Download as pdf or txt
Download as pdf or txt
You are on page 1of 56

L. D.

COLLEGE OF ENGINEERING
AHMEDABAD – 380 015

Name: Er. Maulik Maheshbhai Chauhan

Semester: M.E.-II Enrollment No.: 230280788001

Subject: Physics of Rubber Elasticity (3724008) Year: 2023-24


Certificate

This is to certify that Er. Maulik Maheshbhai Chauhan

Enrollment No. 230280788001 of M.E. Semester-II of Rubber

Technology Department has satisfactorily completed the course

in the subject of Physics of Rubber Elasticity (3724008)

within four walls of L. D. College of Engineering, Ahmedabad -

380015.

Date of Submission : ______________________________________

Staff in-Charge : ______________________________________

Head of Department : ______________________________________

i
Certificate

This is to certify that the above term work of Er. Maulik Maheshbhai

Chauhan University Exam No. 230280788001 of M.E. Semester-II of

Rubber Technology Department is assessed for University

examination on ___________________.

Int. Examiner ______________________________________

Ext. Examiner ______________________________________

ii
3724008 Index

Index
Sr. Page Number Date of Date of Initial
Title From To
No Start Completion of Staff

To carry out the kinetic study of


1 low temperature crystallization 01 21-02-2024 28-02-2024
observed in Rubber.

To study concept of the Rubber


Elasticity by using basic
2 28-02-2024 06-03-2024
fundamentals of
Thermodynamics.

To study about the Kinetic


3 06-03-2024 13-03-2024
theory of Elasticity.

To study about the cohesive


4 energy density related to 13-03-2024 19-03-2024
rubber/polymeric materials.

To study about the elastic


5 properties observed in swollen 27-03-2024 03-04-2024
rubber.

To determine the Crosslinking


density of given vulcanizate
6 03-04-2024 09-04-2024
sample by the swelling method
(Chemical Method).

7 Mulin’s Effect 24-04-2024 01-05-2024

To study about the Statistical


8 properties of Rubber long chain 01-05-2024 15-05-2024
molecule.

To study about the principal


stresses observed in unstrained
9 15-05-2024 22-05-2024
and strained state on Rubber
rectangular block.

10 22-05-2024 29-05-2024

iii
3724008 Practical - 01

Practical - 01

AIM: To carry out the kinetic study of low temperature crystallization observed in
Rubber.

1.1 Introduction
Rubber materials, both natural and synthetic, are essential components in a wide range of
industrial applications due to their unique properties and versatility. Understanding the
crystallization behavior of rubber polymers is crucial for optimizing their performance and
tailoring their properties to meet specific application requirements. Natural rubber (NR),
derived from the latex of the Hevea brasiliensis tree, and synthetic rubber (SR), synthesized
from petroleum-derived monomers, exhibit distinct crystallization characteristics influenced
by their chemical composition, structure, and processing conditions.
Natural rubber consists predominantly of cis-1,4-polyisoprene units, characterized by the
presence of unsaturated double bonds along the polymer backbone. This unique molecular
structure imparts a high degree of chain flexibility and mobility, facilitating crystallization
under appropriate conditions. Crystallization of natural rubber typically occurs upon cooling
from a melt or upon solvent evaporation. At lower temperatures, the molecular mobility of
polymer chains decreases, leading to the alignment and packing of chains into ordered
crystalline domains within the rubber matrix. The crystallization process in natural rubber is
influenced by factors such as temperature, mechanical deformation, and the presence of
nucleating agents or additives.
Synthetic rubbers encompass a diverse range of polymer materials synthesized from
monomers such as styrene, butadiene, isoprene, and ethylene. The crystallization behavior of
synthetic rubber varies significantly depending on the specific type of rubber and its chemical
structure. Some synthetic rubbers, such as polybutadiene (BR), exhibit a higher tendency to
crystallize due to their linear polymer chains and regular packing arrangements. In contrast,
other synthetic rubbers may remain predominantly amorphous or exhibit limited
crystallinity. The crystallization of synthetic rubber can be influenced by polymer
architecture, molecular weight, additives, and processing conditions.
Comparing the crystallization behavior of natural rubber and synthetic rubber reveals
important differences and similarities. Natural rubber, with its cis-1,4-polyisoprene structure,

1
3724008 Practical - 01
tends to crystallize more readily under ambient conditions compared to most synthetic
rubbers. The presence of double bonds and the linear chain structure contribute to the
formation of ordered crystalline domains upon cooling. In contrast, the crystallization
behavior of synthetic rubber is more diverse and depends on the specific chemical
composition and structure of the polymer.
Understanding the crystallization mechanisms in both natural and synthetic rubber is
essential for optimizing rubber formulations, processing techniques, and product
performance. By elucidating the factors influencing rubber crystallization, researchers can
develop advanced materials with tailored properties suitable for a wide range of industrial
applications, including automotive, aerospace, biomedical, and consumer goods sectors.

2
3724008 Practical - 01

1.2 Mechanism behind the Crystallization


1) Molecular Mobility: At higher temperatures, rubber chains exhibit significant mobility
due to thermal energy. As the temperature decreases, molecular mobility decreases,
leading to a transition from a rubbery state to a glassy state. In this glassy state, the
polymer chains are relatively immobile, but some residual mobility persists, enabling the
initiation of crystallization.

2) Nucleation: Crystallization typically begins with the formation of nucleation sites where
polymer chains begin to align and organize into ordered structures. Nucleation can occur
through various mechanisms, including:

a) Homogeneous Nucleation: Spontaneous organization of polymer chains into


crystalline structures throughout the bulk material.

b) Heterogeneous Nucleation: Initiation of crystallization at imperfections or foreign


particles within the rubber matrix, such as dust particles or polymer chain ends.

3) Chain Alignment: Once nucleation sites are formed, polymer chains begin to align along
specific crystallographic directions. This alignment is facilitated by intermolecular forces
such as Van der Waals interactions and hydrogen bonding between polymer chains. As
more chains align, the crystalline regions grow, gradually extending throughout the rubber
matrix.

4) Growth Kinetics: The growth of crystalline regions proceeds as polymer chains continue
to align and pack together more closely. The growth of these crystals proceeds through the
incorporation of additional polymer chains into the existing nuclei. As crystalline regions
expand, they become more prominent within the rubber matrix.

Factors influencing crystal growth include:

a) Diffusion of polymer chains: Chains must diffuse and align in the proper orientation
to join the growing crystal lattice.

3
3724008 Practical - 01
b) Rate of crystallization: The rate at which new polymer chains join the crystal lattice
determines the overall growth kinetics.

c) Temperature: Lower temperatures typically slow down the rate of crystallization,


leading to smaller and more numerous crystalline domains.

5) Morphology of Crystalline Regions: The morphology of crystalline regions can vary


depending on factors such as polymer structure and processing conditions. In natural
rubber, crystalline regions may exhibit a lamellar or spherulitic structure, characterized
by layers or spherical clusters of aligned polymer chains. The size and distribution of
crystalline regions influence the mechanical properties of the rubber material.

Fig 1.1 Stress-strain curve and corresponding diffraction patterns


recorded simultaneously for a vulcanized NR (1.5 phr sulfur).

6) Effects of Additives: The crystallization of rubber can be influenced by the presence of


additives such as nucleating agents, plasticizers, and fillers. Nucleating agents promote the
formation of nuclei, accelerating the crystallization process. Plasticizers and fillers may

4
3724008 Practical - 01
affect chain mobility and hinder or promote crystallization, depending on their chemical
nature and concentration.
7) Effects of Crystallization: The low-temperature crystallization of rubber can significantly
affect its mechanical and thermal properties. Crystalline regions contribute to increased
stiffness, strength, and dimensional stability, particularly at low temperatures. However,
excessive crystallization can lead to brittleness and reduced flexibility, limiting the
applicability of the rubber material in certain environments.

8) Characterization Techniques: Various techniques, including differential scanning


calorimetry (DSC), X-ray diffraction (XRD), and microscopy, can be used to study the
crystallization behavior of rubber materials. These techniques provide insights into the
crystalline structure, morphology, and kinetics of crystallization.

5
3724008 Practical - 01

1.3 Factors affecting the Crystallization

The crystallization of rubber is a complex process influenced by various factors that affect nucleation,
growth, and morphology of crystalline domains within the rubber matrix. Below are the factors
affecting the crystallization of rubber in detail:

1) Polymer Structure and Composition:

• Rubber polymers possess varying chemical compositions and structures, influencing


their crystallization behavior.

• Presence of double bonds: Rubber polymers with unsaturated double bonds, such as
natural rubber (NR), exhibit higher propensity for crystallization compared to fully
saturated synthetic rubbers.

