You are on page 1of 19

Elastic Constant

Elastic constants are the constants describing mechanical response of a material


when it is elastic.

From: Encyclopedia of Condensed Matter Physics, 2005

Related terms:

Young's Modulus, Bulk Modulus, Poisson's Ratio, Shear Modulus, Tensors-


, Anisotropic

View all Topics

Tight-Binding Method in Electronic


Structure
D.A. Papaconstantopoulos, M.J. Mehl, in Encyclopedia of Condensed Matter
Physics, 2005

Elastic Constants
Elastic constants measure the proportionality between strain and stress in a crystal,
provided that the strain is not so large as to violate Hook's law. Computationally,
the elastic constant is determined by applying a strain to a crystal, measuring
the energy versus strain, and determining the elastic constant from the curvature
of this function at zero strain. A given strain is associated with a certain linear
combination of elastic constants. For cubic systems, including diamond, there are
three independent elastic constants, C11, C12, and C44. One linear combination of the
elastic constants is obtained from the bulk modulus

[11]

where V is the unit cell volume. A second combination is most easily obtained by
straining the crystal in the (1 0 0) direction while simultaneously compressing it in
the (0 1 0) direction to conserve the volume, with lengths in the (0 0 1) direction
remaining fixed. If x is the fractional change in the (1 0 0) direction, then,

[12]
where the function, E(x) measures the energy as a function of strain. Finally, one
finds C44 by straining the crystal in the (1 1 0) direction and fixing the volume by
compressing in the (1 1̄ 0) direction. In this case,

[13]

Note that for diamond the calculation of C44 requires the minimization of the total
energy with respect to an internal parameter at each strain.

The calculation of elastic constants assesses the ability of the method to determine
properties not included in the fitting database. The elastic constants of the cubic
materials calculated by the NRL-TB approach are compared to the experimental
values in Table 3. These calculations were performed at the experimental volume
and therefore they are not consistent with the equilibrium value of B found in Table
1. The deviation of the calculated C11 and C12 from the measured values is within
10–15% which is very close to the error one finds when comparing with a direct
evaluation from first-principles calculations.

Table 3. Elastic constants for cubic elements (in GPa)

Element Structure TB Exp.


C11 C12 C44 C11 C12 C44
C dia 1036 48 601 1076 125 576
Al f.c.c. 125 58 26 103 53 28
Si dia 179 73 95 166 64 80
Ca f.c.c. 15 10 14 16 12 8
V b.c.c. 224 106 92 228 119 43
Fe b.c.c. 223 95 78 237 141 69
Cu f.c.c. 139 99 59 156 106 75
Ge dia 133 20 107 131 49 68
Sr f.c.c. 8 3 −3 15 6 10
Nb b.c.c. 277 139 37 246 139 29
Mo b.c.c. 453 147 120 450 173 125
Rh f.c.c. 491 171 260 433 185 206
Pd f.c.c. 233 163 63 227 176 72
Ag f.c.c. 133 86 42 124 93 46
Sn dia 68 30 38 67 36 30
Ta b.c.c. 275 140 78 261 157 82
W b.c.c. 529 170 198 523 203 160
Ir f.c.c. 694 260 348 590 249 262
Pt f.c.c. 380 257 71 347 251 76
Au f.c.c. 195 174 40 189 159 42
Pb f.c.c. 53 35 19 47 39 14
Po sc 128 13 6
All elements for which a set of TB parameters have been constructed and for which
one has a cubic ground state (except manganese) are presented here. A comparison
is made between the results of the TB parametrization and experiment. Calculations
were performed at the experimental volume.

An inspection of Table 3 reveals somewhat larger errors for C44. A solution to


this problem is to fit C44 either to LAPW calculations or to experiment. It should
be noted here that calculations of Cij in the alkaline-earth metals is problematic
because the lattice is extremely soft. One can also note that the method correctly
reproduces the sign of the elastic constant difference C12C44, even in the metals
rhodium and iridium, and the semiconductors Si and Ge where it is negative. It
is worth mentioning that this negative sign cannot be obtained from the standard
embedded-atom method.

