters 3 and 4, progress has
been made on that front for composite materials; see Section 6.5 below.
Mindlin (1963) found that stress concentrations in the presence of holes
are lowered in Cosserat-type versus those in Cauchy-type solids. This was
followed by studies due to Neuber (1966) Kaloni and Ariman (1967), Cowin
(1970a,b) and Itou (1973). On the other hand, an increase of stress concentra-
tions in the vicinity of rigid inclusions was established by Hartranft and Sih
(1965) and Weitsman (1964). Micropolar effects also allow a better analysis of
localization present in failure of solids then that possible in classical continue
(William, 1995).
The case of holes motivated one of the earliest experimental studies of
couple-stress effects by Schijve (1966), who actually found that effect to be in-
significant. However, given the fact that he used aluminum sheets—a macro-
scopically homogeneous material without, say, reinforcing inclusions—his
investigation pertained to couple-stress effects due to the atomic lattice of
aluminum. This is not surprising in view of the fact that couple-stress effects
vanish on scales much larger than the microscale. Indeed, the situation is much
different in, say, a lattice of beams (which may be interpreted as a material
with large holes), if one looks at dependent fields on scales comparable to the
lattice spacing; recall Chapter 3.
Many studies of wave propagation in the context of harmonic distur-
bances have also been conducted. First, in addition to classical dilatational
and shear waves in an unbounded medium, there also exist rotational waves.

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

208 Microstructural Randomness and Scaling in Mechanics of Materials

Next, it turns out that only the dilatational waves propagate non-dispersively
(Nowacki, 1986; Eringen, 1999). In general, this is indicative of various new
dispersion effects in other wave problems, which are not present in classical
continua. In some cases of Cosserat continua, entirely new phenomena arise
such as, for instance, that a layer on top of an elastic half-plane is not necessary
for the propagation of Love waves—in the classical case, a layer is necessary.
Many results on periodic and aperiodic waves were collected by Nowacki
(1986), see also Eringen (1999).
The recent monograph by Dyszlewicz (2004) on micropolar elasticity
collects many new results, including the general methods of integration of
basic equations (Galerkin, Green–Lamé, and Papkovitch–Neuber type), for-
mulations of problems (displacement-rotation and pure stress problems of
elastodynamics), as well as solutions to various boundary value problems
(stationary 2D and 3D problems for a half-space, singular solutions to 2D and
3D elastodynamics and the thermoelastodynamics problems for an infinite
space).
Several workers, in the 1960s, derived micropolar models explicitly from
the microstructure. The work of theoreticians started from lattice-type models
enriched with flexural—in addition to central–interactions (e.g., Askar, 1985;
Banks and Sokolowski, 1968; Woźniak, 1970; Bažant and Christensen, 1972;
Holnicki-Szulc and Rogula, 1979a,b; Bardenhagen and Triantafyllidis, 1994).
From the outset, these models adopted Cosserat-type continua in analyses
of large engineering structures such as perforated plates and shells, or lat-
ticed roofs. There, the presence of beam-type connections automatically led to
micropolar interactions and defined the constitutive coefficients. In principle,
such models have their origin in atomic lattice theories (e.g., Berglund, 1982);
see Friesecke and James (2000) for the latest work in that direction.
Several workers (e.g., Perkins and Thompson, 1973; Gauthier and
Jahsman, 1975; Yang and Lakes, 1982; Lakes, 1983, 1986) have provided ex-
perimental evidence of micropolar effects in porous materials such as foams
and bones. In particular, Lakes (1995) was able to infer micropolar constants
from his experiments, both for centrosymmetric and chiral materials. Another
interesting application in the context of biomechanics was due to Shahinpoor
(1978).
It is also to be noted that composite materials may naturally lead to
Cosserat models where the nonclassical material constants can directly be
calculated from the microstructure; this was done in 1D by Herrmann and
Achenbach (1968). But, a similar task in 2D and 3D has only been undertaken
recently, and this is described in Section 6.5. In more recent years, progress
has been made on derivation of effective (homogeneous) Cosserat models
for heterogeneous composite materials of either Cauchy or Cosserat type. We
point out in Chapter 4 that a central-force lattice (truss of two-force members)
is an example of the former material, while a lattice of beams is an example
of the latter one.
All the studies in the area of stress singularities due to cracks were pre-
ceded by Muki and Sternberg (1965), who studied stress concentrations caused

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

Two- versus Three-Dimensional Micropolar Elasticity 209

by concentrated surface loads or discontinuously distributed surface shear


tractions. Next, Sternberg and Muki (1967) and Bogy and Sternberg (1967)
studied the implications of the couple-stress theory on unbounded concen-
trations of stress and on locally infinite deformation gradients. Basically, it
was found that, depending on a given situation, where the classical elasticity
would predict infinite (singular) stresses, the couple-stress theory may give
either finite stresses or weaker singularities, or have an opposite tendency (see
also Cowin, 1969; Atkinson and Leppington, 1977). This involves a proper
generalization of conservation integrals, which has recently been given in the
setting of couple-stress elasticity (Lubarda and Markenscoff, 2000).
Recently, Griffith’s fracture theory has been generalized to rectilinear and
fractal cracks in micropolar solids (Yavari et al., 2002). In particular, two cases
of the Griffith criterion were considered, depending on whether the effects of
stresses and couple-stresses are coupled or uncoupled, the key finding being
that both cases give equal orders of stress and couple-stress singularities,
which is the same result as that in a classical continuum. Also, the effect
of fractality of fracture surfaces on the powers of stress and couple-stress
singularity was studied.

6.2.2 The Ensemble Average of a Random Local Medium is Nonlocal


As pointed out in Chapter 2, the formal solution for the average of a field
problem governed by a linear random (and local) operator on the domain B
is a deterministic nonlocal operator. This was illustrated in terms of a Fourier-
type heat conduction problem, a result that immediately carries over to an-
tiplane elasticity. Moving to a general setting of linear elastostatics on the
random field of a fourth-rank stiffness tensor Cijkl = {Cijkl (ω, x) ; ω ∈ , x ∈R2 },
the equations governing the average fields in a Cauchy-type continuum are
(Beran and McCoy, 1970a)
 
σij (x) , j = 0

      
σij (x) = ijkl x, x εkl x dx (6.90)
B
   
εkl = uk,l + ul,k /2.

(Here we employ
  σij and εkl to denote symmetric stress and strain tensors.) In
(6.90)2 ijkl x, x is an infinite sum of integrodifferential operators, involving
moments of all orders of the random field Cijkl
         
ijkl x, x = Cijkl + Dijkl x δ x − x + E ijkl x, x , (6.91)
   
where Dijkl x and E ijkl x, x are functions of the statistical properties of Cijkl
and the free-space Green’s function of the nonstatistical problem. Addition
of a deterministic body force field  fi does not change
 the results. When the
fluctuations in Cijkl are small, Dijkl x and E ijkl x, x may be evaluated explic-
itly, and this was done by Beran and McCoy (1970) in the special case of the

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

210 Microstructural Randomness and Scaling in Mechanics of Materials

realizations Cijkl (ω) being locally isotropic, that is, expressed in terms of a vec-
tor random field of two Lamé coefficients {[λ (ω, x) , µ (ω, x)] ; ω ∈ , x ∈R2 };
recall Section 2.3 of Chapter 2.
Next, considering this random field to be statistically homogeneous and
mean-ergodic, one may disregard the contributions of this operator for |x−x |
> lc (the correlation length). Thus, since only the neighborhood within the
distance lc of x has a significant input into the integral (6.85)2 , one may expand
εkl (x ) in a power series about x:
     
εkl x = εkl (x) + xm − xm εkl (x) ,m
   
xm − xm xn − xn
+ εkl (x) ,mn . . . (6.92)
2
so as to obtain

   
σij (x) = ijkl x, x dx εkl (x)
B

  
+ ijkl x, x xm − xm dx εkl (x) ,m + · · · . (6.93)
B

This, in turn, can be rewritten as a sum of local, plus first gradient, plus higher
gradient strain effects:
  ∗ ∗ ∗
σij = Cijkl εkl + Dijklm εkl ,m + E ijklmn εkl ,mn + · · · . (6.94)
    ∗
Thus, B ijkl x, x dx in (6.93) is recognized as the effective stiffness Cijkl ; in-
deed the stiffness of a single realization B (ω) of the random material B. If
one is given the ensemble B of B (ω), then one may determine the microstruc-
∗ ∗
tural statistics, and hence the higher-order approximations Dijklm , E ijklmn , and
so on.