• Branching and side groups: Polymer architecture, including the presence of side chains
or branches, affects chain mobility and packing, influencing the nucleation and growth
of crystalline domains.

2) Temperature:

• Temperature plays a critical role in rubber crystallization. Lower temperatures


generally promote crystallization by reducing molecular mobility and promoting chain
alignment and ordering.

• However, excessively low temperatures can hinder crystallization by slowing down


molecular motion and nucleation kinetics.

3) Rate of Cooling:

• The rate at which the rubber is cooled affects the nucleation and growth kinetics of
crystallization.

• Rapid cooling may lead to the formation of smaller crystalline domains due to
insufficient time for polymer chains to organize into larger structures, whereas slower
cooling allows for more extensive crystal growth.

6
3724008 Practical - 01

Fig 1.2

4) Presence of Nucleating Agents:

• Nucleating agents are additives that promote the formation of nucleation sites,
accelerating the crystallization process.

• Nucleating agents can include organic or inorganic compounds that provide surfaces
for nucleation or induce structural changes in the polymer matrix to facilitate
nucleation.

5) Chain Mobility:

• Molecular mobility of polymer chains is crucial for the nucleation and growth of
crystalline domains.

• Higher chain mobility at elevated temperatures allows for greater diffusion and
reorganization of polymer chains, facilitating crystallization.

6) Presence of Additives:

• The presence of additives such as plasticizers, fillers, and processing aids can
significantly impact rubber crystallization.

• Plasticizers can increase chain mobility, hindering crystallization, while fillers may
promote nucleation by providing nucleation sites or affecting polymer chain
interactions.

7
3724008 Practical - 01
• Processing aids can modify polymer morphology and crystallization kinetics during
manufacturing processes.

7) Mechanical Deformation:

• Mechanical deformation, such as stretching or compression, can induce crystallization


in rubber materials by promoting alignment and reorientation of polymer chains.

• Deformation leads to increased chain mobility and alignment along the direction of
applied stress, facilitating the formation of crystalline domains.

8) Polymer Molecular Weight:

• Polymer molecular weight influences crystallization behavior, with higher molecular


weight polymers typically exhibiting slower crystallization rates due to increased
chain entanglement and reduced mobility.

9) Thermal History:

• The thermal history of rubber, including previous processing steps and storage
conditions, can affect its crystallization behavior.

• Rubber materials subjected to annealing or thermal cycling may exhibit altered


crystallization kinetics and morphology due to changes in chain conformation and
structure.

8
3724008 Practical - 01

1.4 Comparison of Crystallization behaviour in Natural and Synthetic Rubber

Feature Natural Rubber Synthetic Rubber

Crystallization Strain-induced (primary) and temperature-


Primarily cooling-induced (if any)
Type induced (secondary)

Crystallization Mechanical stretching or cooling below a Primarily cooling below the glass
Trigger critical temperature (Tg) transition temperature (Tg)

Varies depending on the specific


Crystalline Orthorhombic unit cell with highly ordered polymer. Can be amorphous or
Structure chains in a cis-1,4 configuration have less ordered structures like
chain folds or fringed micelles.

Varies depending on type and


Smaller crystallites during strain-induced,
Crystal Size processing. Generally smaller than
larger during temperature-induced
natural rubber.

Low to none (amorphous) or very


Crystallinity High (up to 80%)
limited crystallinity

Lower elasticity: Amorphous


Increased elasticity: Crystal regions act like
structure allows for easier chain
physical crosslinks, storing and returning
movement, leading to lower stored
energy during deformation.
energy and return.
Effect on Increased tensile strength: Crystallites resist
Lower tensile strength: Absence of
Properties chain pullout, making the material stronger.
crystallites reduces resistance to
Increased fatigue resistance: Ordered chain pullout.
structure reduces internal friction and
Lower fatigue resistance:
improves resistance to repeated stress.
Disordered structure can lead to

9
3724008 Practical - 01

Hysteresis: Due to energy dissipation during increased internal friction and

deformation/recovery cycle, a loop is fatigue under repeated stress.

observed in stress-strain curves.


Higher heat generation: Internal

Higher melting point: Ordered structure friction during deformation can

requires more energy to disrupt, leading to a lead to increased heat generation.


higher melting point.

10
3724008 Practical - 01

1.5 Thermodynamic Principles of the Crystallization Process

The crystallization process may be described thermodynamically. One may write

∆𝐺 = 𝐺𝑐𝑟𝑦𝑠𝑡 − 𝐺𝑎𝑚𝑜𝑟𝑝ℎ

∴ ∆𝐺 = ∆𝐻 − 𝑇∆𝑆 (1.1)

where 𝐺𝑐𝑟𝑦𝑠𝑡 and 𝐺𝑎𝑚𝑜𝑟𝑝ℎ are the Gibbs free energies of the crystalline and amorphous phases
respectively, ∆𝐻 is the enthalpy of crystallization, and ∆𝑆 is the entropy of crystallization.
Crystallization becomes theoretically possible whenever the ‘melt’ free energy exceeds the ‘crystal’
free energy-enabling a reduction in free energy to result from the crystallization process.
Theoretically this may occur whenever 𝐺𝑚𝑒𝑙𝑡 > 𝐺𝑐𝑟𝑦𝑠𝑡 and will proceed until ∆𝐺 is at a minimum. In
practice crystallization in polymer melts and solutions normally occurs only at lower temperatures
than eqn (1.1) would suggest (i.e., under conditions of some ‘supercooling’) and not uniformly over
the polymer volume but by growth from small crystal nuclei with large specific areas. It follows from
this that ∆𝐺 in eqn (1.1) may be written as

∆𝐺 = ∆𝐺𝐶 − ∑ 𝐴𝛾 (1.2)

where 𝐺𝐶 is the bulk free enthalpy change disregarding surface effects, 𝛾 is the specific surface free
energy, and 𝐴 is the corresponding surface area.

11
3724008 Practical - 01

1.6 Nucleation

There is a characteristic temperature called the equilibrium melting temperature, 𝑇𝑚 , at which all
crystallites will melt when heated very slowly (i.e., under equilibrium conditions). At temperatures
well below the equilibrium melting temperature for zero strain (+30°C for NR) primary crystal
nucleation is likely to be spontaneous and homogeneous.

Embryo nuclei will however have to grow to a critical size and distribution before any macroscopic
effects (such as elastic modulus increase) are observed. There will therefore be a time lag called the
‘induction period’, 𝜏, before the observation of any change in properties. Frisch has derived a general
expression for 𝜏 in terms of the statistical distribution of embryo nuclei, and in general, the induction
period may be written as

𝜏 = ∫ ∆𝑓(𝑖)/I𝑑𝑖 (1.3)

where 𝐼 is the nucleation rate (in nuclei per second), 𝑖 is a parameter characterizing nucleus size, and
∆𝑓(𝑖) is the difference between the steady state distribution of nuclei of size 𝑖 (after long times) and
the initial distribution. As a first approximation, the nucleation rate may be assumed constant
throughout the induction period (when nuclei are of sub-critical size). This effectively assumes time
and size independent crystal surface free energies. Eqn (1.3) may be then simplified to

𝜏 = 𝑁𝐶 /𝐼 (1.4)

where 𝑁𝐶 here represents the increase per mole in the number of nuclei of critical size required to
produce an observable effect (a modulus increase in the present case).

The classical approach to crystal nucleation was developed primarily by Gibbs and rests on the
assumption that fluctuations in the supercooled phase can overcome the energy barrier provided by
the crystal surface. The probability of finding a nucleus of given size at constant volume and energy
is, according to Boltzmann’s law, a function of the entropy change, and may be given by

∆𝑆
𝑃 ∝ 𝑒𝑥𝑝 ( ) (1.5)
𝑘

At constant pressure and temperature this becomes

12
3724008 Practical - 01

−∆𝐺
𝑃 ∝ 𝑒𝑥𝑝 ( ) (1.6)
𝑘𝑇

On this basis, Turnbill and Fisher have used absolute reaction rate theory to derive the following
expression for the rate of crystal nucleation in nuclei per mole per second:

𝑁𝑘𝑇
𝐼= 𝑒𝑥𝑝 (−(∆𝐺1 + ∆𝐺2 )/𝑘𝑇) (1.7)

The term ∆𝐺1 , represents the free enthalpy of crystallization of a nucleus of critical size.
Differentiation of eqn (1.2) with respect to size for the equilibrium morphology has been shown to
lead to the following expression

32𝛾 2 𝛾𝑒 𝑇𝑚2
∆𝐺1 = (1.8)
(ℎ∆𝑇𝜌𝑐 )2

where 𝑇 is the crystallization temperature, 𝑇𝑚 is the equilibrium melting temperature for zero strain,
h is the latent heat of fusion (16.4 Jg¯¹ for NR), ∆𝑇 is the super-cooling. 𝑇𝑚 − 𝑇, (the greater the super
cooling the smaller the critical nucleus size). and 𝜌𝑐 , is the density of the crystalline phase. 𝛾 and 𝛾𝑒
are the surface free energies of the sides and ends of crystals respectively. The greater 𝛾𝑒 is relative
to 𝛾, the more fibrous the nucleus structure.