The resulting elastic constants for the h.c.p. lattice are C11, C12, C13, C33, and C44. The
elastic constants found by the NRL-TB method show larger relative deviations from
experiment than found in the cubic crystals. These were found from TB parameters
fitted only to cubic lattices. It is shown, in the case of Ti, that if one extends the fitting
database to include the h.c.p. lattice as well, there is a dramatic improvement in the
comparison of TB and measured elastic constants.

> Read full chapter

Mechanical Properties: Elastic Behav-


ior
K. Bowman, in Encyclopedia of Condensed Matter Physics, 2005

Isotropic Materials
Elastic constants are the constants describing mechanical response of a material
when it is elastic. The isotropic elastic constants for some common materials are
given in Table 1. If the engineering approximation that this mechanical response is
linear is made, one can define a set of constants that relate any applied stress to the
corresponding strain. When the material can also safely be treated as isotropic, the
number of independent elastic constants required to completely describe the elastic
response of a material is two. Any further elastic constants can then be defined in
terms of the other two.

Table 1. Approximate elastic properties of various materials (and dilute alloys) at


room temperature
Material Young's modulus, E Shear modulus, μ (- Poisson's ratio,
(GPa) GPa)
Aluminum and alu- 69–72 24–26 0.35
minum alloys
Copper and copper al- 125–135 47–50 0.34
loys
Irons and steels 205–215 80–84 0.29
Stainless steels 190–200 75–78 0.33
Titanium and titanium 115 42–44 0.32
alloys
Aluminum oxide 380–390 155–165 0.25
Silicon carbide 440–460 195–200 0.14
Glass 70–90 28–32 0.27
Polyethylene (PE) 0.2–2 a 0.4
Polymethylmethacry- 2–3 a 0.4
late (PMMA)
Polystyrene (PS) 2–4 a 0.35
Boneb 5–30 3–8 0.25–0.5
Tendonb 0.8–1.5

Adapted from Bowman KJ (2004) An Introduction to Mechanical Behavior of Materials.


Wiley; © Wiley. This material is used by permission of John Wiley and Sons, Inc.

a These values are typically about one-third of Young's modulus.

b Bone and tendon are not only strongly anisotropic, but the stiffness also varies
with type, position, and moisture content.

For a simple tension or compression test, the easiest elastic constant to define
is Young's modulus, E. Young's modulus is the elastic constant defined as the
proportionality constant between stress and strain:

[1]

Young's modulus has the same units as stress, which is normally expressed in terms
of pascals (Pa), equivalent to newtons per square meter, or pounds per square
inch (psi). Strain is, of course, a dimensionless proportionality between a change in
length and an original length. The Young's modulus is the slope of the linear elastic
response for a number of materials as shown in Figure 1.

Another elastic constant that can be obtained in a tension or compression test


wherein the strain along the loading direction and orthogonal to it is called Poisson's
ratio, . Poisson's ratio is the ratio between width or diameter strain, 1, of a tension
or compression specimen and the strain along the tension or compression axis, 3.
This dimensionless proportionality constant can be written as

[2]
So far, the possibility of stresses occurring in multiple directions have not been
considered. To do so, one needs to extend the idea of assigning a coordinate system
to the stress state as shown in Figure 2a. If such a multiaxial loading condition is
considered, the relations between a set of normal stresses and the corresponding
strains as given by Hooke's law can be written as

Figure 2. (a) Normal stresses referenced to an orthogonal coordinate system (b) shear
stresses and resulting distortion called shear strain.

[3]

This relation enables a set of normal stresses to be translated into corresponding


strains using just the Young's modulus and Poisson's ratio.

