Anda di halaman 1dari 32

Chapter 3.

Rubber elasticity
3.1. Introduction

Natural rubber is obtained by coagulation of a latex from a tree called Hevea


Brasiliensis. It consists predominantly of cis-1,4-polyisoprene. Fossilised natural
rubber discovered in Germany in 1924 stems back about 50 million years. Columbus
learned during his second voyage to America about a game played by the natives of
Haiti in which balls of an elastic ‘tree-resin’ were used. The word ‘rubber’ is derived
from the ability of this material to remove (rub off) marks from paper, which was
noted by Joseph Priestley in 1770. Rubber materials are not restricted to natural
rubber. They include a great variety of synthetic polymers of similar properties. An
elastomer is a polymer that exhibits rubber elastic properties, i.e. a material that can
be stretched to several times its original length without breaking and which, upon
release of the stress, immediately returns to its original length. Rubbers are almost
elastic materials, i.e. their deformation is instantaneous and they show almost no
creep.

The unique character of rubber was discovered in 1305 by John Gough who
described his experiments and findings as follows:

‘Hold one end of the slip of rubber …. between the thumb and forefinger of each
hand; bring the middle of the piece into slight contact with the lips; …. extend the
slip suddenly; and you will immediately perceive a sensation of warmth in that part
of the mouth that touches it … For this resin evidently grows warmer the further it
is extended; and the edges of the lips possess a higher degree of sensibility, which
enables them to discover these changes with greater facility than other parts of the
body. The increase in temperature, which is perceived upon extending a piece of
Caoutchouc, may be destroyed in an instant, by permitting the slip to contract
again; which it will do quickly by virtue of its own spring, as oft as the stretching
force ceases to act as soon as it has been fully exerted.’ Gough made the following
comment about a second experiment: ‘If one end of a slip of Caoutchouc be
fastened to a rod of metal or wood, and a weight be fixed at the other extremity
….; the thong will be found to become shorter with heat and longer with cold.’

To convince yourself, please make the experiment. You will only need a strip of
rubber, a weight and a hair-dryer. Gough presented no good explanation to the
unexpected findings, i.e. that the stiffness increases with increasing temperature and
that heat is evolved during stretching. It took almost 50 years before the
thermodynamics of the rubber elasticity was formulated.

59
Figure 3.1. Entropy-driven elasticity of rubber materials

Rubbers exhibit predominantly entropy-driven elasticity. This was concluded


already by William Thomson (Lord Kelvin) (1857) and James Prescott Joule (1859)
through measurements of force and specimen length at different temperatures. They
discovered the thermo-elastic effects: (a) a stretched rubber sample subjected to a
constant uniaxial load contracts reversibly on heating; (b) A rubber sample gives out
heat reversibly when stretched. These observations were consistent with the view that
the entropy of the rubber decreased on stretching. The molecular picture of the
entropic force originates from theoretical work during the 1930’s. It was suggested
that the covalently bonded polymer chains were oriented during extension. Further
theoretical development occurred during the 1940's and the stress-strain behaviour
was traced back to the conformational entropy. The view that the long chain
molecules are stretched to statistically less favourable states still prevails (Fig. 3.1).
You can make a very simple demonstration with a piece of rope, which will act as a
model of the polymer chain. Take the ends of the rope with your two hands. If you
keep your hands close, the rope can take many different shapes. If you separate the
ends of the rope, fewer shapes are possible. Hence, the number of shapes the rope can
take decreases with the displacement of the ends. The force acting on the polymer
molecule is equal to the slope of the free energy vs. displacement curve (Fig. 3.1). The
instantaneous deformation occurring in rubbers is due to the high segmental mobility
and thus to the rapid changes in chain conformation of the molecules. The energy
barriers between different conformational states must therefore be small compared to
the thermal energy (RT).
The reversible character of the deformation is a consequence of the fact that
rubbers are lightly crosslinked (Fig. 3.2). The crosslinks prevent the chains from
slipping past each other. The chains between adjacent crosslinks contain typically

60
several hundred main chain atoms. The crosslinks are covalent bonds in conventional
rubbers.

Figure 3.2. Crosslinked rubber. The crosslinks are indicated by filled circles

From a historical perspective, the accomplishment by Charles Nelson Goodyear in


1839 of a method to vulcanize natural rubber with sulphur was a crucial
breakthrough. Sulphur links attached to the cis-1,4-polyisoprene molecules formed the
network structure which is a prerequisite for obtaining elastic properties (Fig. 3.3).
Goodyear got the idea from Nathaniel Hayward, who found that natural rubber
became less sticky by mixing it with sulphur. Goodyear found accidentally that
natural rubber and sulphur reacted at elevated temperatures: a coated piece of fabric
with rubber, sulphur and a few other ingredients was hang close to a stove and the part
in direct contact with the stove had greatly changed its properties. Goodyear
understood the importance of his finding and continued the experimenting. The
discovery of vulcanization led to a large increase in production of natural rubber from
750 tons at 1850 to 6000 tons in 1860. Scotland’s John Dunlop made use of
Goodyear’s discovery in1888 by placing an air-filled rubber tube on his son’s bicycle.
It worked and was the start of a new era.

Figure 3.3. Sulphur bridges in Figure 3.4. Structure of


vulcanized cis-1,4 -polyisoprene thermoplastic elastomers

61
The efforts in developing methods to make synthetic rubber started even before
Staudinger’s formulation of the macromolecular concept. Production of synthetic
rubber dates back to before World War I in Germany. The synthetic rubber research
program was then restarted in Germany in 1926. The development in USA initiated
some years later. Many synthetic rubbers were made in the succeeding decades.

Later development of rubber technology has involved peroxide crosslinking and


thermoplastic elastomers. The latter consists of block-copolymers with hard
segments (physical crosslinks) and flexible segments (Fig. 3.4). The crosslink-
domains can be either in glassy amorphous or semicrystalline states. These materials
can be processed by conventional thermoplastic processing methods at temperatures
above the glass transition temperature or above the crystal melting point of the hard
segment domains.