The second term in eqn (1.7), ∆𝐺2 , describes essentially the mobility of chain segments and has a
temperature dependence similar to that of viscosity. ∆𝐺2 increases rapidly near to the glass transition
temperature, but at higher temperatures near to 𝑇𝑚 eqn (1.7) becomes dominated by ∆𝐺1 . A possible
(and simple) mathematical form for ∆𝐺2 is

𝑐
∆𝐺2 = 𝑘𝑇 ( ) (1.9)
𝑑 + 𝑇 − 𝑇𝑔

where 𝑐 and 𝑑 are material constants, and 𝑇𝑔 , is the temperature below which chain segments have
insufficient mobility to join the crystal structure. For present purposes this latter temperature is
considered equal to the glass transition temperature. Eqn (1.9) is similar in form to the familiar WLF
expression, although the latter cannot be expected to apply to a semi-crystalline polymer without
some modification.

Substituting eqns (1.8) and (1.9) in eqn (1.7) yields

13
3724008 Practical - 01

𝑁𝑘𝑇 𝑏 ∗ 𝑇𝑚 2 𝑐
𝐼 = 𝑒𝑥𝑝 (− ( ) + ) (1.10)
ℎ 𝑘𝑇 ∆𝑇 𝑑 + 𝑇 − 𝑇𝑔

where 𝑏 ∗ is a material constant, and the pre-exponential factor has been ascribed a value of
6.2 × 1012 𝑠 −1 . This leads to the following expression for the induction period:

𝑏𝑇𝑚 2 𝑐
𝜏 = 𝑎 𝑒𝑥𝑝 (( ) + ) (1.11)
𝑇∆𝑇 2 𝑑 + 𝑇 − 𝑇𝑔

The constant a has absorbed the pre-exponential factor and depends on the distribution of nuclei
required to cause the minimum observable modulus increase. Boltzmann's constant has been
absorbed in the constant 𝑏. The constants 𝑐 and 𝑑 describe the restriction in chain mobility as the
crystallization temperature approaches the glass transition.

Eqn (1.11) predicts an infinite induction period at the equilibrium melting temperature, 𝑇𝑚 , and at
𝑑°C below 𝑇𝑔 .

14
3724008 Practical - 01

1.7 Growth rate

The probability at time 𝑡 of a given point in the rubber matrix never being incorporated in a
'crystallite, when the latter grow at a local growth rate, 𝑅, may be expressed by a Poisson probability
distribution

𝑃 = 𝑒𝑥𝑝 (−𝑅) = 𝐼 − 𝑉𝑐 (1.12)

where 𝑉𝑐 is here the volume fraction of crystalline material. If the local growth rate 𝑅 (i.e., that of a
single crystal free to grow without restriction) may be expressed by a power law of the form

𝑅 = 𝐾𝑡 𝑛 (1.13)

where 𝑛 and 𝐾 are constants, then the volume fraction of crystalline material at time 𝑡 is given by

𝑉𝑐 (𝑡) = 𝐼 − 𝑒𝑥𝑝 (−𝐾𝑡 𝑛 ) (1.14)

Relations of this form were proposed by Avrami to describe crystallization in metals. They have been
successfully used to describe crystal growth in polymers by Gent, by Mandelkern (NR and
polyethylene), by Wunderlich (ethylene terephthalate/ethylene scbecate copolymer), and by other
authors.

For polymers, the equilibrium volume fraction of crystalline material, 𝑉𝑐 may be considerably less
than unity, leading eqn (1.14) to become

𝑉𝑐 (𝑡) = 𝑉𝑐 (∞)[𝐼 − 𝑒𝑥𝑝 (−𝐾𝑡 𝑛 )] (1.15)

This may be written in the form

𝑉𝑐 (∞) − 𝑉𝑐 (𝑡)
−𝑙𝑛 ( ) = 𝐾𝑡 𝑛 (1.16)
𝑉𝑐 (∞)

If the increase in elastic modulus is a simple linear function of 𝑉𝑐 at constant strain and temperature,
then the elastic modulus at time 𝑡 may be written as

𝐸(𝑡) = 𝐸(𝑜) + 𝐴𝑉𝑐 (𝑡) (1.17)

15
3724008 Practical - 01
where 𝐴 is a constant of proportionality which may depend on both the crystallization temperature
and the applied test strain. Substitution in eqn (1.16) yields

𝐸(∞) − 𝐸(𝑡)
−𝑙𝑛 ( ) = 𝐾𝑡 𝑛 (1.18)
𝐸(∞) − 𝐸(𝑜)

and

𝐸(∞) − 𝐸(𝑡)
𝑙𝑜𝑔 [−𝑙𝑛 ( )] = log 𝐾 + 𝑛 log 𝑡 (1.19)
𝐸(∞) − 𝐸(𝑜)

A plot of 𝑡 vs. −𝐼𝑛[𝐸(∞) − 𝐸(𝑡)/𝐸(∞) − 𝐸(𝑜)] on logarithmic scales should therefore produce a
straight line of slope 𝑛.

The Avrami constant 𝐾 may be expressed in terms of the induction period

𝐸(∞) − 𝐸(𝜏)
[ ]
𝐸(∞) − 𝐸(𝑜) (1.20)
𝐾 = −𝑙𝑛
𝜏𝑛

16
3724008 Practical - 01

1.8 Dilatometry: Volume Expansion Analytical Method for Rubber

Dilatometry is an analytical method used to measure the volume changes of a material as a


function of temperature.
It is particularly useful for studying the volumetric expansion and contraction of materials,
including rubber, during phase transitions such as crystallization. Here, we focus on using
dilatometry to study the volume expansion of rubber during low-temperature crystallization.

1.8.1 Principle:

Dilatometry measures the change in volume of a sample as it undergoes temperature


variations. This is particularly relevant for studying the crystallization of rubber, where
volume changes due to the formation of crystalline regions can be detected.

Fig 1.3 Dilatometry

1.8.2 Apparatus:

Dilatometer: A device equipped with a temperature-controlled chamber and a sensitive


volume measurement system, typically based on a linear variable differential transformer
(LVDT) or similar sensor.

17
3724008 Practical - 01
Temperature Control System: A cooling unit capable of reaching and maintaining low
temperatures necessary for crystallization.
Rubber Sample: Prepared in a standard shape and size for consistent measurements.

Fig 1.4 Experimental setup of the dilatometer sample chamber

1.8.3 Procedure:

a) Sample Preparation:

Prepare rubber samples of uniform size and shape.


Ensure samples are free from defects and impurities that could affect measurements.

b) Experimental Setup:

Place the rubber sample in the dilatometer.


Connect the dilatometer to the temperature control system.

c) Cooling Phase:

Cool the sample to the desired low temperature at a controlled rate. This can be done using a
liquid nitrogen bath or a cryogenic chamber.
Monitor the volume changes continuously as the temperature decreases.

18
3724008 Practical - 01
d) Isothermal Crystallization:

Maintain the sample at a constant low temperature to allow crystallization to occur.


Record the volume change over time as crystallization progresses.

e) Heating Phase:

After crystallization is complete, gradually heat the sample back to room temperature.
Monitor the volume changes during heating to study the melting behavior of the crystalline
regions.

Fig 1.5 Optical dilatometry – monochrome image taken during a heating process

1.8.4 Data Analysis:

a) Volume-Temperature Relationship:

Plot the volume change against temperature to identify the onset of crystallization (volume
decrease) and melting (volume increase).

19
3724008 Practical - 01
b) Kinetics of Crystallization:

Analyze the volume change over time at a constant low temperature to determine the
crystallization rate.
Use models such as Avrami’s equation to fit the crystallization data and extract kinetic
parameters.

c) Degree of Crystallinity:

Calculate the degree of crystallinity by comparing the volume change during crystallization to
the total possible volume change.

d) Thermal Expansion Coefficient:

Determine the thermal expansion coefficient of the rubber in both the amorphous and
crystalline states.

1.8.5 How does optical dilatometry work in a measuring device?

In thermo-optical analysis with a dilatometer equipped with a camera, the sample is


irradiated with light from one side and the shadows are recorded on the other side.
Based on the data obtained, a computer calculates the change in length of the sample.

