Elasticity, Rubber Like

  • Upload
    ash1968

  • View
    218

  • Download
    0

Embed Size (px)

Citation preview

  • 7/28/2019 Elasticity, Rubber Like

    1/37

    ELASTICITY, RUBBER-LIKE

    Introduction

    Elasticity is the reversible stressstrain behavior by which a body resists and

    recovers from deformation produced by a force (1). This behavior is exhibited by

    rubber-like materials in a unique and extremely important manner. Unlike metals

    or glasses, they canundergo very large deformations without rupture (and arethus

    similar to liquids) and then can come back to their original shape (as do solids).

    This exceptional ability was investigated in 1805 (2,3), well before the concept of

    polymers as long-chain molecules was established in the 1920s. It was observed

    that a rubber gave off heat upon stretching and, when submitted to a constant

    load, became shorter as temperature was increased. Later, Kelvin (4) derived the

    thermodynamic laws of elasticity, and more careful experiments were performed

    by Joule (5). Many books and review papers have been published on the subject

    (636).

    Deformation and Mechanical Testing

    When a body is submitted to external stress or pressure, it undergoes a change

    in shape or volume. Mechanical testing of rubber involves application of a force

    to a specimen and measurement of the resultant deformation or application of adeformation and measurement of the required force. The results are expressed in

    terms of stress and strain and thus are independent of specimen geometry. Stress

    is the force f per unit original cross-sectional area A, and strain the deformation

    per unit original dimension. The units of stress are Newtons per square meter

    (N/m2 or Pa); strain is dimensionless. The ratio of stress to strain is called the

    216Encyclopedia of Polymer Science and Technology. Copyright John Wiley & Sons, Inc. All rights reserved.

  • 7/28/2019 Elasticity, Rubber Like

    2/37

    Vol. 2 ELASTICITY, RUBBER-LIKE 217

    A

    L

    f

    f

    L0

    (a) (b)

    Fig. 1. Uniaxial extension of a bar-shaped sample. (a) No load; (b) tensile load. A = area;f = force; L0 = original length; L = length.

    modulus. A material is Hookean when its modulus is independent of strain (typical

    of a metal spring or wire); elastomers are Hookean only in the range of very small

    deformations.

    The simplest deformations are isotropic compression under hydrostatic pres-

    sure, uniaxial extension (or compression), and simple shear. They are used to

    determine the bulk modulusK, Youngs modulusE, and shear modulus G, respec-

    tively (3744). Elastomer testing for commercial applications is highly dependent

    on the method, the conditions (eg, strain rate, temperature), and the shape of the

    samples. Therefore, mechanical tests have been standardized for uniformity and

    simplicity by the American Society for Testing and Materials (ASTM) (45).

    Isotropic Compression. The volume of a sample decreases from V0 to Vwhen submitted to a hydrostatic pressure p. The bulk modulus is the ratio of this

    pressure to the volume change per unit volume V/V0:

    K= p/(V/V0) (1)

    where V = V0 V; K is the reciprocal of the compressibility and can be deter-

    mined by direct measurement of compressibility (46) or velocity of longitudinal

    elastic waves (sound) or by theoretical calculations (39).Uniaxial Extension. A rubber strip of original length L0 is stretched uni-

    axially to a length L, as illustrated in Figure 1. The stretch and elongation areL =L L0 and =L/L0, respectively. The strain (also known as the relative

    deformation, linear dilation, or extension) and the elongation or extension ratio are related by

    = L/L0 = 1 (2)

  • 7/28/2019 Elasticity, Rubber Like

    3/37

    218 ELASTICITY, RUBBER-LIKE Vol. 2

    The lengthwise extension is in general accompanied by a transverse contraction.

    The ratio of the change w in width per unit width w0 to the change in length per

    unit length is called Poissons ratio (47):

    p = (w/w0)/(L/L0) (3)

    Ideal rubbers and liquids deform at constant volume, for which p is equal to 0.5.

    Poissons ratio for real elastomers may be experimentally determined by applying

    extensometers in the transverse and axial directions of a sample. Approximate

    values of p thus determined are 0.33 for glassy polymers, 0.4 for semicrystalline

    polymers, and 0.49 for elastomers. Youngs modulus E is the ratio of normal stress

    f/A to corresponding strain :

    E = (f/A)/ (4)

    Generally, stressstrain curves deviate markedly from a straight line, as illus-

    trated in Figure 2. The Youngs modulus at small deformation is the slope tan ofthe stressstrain curve at the origin (initial tangent modulus). It can be measured

    in flexure or by tensile experiments (48).

    Tension testing of a vulcanized elastomer also permits the determination

    of the tensile strength, which is the maximum stress applied during stretching

    a specimen to rupture; the corresponding rupture strain is called the maximum

    extensibility (4952). Typical values are given in Table 1 (14). Dumbbell and ring

    specimens can be used.

    6

    6

    5

    5

    4

    4

    3

    3

    2

    f/A,

    N/mm2

    2

    1

    10

    7 8

    Fig. 2. Stressstrain curve of a non-Hookean solid; typical curve for an elastomer inuniaxial extension.

  • 7/28/2019 Elasticity, Rubber Like

    4/37

    Vol. 2 ELASTICITY, RUBBER-LIKE 219

    Table 1. Properties of a Typical Rubber, Metal, and Glassa

    Property Rubber Metal Glass

    Breaking extension, % 500 Large, plastic 3

    Elastic limit, % 500 2 3Tensile strength, MN/m2b 10c 5 104 5 103

    aRef. 14.bTo convert MN/m2 (MPa) to psi, multiply by 145.c>103 if based on area at break.

    Forcegauge

    Recorder

    CathetometerConstant T

    N2

    Fig. 3. Schematic diagram of a typical apparatus used to measure elastomer stress as afunction of strain (21). T = Temperature.

    A typical apparatus used to measure the equilibrium stress of an elongated

    network as a function of strain and temperature is shown in Figure 3 (21). The

    rubber strip is held between two clamps and maintained under a protective atmo-

    sphere of nitrogen. The sample length, required to characterize its deformation,

    is obtained by means of a cathetometer or traveling microscope (the central test

    section of the sample is delineated by ink marks applied before loading). Values

    of the force are obtained from a calibrated stress gauge, the output of which isdisplayed as a function of time on a standard recorder. Measurements are made

    at elastic equilibrium; the influence of temperature can also be studied. Another

    example of a stretching device is an automatic stress relaxometer (53).

    Simple Shear. Simple shear, illustrated in Figure 4, is a deformation inwhich the height H and surface A of the sample are held constant. The shear

  • 7/28/2019 Elasticity, Rubber Like

    5/37

    220 ELASTICITY, RUBBER-LIKE Vol. 2

    A

    H

    y

    xf

    U

    y

    A

    H

    f

    x

    (a) (b)

    Fig. 4. Simple shear of a rectangular sample. (a) No deformation; (b) simple shear.A = area; H = height; f = force; U = linear displacement.

    modulus or rigidity G is expressed by the relations

    G = (f/A)/(U/H) = (f/A)tan f/A (5)

    where is the shear strain and f/A the shear stress. As in the case of Youngs

    modulus, the shear modulus (41,42) is a material constant if stress and strain are

    directly proportional. If they are not, the shear modulus at small deformation is

    employed; it is defined as the slope at the origin of the stressstrain curve. The

    shear modulus is generally measured on a cylindrical specimen and a tension-

    testing machine.

    Typical values of moduli and Poissons ratios for metals, ceramics, and poly-

    mers are given in Table 2 (39). Moduli of rubbers are strikingly low, and even

    those of other types of polymers are of the order of one tenth of those of metals.

    However, compared at equal weights, ie, ratio of modulus to density , polymers

    (except rubbers) compare favorably.Relationships between Moduli and Poissons Ratio. On the basis of

    the theory of elasticity of isotropic solids, the moduli and Poissons ratio are inter-

    related (54) as follows:

    E = 2G(1 + p) = 3K(1 2p) (6)

    For elastomers, the Poissons ratio is nearly 0.5, and thus E 3G.

    Conditions for Rubber-like Elasticity

    Long, Highly Flexible Chains. Elastomers consist of polymeric chains

    which are able to alter their arrangements and extensions in space in response toan imposed stress. Only long polymeric molecules have the required exceedingly

    large number of available configurations.

    It is necessary that all the configurations are accessible; this means that

    rotation must be relatively free about a significant number of the bonds joining

    neighboring skeletal atoms.

  • 7/28/2019 Elasticity, Rubber Like

    6/37

    Table 2. Poissons Ratio, Moduli, and Density of Metals, Ceramics, and Polymersa

    Youngs modulus E, Shear modulus G, Bulk modulus K, Density , E

    Material p 109 N/m2 b 109 N/m2b 109 N/m2 b g/cm3

    Metals

    Cast iron 0.27 90 35 66 7.5

    Steel (mild) 0.28 220 86 166 7.8

    Aluminum 0.33 70 26 70 2.7

    Copper 0.35 120 44.5 134 8.9

    Lead 0.43 15 5.3 36 11.0

    Mercury 0.5 0 0 25 13.55 Inorganics

    Quartz 0.07 100 47 39 2.65

    Vitreous silica 0.14 70 30.5 32.5 2.20

    Glass 0.23 60 24.5 37 2.5

    Granite 0.30 30 11.5 25 2.7

    WhiskersAlumina 2000 1000 667 3.96

    Carborundum 1000 500 333 3.15

    Graphite 1000 500 333 2.25 Polymers

    Polystyrene 0.33 3.2 1.2 3.0 1.05

    Poly(methyl Methacrylate) 0.33 4.15 1.55 4.1 1.17

    Nylon-6,6 0.33 2.35 0.85 3.3 1.08

    Polyethylene (low density) 0.45 1.0 0.35 3.33 0.91

    Ebonite 0.39 2.7 0.97 4.1 1.15

    Rubber 0.49 0.002 0.0007 0.033 0.91 Liquids

    Water 0.5 0.0 0.0 2.0 1.0

    Organic liquids 0.5 0.0 0.0 1.33 0.9

    aRef. 39. Courtesy of Elsevier.bN/m2 = Pa. To convert Pa to psi, multiply by 0.000145.

    221

  • 7/28/2019 Elasticity, Rubber Like

    7/37

    222 ELASTICITY, RUBBER-LIKE Vol. 2

    Fig. 5. Segment of an elastomeric network.

