You are on page 1of 30

Chapter 6.

Elastic behaviour of solids

6.1. Objectives
• To study the relationship between the forces acting on a solid and the
resulting deformations.
• To introduce the concept of modulus of elasticity.
• To identify the atomic parameters which determine the elastic behav­
iour of solids.
• To calculate the theoretical limit for the strength of materials.
• To analyse the viscoelastic behaviour of certain materials and to intro­
duce the concept of linear viscoelasticity.

Any object subjected to a mechanical force deforms. This deformation, which


leads to the displacement of the atoms from their equilibrium position, produces
forces of retraction opposing the deformation. These tend to re-establish the solid in
its original form when the force is no longer applied.
The behaviour of solids subjected to mechanical stresses is determined by the
nature of the interatomic forces. In this chapter, only the elastic behaviour of ideal
solids submitted to small deformation (generally < 0.1 %) will be studied in order to
simplify the presentation of the concepts introduced. In chapter 11, the tensile beha­
viour of real materials will be studied at high deformation.

6.2. Elastic deformation of solids


6.2.1. Introduction
The theory of elasticity treats the relationship between forces applied to an ob­
ject and the resulting deformations. In practice, the analysis of the elastic behaviour
of a material is reduced to the study of simple deformations and the determination of
the corresponding elastic constants. These simple deformations are uniaxial elonga­
tion, simple shear and uniform (or hydrostatic) compression.

6.2.2. Uniaxial elongation


When a prismatic sample (figure 6.1.) is subjected along χ an external force F , it
undergoes an uniaxial elongation proportional to its initial length x . 0

© 2002 Editions scientifiques et medicales Elsevier. All rights reserved.

Introduction to Materials Science


122 Introduction to Materials Science

ύ
0

Figure 6.1. Uniaxial elongation of an elastic bar of initial cross-section S =yo z and of
0 0

initial length xo. This figure is schematic and does not show the lateral contraction, asso­
ciated with the elongation of the sample during the application of the force

This elongation produces a force of retraction F in the solid equal in absolute


r

value and opposite in direction to the appliedforce F:


F + F =0 r

The force of retraction F and the elongation Δχ are expressed in general as a func­
r

tion of the initial dimensions of the prismatic sample. The relative force per unit
cross section of the sample is the (nominal) stress:

(6.1.)
x
s 0

The ratio of the increase in length Ax, to the initial length x defines the (nomi­
0

nal) strain £ : x

ε= — (6.2.)
χ
x0

The nominal stress (strain) is the stress (strain) determined taking into account
the initial dimensions of the sample. The real stress (strain) considers the dimension
of the deformed sample. The difference between nominal and real values is only
significant at high deformation and will be discussed in chapter 11.
At small deformation (for most materials, e < 0.1 % ) , a linear relationship exists
x

between the stress σ and the strain e \


χ x

σ = Εε
χ χ
(6.3.)

This expression defines Hooke's Law, in which Ε is the elastic modulus in ten­
sion or Young's modulus characterising the resistance (rigidity) of the solid to
uniaxial deformation. In table 6.2., the values of the elastic modulus for a number of
Elastic behaviour of solids 123

Table 6.2. Typical values of the modulus of elasticity Ε for different


materials at ambient temperature.

Material Ε Material Ε
[GPa] [GPa]

Diamond 1000 Window glass 70


Silicon carbide (SiC) 450 Aluminium 70
Tungsten 400 Concrete 50
Alumina (A1 0 )2 3 400 Magnesium 40-45
Carbon fibre 300 Wood agglomerate 7
Magnesia (MgO) 250 Epoxy resins (cross-linked) 2.8-4.2
Steel 210 Polystyrene 2
Copper 125 Polyamide 6-6 2
Brass, bronze 110 Polypropylene 1.5
Glassy silica ( S i 0 ) 2 95 Medium density polyethylene 0.7
Gold 80 Rubbers -0.001

materials are given. They are expressed in GPa or G N m ( 1 0 ^ m~ ). The variation2 2

observed from diamond to rubber extends over six orders of magnitude. The value
for the modulus of elasticity is a function of the bond energies, the nature of the re­
tractive elastic forces, and the structure (amorphous or crystalline) of the material.

6.2.3. Lateral contraction and Poisson's Ratio


For most materials; the elongation Ax of the sample along the tensile axis leads
to an increase in volume. In the case of elastic deformation, this volume increase is
partially compensated by lateral contractions, Ay and Δζ, of the sample in the y and ζ
transverse directions perpendicular to the tensile axis.
The strains in the y and ζ directions are given by:
Ay Az
£ = — et ε = —
y ζ (6.4.)
yo z 0

For an isotropic material, £ and e are equal. y z

The effect of the lateral contraction is generally measured relative to the tensile
strain. Thus Poisson'sRatio ν is defined as:

V = -3l =-£- (6.5.)

where the negative sign makes ν positive. It is easy to relate the variation in volume
during uniaxial extension to Poisson's Ratio v. The relative variation in volume Δ is
written as:

(6.6.)

with
Vo = xoyoz 0 (6.7.)

and
124 Introduction to Materials Science

(6.8.)
XQ
0
yo

By considering only small strains, it is possible to neglect the infinitely small second
and higher order terms. This means that, taking into account equations 6.6, 6.7, 6.8, it
is possible to write:

Ax Ay Az
A
(6.9.)
A= + — + : £ x + £ y + £ z

XQ
yo

By introducing into equation (6.9.) the value of ν taken from (6.5.), the following
relationship is obtained:

A=£ (l-2v)
x
(6.10.)

The various values of Poisson's Ratio ν are summarised in figure 6.3. The upper
limit for the value of ν is 0,5 for an isotropic material. This corresponds to a zero
variation in volume during extension (A = 0). This limiting value is almost obtained
for rubber (v = 0.49) which deforms under tension in an elastic manner almost with­
out any increase in volume. For metals, Poisson's Ratio is close to 0.35 and for ce­
ramics, υ is generally between 0.07 and 0.27.

Metals Ceramics Polymers


0,5
Nat rubber AV = 0

Pb PE
0,4
Ag
Cu AV>0
Al PMMA
PS, PA 6-6
0,3
Fe, Steel, W

Mineral glass
0,2 A1 0 ,WC2 3

MgO
Amorphous S i 0 2

0,1
Diamond

0,0

Figure 6.3. Indicative values of Poisson's Ratio ν for different types of materials for
elastic strains at ambient temperature.
Elastic behaviour of solids 125

6.2.4. Simple shear

Shear forces play an important role in the behaviour of materials. Shear forces
are involved when brake blocks are applied to the rim of a bicycle wheel or to the
jaws of a car disc brake.
An example of deformation under shear is shown in figure 6.4. A prismatic
specimen is fixed at one face of the surface .SO to a rigid support, while on the oppo­
site face a force F is applied parallel to the plane xy. The ratio between F and So is the
shear stress r = F/So. The shear stress induces a shear strain which is measured by the
shear angle γ = Ay/z . There exists for this deformation, which occurs at constant
0

volume, a linear relationship between the stress per unit area of surface, the shear
stress r a n d the shear angle γ:

τ = Gtgy- Gy (6.11.)

since for small deformations, t g / ~ γ= Ay/z . G is the shear modulus.