1.8.6 Advantages of Dilatometry:

Direct measurement of volume changes provides clear insights into the phase transition.
High sensitivity to small volume changes makes it suitable for detecting subtle crystallization
events.
Non-destructive method allowing further analysis of the same sample using other techniques.

1.8.7 Considerations and Limitations:

Ensure proper calibration of the dilatometer for accurate volume measurements.


Sample preparation and handling must be consistent to avoid measurement errors.
Dilatometry primarily measures bulk volume changes; complementary techniques (e.g., DSC,
XRD) may be needed to provide molecular-level insights.

20
3724008 Practical - 07

Practical - 02

AIM: To study concept of the Rubber Elasticity by using basic fundamentals of


Thermodynamics.

2.1 About Rubber Elasticity

Early attempts to account for the mechanical properties of rubber in terms of classical
concepts of the molecular structure of matter encountered overwhelming difficulties. In the
first-place natural rubber (the only type of rubber then available) appeared both unique and
enigmatic.
Yet its molecular constitution as then understood could hardly have been simpler.
Other known hydrocarbons of apparently entirely similar chemical constitution were
invariably ordinary liquids or solids.
The classical solid was envisaged as an array of atoms (or molecules) maintained in fixed
mean relative positions by well-defined interatomic forces.
Application of a stress (e.g. a tensile force) to such a structure leads to a disturbance of the
equilibrium which calls into play a counterbalancing internal stress.
Owing to the very strong dependence of such interatomic forces on the distance between
atoms, such a mechanism is incapable, even theoretically, of supporting extensions of
magnitude greater than about 10 per cent; it cannot conceivably serve as a basis for the
interpretation of deformations of 100 times this magnitude.
The basic difficulty can in principle be circumvented, without entirely breaking away from
classical concepts, by postulating either some sort of open network structure, or alternatively
a helical or coil-spring type of molecule. In either system large total deformations may be
obtained without the introduction of large strains into the elastic elements of the structure.
This attributed the essential elasticity to a sort of network of micelles or molecular aggregates
based on proteins or resins derived from the outer sheath of the latex globule, this network
being suspended in a semi-liquid medium formed of rubber hydrocarbon of lower molecular
weight.
Among theories of the second type, the helicalspring theory of Fikentscher and Mark (1930)
was the most notable.

21
3724008 Practical - 07
In this, the retraction tendency was associated with the residual forces between neighbouring
turns of the helix which was believed to represent the configuration of the polyisoprene chain.
A rather similar theory by Mack (1934) envisaged a folded configuration of the molecule
which was maintained by the agency of forces between neighbouring hydrogen atoms.
This model permitted an extension of 300 per cent in passing from the closely packed to the
fully extended form by rotation about single bonds in the chain structure.
A further extension of the assembly of molecules, leading to a total extensibility of 600 per
cent, was accommodated by allowing for a rotation of the extended chain from their initially
random directions into the direction of the applied strain.

22
3724008 Practical - 02

2.2 Thermodynamic Analysis

For the more explicit examination of these thermoelastic phenomena it is necessary to


develop relations between force, length, and temperature on the one hand and the
thermodynamic quantities, internal energy and entropy, on the other.
The required relations are obtainable directly from the first and second laws of
thermodynamics.
From the first law of thermodynamics the change in internal energy 𝑑𝑈 in any process is given
by

𝑑𝑈 = 𝑑𝑄 + 𝑑𝑊 (2.1)

where 𝑑𝑄 and 𝑑𝑊 are respectively the heat absorbed by the system and the work done on it
by the external forces.
The second law defines the entropy change dS in a reversible process by the relation

𝑇𝑑𝑆 = 𝑑𝑄 (2.2)

and hence from (2.1), we have for a reversible process

𝑑𝑈 = 𝑇𝑑𝑆 + 𝑑𝑊 (2.3)

In discussing the equilibrium of a system which is subject to reversible changes (e.g. elastic
deformations) it is convenient to introduce the Helmholtz free energy A, defined by the relation

𝐴 = 𝑈 − 𝑇𝑆 (2.4)

For a change taking place at constant temperature, we have then

𝑑𝐴 = 𝑑𝑈 − 𝑇𝑑𝑆 (2.5)

Combining this equation with (2.3) we obtain the standard thermodynamic result

𝑑𝐴 = 𝑑𝑊 (constant T) (2.6)

23
3724008 Practical - 02
which signifies that in a reversible isothermal process the change in Helmholtz free energy is
equal to the work done on the system by the applied forces.
In most thermodynamic subsequently developed with particular reference to gases and
liquids, for which the significant variables include pressure 𝑝 and volume 𝑉. The work done
on the system in a small displacement is then written as

𝑑𝑊 = −𝑝 𝑑𝑉

In discussing problems related to the elasticity of solids, on the other hand, we are concerned
primarily with the work done by the applied stress, corresponding, for example, to a tensile
force 𝑓 acting on a specimen of length 𝑙, in which case the work done in a small displacement
is

𝑑𝑊 = 𝑓 𝑑 𝑙 (2.7)

When in addition a hydrostatic pressure (e.g. atmospheric pressure) is also present the total
work done by the applied forces becomes

𝑑𝑊 = 𝑓 𝑑 𝑙 − 𝑝 𝑑 𝑣 (2.7a)

For strict accuracy it is necessary to take account of both terms on the right-hand side of
(2.7a).
But in the case of rubbers, the volume change 𝑑𝑉 is usually very small, and if 𝑝 is the
atmospheric pressure the term 𝑝 𝑑 𝑉 is less than 𝑓 𝑑𝑙 by a factor of 10-3 to 10-4.
As a first approximation we may therefore neglect this term and use eqn (2.7), which is strictly
accurate only at zero applied pressure or under constant-volume conditions, in place of the
exact expression (2.7a) (the more accurate analysis, not involving this assumption).
By making use of eqns (2.6) and (2.7) the tension may then be expressed in the form

𝜕𝑊 𝜕𝐴
𝑓= ( ) = ( ) (2.8)
𝜕𝑙 𝑇 𝜕𝑙 𝑇

which shows that the tensile force is equal to the change in Helmholtz free energy per unit
increase in length of the specimen.

24
3724008 Practical - 02
The significance of the important relation (2.8) may be better appreciated by reference to Fig.
2.1, which represents diagrammatically the variation of Helmholtz free energy for an elastic
body as a function of its length 𝑙.

Fig 2.1 Dependence of Helmholtz free energy 𝐴 and force 𝑓 on length 𝑙 for
specimen subjected to uniaxial extension or compression. 𝑙0 = unstrained length.

The unstressed state is such that the Helmholtz free energy is a minimum, so that (𝜕𝐴/𝜕𝑙) 𝑇 =
0 when 𝑙 = 𝑙0 ,. If 𝑙 is greater than 𝑙0 , (𝜕𝐴/𝜕𝑙) 𝑇 , is positive, corresponding to a tensile force,
while if 𝑙 is less than 𝑙0 , (𝜕𝐴/𝜕𝑙) 𝑇 , is negative, corresponding to a compressive force. (It should
be noted, however, that the force-deformation relation will not in general be linear, except for
very small strains.)
The tension, like the free energy, may be expressed as the sum of two terms (from eqn (2.5)),
thus

𝜕𝐴 𝜕𝑈 𝜕𝑆
𝑓 = ( ) = ( ) −𝑇( ) (2.9)
𝜕𝑙 𝑇 𝜕𝑙 𝑇 𝜕𝑙 𝑇

of which the first represents the change in internal energy and the second the change in
entropy, per unit increase in length.
The second term is related to the temperature coefficient of tension, as will now be shown.
Writing eqn (2.4) in differential form, we have for any change (i.e. not necessarily isothermal)

25
3724008 Practical - 02

𝑑𝐴 = 𝑑𝑈 − 𝑇𝑑𝑆 − 𝑆𝑑𝑇

whilst from (2.3) and (2.7)

𝑑𝑈 = 𝑓 𝑑𝑙 + 𝑇𝑑𝑆

Combination of these two equations gives

𝑑𝐴 = 𝑓 𝑑𝑙 − 𝑆𝑑𝑇

Hence, by partial differentiation,

𝜕𝐴 𝜕𝐴
( ) = 𝑓 ; ( ) = −𝑆 (2.10)
𝜕𝑙 𝑇 𝜕𝑇 𝑙

By a well-known property of partial differentials

𝜕 𝜕𝐴 𝜕 𝜕𝐴
( ) = ( )
𝜕𝑙 𝜕𝑇 𝑙 𝜕𝑇 𝜕𝑙 𝑇

and hence, from eqns (2.10),

𝜕𝑆 𝜕𝑓
( ) = −( ) (2.11)
𝜕𝑙 𝑇 𝜕𝑇 𝑙

Eqn (2.11) gives the entropy change per unit extension in terms of the measurable quantity
(𝜕𝑓/𝜕𝑇)𝑙 , the temperature coefficient of tension at constant length. Insertion of this relation
in (2.9) gives for the corresponding internal energy change

𝜕𝑈 𝜕𝑓
( ) = 𝑓 −𝑇( ) (2.12)
𝜕𝑙 𝑇 𝜕𝑇 𝑙

The relations (2.11) and (2.12) are of fundamental importance in rubber elasticity, since they
provide a direct means of determining experimentally both the internal energy and entropy
changes accompanying a deformation.