A type of stress called a shear stress, , can also be defined which can produce a shear
strain, . For shear stresses, the force is applied across the area and it results in a
shearing of the material. Just as for normal stresses, shear loading can result in a
mechanical response that is elastic and nearly linear, but the shape change is called
a shear strain as shown in Figure 2b. The specific case in Figure 2b shows 4, which
consists of a shear displacement in the y-direction on the z-face and in the negative
y-direction on the negative z-face. To have a stable distortion, a similar pair of forces
would need to be applied on the y-faces to cancel the rotational moment imposed
by 4. The elastic constant that describes the linear relation between and is called
the shear modulus, μ. So, one can write

[4]

For isotropic materials, the relation between these three elastic constants is

[5]

Values for all three elastic constants are given for a number of materials at room
temperature in Table 1. The values are given as ranges to represent variations
expected from alloying.

Because elastic constants are related to bonding, they generally scale with melting
temperature and the type of bonding. The strength and the type of bond determines
the relationship between an applied force and changes in the distance between
atoms. The spacing between atoms or molecules in a material that is not under
stress comes about from a balance between attractive and repulsive forces that
prevent the atoms or molecules from getting too close, and attractive forces that
bring the atoms together. The sum of these two opposite forces is called the
interatomic potential. Application of a stress to the material changes the balance
point. Because the relationship between force and displacement is nearly linear,
the elastic properties of materials can be described using linear relationships. As
materials increase in temperature, they undergo increased thermal vibration that
leads to thermal expansion. This increase in lattice constants also leads to a reduction
in elastic constants with increasing temperature as indicated by Figure 3.

Figure 3. The dependence of Young's modulus on temperature for a number of


materials. The curve for iron shows the values of Young's modulus for the three
phases found in iron as a function of temperature. (Adapted from Bowman KJ (2003)
An Introduction to Mechanical Behavior of Materials. Wiley; © Wiley. This material is
used by permission of John Wiley and Sons, Inc.)

> Read full chapter

Hydrogen–Metal Systems: Elastic Prop-


erties
R.G. Leisure, in Encyclopedia of Materials: Science and Technology, 2004

1.2 Hydrogen–Metal Systems


Elastic constants are of special importance for hydrogen–metal systems. For an
isotropic material with free surfaces, the long-range, elastic interaction energy of
one interstitial hydrogen atom with the strain field of all the other hydrogen atoms
is (Alefeld 1972, Siems 1970)
(3)

where P is an element of the dipole moment tensor, B is the bulk modulus, is the
density of H atoms, and

(4)

where is Poisson's ratio. For a homogeneous distribution of centers of dilatation,


the dipole moment tensor is . The nH hydrogen atoms per unit volume produce a
fractional change in volume given by (Peisl 1978)

(5)

As a result Eqn. (3) can be written as

(6)

where is the volume associated with each H (metal) atom and is the ratio of
H atoms to metal atoms. The elastic constants are needed to evaluate and B;
vM is determined from the density of the metal, and vH is often known from
measurements of the lattice expansion on absorption of hydrogen. The energy U
contributes to the enthalpy of formation of the hydride.

When the hydride forms, the lattice constants of the hydride are usually different
from those of the host metal, producing misfit strains. These strains result in an
elastic energy that affects the formation, decomposition, and morphology of hydride
precipitates (Puls 1984). There have been a number of theoretical treatments of the
elastic energy associated with both coherent and incoherent precipitates (Lee and
Johnson 1978, Lee et al. 1977). The elastic energy of a coherent hydride precipitate
can be calculated with the method of Eshelby (1957, 1961). The formulation of the
Eshelby method by Onaka et al. (1995) is convenient to use. This formulation allows
for both shear and dilatational misfit strains as well as different elastic constants for
the precipitate and the host metal; the elastic moduli of both the precipitate and the
host metal are taken as isotropic. In principle, the full, anisotropic elastic constant
matrix should be used, but detailed calculations show that if the precipitate is either
much softer or harder than the host metal, the results with the full elastic constant
matrix are similar to those for isotropic elasticity (Lee 1977). If the precipitates
become incoherent due to the generation of misfit dislocations, or if precipitation
at defects is important, the situation is quite different (Tanaka et al. 2000).