0

Figure 6.4. Simple shear. The force applied per unit area is the shear stressr = F/SO and γ
is the shear angle (shear strain).

6.2.5. Uniform (hydrostatic) compression


Uniform compression constitutes the third type of simple deformation. This
occurs when a solid is subjected to a hydrostatic pressure (positive or negative). A
linear relationship exists, at small volume change, between the hydrostatic pressure
(stress)p, and the relative change in volume Δ = AVIV, given by:

ρ = -ΚΔ (6.12.)

In this expression, Κ is the bulk modulus. The negative sign in this expression is due
to the fact that the change in volume is negative for a positive p.
126 Introduction to Materials Science

6.2.6. Relationships between the various elastic moduli


Three different moduli, E, G and K, which have been defined, allows the elastic
behaviour of a material to be characterised. These three constants result from the
proportionality, which exists between stress and strain for these three types of simple
deformation:
σ= Εε
χ - Gy (6.13.)
ρ--ΚΑ

As already indicated, these three expressions of Hooke's Law are only valid for
small deformations (linear elasticity). This limit for linear elastic strain is approxi­
mately equal to 0.1 % for high modulus materials such as metals. Above this limit the
phenomena of permanent (plastic) deformation appears in a large number of cases
(chapter 12).
As will be seen latter, for elastomers it is not isolated atoms, which are displaced
but segments of chains. Elastomers are characterised by very low values for the Ε
and G moduli (of the order of 0.01 to 0.001 GPa). They are materials very deform-
able by uniaxial elongation and in shear. In uniform (hydrostatic) compression, elas­
tomers behave like high modulus materials or like liquids, with a bulk modulus Κ
greater than 1 GPa.
The three moduli E, G and K, as well as Poisson's Ratio ν are related to each
other by the following equations:

E =- ^ ^ = 3K(l-2v) = 2G(l + v) (6.14.)

In other words, among the four elastic constants, only two are independent.
These relationships are only valid for isotropic solids. For the most general case of an
anisotropic solid, twenty-one independent elastic constants exist (Cottrell, 1964).
In isotropic solids, simple shear and uniform compression make up two types of
elementary deformation, because in one case the volume and in the other case the
shape of the sample remains constant. Uniaxial elongation is not a elementary de­
formation because, except in the case of rubber, it causes both the volume and the
shape of the sample to vary simultaneously.

Table 6.5. Relationships between K, G and Ε

Physical State Κ G,E

Liquid large 0
Rubber large small (E, G « K)
Crystals, glasses large large (£, G~X)

In table 6.5, the relative values of the elastic constants are summarised. For
rubber, Κ is very high compared to Ε and G, i.e. for mechanical deformations, the
elastomers can be considered to be essentially incompressible bodies and Poisson's
Ratio ν is practically equal to 0.5. Equation 6.14 reduces then to:
E~3G (K^oo) (6.15.)
Elastic behaviour of solids 127

6.3. Thermodynamics - Atomic origin of elasticity


6.3.1. Thermodynamics of isothermal uniaxial elongation

In section 6.2., consideration was limited to the macroscopic aspects of the elas­
tic behaviour of solids. It is useful to determine the relationships existing between the
force of retraction, the length and the temperature of a sample subjected to deforma­
tion and the magnitude of the fundamental thermodynamic parameters, i.e. free en­
ergy (free enthalpy), internal energy (enthalpy) and entropy.
Consider an elastic bar (figure 6.6.) of initial length / , maintained at a length
0

/ > /o by the force F. The application of the first and second laws of classical thermo­
dynamics to this deformation will now be examined. The variation of the internal
energy dUresulting from this elementary deformation is given by:
dU=dq-dW (6.16.)

where dq is equal to the amount of heat absorbed by the system during deformation
and dW represents the work done. By considering only the case of reversible proc­
esses, the thermodynamic effect dq is related to the entropy variation dS by:

dq=TdS (6.17.)

The work dW done by the system breaks down into two terms related to the work
done by the elastic force of retraction and to the work of the pressure p\

F + dF

dl

ίο

F + dF

Figure 6.6. Uniaxial extension of an elastic bar, maintained at a length / by a force F and
then extended by an increment d/ by an increase in force of dF.
128 Introduction to Materials Science

dW=-F dl+pdV r (6.18.)


It should be recalled that the elastic force of retraction F is equal in absolute valuer

and of opposite direction to the applied force F:

F = -F
r (6.19.)
The variation in internal energy during the reversible uniaxial deformation is there­
fore given by:
dU = TdS+F dl-pdV r (6.20.)

If the free energy A (Helmholtz's function) is defined in the usual way:


A-U-TS (6.21.)
then the variation in the free energy dA can be related to the entropy and to the work
carried out by the system. Therefore:

dA = - SdT -pdV + Frdl (6.22.)

If the deformation occurs at constant Τ and V, the expression (6.22.) reduces to


dA = F dl or ( 3 A / 3 / ) = F . For constant Τ and V and taking into consideration
r rv r

equation (6.21.), then:

( 3/ ) r f V ( 3/ ) r v (3/ \ v (6.23.)

The elastic force of retraction, at constant Τ and V, is therefore equal to the in­
crease in free energy per unit length of elongation of the specimen. The elastic force
of retraction can be split into two components, the internal or enthalpic force of re­
traction F ,i , and the so-called entropic force of retraction F . The mechanical en­
r r<e

ergy brought to the system by the deformation can be stored in the form of an inter­
nal energy increase resulting from the modification of the interatomic distances or
the bond angles. The energy can also be dissipated into the environment as heat with
a corresponding reduction in the system entropy accompanied by an increase in order
resulting from an orientation of the chains.
As A is a state function, dA is an exact differential, i.e.:

" IdA\' 3 f

(6.24.)
a/1 Τ ,3/Jr.
and so:
JdS\ JdFA (6.25.)