26
3724008 Practical - 02
Their application may be illustrated by reference to Fig. 2.2, in which the curve CC’ represents

Fig 2.2 Slope and intercept of force-temperature curve.

the variation with temperature of the force at constant length, which may or may not be
linear.
The slope of this curve at a point P is (𝜕𝑓/𝜕𝑇)𝑙 , which from eqn (2.11) gives the entropy
change per unit extension (𝜕𝑆/𝜕𝑙) 𝑇 , for an isothermal extension at the temperature 𝑇.
In a corresponding way, the intercept 0A of the tangent to the curve at P on the vertical axis
𝑇 = 0 is 𝑓 − (𝜕𝑓/𝜕𝑇)𝑙 , which by eqn (2.12) is equal to the change of internal energy per unit
extension.
Thus, the internal energy and entropy contributions to the force at any given value of the
extension are directly obtainable from the experimental force-temperature plot.
In particular, if the force-temperature plot is linear both internal energy and entropy terms
are independent of temperature.

If in addition the force-temperature relation is represented by a straight line passing through the
origin, the internal energy term is zero, i.e. the elastic force arises solely from the change in entropy
on extension.

27
3724008 Practical - 03

Practical - 03

AIM: To study about the Kinetic theory of Elasticity.

3.1 Early Theories:

Aimed to explain long-range elasticity based on the only common rubber at the time, i.e.,
natural rubber.
Interpretations of mechanical properties were unconvincing.
Failed to account for thermoelastic properties.

3.2 Recognition of Broader Class:

Progress in understanding rubber elasticity acknowledged similarities between natural


rubber and other materials (e.g., gelatin, muscle fibers, silk).
These materials, though diverse in chemical constitution, exhibited similar properties and
were broadly classed as colloids.

3.3 Insights from Wohlisch (1926):

Observed similarities between the contraction of tendons (collagen) on heating and stretched
(crystalline) raw rubber.
Suggested thermal agitation of crystals, micelles, or rod-like molecules as the cause.

3.4 Advances in Polymer Understanding:

Refinement in measuring high molecular weights (100,000-1,000,000), typical value for


rubber ~350,000.
Emergence of the concept of polymers as genuine chemical entities.
Recognition that large molecules could not be rigid structures; internal vibrations and
rotations were expected due to thermal fluctuations.

28
3724008 Practical - 03

3.5 Contributions by Haller (1931):

Calculated that thermal fluctuation of bond lengths and valence angles in a paraffin chain
could cause significant chain curvature.
Initial calculations ignored the critical factor of bond rotation.

3.6 Meyer, von Susich, and Valko (1932) and Karrer (1933):

Developed the now generally accepted theory of rubber elasticity.


Based theory on thermal energy of atoms in long-chain molecules causing greater vibrations
perpendicular to the chain.
This leads to repulsive pressure between parallel or extended chains, equivalent to
longitudinal tension.
Stretched rubber retracts to an irregular, statistically-determined arrangement of molecules,
indicating rubber-like elasticity is due to the chain's ability to adopt a random form through
energy interchange with surrounding atoms.

3.7 Kinetic Theory and Thermo-elasticity:

Kinetic theory indicated tension in stretched rubber (at constant length) should be
proportional to absolute temperature.
Confirmed by later experiments (Meyer and Ferri).
Meyer et al. highlighted the similarity in elastic properties among various materials but
recognized quantitative differences and the significance of crystallization in determining
mechanical properties.

3.8 Karrer’s Muscle Fiber Theory (1933):

Applied similar principles to muscle fibers, suggesting muscle is part of a class of long-chain
structures, including rubber.
Noted Brownian motion causing random chain conformation unless constrained by tensile
force.

29
3724008 Practical - 03
Related random chain conformation to ‘maximum mechanical chaos’ or maximum entropy
(Meyer’s concept).

3.9 Thermodynamic Implications:

Karrer showed muscle work should be accompanied by heat absorption or temperature drop
in an adiabatic process (calculated fall for frog's leg muscle: 0.006°C).
Interest in thermodynamic implications traced back to Joule and Kelvin’s mid-19th century
work on heat changes and temperature effects on elastic retractive force.

3.10 Acceptance of Rubber Elasticity Theory:

Post-publication of Meyer’s theory, there was a gradual acceptance due to:


Initial skepticism due to previous theories with little experimental backing.
Difficulty separating essential elasticity phenomena from thermomechanical effects like
strain-induced crystallization.
Revolutionary nature of the kinetic theory and reluctance to accept the concept of high
polymers.

3.11 Impact of Synthetic Rubbers:

The advent of synthetic rubbers highlighted the theory's significance.

The concept of rubber-like elasticity became attributed to a general type of molecular structure
rather than the specific constitution of the polyisoprene chain.

30
3724008 Practical - 04

Practical - 04

AIM: To study about the cohesive energy density related to rubber/polymeric


materials.

Cohesive Energy Density is a measure of the internal energy required to separate the
molecules of a material. It is particularly important in understanding the physical properties
of rubber and polymeric materials, such as solubility, swelling behavior, and mechanical
strength. The Cohesive Energy Density is related to the van der Waals forces and other
intermolecular forces within the material.
Here considered the specific differences in swelling behaviour among the various polymer-
liquid systems to be capable of representation by means of the empirical parameter 𝜒, which
we saw was related primarily, though not entirely, to the heat of mixing or energy associated
with the interactions between polymer and liquid molecules. In this section we examine in
more detail the nature of this interaction and consider alternative methods of representing its
numerical magnitude.
The most important consideration in this connection is that the energetic interactions
between polymer and solvent molecules, as previously noted, are not specifically related to
the polymeric nature of the polymer component but arise from the local fields of force
between neighboring atoms; they are therefore likely to resemble very closely the energetic
interactions between corresponding pairs of low-molecular-weight liquids of comparable
chemical constitution.
Bearing this in mind, Gee attempted to apply the concept of the cohesive-energy density to
obtain a semi-quantitative understanding of the role of intermolecular interactions in the
determination of swelling properties. The cohesive-energy density is defined as the energy
required to separate all the molecules in a given material from one another; its value (per
mole) is equal to (𝐿 − 𝑅𝑇)/𝑉, where 𝐿 and Vare respectively the molar latent heat of
evaporation and molar volume. For many non-polar liquid mixtures, it has been found that
the heat of dilution ∆𝐻1 (for component 1) may be represented by the expression

1/2 1/2 2 (4.1)


∆𝐻1 = 𝑘𝑉1 (𝑒1 − 𝑒2 ) 𝑣22

31
3724008 Practical - 04
in which 𝑒1 and 𝑒2 are the respective c.e.d. values for the two liquids, 𝑣2 is the volume fraction
of component 2, and 𝑘 is a numerical factor. According to this equation, the heat of dilution is
always positive, and passes through a minimum (zero) when 𝑒1 = 𝑒2 .
Adapting this idea to the case of a polymer-liquid mixture, Gee suggested that, for a given
polymer swollen in a variety of liquids, the swelling should be a maximum when ∆𝐻1 is a
minimum, i.e. for that liquid whose Cohesive Energy Density. is equal to that of the polymer.
He found that a plot of the volume swelling ratio for a given vulcanized rubber against the
Cohesive Energy Density. of the swelling liquid yielded a curve with a pronounced maximum;
the position of this maximum was therefore taken to be a measure of 𝑒2 , the Cohesive Energy
Density. of the polymer, which of course cannot be measured directly. In this way he derived
the value 𝑒2 = 63.7 cal cm-3 (266 J cm-3) for natural rubber. A plot of swelling against
1/2 1/2 1/2
𝑉1 (𝑒1 − 𝑒2 ), in accordance with eqn (3.1), then yielded a curve having the form shown
in Fig 3.1.

Fig 4.1 Relation of swelling to cohesive-energy density

32
3724008 Practical - 04

The solubility parameter

Equations such as (3.1) may be presented rather more conveniently in terms of a parameter
known as the ‘solubility parameter’, which is defined as the square root of the Cohesive Energy
Density, and is denoted by the symbol 𝛿. Numerical values of this parameter for a large
number of polymers, obtained by a method similar to that of Gee, and also for solvents of
various types, have been collected by Sheehan and Bisio (1966); a selection from these figures
is given in Table 3.1. The values of 𝛿 for the various solvents change systematically with their
chemical constitution and tend to increase with increasing polarity. The best solvents (or
swelling agents) for a given polymer are those whose 𝛿 -values are nearest to that of the
polymer, and which therefore are most closely related in chemical structure to the polymer.