    Network Structure. Chains must be joined by permanent bonds calledcross-links, as illustrated in Figure 5. The network structure thus obtained is

    essential so as to avoid chains permanently slipping by one another, which would

    result in flow and thus irreversibility, ie, loss of recovery. These cross-links may

    be chemical bonds or physical aggregates, eg, glassy domains in multiphase block

    copolymers (55,56).

    At the end of the cross-linking process, the topology of the mesh is composedof the different entities represented in Figure 6 (16,5759). An elastically active

    junction is one joined by at least three paths to the gel network (60,61). An ac-

    tive chain is one terminated by an active junction at both its ends. Rubber-like

    elasticity is due to elastically active chains and junctions. Specifically, upon de-

    formation the number of configurations available to a chain decreases and the

    resulting decrease in entropy gives rise to the retractive force.Weak Interchain Interactions. Apart from the effects of the cross-links,

    the molecules must be free to move reversibly past one another, that is, the in-

    termolecular attractions known as secondary or van der Waals forces, which exist

    between all molecules, must be small. Specifically, extensive crystallization should

    not be present, and the polymer should not be in the glassy state.

    Differences between Elastomers and Metals

    Elastomers and metals differ greatly with regard to deformation mechanisms (26,

    62,63). Metals and minerals are formed of atoms arranged in a three-dimensional

    crystalline lattice, joined by powerful valence forces operating at relatively short

  • 7/28/2019 Elasticity, Rubber Like

    8/37

    Vol. 2 ELASTICITY, RUBBER-LIKE 223

    (a) (b)

    (d)(c)

    Fig. 6. Structural features of a network. (a) Elastically active chain; (b) loop; (c) trappedentanglement; (d) chain end.

    range. Deformation of such materials involves changes in the interatomic distance,

    which requires large forces; hence the elastic modulus of these materials is very

    high. After a small deformation, slippage between adjacent crystals occurs at the

    yield point, and the deformation increases much more rapidly than the stress and

    becomes irreversible or plastic. The primary effect of stretching a metal short of

    this yield point is the increase Em in energy caused by changing the distance

    d of separation between metal atoms. The sample recovers its original lengthwhen the force is removed, since this process corresponds to a decrease in energy.

    Heating increases oscillations about the minimum of the asymmetric potential

    energy curve and thus causes the usual volume expansion.

    Stretching of elastomers does not involve any significant changes in the in-

    teratomic distances, and therefore the forces required are considerably lower. The

    number of available configurations for a network chain is reduced in the defor-

    mation process. After suppression of the stress, the specimen recovers its original

    shape, since this corresponds to the most disordered state. Thus the retractive

    force arises primarily from the tendency of the system to increase its entropy

    toward the maximum value it had in the undeformed state. At high elongations,

    stressstrain curves turn upward, a behavior very unlike that of metals. The num-

    ber of available configurations is drastically reduced in this region, and chains

    reach the limits of their extensibility. For polymers with regular structures, crys-tallization may also be induced. Further elongation may then require deformation

    of bond angles and length, which requires much larger forces.

    At constant force, heating increases disorder, forcing the sample in the direc-

    tion of the state of disorder it had at a lower temperature and smaller deformation.

    The result is therefore a decrease in length.

  • 7/28/2019 Elasticity, Rubber Like

    9/37

    224 ELASTICITY, RUBBER-LIKE Vol. 2

    Extension of Elastomers and Compression of Gases

    The extension of elastomers and the compression of gases are associated with a

    decrease in entropy. Thus, similar behavior is expected in regard to (adiabatic)

    deformation and heating. The work of deformation of a gas is dW = pdV, wherep is the pressure and V the volume. For an elastomer it is dW=fdL, wheref is the

    force andL the length; the term pdV is not taken into account, since there is only

    a negligible change in rubber volume during stretching. These observations can be

    used to explain Goughs experiments by making use of classical thermodynamic

    principles (64).

    Statistical Distribution of End-To-End Dimensions of a Polymer Chain

    Before treating the statistical properties of a network, the statistics of a single

    chain must be considered, mainly to establish the relationship between the num-

    ber of configurations and the deformation (12,16,6572). A polymeric chain isconstantly changing its configuration by Brownian motion. Statistical methods

    and idealized models permit calculation of the average properties of such a chain.The Freely Jointed Chain. This type of idealized chain consisting of n

    links of length l is represented in Figure 7, where r is the end-to-end distance.

    When the chain is completely extended, R = nl. The chain may assume many

    configurations, each associated with an end-to-end distance ri. In statistics, it is

    equivalent to consider a molecule at different times or an assembly ofNmolecules

    at the same time. An average quantity describing the assembly is the mean-squareend-to-end distance r2 defined by

    r2 =1

    N

    N

    i = 1

    r2i (7)

    where ri is the end-to-end distance of the ith chain. The vector ri is the sum of the

    link vectors Ij:

    ri =n

    j = 1

    Ij (8)

    Ij

    r

    Fig. 7. Ideal chain formed with n links of length l. r = End-to-end distance, Ij = linkvector.

  • 7/28/2019 Elasticity, Rubber Like

    10/37

    Vol. 2 ELASTICITY, RUBBER-LIKE 225

    and r2i is the scalar product of ri with itself:

    r2i =n

    j = 1

    I2j + 2k

  • 7/28/2019 Elasticity, Rubber Like

    11/37

    226 ELASTICITY, RUBBER-LIKE Vol. 2

    0.04

    0.03

    0.02

    0.01

    00 10 20 30 40 50

    r, nm

    W(r),nm1

    Fig. 8. Radial distribution function W(r) of the chain displacement vectors for chainmolecules consisting of 104 freely jointed segments, each of length l = 0.25 nm; W(r) isexpressed in nm 1 and r in nm (3). Courtesy of Cornell University Press.

    take into account the significant geometric and conformational differences known

    to exist among different types of polymer chains (83). The distribution obtained

    from a more general treatment is of the more complicated form (8486) shown as

    follows:

    W(r)dr = const.exp

    r0

    L 1(r/nl)dr/l

    4r2dr (16)

    where L 1 is the inverse Langevin function and

    L(u) = coth u 1/u (17)

    Equation (16) can be expanded in the series

    W(r)dr = const.exp{n[(3/2)(r/nl)2

    + (9/20)(r/nl)4

    + (99/350)(r/nl)6

    + ]}4r2dr

    (18)

    and the Gaussian distribution (eq. (13)) recovered for r nl. A comparison of the

    Gaussian and inverse Langevin distributions for n = 6 is shown in Figure 9 (7).

    Chain with Bond-Angle Restrictions. Although the chain in the afore-mentioned treatment was assumed to be freely jointed, a real polymer chain has

    fixed bond angles . Therefore, the second term in equation (9) is no longer zero.

    The scalar product of two vectors IiIj is l2 cos|j i|. It can be shown (9,73,74) that

    r2 nl2(1 cos )/(1 + cos ) (19)

  • 7/28/2019 Elasticity, Rubber Like

    12/37

    Vol. 2 ELASTICITY, RUBBER-LIKE 227

    0.4

    0.3

    0.2

    0.1

    0.00 1 2 3 4 5 6

    W(r)

    A

    B

    Fig. 9. Distribution function W(r) for a six-link random chain, with length l = 1 in arbi-trary units: A, Gaussian limit; B, inverse Langevin function (7). Courtesy of Oxford Uni-versity Press.

    In the particular case of a tetrahedrally bonded chain, = 109.5 and

    r2 = 2nl2, twice the value for a freely jointed chain. Thus, a real chain is quite

    different from this ideal representation. Nevertheless, any flexible real chain can

    be represented by a simple model which is the statistical equivalent of a freely

    jointed chain (87,88). The two conditions are that the real and freely jointed chains

    have the same mean-square end-to-end distance and the same length at complete

    extension:

    r2 = r2e = nel2e (20)

    and

    R=Re = nele (21)

    Thus, only one model chain is equivalent to the real one, obeying the same

    Gaussian distribution function and composed of ne segments of length le given

    by

    ne =R2/r2 (22)

    le = r2/R (23)

  • 7/28/2019 Elasticity, Rubber Like

    13/37

    228 ELASTICITY, RUBBER-LIKE Vol. 2

    Chain with Bond-Angle Restrictions and Hindered Rotations. Theangle of rotation of a single bond around the axis formed by the preceding one

    may be restricted by steric interferences between atoms. When n is large and the

    average value of cos , ie, cos , is not too close to unity, the following relationship

    can be established (89,90):

    r2 = nl2

    1 cos

    1 + cos

    1 + cos

    1 cos

    (24)

    When the rotation is not hindered, ie, when cos = 0, equation (24) is equivalent

    to equation (19).

    The conformations of polymer chains may be generated by a Monte Carlo

    simulation method. The dimensions of linear chains, unperturbed by excluded-

    volume interactions, have been measured in various solvents by light scattering, x-

    ray small-angle scattering, and dilute-solution viscosity measurements. They are

    reported most comprehensively in the Polymer Handbook (91,92). A powerful tool

    to investigate chain conformations in unswollen and swollen melts and networksis small-angle neutron scattering (sans) (67,9397).

    The Equation of State for a Single Polymer Chain. The variation inthe Helmholtz free energy is the negative of the work of deformation in isothermal

    elongation

    dA = dW= fdr (25)

    with

    dA = dU TdS (26)

    The tensile force f on a polymer chain for a given length r is

    f =

    A

    r

    T

    =

    U

    r

    T

    T

    S

    r

    T

    (27)

    In freely jointed and freely rotating model chains, no rotation is preferred;

    therefore, the internal energy is the same for all the conformations. Then,

    f = T

    S

    r

    T

    (28)

    The entropy is given by the Boltzmann relation

    S= k ln (29)

    where k is the Boltzmann constant, and , the total number of configurations

    available to the system, is proportional to W(r) (eq. (13)). Calculation of the force

    gives

    f = 2kTb2r (30)

  • 7/28/2019 Elasticity, Rubber Like

    14/37

    Vol. 2 ELASTICITY, RUBBER-LIKE 229

    f = 2kTr/r2 (31)

    Thus,f is directly proportional to the absolute temperature and to r, which means

    the chain acts as a Hookean spring with modulus 3kT/r2

    . The use of a non-Gaussian distribution for leads to (71,85,86)

    f = (kT/l)L 1(r/nl) (32)

    Classical Thermodynamics of Rubber-like Elasticity

    In experiments concerning the relationships between length, temperature, and

    force, usually the change in force with temperature at constant length is recorded

    (53,98101). It is therefore necessary to extend the thermodynamic treatment of

    the elasticity. Moreover, the force is notpurelyentropic, and theenergetic contribu-

    tion carries useful information on the dependence on temperature of the average

    end-to-end distance of the network chains in the unstrained state (21,102). It istherefore important to know how to deduce these quantities from a thermoelastic

    experiment.