6.3 Planar Cosserat Elasticity


6.3.1 First Planar Problem
There are, in general, two planar problems of Cosserat elasticity (Nowacki,
1986):
1. The so-called first planar problem with u = (u1 , u2 , 0) and ϕ =
(0, 0, ϕ3 ), which is a generalization of the classical planar elasticity.
2. The so-called second planar problem u = (0, 0, u3 ) and ϕ = (ϕ1 , ϕ2 , 0),
which is a generalization of the classical antiplane elasticity.
Focusing on the first planar problem in static setting, from (6.12) and (6.7),
the equilibrium equations become

τ11,1 + τ21,2 = 0 τ12,1 + τ22,2 = 0 τ12 − τ21 + µ13,1 + µ23,2 = 0, (6.95)

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

Two- versus Three-Dimensional Micropolar Elasticity 211

while the kinematic relations (6.19) are

γ11 = u1,1 γ22 = u2,2 γ12 = u2,1 − ϕ3 γ21 = u1,2 + ϕ3


(6.96)
κ13 = ϕ3,1 κ23 = ϕ3,2 ,

and these satisfy the compatibility equations

γ21,1 − γ11,2 = κ13 γ22,1 − γ12,2 = κ23 κ23,1 = κ13,2 . (6.97)

In the isotropic planar Cosserat medium, compliances of (6.27) become

(1) 1  (2)
Sijkl = (S + P)δik δjl + (S − P)δil δ jk + ( A − S)δij δkl Si3k3 = δik M, (6.98)
4
where A, S, P, and M are four independent planar Cosserat constants, defined
in Ostoja-Starzewski and Jasiuk (1995):

1 1 1 1 1
A= = S= P= M= . (6.99)
κ λ+µ µ α γ +ε

Note that A and S define planar bulk and shear compliances of classical
elasticity (Dundurs and Markenscoff, 1993), while P and M are two additional
Cosserat constants; in the couple-stress elasticity P = 0. The restriction that
the strain energy be nonnegative implies the following inequalities:

0 ≤ A≤ S 0 ≤ P 0 ≤ M. (6.100)

In the case of orthotropy for plane Cosserat elasticity, constitutive equa-


tions (6.27) become
(1) (1) (1) (1)
γ11 = S1111 τ11 + S1122 τ22 γ22 = S2211 τ11 + S2222 τ22
(1) (1) (1) (1)
γ12 = S1212 τ12 + S1221 τ21 γ21 = S2112 τ12 + S2121 τ21 (6.101)
(2) (2)
κ13 = S1313 µ13 κ23 = S2323 µ23 .

In the above, given (6.26), we have


(1) (1) (1) (1)
S1122 = S2211 S1221 = S2112 . (6.102)

Because for the couple-stress formulation γ12 = γ21 (recall equation 6.39),
we must have
(1) (1) (1) (1)
S1212 = S2112 S1221 = S2121 . (6.103)

This, combined with (6.102)2 above implies


(1) (1) (1) (1)
S1212 = S2112 = S1221 = S2121 , (6.104)

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

212 Microstructural Randomness and Scaling in Mechanics of Materials

so that the constitutive relations (6.101) take on a simpler form


(1) (1) (1) (1)
γ11 = S1111 τ11 + S1122 τ22 γ22 = S1122 τ11 + S2222 τ22
(1)
γ12 = γ21 = S1212 (τ12 + τ21 ) (6.105)
(2) (2)
κ13 = S1313 µ13 κ23 = S2323 µ23 .

Finally, for the special type of orthotropy (symmetric, referred to in Section


6.5 below) we have two additional simplifications
(1) (1) (2) (2)
S1111 = S2222 S1313 = S2323 . (6.106)

Thus, the constitutive law for such an orthotropic and symmetric planar
(1)
couple-stress model involves four independent compliance components: S1111 ,
(1) (1) (2)
S1122 , S1212 , and S1313 .

6.3.2 Characteristic Lengths in Isotropic and Orthotropic Media


In the early 1960s when the Cosserat models began to undergo a revival fol-
lowing half a century of dormancy after invention by the Cosserats, several
people realized that, contrary to classical elasticity, an intrinsic length scale
was involved in the governing equations. It was denoted l, and called a char-
acteristic length. Let us now see how this l can be arrived at. Following Mindlin
(1963) and Schäfer (1962), we employ a stress function formulation, which for
the planar Cosserat (as well as the couple-stress) elasticity involves two stress
functions, φ and ψ

τ11 = φ,22 − ψ,12 τ22 = φ,11 + ψ,12


τ12 = −φ,12 − ψ,22 τ21 = −φ,12 + ψ,11 (6.107)
µ13 = ψ,1 µ23 = ψ,2 .

Note that φ is the Airy stress function known from the classical elastostatics.
Recall also that, for the isotropic planar Cosserat elasticity, the compatibility
conditions in terms of φ and ψ are given by (e.g., Nowacki, 1986)
   
P+S 2 A+ S 2 P+S 2 A+ S 2
ψ− ∇ ψ =− ∇ φ,2 ψ− ∇ ψ = ∇ φ,1 .
4M ,1 4M 4M ,2 4M
(6.108)

These are the Cauchy-Riemann conditions for the functions A+S4M


∇ 2 φ and
[ψ − 4M ∇ ψ], so that we actually have two harmonic functions
P+S 2

 
P+S 2
∇ 2∇ 2φ = 0 ∇ 2 ψ − ∇ ψ = 0. (6.109)
4M

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

Two- versus Three-Dimensional Micropolar Elasticity 213

The coefficient ( P + S)/4M appearing above has the dimension of length


squared and has thus led to a definition of characteristic length l via

P+S S(1)
l2 = ≡ 1212
(2)
. (6.110)
4M S1313

Note: In the couple-stress theory P = 0 in equations (6.107–6.109).


For the orthotropic Cosserat elasticity case, the compatibility conditions
(6.97) result in
(1) (1) (1) (1)
[S1111 − S1221 − S1122 ]ψ,122 + S2121 ψ,111
(1) (1) (1) (1) (2)
−[S1122 + S2121 + S1221 ]φ,112 − S1111 φ,222 = S1313 ψ,1

(1) (1) (1) (1)


[S2222 − S1221 − S1122 ]ψ,112 + S1212 ψ,222
(1) (1) (1) (1) (2)
+[S1122 + S1212 + S1221 ]φ,122 + S2222 φ,111 = S2323 ψ,2 , (6.111)

which suggest the following definitions of four characteristic lengths (Bouyge


et al., 2002):
 
 (1)  (1)
 S − S(1) − S(1) S
l1 =  1111 1221
(2)
1122
l2 =  2121
(2)
S1313 S1313
  (6.112)
 (1)  (1)
S (1)
− S1221 (1)
− S1122 S
l3 =  2222 (2)
l4 =  1212
(2)
.
S2323 S2323

In the special case of plane isotropic Cosserat elasticity, the following


relations hold
(1) (1) (1) (1) (1) (1) (1) (1)
S1111 − S1221 − S1122 = S2121 = S2222 − S1221 − S1122 = S1212 , (6.113)

and equations (6.112) reduce to a single characteristic length defined in (6.110).