> Read full chapter

MULTIPHYSICS BEHAVIORS
STEPHEN C. COWIN, in Handbook of Materials Behavior Models, 2001

10.10.4 THE SOURCE OF THE ELASTIC CONSTANT DATA


FOR CANCELLOUS BONE
The elastic constant results for cancellous bone presented here are based upon an
analysis of a database consisting of 141 human cancellous bone specimens. This
database, reported by van Rietbergen et al. [18, 19] and Kabel et al. [9], is superior to
previous databases because the authors provide the entire set of anisotropic elastic
constants without an a priori assumption of a particular material symmetry and
without an assumption of the direction in which the maximum Young's modulus
occurs. This database is unique in many different ways, the most important of which
is the large number of specimens and its method of construction, but particularly
because it is not based entirely on measurements of real specimens. The database
of elastic constants of 141 human cancellous bone specimens employed here was
constructed by imaging real specimens and then computationally determining
their elastic constants. This cyberspace method of construction is thought to be
more accurate than the conventional mechanical testing procedures for evaluating
the elastic constants of human cancellous bone. The determination of the elastic
constants of cancellous bone by conventional mechanical test procedures is very
difficult. The basic problem is that, because of the size of the human body, it
is difficult to obtain specimens of cancellous bone that are more than 5-mm
cubes. The logical way to test small cubes such as these is by compression testing.
However, compression testing is highly inaccurate for cancellous bone because of
(1) the frictional end effects of the platens, (2) the near impossibility of identifying, a
priori, the grain directions in a bone specimen and thus of cutting a specimen in the
grain directions, (3) the stiffening effect of the platens on the bone near the platens,
and (4) the unpredictable inhomogeneity of the specimen.

The construction of the database of elastic constants of 141 human cancellous bone
specimens employed here is a relatively inexpensive method of determining the
full set of anisotropic elastic constants for a small specimen of cancellous bone
by a combination of imaging the specimen [9,10,14] and subsequent evaluation of
the effective elastic constants using computational techniques based on the finite
element by van Rietbergen et al. [18,19]. Once the image of the specimen was in the
computer and a finite element mesh was generated, a sequence of loadings [18,19]
was applied to the specimen and the responses were determined. The sequence of
loadings was sufficient in number to determine all 21 elastic constants. Thus no
material symmetry assumptions were made in the determination of the constants.
Quantitative stereological programs were used to determine the solid volume frac-
tion of each specimen. These are the data employed in the determination of the
elastic constants for cancellous bone recorded in following text.
In this method, the actual matrix material of the trabeculae comprising the bone
specimen is assumed to have an axial Young's modulus Et. The value of Et may be
fixed from a knowledge of the axial Young's modulus for the tissue, or from the shear
modulus about some axis, or by measuring the tissue modulus Et itself. For purposes
of numerical calculation, Et was taken to be 1 Gpa [9,10,14,18,19]. However, since
these are linear finite element (FE) models, the FE results can be scaled for any other
modulus by multiplying the results with the new value of Et (in GPa). The tissue
modulus Et thus is a scale factor that magnifies or reduces all the elastic constants.
The cancellous bone elastic constant results are presented here as multiples of Et.

> Read full chapter

Definition of macro- and microstress-


es and their separation
Viktor Hauk, in Structural and Residual Stress Analysis by Nondestructive Methods,
1997

2.071c Textured material


The XEC are substituted by the stressfactor Fij. They are in addition dependent on the
direction ( , ). Experimentally they are evaluated by a uniaxial tension or bending
test. Also the stress factors can be calculated using the monocrystal data and a
model of crystallite coupling (Voigt, Reuss, Eshelby-Kröner) and the ODF as the
weight function, Chapter 2.13. For non-textured materials the Fij correspond with
combinations of XEC.