Table 4.1

Values of solubility parameter 𝜹 (cal cm-3) 1/2

Polymers

Natural rubber 8.10

Butyl rubber 7.84

Polybutadiene 8.44

Neoprene (polychloroprene) 8.85

Butadiene-styrene (28.5 per cent styrene) 8.33

Butadiene-acrylonitrile (18 per cent acrylonitrile) 8.70

Butadiene-acrylonitrile (39 per cent acrylonitrile) 10.30

Polyethylene 7.94

Solvents

Hexane 7.33
33
3724008 Practical - 04

Decane 7.77

Cyclohexane 8.25

Benzene 9.22

Toluene 8.97

Chloroform 9.30

Carbon tetrachloride 8.63

Acetone 9.74

Ethanol 12.97

Methanol 14.52

Thus, hydrocarbon rubbers (natural rubber, butyl rubber, polybutadiene) are readily soluble
in hydrocarbon solvents (petrol, benzene, etc.), but insoluble in polar liquids such as acetone
and alcohol. For the more polar butadiene-acrylonitrile rubbers the value of 𝛿 increases with
increasing acrylonitrile content, with a consequent increasing resistance to absorption of
hydrocarbon solvents such as petrol and lubricating oils.
On the basis of experimental data for non-polar polymer-solvent systems Blanks and
Prausnitz (1964) estimated the mean numerical value of 𝜒𝑠 at 0.34. This value may be used,
in conjunction with an equation given by Shvarts (1958), namely,

1
𝛿1 = 𝛿2 ± {(𝑅𝑇/𝑉1 )(𝜒 − 𝜒𝑠 )}2 (4.2)

which is substantially the same as (4.1), to relate the values of 𝛿 and 𝜒 for a variety of liquids,
and has proved moderately successful in accounting for the experimental data (Sheehan and
Bisio 1966).

34
3724008 Practical - 05

Practical - 05

AIM: To study about the elastic properties observed in swollen rubber.

The effect of the swelling on the mechanical properties of a cross-linked rubber. To define the
problem, it is assumed that the sample in the unswollen state is in the form of a cube of unit
edge length, and that it contains 𝑁 chains (per unit volume).
The degree of swelling will be defined in terms of the volume fraction 𝑣2 of rubber in the
mixture of rubber and liquid; the volume swelling ratio, referred to the dry state, will thus be
1/𝑣2 , and the corresponding linear dimensions of the swollen sample (linear swelling ratio)

will be 𝜆0 = 1/𝑣21 3 .
In the following argument we shall not be concerned with the nature of the swelling liquid,
nor with the question of the equilibrium degree of swelling and the factors by which it is
determined; the parameter v2 will be introduced merely as a means of defining the state of
swelling of the network, regardless of whether or not this state is the equilibrium state with
respect to the absorption of liquid.
The swelling process itself corresponds to an isotropic expansion of the network, and will
therefore be accompanied by a reduction in the network entropy.
On the application of a stress to the swollen rubber there will be a further reduction of entropy
due to the deformation of the already swollen network.
The total reduction of entropy involved in the transformation from the initial unswollen
unstrained state to the final strained swollen state will thus be the sum of two terms: one
associated purely with the swelling, and the other with the subsequent strain.
Our present interest will be limited to a consideration of the second of these two terms, this
being the one by which the mechanical properties of the swollen rubber are determined.
The expression for the entropy of deformation, though originally worked out for the case of a
constant-volume deformation, is equally applicable to the more general case in which this
restriction is not introduced, since its derivation is not in any way dependent on this
limitation.
The total entropy change ∆S’0 in passing from the unstrained unswollen state to the strained
swollen state for the most pure homogeneous strain may therefore be written in the form

1
∆𝑆’0 = − 𝑁𝑘 (𝑙12 + 𝑙22 + 𝑙32 – 3) (5.1)
2

35
3724008 Practical - 05
where l1, l2 and l3 are the lengths of the edges of the original unit cube, i.e. the principal
extension ratios referred to the unswollen state. The change of entropy ∆S0 associated with
the initial isotropic swelling, corresponding to the linear extension ratio λo,

2
1 1 −
∆𝑆0 = − 𝑁𝑘 (3𝜆20 − 3) = − 𝑁𝑘 (3𝑣2 3 – 3) (5.2)
2 2

The entropy of deformation ∆S’ of the swollen network is the difference between these two
quantities, i.e.

2
1 −
∆𝑆 ′ = ∆𝑆’0 − ∆𝑆0 = − 𝑁𝑘 (𝑙12 + 𝑙22 + 𝑙32 – 3𝑣2 3 ) (5.3)
2

We now define extension ratios λ1, λ2, and λ3 with reference to the unstrained swollen state,
so that l1 = λ1 λ0 = λ1v2-1/3, etc., and obtain

2
1 −
∆𝑆 = − 𝑁𝑘𝑣2 3 (𝜆12 + 𝜆22 + 𝜆23 – 3)

(5.4)
2

This is the entropy of deformation per unit volume of the original unswollen rubber. We
require the entropy of deformation ∆S per unit volume of the swollen rubber; this is given by

1
1
∆𝑆 = 𝑣2 ∆𝑆 ′ = − 𝑁𝑘𝑣23 (𝜆12 + 𝜆22 + 𝜆23 – 3) (5.5)
2

The corresponding strain-energy function for the swollen rubber thus becomes

1
1
𝑊 = −𝑇∆𝑆 = 𝑁𝑘𝑣23 (𝜆12 + 𝜆22 + 𝜆23 – 3) (5.6)
2

This equation represents the properties of the swollen rubber in terms of the extension ratios
measured in the swollen state and the volume fraction of rubber v2. It should be noted,
however, that N is the number of chains per unit volume of the unswollen rubber.
On putting v2 = 1, this result (5.6) reduces to that previously obtained for a dry rubber.

36
3724008 Practical - 05

5.1 Comparison with dry rubber

Comparison of the result expressed by equation with the corresponding expression for the
unswollen rubber shows the dependence of the stored energy on strain to be of the same form
in both cases, the only difference being in the value of the shear modulus. If G and G’ are the
respective moduli in the unswollen and in the swollen states, then

1
𝜌𝑅𝑇 13

𝐺 = 𝐺𝑣23 = 𝑣 (5.7)
𝑀𝑐 2

where ρ is the density in the unswollen state. The resultant stress-strain relations for the
swollen rubber, which are of the type

1
𝑡1 − 𝑡2 = 𝐺𝑣23 (𝜆12 − 𝜆22 ) = 𝐺′ (𝜆12 − 𝜆22 ) (5.8)

show a corresponding difference.


These results imply that the only effect of the swelling is to reduce the modulus in inverse
proportion to the cube root of the swelling ratio, without changing the form of the stress-
strain relations.

5.2 Alternative formula for network entropy

By the use of a different method of calculating the entropy of deformation, based on the
consideration of the entropy of formation of the network from a corresponding set of
independent chains, Flory (1950) and Wall and Flory (1951) have obtained the result

1
∆𝑆′0 = − 𝑁𝑘 {𝑙12 + 𝑙22 + 𝑙32 − 3 − 𝑙𝑛(𝑙1 𝑙2 𝑙3 )} (5.9)
2

which differs from (5.1) by the inclusion of the term −ln(l1 l2 l3 ).


The method of derivation of this formula has been the subject of much discussion, which,
however, remains inconclusive.

37
3724008 Practical - 05
Fortunately, the two formulae (5.1) and (5.9) become identical for the case of an unswollen
rubber, for which ln(l1 l2 l3 ) = 0, while for a swollen rubber the effect of the additional term is
in most practical cases rather small.

38
3724008 Practical - 06

Practical - 06

AIM: To determine the Crosslinking density of given vulcanizate sample by the


swelling method (Chemical Method).

Determining the crosslinking density of a vulcanizate sample by the swelling method involves
several steps. This chemical method is based on the theory that the degree of swelling of a
crosslinked polymer in a solvent is related to its crosslink density.

6.1 Materials Needed


Vulcanizate sample, Solvent (commonly used: toluene or benzene), Analytical balance,
Desiccator and Container for swelling.