    The change in internal energy during stretching an elastic body is

    dU= dQ dW (33)

    where dQ is the element of heat absorbed by the system and dW the element of

    work done by the system on the surroundings. In a reversible process,

    dQ = TdS (34)

    where S is the entropy of the body. The work dW can be expressed as the sum

    dW= pdV+ fdL (35)

    where p is the equilibrium external pressure, dV the volume dilation accompany-

    ing the elongation of the elastomer, and f the equilibrium tension. Thus,

    dU= TdS pdV+ fdL (36)

    At constant pressure, the enthalpy change is

    dH= dU+ pdV= TdS+ fdL (37)

    A deformation dL at constant pressure and temperature induces a retractive force

    f =

    H

    L

    T,p

    T

    S

    L

    T,p

    (38)

    Expression 38 is one of the forms of the thermodynamic elastic equation of state.

    Measurements of stress at constant length as a function of temperature have been

  • 7/28/2019 Elasticity, Rubber Like

    15/37

    230 ELASTICITY, RUBBER-LIKE Vol. 2

    20 40 60 800

    0.1

    0.2

    0.3

    0.4

    0.5

    Stress,

    N/mm

    2

    60

    50

    40

    30

    20

    15

    10

    53

    1

    Temperature, C

    Fig. 10. Stresstemperature curves for sulfur-vulcanized natural rubber (99,103). Cour-tesy of the American Chemical Society.

    performed (see Fig. 10) (97,103105). All the curves appear to be straight lines;

    the slope increases with increasing elongation, but for very small elongations, the

    slopes can be negative. This phenomenon, called the thermoelastic inversion, is

    due to the volume expansion occurring in any elastomer. The condition of constant

    length does not correspond to constant elongation as well, since the samples un-

    strained reference length changes with temperature. The inversion is suppressed

    by correction of the original length (11).

    The energetic and entropic contributions to the force in the intramolecular

    process of stretching the chains can be obtained in experiments where there is no

    other energetic contribution resulting from changes in (intermolecular) van der

    Waals forces. Therefore, these experiments must be performed at constant volume.

    A basic postulate of the elasticity of amorphous polymer networks is that the stress

    exhibited by a strained polymer network is assumed to be entirely intramolecular

    in origin. That is, intermolecular interactions play no role in deformations at

    constant volume and composition.

    An equation similar to equation (38) is obtained for the elastic force measuredat constant volume:

    f =

    U

    L

    T,V

    T

    S

    L

    T,V

    (39)

  • 7/28/2019 Elasticity, Rubber Like

    16/37

    Vol. 2 ELASTICITY, RUBBER-LIKE 231

    The variation in the Helmholtz free energy has the following expression:

    dA = SdT pdV+ fdL (40)

    The second derivative obtained by differentiating ( AL

    )V,T with respect to T at

    constant V and L is identical with that obtained by differentiating ( A T

    )V,L with

    respect to L at constant V and T:

    f

    T

    V,L

    =

    S

    L

    V,T

    (41)

    Thus, equation (39) can be written as

    f =

    U

    L

    T,V

    + T

    f

    T

    V,L

    (42)

    The energetic and entropic components of the elastic force, fe andfs, respectively,are obtained from thermoelastic experiments using the following equations:

    fe =

    U

    L

    T,V

    = f T

    f

    T

    V,L

    (43)

    fs = T

    S

    L

    T,V

    = T

    f

    T

    V,L

    (44)

    An example of thermoelastic data is given in Figure 11 (99). The change in

    entropy with elongation up to 350% is responsible for more than 90% of the total

    stress at room temperature, whereas the contribution of internal energy is lessthan 10%. Thus, the restoring force is due almost entirely to the tendency of the

    extended rubber molecules to return to their unperturbed, random conformational

    states. Above 350%, crystallization appears. This is a specific feature of stereo-

    regular rubbers, such as natural rubber which is capable of crystallization. Data

    concerning most of thepolymers studied in this mannerare reviewed in References

    21 and 106.

    Statistical Treatment of Rubber-like Elasticity

    A network is an ensemble of macromolecules linked together, each of them re-

    arranging its configurations by Brownian motion. Classical thermodynamics ex-

    plains the behavior of elastomers with regard to force, temperature, pressure, andvolume, but does not give the relationship between the molecular structure of the

    network and elastic quantities such as the moduli. Therefore, statistical mechan-

    ics was introduced in the 1940s (16,86,87,107109), and its theoretical predictions

    were tested (110112). Because of the complexity of network structures, two mod-

    els based on affine and phantom networks were studied. The cross-linking points

  • 7/28/2019 Elasticity, Rubber Like

    17/37

    232 ELASTICITY, RUBBER-LIKE Vol. 2

    Elongation, %

    0

    0

    0.2

    0.2

    0.4

    0.6

    0.8

    1.0

    1.2

    1.4

    1.6

    1.8

    2.0

    2.2

    2.4

    50 100 150 200 250 300 350 400

    Stress,

    N/mm

    2f

    fe

    fs

    Fig. 11. Plot of the energetic (fe) and entropic (fs) contributions to the stress at 20C

    for sulfur-vulcanized rubber (99). N/mm2 = MPa; to convert MPa to psi, multiply by 145.Courtesy of the American Chemical Society.

    of the affine network are fixed, whereas those of the phantom network can un-

    dergo fluctuations independent of their immediate surroundings. The effects of

    the macroscopic strain are transmitted to the chains through these junctions at

    which the chains are multiply joined, with the internal state of the network system

    specified in terms of the positions of the cross-linkages. Deformation transforms

    the arrangements of these points. The system is represented by the vectors ri,

    each of which connects the two ends of a chain. For a given end-to-end vector ri,

    the number of available configurations is directly proportional to the probability

    W(ri) or relative number of configurations. Representation of W(r) by a Gaussian

    function according to equation (13) fails as r approaches the maximum extension

    rmax. Experimental evidence of deviations from the Gaussian theory have been

    reported (24,113116) and non-Gaussian theories based on an expression for W(r)

    similar to equation (16) have been developed (84,85,117119). These theories have

    the disadvantage of containing parameters that can be determined only by com-parisons between theory and experiments, specifically , the number density of

    chains in the network, and n, the number of statistical links.

    The tensile force is expressed by

    f/A0 = (1/3)vkTn1/2L

    1(n 1/2) 3/2L 1( 1/2n 1/2)

    (45)

  • 7/28/2019 Elasticity, Rubber Like

    18/37

    Vol. 2 ELASTICITY, RUBBER-LIKE 233

    where

    L 1(t) = 3t + (9/5)t3 + (297/175)t5 +

    and A0 is the cross-sectional area of the unstrained sample. The number n of ran-

    dom links may be obtained by birefringence and stressstrain measurements, and

    from this result, an estimation of the number of monomer units in the equivalent-

    random link (of the KuhnGrun theory) of the polymeric chain (120). Another

    approach to a non-Gaussian theory utilizes information provided by rotational iso-

    meric state theory on the spatial configurations of chain molecules (83), including

    most of those used in elastomeric networks. Specifically, Monte Carlo calculations

    (121123) based on the rotational isomeric state approximation are used to sim-

    ulate spatial configurations followed by distribution functions for the end-to-end

    separation r of the network chains (124).

    The theory in the Gaussian limit has been refined greatly to take into ac-

    count the possible fluctuations of the junction points. In these approaches, the

    probability of an internal state of the system is the product of the probabilitiesW(ri) for each chain. The entropy is deduced by the Boltzmann equation, and the

    free energy by equation (26). The three main assumptions introduced in the treat-

    ment of elasticity of rubber-like materials are that the intermolecular interactions

    between chains are independent of the configurations of these chains and thus of

    the extent of deformation (125,126); the chains are Gaussian, freely jointed, and

    volumeless; and the total number of configurations of an isotropic network is the

    product of the number of configurations of the individual chains.

    Affine Networks. Diffusion of the junctions about their mean positionsmay be severely restricted by neighboring chains sharing the same region of

    space. The extreme case is the affine network where fluctuations are completely

    suppressed, and the instantaneous distribution of chain vectors is affine in the

    strain. The elastic free energy of deformation is then given by

    Ael(aff) = (1/2)vkT(I1 3) (v )kTln(V/V0) (46)

    where I1 is the first invariant of the tensor of deformation

    I1 = 2x +

    2y +

    2z (47)

    The quantities x, y, and z are the principal extension ratios, which specify the

    strain relative to an isotropic state of reference having volume V0; V is the volume

    of the deformed specimen, and is the number of linear chains whose ends are

    joined to multifunctional junctions of any functionality > 2. (The functionality

    of a junction is defined as the number of chains connected to it.) The cycle rank (102) represents the number of chains that have to be cut to reduce the network

    to an acyclic structure or tree (127). The cycle rank is the difference between the

    number of effective chains and effective junctions of functionality > 2:

    = (48)

  • 7/28/2019 Elasticity, Rubber Like

    19/37

    234 ELASTICITY, RUBBER-LIKE Vol. 2

    Isotropic state Deformed state

    V

    L

    V0

    L0

    Fig. 12. Simple extension of an unswollen network prepared in the undiluted state (99).Courtesy of the American Chemical Society.