By introducing (6.25.) into (6.23.), the following relation can be deduced for F : r

(6.26.)
Elastic behaviour of solids 129

The relationship (6.26.) is a form of the equation of state for isothermal reversi­
ble elastic deformation. It shows that by studying the variations in the force of re­
traction as a function of temperature, the internal and entropic contributions to the
force of retraction can be determined.
For all materials except elastomers, the volume increases during uniaxial elon­
gation but since this variation is small, expression (6.26.) remains valid, to a first
approximation, for all materials.
This study of the thermodynamics of elastic deformation leads to two limiting
cases: the ideal crystal and the elastomer that will be introduced in the course of this
chapter. For an ideal crystal, the force of retraction remains practically constant as a
function of the temperature. This means that it varies with internal energy or enthalpy
changes. For rubber, there is an increase in the force of retraction proportional to the
absolute temperature Τ and this has therefore an almost exclusively entropic origin.
There is considerable similarity in behaviour between a gas deformed under
compression and rubber deformed in tension. In a perfect gas subjected to isothermal
reversible compression, the entropy per unit volume varies proportionally with pres­
sure:

(6.27.)
P 1
\dV;
T [dT;
V

In the same way, ideal rubber is defined as a material characterised by an elastic


force of retraction exclusively entropic and given by the following expression:

(6.28.)
{dl) y
T {dT) ltV

This analogous behaviour of the perfect gas and ideal rubber is not fortuitous.
Gas molecules heat up when they are rapidly compressed and cool down during
expansion. A piece of ideal rubber heats up when stretched and cools down when the
deformation is released. This thermal effect is easy to explain: during the isothermal
deformation of an ideal elastomer, there is virtually no interaction between the chains
and the internal energy does not depend on the degree of elongation. In this case, the
elongation of the sample is only accompanied by a reduction in entropy, leading to
heat emission into the thermostatic bath surrounding the sample. During the opposite
deformation, entropy increases with heat absorption. In practice, it is observed that
the current elastomers have an elastic behaviour close to that of ideal rubber.
The heating up of an elastomer during elongation should not be confused with
the heating of a car tyre subjected to dynamic deformation induced by the movement
of the vehicle. In effect, during the cycle of deformation experienced by the tyre, the
phenomena of elongation and retraction entirely compensate each other and the
global calorific effect should be zero for an ideal rubber. In reality, a hysteresic effect
(chapter 11) occurs and the loading curve is not identical to that for unloading. A
fraction of the energy put into the system during elongation of the rubber is
dissipated in the form of heat during retraction. The internal heating of tyres is
related to the phenomena of internal friction created by the rubbing of segments of
chains against one another, exactly as molecular friction in a deformed fluid leads to
the Joule effect. This effect is not considered in the theory of elasticity presented
130 Introduction to Materials Science

here. This non-elastic response of the rubber also explains the shock-absorbing
capacity of rubber blocks which isolate machine tools from their mountings.

6.3.2. Relation between phenomena occurring on an atomic scale and thermo­


dynamic variable
The force of retraction is always induced by the displacement of atoms from
their equilibrium positions. In solids with enthalpic deformation behaviour (metals,
crystalline ceramics, mineral or organic glasses, highly cross-linked thermoset poly­
mers etc.), the force of retraction (figure 6.7. (a)) results from small displacements of
the atoms from their equilibrium positions. The cohesive energy of these solids being
high, the elastic forces of retraction are very strong, the modulus of elasticity is high
and the elastic domain limited (ε < 0.1 % ) . In this case, the deformation is associated
with a significant increase in internal energy (or enthalpy) of the system, whereas the

(a)

CO

Cu
ο

0,2 0,4 0,6 (%)


200 400 600 (%)
Deformation ε —> Deformation ε
(b) (d)
Figure 6.7. Atomic mechanisms involved in the elasticity of solids and their relationships
with internal energy U and entropy S variations. Deformation mechanisms (a) and inter­
nal energy variation (b) in enthalpic solids (metals and ceramics). Deformation mecha­
nisms (c) and entropy variation (d) in entropic solids (elastomers). The different scales of
deformation ε between (b) and (d) should be noted.
Elastic behaviour of solids 131

configuration remains practically unchanged since the atoms hardly move from their
equilibrium positions (figure 6.7. (b)).
The solids with entropic elasticity (elastomers) (figure 6.7. (c)) are made up of
an assembly of polymeric chains with a small number of cross-links (~1 cross-linked
unit per 100 structural units in general). This three-dimensional structure ensures the
reversibility of the deformation. The cohesion between the rubbery chains is weak
and of the same order of magnitude as that found for volatile molecular liquids (~4 kJ
m o f ) ; this value is some 100 times smaller than covalent cross-linked bonds.
1

Consequently in elastomers, it is not individual atoms, which are displaced but seg­
ments of chains. In this type of material, the forces of retraction are induced by mi-
crobrownian movements which continuously modify, in a random way, the position
of diverse segments of chains and which therefore oppose any orientation of the
chains under the action of an external force. The forces of retraction created by these
microbrownian motions are very weak and they increase in proportion with the ab­
solute temperature T. This explains why elastomers are easily deformed and have
very low elastic moduli (E ~ 1 to 10 MPa), the latter increasing as a function of T.
As previously mentioned, this study has been restricted to the perfect crystal
(slightly deformable) and an ideal rubber (very deformable). It should also be noted
that the majority of current materials are solids with enthalpic elasticity. Their elastic
moduli vary by some three orders of magnitude, from diamond (Ε ~ 10 GPa) to the 3

organic glasses (Ε ~ 2 GPa). This variation mainly depends on the nature of the
bonds. This explains the difference in the elastic modulus observed between mineral
glass (Ε ~ 70 GPa) and organic glass (E = 2 GPa). The cohesion between the polymer
chains of organic glasses is made by secondary bonding (-40 kJ mol" ), while only1

strong ionic or polar covalent bonds exist in mineral glasses. The mechanical proper­
ties are also influenced by the atomic structure and so amorphous materials generally
have somewhat lower elastic moduli than crystalline solids.
Semi-crystalline polymers (chap. 5 and 10) have a microcrystalline structure
dispersed in an amorphous phase. If the amorphous phase is glassy (T< Tg), the
modulus of these materials is of the same order of magnitude as that of organic
glasses (Ε ~ 2 to 3 GPa). When the amorphous phase is rubbery (T> Tg), the elastic
modulus varies between 0.2 and 1.5 GPa as a function of the degree of crystallinity.
In this class of materials, the elastic force of retraction can simultaneously have en­
thalpic and entropic components.

6.3.3. Elastic deformation of a perfect crystal


This calculation will be restricted to the case of a crystalline solid having a
primitive cubic unit cell Ρ (figure 3.8.); a spherical atom of diameter r (section 2.4)
0

fills each node. Each atom occupies a cubic volume roughly equal tor . The study is
0
3

limited to the deformation of a prismatic crystal in tension (figure 6.8.). To simplify


the calculation, a tensile axis parallel to an edge of the basic network is chosen. Thus,
from expression (6.29.), the strain £ of the solid is the same on both the macroscopic
x

and atomic scales.


Ax _ Ar
(6.29.)
x
£
x ~ ~~
0
x r
0
132 Introduction to Materials Science

Figure 6.8. Uniaxial deformation of a prismatic crystal of section S and length x in 0 0

which the basic unit cell is primitive cubic and the structural unit is an atom of diameter
r . S = r is the area occupied by an atom. The tensile axis is parallel to the edge of the
0 a 0
2

basic unit cell.