For the planar orthotropic couple-stress case, the compatibility equations
(6.113) yield

(1) (1) (1) (1)


[S1111 − S1212 − S1122 ]ψ,122 + S1212 ψ,111
(1) (1) (1) (2)
− [S1122 + 2S1212 ]φ,112 − S1111 φ,222 = S1313 ψ,1

(1) (1) (1) (1)


[S2222 − S1212 − S1122 ]ψ,112 + S1212 ψ,222
(1) (1) (1) (2)
+ [S1122 + 2S1212 ]φ,122 + S2222 φ,111 = S2323 ψ,2 . (6.114)

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

214 Microstructural Randomness and Scaling in Mechanics of Materials

Thus, the four characteristic lengths are


 
 (1)  (1)
 S − S(1) − S(1) S
l1 =  1111 1212
(2)
1122
l2 =  1212
(2)
S1313 S1313
 
 (1)  (1)
S (1)
− S1212 (1)
− S1122 S
l3 =  2222 (2)
l4 =  1212
(2)
, (6.115)
S2323 S2323

whereupon equations (6.114) can be rewritten as

(1) (1) (1)


S1122 + 2S1212 S1111
l12 ψ,122 + l22 ψ,111 − (2)
φ,112 − (2)
φ,222 = ψ,1
S1313 S1313
(1) (1) (1)
S1122 + 2S1212 S2222
l32 ψ,112 + l42 ψ,222 + (2)
φ,122 + (2)
φ,111 = ψ,2 . (6.116)
S2323 S2323

For the special case of plane orthotropic couple-stress case with symmetry,
in view of (6.106), there are only two characteristic lengths
 
 (1)  (1)
 S − S(1) − S(1) S
l1 =  1111 1212
(2)
1122
l2 =  1212
(2)
. (6.117)
S1313 S1313

6.3.3 Restricted Continuum vis-à-vis the Micropolar Model


It is now recalled that a restricted continuum (couple-stress) Cosserat model,
with the limitation (6.38), was introduced in the past as a simplified, and
somewhat restricted, version of the general micropolar case. First, let us note
that, with the definition (6.117), the counterpart of equations (6.108) is

[ψ − l 2 ∇ 2 ψ],1 = −2 (1 − ν) l 2 ∇ 2 φ,2 [ψ − l 2 ∇ 2 ψ],2 = −2 (1 − ν) l 2 ∇ 2 φ,1 ,


(6.118)

and (6.109) holds with ( P + S) /4M replaced by S/4M. The length l appearing
in (6.118) is

S 2 (1 + ν) B
l2 = ≡ . (6.119)
4M E

where B is the modulus of curvature and µ, E, and ν are the shear modulus,
Young’s modulus, and Poisson’s ratio of classical elasticity, respectively.
To illustrate the distinction between both models we focus now on the
problem of a hole in an infinite body in plane strain under uniaxial tension p.
Mindlin (1963) found that in a couple-stress material the maximum stress σm

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

Two- versus Three-Dimensional Micropolar Elasticity 215

(or stress concentration factor)

3+ F
σm = p (6.120)
1+ F
where, in our notation,
4 (S + A) /S
F = (6.121)
4 + (a /l) + 2 (a /l) K 0 (a /l) /K 1 (a /l)
2

where l is defined by (6.119) and a is the hole radius; also, K 0 and K 1 are
the modified Bessel functions of the second kind of orders zero and one,
respectively.
On the other hand, somewhat later Kaloni and Ariman (1967) solved the
same problem for a micropolar elastic body with the result

4 (S + A) / (S + P)
F = (6.122)
4 + (a /l) 2 + 2 (a /l) K 0 (a /l) /K 1 (a /l)

where l is defined by (6.110).


Finally, we would like to point out that if we set

1
M= P = 0, (6.123)
4B
in (6.121) and other pertinent micropolar formulas, we recover Mindlin’s
couple-stress result. Also, when the details of the microstructure become
much smaller than the hole radius, i.e. l → 0, then F → 0, and σm → 3 p,
which recovers the classical elasticity result.
To see a continuous transition from the classical elasticity to both Cosserat
models, it is convenient at this stage to bring in, after Cowin (1969, 1970a, b),
a nondimensional constant
 
α S
N= = 0 ≤ N ≤ 1. (6.124)
α+µ S+ P

Then, following Cowin, F is given by

8 (1 − ν) N2
F = (6.125)
4 + ( NL) 2 + 2NL KK 01 (( NL)
NL)

whereby, in our notation,



A+ S a 4M
1−ν = L= =a . (6.126)
2S l S
Indeed, for this example, the case N = 0 leads to the classical elasticity solu-
tion, while N = 1 gives the limiting case of couple-stress theory (6.121). The
micropolar unrestricted theory follows for any value of N between 0 and 1.

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

216 Microstructural Randomness and Scaling in Mechanics of Materials

TABLE 6.1
A Comparison of Various Notations for Micropolar Compliances in the First
Planar Problem
Compliance Our Notation Nowacki Eringen Mindlin
1 1 2(1+ν)
Shear compliance 1/G S µ µ E +κ E /2 E
1 1 1−2ν
Plane strain bulk compliance A λ+µ λ E +µ E +κ E /2 G
1 2
Bulk compliance P α κE 0
1 1 1
Bending or curvature compliance M γ +ε γE 4B
P+S (µ+α)(γ +ε) γ E (µ E +κ E ) B
Characteristic length (square of) l 2 4M 4µα κ E (2µ E +κ E ) G

Note: Mindlin’s column refers to the restricted model.

However, in general, the limit N = 0 needs to be used with caution (Lakes,


1985) because the zero value of N does not automatically imply that all the
micropolar stiffnesses vanish (i.e., the compliances go to infinity), and, in fact,
the microstructural degrees of freedom may remain.
Finally, for the sake of reference we provide in Table 6.1 a comparison
between several different notations employed in the works referenced here.
A similar table linking the notation of other references is included in Cowin
(1970a).

6.4 The CLM Result and Stress-Invariance


6.4.1 Isotropic Materials
We now allow the micropolar solid to be inhomogeneous by taking all the
material coefficients in (6.99) to be class C 2 functions of x, and assume it is
simply connected (i.e., no holes are present). Now, note that the compatibility
equations (6.97) of Section 6.3 can be written as

γ22,11 + γ11,22 = (γ12 + γ21 ) ,12


 
γ12,22 − γ21,11 = (γ22 − γ11 ) 12 − κ13,1 − κ23,2 (6.127)
κ23,1 = κ13,2 .

Substituting (6.98) of Section 6.3 into the compatibility condition (6.127)1 ,


and using (6.95)1 and (6.95)2 of Section 6.3, we obtain, after rather lengthy
manipulations,

A+ S
∇ 2[ (τ11 + τ22 )] − [S,1 τ11 ],1 − [S,2 τ22 ],2 − [S,1 τ12 ],2 − [S,2 τ21 ],1 = 0.
2
(6.128)

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

Two- versus Three-Dimensional Micropolar Elasticity 217

Proceeding similarly with respect to the compatibility condition (6.127)2 , we


obtain, again after much algebra,

P+S
∇ 2[ (τ12 − τ21 )] + [S,1 τ11 ],2 − [S,2 τ22 ],1 − [S,1 τ12 ],1 − [S,2 τ21 ],2
2
+ 2 [Mµ13 ],1 + 2 [Mµ23 ],2 = 0. (6.129)

The third compatibility condition (6.127)3 yields

[Mµ23 ],1 − [Mµ13 ],2 = 0. (6.130)

Let us note here that (6.128) simplifies to equation (5.36) of Chapter 5 in


the case of classical elasticity where τ12 = τ21 . As before, we ask the question:
“Supposing that A, P, S, and M are changed to some A, P, S and M, then
under what restrictions would the original stress field τ11 , τ22 , τ12 , τ21 , µ13 , µ23
remain unchanged?” An examination of (6.95) implies that we must have

A + S = m ( A + S) S,1 = mS,1 S,2 = mS,2


(6.131)
S,11 = mS,11 S,22 = mS,22 S,12 = mS,12 ,

where m is an arbitrary scalar. Next, note that (6.131) leads to

A =m A + b S = mS − b, (6.132)

where b is an arbitrary constant restricted by the requirement that the com-


pliances be non-negative.
By a similar reasoning, the compatibility equation (6.127)2 implies

P = nP + c S = nS − c M = nM, (6.133)

where c is an arbitrary constant restricted by the requirement that the com-


pliances be non-negative.
Finally, the compatibility condition (6.127)3 involves a micropolar compli-
ance M and its first derivatives, and so we have the following conditions:

M = nM M,1 = nM,1 M,2 = nM,2 . (6.134)

Considering all the above results, it is seen that they are consistent pro-
viding m = n and b = c. It thus follows that the stress will be invariant if the
following shifts in material compliances are taken:

A =m A + c P = mP + c S = mS − c M = nM. (6.135)

In the terminology of the CLM result (Chapter 5), both materials are equivalent.
Clearly, equations (6.135) represent a constant shift in three out of four material
parameters, and this is a weaker shift than the linear one: (5.39) in Chapter 5.
Note: The second planar problem does not admit a shift.