    For a perfect network of functionality ,

    = (1 2/) (49)

    Simple Extension. The most general case is an extension with change insample volume, as illustrated in Figure 12. The strain along the direction of

    stretching is given by x as

    = x =L/L0 (50)

    Since the volume changes from V0 to V,

    xyz = V/ V0 (51)

    Because there is symmetry about the x axis, y = z. Combining equations (50)

    and (51) leads to

    y = z = (V/V0)1/2

    (52)

    The force f is the derivative of the free energy with respect to length; specifically,

    f =

    A

    L

    T,V

    = (1/L0)

    A

    T,V

    (53)

    Using equation (46),

    f = (vkT/L0)

    V/

    V02

    (54)

    It is also possible to deduce the expression for the force as a function of the ex-

    tension measured relative to the length Lvi of the unstretched sample when its

    volume is fixed at the same volume V as occurs in the stretched state:

    =L/Liv = L0/Liv (55)

    and

    L0 =Liv(V/V0)

    1/3 (56)

  • 7/28/2019 Elasticity, Rubber Like

    20/37

    Vol. 2 ELASTICITY, RUBBER-LIKE 235

    Equation (54) is then transformed to

    f =

    vkT/Liv

    (V/V0)

    2/3( 1/2). (57)

    Energetic Contributions. Equation (57) can be used for the molecular in-terpretation of the ratio fe/f. Thus, equation (43) can be rewritten as

    fe = f T

    f

    T

    V,L

    = f T

    ln(f/T)

    T

    V,L

    (58)

    or

    fe

    f= T

    ln(f/T)

    T

    V,L

    (59)

    The derivative of f at constant volume and length, thus at constant , can be

    obtained from equation (57), with the result

    fe/f = (2/3)T(dlnV0/dT) = T

    dln r20

    dT

    (60)

    since V0 is the volume of the isotropic state so defined that the mean square of

    the magnitude of the chain vectors equals r20, the value for the free unperturbed

    chains.

    The intramolecular energy changes arising from transitions of chains from

    one spatial configuration to another are, by equation (60), directly related to

    the temperature coefficient of the unperturbed dimensions. It is interesting to

    compare the thermoelastic results for polyethylene (128), fe/f = 0.45, and

    poly(dimethylsiloxane) (129), fe/f = 0.25. The preferred (lowest energy) confor-

    mation of the polyethylene chain is the all-trans form, since gauche states at rota-

    tional angles of120 cause steric repulsions between CH2 groups (83). Since thisconformation has the highest possible spatial extension, stretching a polyethylene

    chain requires switching some of the gauche states, which are, of course, present in

    the randomly coiled form, to the alternative trans states (106,130). These changes

    decrease the conformational energy and are the origin of the negative type of ide-

    ality represented in the experimental value of fe/f. (This physical picture also

    explains the decrease in unperturbed dimensions upon increase in temperature.

    The additional energy causes an increase in the number of the higher energy

    gauche states, which are more compact than the trans ones.) The opposite be-

    havior is observed in the case of poly(dimethylsiloxane) (26). The all-trans form

    is again the preferred conformation; the relatively long Si O bonds and the un-

    usually large Si O Si bond angles reduce steric repulsion in general, and the

    trans conformation places CH3 side groups at distances of separation where they

    are strongly attractive (83,129,131). Because of the inequality of the Si O Siand O Si O bond angles, however, this conformation is of very low spatial ex-

    tension. Stretching a poly(dimethylsiloxane) chain therefore requires an increase

    in the number of gauche states. Since these are of higher energy, this explains

    the fact that deviations from ideality for these networks are found to be positive

    (106,129,130).

  • 7/28/2019 Elasticity, Rubber Like

    21/37

    236 ELASTICITY, RUBBER-LIKE Vol. 2

    Isotropic drystate 1

    Isotropic swollenstate 2

    Deformed swollenstate 3

    V

    L

    V

    L0

    V0

    Li

    Fig. 13. Simple extension of a swollen network.

    Simple Extension of Swollen Networks. The force required to deform anelastomeric sample, from state 2 to state 3 in Figure 13, is given by equation (57),

    in which =L/Liv is the extension ratio for the isotropic swollen state relative to

    the deformed swollen state. The ratio V0/V is the volume fraction v2 of polymer

    in the swollen system, and A0 is the cross-sectional area of the dry sample, which

    is generally measured before the experiment. The area of the swollen sample is

    given by

    Aiv = A0(V/ V0)2/3

    (61)

    Equation (56), of course, still holds, and L0A0 = V0. In an experiment, the force f

    is measured as a function of elongation. Under these conditions,

    f/A0 = (kT/L0A0)v 1/3

    2 ( 1/2) (62)

    Hence, the quantity [ f]v1/3

    2 /A0( 1/2) should be a constant, independent of

    the degree of swelling, and be equal to vkT/V0. As shown later [f] is commonly

    plotted versus 1/ to determine deviations from theory (112,132,133).

    Simple Extension of Networks Cross-linked in the Diluted State. The

    polymer is dissolved, cross-linked, and dried. The stressstrain measurementsare carried out on the dry network, as illustrated in Figure 14. Equation (57)

    also holds for this case. If Aiv is the cross-sectional area of the undeformed dry

    specimen, the force per unit undeformed area is

    f/Aiv = (kT/ V)(V/V0)2/3

    ( 1/2) (63)

    Reference statefor cross-linking

    process

    Dry, undeformedstate

    Dry, deformedstate

    L0

    V0 V V

    LLi

    Fig. 14. Simple extension of an unswollen network cross-linked in the diluted state(V0 > V).

  • 7/28/2019 Elasticity, Rubber Like

    22/37

    Vol. 2 ELASTICITY, RUBBER-LIKE 237

    which is (V/V0)2/3 times the tension of a network cross-linked in the dry state

    (111,134).The Affine Shear Modulus. For V = V0, equation (54) becomes

    f/A0 = (kT/V0)( 1/2) (64)

    Making use of equation (2) gives = 1 + . For small values of , 1, equation

    (64) may be written as

    f/A0 = 3(kT/ V0) (65)

    The tensile modulus Eaff is three times the shear modulus Gaff, since Poissons

    ratio for elastomers is close to 0.5. Specifically,

    Eaff= 3vkT/ V0 = 3Gaff (66)

    and hence,

    Gaff = vkT/ V0 (67)

    Ifv/V0 is the molar number density of chains [equal to /Mc for a perfect network,

    where is the elastomer density and Mc is the molecular weight between cross-

    link points (in g/mol)], then

    Gaff= vRT/V0 (68)

    where R is the gas constant. In the remaining material, v is a molar quantity

    unless it is followed by the Boltzmann constant k.For an affine perfect network,

    Gaff = RT/Mc (69)

    As a numerical example, the affine shear modulus of a perfect poly

    (dimethylsiloxane) tetrafunctional network of density = 0.97 g/cm3 and Mc =

    11,300 g/mol is Gaff = 0.212 106 N/m2. However, imperfections such as chain

    ends exist in typical networks, as illustrated in Figure 6. The following correc-

    tion can be made to account for this circumstance (16). Before the cross-linking

    reaction, it is assumed that vm chains of length Mn are present in the melt, with

    vm = /Mn. Each chain has two ends, thus there are 2/Mn chain ends. After the

    cross-linking process, there are v0 chains of length Mc, v0 = /Mc, along with the2/Mn chain ends. Therefore the number of chains that are elastically effective is

    v = /Mc 2/Mn = (/Mc)(1 2Mc/Mn) (70)

    This correction is small for Mc Mn, which is frequently the case.

  • 7/28/2019 Elasticity, Rubber Like

    23/37

    238 ELASTICITY, RUBBER-LIKE Vol. 2

    Phantom Networks. In this idealized model, chains may move freelythrough one another (135). Junctions fluctuate around their mean positions be-

    cause of Brownian motion, and these fluctuations are independent of deforma-tion. The mean square fluctuations r2 of the end-to-end distance r are related

    to the mean square of the end-to-end separation of the free unperturbed chainsr2(102,136,137) by

    r2/r20 = 2/ (71)

    The fluctuation range (r2 = r20/2 for a tetrafunctional network) is generally quite

    large and of considerable importance. The instantaneous distribution of chain

    vectors ris not affine in the strain because it is the convolution of the distribution

    of the affine mean vector rwith the distribution of the fluctuations r, which are

    independent of the strain. The elastic free energy of such a network is

    Ael(ph) = (1/2)kT(I1 3) = (1/2)( )kT(I1 3) (72)

    In simple extension, the equivalent of equation (62) is now

    f/A0 = ( )kTV 1

    0 v 1/3

    2 ( 1/2) (73)

    and the shear modulus is given by

    Gph = (v )RT/ V0 (74)

    For a perfect network of functionality , equation (49) holds, and therefore,

    Gph = (1 2/)vRT/ V0 (75)

    Gph = (1 2/)RT/Mc (76)

    The relationship between affine and phantom moduli is then

    Gph = (1 2/)Gaff (77)

    For example, Gaff is twice Gph for a perfect tetrafunctional network.

    Comparisons With Experimental Results. Stressstrain measure-ments in uniaxial extension can be compared with the prediction of an affine

    Gaussian network (eq. (64)), as illustrated in Figure 15 (113). The Gaussian re-

    lationship in the affine limit is valid only at small deformations. The best fit is

    obtained using Gaff = 0.39 MN/m2 (56.6 psi) (7); deviations occur when > 1.5.The experimental curve may nevertheless be well represented by adjusting the

    parameters of the non-Gaussian stressstrain relationship (eq. (45)) (84,117,138).

    These disagreements between experiments and the simple predictions of statis-

    tical mechanics have led some workers to develop a phenomenological theory of

    rubber-like elasticity.

  • 7/28/2019 Elasticity, Rubber Like

    24/37

    Vol. 2 ELASTICITY, RUBBER-LIKE 239

    Extension ratio

    Tensileforceperunitunstrainedarea,

    N/

    mm

    2

    Theoretical

    1 2 3 4 5 6 7 80.0

    1.0

    2.0

    3.0

    4.0

    5.0

    6.0

    7.0

    Fig. 15. Stressstrain isotherms in simple elongation; comparison of experimental curve(open circles) with theoretical prediction (eq. (64)) (113). To convert N/mm 2 to psi, multiplyby 145. Courtesy of The Royal Society of Chemistry.