For isotropic materials, the transverse deformation is similar on both the micro­
scopic scale and on the macroscopic level:
_ Ar _ Ar y z
(6.30.)
0
r
r 0

This is known as an affine deformation. In crystalline solids, this behaviour is


only observed for small deformations (ε < 0.1 %). χ

In elastically deformed crystalline solids, the force of retraction is induced by


the atoms being displaced from their equilibrium positions. When only small defor­
mations are considered, the elastic force of retraction f , between two neighbouring x

atoms parallel to the direction of deformation χ is proportional to the interatomic


distance Ar : x

f
x = CAr x (6.31.)

fx represents the microscopic retraction force acting between two atoms ,each occu­
pying a section S = r (figure 6.8.). It should be noted that f, represents, in reality, a
a 0
2
x

difference in force Δ f relative to the equilibrium position (f = 0). The macros­


x iX

copic force of retraction F , which is related to the microscopic force of retraction f ,


r x

is proportional to the number of atoms found in the section So.f.x represents a dif­
ference in force Δ f relative to the equilibrium position if,χ 0). C is a constant of a
x
=

function proportional to the bond energy. Since each atom roughly occupies a section
S = r (figure 6.8.), the macroscopic retractive force F is proportional to the number
a 0
2
r

of atoms found in the section So:

(6.32.)
0
r
0
r

To simplify the notation, in ( 6 . 3 2 . ) h a s been used for/ ,x and Δ r for Δ r . As: r x

_ Ax _ Ar
(6.29.)
x

x
0 r
0
Elastic behaviour of solids 133

then:
^ SftC Ax
F =-^ r (6.33.)

Introducing the stress σ , then: χ

σ χ = ^ =- ε χ (6.34.)

In this expression, C/r is the modulus of elasticity or Young's modulus E, which was
0

defined in (6.3.) in introducing Hooke's Law (§ 6.2.2.):


σ = Εε
χ χ (6.3.)

6.3.4. Theoretical strength and fracture energy


In solids with enthalpic elasticity, the force of retraction is related to the bond
energy. The following calculation relates tensile strength to the modulus of elasticity.
For ionic solids, an equation was established, in chapter 2, which relates the variation
of the bonding force between two ions of opposite sign as a function of the distance
(2.20). When a pair of ions is subjected to elastic deformation, the force of retraction
has (with the exception of the sign) the same mathematical form as the expression
(2.20). It can thus be written:

f r 8

f =JL i_LL (6.35.)


Jr 2

r represents the interatomic distance after deformation and r is the distance at which0

the cohesion and repulsion forces are in equilibrium (f = 0). If only small strains are
considered, it is possible to write the following approximate expression:

f ~df =^-dr
r r (6.36.)
ro
For ionic solids, this expression permits the rigidity constant, C of the bond and
the elastic modulus to be calculated. Introducing f into (6.31) and (6.3.) respectively,
r

the value of C is obtained:

C=H (6.37.)

and the modulus of elasticity E:

£ =^ = *ί (6.38.)

The force of retractions β increase significantly when the deformation ε is in­ χ

creased (figure 6.10.). The force of retraction passes through a maximum for a value
of strain equal to e , . The maximum value of the force of retraction/^, which deter­
x m

mines the theoretical tensile strength of the material, is calculated by putting the first
derivative of (6.35) equal to zero, giving:
134 Introduction to Materials Science

e , = 0.22
x m (6.39.)

The value of e (22 %) is high and the bond strength f , deviates considerably from
x>m r m

the linear relationship (6.36.)· From the equation (6.35.), which gives the variation in
the force of retraction as a function of the interatomic distance r, the maximum
strength of the material R is calculated by:
max

±
X_
Rn :0,54 (6.40.)
f
ro4

2 Jr,m
By substituting equation (6.38) in (6.40), R can be calculated as a function of the
m x

modulus of elasticity E:
E_ (6.41.)
RfT\ il Y
15
A similar relationship can be calculated for other materials with strong bonds.
This relationship is approximate, as it only takes into account the interactions be­
tween pairs of atoms.
Is it possible to obtain maximum strengths as high as those predicted theoreti­
cally? In practice, the maximum strength R corresponds to the stress in the material
m

at the onset of plastic deformation, i.e. the yield strength R (chapter 11.). Figure 6.9. e

gives the value for the ratio RJE for the three categories of materials. In certain cases
(diamond), very high maximum strengths are achieved close to those predicted theo­
retically. This is equally true for metal whiskers, which are fine needles (some μπι in
diameter), obtained under special conditions of crystalline growth which confers on
them a defect-free crystalline structure. For example, an iron whisker attains an ex­
traordinary strength of 11 GPa with a maximum elastic elongation of 5 %. This is
very close to the theoretical value (R = Ε/15 ~ 14 GPa). On the other hand, the best
max

Metals Ceramics Polymers

Theoretical limit
10"
Trichites Diamond ΡΕ,ΕΡ,ΡΑ
SIC PPMA
Ti Alloys Al 0 ,Si N
ι ο­ 2 3 3 4
Steels
MgO
Al Alloys
Cu Alloys
ίο- 3

Concrete, cement
Pure metals
1 1 ( r

3
Ultrapure
Ö 10" : metals

10" 6

Figure 6.9. Ratio RJE for different types of materials. The theoretical limit of R / max Ε is
equal to 1/15 (after Ashby and Jones, 1980).
Elastic behaviour of solids 135

steels currently used have a maximum strength of only about a quarter of the theo­
retical value. A steel wire, which in theory should be able to support stresses of the
order of 14000 MPa (1400 kg mm" ), in practice does not generally withstand more
2

than 2400 MPa (240 kg m m ) . -2

Polymers also have strengths which approach the theoretical limit R = E/\5. It max

is important to note, however, that, in this case, the strength of these materials is low
because the modulus of elasticity is significantly less than that of other materials, due
to the presence of weak secondary bonds.
As may be seen in figure 6.9, calculation leads, in the case of metals and alloys,
to theoretical strengths which are often hundreds of times higher than measured
values (10000 times for pure metals). This arises because the role played by defects
in lowering material strength has been neglected in the calculations. These defects
will be studied in the next chapter. It is by taking into consideration the predominant
role played by these defects, in particular by dislocations, that theory and experience
can be reconciled. It should be noted that it was this divergence between theory and
practice that was at the origin of the discovery of dislocations.
Fracture energy is another important characteristic of materials. The energy at
the fracture of a bond U, is given by the area under the curve of the variation in the
force of retraction f as a function of deformation (figure 6.10.):

Ur="fr,xar (6.42.)