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

218 Microstructural Randomness and Scaling in Mechanics of Materials

6.4.2 Anisotropic Materials and the Null-Lagrangian


In the case of an anisotropic material the starting point is provided by
relations (6.27). A substitution into the compatibility relations (6.127) and
a subsequent inspection leads us to conclude that the stress field remains
(1) (2)
unchanged when the material constants are modified from Sijkl and Sijkl to
(1) (1) (1) I (2) (2) (1) I
Sijkl = Sijkl + Sijkl (, −, ) and Sijkl = Sijkl , providing Sijkl (, −, ) is the
(2)
shift defined by (6.135) for isotropic materials, and Sijkl undergoes no shift.
(1)
Let us first note that the compliance Sijkl of an isotropic material is a fol-
lowing function of three constants κ, µ, and α:
   
(1) I 1 1 1 1 1 1 1
Sijkl (κ, µ, α) = + δik δjl + − δil δ jk + − δij δkl .
4 µ α µ α κ µ
(6.136)

Equivalently, noting (6.99), this can also be given in terms of A, P, S as

(1) I 1 
Sijkl ( A, P, S) = (S + P) δik δjl + (S − P) δil δ jk + ( A − S) δij δkl . (6.137)
4

If we let κ = , µ = −, and α =  in (6.136), it follows that the shift tensor


is given as

(1) I 1  
Sijkl (, −, ) = δij δkl − δil δ jk . (6.138)
2

Let us now compare this with the CLM shift tensor of classical elasticity
(5.49) with the rotation (5.50) of Chapter 5. This leads to a question: “What is
(1) I
the meaning of Sijkl (, −, )?” The answer is obtained from a consideration
of a new rotation tensor defined as

(1) I 1 (1)
R(1)
ijkl = δij δkl − δil δ jk Sijkl (, −, ) = R , (6.139)
2 ijkl

showing that R(1)ijkl σkl is also a right-angle rotation of τkl .


The foregoing development allows us to present the “shift-result” in terms
of the null-Lagrangian. First, let us recall from the theory of micropolar elas-
ticity (Nowacki, 1986) that two stress functions φ and ψ can be introduced
such that [recall (6.107)]

τij = R(1)
ijkl φ,kl − Rik ψ,k j µi3 = ψ,i , (6.140)

where Rik was specified in (5.54) of Chapter 5. The strain energy density

(1) (2) (1) (2)


W(Sijkl , Sijkl ) = τij Sijkl σkl + µi3 Si3k3 µk3 (6.141)

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

Two- versus Three-Dimensional Micropolar Elasticity 219

may now be written as


   
(1) (2) (1)  (1) 
W Sijkl , Sijkl = R(1)
ijkl φ,kl − Rik ψ ,k j Sijmn Rmnpq φ, pq − Rmp ψ, pm
(2)
+ µi3 Si3k3 µk3 . (6.142)

First, let us note that the minimization of (6.142) via the Euler–Lagrange
equations of (6.141) results in the compatibility equations in terms of φ and
ψ for a general anisotropy. On the other hand, the energy density of the shift
in compliance becomes
 
(1) I
W Sijkl (, −, ) , Oijkl
 
(1) I  (1) 
= R(1) ijkl φ ,kl − Rik ψ ,k j Sijmn Rmnpq φ, pq − Rmp ψ, pm , (6.143)

where Oijkl is a null tensor. Now, observing that

R(1) (1) (1)


klij Rijmn Rmnpq = Rklpq
(1)
R(1) (1)
ijkl Rijmn = δkm δln , (6.144)

we find
  1  2
(1) I
W Sijkl (, −, ) , Oijkl = φ,11 φ,22 − φ,12
2
   
+ ψ,12 φ,22 − φ,11 + φ,12 ψ,22 − ψ,11 . (6.145)

It is interesting to note here that:

1. The first term in the square brackets is the same as that in the classical
elasticity [our (5.59) in Chapter 5, or equation (35) of Cherkaev
et al. (1992)].
2. The second and third terms represent the coupled contribution of φ
and ψ potentials.
3. The energy (6.145) can also be written as the divergence of a vector
field v,k such that

1 
vk = φ,l R(1)
klpq φ, pq − φ,l Rkr ψ,rl − φ,i j Rik ψ, j
2 
 
+ ψ, j R jk Rnr − Rnk Rjr ψ,rn
  
1 φ,1 φ,22 − φ,2 φ,12 − φ,1 ψ,12 − φ,2 ψ,22 + φ,21 ψ,1 + φ,22 ψ,2
=   ,
2 φ,2 φ,11 − φ,1 φ,12 + φ,1 ψ,11 + φ,2 ψ,12 − φ,11 ψ,1 − φ,12 ψ,2
(6.146)

where, again, the first term in each of the square brackets can be
recognized as that of classical elasticity. It follows now that vk,k = 0,

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

220 Microstructural Randomness and Scaling in Mechanics of Materials

(1) I
or that the Euler–Lagrange equations for W(Sijkl (, −, ), Oijkl )
are satisfied identically, implying that it is a null-Lagrangian.

6.4.3 Multiply Connected Materials


The requirement that B be simply connected ensures that the stress field σ is
single valued. Generalizing the results of Section 5.2.3 of Chapter 5, we have
 
∂ϕ3 ∂ϕ3
3 = d x1 + d x2
∂ x1 ∂ x2
   
∂ε21 ∂ε11 ∂ε22 ∂ε12
= − d x1 + − d x2 , (6.147)
∂ x1 ∂ x2 ∂ x1 ∂ x2

 
∂ε11 ∂ε11
D1 + y0 3 = − x1 d x1 + d x2
∂ x1 ∂ x2
   
∂ε11 ∂ε22 ∂ε12 ∂ε21
− x2 d x1 − − − d x2 , (6.148)
∂ x2 ∂ x1 ∂x2 ∂x2

 
∂ε22 ∂ε22
D2 − x0 3 = − x2 d x1 + d x2
∂ x1 ∂x2
   
∂ε22 ∂ε12 ∂ε21 ∂ε11
− x1 d x2 + + − d x1 , (6.149)
∂ x1 ∂x1 ∂ x1 ∂ x2

where D1 and D2 are dislocation vectors and 3 is a disclination. (Nowacki,


1986; Takeuti, 1973). If, using (6.98), we express the above equations in terms
of stresses, they take on the following forms:
  
∂ ∂S ∂S
3 = [( A + S) (σ11 + σ22 )] ds = 2 t1 ds − 2 t2 ds
∂n ∂x1 ∂ x2
 
∂ ∂
+ (S + P) (σ21 − σ12 ) ds + (σ12 − σ21 ) (S − P) ds, (6.150)
∂s ∂s

  
∂ ∂
4 ( D1 + y0 3 ) = x2 − x1 ( A + S) (σ11 + σ22 ) ds
∂n ∂s
  
∂S ∂S
=2 St2 ds − 2 x2 t1 + t2 ds
∂ x1 ∂ x2

∂S
−2 x2 (σ21 − σ12 ) ds, (6.151)
∂s

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

Two- versus Three-Dimensional Micropolar Elasticity 221

  
∂ ∂
4 ( D2 − x0 3 ) = x1 + x2 ( A + S) (σ11 + σ22 ) ds
∂n ∂s
  
∂S ∂S
=2 St1 ds + 2 x1 t1 + t2 ds
∂ x1 ∂ x2

∂S
+2 x1 (σ21 − σ12 ) ds, (6.152)
∂s

where n and s denote the outer unit normal and arc length of the hole
boundary.
With reference to Section 6.3.3, for the problem of an infinite plate with
hole, we have these conclusions:
1. For classical elasticity, σm as well as the entire stress field are inde-
pendent of elastic constants, say A and S (or µ and ν).
2. For a pseudo-continuum which has three constants— A, S, and M
(or µ, ν and B)—the stress field depends on two combinations of
these constants, such as ( A + S) /S (or ν) and l 2 = S/4M, and thus
no shift is possible here.
3. For an unrestricted continuum, which has four constants, A, S, P,
and M, the dependence is on two independent combinations of the
elastic constants ( A + S) / ( P + S) and l 2 = (S + P) /4M, which, in
light of (6.134), allow a shift.
Note: Setting P = 0, we get the pseudo-continuum. In this case S + P
becomes S, and S by itself is not invariant under shift.