In the case of materials with strongly oriented crystallites there is no common


relationship of the strain-, stress-free direction with the appropriate parameters.
As Fig. 1 shows, there exist more than one point of the strain-stress-independent
direction in the D-vs.-sin2 diagram /28/. Kind and intensity of texture are the
influences. More on that subject in Chapter 2.16.
Figure 1. D{100} and relative intensity versus sin2 of a textured steel strip, left: de-
creasing loads after regression analysis, right: points at maximum load and straight
lines after linear-regression analysis /28/.

> Read full chapter

ELASTIC CONSTANTS OF DISOR-


DERED TERNARY CUBIC ALLOYS
Craig S. Hartley, in Nano and Microstructural Design of Advanced Materials, 2003

ABSTRACT
Single crystal elastic constants of disordered alloys having body-centered cubic and
face-centered cubic Bravais lattices can be calculated as functions of composition by
modelling the lattice as a virtual crystal. The technique is based on the method of
long waves applied to the virtual crystal. The three independent elastic constants are
related to four, axisymmetric force constants (ASFC) for first and second neighbors.
The ASFC are defined in terms of the first and second derivatives of a three-pa-
rameter, virtual pair potential, which is determined from the corresponding pair
potentials of like and unlike atom pairs in the crystal weighted by the probabilities
of their existence in the first and second neighbor shells. This technique permits
calculation of single crystal elastic constants of multicomponent disordered alloys
using data obtained from elastic constants of terminal solid solutions of binary alloys
and pure elements when the elements have the same crystal structure as the alloys.
An illustration of the technique is given for ternary alloys of copper, aluminium and
nickel.

> Read full chapter

Mechanical Properties of Silicon Mi-


crostructures
Maria Ganchenkova, Risto M. Nieminen, in Handbook of Silicon Based MEMS
Materials and Technologies (Second Edition), 2015

9.1.2.1 Harmonic Approximation and Room Temperatures


The elastic constants are the system energy second derivatives with respect to the
strains, defined as:

(9.5)

where u is the elastic displacement. Indeed, when the strains are sufficiently low,
one can approximate the elastic energy ΔE stored in a uniformly strained crystal as:

(9.6)

where Cij is the tensor of elastic constants, V is the crystal volume, the Einstein
summation rule over the repeated subscripts is implied, and Voigt’s notation for
the strains is used. The quadratic dependence of elastic energy on strains is usually
called harmonic approximation. The total number of independent components of
the elastic constant tensor depends on the symmetry of the crystal. In particular,
crystalline silicon has cubic symmetry, which is fully characterized by only three
independent constants:

(9.7)

The elastic constants allow one to calculate the components of the elastic stress from
the known elastic strain according to Hooke’s law:

(9.8)

(Note that here the stress and strain components are not in the Voigt but in the
standard notation.)
The usual way for measuring the elastic constants is the monitoring of ultrasonic
velocities as a function of the external static pressure [29]. The first measured values
of silicon elastic constants were reported by McSkimin et al. [30,31]. Later these
constants were defined more accurately by Hall [32] (Table 9.4).

Table 9.4. Comparison of Theoretical and Experimental Values for Second-Order


Elastic Constants for Crystalline Silicon (in GPa)

Method Poten- Refer- C11 C12 C44 μ B Y


tial ence
Theoretical calculations
MD SW [33] 161 80.5 60.2 0.291 46.9 108 138

Tersoff 141 74.0 68.9 0.261 45.3 98.1 138


Lenosky 164 83.0 70.5 0.274 50.5 110 151
SWmod 161 80.5 80.4 0.250 53.6 108 161
Baskes 161 63.3 76.9 0.221 58.2 97.6 164
MS EDIP TB [34] 172 64.7 72.8 0.233 65.1 100.5 161