By multiplying the fracture energy U by the number of bonds broken per unit
R

area of the solid rupture surface, the specific fracture energy G can be calculated. C

The separation of the sample into two parts during rupture forms two fracture
surfaces, each one with a specific surface energy y. If the creation of two surfaces
were the only phenomenon accompanying the rupture, then the specific fracture
energy G would be double the specific surface energy /.
C

G = 2y (6.43.)
C

ο 0,2I 0,4 0,6 0,8


Deformation ε = Ar/r
χ 0

Figure 6.10. Variation of the ratio of the retraction force f to the maximum value f , of
r r m

the ionic bond as a function of the deformation ε . χ


136 Introduction to Materials Science

Measurement of the specific fracture energy G in brittle materials such as ce­


c

ramics gives values close to theory (figure 6.11.). On the other hand, ductile materi­
als such as metals and certain polymers have fracture energies from 10 to 10 times 3 6

higher than theory. This indicates the existence of other energy absorption processes,
linked mainly to plastic deformation (chapters 7. and 11.). This highlights a very
important characteristic of ductile materials, which will be examined latter (chapter
14.). There are few materials, which reach the theoretical limit of 2y. In practice,
there are almost always other mechanisms (chapter 14.) dissipating energy, thereby
increasing the fracture energy above the theoretical value predicted by (6.43.).

Metals Ceramics Polymers


10 3

Ductile Metals

10 2 Steels
Ti alloys

Al alloys
η 10 PP
ε PE
PA
PS
PMMA
EP
Be
I ίο- 1

S3N4
UP

sic
MgO
ΚΤ- Cement
AUO3
Glass

ΐσ-

io-
Figure 6.11. Measured values for the specific fracture energy G for different classes of
c

materials (after Ashby and Jones, 1980).

6.3.5. Elastic deformation of ideal rubber


As already seen in figure 6.7. (c), an elastomer consists of an assembly of chains
linked to each other by cross-links which are introduced by vulcanisation after shap­
ing of the material. Microbrownian motions constantly excite the polymer segments
in an elastomer, and their amplitude is proportional to the absolute temperature. The
kinetic theory of elasticity in rubber shows that the microbrownian motion, which
excites an elastic chain linked at both ends by cross-links (figure 6.12.), induces a
force pulling the chain ends together. This force of retraction, which is entropic in
origin, is proportional to the distance / separating the two cross-links situated at the
ends of the elastic chains and is given by the expression:

Ρ - (6.44.)
3 K T L

r
~ Ϊ 2
Elastic behaviour of solids 137

Crosslink
Crosslink

Figure 6.12. Elastic retraction force between the ends of a elastomer segment.

The factor k is the Boltzmann s constant and Ϊ represents the mean quadratic dis­
1

tance between the elastic chain ends before the introduction of the cross-links (i.e.
before vulcanisation).
In the undeformed state, the retractive forces acting on the cross-links of the
network are in equilibrium. When external tensile forces are applied to the rubber
sample, the distance separating the ends of each elastic segment is increased. For the
chain represented in figure 6.12., when the distance / between the two ends of the
segment is increased by an increment Al parallel to /, the elastic retraction force is
increased by a value Δ f given by the expression:
r

» r 3kT (6.45.)

During deformation, the mean distance between the cross-links varies in the
same proportion as the macroscopic deformation. Therefore, an affine deformation
occurs here, as well as in the case of the deformation of a monocrystal described in
paragraph 6.3.3. In an elastomer, it is not the chain atoms, which are displaced in an
affine manner, but it is rather the cross-links.
For high modulus materials (metals, ceramics, glassy or semi-crystalline poly­
mers), there exists, for small deformations (-0.1 % ) , a linear relationship between the
stress, i.e. the retraction force per unit area, and the strain. The elastic region for
these materials is very small. More important deformation leads to plastic deforma­
tion or rupture (§ 11.2.3.). For elastomers, which are much more deformable materi­
als, the elastic region is larger. In figure 6.13., the stress-strain curve is shown for
rubber in uniaxial compression and tension. The deformation is completely reversible
up to levels of several hundreds of per cent.
By generalising (6.45.) to a rubber sample, a relationship between the nominal
stress and the strain can be calculated from the kinetic theory of rubber elasticity.

(6.46)

As shown in figure 6.13., the experimental points fit the theoretical curve for
relative strains λ = x/x from 0.4 (uniaxial compression) to 1.5 (uniaxial extension).
0

The Young's modulus Ε is equal to the tangent to the stress - strain curve at
λχ = 1. From the equation of the tangent, it can be demonstrated that the modulus of
elasticity Ε is proportional to the number η of moles of elastic chains per unit volume
and to the absolute temperature T:
E = 3nRT (6.47.)
138 Introduction to Materials Science

0,8

^^cf^y Extension
0,0

-0,8

/ Uniaxial
/ Compression

-2,4

I
0,5 1,0 1,5 2,0
Relative deformation λ -χ /χ
χ 0

Figure 6.13. Elastic response of a natural rubber vulcanised with 8 % sulphur, in uniaxial
compression and extension. The stress is expressed in relation to the area of the
undeformed sample (nominal stress). The continuous line shows the theoretical curve
calculated from the gaussian theory of rubbery elasticity; the points correspond to
experimental results. (After Treloar, 1975).

In (6.47), R is the ideal gas constant, η is related to the density p . The molecular
weight Me of the elastic segments is given by:

n=-£- (6.48.)
e

Finally, the following expression is obtained for the modulus of elasticity of rubber:

E = 3 p R T
(6.49.)
Me

To a first approximation, the modulus of elasticity of an elastomer does not


depend directly on the chemical structure. The modulus of elasticity Ε increases
proportionally with the absolute temperature, as does the force of retraction. As al­
ready shown, this is a result of the almost exclusively entropic origin of rubbery
elasticity. The mechanical deformation leads to an alignment of the elastic chains
accompanied by a reduction in the entropy of the system (increased order). During
the deformation of rubber, the bond angles and the distance between the neighbour­
ing atoms do not vary. The internal energy (or enthalpy) therefore remains practically
constant irrespective of the deformation (figure 6.7. (d)).
Elastic behaviour of solids 139

6.4. Relationship between stress and strain in viscoelastic


materials
6.4.1. Viscous behaviour of Newtonian fluids
It has been seen in paragraph 6.2.2. that when an elastic body is deformed in
uniaxial tension, the stress varies proportionally with the strain (deformation), the
slope of this curve defines the Young's modulus. For Newtonian fluids subjected to
simple shear, a linear relationship exists between the stress and the deformation
(shear) rate defining the viscosity η:

σ = η ^ (6.50.)
1
dt
It should be noted that in hydrostatic compression, the Newtonian fluids exhibits
elastic behaviour (E = G =~ 0 ; Κ » 0).