6.4.4 Applications to Composites


6.4.4.1 Two-Phase Materials
We continue to generalize the results of Chapter 5. Thus, when the planar
body is made up of two or more phases, we must also consider the inter-
face boundary conditions. Assuming perfect bounding between micropolar
phases (1 and 2), they have the following (classical) form in the curvilinear
coordinate system (n, s, x3 ):

τnn
(1)
= τnn
(2)
τns
(1)
= τns
(2)
µ(1) (2)
n3 = µn3
(6.153)
n = un
u(1) (2)
s = us
u(1) (2)
ϕ3(1) = ϕ3(2) .

Alternately, using the boundary conditions proposed by Dundurs (1996),


we have
τnn
(1)
= τnn
(2)
τns
(1)
= τns
(2)
µ(1) (2)
n3 = µn3

∂ϕ3
(1)
∂ϕ3
(2) (6.154)
κn(1) = κn(2) γss(1) = γss(2) = ,
∂s ∂s

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

222 Microstructural Randomness and Scaling in Mechanics of Materials

where κn is the change in curvature of boundary curve


∂ ∂γss
κn = (γns + γsn ) − − κγnn , (6.155)
∂s ∂n
and γss is the stretch strain. The advantage of using this second set of boundary
conditions is that they can be expressed in terns of stresses, and thus the
dependence of the solution on the micropolar constants can be seen more
easily. Using the constitutive relations (6.98), which remain of the same form
in (n, s, x3 ), and the equilibrium equation
∂τnn 1 ∂
+ (τns + τsn ) + κ (τss − τnn ) = 0, (6.156)
∂n 2 ∂s
the boundary condition (6.154)4 , which implies the continuity of a change in
curvature, becomes, in view of (6.155) and again (6.98),

∂   ∂   ∂τ (1) ∂τ (2)
( A2 + S2 ) τss(2) − ( A1 + S1 ) τss(1) + sn ( A1 + S1 ) − sn ( A2 + S2 )
∂n ∂n ∂s ∂s
∂ S1 ∂ S2 ∂ ∂τns
(1)
+ 2τsn
(1)
− 2τsn
(2)
+ 2τns
(1)
(S1 − S2 ) − 2 (S1 − S2 )
∂s ∂s ∂s ∂s

+ τnn
(1)
[( A2 − A1 ) − (S2 − S1 ) + 2 ( A2 − A1 ) κ] = 0. (6.157)
∂n
Now, taking note of (6.99), the continuity of stretch strain (6.154)5 implies

( A2 + S2 ) τss(2) − ( A1 + S1 ) τss(1) + τnn


(1)
( A2 − A1 ) − (S2 − S1 ) = 0. (6.158)

Finally, noting ∂ϕ3 /∂s = κs3 , we observe that (6.154)6 implies

M1 µ(1) (2)
s3 − M2 µs3 = 0. (6.159)

Note that these boundary conditions are invariant under the shift (6.135).
Thus, if the multiphase material is simply connected (i.e., contains intrusions),
the governing equations in terms of stresses are (6.95) and (6.128–6.130) for
each phase, and these are invariant under traction loading and boundary
conditions (6.154), or, equivalently, (6.154)1−3 and (6.157–6.159). However, if
the material is multiply connected, we also need Cesàro integrals that involve
the continuity of displacements.

6.4.4.2 Effective Moduli of Composites


We now move to the determination of effective properties of composites,
which in the case of classical elasticity turned out to nicely involve the CLM
result. We thus examine the shift in macroscopically effective compliance
(1)eff (2)eff
tensors Sijkl and Sijkl connecting the volume averaged stress and strain
tensors
(1) (2)
ε ij = Sijkl σ kl κ ij = Sijkl µkl , (6.160)

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

Two- versus Three-Dimensional Micropolar Elasticity 223

wherein denotes a volume average. We proceed here in a fashion analogous to


that of Cherkaev et al. (1992) with a restriction to materials with continuously
varying (class C 2 ) properties.
Now, stresses τ11 , τ22 , τ12 , τ21 , µ13 , µ23 are the same in two equivalent
(1) (2) (1) (2)
materials (Sijkl (x), Sijkl (x)) and ( Sijkl (x), Sijkl (x)), so that the strain fields γij and
γij satisfy the relation
(1) (1) (1) I (1) I
γij = Sijkl τkl = Sijkl τkl + Sijkl (, −, ) τkl = γij + Sijkl (, −, ) τkl .
(6.161)
(1) I
Volume averaging (6.161), and noting that Sijkl (, −, ) is independent of
position, we find
(1)eff (1)eff
(1) I
γ ij = Si jkl τ kl = Si jkl τ̄kl + Sijkl (, −, ) τ̄kl . (6.162)

which shows that the effective compliance tensor of the second material (with
hat) is given by that of the first material (without hat) plus the shift given by
(6.138)
(1)eff (1)eff (1) I
Sijkl = Sijkl + Sijkl (, −, ) . (6.163)

We conclude (by inspection) that there is no shift in the second effective com-
pliance tensor and
(2)eff (2)eff
Sijkl = Sijkl . (6.164)

As mentioned at the beginning of this section, this conclusion holds for simply
connected inhomogeneous media with twice-differentiable properties.

6.4.5 Extensions of Stress Invariance to Presence


of Eigenstrains and Eigencurvatures
6.4.5.1 Basic Concepts
Here we extend the results of Section 5.2.3, Chapter 5. First note that one
can generalize the elasticity with eigenstrains to Cosserat elasticity. Thus, in
analogy to (5.74), the total strain γij is the sum of the elastic strain gij and the
eigenstrain γij∗ ,

γij = gij + γij∗ (6.165)

while the total (generally nonsymmetric) curvature κij is the sum of the elastic
curvature kij and the eigencurvature κij∗

κij = kij + κij∗ (6.166)

The eigenstrain γij∗ and eigencurvature κij∗ are inelastic strains and curva-
tures, respectively. The total strain γij and total curvature κij must satisfy

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

224 Microstructural Randomness and Scaling in Mechanics of Materials

compatibility equations

γli,n − γni,l − kni κlk + kli κnk = 0 κli,n = κni,l . (6.167)

From (6.165 and 6.166) we have


(1)   (2)  
(1)
τij = Cijkl gkl = Cijkl γkl − γkl∗ (2)
µij = Cijkl kkl = Cijkl κkl − κkl∗ . (6.168)

The inverses of (6.168) are


(1) (2)
gij = Sijkl τkl kij = Sijkl µkl . (6.169)

Note that all the quantities may depend on the spatial position x (≡ xi ).
We extend the assumption that the material is free from any external forces
and surface constraints to Cosserat elasticity with eigenstrains and eigencur-
vatures. If these conditions of free surface are not satisfied, the force-stress
and couple-stress fields can be obtained by a superposition of the force-stress
and couple-stress of a free body and the stress obtained from the solution of
a given boundary value problem with non zero external forces or boundary
conditions.
The force-stresses and couple-stresses must satisfy the equations of equi-
librium [assume no body and inertia forces in (6.12) and (6.17)] in B

