Anda di halaman 1dari 11

9.

10

SOME REFINEMENTS TO RUBBER ELASTICITY

459

Figure 9.18 MooneyRivlin plot for sulfur-vulcanized natural rubber. Solid line, equation (9.48); dotted line, equation (9.63).

9.10

SOME REFINEMENTS TO RUBBER ELASTICITY

The statistical theory of rubber elasticity has undergone signicant and continuous renement, resulting in a series of correction terms. These are sometimes omitted and sometimes included in scientic and engineering research, as the need for them arises. In this section we briey consider some of these. 9.10.1 The Inverse Langevin Function

The Gaussian statistics leading to equation (9.34) are valid only for relatively small strainsthat is, under conditions where the contour length of the chain is much more than its end-to-end distance. In the region of high strains, where 1 1 the ratio of the two parameters approaches 3 to 2 , this limit is exceeded. Kuhn and Grn (65) derived a distribution function based on the inverse Langevin function. The Langevin function itself can be written L( x) = coth x 1 x (9.67)

and was rst applied to magnetic problems. In this case L(b ) = r n l (9.68)

where n is the number of links of length l. (It must be pointed out that the quantity n in this case need not be identical with the number of mers in the chains.) Thus the quantity nl represents a measure of the contour length of the chain, and r/nl is the fractional chain extension. Of course, for the inverse Langevin function of interest here,

460

CROSS-LINKED POLYMERS AND RUBBER ELASTICITY

L-1

r = b n l

(9.69)

The stress of an elastomer obeying inverse Langevin statistics can be written (42,63) a 1 1 2 - a -3 2 L-1 s = nRT (n) L-1 1 2 1 2 1 2 ( n ) a ( n ) (9.70)

At intermediate values of a (and hence of nl), equation (9.70) predicts a sharp upturn in the stress at as greater than 4, as observed in experiments. Because of the complexity of equation (9.70), the Gaussian-based (9.34) is preferred where possible. 9.10.2 Cross-link Functionality

In order to form a network, at least some of the mers need to have a functionality greater than 2; that is, more than two chain portions must emanate from those mers. In the structure depicted in Figure 9.3, the functionality of each cross-link is 4. When divinyl benzene or sulfur is used as a cross-linker, the functionality will indeed be 4. Suppose, however, that glycerol is used as the cross-linker in the synthesis of a polyester. Then the functionality of the cross-link site will be 3. Use of trimethylol propane trimethacrylate or pentaerythritol tetramethacrylate results in functionalities of 6 and 8, respectively (66). Duiser and Staverman (67) and Graessley (70) have shown that the front factor depends on the functionality of the network. Representing the network functionality as f *, equation (9.34) can be written s= ri 2 1 f * -2 nRT a- 2 f* 2 a r0 (9.71)

For tetrafunctional cross-links, dened as four chain segments emanating from each cross-link site [the same type as obtained with the use of divinyl benzene (see Figure 9.3)], f * = 4, equation (9.71) predicts one-half the stress that equation (9.34) predicts. Another way of writing the correction for cross-link functionality is (41,69) s = (n - m )RT ri 2 1 a- 2 2 a r0 (9.72)

where n and m are the number densities of elastically active strands and junctions. A junction is elastically active if at least three paths leading away from

9.10

SOME REFINEMENTS TO RUBBER ELASTICITY

461

it are independently attached to the network. A strand, meaning a polymer chain segment, is elastically active if it is bound at each end by elastically active 1 junctions (70). Equation (9.72) also predicts a front-factor correction of 2 for a tetrafunctional network, since there are half as many cross-links as there are chain segments. 9.10.3 Network Defects

There are two major types of network defects: (a) the formation of inactive rings or loops, where the two ends of the chain segment are connected to the same cross-link junction, and (b) loose, dangling chain ends, attached to the network by only one end (7174) (see Figure 9.19). Both of these defects tend to decrease the retractive stress, because they are not part of the network. The equation in use to correct for dangling ends may be written 2 Mc ri 2 1 s = nRT 1 a- 2 2 M r0 a (9.73)

where Mc is the molecular weight between cross-links and M is the primarychain molecular weight. Where M > > Mc, the correction becomes negligible. 9.10.4 Volume Changes and Swelling

If a polymer network is swollen with a solvent (it does not dissolve), equation (9.32) may be rewritten a xa ya z = 1 v2 (9.74)

where v2 is the volume fraction of polymer in the swollen material. Of course, v2 is less than unity, and axayaz is larger than unity, as is commonly experienced.

Figure 9.19 Network structure and defects: (a) elastically active chain, (b) loop, and (c) dangling chain end.