3.2. Thermo-elastic behaviour and thermodynamics: energetic and entropic elastic


forces

Fig. 3.5 shows the classical data for sulphur-vulcanized natural rubber of Anthony,
Caston and Guth (1942). At small strains, typically less than ! = L L0 < 1.1 (L and L0
are the lengths of the stressed and unstressed specimen, respectively), the stress at
constant strain decreases with increasing temperature, whereas at λ values greater than
1.1, the stress increases with increasing temperature. This change from a negative to a
positive temperature coefficient is referred to as thermoelastic inversion. Joule
observed this effect much earlier (1859). The reason for the negative coefficient at
small strains is the positive thermal expansion and that the curves are obtained at
constant length. An increase in temperature causes thermal expansion (increase in L0
and also a corresponding length extension in the perpendicular directions) and
consequently a decrease in the true λ at constant L. The effect would not appear if L0
was measured at each temperature and if the curves were taken at constant λ (relating
to L0 at the actual temperature). The positive temperature coefficient is typical of
entropy-driven elasticity as will be explained in this section.

The reversible temperature increase that occurs when a rubber band is deformed
can be sensed with your lips, for instance. It is simply due to the fact that the internal
energy remains relatively unchanged on deformation, i.e. dQ=-dW (when dE=0). If
work is performed on the system, then heat is produced leading to an increase in
temperature. The temperature increase under adiabatic conditions can be substantial.
Natural rubber stretched to λ=5 reaches a temperature, which is 2-5 K higher than that
prior to deformation. When the external force is removed and the specimen returns to
its original, unstrained state, an equivalent temperature decrease occurs.

62
Figure 3.5.. Stress at constant length as a function of temperature for natural rubber.
The extension ratio (L/L0) referring to a universal value for L0, is shown adjacent to
each line. Drawn after data of Anthony, Caston and Guth (1942).

It is a matter of importance to separate the elastic force into entropic and energetic
contributions. A stress acting on the rubber network will stretch out and orient the
chains between the crosslink joints. This will decrease the entropy of the chains and
hence give rise to an entropic force. The change in chain conformation is expected to
change the intramolecular internal energy. The packing of the chains may also change
affecting the intermolecular-related internal energy. Both the intra- and
intermolecular potentials contribute to the energetic force. The following
thermodynamic treatments yield expressions differentiating between the entropic and
energetic contributions to the elastic force.

According to the first and second laws of thermodynamics, the internal energy
change (dE) in a uniaxially stressed system exchanging heat (dQ) and deformation
and pressure volume work (dW) reversibly is given by:

dE = T dS - p dV + f dL (3.1)

where dS is the differential change in entropy, p dV is the pressure volume work and
f dL is the work done by deformation. It is appropriate to point out that the force is
obviously a vector (denoted f) but in this treatment is treated as a scalar (denoted f;
being the absolute value of the vector). The Gibbs free energy (G) is defined as:

G = H-TS = E + pV - TS (3.2)

63
where H is the enthalpy. Differentiating Eq. (3.2) gives:

dG = dE + p dV + V dp - T dS - S dT (3.3)

Insertion of Eq. (3.1) in Eq. (3.3) gives:

dG = f dL + Vdp - SdT (3.4)

The partial derivatives of G with respect to L and T are:

" !G %
$$ '' = f (3.5)
# !L & p,T

" !G %
$$ '' = ( S (3.6)
# !T & L, p

G is a state function, which means that the order of derivation is unimportant:

" " % % " " % %


$ ! $ !G ' ' = $ ! $ !G ' ' (3.7)
$ !T $# !L '& ' $ !L $# !T '& '
# p,T & p, L # L, p & p,T

By combining Eqs. (3.5-3.7), the following expression is obtained:

" !f % " !S %
$$ '' = ( $$ '' (3.8)
# !T & L, p # !L & p,T

The partial derivative of G with respect to L at constant p and constant T (from


Eq. (3.2) is :

" !G % " % " %


$$ '' = $$ !H '' ( T $$ !S '' (3.9)
# !L & p,T # !L & p,T # !L & p,T

64
Eq. (3.8) in Eq. (3.9):

" !H % " !f %
f = $$ '' + T $$ '' (3.10)
# !L & p,T # !T & p,L

The derivative of H with respect to L at constant p and constant T (from Eq. (3.2):

" !H % " % " %


$$ '' = $$ !E '' + p$$ !V '' (3.11)
# !L & p,T # !L & p,T # !L & p,T

Experiments show that the volume is approximately constant during deformation,


( !V !L ) p,T
" 0 . Hence,

" !H % " %
$$ '' = $$ !E '' (3.12)
# !L & p,T # !L & p,T

and

" !E % " !f %
f = $$ '' + T $$ '' (3.13)
# !L & p,T # !T & p,L

The first term, (!E !L ) p,T , is associated with the change in internal energy

accompanying deformation at constant pressure and temperature. The second term


originates from changes in entropy by deformation; note that
(!f !T )L, p = "(!S !L) p,T . The entropy- and internal energy-components of the elastic
force are not only associated with the orientation of the chains. An additional and
important contribution originates from the change in volume:

" !E % " % " % " %


$$ '' = $$ !E '' + $$ !E '' $$ !V '' (3.14)
# !L & p,T # !L & T,V # !V & T ,L # !L & p,T

Typical of rubbers is that (!V !L ) p,T is small. The change in internal energy
associated with volume change (!E !V )T , L is however substantial. Eq. (3.13) can be

65
applied to stress-strain data taken at constant pressure and the separation into entropy-
and internal energy-components of the elastic force is readily made. The drawback is
that internal energy-related force component has two parts, one is due to the energy
change associated with the change of conformation and the other is due to change in
volume on deformation.

Figure 3.6. Entropic and energetic contributions of the tensile elastic force at
constant as a function of extension ratio (λ) for natural rubber. Drawn after data from
Wood and Roth (1944). The method used to resolve these components is shown in the
insert figure.

Physically more interesting is to consider deformation at constant volume in order


to view only the direct effects of orientation on entropy and internal intramolecular
energy. An analogous expression to Eq. (3.13) can be derived for constant volume
conditions:

" !E % " !f %
f = $$ '' + T $$ '' (3.15)
# !L & V ,T # !T & V ,L

This equation is difficult to apply in an experiment with an uniaxially stretched


rubber specimen, because the hydrostatic pressure has to be adjusted to keep the
volume constant to counteract changes in volume caused by the stress-strain work.
However, such an experiment was carried out by Allen et al. (1963).