τji,j = 0 ijk τ jk + µji,j = 0 i, j = 1, 2, 3, (6.170)

and zero force-traction or couple-traction free boundary conditions on ∂B

τji n j = 0 µji n j = 0. (6.171)

In analogy to Mura’s approach in classical elasticity, by substituting (6.168)


into (6.170) and assuming a homogeneous material, we have
(1) (1) ∗ ∗ (1)  ∗

Cjikl γkl, j = Cjikl γkl, j C (2) (2)
jimn κmn, j = C jimn κmn, j − ijk C jkmn γmn − γmn ,
(6.172)

and by substituting (6.168) into (6.171) we obtain


(1) (1) ∗ (2) (2) ∗
Cjikl γkl n j = Cjikl γkl n j Cjikl κkl n j = Cjikl κkl n j . (6.173)

Note that in the absence of eigenstrains (γ ∗ = 0), the left-hand side of


(6.172)1 corresponds to τji,j , and the left-hand side of (6.173)1 to τji n j . Thus,

(6.172)1 , is in the form τji,j = −Xi where Xi = −Cjikl γkl, j and (6.173)1 is in the

form τji n j = ti where ti = Cjikl γkl n j . Therefore, the contribution of eigenstrain
γ ∗ to the equations of equilibrium (6.170) is mathematically equivalent to a
body force, and in the boundary conditions (6.171)1 is similar to a surface
force.
Next, in the absence of eigencurvatures (κ ∗ = 0) the left-hand side of
(6.172)2 corresponds to ijk τjk + µji,j , whereas the left-hand side of (6.173)2 to
µji n j . Thus, (6.172)2 is in the form ijk τjk + µji,j = −Yi where the couple-body

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

Two- versus Three-Dimensional Micropolar Elasticity 225

∗ ∗
force Yi = −Cijkl γkl, j and (6.173)2 is in the form τji n j = ti where ti = C jikl γkl n j .

Therefore, the contribution of eigenstrain γij to the equations of equilibrium
(6.170)2 is mathematically equivalent to a body force, while their contribution
to the boundary conditions (6.171)2 is similar to a surface force.
In the next sections we focus on the planar elasticity with eigenstrains,
assuming isotropy in elastic properties. In addition, we relax the bound-
ary condition (6.171) and admit non-zero tractions to make the formulation
more general. This will not change our conclusions on the reduced parameter
dependence.
Note that the special case of uncoupled micropolar thermoelasticity with
eigenstrains εij∗ is defined as

γij∗ = αij T αij = 0 if i=


 j i, j = 1, 2, 3, (6.174)

where αij is a thermal expansion coefficient and T is temperature change.


We will refer to this special case in examples.
We demonstrated the reduced parameter dependence in the in-plane stress
fields in the problems governed by plane Cosserat elasticity with eigenstrains
and eigencurvatures, if eigenstrains and curvatures satisfy certain conditions.
Note that no conditions are needed for the plane stress case for the form
of shift with m = 1. These results can be applied for two-phase materials
to linear planar uncoupled micropolar thermoelasticity, where eigenstrains
are uniform and represent the product of the thermal expansion coefficient
and temperature change. The analysis can also be extended to multiphase
materials and inhomogeneous multiply connected materials.

6.4.5.2 Inhomogeneous Materials


Here we focus on 2D boundary value problems involving applied force- and
couple-tractions

ti = τji n j mi = µji n j on ∂B. (6.175)

It is also useful to write the constitutive laws (recall Section 6.3.1)

A+ S S ∗ ∗
γ11 = (τ11 + τ22 ) − τ22 + γ11 + ηγ33
4 2
A+ S S ∗ ∗
γ22 = (σ11 + σ22 ) − τ11 + γ22 + ηγ33
4 2
S P ∗
γ12 = (τ12 + τ21 ) + (τ12 − τ21 ) + γ12 (6.176)
4 4
S P ∗
γ21 = (τ12 + τ21 ) − (τ12 − τ21 ) + γ21
4 4
∗ ∗
κ13 = Mµ13 + κ13 κ23 = Mµ23 + κ23 ,

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

226 Microstructural Randomness and Scaling in Mechanics of Materials

where
1 1 1
S= P= M= , (6.177)
µ α γ +ε
and for plane strain
1 1 λ
A= = , η= , (6.178)
κplane-strain λ+µ 3K − λ
while for plane stress
1 µ (3λ + 2µ)
A= = , η = 0. (6.179)
κplane-stress λ + 2µ
If we express λ in terms of Poisson’s ratio v
2vµ
λ= , (6.180)
1 − 2v
then, for plane strain

1 1 − 2v 1 A
A= = , η=v= 1− , (6.181)
κplane-strain µ 2 S
while, for plane stress
1 1−v
A= = , η = 0. (6.182)
κplane-stress (1 + v) µ
Now, with the assumption of both compliances and eigenstrains being
smooth functions of position, it follows from the substitution of (6.176) into
(6.127)1 , and in light of (6.171)1 , that the first compatibility condition in
(6.127) is
1 2 1 1 1
∇ [( A + S) (τ11 + τ22 )] − [S,1 τ11 ],1 − [S,2 τ22 ],2 − [S,1 τ12 ],2
4 2 2 2
1 ∗ ∗
 ∗ ∗
 ∗
− [S,2 τ21 ],1 = −γ11,22 − γ22,11 + γ12 + γ21 ,12
− ∇ 2 ηγ33
2
∗ ∗ ∗
−2η,1 γ33,1 − 2η,2 γ33,2 − η∇ 2 γ33 . (6.183)

Similarly, the compatibility condition (6.127)2 yields


1 2 1 1 1
∇ [( P + S)] + (S,1 τ11 ) ,2 − (S,2 τ22 ) ,1 − (S,1 τ12 ) ,1
4 2 2 2
1 ∗ ∗ ∗ ∗
+ (S,2 τ21 ) ,2 = −γ11,12 + γ22,12 − γ12,22 + γ21,11
2
∗ ∗
− ( Mµ13 ) ,1 − κ13,1 − ( Mµ23 ) ,2 − 2κ23,2 . (6.184)

Finally, from the compatibility condition (6.127)3 we have


∗ ∗
( Mµ23 ) ,1 + κ23,1 = ( Mµ13 ) ,2 + κ13,2 . (6.185)

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

Two- versus Three-Dimensional Micropolar Elasticity 227

Next, following Ostoja-Starzewski and Jasiuk (1995), we seek the condi-


tions for invariance in planar force-stresses and couple-stresses with respect
to shift in compliances. We first state the conditions for no eigenstrains and
eigencurvatures. Equations (6.183–6.185) remain invariant if the compliances
A, P and S undergo a shift

A =m A + c P = nP + c S = mS − c M = mM. (6.186)

Under such a linear shift the force-stress does not change in the absence of
eigenstrains and eigencurvatures. Next, we investigate what conditions are
needed to be satisfied in the presence of eigenstrains and eigencurvatures. In
this analysis, in addition to the plane stress and plane strain cases, which lead
to different results, the distinction is made between the cases when m = 1 and
m=  1.
For the plane stress case and m = 1, (6.183–6.185) remain unchanged under
the linear shift (6.186), i.e., the planar stress components remain unchanged,
and thus there is a reduced parameter dependence.
For the plane stress case and m =  1, (6.183–6.185) give the following
conditions on eigenstrains and eigencurvatures:
∗ ∗
 ∗ ∗

γ11,22 + γ22,11 − γ12 + γ21 ,12
=0
∗ ∗ ∗ ∗ ∗ ∗
γ11,12 − γ22,12 + γ12,22 + γ21,11 + κ13,1 + κ23,2 =0 (6.187)
∗ ∗
κ23,2 = κ13,1 .