[9] 167 66.5 75.4 0.232 65.4 100.1 161

[10] 150 56.3 89.0 0.177 72.2 87.6 170

[11] 154 69.9 69.0 0.252 58.3 98.1 146

[12] 128 75.3 58.3 93.0

LDA [33,35] 159 67.3 80.0 0.224 57.2 97.9 140

162 63.5 77.3 0.221 66.1 96.4 161


GGA [33] 151 60.6 79.2 0.209 56.5 90.6 137

Experimental measurements
T=300 K [30,31] 165.64 63.94 79.51 0.218 68 97.84

T=300 K [32] 165.70 63.9 79.6 0.218 60.5 97.8

T=below [29] 167.72 64.98 80.36 99.23


100 K

There exist several theoretical techniques for the calculation of elastic constants,
including molecular statics (MS), molecular dynamics (MD), and Monte-Carlo (MC).
The interactions between silicon atoms are described either with empirical inter-
atomic potential (the best-known potentials are those of Stillinger–Weber (SW)
[13,14], and Lenosky [36], as well as the modified embedded atom method (MEAM)
potential of Baskes [37] and the environment-dependent (EDIP) potential [38] or
in the framework of TB and ab initio formulations. Typically, the elastic constants
are determined as strain derivatives of the calculated stresses or energy. By now,
quite a number of theoretical calculations of these values have been reported
[9–12,33–35] (Table 9.4). The calculated elastic constants are sensitive to the choice
of the interatomic interaction law, but most of them reproduce the experimental
data reasonably well.

The crystal bulk modulus B, defined as:

(9.9)

and the reverse value—compressibility, K=1/B, can be determined by imposing


hydrostatic loading scheme (i.e., 1= 2= 3). The experimentally measured values of B
and K for crystalline silicon are 97.84 GPa and 1.02 GPa−1, respectively [32], whereas
the calculated values for B vary within 90–110 GPa.

For practical characterization of material elastic properties we often use other pa-
rameters, such as the shear modulus, μ, Poisson’s ratio, , and Young’s modulus,
E, which can be expressed in terms of the elastic constant tensor components (in
Voigt’s notation) as:

(9.10)

(9.11)

and

(9.12)

The deviation of the crystal elastic properties from full isotropy is characterized by
either the anisotropy ratio, A, or anisotropy factor, H:

(9.13)

which in silicon are measured to be 1.56 and 574 GPa, respectively.

The Voigt formulation is always appropriate for monocrystalline materials. However,


for polycrystalline materials it can be used only when the grains of different orien-
tations have the same stress. Otherwise, when the grains have the same state of
strain or when long-range internal stress fields are involved, the Reuss formulation
operates better [8].

> Read full chapter

Theory of Vibration
DAN B. MARGHITU, ... DUMITRU MAZILU, in Mechanical Engineer's Handbook,
2001
Solution
The elastic constants are determined using Table 4.4. The natural frequency modes
for some continuous systems are given in Table 4.5, where is specific mass and A
is the area of the section.

Table 4.5.

The natural frequencies are computed with the relation

For the model in Fig. 4.15a, one can write

Equalizing the preceding equation with the natural frequency 1 (n = 1), position 1
in Table 4.5, gives

But m = m and A = m/(a + b). The coefficient of reduction will be

For Fig. 4.15b one can write

and from position 3 in Table 4.5,

For Fig. 4.15c,

and from position 4 in Table 4.5,

For Fig. 4.15d, one can use the same procedure.

For Fig. 4.15e, one can calculate two coefficients of reduction 1 with respect to
and 2 with respect to . If m1 is the mass of the AC section, 1 is determinated with
the relation from Fig. 4.15a, where a → c, and b → a:

If m2 is the mass of the AB section, a2 is determined using the relation from Fig.
4.15e, and one can write 2 = 0.2427 and .

The situation presented in Fig. 4.15f can be calculated using the same procedure.