66
The difficult problem of eliminating the effect of volume changes on the internal
energy was tackled much earlier by Elliot and Lippmann and Gee. They showed that it
was possible to derive the change in internal energy at constant volume from stress-
strain measurements obtained at constant pressure according to the following
expression:

" !E % " !f %
$ ' ( f ) T$ ' (3.16)
# !L & V ,T # !T & p,*

where ! = L L0 , L0 being the length of the specimen at zero stress and temperature T.
The application of Eq. (3.16) to the early experimental data for natural rubber
obtained by Anthony et al. and Wood and Roth (1944) showed that internal energy
contribution to the elastic force at constant volume is small at λ < 2.7 (Fig. 3.6).
Treloar (1975) and Mark (1984) have collected f e f data (fe is the force component
relating to the change in internal energy at constant volume and f is the total force) for
natural rubber: f e f =0.18±0.03 (λ ≤2). Mark concludes from gathering data from
literature that f e f is not influenced by dilution, i. e. swelling of the network polymer
in a low molar mass solvent (Table 3.1). Thus, f e f is controlled by the
intramolecular energetics, i.e. the energy differences between different
conformational states. Table 3.1 shows results from experiments on a number of
polymers. Both negative and positive values of f e f are found.

Table 3.1. Energetic stress ratio of a few polymers (Mark, 1984)


Polymer Diluent v2a ( fe f ) b
Polyethylene none 1.00 -0.42
" n-C30H62 0.50 -0.64
" n-C32H64 0.30 -0.50
Natural rubber none 1.00 0.17
" " n-C16H34 0.34-0.98 0.18
" " decalin 0.20 0.14
Poly(dimethyl-
siloxane) none 1.00 0.25
Trans (1,4 - poly-
isoprene) none 1.00 -0.10
" decalin 0.18 -0.20
a
Volume fraction of polymer in network
b
Obtained at constant volume

67
Polyethylene shows a negative f e f value (-0.42). During stretching of crosslinked
(molten) polyethylene obviously a large entropy force builds up and the internal
energy (at constant volume) decreases because many gauche conformers are
transferred into trans states. The energetic force must then be negative.

Other polymers such as natural rubber and poly(dimethyl siloxane) exhibit positive
f e f values, i.e. the extended conformation is of higher energy than the unstrained
structure. The low-energy conformation of poly(dimethyl siloxane) is all-trans, but
this gives the chain a non-extended (‘circular’) form due to the difference in bond
angles for O-Si-O and Si-O-Si.

The ratio f e f at constant volume is thus related to the intramolecular energy of


the polymer chains and it can be shown [details about the derivation are presented by
Treloar (1975)] that:

! d ln r 2
! fe $
## && #
= T#
( 0 ) $& (3.17)
" f % V = constant # dT &&
" %

The temperature coefficient of the dimension of the unperturbed polymer


(( ) )
molecules, d ln r2 0 dT obtained from stress-strain data for a range of crosslinked

polymers is in accordance with estimates from viscometry of the chain dimensions in


theta solvents at different temperatures. Polyethylene shows negative f e f and
(d( ln r ) dT ) . The trans-content in polyethylene becomes progressively lower with
2
0

increasing temperature and hence the size of the random coil decreases with
increasing temperature.

3.3. The statistical mechanical theory of rubber elasticity

The early molecular-based statistical mechanical theory was developed by Wall


and Flory and Rehner, with the simple assumption that chain segments of the network
deform independently and on a microscopic scale in the same way as the whole
sample (affine deformation). The crosslinks are assumed to be fixed in space at
positions exactly defined by the specimen deformation ratio. This model is referred to
as the ‘affine network model’. James and Guth allowed in their ‘phantom network
model’ a certain fluctuation of the crosslinks about their average affine deformation
positions. These two theories are in a sense ‘limiting cases’ with the affine network
model giving an upper bound modulus and the phantom network model theory the

68
lower bound. It is important to emphasize that both models assumes that the chains of
the network behave like phantom chains, i.e. the dimensions of these chains are
unperturbed by excluded-volume effects. This assumption has later been confirmed by
small-angle neutron scattering (SANS) of labelled (deuterated) amorphous samples.

Figure 3.7 shows schematically the difference between the affine network model
and the phantom network model. The affine network model assumes that the junction
points (i.e. the crosslinks) have a specified fixed position defined by the specimen
deformation ratio (λ). The chains between the junctions points are however free to
take any of the great many possible conformations. The junction points of the
phantom network are allowed to fluctuate about their mean values (shown in Fig. 3.8
by the unfilled circles) and the chains between the crosslinks to take any of the great
many possible conformations.

We will derive an equation relating stress and strain in a rubber on the basis of
the affine network model. This model is based on the following assumptions: (1) The
chains between crosslinks can be represented by Gaussian statistics of phantom
chains. (2) The free energy of the network is the sum of the free energies of the
individual chains. (3) The positions of the crosslinks are changed precisely according
to the macroscopic deformation, i.e. deformation is affine. (4) The unstressed network
is isotropic. (5) The volume remains constant during deformation.

Figure 3.7. Schematic representation of the deformation of a network according to the


affine network model and the phantom network model. The unfilled circles indicate
the position of the crosslinks assuming affine deformation (phantom network).

69
Figure 3.8. Affine deformation of a single chain from unstressed state r0 =(xo,y0,z0) to
stressed state r =(λ1x0,λ2y0,λ3z0).

The words ‘chains of network’ or ‘Gaussian chains of network’ are frequently used
in this section. The two are synonymous and mean the piece of the network between
two adjacent crosslinks.

The derivation of the stress-strain equation goes through some elementary steps.
The Gaussian distribution function for the end-to-end distance expresses the
probability of finding the end of a chain at a certain position (x,y,z) with respect to the
other chain end found at (0,0,0). This equation expresses, to phrase it differently, the
number of conformations a chain can take provided that the chain ends are in (0,0,0)
and (x,y,z). It is then possible to calculate the Helmholtz free energy (the free energy
associated with constant volume) for a single chain and by adding the contributions
from all individual chains of the network, also for the network. The stress-strain
equation is finally obtained by taking the derivative of the Helmholtz free energy with
respect to length. Let us start by showing the Gaussian function describing the
distribution of the chain end position:

32 3 x2 +y2 +z
( )
" 3 %
(
2 r2
P(r ) = $ 2 ' e 0
(3.18)
# 2! r 0 &

where r2 is the average end-to-end distance of the phantom chains. The Helmholtz
0

free energy (A) is related to the conformational partition function Z(r) for a given
chain with the end-to-end vector (r) according to the following equation:

A(r ) = ! kT ln Z(r ) = ! kT ln P( r)Z( ) (3.19)

70
where Z is the partition for the unconstrained chain (taking all possible r values). The
following expression is obtained by inserting Eq. (3.18) in Eq. (3.19):