For the plane strain case and m = 1, (6.183) remains invariant under the
shift (6.186) when
∗ ∗ ∗ ∗
∇ 2 ηγ33 + 2η,1 γ33,1 + 2ηγ33,2 + η∇ 2 ε33 = 0. (6.188)

For the special case of uniform eigenstrains, the condition above is satisfied
provided that

γ33 =0 or ∇ 2 η = 0, (6.189)

while for the case of a homogeneous material the condition (6.188) is


satisfied if

∇ 2 γ33 =0 or η = 0. (6.190)

For the case of plane strain and m =


 1, the following conditions need to be
satisfied
∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗
γ11,22 + γ22,11 − (γ12 + γ21 ) ,12 + ∇ 2 ηγ33 + 2η,1 γ33,1 − 2η,2 γ33,2 + η∇ 2 γ33 =0
∗ ∗ ∗ ∗ ∗ ∗
(γ11 − γ22 ) ,12 + γ12,22 + γ21,11 + κ13,1 + κ23,2 =0 (6.191)
∗ ∗
κ23,1 = κ13,2 .

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

228 Microstructural Randomness and Scaling in Mechanics of Materials

6.5 Effective Micropolar Moduli and Characteristic Lengths


of Composites
6.5.1 From a Heterogeneous Cauchy to a Homogeneous
Cosserat Continuum
Any of the lattices considered in Chapter 3 can be viewed as a planar, two-
phase composite material: one phase (1) is the solid that makes up the beams
and another (2), quite trivially, is the vacuum in the pores. Clearly then, the
mismatch of elastic properties, that is the ratio of moduli E (1) /E (2) , is in-
finite (Ostoja-Starzewski et al., 1999). Starting from this consideration, one
may now consider “nontrivial” two-phase composites made of two kinds of
solids having a finite mismatch, and generalizing the previously established
method (i.e., that for lattices) for passage from heterogeneous Cauchy to a
homogeneous Cosserat continuum; Fig. 6.3. This passage is done according
to the following equality:

1 V  0 (1) 0 0 (2) 0

εij Cijkl εkl dV = γ C γ + κi3 Ci3k3 κk3 , i, j, k, l = 1, 2, (6.192)
2 V 2 ij ijkl kl

where the left-hand side is the total elastic strain energy stored in the unit
cell of the matrix-inclusion composite (a function of Cauchy strain fields εij ),
while the right-hand side is the energy of a Cosserat continuum (a function of
volume-average strains εij0 and curvatures κi3 0
of the unit cell). V is the volume
of the unit cell B L . Cijkl is the elastic stiffness of the composites’ constituents,
(1) (2)
while Cijkl and Ci3k3 are the sought (effective) micropolar stiffnesses.
The key issue concerns the choice of loading on the periodic unit cell. Fol-
lowing Forest (1989, 1999) and Forest and Sab (1998), the appropriate periodic
boundary conditions in terms of displacements are generally nonlinear:

ui (x) = Ai + Bij x j + Cijk x j xk + Dijkl x j xk xl (6.193)

(a) (b) (c)

FIGURE 6.3
(a) A periodic, globally orthotropic, matrix-inclusion composite, of period L, with inclusions of
diameter d arranged in a square array; (b) a periodic unit cell with soft inclusions at corners;
(c) a periodic unit cell with a stiff inclusion at the center.

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

Two- versus Three-Dimensional Micropolar Elasticity 229

where vi takes an equal value on corresponding points of the boundary. The


field (6.193) satisfies the Hill condition generalized to the Cosserat material
2
εij σij = γij τij + µij κij with τij = σij0 µij = e ikl xk σl j . (6.194)
3
In the notation and language of homogenization theory (e.g., Bensoussan
et al., 1978), the Cauchy material in the unit cell is described by a fast co-
ordinate x, while the Cosserat material is described by the slow coordinate y.
Accordingly, the displacement, stress and strain fields in the Cauchy conti-
nuum are
u σ ε, (6.195)

while the fields of displacement, rotation, force-stress, couple-stress, strain,


and rotation in the Cosserat continuum are

U   M E K. (6.196)

We now rewrite (6.194) as


2
εij σij = ij E ij + Mij K ij with ij = σij0 Mij = e ikl xk σl j . (6.197)
3
In the general setting, we have a micromorphic continuum with U (dis-
placement) and χ (microdeformation) fields, and these are determined from
a minimization problem
 
(U, χ ) = min |u (x) − U − χ · (x − X) |2 (6.198)

The solution is

U (X) = u (x) χ (x) = u⊗ (x − X) · A−1 , (6.199)

with A = (x − X) ⊗ (x − X) and respective gradients being

U ⊗ ∇X = u ⊗ ∇ x χ ⊗ ∇x = (u ⊗ x) ⊗ ∇x · A−1 − U ⊗ A−T . (6.200)

Although none of our boundary conditions were of periodic type, the situ-
ation changes when an unrestricted model is used. Indeed, such a derivation
has been done in Forest and Sab (1998) by extending the homogenization
method (e.g., Sanchez-Palencia and Zaoui, 1987). The loading on ∂ B L in 2D is
then effected by boundary conditions involving polynomials of the general
form
 
u1 (x) = B11 x1 + B12 x2 − C23 x22 + 2C13 x1 x2 + D12 x23 − 3x12 x2
 
u2 (x) = B12 x1 + B22 x2 − C13 x12 + 2C23 x1 x2 − D12 x13 − 3x1 x22 . (6.201)

Upon a comparison of this with equations (6.193–6.196), we observe that the


derivation of the restricted model involves a second-order polynomial, while
that of the unrestricted one requires a third-order polynomial. Work has also

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

230 Microstructural Randomness and Scaling in Mechanics of Materials

been done on homogenization of a heterogeneous Cosserat-type continuum


by a homogeneous one (Forest, 1999; Forest et al., 1999); see also Forest et al.
(2000, 2001).
Most recently, a related homogenization procedure was outlined by Onck
(2002) for the derivation of the micropolar model. In particular, a loading
via a skew-symmetric part of the strain tensor, applied in terms of bound-
ary rotations, was proposed to grasp the effect of the difference between the
micro-rotation ϕi and the macrorotation e ijk uk, j /2; recall the restricted model’s
condition (6.37), where that difference is set to zero.
Note: A major outstanding challenge is to develop boundary conditions for the
homogeneous Cosserat from a heterogeneous Cauchy continuum. This should
be done through a scheme somewhat analogous to that developed so far for
the effective micropolar moduli.

6.5.2 Applications
A few years ago we computed effective micropolar moduli for planar matrix-
inclusion composites arranged in periodic arrays: triangular (Bouyge et al.,
2001) and square (Bouyge et al., 2002), Figure 6.4, using a finite element
method. Several different boundary conditions—ranging from displacement-
type to traction-type, and various combinations thereof—were used. For
(1)
example, using displacement boundary conditions, we determine Cijkl from
three tests:
1. Uniaxial extension:

u1 (x) = 0 u2 (x) = ε22 x2 ∀x ∈ ∂ B (6.202)


(1)
gives C2222 (= 2U cell /V when we set ε22 = 1). For our composite,
(1) (1)
C1111 = C2222 due to the symmetry of the square arrangement.
2. Biaxial extension:

u1 (x) = ε11 x1 u2 (x) = ε22 x2 ∀x ∈ ∂ B (6.203)


(1) (1)
yields 2C1111 + C1122 (= 2U cell /V when we set ε11 = ε22 = 1).
3. Shear test:

u1 (x) = ε12 x2 u2 (x) = 0 ∀x ∈ ∂ B (6.204)


(1)
yields C1212 (= 2U cell /V when we set ε12 = 1).
(2)
4. Finally, to determine Ci3k3 , we conduct the fourth test, the bending
test:
x12
u1 (x) = −κ13 x1 x2 u2 (x) = κ13 ∀x ∈ ∂ B (6.205)
2
(2) (2) (2)
gives C1313 . Note that in our study C1313 = C2323 due to the symmetry
of square arrangement.