39
3724008 Practical - 06

6.2 Procedure
a] Sample Preparation
Cut the vulcanizate sample into small pieces, typically around 1 cm² each.
Weigh the dry sample accurately using an analytical balance (𝑊𝑑 ).

b] Swelling Process
Immerse the vulcanizate sample in a suitable solvent (e.g., toluene) at room temperature.
Allow the sample to swell until it reaches equilibrium. This may take 24 to 48 hours.
Ensure that the solvent completely covers the sample during this period.

c] Measuring Swollen Sample


Remove the swollen sample from the solvent.
Blot the sample gently with filter paper to remove surface solvent without squeezing the
sample.
Weigh the swollen sample immediately (𝑊𝑠 ).

d] Calculate the Degree of Swelling (Q)


The degree of swelling (𝑄) is calculated using the formula:

𝑊𝑠 − 𝑊𝑑
𝑄= (6.1)
𝑊𝑑

where 𝑊𝑠 is the weight of the swollen sample, and 𝑊𝑑 is the weight of the dry sample.

e] Determining Crosslink Density


Use the Flory-Rehner equation to calculate the crosslink density. The equation is:

40
3724008 Practical - 06

ln(1 − 𝑉𝑟 ) + 𝑉𝑟 + 𝜒𝑉𝑟2
𝑣= − 1/3 (6.2)
𝑉1 (𝑉𝑟 − 𝑉𝑟 /2)

where: 𝑣 is the crosslink density (moles of crosslinks per unit volume).


𝑉𝑟 is the volume fraction of polymer in the swollen state, given by:

1
𝑉𝑟 = (6.3)
1 + 𝑄(𝜌𝑝 /𝜌𝑠 )

where 𝜌𝑝 is the density of the polymer and 𝜌𝑠 is the density of the solvent.
𝜒 is the polymer-solvent interaction parameter, which can be found in literature or
determined experimentally.
𝑉1 is the molar volume of the solvent.

Vulcanized rubber does not dissolve but swells in solvents and the extent of swelling depends
on cross link density. A tightly cured sample will swell less than a lightly cured sample.
Equilibrium swelling in benzene can be used to determine the crosslink density

𝑁𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑁𝑒𝑡 𝑤𝑜𝑟𝑘 𝑐ℎ𝑎𝑖𝑛𝑠


𝐶𝑟𝑜𝑠𝑠𝑙𝑖𝑛𝑘 𝐷𝑒𝑛𝑠𝑖𝑡𝑦 (𝑉𝑒 ) = (6.4)
𝑉𝑜𝑙𝑢𝑚𝑒 𝑜𝑓 𝑡ℎ𝑒 𝑢𝑛𝑠𝑤𝑜𝑙𝑙𝑒𝑛 𝑛𝑒𝑡 𝑤𝑜𝑟𝑘

The Crosslink Density is determined as follows:


The test sample is extracted in acetone for 16 hour to remove any soluble material and then
dried at 70°C to remove acetone. Its weight is recorded. The sample is allowed to swell in
benzene for 7 days at 25°C.
Assuming that there are 2 crosslinks for each elastically effective network chain; the crosslink
density can be calculated using following equation:
The volume fraction of the polymer in the swollen network is calculated by assuming addivity
of volumes and is used to arrive at the crosslink density by the Flory-Rehner Equation:

−[log 𝑛 (1 − 𝑉𝑟 ) + 𝑉𝑟 + (𝑋1 𝑉𝑟 2 )]
(𝑉𝑒 ) = 1 (6.5)
𝑉1 [(𝑉𝑟 3 − 𝑉𝑟 )/2]

41
3724008 Practical - 06
where,

𝑉𝑒 = Effective number of chains in a real network per unit volume [Crosslink density (mol/cm3)]

𝑉𝑟 = Volume fraction of rubber in the swollen network.

𝑊𝑒𝑖𝑔ℎ𝑡 𝑜𝑓 𝑑𝑟𝑦 𝑟𝑢𝑏𝑏𝑒𝑟 𝑊𝑒𝑖𝑔ℎ𝑡 𝑜𝑓 𝑠𝑜𝑙𝑣𝑒𝑛𝑡 𝑎𝑏𝑠𝑜𝑟𝑏𝑒𝑑 𝑏𝑦 𝑟𝑢𝑏𝑏𝑒𝑟


i.e. = +
𝐷𝑒𝑛𝑠𝑖𝑡𝑦 𝑜𝑓 𝑑𝑟𝑦 𝑟𝑢𝑏𝑏𝑒𝑟 𝐷𝑒𝑛𝑠𝑖𝑡𝑦 𝑜𝑓 𝑠𝑜𝑙𝑣𝑒𝑛𝑡

𝑉1 = Molar volume of solvent (cm3/mol), 𝑋1 = Polymer solvent interaction parameter (NR & Benzene
= 0.437, SBR & Benzene = 0.442, NBR-Med ACN & Benzene = 0.390).

Ensure that the sample is completely dry before the initial weighing.
Handle the swollen sample carefully to avoid any loss of solvent before weighing.
The accuracy of the calculation depends on the precision of the weights and the correctness
of the polymer-solvent interaction parameter.
By following these steps, you can determine the crosslinking density of a vulcanizate sample
using the swelling method.

42
3724008 Practical - 07

Practical – 07

AIM: Mulin’s Effect

7.1 Stress-Strain Properties

Test sheets vulcanized for various cure times at a particular vulcanization temperature are
prepared.
The test samples are tested for stress-strain properties using a tensile testing machine.
The values of the modulus at 100, 200, 300 % elongation, the ultimate tensile strength,
elongation at break, tear strength, etc. are plotted against the vulcanization time at the
selected temperature.
The values of each cure time represent the state or degree of vulcanization.
A plot of tensile strength Vs cure time reveals that the tensile strength values initially increase
with the cure time and then pass through an optimum value before showing a decreasing
trend as the cure time increases.
The cure time at optimum value of tensile strength is the state of optimum vulcanization at
the vulcanization temperature as far as tensile strength is concerned.
The optimum state of vulcanization could be different for different properties and according
to end-property requirements the optimum state of vulcanization is determined.
Measurement of tensile modulus at 100% elongation (E) is directly proportional to the
crosslink density (1/2 Mc) which is expressed by following relationship.

3𝑃𝑟 𝑅𝑇
𝐸=
2𝑀𝑐

where, 𝑃𝑟 = Density of Rubber,


𝑅 = Universal gas constant,
𝑇 = Absolute Test Temperature,
𝑀𝑐 = Number average relative mol. mass between the cross links.
The accuracy of 100% modulus measurements may raise some concern and hence often 300%
modulus is measured for day-to-day quality control in a factory.

43
3724008 Practical - 08

Practical - 08
AIM: To study about the Statistical properties of Rubber long chain molecule.

8.1 Introduction
The study of the statistical properties of the long-chain molecule is a complex one, and may
be carried to various stages of refinement.
In the present context we shall be concerned primarily with those statistical properties which
are common to all rubber-like molecules, and which may conveniently be dealt with in terms
of an idealized chain of freely rotating links.
The statistical properties of such an idealized model involve geometrical considerations only,
and may be dealt with quite simply.
The more realistic treatment of an actual molecule, however, requires that the effects of
energetic interactions between different segments of the same chain should be taken into
account.
This is a problem of considerably greater complexity, which falls outside the scope of the
present work.
Even with the simplest model of the chain the degree of mathematical difficulty depends on
the range of extension considered.
If only moderate extensions are involved a very simple statistical treatment, known as the
Gaussian treatment, is sufficient.
But for the higher range of extensions this treatment becomes increasingly inadequate as the
distance between the two ends of the chain approaches the fully extended length, and a more
elaborate non-Gaussian treatment has to be used.
The Gaussian treatment for the single chain will be developed in the present chapter, and its
application to the problem of the network in the one which follows.

44
3724008 Practical - 08

8.2 Statistical form of long-chain molecule


The statistical form of the long-chain molecule may be illustrated by considering an idealized
model of the poly-methylene or paraffinic type of chain (CH2)n, in which the angle between
successive bonds (i.e. the valence angle) is fixed but complete freedom of rotation of any given
bond with respect to adjacent bonds in the chain is allowed. This is illustrated in Fig. 8.1, in
which the first two bonds C1C2 and C2C3 are represented as lying in the plane of the paper.

Fig 8.1 Rotation about bonds in paraffin-type molecule.

The third bond, C3C4, will in general not lie in this plane but will rotate in a random manner
about the bond C2C3 as axis.
Similarly, C4C5, will rotate about C3C4, and so on.
The chain will thus take up an irregular or randomly kinked form (Fig. 8.2(b)) in which the
distance between its ends is very much less than that corresponding to the outstretched or
planar zig-zag form (Fig. 8.2(a)).
The actual conformation will be subject to continual fluctuation due to thermal agitation and

Fig 8.2 (a) Planar zig-zag; (b) randomly kinked chain.