) 3 # 3 & 3kT ( x + y + z ) ,.
2 2 2

A(r ) = ! kT + ln Z + ln% ( ! (3.20)


+* 2 $ 2" r 2 0 ' 2 r2 0 .-

The end-to-end vector r0 = ( x 0 , y 0 , z0 ) characterizes the unstressed state of a single


chain in the network (Fig. 3.8). It may be noted that energetic effects are considered
by the inclusion of r 2 0 . The end-to-end vector r = ( x, y, z) corresponding to the
stressed state of the sane single chain is related to r0 through the deformation ratios
(!1 , !2, !3 ) according to:

x = !1 x 0 ; y = ! 2 y0 ; z = ! 3 z0 (3.21)

The free energies of the chain before (A0) and after (A) the stress has been applied
are:

" 3 (x 2 + y 2 + z 2 %
A0 = C + kT $
( 0 0 0 )' (3.22)
$ 2 r2 '
# 0 &

# 3 ( "2 x 2 + "2 y 2 + "2 z 2 &


! A = C + kT %
( 1 0 2 0 3 0 )( (3.23)
% 2 r2 (
$ 0 '

) 3 # 3 & ,.
! +
where C = !kT ln Z + ln % (!1 , !2 , !3 ) .
( , which is independent of The
+* 2 $ 2" r 2 0 ' .-

difference in free energy between the two states is:

% $2 #1 x 2 + $2 #1 y 2 + $2 #1 z 2 (
"A = A # A0 = 3kT '
( 1 ) 0 ( 2 ) 0 ( 3 ) 0* (3.24)
' 2 r2 *
& 0 )

! The change in free energy of the network (∆AN) is the sum of the contributions of
all chains of the network. It is here assumed that deformation is affine, i.e. all chains
are equally deformed according to the deformation ratios (!1 , ! 2 , !3 ) :

71
N N N
% 2 (
' ( #1 $ 1)" x 0 + (# 2 $ 1)" y0 + (# 3 $ 1)" z20 *
2 2 2 2
N
!AN = " !A = 3kT ' 1
2
1 1
* (3.25)
1 '' 2r 0 **
& )

It is assumed that the undeformed network is isotropic, i.e.:

N N N
2
!x 0 = ! y 02 = ! z20 (3.26)
1 1 1

The sum of the squares of the x, y and z components must be equal to:

N N N N

! x 02 + ! y 02 + ! z02 = ! r02
1 1 1 1 (3.27)

Insertion of Eq. (3.27) in Eq. (3.26) yields:

1 N 2 N r
2
N N N
2 2 2
!1 x = !1 y = !1 z = 3 !1 r0 = 3
0 0 0
0
(3.28)

where N is the number of chains of the network. Insertion of Eq. (3.28) in Eq. (3.25)
gives:

1
!A = NkT( "21 + " 22 + "23 # 3) (3.29)
2

Eq. (3.29) is general and is not restricted to any particular state of strain.
Let us derive a stress-strain equation for a rubber specimen subjected to constant
uniaxial stress. The extension ratio along the stress vector is denoted ! 1 = ! . It may
also be assumed that the transverse deformation ratios are equal: ! 2 = ! 3 = ! t . The
transverse deformation ratio can be calculated assuming constant volume (∆V= V0–V=
0) on deformation:

!V = "1 x 0 # " 2 y 0 # "3 z0 $ x 0 # y 0 # z0 = 0

72
! 1! 2 ! 3 " x 0 y 0 z0 # x 0 y0 z0 = 0
! 1! 2 ! 3 = 1 (3.30)

considering that ! 1 = ! and ! 2 = ! 3 = ! t :

! " !2t = 1 # ! t = 1 ! (3.31)

Hence, the following deformation ratios are obtained: ! , 1 ( !, 1 )


! . The force

(f) is obtained by inserting the deformation ratios in Eq. (3.29) and then take the
derivative of the free energy with respect to the length (L):

# ! ("A)& # ! ( "A) & # !) &


f =% ( =% ( (3.32)
$ !L ' T ,V $ !) ' T ,V $ !L ' T ,V

! $& NkT $ 2 2 '' $ ' $ '


&& " + # 3)) )) * ! && L )) = NkT && " # 12 ))
f = (3.33)
!" &% 2 % " ( ( !L % L0 ( L0 % " (

# A & nRT # &


f = ! %% 0 (( = %% " ) 12 (( (3.34)
$"' L0 $ " '

where σ is the real stress, A0 is the original specimen cross-sectional area, L0 is the
original length of the specimen parallel to the stress, n is the number of moles of
Gaussian chains and R is the gas constant. After simplification, the following stress-
strain equation is obtained:

nRT $ 2 1 & 1
!= %
" # ' = Ne0 RT $% "2 # &' (3.35)
V0 " "

where V0 is the volume of the specimen, Ne0 = n V0 is the number of moles of


Gaussian chains in the network per unit volume of rubber. The stress at a given
extension ratio is thus proportional to Ne0 . Another way of expressing the crosslink
density is:

73
n ! nMc $ ! 1 $ ! m0 $ ! 1 $ (
=# & '# & = # & '# &= (3.36)
V0 " V0 % " Mc % " V0 % " Mc % Mc

where M c is the number average molar mass of the Gaussian chains of the network
and ρ is the density. The equation relating the true stress and the extension ratio
becomes after insertion of Eq. (3.39) in Eq. (3.38):

" RT %' 2 1 (*
!= '# $ * (3.37)
Mc & #)

The ‘modulus’, !RT M c , is proportional to the temperature (in Kelvin). This is


typical of entropy-elastic materials. The other important aspect of eq. (3.37) is that the
modulus is inversely proportional to Mc . Rubbers with a high crosslinking density,
i.e. low M c , are stiff. Fig. 3.9 shows the thermoelastic behaviour of the ideal entropy-
elastic rubber material. All lines meet in the origin at 0 K.

Figure 3.9. Stress as a function of Figure 3.10. Schematic


temperature at constant λ representation of a tetrahedral
according the affine network model. network.