    Phenomenological Theory. Continuum mechanics is used to describemathematically the stressstrain relations of elastomers over a wide range in

    strain. This phenomenological treatment is not based on molecular concepts, but

    on representations of observed behavior (132,139143). The main goal is to find an

    expression for the elastic energy W stored in the system (assumed to be perfectly

    elastic, isotropic in its undeformed state, and incompressible), analogous to thefree energy of the statistical treatment. The condition of isotropy in the unstrained

    state requires that Wbe symmetrical with respect to the three principal extension

    ratios x, y, z. A rotation of the material through 180, ie, a change of sign of

    two of the i (i = x, y, z), does not alter W (144). The three simplest even-powered

    functions satisfying these conditions are the strain invariantsI1,I2, andI3 defined

    as

    I1 = 2x +

    2y +

    2z

    I2 = 2x

    2y +

    2y

    2z +

    2z

    2x

    I3 = 2x

    2y

    2z

    (78)

    The most general form of the strainenergy function, which vanishes at zero

    strain, for an isotropic material is the power series

  • 7/28/2019 Elasticity, Rubber Like

    25/37

    240 ELASTICITY, RUBBER-LIKE Vol. 2

    W=

    i,j,k = 0

    Ci jk(I1 3)i(I2 3)

    j(I3 1)

    k(79)

    In equation (79), the experimental results obtained at sufficiently small strainsare well represented with two nonzero terms, C100 and C010:

    W= C100(I1 3) + C010(I2 3) (80)

    In uniaxial extension, the so-called MooneyRivlin equation is obtained (17):

    f1/32 /A0 = 2(C1 + C2/)( 1/

    2) (81)

    A convenient and standard way to treat stressstrain data is to plot the reduced

    force

    [f]f1/3

    2 /( 1/2) (82)

    versus 1/, where f is the nominal stress f/A0 and 2 the volume fraction of

    polymer in the network, if swollen (Fig. 16).In this scheme, the affine and phantom networks are represented by hor-

    izontal lines (eqs. (62) and (73)) from equations [f] = vRT/V0 and [f] =

    (v )RT/V0, respectively. It has been reported that C2 decreases as 2 decreases,

    whereas C1 is approximately constant, as illustrated in Figure 16 (112). In view

    0.20

    0.29

    0.40

    0.55

    0.74

    0.4 0.6 0.8 1.00.125

    0.150

    0.175

    0.200

    0.225

    1/

    [f],N/mm

    2

    2 1.00

    Fig. 16. Plot of [f] versus 1/ for a natural rubber vulcanizate swollen in benzene todemonstrate the influence of v2 on C2 (114). To convert N/mm

    2 to psi, multiply by 145.Courtesy of The Royal Society of Chemistry.

  • 7/28/2019 Elasticity, Rubber Like

    26/37

    Vol. 2 ELASTICITY, RUBBER-LIKE 241

    of experimental results on swollen networks (69,114,145,146), the following form

    of the strain energy per unit volume of dry network is to be preferred:

    W/ V0 = C1(I1 3)+C2(I2I 1+m/2

    3 3) (83)

    The reduced force can be deduced (145) as [ f] = 2C1+2C2v(4/3) m

    2 1. The pa-

    rameter C1 of the swollen network is equal to C1 of the dry network, whereas

    C2 (swollen) is equal to C2 (dry) v(4/3) m

    2 . Experimentally, m was found to be 0 or12

    (145). The constant C1 increases with the cross-linking density of the rubber

    (114,147), and 2C1 + 2C2 is approximately the shear modulus at small defor-

    mation (148). Although there have been many attempts (149155), a molecular

    explanation of C1 and C2 has been achieved only recently, as is described later.

    The experimental error range is of great importance. A 1% error in the de-

    termination of has a tremendous effect on the Mooney curve when 1/ > 0.9;

    this part of the curve is therefore highly unreliable (156).

    Statistical Theory of Real Networks. Affine and phantom networks areextreme limits. Stressstrain measurements in uniaxial extension have revealed

    that the behavior of real networks is between these limits. A theoretical attempt

    has been made to account for this dependence of [f] on 1/ in terms of a gradual

    transition between the affine and phantom deformations (102,157), and a molec-

    ular theory has been formulated (102,158161). In this model, the restrictions of

    junction fluctuations due to neighboring chains are represented by domains of con-

    straints. At small deformations, the stress is enhanced relative to that exhibited

    by a phantom network. At large strains, or high dilation, the effects of restrictions

    on fluctuations vanish and the relationship of stress to strain converges to that for

    a phantom network. In a later theoretical refinement, the behavior of the network

    is taken to depend on two parameters. The most important is which measures

    the severity of entanglement constraints relative to those imposed by the phantom

    network. Another parameter takes into account the nonaffine transformation ofthe domains of constraints with strain. Topological and mathematical treatment

    leads to the expression

    [f] = fph(1 + fc/fph) (84)

    where fph is the usual phantom modulus and fc/fph is the ratio of the force due to

    entanglement constraints to that for the phantom network. A specific expression

    for fc/fph in uniaxial extension is

    fc/fph = (/)

    K

    21

    2K

    22

    /( 1/2) (85)

    where 1 = v 1/3

    2 ,2 = 1/2v

    1/3

    2 , and / is the ratio of the number of effective

    junctions to the cycle rank. For a perfect network, /=2/( 2). The function

    K(x2) is given by

    K(x2) =B[.

    B (B + 1) 1

    +g(.

    g B +g.

    B )(gB + 1) 1] (86)

  • 7/28/2019 Elasticity, Rubber Like

    27/37

    242 ELASTICITY, RUBBER-LIKE Vol. 2

    with

    g =x2[1/ + (x 1)].

    g = 1/ (1 3x/2)B = (x 1)(1 +x x2)/(1 +g)

    2

    .B =B

    (2x(x 1))

    1 2

    .g (1 +g)

    1+ (1 2x)[2x(1 +x x2)]

    1

    It is interesting to note that junction fluctuations increase in the direction of

    stretching but decrease in the direction perpendicular to it. Therefore the modu-

    lus decreases in the direction of stretching, but increases in the normal direc-

    tion since the state of the network probed in this direction tends to be more

    nearly affine. The curve of [f] versus 1/ is sigmoidal. The parameters and of

    poly(dimethylsiloxane) networks are determined in Figure 17 (155); the intercept

    of the sigmoidal curves is the phantom modulus. This FloryErman theory has

    been compared successfully with such experiments in elongation and compression

    (155,162,162166). It has not yet been extended to take account of limited chain

    extensibility or strain-induced crystallization (167).

    0.20

    0.12

    0.08

    0.04

    0 0.2 0.4 0.6 0.8 1

    Mn 4000 0.1 0

    Mn 9500 3.4 0 Mn 18500 30 0.01

    Mn 25600 24.5 0

    Mn 32900 16.2 0

    31.5 0.01

    27.5 0 23 0

    21.5 0

    1/

    [f*],N/mm2

    Fig. 17. Determination of the parameters and of the FloryErman theory for perfecttrifunctional poly(dimethylsiloxane) networks (155). To convert N/mm2 to psi, multiply by145. Courtesy of Springer-Verlag.

  • 7/28/2019 Elasticity, Rubber Like

    28/37

    Vol. 2 ELASTICITY, RUBBER-LIKE 243

    Predictions for the Parameters and . The parameter is not far fromzero, which is to be expected since the surroundings of junctions cause their defor-

    mation to be nearly affine with the macroscopic strain. The primary parameter is defined as the ratio of the mean-square junction fluctuations in the equivalent

    phantom network, ie, in the absence of constraints, to the mean-square junction

    fluctuations about the centers of domains of entanglement constraints (in the

    absence of the network) in the isotropic state. Thus in a phantom network, the

    absence of constraints leads to = 0. In an affine one, the complete suppression

    of fluctuations is equivalent to = . It has been proposed that should be pro-

    portional to the degree of interpenetration of chains and junctions (165). Since an

    increasing number of junctions in a volume pervaded by a chain leads to stronger

    constraints on these junctions, was taken to be

    =I

    r203/2

    (/V0) (87)

    where I is a constant of proportionality, (r20

    )3/2

    is assumed to be proportional to

    the volume occupied by a chain, and /V0 is the number of junctions per unit

    volume. The mean-square unperturbed dimension r20 can be taken proportional to

    Mc (95), the molecular weight between cross-links, and /V0 to M 1

    c for a perfect

    tetrafunctional network; therefore M1/2c (93,155,165).

    Swelling Equilibrium. The isotropic swelling of a cross-linked elastomerby a liquid has two important opposing effects: the increase in mixing entropy of

    the system because of the presence of the small molecules, and the decrease in con-

    figurational entropy of the network chains by dilation. Therefore, an equilibrium

    degree of swelling is established, which increases as the cross-linking density de-

    creases (168170). The free energy change A for this process is usually assumed

    to be separable into the free energy of mixing, Am, and the elastic free energy

    Ael,

    A= Am + Ael (88)

    although questions have been raised with regard to this separability (171,172).

    The contribution Am has been calculated with the help of a lattice model (173,

    174). The other contribution Ael is given in the later FloryErman theory (161)

    by

    Ael = Ael(ph) + Ac (89)

    where Ael(ph) is given by equation (72); Ac is the additional term which ac-

    counts for the constraints and is given by

    Ac = kT/2

    i=x,y,z

    {[1 +g(i)]B(i) ln[(B(i) + 1)(g(i)B(i) + 1)]}. (90)

    The i are the principal extension ratios, andg andB have been defined in equation

    (86).

  • 7/28/2019 Elasticity, Rubber Like

    29/37

    244 ELASTICITY, RUBBER-LIKE Vol. 2

    The chemical potential of the solvent in the swollen network is

    1 01 =N

    Am

    n1

    T,p

    +N

    Ael

    T,p

    n1

    T,p

    (91)

    where n1 is the number of moles of solvent, and the isotropic extension ratio is

    = x = y = z = [(n1V1 + V0)/V0]1/3

    = v 1/3

    2 (92)

    where V1 is the molar volume of the solvent and V0 the volume of the dry network.

    At swelling equilibrium, 1 = 01. Hence, the standard expression for Am (3) leads

    to

    ln(1 2m) + 2m + 22m =

    Ael

    T,p

    n1

    T,p

    (RT) 1

    (93)

    where v2m is the volume fraction of polymer at swelling equilibrium. The inter-

    action parameter 1 may be determined, for example, by vapor pressure or by

    osmometry measurements (145,175178). It depends on the concentration of poly-

    mer in the polymersolvent system (175). Using the FloryErman expression for

    the elastic energy and assuming the parameter to be independent of swelling

    (isotropic swelling does not change the relative topology of the network), equation

    (93) (with the left-hand side abbreviated as H) becomes

    H= (/ V0)V1v1/3

    2m

    1 + (/)K

    v

    2/3

    2m

    (94)

    The molecular weightMc between cross-links of a perfect network is then obtained

    by combination of equations (48),(49), and (94), with v/V0 = /Mc:

    Mcr = (2/ 1)V11/32m

    1 + (/2 1)

    1K

    V1 2/3

    2m

    H (95)

    where the subscript r is employed here for real networks.