45
3724008 Practical - 08
hence cannot be defined explicitly, but it is possible to specify some of the properties of the
system in statistical terms, or in terms of certain average values.
Thus, for example, it may be shown that the root-mean-square (r.m.s.) value of the distance 𝑟
between the ends of a chain containing 𝑛 bonds (where 𝑛 is a large number) is given by

1
1 1 1 + 𝑐𝑜𝑠𝜃 2
(𝑟̅̅̅2 )2 = 𝑙𝑛2 ( ) (8.1)
1 − 𝑐𝑜𝑠𝜃

where 𝑙 is the bond length and 𝜃 the supplement to the valence angle. Taking the valence angle
1 1 1
to be 109 2 °, 𝜃 = 70 2 °, and therefore 𝑐𝑜𝑠𝜃 = 3, hence for this case

1 1
(𝑟̅̅̅2 )2 = 𝑙(2𝑛)2 (8.1a)

Fig 8.3 Form of 1000-link poly-methylene chain according to the statistical theory.

46
3724008 Practical - 08
This result illustrates the quite general conclusion that the mean (or r.m.s.) dimensions of any
chain of this kind are proportional to the square root of the number of bonds or links which it
contains.
The general form of the idealized freely rotating chain is more clearly indicated in Fig. 8.3,
which represents an actual wire model of a poly-methylene chain containing 1000 links.

In the construction of this model the links were set at the required valence angle, but the
position of each successive link in the circle of rotation was chosen at random by the throw of
a die.
This gave an equal probability for each of six equally spaced positions in the circle, which may
be regarded as a sufficiently close approximation to complete randomness.
The photographic reproduction in Fig. 8.3 is, of course, equivalent to a two-dimensional
projection of the actual three-dimensional form.
The particular conformation obtained is just one of a very large number of equally probable
forms (6 to be precise) which might have been produced by this method of construction, but
it happens to be a fairly ‘average’ sample, the end-to-end distance being within about 10 per
cent of the calculated r.m.s. value (eqn (8.1a)).

47
3724008 Practical - 08

8.3 Polyisoprene chains

The calculation of the mean-square length of a molecule is not limited to the case where all
the bonds and valence angles are equal.
A general method of treating more complicated structures has been worked out by Wall
(1943) and applied to the polyisoprene chain in which there is one double bond per four chain
carbon atoms.
The two single bonds adjacent to the double bond are fixed in a plane; apart from this
restriction, freedom of rotation about single bonds is assumed.
Wall considered both natural rubbers, in which the single bonds have the cis-configuration
with respect to the double bond, and gutta-percha, which is otherwise similar to natural
rubber, but in which the trans -configuration occurs.
For the former he obtained, on the basis of accepted values of bond lengths and valence angles,
the numerical result

1 1
(𝑟̅̅̅2 )2 = 0.201 ∙ 𝑛2 𝑛𝑚 (8.2)

where 𝑛 (assumed large) is the total number of bonds in the chain (including double bonds).

48
3724008 Practical - 09

Practical - 09
AIM: To study about the principal stresses observed in unstrained and strained state
on Rubber rectangular block.

Fig 9.1 Pure homogeneous strain: (a) the unstrained state; (b) the strained state.

In the state of pure homogeneous strain depicted in Fig. 9.1, there are three principal stresses
acting in directions parallel to the principal axes of strain on planes corresponding to the faces
of the rectangular block.
If these stresses are denoted by 𝑡1 , 𝑡2 , and 𝑡3 , the problem is to determine their magnitude in
terms of the corresponding principal extension ratios 𝜆1 , 𝜆2 , and 𝜆3 .
The derivation of these general stress-strain relations is based on the expression for the
elastically stored energy, and makes use of the condition for constancy of volume.
It is important to distinguish between true and nominal stresses. We shall define 𝑡1 , 𝑡2 , and 𝑡3
as the forces per unit strained area, or true stresses, and use the symbols 𝑓1 , 𝑓2 , and 𝑓3 to
represent the corresponding forces per unit unstrained area, or nominal stresses.
For a body initially in the form of a unit cube, the force 𝑓1 acts on an area 𝜆2 𝜆3 measured in the
strained state; the corresponding true stress is therefore,

𝑓1
𝑡1 = = 𝜆1 𝑓1 (9.1)
𝜆2 𝜆3

with similar expressions for 𝑡2 , and 𝑡3 .


To determine the forces, the method used is to equate the change in the stored energy in any
variation of 𝜆1 , 𝜆2 , and 𝜆3 to the work done by the applied forces.

49
3724008 Practical - 09
However, these three extension ratios cannot be varied independently of one another, since
they are connected by the constant-volume relation (9. 1), which implies that if any two of
these are chosen as independent variables the third is then determined.
Thus, if 𝜆1 and 𝜆2 are chosen as independent variables we have 𝜆3 = 1/𝜆1 𝜆2 .
The strain energy function may thus be written

1
𝑊= 𝐺(𝜆1 2 + 𝜆2 2 + 1/𝜆1 2 𝜆2 2 ) (9.2)
2

Let us consider, for simplicity, the case when only two forces 𝑓1 and 𝑓2 are applied, so that 𝑓3 =
t1 = 0 (Fig. 4.3). If the extension in the direction 𝜆1 is increased by the amount 𝑑𝜆1 , while 𝜆2

Fig 9.2 Two-dimensional extension under the action of forces 𝑓1 and 𝑓2 .

is held constant, the only work done by the applied forces is that done by the force 𝑓1 .
Assuming the unstrained block to be in the form of a unit cube we have then

𝑑𝑊 = 𝑓1 𝑑𝑙 = 𝑓1 𝑑𝜆1 (9.3)

where dl is the increase in length. From (9.2) we have also

𝑑𝑊 = (𝜕𝑊/𝜕𝜆1 )𝑑𝜆1 = 𝐺(𝜆1 − 1/𝜆1 3 𝜆2 2 )𝑑𝜆1 (9.4)

Equating these values of dW we obtain

𝑓1 = 𝐺(𝜆1 − 1/𝜆1 3 𝜆2 2 ) (9.5)

50
3724008 Practical - 09
The corresponding principal stress 𝑡1 , from (9.1), is

𝑡1 = 𝜆1 𝑓1 = 𝐺(𝜆1 2 − 1/𝜆1 2 𝜆2 2 ) = 𝐺(𝜆1 2 −𝜆3 2 ) (9.6)

A similar result is obtained for 𝑡2 . The three principal stresses thus become

𝑡1 = 𝐺(𝜆1 2 −𝜆3 2 ); 𝑡2 = 𝐺(𝜆2 2 −𝜆3 2 ); 𝑡3 = 0 (9.7)

To obtain the general solution corresponding to the case when t 3 is not zero, let us
superimpose on to the above stress system a negative hydrostatic pressure −𝑝 (i.e. a
hydrostatic tensile stress +𝑝).
Since in accordance with Assumption 3 the volume must remain constant this hydrostatic
stress can have no effect on the state of strain.
The stresses, however, are each increased by the amount 𝑝 to give

𝑡1 = 𝐺(𝜆1 2 −𝜆3 2 ) + 𝑝; 𝑡2 = 𝐺(𝜆2 2 −𝜆3 2 ) + 𝑝; 𝑡3 = 𝑝 (9.8)

The interpretation of eqns (9.8) is that for a material which is incompressible with respect to
volume the stresses are indeterminate to the extent of an arbitrary hydrostatic pressure 𝑝.
Only the differences between any two of the principal stresses, which are of course unaffected
by the addition of a hydrostatic stress, are determinate.
From either (9.7) or (9.8) these are given by

𝑡1 − 𝑡2 = 𝐺(𝜆1 2 −𝜆2 2 ),

t 2 − t 3 = G(𝜆2 2 −𝜆3 2 ), (9.9)

t 3 − t1 = G(𝜆3 2 −𝜆1 2 ).

In most practical problems the fact that the principal stresses are, in general, indeterminate
does not lead to any difficulty, since one of them can usually be obtained from the boundary
conditions.
In particular, if one of the surfaces is stress-free as, for example, in the case of simple extension,
the corresponding principal stress is zero, and the indeterminacy is automatically removed.

51
3724008 Practical - 09
For a simple extension corresponding to an extension ratio 𝜆1 , we have 𝑡2 = 𝑡3 = 0, and also,
from another equation, 𝜆2 2 = 𝜆3 2 = 1/λ1 .
Substitution into the first of eqns (9.9) then yields

𝑡1 = 𝐺(𝜆1 2 − 1/𝜆1 ) (9.10)

Making use of (9.1) the corresponding force per unit unstrained area becomes

𝑓1 = 𝐺(𝜆1 − 1/𝜆1 2 ). (9.10a)

52

You might also like