> Read full chapter

Highly Porous Metals and Ceramics


A.E. Markaki, ... T.W. Clyne, in Encyclopedia of Condensed Matter Physics, 2005
Elastic Constants and Stiffness-Critical Lightweight Structures
The elastic constants of porous materials can be predicted using conventional
composite theory models, with the “reinforcement”-ascribed zero stiffness (and
zero Poisson ratio). For example, Figure 3 shows how axial and transverse Young's
modulus and Poisson ratio are predicted to vary with void content, and with void
aspect ratio, according to the Eshelby method, which is based on assuming that el-
lipsoid-shaped inclusions (pores) are dispersed throughout the matrix. The following
equation gives the elastic constants

Figure 3. Predicted dependence on void content f, and void aspect ratio s, of the
Young's modulus and Poisson ratio, obtained using the Eshelby method (eqn [1]).

[1]

where C is the stiffness tensor, the subscripts m and i refer to matrix and inclusions
(pores), S is the Eshelby tensor (dependent on inclusion aspect ratio s), I is the identity
tensor and f is the volume fraction of inclusions (pores). The model is known to be
unreliable for high inclusion contents, but it, nevertheless, serves to illustrate a few
simple points.

First, the specific stiffness of porous materials differs little from that of the matrix,
since the Young's modulus falls off in a broadly similar way to the density (rule of
mixtures line). In fact, apart from the axial stiffness of a material with aligned, highly
elongated pores, which approximates closely to the rule of mixtures line, the stiffness
falls off more sharply than the density, so the specific stiffness of porous materials
is, in general, lower than that of corresponding monolithic materials.
Second, porous materials can be quite significantly anisotropic in stiffness, if the
pores are aligned and elongated. This is the case in many natural cellular materials.
In fact, it is clear from the plot that the pores do not need to be highly elongated for
this to be the case, since the s = 3 curves are fairly close to those for infinite aspect
ratio. However, it is also worth noting that the plot suggests that elastic anisotropy
is unlikely to be dramatic (greater than a factor of 2 or so). Some assumptions,
such as those of the equal stress model, lead to much greater predicted anisotropy,
but these are known to be very unreliable, at least for cylindrical inclusions (pores).
On the other hand, experimental data for transverse stiffness often suggest that it
is much lower than in the axial direction. For example, typical woods (with a void
content of 20–40%) are often quoted as having an axial (parallel to grain) stiffness
of 10 GPa, but a transverse value of only 1 GPa. For bone, the quoted anisotropy
is usually lower, but is still substantial in some cases. Partial explanation for such
discrepancies is often related to difficulties in measuring a true transverse stiffness.
The strength (and creep resistance) is often dramatically lower under transverse
loading (see below), so that inelastic straining commonly occurs at low applied loads
during stiffness measurement. However, it should also be recognized that this type
of model, which takes no account of the distribution of voids, and assumes that
they have no contiguity (i.e., that the cells are all closed), may be very unreliable for
high-void contents and for open-cell materials. Finally, it must be noted that these
models assume the base material to be isotropic. For many natural materials, such
as wood, the cell walls themselves exhibit high degrees of structural alignment, and
hence are highly anisotropic, both elastically and in terms of strength. Obviously,
such effects should be taken into account when predicting the effect of the presence
of pores.

It can also be seen in Figure 3 that the Poisson ratio may vary significantly when
voids are introduced, although the relative changes are predicted to be smaller than
those for stiffness. However, it should again be emphasized that, particularly at
high-void contents, and with open-cell structures, geometrical aspects can assume
much greater importance, and the “mean field” approach of the Eshelby method
is certainly not well-suited for prediction of elastic constants at high-void contents.
For highly porous materials of this type, geometrically specific models may be more
appropriate. For example, consider the plots shown in Figure 4. These relate to
porous materials composed of an isotropic (random) array of slender cylindrical
fibers, bonded together at junction points – see Figure 2d. A model has been
developed for such material, based on assuming that deformation arises solely from
elastic-bending deflections of the fiber segments between junctions. The following
simple equation is derived for the normalized Young's modulus
Figure 4. Predicted Young's modulus of an open-network porous material compris-
ing an isotropic array of bonded fibers, as a function of the aspect ratio L/D, of fiber
segments between junctions, according to eqn [2]. Curves are plotted for several
void contents (f) and some corresponding experimental data are also shown.