    For a network deforming affinely, = , K(2) = 1 2, and

    Mca = V11/3

    2m

    1 2

    2/3

    2m /

    H (96)

    For a phantom network, = 0, K(2) = 0, and

    Mcp = (2/ 1)V11/3

    2m

    H (97)

    Determination of the Degree of Cross-linking by StressStrainMeasurements and by Swelling Equilibrium

    Characterization of network structures is often the main objective of theoretical

    and experimental works in the field of rubber elasticity (179183). Simple experi-

    ments such as swelling equilibrium have been extensively used. However, most of

    the experimental swelling results on cross-linked polymers have been interpreted

    using the FloryRehner expression for an affinely deforming network (6,184186).

  • 7/28/2019 Elasticity, Rubber Like

    30/37

    Vol. 2 ELASTICITY, RUBBER-LIKE 245

    The modern theory of real networks now permits a more accurate determina-

    tion of network structures through use of equations (84),(87), and (95) (187204).

    Stressstrain measurements can be analyzed as shown in Figure 17. The phantom

    modulus thus determined leads to andMc through equations (75) and (76) (189).

    Swelling equilibrium data are similarly analyzed through equation (94), with the

    parameter given by equation (87) (189).

    If Mc and have been determined previously from stressstrain measure-

    ments, then the interaction parameter 1 may be calculated through equation

    (95) (187,190).

    Entanglements

    Entanglements have been introduced in the later FloryErman theory as con-

    straints that restrict junction fluctuations. Another viewpoint considers entangle-

    ments to act as physical cross-links, being based, in part, on the observation that

    linear polymers of high molecular weight exhibit a storage modulus G(), which

    remains relatively constant over a wide range of frequencies (205). This plateaumodulus G0N is independent of chain length for long chains and is insensitive to

    temperature. Since it varies with the volume fraction of polymer in concentrated

    solutions, it could be due to pair-wise interactions between chains (20), and a uni-

    versal law has been proposed for the dependence of G0N on the chemical structure

    of the polymer (206). During the cross-linking process, some of such interactions

    or entanglements could be trapped in the network and act as physical junctions.

    This conclusion has been tested by irradiation cross-linking of already deformed

    networks (207209), and then measuring the dimensional changes.

    In the simplest phenomenological approach for rubber-like elasticity of

    trapped entanglements at small deformation (210,211), the shear modulus is

    taken to be the sum of two terms:

    G = Gc + Ge Te (98)

    where Gc is the contribution of the chemical cross-links. Taking into account the

    restrictions of junction fluctuations, as in the Flory theory, leads to

    Gc = (v h)RT (99)

    in which the empirical parameter h was introduced (153). Its value, between 0

    (affine) and 1 (phantom), characterizes the nature of the networks at small de-

    formation; h can also be expressed as a function of the Flory parameters and

    (155,212). The additional contribution GeTe is said to arise from permanently

    trapped (interchain) entanglements in the network. The modulus Ge is thought to

    have a value close to G0N, and Te, the trapped entanglement factor, is the proba-

    bility that all of the four directions from two randomly chosen points in the system,

    which may potentially contribute an entanglement, lead to the gel fraction.The idea of entanglements acting as physical junctions was originally de-

    veloped in the literature to explain deviations from the predictions for affine

    networks (185,213). The possibility of a contribution at equilibrium caused by

    trapped entanglements has been tested with model networks, ie, those prepared

    in such a way that the number and functionality of the cross-links are known. A

  • 7/28/2019 Elasticity, Rubber Like

    31/37

    246 ELASTICITY, RUBBER-LIKE Vol. 2

    typical, highly specific reaction used for this purpose is the end-linking of function-

    ally terminated polymer chains. Specific examples would be hydroxyl-terminated

    or vinyl-terminated chains of poly(dimethylsiloxane), [Si(CH3)2O]x , end-linked

    with an organic silicate or silane (214,215). A considerable body of published data

    has, in fact, been interpreted as proving the existence of a trapped-entanglement

    contribution in this kind of network (216222). Typically, the network charac-

    teristics, eg, number of chains and junctions, extent of cross-linking reaction p,

    trapped entanglement factor Te, and effective functionality e, were calculated by

    the branching theory after measurement of the network sol fraction. The main

    assumption of this probabilistic method is that the sol fraction, after subtraction

    of the amount of nonreactive species as determined by gel permeation chromatog-

    raphy analysis (218), is composed only of primary reactive chains (223). On the

    contrary, however, the sol fraction is a complicated mixture of reactive, unreactive,

    reacted, and unreacted molecules. For example, simulation of nonlinear polymer-

    izations has shown that about half of the sol molecules are (reacted) cyclics (224).

    Their presence indicates that the value of the extent of the cross-linking reac-

    tion, formerly calculated with the assumption that the sol fraction is composed ofnonreacted reagents, has to be significantly increased. As a result, many model

    networks can be considered as nearly perfect. This casts some doubt on the results

    interpreted as showing an entanglement contribution at equilibrium. If these net-

    works are considered as perfect, such a contribution does not seem to be important

    (155,225). Entanglement contributions have been reported for polybutadiene (153)

    and ethylenepropylene copolymer (154) networks prepared by radiation-induced

    cross-linking. Again some doubt exists on the method used to characterize the

    network structure.

    Molecular models treating entanglements as interstrand links that are free

    to slip along the strand contours have been developed (226228) and tube mod-

    els have been investigated (229,230). These approaches have been reviewed in

    Reference 27.

    The question of entanglements is still controversial. It has been postulated(160) that an entanglement cannot be equivalent to a chemical cross-link. Con-

    tacts between a pair of entangled chains are transitory and of short duration owing

    to the diffusion of segments and associated time-dependent changes of configura-

    tions. Such trapped entanglements as previously described are possibly of minor

    importance in equilibrium stress measurements.

    BIBLIOGRAPHY

    Elasticity, Rubber-like in EPSE 2nd ed., Vol. 5, pp. 365408, by J. P. Queslel and J. E.

    Mark, University of Cincinnati.

    1. D. R. Lide, ed., Handbook of Chemistry and Physics, 62nd ed., CRC Press, Cleveland,

    Ohio, 19811982, p. F-91.2. J. Gough, Mem. Lit. Phil. Soc. Manchester 2nd Ser. 1, 288 (1805).

    3. P. J. Flory, Principles of Polymer Chemistry, Cornell University Press, Ithaca, N.Y.,

    1953, p. 434.

    4. Lord Kelvin, Q. J. Math. 1, 57 (1857).

    5. J. P. Joule, Philos. Trans. R. Soc. London, Ser. A 149, 91 (1859).

    6. Ref. 3, Chapt. XI.

  • 7/28/2019 Elasticity, Rubber Like

    32/37

    Vol. 2 ELASTICITY, RUBBER-LIKE 247

    7. L. R. G. Treloar, The Physics of Rubber Elasticity, 3rd ed., Clarendon Press, Oxford,

    U.K., 1975.

    8. A. V. Tobolsky, Properties and Structure of Polymers, John Wiley & Sons, Inc., New

    York, 1960.

    9. F. Bueche, Physical Properties of Polymers, Wiley-Interscience, New York, 1962, pp.37, 225.

    10. A. R. Payne and J. R. Scott, Engineering Design with Rubber, MacLaren and Sons,

    London, 1960, p. 7.

    11. D. J. Williams, Polymer Science and Engineering, Prentice-Hall, Englewood Cliffs,

    N.J., 1971, Chapt. 9.

    12. A. N. Gent, B. Erman, and J. E. Mark, in J. E. Mark, B. Erman, and F. R. Eirich, eds.,

    Science and Technology of Rubber, 2nd ed., Academic Press, Inc., New York, 1994, pp.

    1, 189.

    13. J. J. Aklonis and W. J. MacKnight, Introduction to Polymer Viscoelasticity, 2nd ed.,

    John Wiley & Sons, Inc., New York, 1983, p. 102.

    14. J. W. S. Hearle, Polymers and Their Properties, Vol. 1, John Wiley & Sons, Inc., New

    York, 1982, p. 153.

    15. P. Meares,Polymers: Structure and Bulk Properties, Van Nostrand Reinhold Co., Inc.,

    New York, 1965, p. 160.16. P. J. Flory, Chem. Rev. 35, 51 (1944).

    17. W. R. Krigbaum and R.-J. Roe, Rubber Chem. Technol. 38, 1039 (1965).

    18. K. Duek and W. Prins, Adv. Polym. Sci. 6, 1 (1969).

    19. K. J. Smith Jr. and R. J. Gaylord, Rubber Age 107(6), 31 (1975).

    20. W. W. Graessley, Adv. Polym. Sci. 16, 1 (1974).

    21. J. E. Mark, J. Polym. Sci. Macromol. Rev. 11, 135 (1976).

    22. R. G. C. Arridge, Adv. Polym. Sci. 46, 67 (1982).

    23. K. Duek, Adv. Polym. Sci. 44, 164 (1982).

    24. L. R. G. Treloar, Rep. Prog. Phys. 36, 755 (1973).

    25. J. E. Mark, J. Educ. Mod. Mater. Sci. Eng. 4, 733 (1982).

    26. J. E. Mark, J. Chem. Educ. 58, 898 (1981).

    27. B. E. Eichinger, Ann. Rev. Phys. Chem. 34, 359 (1983).

    28. M. C. Boyce and E. M. Arruda, Rubber Chem. Technol. 73, 504 (2000).

    29. K. J. Smith Jr., Comprehensive Polymer Science, 2nd supplement, Elsevier, Oxford,1996.

    30. O. H. Yeoh, Comprehensive Polymer Science, 2nd supplement, Elsevier, Oxford,

    1996.

    31. A. N. Gent, in A. N. Gent, ed., Engineering with Rubber: How to Design Rubber Com-

    ponents, Hanser, Munich, 1992, p. 33.

    32. J. E. Mark, M. A. Sharaf, and B. Erman, in S. M. Aharoni, ed., Synthesis, Character-

    ization, and Theory of Polymeric Networks and Gels, Plenum, New York, 1992.