6.4.2. Characterisation of the mechanical behaviour of materials


The mechanical behaviour of a substance can be characterised in a relatively
simple manner by applying a step deformation:
ε*(0 = ε*,ο«(0 (6.51.)

where e $is a constant and u{t) a step function where:


x

Owheni < 0
u(t) = (6.52)
1 when t > 0

The response of an elastic material to a step deformation (figure 6.14. (a)) is also a
step function (figure 6.14. (b)):
ο(ή = Εε Μί) χ (6.53.)
while in the case of a viscous fluid (figure 6.14. (c)), the response is a Dirac impulse:

σ(0 = ηε , —= ηε , ό\ήχ 0 χ 0 (6.54.)


dt
where ö\t) i s th e Dira c function .
Elastic material s an d viscou s fluid s constitut e tw o limitin g cases . Substance s
exist whic h sho w intermediat e behaviour . Thes e ar e th e viscoelasti c materials . Whe n
a viscoelastic body i s subjecte d t o a ste p deformatio n £*(/ ) = e w(0>th e stres s whic h
Xj0

must b e applie d t o kee p th e deformatio n constant , diminishe s progressivel y ove r tim e


(figure 6.14 . (d)) .
This behaviou r appear s principall y i n th e cas e o f thermoplasti c o r slightl y cross -
linked organi c polymers . I t i s du e t o th e anisotrop y o f th e fiel d o f inter - an d intra -
molecular forces , whic h i s induce d b y th e presenc e o f strong , covalen t chain - bond -
ing force s alongsid e relativel y wea k interchai n bonding . Whe n a mechanica l stres s i s
applied, th e chain s d o no t mov e instantaneousl y t o th e ne w equilibriu m position s an d
the mechanica l propertie s evolv e ove r time .
140 Introduction to Materials Science

Figure 6.14. Response of different mediums to an instantaneous step deformation (a): (b)
elastic material; (c) viscous fluid and (d) viscoelastic material where η is the viscosity of
the fluid and b\t) a Dirac impulse.

In viscoelastic materials, the strains and stresses measured at a particular time /


depend on the whole history of strains and stresses previously experienced by the
material. This is known faded memory, as it retains a memory for the history of the
stresses or strains previously experienced in the course of manufacture or in use, but
the effect of these stresses or strains dies away progressively.
Viscoelastic behaviour is also exhibited in a large number of metals and ceram­
ics when they experience temperatures close to the melting point or, for amorphous
materials, close to the glass transition temperature. At these high temperatures (T >
0.5 T , in degrees Kelvin), thermal agitation becomes significant. Polymers and cer­
m

tain metals (Sn, Pb) melting at temperatures below 600 Κ are already at an elevated
temperature under ambient conditions (300 K), and this gives rise to the appearance
of viscoelastic behaviour (chapter 12.).
Mechanical models are sometimes used to represent viscoelastic behaviour. The
behaviour of the ideal elastic solid and the viscous fluid are represented respectively
by a spring and a dashpot (figure 6.15. (a) and (b)). By combining springs and dash-
pots in series or in parallel, it is possible to simulate all types of viscoelastic beha-
Elastic behaviour of solids 141

viour. In figures 6.15. (c) and (d) two basic mechanical models are shown: that of
Maxwell (figure 6.1. (c)) and that of Voigt-Kelvin (figure 6.15. (d)). Below the me­
chanical models the different mechanical equations are given describing their be­
haviour. As an example, the Maxwell model, consisting of a spring and a dashpot
placed in series, will be considered; a stress σ is applied during a very short time dt.
During the application period, the first term on the right of the Maxwell equation is
much more significant and the Maxwell equation reduces to:

te lfo
= (6.55.)
at Ε dt
Equation (6.55.) is equivalent to that of an elastic solid. After much longer times than
that necessary for the application of the stress, the first term on the right hand side of
the Maxwell equation is zero and the equation reduces to that for a Newtonian fluid:

^£ = £ (6.56.)
dt η

This simple example shows that, as a function of the measurement time, a vis­
coelastic body behaves either as an elastic solid or as a viscous fluid. The Maxwell
model and that of Voigt-Kelvin (figure 6.15. (d)) are in general too simple. To repre­
sent quantitatively viscoelastic behaviour, a combination of the Maxwell and Voigt-
Kelvin models should be used. It must be emphasised that a large number of me­
chanical models exist which are capable of representing the behaviour of a
viscoelastic material in a quantitative manner without having any direct molecular
significance.

σ
σ
σ σ

1
σ
e [iL
τ σ
ν

σ
>
I ¥ σ

σ=Εε άε 1 άσ σ
σ = εΕ η ^
άΐ " Ε dt
+
n +

(a) (b) (c) (d)

Figure 6.15. Elementary mechanical models used to simulate viscoelastic behaviour: (a)
elastic solid: spring; (b) viscous fluid: damper (dashpot); (c) Maxwell model: spring and
dashpot in series; (d) Voigt-Kelvin model: spring and dashpot in parralel.

6.4.3. Extension of Hooke's Law to linear viscoelasticity


In elasticity, the relation between stress and strain is independent of the manner
in which the stress or the strain is applied. In viscoelasticity, this is not the case and it
142 Introduction to Materials Science

is indispensable to define the experimental protocol used to apply the stress or the
strain. Thus, the concept of deformation mode is defined. In this treatment, the study
will be limited to two deformation modes: relaxation and creep.
Stress relaxation consists of applying a step deformation to a material and ob­
serving the evolution of the stress as a function of time (figure 6.16. (a)). In a creep
experiment, stress is applied as a step function and the evolution of the strain is
studied (figure 6.16. (b)). Relaxation and creep are two important deformation
modes. It is important to avoid confusion between deformation modes and deforma­
tion types, which are uniform compression, simple shear and uniaxial extension. In
the general theory of elasticity, the deformation types are related the form of the
stress and strain tensors while the deformation modes are related to their time de­
pendence.

Ö
•ess
1
in

t - t —-
t
Retarded stress

Figure 6.16. Two deformation modes encountered in viscoelasticity: (a) stress relaxation;
(b) creep.

It is possible to apply the various deformation modes such as stress relaxation or


creep to each deformation type (uniform compression, simple shear, uniaxial exten­
sion) previously introduced (section 6.2.). Stress relaxation in uniaxial extension can
be taken as an example. In this case, ε (ή = ε*, and, from this experiment the relaxa­
χ 0

tion modulus E (t) can be defined:


r

E (t) = ^
r (6.57.)
£
x,0

In a creep experiment, it is a stress step function which is applied: (σ*(0 = σ ,ο). χ

The strain then varies as a function of time and the creep compliance J£t) is defined
by the relationship:

J ()
c t = £M (6.58.)

For elastic materials, the creep compliance is the inverse of the relaxation modulus
since o and £ are constants:
x x
Elastic behaviou r o f solid s 143

Er=°Ä_ = 2*&. = ± (6 .59.)