[2]

where L/D is the aspect ratio of fiber segments between joints. These stiffnesses,
both measured and predicted, are well below those predicted by the Eshelby method
for the void contents concerned. In general, open-cell materials are less stiff than
closed-cell ones with similar void contents, while open-network materials are the
least stiff. These discrepancies highlight some of the limitations of applying con-
ventional composite theory approaches to highly porous materials, particularly if
they have open-cell or open-network structures. It might also be noted that the
model used to produce the plots in Figure 4 predicts a Poisson ratio value of 0.32,
independent of void-content and fiber segment aspect ratio.

Finally, it should be emphasized that, while porous materials do not have superior
specific stiffness to monolithic materials under uniaxial loading, they can exhibit
dramatic advantages when subjected to bending moments or torques. The reason
for this is that “bending stiffness” and “torsional stiffness” are raised, not only by the
material having a higher Young's modulus, but also by it being located further from
the neutral axis or neutral plane. Depending on the exact boundary conditions, the
appropriate figure of merit for stiffness-critical weight minimization (E/ for uniaxial
loading) is commonly E1/2/ or E1/3/ for bending, which can make highly porous
materials attractive for construction of many types of stiff, lightweight structures. A
simple example of this is sandwich panels, for which optimizing the design of highly
porous core material is an important issue.

> Read full chapter

X-ray elastic constants (XEC)


Viktor Hauk, in Structural and Residual Stress Analysis by Nondestructive Methods,
1997

2.134h Comparison of XEC-calculations on multiphase mate-


rials using different model assumptions
A study on the XEC of two-phase materials was done early using different kinds of
structures /51/. The details of the different assumptions, models and the calculation
itself are handled at the appropriate places of this book. Here, as an example, the
XEC of Cu-Fe materials will be discussed taking into account the XEC of Cu{331} and
Fe{211}. Fig. 21 /51/ demonstrates the XEC versus the Cu-Fe composition with the
structure or model as parameter. The margins are again the limiting assumptions
of homogeneous strain and homogeneous stress, here for the interaction between
the phases. Here, reference will be made to two similar diagrams but XEC versus
orientation parameter are plotted with the Cu-Fe composition as parameter /48,56/.
This figure is shown in section 2.134c.
Figure 21. XEC of two-phase Cu-Fe materials /51/,— homogeneous stress, —
sintered material: a matrix, b inclusion; …. monocrystal in two-phase matrix; -..-
layer-substrate material; - · - fiber-reinforced composite (Cu matrix, Fe fiber), ___
homogeneous strain.

The suppositions, models, and references of the XEC-calculations are summarized


in Table 4.

Table 4. Publications on XEC-calculation of nontextured materials and materials with


preferred orientation.

material supposition model reference


nontextured; S1,

one phase cubic Voigt, /%14,%15,%3//24/-


Reuss,Eshelby-K- /18/
röner,shear
modulus known
hexagonal V, R /46/
all systems V, R, E+K+Kneer /25/
two-/multiphase phase-XEC known VR /50//51/
macroscopic E known E + K + Kn /48,56/

two phase layer-sub- phase-XEC known Oldroyd-StroppeV- /51//51//57//52/


strate fiber reinforces
pores VOldroyd-Stroppe
textured; Fij
one phase ODF VRE + K + Kn /60,61,37,42/-
/28,29,30,61,37,42//-
42,58/
two-/multiphase ODF, cubic, macro- E + K + Kn /42/
scopic, E known

> Read full chapter

ScienceDirect is Elsevier’s leading information solution for researchers.


Copyright © 2018 Elsevier B.V. or its licensors or contributors. ScienceDirect ® is a registered trademark of Elsevier B.V. Terms and conditions apply.

You might also like