    33. A. Kloczkowski, J. E. Mark, and B. Erman, Macromolecules 28, 5089 (1995).

    34. J. E. Mark and B. Erman, in R. F. T. Stepto, ed., Polymer Networks, Blackie Academic

    Chapman and Hall, Glasgow, 1998.

    35. B. Erman and J. E. Mark,in W. Brostow, ed.,Performance of Plastics, Hanser, Munich,

    2000.

    36. J. E. Mark,in C. Carraherand C. Craver, eds.,Applied polymer Science21st Century,

    American Chemical Society, Washington, D.C., 2000.

    37. R. P. Brown, Elastomerics 115, 46 (1983).

    38. L. E. Nielsen, Mechanical Properties of Polymers and Composites, Vol. 1, Marcel

    Dekker, Inc., New York, 1974, pp. 5, 39.

    39. D. W. van Krevelen, Properties of Polymers: Correlations with Chemical Structure,

    Elsevier, Amsterdam, 1972, p. 147.

  • 7/28/2019 Elasticity, Rubber Like

    33/37

    248 ELASTICITY, RUBBER-LIKE Vol. 2

    40. F. S. Conant, in M. Morton, ed., Rubber Technology, Van Nostrand Reinhold Co., Inc.,

    New York, 1973, p. 114.

    41. P. G. Howgate, in C. Hepburn and R. J. W. Reynolds, eds., Elastomers: Criteria for

    Engineering Design, Applied Science, London, 1979, p. 127.

    42. E. R. Praulitis, I. V. F. Viney, and D. C. Wright, in Ref. 41, p. 139.43. Ref. 10, pp. 123, 225.

    44. Ref. 15, p. 179.

    45. S Standards, Rubber, Natural and SyntheticGeneral Test Methods, and Materi-

    alscos Part 09:01, American Society for Testing and Materials, Philadelphia, Pa.

    46. W. K. Moonan and N. W. Tschoegl, Macromolecules 16, 55 (1983).

    47. O. Posfalvi, Kautsch. Gummi Kunstst. 35, 940 (1982); L. M. Boiko, Ind. Lab. 49, 538

    (1983).

    48. E. Galli, Plast. Compounding 5, 14 (1982).

    49. T. L. Smith, Polym. Eng. Sci. 17, 129 (1977).

    50. R. R. Rahalkar, C. U. Yu, and J. E. Mark, Rubber Chem. Technol. 51, 45 (1978).

    51. A. L. Andrady and co-workers, J. Appl. Polym. Sci. 26, 1829 (1981).

    52. A. J. Chompff and S. Newman, eds., Polymer Networks: Structural and Mechanical

    Properties, Plenum Press, New York, 1971, pp. 111, 193.

    53. M. C. Shen, D. A. McQuarrie, and J. L. Jackson, J. Appl. Phys. 38, 791 (1967).54. Ref. 13, p. 29.

    55. W. O. Murtland, Elastomerics 114, 19 (1982).

    56. A. Whelan and K. S. Lee, eds., Development in Rubber Technology3 Thermoplastic

    Rubbers, Applied Science Publishers, London, 1982.

    57. D. S. Pearson and W. W. Graessley, Macromolecules 11, 528 (1978).

    58. P. J. Flory, Macromolecules 15, 99 (1982).

    59. L. C. Case and R. V. Wargin, Makromol. Chem. 77, 172 (1964).

    60. J. Scanlan, J. Polym. Sci. 43, 501 (1960).

    61. L. C. Case, J. Polym. Sci. 45, 397 (1960).

    62. Ref. 10, p. 15.

    63. Ref. 40, p. 7.

    64. L. K. Nash, Elements of Classical and Statistical Thermodynamics, Addison-Wesley,

    Reading, Mass., 1970.

    65. Ref. 6, p. 399.66. Ref. 13, p. 85.

    67. P. G. De Gennes, Scaling Concepts in Polymer Physics, Cornell University Press,

    Ithaca, N.Y., 1979, p. 29.

    68. P. J. Flory, Pure Appl. Chem. 52, 241 (1980).

    69. G. Gee, Polymer 7, 373 (1966).

    70. A. V. Tobolsky, D. W. Carlson, and N. Indictor, J. Polym. Sci. 54, 175 (1961).

    71. T. L. Hill, An Introduction to Statistical Thermodynamics, Addison-Wesley, Reading,

    Mass., 1962, p. 214.

    72. F. A. Bovey, Chain Structure and Conformation of Macromolecules, Academic Press,

    Inc., New York, 1982, p. 185.

    73. F. T. Wall, J. Chem. Phys. 11, 67 (1943).

    74. H. Eyring, Phys. Rev. 39, 746 (1932).

    75. P. Debye, J. Chem. Phys. 14, 636 (1946).

    76. B. H. Zimm and W. H. Stockmayer, J. Chem. Phys. 17, 1301 (1949).

    77. S. Chandrasekhar, Rev. Mod. Phys. 15, 1 (1943).

    78. W. Kuhn, Kolloid Z. 68, 2 (1934).

    79. E. Guth and H. Mark, Monatsh. Chem. 65, 93 (1934).

    80. M. V. Volkenshtein, Configurational Statistics of Polymeric Chains, Wiley-

    Interscience, New York, 1963.

  • 7/28/2019 Elasticity, Rubber Like

    34/37

    Vol. 2 ELASTICITY, RUBBER-LIKE 249

    81. Ref. 13, p. 90.

    82. Lord Rayleigh, Philos. Mag. 37, 321 (1919).

    83. P. J. Flory, Statistical Mechanics of Chain Molecules, Wiley-Interscience, New York,

    1969.

    84. M. C. Wang and E. Guth, J. Chem. Phys. 20, 1144 (1952).85. W. Kuhn and F. Grun, Kolloid Z. 101, 248 (1942).

    86. H. M. James and E. Guth, J. Chem. Phys. 11, 470 (1943).

    87. W. Kuhn, Kolloid Z. 76, 256 (1936).

    88. W. Kuhn, Kolloid Z. 87, 3 (1939).

    89. R. A. Jacobson, J. Am. Chem. Soc. 54, 1513 (1932).

    90. W. H. Hunter and G. H. Woollett, J. Am. Chem. Soc. 43, 135 (1921).

    91. J. Brandrup and E. H. Immergut, eds., Polymer Handbook, Part IV, Wiley-

    Interscience, New York, (1975), p. 1.

    92. R. Ullman in J. E. Mark and J. Lal, eds., Elastomers and Rubber Elasticity, ACS,

    Washington, D.C., (1982), p. 257.

    93. R. Ullman, Macromolecules 15, 1395 (1982).

    94. H. Benot and co-workers, J. Polym. Sci. Polym. Phys. Ed. 14, 2119 (1976).

    95. M. Beltzung and co-workers, Macromolecules 15, 1594 (1982).

    96. M. Beltzung, J. Herz, and C. Picot, Macromolecules 16, 580 (1983).97. S. Candau, J. Bastide, and M. Delsanti, Adv. Polym. Sci. 44, 27 (1982).

    98. Ref. 52, pp. 23, 47.

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

    100. K. J. Smith Jr., A. Greene, and A. Ciferri, Kolloid Z. 194, 49 (1964).

    101. U. Bianchi and E. Pedemonte, J. Polym. Sci. Part A 2, 5039 (1964); G. Allen, U.

    Bianchi, and C. Price, Trans. Faraday Soc. Part II59, 2493 (1963); J. Bashaw and K.

    J. Smith Jr., J. Polym. Sci. Part A-2 6, 1041 (1968); Yu. K. Godovsky, Polymer 22, 75

    (1981).

    102. P. J. Flory, Proc. R. Soc. London, Ser. A 351, 351 (1976).

    103. Ref. 11, p. 261.

    104. Ref. 6, p. 489.

    105. Ref. 11, p. 263.

    106. J. E. Mark, Rubber Chem. Technol. 46, 593 (1973).

    107. P. J. Flory and J. Rehner, J. Chem. Phys. 11, 512, 521 (1943).108. F. T. Wall, J. Chem. Phys. 10, 485 (1942).

    109. L. R. G. Treloar, Trans. Faraday Soc. 39, 36, 241 (1943).

    110. J. E. Mark, J. Am. Chem. Soc. 92, 7252 (1970).

    111. R. M. Johnson and J. E. Mark, Macromolecules 5, 41 (1972).

    112. C. U. Yu and J. E. Mark, Macromolecules 6, 751 (1973).

    113. L. R. G. Treloar, Trans. Faraday Soc. 40, 59 (1944).

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

    115. A. N. Gent and V. V. Vickroy, J. Polym. Sci. Part A2 5, 47 (1967).

    116. A. L. Andrady, M. A. Llorente, and J. E. Mark, J. Chem. Phys. 72, 2282 (1980).

    117. L. R. G. Treloar and G. RidingProc. R. Soc. London, Ser. A 369, 261 (1979).

    118. W. Kuhn and F. Grun, J. Polym. Sci. 1, 3 (1946).

    119. L. R. G. Treloar, Trans. Faraday Soc. 43, 277, 284 (1947).

    120. R. G. Morgan and L. R. G. Treloar, J. Polym. Sci. Part A-2 10, 51 (1972).

    121. D. Y. Yoon and P. J. Flory, J. Chem. Phys. 61, 5366 (1974).

    122. P. J. Flory and V. W. C. Chang, Macromolecules 9, 33 (1976).

    123. J. C. Conrad and P. J. Flory, Macromolecules 9, 41 (1976).

    124. J. E. Mark and J. G. Curro, in R. A. Dickie and S. S. Labana, eds., Characterization of

    Highly Cross-Linked Polymers, American Chemical Society, Washington, D.C., 1983.