£χ,0 Jc

In the case of viscoelastic materials, E (t) and 1/ J (i) are distinct functions of time.
r c

The expressions (6.57.) and (6.58.) are a generalisation of Hooke's law, already
defined in (6.3.) for a purely elastic body. This type of law is only applicable up to a
certain value of stress or strain known as the limit of linear viscoelasticity. This limit
is determined by a series of experiments where the initial stress (in creep) or strain
(in relaxation) is progressively increased. To illustrate this concept, a procedure used
to determine the linear viscoelasticity limit for stress relaxation is described in figure
6.17. For example, four relaxation experiments are carried out by applying increasing

(c)

Linear
viscoelasticity
limit

Strain Εχ

Figure 6.17. Experimental determination of the limit of linear viscoelasticity: (a) applica­
tion of step deformations of increasing amplitude; (b) relaxation curves for the corre­
sponding applied stresses. Stress - strain curve determined from (b) for different relaxa­
tion times. The relaxation modulus for the different times considered is shown in (c).
144 Introduction to Materials Science

step deformations (figure 6.17. (a) and (b)). A certain number of values for time
(t\... t ) are selected and, for each value of t, a stress - strain curve (figure 17. (c)) is
5

plotted. The limit of linear viscoelasticity can be determined from the linear portion
of these curves.
The limit of viscoelasticity varies as a function of the modulus value. For po­
lymeric materials having a modulus larger than 1 GPa, the value is less than 1 %. For
polymers with a modulus of the order of 1 MPa, it exceeds 10 %. The treatment that
has been developed can be extended to simple shear and to uniform compression.
The viscoelastic behaviour in uniform compression, however, differs notably from
that encountered in uniaxial tension and simple shear. In uniform compression, the
viscoelastic behaviour is much reduced since the variation between the bulk modulus
of a glassy polymer and of a rubbery polymer is of the order of 50 %. The Young's
modulus and the shear modulus diminish by a factor of between 100 and 1000 under
these conditions between a glassy and a rubbery polymer.

6.4.4. Bolzmann's Principle


In o r d e r to calculate the relation between stress and strain for viscoelastic mate­
rials subjected to more complex stresses and strains, it is necessary to introduce the
concept of the linear superposition of effects, known in viscoelasticity under the
name of Bolzmann's superposition principle.
This was introduced in 1876 for creep. Boltzmann observed that the amount of
creep was a function of all the stresses previously applied to the sample and the effect
of the application of each stress was additive.
The method of introducing this principle is relatively simple (figure 6.18.). If a
viscoelastic material is subjected at time zero to a stress σ ,ο after time / the strain
χ ;

induced by this stress is equal to:


£ (ή=σ ,ο
χ χ Jc(t) (6.60.)

Figure 6.18. Bolzmann's principle: the strains e {t) = (t- t, )o j are additive in the same
Xti x

way as the stresses σ*,,.


Elastic behaviour of solids 145

If a stress σ is applied at time h to another sample, the strain equation is given


χ

by:
£ {ί)=σ
χ χΛ Jdt-h) (6.61.)

The successive application of two stresses σ , and σ to the same sample at χ 0 χΛ

times t = 0 and t = U will now be considered. Bolzmann's superposition principle


postulates that the two stresses are acting independently and that the overall strain is
the sum of the two separate strains:
£ (t) = G ,oJ (t)
x x c + o ,iJc(t
x - ti) (6.62.)

This superposition of strains is represented in figure 6.18.


The equation (6.62.) can be generalised to the case of any particular stress by
breaking it down into the sum of elementary stresses. So that, if between the times τ
and τ + dr, the stress varies by an increment da , this elementary stress da will in­ x x

duce after a time t > τ, a strain de of value: x

d£ = J (t-T)da
x c x (6.63.)

The total deformation at time t is obtained by the integration of (6.63.):

ε χ = μ (ί-τ)άσ € χ = J J c ( t - r ) ^ d t (6.64.)
0 -co dr

A time value of minus infinity (-«>) is taken as the lower integration limit, to show
that all previous stresses applied to the sample should be taken into consideration.
In an entirely analogous way, from a knowledge of the strain history ε*(ί), the
stress at time / can be calculated:
* d£
σ (ή=χ j£(i- )-idT T (6.65.)
-co at
It should be noted that the functions σ and £ are continuous and derivable functions χ x

of time.
By applying Bolzmann's principle, the relationship between the characteristic
functions E (t) and J (t) can be calculated and the following approximate expression
r c

obtained:
E (t)= r
S h l ( m 7 C ) 1
(6-66.)
mn J (t)
c

In this expression, m represents the slope of the function, log J (t) - log t. When c

the slope m tends to zero, the function §m(mn)lmii = 1, then E (t) = l/J (t). This r c

situation corresponds to the limiting case of ideal elastic behaviour, which is a par­
ticular case of viscoelastic behaviour.

6.5 Summary and conclusions


In this chapter, the study of the behaviour of materials subjected to mechanical
stresses has been studied in an elementary manner. All forces (stresses) applied to a
146 Introduction to Materials Science

material result in a deformation (strain) inducing a change in the relative position of


the atoms. This deformation leads to the appearance of a force of retraction, which
tends to restore the sample to its original form. If, for a given deformation, the force
of retraction is constant over time, the material has an elastic behaviour. By consid­
ering only small deformations, there is proportionality between stress and strain (lin­
ear elasticity: Hooke's law).
It is possible to relate stress and macroscopic deformations to interatomic
bonding forces by considering two model materials: the perfect crystal and the ideal
rubber. A thermodynamic study has shown that the elastic retraction force is associ­
ated with an increase in internal energy (or enthalpy) of the solid or with a decrease
in its structural disorder (entropy). This corresponds to two limiting cases described
in this chapter, i.e. enthalpic solid (metals, ceramics, thermoplastics) and entropic
solids (rubbers or elastomers).
A certain number of materials exist (organic polymers at ambient temperatures,
metals exposed to high temperatures) which no longer obey this ideal elastic behav­
iour. In fact, these materials have a behaviour characterised as viscoelastic, i.e. in­
termediate between that of an elastic solid and that of a viscous fluid. By introducing
the concept of linear viscoelasticity, it is possible to treat this phenomenon in a
quantitative manner and to generalise Hook's law.
Finally, with the help of simple calculations, the theoretical maximum strength
of materials has been estimated and their fracture energy determined. Theory and
practice do not agree in a large number of cases. This lack of agreement, which is
particularly significant for metallic materials, is related, as will be shown in chapter
7, to the existence of crystal defects and, in particular, dislocations. The movement
of dislocations leads to the dissipation of energy, which increases the fracture energy.

6.6. Illustrative example: the ultra-light aircraft or the


realisation of Icarus' dream
Since Leonardo da Vinci, the various efforts made by man to fly by his own
power have in general ended in failure. The recent conquest of air by manpowered
aircraft owes its success to a large extent to the availability of very strong and very
light materials.
It was in 1977 that a Californian team took the British Royal Aeronautical Soci­
ety's Kremer prize by carrying out a take-off and a closed circuit flight between two
points half a mile (800 m) apart with the 32 kg Gossamer Condor (figure 6.19.).
Two years later, an improved version, the Gossamer Albatross (wingspan of
30 m and weighing 25 kg) crossed the English Channel and succeeded in flying a
distance of 36 km in almost 3 hours.
The Gossamer Albatross consists of a cabin suspended from a large braced wing
with the wingspan of a 100-seat airliner. A directional sail is fitted at the front and
the pedal-operated propeller, 4 m in diameter is fixed to the rear of the cabin.
With the exception of the braces, cables, and the propeller drive mechanism, the
structure of the Gossamer Albatross is completely made from organic materials. The
main strengthening member is made of epoxy-carbon fibre composite tubes produced
by filament winding. The ribs of the wing and the propeller are of expanded polysty-
Elastic behaviour of solids 147