What is the relationship between the number of chains and the number of
crosslinks? A particular simple and common case is the tetrahedral network with
four chains meeting in each junction point (Fig. 3.10). The number of chains in the
network with ! = N " N junctions is Ne0 = N ! N (vertical) + N ! N ( horisontal) .
Hence, for an infinite tetrahedral network the following relationship holds:

74
! N e0 = 1 2 (3.38)

It can be shown that the following general equation relates the number of crosslinks
(ν) to the number of chains and the crosslink functionality (ψ= the number of chains
emanating from a junction point):

2N e0
!= (3.39)
"

which after insertion in Eq. (3.35) gives:

"#RT &( 2 1 )+
!= ($ % + (3.40)
2 ' $*

The other early statistical mechanical theory, the phantom network model of James
and Guth permits fluctuations of the junction points about average positions
prescribed by the macroscopic deformation. The derivation of the free energy-strain
equation is here complicated by the fact that the chains of the network are coupled and
the probability function for the network is the product of the probability functions of
the individual chains. The lengthy derivation is not shown here but some important
results of the derivation are highlighted. The average positions of the junctions are
deformed affinely and the average fluctuation of any given junction is independent of
strain and can be described by a Gaussian probability function. The average force on a
chain is the same as they were fixed at their most probable position. The forces
exerted by the network are the same whether any given junction is treated as free, or
as fixed at its most probable position. In all, these findings justify much of the
assumptions made in affine network model and it is no surprise that the free energy
equations and the derived stress-strain equations are similar for the two models. The
free energy-extension ratio expression derived by James and Guth is:

$ 2 ' NkT 2
!A = && 1 " ))
% #( 2
(*1 + *22 + *23 " 3) (3.41)

which, for an infinite tetrahedral network (ψ=4) becomes:

75
$ 2 ' NkT 2 1
"A = &1# )
% 4( 2
( *1 + *22 + *23 # 3) = NkT ( *12 + *22 + *23 # 3)
4 (3.42)

! which is precisely half of the value predicted by the affine network model (cf. Eq.
(3.29). The following true stress-extension ratio equation is obtained for the phantom
network in the case of a uniaxial stress:

$ 2 ' #*RT $ 2 1'


! = && 1" )) && + " )) (3.43)
% #( 2 % +(

For the tetrahedral network with the crosslink functionality (ψ) of 4, the phantom
network model predicts a modulus equal to ! RT , which is 1/2 of the modulus
predicted by the affine network model ( 2!RT ; insert ψ=4 in Eq. 3.40).

It is common in the literature to show the stress-strain equation in another form:


nominal stress (σN) defined as force divided by the original cross-sectional area.
Provided that the volume is constant during deformation, the nominal stress and the
true stress (σ, force divided by actual cross-sectional area) are related as follows:

"
"N = (3.44)
#

which after insertion in eq. (3.35) gives:


!
% 1(
N e0 RT' #2 $ *
& #) % 1(
"N = = N e0 RT' # $ 2 *
# & # ) (3.45)

(3.37, 3.40 and 3.43). The nominal stress is approximately proportional to λ at high λ
! because then ! >> 1 ! 2 , It may be also noted that the force is always proportional to
the nominal stress and hence also the force is also proportional to (! " 1 !2 )
according to the statistical mechanical theory.

3.4. Comparison of predictions made by theory and experimental data

Fig. 3.11 shows how theory compares with experimental nominal stress-extension
ratio data for natural rubber and silicone rubber. The modulus G = KNe0 RT - for
tetrahedral network is K=1 (affine network model) and 1/2 (phantom network model)
– is basically used an adjustable parameter fitting the theoretical equations to the

76
experimental data. The theoretical equation captures the trend in the experimental data
in compression (λ<1) and at low extension ratios (λ<1.2). At higher extension ratios
the experimental data fall below the theoretical curve. This behaviour is quite general
to rubber materials and it indicates a flaw in the fundamental statistical mechanical
theory.

Figure 3.11. Nominal stress-extension ratio data for natural rubber and poly(dimethyl
siloxane) collected by Higgs and Gaylord (1990). The data are scaled to fit the
relation [! ] = (" # "#1 ) near λ=1.

Figure 3.12. Mooney diagram (reduced modulus as a function of the reciprocal


extension ratio) for natural rubber swollen in benzene. The volume fraction of
polymer (v2) in the swollen network is shown adjacent to each line. The insert figure
shows C2 (constant in the Mooney equation) as a function of volume fraction of rubber

77
for natural rubber, butadiene-styrene rubber and butadiene-acrylonitrile rubber.
Drawn after data from Gumbrell et al. (1953).

Figure 3.13. Schematic Mooney diagram showing the upper bound (affine network)
and lower bound (phantom network) and the gradual change between the two
according to the constrained-junction model. The trend in experimental data is
displayed with the thick line. Drawn after Mark (1993).

The Mooney equation, which was derived from the assumption that Hooke’s law
is obeyed in simple shear, is in accordance with the experimental data in this
particular extension ratio range (1<λ<2):

% 1 (% C (
" N = 2' # $ 2 *'C1 + 2 * (3.46)
& # )& #)

! where C1 and C2 are empirical constants. It may be noted that the Mooney equation is
identical with the equations derived from statistical mechanics when C2=0. In a
[ ]
Mooney diagram, " N 2( # $1/ #2 ) plotted versus 1/λ, C2 is obtained as the slope
coefficient and C1+C2 is the intercept at 1/λ=1. The reduced stress or modulus
[ ]
E* = " N v1/2 3 ( # $1/ #2 ) is often used. The factor v21/ 3 considers that the number of
load-bearing
! chains is reduced in a swollen system. The modulus changes from
2(C1+C2) at λ=1 to 2C1 at λ=∞. The ‘built-in’ gradual decrease in modulus of the
! Mooney equation makes it very useful in capturing the experimental data (Fig. 3.12).
Gumbrell et al. showed for a range of rubbers that C1 depended on the crosslink
density basically according to the statistical mechanical theory, whereas C2 remained
approximately constant. The stress-strain behaviour of rubbers swollen in organic
solvents shows some interesting general features (Fig. 3.13): C1 is practically
independent of the degree of swelling, whereas C2 decreases with increasing degree of
swelling approaching C2 = 0 at a volume fraction of rubber in the swollen system of

78
0.2. Hence, highly swollen rubbers obey the stress-strain behaviour according to the
fundamental statistical mechanical theories.

The gradual decrease in modulus with increasing extension ratio at moderate


extensions captured by the Mooney equation suggests that the response of the network
mediates between affine deformation at low extension ratios to phantom-network-type
deformation at high extension ratios (Fig. 3.13). The two models, affine network and
phantom network may be considered as limiting cases. It seems thus that the
fluctuations of the junctions are suppressed at low extensions and that further
extension makes the junctions more mobile to fluctuate. Flory and coworkers
proposed the constrained junction model, which introduced a constraint parameter
(! ), which at high extension takes the value 0 (phantom network) and at low
extensions (λ≈1) takes the value ∞ (affine network). This development of the
fundamental statistical mechanical theories bringing one additional adjustable
parameter is thus capable of describing the stress-strain data in the intermediate
extension ratio range (1.2<λ<2).