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

Two- versus Three-Dimensional Micropolar Elasticity 231

u1 (x) = 0, u2 (x) = y22x2

u1 (x) = y11x1, u2 (x) = y22x2

u1 (x) = y12 x2, u2 (x) = 0

x12
u1 (x) = –x1x2κ13, u2 (x) = κ
2 13

FIGURE 6.4
(1) (1) (1) (2)
Tests for the determination of constants C2222 , C1122 , C1212 , and C1313 of a periodic composite with
circular inclusions in a square arrangement under displacement boundary conditions (Bouyge
et al., 2002). Left (right) column corresponds to the inclusion at the corner (center). Inclusions
can be seen from the mesh pattern.

The resulting deformation modes for the above four tests under displace-
ment boundary conditions are shown in Figure 6.4. Two distinct situations are
considered here depending on whether the inclusion is softer or stiffer than
the matrix. In the first case, the inclusion is located at the corner, whereas in
the second it is located at the center. Typical results for effective moduli are
(1)
shown in Figure 6.5 in terms of C1212 for a wide range of Poisson’s ratio of the
matrix. In the special case of no mismatch in the properties we recover a ho-
mogeneous medium of Cauchy type, whereby the composite microstructure

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

232 Microstructural Randomness and Scaling in Mechanics of Materials

0.6

0.5

0.4

dd
C1212

0.3 dp
tt

0.2

0.1

0
0.0001 0.001 0.01 0.1 10 100 1000 10000
Mismatch

FIGURE 6.5
(1)
The effective moduli C1212 , normalized by E m , from three types of boundary conditions—
displacement (dd) displacement-periodic (dp), and traction (tt)—plotted as functions of the stiff-
ness ratio E i /E m for the case of Poisson’s ratios ν m = ν i = 0.3, at inclusions volume fraction of
18.4%. (From Bouyge et al., 2002. With permisssion.)

disappears and no Cosserat continuum is to be set up. Note that when the
inclusion is softer, as well as stiffer, than the matrix, the micropolar model
provides a better representation of the mechanics of the composite than the
classical model. Indeed, this was brought out by the experiments of Mora and
Waas (2000) on honeycombs with either porous or very stiff inclusions.
In the case of traction boundary conditions, we use


1 V  0 (1) 0 0 (2)

σij Sijkl σkl dV = τij Sijkl τkl + µi3 Si3k3 µ0k3 , i, j, k, l = 1, 2, (6.206)
2 V 2

where on the left we have the total complementary strain energy in the unit cell
(a function of Cauchy stresses σij ), and on the right we have the complemen-
tary strain energy of a couple-stress continuum (a function of volume-average
stresses σij0 and couple-stresses µi3
0
of the unit cell). Here Sijkl (inverse of Cijkl ) is
(1) (2)
the microscale elastic compliance, while Sijkl and Si3k3 are the sought effective,
micropolar compliances.
Summing up, for the restricted (or Koiter) model of the composite, the
micropolar moduli are bounded from above and below, respectively, by dis-
placement and traction boundary conditions. In fact, as these bounds are
wide, we recommend three mixed types of loadings to get tighter results.
On the other hand, the characteristic lengths are highly insensitive to the

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

Two- versus Three-Dimensional Micropolar Elasticity 233

L L L
u2 u2 u2 u2 u2 u2
–1. E-004
–9.0

–9.0E
–1.0

–8
0E-

3 3

.0
3 0
-00 -00 -0

E-
003

E-00
.0E .0E .0E

- 00 4

00
–1 1 –1

4
4 4
– -00 –7.0 -00

3
04
–8

E-0 –8 .0E E-0 .0E


.0 –9 04 –9
.0 E

–9 . 0E
4 004 4
-0

–7 -00 -0
0E-
–6.0 -00
04

E-00 0E
.0E .0E –7 04
–8. 4 –8.
-0 –8 . 0E
04 -0 –5.0
E-00
–6

4
04
– 6 -004

4
–6
4

00

-00

04
04
.0E

–4.

.0E
00

-0
-0
.0E
-0

0E-

.0E
.0E

E-

.0E
.0 E
–4

-00
0
04

04
.0

–7
.0E

–7
–7

–6
–6
4
-00
4
Cosserat Cauchy Cosserat Cauchy Cosserat Cauchy
ℓ/L = 2/100 ℓ/L = 2/10 ℓ/L = 2/1

FIGURE 6.6
Contour lines of vertical components of displacement in a panel loaded as shown, for various
scale ratios l/L, where l is the brick length and L is the macroscopic load print. (From Trovalusci
and Masiani, 2003. With permission).

mismatch in moduli, especially in the case of stiff inclusions, and this must
be contrasted with the sensitivity of moduli.
An interesting comparison of boundary value problems set up on struc-
tures made of Cauchy vis-à-vis Cosserat materials was conducted by
Trovalusci and Masiani (2003). Their research is motivated by the mechanics
of block-type masonry structures, whose stability (and, therefore, safety) is of
primary concern in places rich with ancient architecture like Italy and Greece.
Figures 6.6 and 6.7 show comparisons of symmetric boundary value problems

L L L
W12 (skw H)12 W12 (skw H)12 W12 (skw H)12

06 06 06
E-0 E-0 E-0
2.00 2.00 06 2.00
–2

–2

–1.0E-0
.0

.0E

-006 06 -006
E-

1.00E 1.00E-0 1.00E


-0
00

–1.0E-0

06
6

007
–1.0E-0

E-
–5.0
06

Cosserat Cauchy Cosserat Cauchy Cosserat Cauchy


06

–1.0E-0

–1.0E-0
–2.0

–5
–2.

06 6
06

-0 00
.0E

E-
-0
0E-

0E
E-00

6 0 006
E

6
06

1.0 -00 00
-00

1.0 0E-
00

–1.0E
06

0E 0E-
006

-006 2.0
1.

2.0 2.0
7
6

ℓ/L = 2/100 ℓ/L = 2/10 ℓ/L = 2/1

FIGURE 6.7
Contour lines of components of microrotation (Cosserat model) and macrorotation (Cauchy
model) in a panel loaded as shown, for various scale ratios l/L, where l is the brick length and
L is the macroscopic load print.

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

234 Microstructural Randomness and Scaling in Mechanics of Materials

(a)

t
d

(b)

FIGURE 6.8
(a) Microstructure of trabecular bone at the so-called mesoscale level of Figure 7.15 in Chapter 7;
(b) Unit cell of an idealized periodic model following Gibson and Ashby (1988). (From Yoo and
Jasiuk, 2006. With permission).

of a plate-type structure, modeled in either of two fashions. It is clear that,


as the ratio l/L increases, the discrepancies between both models tend to
increase.
Another example of determination of micropolar moduli has recently been
developed in studies of trabecular bone (Yoo and Jasiuk, 2006), see Figure 6.8.
It appears to be a first-ever prediction and evaluation of apparent couple-
stress moduli for a 3D periodic orthotropic material.

© 2008 by Taylor & Francis Group, LLC


P1: Binod
May 28, 2007 18:6 C4174 C4174˙C006

Two- versus Three-Dimensional Micropolar Elasticity 235

Problems
1. Derive the local equations of motion of a micropolar continuum
from the global equations of conservation of linear and angular
momenta.
2. Generalize the Mohr circle concept and analysis to plane stress mi-
cropolar elasticity.
3. Prove the inequalities (6.32). Hint: use the Sylvester theorem.
4. Determine the convolution operation involved in the equation (6.60).
5. With reference to equations (6.72), demonstrate that ,  and H are
dispersive waves.
6. Examine the implications of equation (6.79).
7. With reference to equation (6.80), consider an isotropic, hemitropic
medium. Develop the corresponding form of the free energy func-
tion and obtain restrictions on all the elastic constants, more general
than those of (6.32). Hint: introduce three new elastic constants.
8. Extend Kirchhoff’s uniqueness proof from the setting of linear elas-
ticity to linear micropolar elasticity.
9. Outline a theory of micropolar media for finite motions and strains.
10. Formulate the Clausius-Duhem inequality for micropolar elastic-
dissipative solids, and then outline a formulation of thermomechan-
ics with internal variables. Consult other sources as necessary.

© 2008 by Taylor & Francis Group, LLC

You might also like