Figure 6.19. The Gossamer Condor making the closed circuit flight for the Kremer prize
(after Grosser, photo: T.Akoma, 1981)

rene reinforced with carbon fibres. The extreme end of the ribs of the wing towards
the trailing edge is made of aromatic polyamide fibres (Kevlar® - figure 5.11) and the
wing is an ultra thin and ultra light sheet of poly(ethylene terephthalate) (Mylar® -
figure 5.6.).
The significant improvements in performance realised during the two years
between the Gossamer Condor and the Gossamer Albatross resulted from the choice
of materials. The Gossamer Condor was constructed of aluminium aviation tubing
and balsa wood. The Gossamer Albatross was almost entirely built from synthetic
organic materials.
To obtain the highest performance in flight, the structure has to be as rigid and
as light as possible. It can be shown that for a given degree of rigidity in flexure of a
beam, the larger the ratio 4Ejp then the lighter the mass of the structure. In this
relationship, ρ is the density and Ε the modulus of elasticity.
The figure 6.20 shows the relative positions of various materials in a modulus of
elasticity - density diagram. The dotted line gives the values of Ε and ρ correspond­
ing to constant values (-jE/p = Q for the performance index. If only two properties,
Ε and p , are taken into account, all the materials which lie on the line have, in princi­
ple, the same performance.
The dotted line in figure 6.20, which is shown for a value of constant C close to
8 ((GPa) tm~ ), allows the most interesting materials for aviation to be identified.
05 3

As can be seen in figure 6.20; wood, carbon fibre reinforced composites (CFRC),
some ceramics and diamond seem most attractive.
However, the final choice of a material is not made exclusively on a modulus of
elasticity and density basis. Other properties and other aspects such as the cost and
the ease of fabrication of large parts must also to be considered. Inevitably, this
eliminates diamond from materials suitable for aeronautics. Ceramics in general are
too brittle for this use. They cannot sustain mechanical and thermal shocks. Wood is
inhomogeneous because of the presence of knots and it is sensitive to humidity.
However, wood was widely used in the early days of aviation.
148 Introduction to Materials Science

The interesting position of composites based on carbon fibres or aromatic poly­


amide (Kevlar) fibres (figure 6.20.) makes them attractive to the aviator, which ex­
plains their increasing use in commercial aviation. In high performance gliders, the
use of composites has been widespread for many years. In the area of materials, ul­
tra-light aviation is evolving much more rapidly than commercial aviation.
In the future, progress linked to research in materials science combined with
computer aided design will very likely lead to a significant improvement in the per­
formance of materials used in the field of aeronautics.

0,1 1,0 1 0

Density (t/m )
3

Figure 6.20. Selection chart for materials for an aircraft wing. The dotted line gives the
values of Ε and ρ corresponding to constant values V£/p = C) for the performance index,
(after Ashby, 1993).

6.7. Exercises
6.7.1. A filament of polyamide (Nylon®) 66 (modulus of elasticity Ε = 2GPa) of 1
mm diameter is deformed elastically by a 50 Ν load. Calculate the deformation of the
filament under the action of this load as well as the elastic energy stored in the fila­
ment as a result of the deformation.
6.7.2. Determine the value of the Poisson's Ratio ν of a material, which deforms at
constant volume. In which type of material is this behaviour encountered?
6.7.3. During a free fall "bungee" jump (i.e. attached to an elastic cord) the potential
energy gained by the human body is equal to U = mgl. In this expression, m is equal
to the mass, g is the acceleration due to gravity (9.81 m s ) and / the length of the
-2
Elastic behaviour of solids 149

elastic cord (20 m) of which the diameter is 20 mm and the Young's modulus
0.01 GPa. Calculate the extension of the elastic during a jump with an individual of
65 kg. To simplify the calculation, the height of the free fall can be considered as the
length of the elastic corde.
6.7.4. A tensile stress a i s applied along the axis of a cylindrical aluminium bar of 15
mm diameter. Assuming that the deformation is totally elastic, calculate the maxi­
mum applied stress which would limit the lateral contraction of the cylindrical bar to
2 μιη. Poisson's ratio ν for aluminium is equal to 0.35 and the modulus Ε is 70 GPa.
6.7.5. A iron block has a volume V = 1 d m at atmospheric pressure. What would be
0
3

the change in the volume of this block if it were submerged at 1500 m below sea
level? It can be assumed that the temperature of the water remains constant. The
density ρ of seawater is equal to 1.025 t m" and the compression modulus Κ of iron
3

is 71.7 GPa.
6.7.6. Explain why the temperature of a rubber band increases when it is stretched.
6.7.7. The different forces acting on the tubes of a bicycle frame can be represented
by studying the case of a beam with a circular cross-section of diameter d and length
L subjected to a force F at its free end. In the elastic region, the deflection δ is:

s = F4Ü
3End 3

where Ε represents the Young's modulus. Neglecting the actual weight of the beam,
show that wood is preferable to aluminium to minimise the mass Μ of the beam for
the same value of tensile rigidity. It is given that the elasticity modulus of wood Ε in
the direction parallel to the fibres is 10 GPa and the density is approximately equal to
0.5 g cm" . For what reason are bicycles no longer made of wood?
3

6.7.8. What is the class of materials of which the specific fracture energy G is clos­
c

est to the theoretical value, equal to twice the specific surface energy y?
6.7.9. A sample of rubber, 12 cm long and with a square cross-section of l x l cm, is
stretched at 20 °C to a length of 30 cm by applying a stress of 2 MPa. Calculate:
• The Young's modulus of this elastomer and the number of moles of
elastic segments per cubic centimetre;
• The stresses which must be applied to this sample to stretch it to a
length of 20 cm at 20 °C and to a length of 30 cm at 100 °C.
6.7.10. Calculate the Young's modulus of an elastomer obtained by the vulcanisation
of a polybutadiene with a molecular weight M„ of 10 gmol" and after vulcanisation
5 1

with a molecular weight for the elastic segments M equal to 5-10 gmol" .
c
3 1

6.8. References and complementary reading


M. F. ASHBY, Materials Selection in Mechanical Design, Butterworth-Heinemann, Oxford, 1993.
A. H. COTTRELL, The Mechanical Properties of Matter, John Wiley, New York, 1964.
R. W. HERTZBERG, Deformation and Fracture, Mechanics of Engineering Materials, 4th ed., John
Wiley, New York, 1996.
150 Introduction to Materials Science

M. GROSSER, Gossamer Odyssey, The Triumph of Humanpowered Flight, Houghton Mifflin, Boston,
1981.
L. R. G. TREOLAR, The Physics of Rubber Elasticity, 3rd ed., Clarendon, Oxford, 1975.
I. M. WARD, D. W. HADLEY, An Introduction to Mechanical Properties of Solid Polymers, Wiley,
Chichester (U.K.), 1993.

You might also like