An even more critical test of the theory would be to see if the modulus predicted by
the theory is according to the experimental data. The simplicity (and beauty) of the
statistical mechanical theory is that the modulus is related to only a single material
parameter ( Ne0 ) . The problem is that this quantity is seldom known by independent
measurements. In fact, the crosslink density is often determined by applying the
statistical mechanical theory to experimental stress-strain data. An obvious approach
to the problem is use ‘selective’ chemistry to produce model networks of known
crosslink density (exact networks). The issue is still not resolved. Some model
networks show moduli according to the basic statistical mechanical theories; the
stress-strain characteristics of real networks is between the limits set by the affine
network (upper limit in modulus) and phantom network (lower limit in modulus) with
the closer resemblance to the affine network at low extension ratios. Other researchers
have reported experimental moduli higher than those predicted by the statistical
mechanical theory. A tentative explanation to discrepancy between theory and
experimental data is the presence of trapped chain entanglements not accounted for in
the basic statistical mechanical theory.

Let us return to Fig. 3.11. At very high extension ratios (λ>3) the experimental
nominal stress – extension ratio data show an upturn which is clearly inconsistent with
the fundamental statistical mechanical theory which predict a linear relationship
between these quantities at high extension ratios. Two different explanations to the
upturn have been proposed:

79
b) The Gaussian probability function that describes the chain statistics is a
reasonable accurate approximation at low extension ratios but certainly not
when the chains of the network become highly extended. The refinement of the
rubber elastic theory considering the non-Gaussian statistics is discussed in
section 3.6.

(b) The orientation of the chains molecules of the network reduces their entropy
(Smelt) and the equilibrium melting point ( Tm0 = !H !S , where ∆H is the heat of
fusion and ∆S is the change in entropy on melting) should increase because
!S = Smelt " Scrystal is reduced by orientation. Hence, the effective degree of
supercooling, !T = Tm0 " T increases with increasing extension ratio and the
tendency for regular polymers like natural rubber to crystallise increases with
the extension ratio. Goppel showed that natural rubber crystallised at λ≈4 when
stretched at room temperature. Treloar recorded the rate of crystallisation by
measurement of the density of stretched natural rubber at 0°C and found that
crystallisation occurred at all extension ratios, but the rate of crystallisation
greatly increased in the highly stretched rubbers. It is reasonable to assume that
a semicrystalline rubber is stiffer than its fully amorphous analogue and in fact,
Flory suggested that the upturn in the force-extension ratio curve was due to
crystallisation. The minor change in the force-extension ratio curvature on
heating natural rubber to 100°C, where basically no crystallinity remains,
reported by Wang and Guth and the X-ray diffraction data of Smith et al.
suggest, however, that the upturn in the force-extension ratio curve is dominated
by the non-Gaussian effect.

3.5. Deviations from classical statistical theories for finite-sized and entangled
networks

In the classical theories, it is assumed that the network is infinite, i.e. that no loose
chain ends exist. Loose chain ends transfer stress less efficiently than the other parts
of the network and it may be assumed that they have no contribution to the elastic
force. In the classical treatments it assumed that the chains are phantom chains and
that they can take any conformation possible to a single phantom chain in vacuum.
This is not a true picture because chain entanglements may be present. Two different
kinds of junctions may thus exist: chemical crosslinks and trapped chain
entanglements.

Let us start by considering the effect of chain ends and to see how it affects the
stress-strain function. Each linear molecule prior to crosslinking has two chain ends
and these form cilia in the network structure even after vulcanisation (Fig. 3.14). The

80
number of chain ends (cilia) per unit volume in the network is Nce = 2 !N A M , where
NA is the Avogadro number and M is the number average molar mass of the polymer
prior to crosslinking. The number of load-bearing chains (Ne) per unit volume is:

"NA 2 "N A
Ne = Ne0 ! Nce = ! (3.47)
Mc M

where Ne0 is the number per unit volume of load-bearing chains in an infinite network.

The ratio Ne Ne0 becomes:

Ne !N A Mc " 2 !NA M 2Mc


0 = =1 " (3.48)
Ne !NA Mc M

Figure 3.14. Chain ends before and after vulcanization.

The stress-strain equation for a uniaxially stressed rubber can be modified to take
into account that only the fraction Ne Ne0 of the Gaussian chains carries the load:

"N % 1 " 2Mc % " 2 1 % * RT " 2Mc % " 2 1 %


! = Ne0 $ e0 ' RT "# (2 ) %& = Ne0 RT $ 1 ) ' ( ) &= $ 1) ' ( ) &
# Ne & ( # M &# ( Mc # M &# (
(3.49)

The second factor, i.e. (1 ! 2Mc M ) , becomes important when M is of the same

order of magnitude as Mc .

Other types of network defects also exist, physical crosslinks and intramolecular
crosslinks (loops), as displayed in Fig. 3.15. The physical crosslinks may be
permanent with a locked-in conformation (Fig. 3.15; case a) or temporary by
entanglement (Fig. 3.15, case b). The presence of the latter type leads to visco-elastic

81
behaviour, i.e. to creep and stress relaxation. Intramolecular crosslinks decrease the
interconnectivity of the network and reduce the number of load-carrying chains
(Fig. 3.15, case c).

Figure 3.15. Imperfections in real networks: (a) trapped chain entanglement; (b)
temporary chain entanglement; (c) intramolecular crosslink.

The role of the chain entanglements is still under debate. The early work was due
to Mullins, who argued that the number of trapped (such which cannot disentangle)
entanglements should increase with increasing degree of crosslinking. Mullins
proposed that the number of effective crosslinks should be the sum of chemical (νchem)
and trapped entanglements (νent):

$ #Mc '
! e = (! chem + ! ent )& 1" ) (3.50)
% M (

where β is an empirical constant. The chain ends not carrying load are accounted for
in the second factor in Eq. (3.50). By inserting Eq. (3.50) in the expression for C1 of
the Mooney equation, the following expression was obtained:

$ #Mc ( chem) '


( %
)
C1 = C1Mc (chem) + ! & 1"
M
)
(
(3.51)

where C1Mc ( chem ) is the C1 value corresponding to the pure chemical crosslinks
[described by Mc (chem ) ] and 1/M=0, and α is the term that is related to the
entanglements. Eq. (3.51) could be fitted (two adjustable parameters: α and β) to
experimental data of crosslink density by chemical methods and stress-strain
measurements. The presence of trapped entanglements in networks was further
supported by the more recent data of Gottlieb et al. and Dossin and Graessley of the

82
shear modulus as a function of crosslink density display non-zero intercepts, i.e. a
certain shear modulus at zero-concentration of chemical crosslinks.