    125. P. J. Flory, Pure Appl. Chem. Macromol. Chem. 33(Suppl. 8), 1 (1973).

  • 7/28/2019 Elasticity, Rubber Like

    35/37

    250 ELASTICITY, RUBBER-LIKE Vol. 2

    126. P. J. Flory, Rubber Chem. Technol. 41, 641 (1968).

    127. F. Harari, Graph Theory, Addison-Wesley, Reading, Mass., 1971.

    128. A. Ciferri, C. A. J. Hoeve, and P. J. Flory, J. Am. Chem. Soc. 83, 1015 (1961).

    129. J. E. Mark and P. J. Flory, J. Am. Chem. Soc. 86, 138 (1964).

    130. J. E. Mark, Macromol. Rev. 11, 135 (1976).131. J. E. Mark, J. Chem. Phys. 49, 1398 (1968).

    132. M. Mooney, J. Appl. Phys. 11, 582 (1940).

    133. D. S. Chiu, T.-K. Su, and J. E. Mark, Macromolecules 10, 1110 (1977).

    134. M. A. Llorente and J. E. Mark, J. Chem. Phys. 71, 682 (1979).

    135. H. M. James and E. Guth, J. Chem. Phys. 15, 669 (1947).

    136. B. E. Eichinger, Macromolecules, 5, 496 (1972).

    137. W. W. Graessley, Macromolecules 8, 865 (1975).

    138. L. R. G. Treloar, Trans. Faraday Soc. 50, 881 (1954).

    139. Ref. 7, p. 211.

    140. Ref. 13, p. 123.

    141. P. J. Blatz, S. C. Sharda, and N. W. Tschoegl, Trans. Soc. Rheol. 18, 145 (1974).

    142. R. W. Ogden, Proc. R. Soc. London, Ser. A 326, 565 (1972).

    143. K. Tobish, Colloid Poly. Sci. 257, 927 (1979).

    144. R. S. Rivlin, Philos. Trans. R. Soc., Ser. A 241, 379 (1948).145. P. J. Flory and Y. Tatara, J. Polym. Sci., Polym. Phys. Ed. 13, 683 (1975).

    146. L. Mullins, J. Appl. Polym. Sci. 2, 257 (1959).

    147. J. E. Mark, Rubber Chem. Technol. 48, 495 (1975).

    148. B. Erman, W. Wagner, and P. J. Flory, Macromolecules 13, 1554 (1980).

    149. W. J. Bobear, Rubber Chem. Technol. 40, 1560 (1967).

    150. D. R. Brown, J. Polym. Sci., Polym. Phys. Ed. 20, 1659 (1982).

    151. R. L. Carpenter, O. Kramer, and J. D. Ferry, Macromolecules 10, 117 (1977).

    152. J. D. Ferry and H. C. Kan, Rubber Chem. Technol. 51, 731 (1978).

    153. L. M. Dossin and W. W. Graessley, Macromolecules 12, 123 (1979).

    154. D. S. Pearson and W. W. Graessley, Macromolecules 13, 1001 (1980).

    155. J. P. Queslel and J. E. Mark, Adv. Polym. Sci. 65, 135 (1984).

    156. Ref. 13, p. 126.

    157. G. Ronca and G. Allegra, J. Chem. Phys. 63, 4990 (1975).

    158. P. J. Flory, J. Chem. Phys. 66, 5720 (1977).159. B. Erman and P. J. Flory, J. Chem. Phys. 68, 5363 (1978).

    160. P. J. Flory, Polymer 20, 1317 (1979).

    161. P. J. Flory and B. Erman, Macromolecules 15, 800 (1982).

    162. B. Erman and P. J. Flory, J. Polym. Sci., Polym. Phys. Ed. 16, 1115 (1978).

    163. H. Pak and P. J. Flory, J. Polym. Sci., Polym. Phys. Ed. 17, 1845 (1979).

    164. B. Erman, J. Polym. Sci., Polym. Phys. Ed. 19, 829 (1981).

    165. B. Erman and P. J. Flory, Macromolecules 15, 806 (1982).

    166. B. Erman, J. Polym. Sci., Polym. Phys. Ed. 21, 893 (1983).

    167. L. R. G. Treloar, Br. Polym. J. 14, 121 (1982).

    168. Ref. 7, p. 128.

    169. J. P. Queslel and J. E. Mark, Polym. Bull. 10, 119 (1983).

    170. G. Rehage, Rubber Chem. Technol. 39, 651 (1966).

    171. R. W. Brotzman and B. E. Eichinger, Macromolecules 15, 531 (1982).

    172. R. W. Brotzman and B. E. Eichinger, Macromolecules 16, 1131 (1983).

    173. Ref. 6, pp. 495, 541.

    174. P. J. Flory, J. Chem. Phys. 10, 51 (1942).

    175. B. E. Eichinger and P. J. Flory, Trans. Faraday Soc. 64, 2035, 2053, 2061, 2066 (1968).

    176. P. J. Flory and H. Daoust, J. Polym. Sci. 25, 429 (1957).

    177. W. R. Krigbaum and P. J. Flory, J. Am. Chem. Soc. 75, 1775 (1953).

  • 7/28/2019 Elasticity, Rubber Like

    36/37

    Vol. 2 ELASTICITY, RUBBER-LIKE 251

    178. C. H. Baker and co-workers, Polymer 3, 215 (1962).

    179. B. Erman and J. E. Mark, Structures and Properties of Rubber-like Networks, Oxford

    University Press, New York, 1997.

    180. J. E. Mark and B. Erman, Rubber-like Elasticity: A Molecular Primer, Wiley-

    Interscience, New York, 1988.181. J. E. Mark and B. Erman, Elastomeric Polymer Networks, Prentice Hall, Englewoods

    Cliffs, N.J., 1992.

    182. J. E. Mark and co-workers, Physical Properties of Polymers, American Chemical So-

    ciety, Washington, D.C., 1993.

    183. L. H. Sperling, Introduction to Physical Polymer Science, 2nd ed., Wiley-Interscience,

    New York, 1992.

    184. L. A. Woods, J. Res. Nat. Bur. Stand. Sect. A 80, 451 (1976).

    185. G. C. Moore and W. F. Watson, J. Polym. Sci. 19, 237 (1956).

    186. B. Ellis and G. N. Welding, Rubber Chem. Technol. 37, 571 (1964).

    187. B. Erman and P. J. Flory, Macromolecules 16, 1607 (1983).

    188. J. P. Queslel and J. E. Mark, Adv. Polym. Sci. 71, 229 (1985).

    189. J. P. Queslel, Rubber Chem. Technol. 62, 800 (1989).

    190. J. P. Queslel and J. E. Mark, Rubber Chem. Technol. 63, 46 (1990).

    191. J. P. Queslel and J. E. Mark, J. Chem. Phys. 82, 3449 (1985).192. B. Erman and L. Monnerie, Macromolecules 22, 3342 (1989).

    193. J. P. Queslel and J. E. Mark, Eur. Polym. J. 22, 273 (1986).

    194. J. P. Queslel and L. Monnerie, Makromol. Chem., Macromol. Symp. 30, 145 (1989).

    195. J. P. Queslel, F. Fontaine, and L. Monnerie, Polymer 29, 1086 (1988).

    196. B. Erman and J. E. Mark, Macromolecules 20, 2892 (1987).

    197. J. P. Queslel, P. Thirion, and L. Monnerie, Polymer 27, 1869 (1986).

    198. L. Y. Shy and B. E. Eichinger, Macromolecules 19, 2787 (1986).

    199. M. A. Sharat, A. Kloczkowski, and J. E. Mark, Comput. Theor. Polym. Sci. 11, 251

    (2001).

    200. J. P. Queslel and J. E. Mark, J. Chem. Educ. 64, 491 (1987).

    201. I. Bahar and B. Erman, Macromolecules 20, 1696 (1987).

    202. J. P. Queslel and J. E. Mark, Polym. J. 18, 263 (1986).

    203. M. A. Sharaf and co-workers, Comput. Polym. Sci. 2, 84 (1992).

    204. J. E. Mark and B. Erman, Comput. Polym. Sci. 5, 37 (1995).205. J. D. Ferry, Viscoelastic Properties of Polymers, 2nd ed., John Wiley & Sons, Inc., New

    York, 1970.

    206. W. W. Graessley and S. F. Edwards, Polymer 22, 1329 (1981).

    207. R. L. Carpenter, H. C. Kan, and J. D. Ferry, Polym. Eng. Sci. 19, 266 (1979).

    208. W. Batsberg and O. Kramer, J. Chem. Phys. 74, 6507 (1981).

    209. S. Granick and J. D. Ferry, Macromolecules 16, 39 (1983).

    210. N. R. Langley, Macromolecules 1, 348 (1968).

    211. N. R. Langley and K. E. Polmanteer, J. Polym. Sci., Polym. Phys. Ed. 12, 1023 (1974).

    212. M. Gottlieb, J. Chem. Phys. 77, 4783 (1982).

    213. B. M. E. van der Hoff and E. J. Buckler, J. Macromol. Sci. Chem. 1, 747 (1967).

    214. J. E. Mark, Rubber Chem. Technol. 54, 809 (1981).

    215. J. E. Mark, Adv. Polym. Sci. 44, 1 (1982).

    216. M. Gottlieb and co-workers, Macromolecules 14, 1039 (1981).

    217. M. Gottlieb, C. W. Macosko, and T. C. Lepsch,J. Polym. Sci., Polym. Phys. Ed. 19, 1603

    (1981).

    218. C. W. Macosko and G. S. Benjamin, Pure Appl. Chem. 53, 1505 (1981).

    219. K. O. Meyers, M. L. Bye, and E. W. Merrill, Macromolecules 13, 1045 (1980).

    220. K. A. Kirk and co-workers, Macromolecules 15, 1123 (1982).

    221. E. M. Valles and C. W. Macosko, Macromolecules 12, 673 (1979).

  • 7/28/2019 Elasticity, Rubber Like

    37/37

    252 ELASTICITY, RUBBER-LIKE Vol. 2

    222. E. M. Valles and C. W. Macosko, Rubber Chem. Technol. 49, 1232 (1976).

    223. D. R. Miller and C. W. Macosko, Macromolecules 9, 206 (1976).

    224. Y. K. Leung and B. E. Eichinger, Prepr. Div. Polym. Mater. 48, 440 (1983).

    225. J. P. Queslel and J. E. Mark, J. Polym. Sci., Polym. Phys. Ed. 22, 1201 (1984).

    226. W. W. Graessley and D. S. Pearson, J. Chem. Phys. 66, 3363 (1977).227. W. W. Graessley, Adv. Polym. Sci. 47, 67 (1982).

    228. R. C. Ball and co-workers, Polymer 22, 1011 (1981).

    229. R. J. Gaylord, Polym. Bull. 8, 325 (1982).

    230. G. Marrucci, Macromolecules 14, 434 (1981).

    J. P. QUESLEL

    Manufacture Michelin, CERL GPA

    J. E. MARK

    University of Cincinnati