3.6. Large deformations when the Gaussian approximation is not valid

The pronounced upturn in the nominal stress-extension ratio (λ) graph appearing at
λ>3 is a general feature characteristic of basically all rubbers (Fig. 3.11). The
statistical mechanical theories based on Gaussian chains predict that the nominal
stress should be proportional to the extension ratio in this extension ratio range. It is
now generally accepted that these early theories neglect the finite extensibility of the
network. To phrase it differently: the Gaussian distribution function is a poor
descriptor of the chain statistics of highly extended chains. Fig. 3.16 presents a
comparison between results obtained by Monte Carlo simulation of a 50-carbon chain
and the theoretical (Gaussian) distribution. It should be noted that P(r ) = P (r) ! 4"r 2 ,
where P(r) is given by eq. (3.21). The Gaussian distribution function predicts too high
probabilities for the extended conformations as is illustrated by Fig. 3.16. However, at
low extension ratios (below and around the maximum of the P(r) distribution), the
theoretical distribution is in accordance with the results obtained by Monte Carlo
simulation.

Figure 3.16. Radial distribution function [P(r)] of a 50-carbon-alkane at 400 K


obtained by Monte-Carlo simulation using Lattice™ (Nairn, 1998) assuming that the
chains are unperturbed by excluded volume represented as a histogram. The
Gaussian approximation is shown as the continuous line.

83
The above is not a shortcoming of statistical mechanics but rather of the Gaussian
approximation, which is valid only at low chain extension ratios. The problem can be
‘solved’ by considering non-Gaussian chain statistics. Kuhn and Grün treated the
problem for a single chain and they showed that the radial distribution function is
approximately given by:

"" r % " ( %%
ln P(r) = C ! n$$ $ ' ( + ln $$ '' ' (3.52)
# # nl & # sinh ( & '&

where C is a constant and β is a parameter related to the extension of the chain (r/nl):

r 1
= coth " # = L (" ) (3.53)
nl "

where L is the Langevin function. The parameter β is explicitly given by:


!

" = L#1 ( r nl) (3.54)

where L -1 is the inverse Langevin function. Eq. (3.52) can be written in a more useful
!
form by making use of series expansion:

( 3 " r % 2 9 " r % 4 99 " r % 6 +


ln P(r) = C ! n* $ ' + $ ' + $ ' + .....- (3.55)
*) 2 # nl & 20 # nl & 350 # nl & -,

where C is a constant. For small chain extension ratios (i.e. small r/nl values) the
higher terms of the series can be neglected and

3r2
ln P(r) = C ! (3.56)
2nl 2

which is identical with the Gaussian distribution function. The force-extension ratio
equation can be obtained by first by transforming P(r) into P(r), calculating the
entropy using the Boltzmann entropy equation [S=k ln P(r)] and by differentiating the
entropy with respect to r [ f = !T ("S "r ) ]:

" kT % " r %
f = $ 'L(1$ ' (3.57)
# l & # nl &

84
!
which can be expanded to:

kT ' !r$ 9 ! r $ 3 297 ! r $ 5 1539 ! r $ 7 *


f = )3# & + # & + # & + # & + ...., (3.58)
l )( " nl % 5 " nl % 175 " nl % 875 " nl % ,+

The first term in Eq. (3.58) corresponds to the Gaussian approximation, i.e. the
linear portion of the f-r curve. The following terms produce a pronounced upturn at
higher extensions in the force-extension curve (Fig. 3.17).

Figure 3.17. Force as a function of extension for a single polymer molecule according
to the non-Gaussian theory (eq. (3.58). The broken line shows the Gaussian
approximation. Drawn after Treloar (1975).

Figure 3.18. Nominal stress (divided by NkT) as a function of extension ratio for the
three-chain model with different number of bonds (n) as shown adjacent to each curve
The lower line shows the Gaussian approximation. Drawn after Treloar (1975).

85
Figure 3.19. Nominal stress as a function of the extension ratio for natural rubber.
The points are experimental data and the continuous line shows the best fit of eq.
(3.59) to the experimental data. The lower line shows the Gaussian approximation.
Drawn after Treloar (1975).

It is much more of a challenge to develop a non-Gaussian theory for polymer


networks. It should be remembered that the previous mathematical treatment is
concerned with a single molecule. Treloar (1975) stated that the conclusions arrived
for networks must be regarded as somewhat uncertain in the quantitative sense.
Simplest but still very useful is the three-chain model. The network is here replaced
by sets of chains parallel to three orthogonal axes; x, y and z. The entropy of the
system is the sum of the entropies of the three chains affinely deformed. The entropy
of each chain is obtained by transforming P(r) into entropy and in the simple case of
uniaxial deformation inserting the following r values: rx=λr0 and ry=rz=λ−1/2r0 in the
entropy equation. The force-extension ratio expression is obtained according to the
method used for the simple Gaussian network:

NkT + #1% $ ( # 3 #1% 1 (.


f = " n -L ' * # $ 2L ' *0 (3.59)
3 , & n) & $n )/

The three-chain model predicts that the onset of curvature in the force-extension
!
ratio diagram occurs at very low extension ratios for short molecules (dense networks)
and that this extension ratio progressively increases with increasing molar mass of the
chains (Fig. 3.18). Very short chains, e.g. n= 5 or less, the mean chain extension even
in the unstrained state already exceeds that for which the Gaussian approximation is
valid. For networks of such short chains, a non-Gaussian treatment is essential for the
accurate representation stress-strain even at the smallest strains (Fig. 3.18). Fig. 3.19
shows that Eq. (3.59) is capable of fitting experimental data at high extensions (λ>3).

86
3.7. Summary

Conventional rubbers are crosslinked amorphous polymers well above their glass
transition temperature. The elasticity of rubbers is predominantly entropy-driven
which leads to a number of spectacular phenomena. The stiffness increases with
increasing temperature. Heat is reversibly generated on deformation.

A more detailed analysis shows that the elastic force originates both from changes
in conformational entropy and changes in the internal energy. The latter are normally
small and at constant volume they relate to changes in conformational energy.
Polymers like polyethylene with an extended conformation as their low energy state
exhibit a negative energetic force contribution, whereas other rubbers, a notable
example is natural rubber, show positive energetic forces.

Statistical mechanical models have proven useful in describing the stress-strain


behaviour of rubbers. The affine network model assumes that the network consists of
phantom Gaussian chains and that the positions of the junctions are prescribed by the
macroscopic deformation. The phantom network model assumes that the positions of
the junctions fluctuate about their mean positions prescribed by the macroscopic
deformation ratio. The change in Helmholtz free energy (∆A) on deformation at
constant temperature (T) is due to decrease in the number of possible conformations
of the chains of the network:

!A = KNkT ("21 + "22 + "23 # 3) (3.60)

where K is dimensionless number different for the different models, N is the number
of chains in the network, k is the Boltzmann constant, T is the temperature (in K) and
λi are the extension ratios along the axes of an orthogonal coordinate system. Eq.
(3.60) can be used to obtain the stress-strain equation for different stress states. For
the case of a rubber specimen subjected to a constant uniaxial stress, the following
true stress (σ)-strain (λ) expression holds:

1
! = KN e0 RT $% "2 # &' (3.61)
"

where K is dimensionless number different for the different models, Ne0 is the number
of moles of Gaussian chains in the network per unit volume of rubber and R is the gas
constant. The statistical mechanical theories thus predict that the stiffness of a rubber

87
increases with the crosslink density and the temperature. The theoretical equation
captures the trend in experimental stress-strain data of unfilled rubbers in compression
(λ<1) and at low extension ratios (λ<1.2). At higher extension ratios the experimental
data fall below the theoretical curve. The Mooney equation which was derived from
continuum mechanics is however capable of describing the nominal stress-strain data
at λ between 1 and 2:

% 1 (% C (
" N = 2' # $ 2 *'C1 + 2 * (3.62)
& # )& #)

! where C1 and C2 are empirical constants. It may be noted that the Mooney equation is
identical with the equations derived from statistical mechanics when C2 = 0.
Experiments have shown that a range of different rubbers that C1 depends on the
crosslink density basically as the modulus does according to the statistical mechanical
theory, whereas C2 is approximately constant. Stress-strain data of rubbers swollen in
organic solvents analysed by the Mooney equation show interesting results: C1 is
practically independent of the degree of swelling, whereas C2 decreases with
increasing degree of swelling approaching C2 = 0 at v2 ≈ 0.2. Hence, highly swollen
rubbers (v2 ≤0.2) behave according to the statistical mechanical theories. Stress-strain
data described by the Mooney equation suggests that the networks at low extension
(λ=1.0-1.2) deform near affinely whereas at higher extensions (λ=1.5-2) the
fluctuations of the junctions increases (phantom network behaviour).

None of Gaussian statistical mechanical theories is adequate to describe the stress-


strain behaviour at large strains (λ>3 to 4). The pronounced upturn in the stress-strain
curve can be accounted for by the finite extensibility of the network not accounted for
by the Gaussian distribution that describes the statistics of the phantom chains. In
addition, some rubbers crystallise during extension and a smaller part of the increase
in stiffness at high extension ratios is due to the presence of crystallites in the
material.

Loose chain ends, temporary and permanent chain entanglements, intramolecular


crosslinks are complications not directly addressed by the classical statistical
mechanics theory. Chain ends and intramolecular crosslinks do not contribute to the
elastic force, whereas chain entanglements add more junction points to the covalent
network.

88
3.8. Exercises

3.1. The rubber in a blown-up balloon is biaxially stretched. Derive the force-strain
relationship under the assumption that the rubber follows the Gaussian statistical
theory of rubber elasticity.

3.2. Derive the relationship between the internal pressure (p) and the degree of
expansion (α) of the balloon. Assume the validity of the ideal gas law (pV = nRT).

3.3. At what α value has the internal pressure a maximum?

3.4. Suppose the balloon has a small nose. Is it possible to get the nose to expand to
the same degree as the rest of the balloon?

3.5. Many rubber materials exhibit time-dependent mechanical properties (Fig. 3.20).
Make a list of possible reasons.

3.6. Polyethylene can be crosslinked by decomposition of organic peroxides,


hydrolysis of vinyl-silane grafted polyethylene or by high-energy (β or γ) irradiation.
Design a suitable experiment to determine the crosslink density and present the
relevant equations.

Figure 3.20. Results from measurements of continuous stress relaxation of nitrile


rubber (low network density). Drawn after data from Björk (1988).

3.7. Calculate by using the affine network model the modulus at room temperature of
natural rubber (ρ ≈ 970 kg m-3) crosslinked with n molar fraction of organic peroxide.
Assume that each peroxide molecule results in one crosslink.

3.8. Calculate the temperature increase occurring in natural rubber with M c = 5 000 g
mol-1 when it is stretched to λ=5 at room temperature. Use the following data: ρ =
970 kg m-3, cp= 1900 J kg-1 K-1.

89
3.9. References

Anthony, P. C., Caston, R. H. and Guth, E. (1942) J. Phys. Chem., 46, 826.

Björk, F. (1988) Ph.D. Thesis: Dynamic stress relaxation of rubber materials,


Department of Polymer Technology, Royal Institute of Technology, Stockholm,
Sweden.

Gumbrell, S. M., Mullins, L. and Rivlin, R. S. (1953) Trans. Faraday Soc., 49, 1495.

Higgs, P. G. and Gaylord, R. J. (1990) Polymer, 31, 70.

Mark, J. E. (1984) The rubber elastic state, in Physical Properties of Polymers, Mark,
J. E. (ed.), American Chemical Society, Washington, DC.

Mark, J. E. (1993) The rubber elastic state, in Physical Properties of Polymers, 2nd ed.,
Mark, J. E. (ed.), American Chemical Society, Washington, DC.

Nairn, J. A. (1998) Lattice™ 7.0, Random walk simulations of polymer molecules on


a tetrahedral lattice, University of Utah, Salt lake City, USA.

Treloar, L. R. G. (1975) The Physics of Rubber Elasticity, 3rd ed., Clarendon, Oxford.

Wood, L. A. and Roth, F. L. (1944) J. Appl. Phys., 16, 781.

90

Anda mungkin juga menyukai