462

CROSS-LINKED POLYMERS AND RUBBER ELASTICITY

The detailed effect has two parts:


2 1. Effect on the front factor, ri2/r0 . The quantity ri2 increases with the volume V to the two-thirds power, since ri itself is a linear quantity. Of course, 2 r0 remains constant. Thus

ri 2 V 2 =V r0 s 0

2 3

ri 2 1 ri 2 2 = 23 2 r0 v2 r0

(9.75)

where the subscript s refers to the swollen state and where V0 is the volume of the unswollen polymer. 2. Effect on the number of network chain segments concentration, n. The quantity n decreases with volume: V0 n = ns V where ns is the chain segment concentration in the swollen state. V0 n = v2 n V (9.77) (9.76)

Incorporating the right-hand sides of equations (9.75) and (9.77) into equation (9.34) leads to an equation of the form (75,76)
3 s = nRTv1 2

ri 2 1 a- 2 2 a r0

(9.78)

The stress, dened as force per unit actual cross section, is decreased by v1/3 2 , since the number of chains occupying a given volume has decreased. When the volume change caused by deformation alone is considered (usually less than 1%), the equation of state can be written (7780) s = nRT V0 V
2 3

ri 2 1 V a- 2 2 V0 a r0

(9.79)

9.10.5

Physical Cross-links

9.10.5.1 Trapped Entanglements So far the discussion has been restricted to ordinary covalent cross-links. There are, however, several types of physical cross-links that exist as permanent loops or entanglements exist-

9.10

SOME REFINEMENTS TO RUBBER ELASTICITY

463

ing in the network structure. (They may slide, however, yielding a mode of stress relaxation also.) Three types of trapped entanglements are shown in Figure 9.20 (8186).They each portray the same phenomenon, but with increasing rigor of denition. Early works referred to the chemical and physical cross-links in a simple manner, s = (nc + n p )RT ri 2 1 a- 2 2 a r0 (9.80)

where nc and np are the concentration of chains bound by chemical and physical cross-links, respectively. It had been established early, for example, that the retractive stress was higher than expected by nearly a constant amount; indeed, for short relaxation times even linear polymers above Tg behaved as if they had some type of cross-linking (87,88) (see also Section 8.2). Figure 9.21 illustrates the rubbery plateau (see Section 8.2) for a dynamic mechanical study of polystyrene as a function of frequency. The plateau shear modulus, near 3 106 dynes/cm2, corresponds to a number of active network chains of near 1 10-4 mol/cm3, nearly independent of the molecular weight of the polymer. A more recent approach uses the concept of the potential entanglements that have been trapped by the cross-linking process. Langley (85) denes the quantity Te as the fraction (or probability) that an entanglement is trapped in this manner. Two theories, developed by Flory (89) and Scanlan (90), yield the calculation of chemical cross-links (58,59). For Florys theory, the total number of effective cross-links is ntot = nc WgTe1 2 + neTe (9.81)

where ne is the concentration of potential entanglement strands and Wg is the weight fraction of gel. For Wg = 1, equation (9.81) reduces to nc + np [see equation (9.79)].

Figure 9.20 Three types of trapped entanglements: (a) The Bueche trap (8186), (b) the Ferry trap (84), and (c) the Langley trap (77). The black circles are chemical cross-link sites. After Ferry (86).

464

CROSS-LINKED POLYMERS AND RUBBER ELASTICITY

Figure 9.21 Shear storage modulus versus frequency for narrow molecular-weight polystyrenes at 160C. Molecular weights range from MW = 8900 g/mol (L9) to MW = 581,000 g/mol (L18) (88).

With the Scanlan criterion, ntot = 1 n T 1 2 (3Wg - Te1 2 ) + neTe 2 c e (9.82)

which also reduces to nc + np for Wg = 1. The value of these relationships, of course, is that for real networks, Wg < 1, and the way is open to evaluate nc and ne. Some calculations are shown in Figure 9.22. As they illustrate, the effective (permanent) physical cross-links start out at zero when the system is linear and increase rapidly to a plateau level (91). 9.10.5.2 The Phantom Network It must be remarked that considerable controversy exists over the existence of physical cross-links (40,41,87,92). A theory has been proposed by Flory (40,41) using mathematics of a simplied network, called the phantom network. The model consists of a network of Gaussian chains connected in any arbitrary manner. The physical effect of the chains is assumed to be conned exclusively to the forces they exert on the junctions to which they are attached. For a perfect phantom network of functionality f *, the front factor contains the term (f * - 2)/f *, leading to equation (9.71), which considers chemical crosslinks of arbitrary functionality, but no physical cross-links.

9.10

SOME REFINEMENTS TO RUBBER ELASTICITY

465

Figure 9.22 Contributions of chemical cross-links and trapped entanglements to the total cross-link level, for polystyrene, Mn = 5 105 g/mol, MW = 1 106 g/mol, based on equation (9.81) (91).

Flory (40) argues that the presence or absence of the term ( f * - 2)/f * may depend on the magnitude of the strain. At small deformations the displacement of junctions may conform more nearly to the older assumptions (i.e., afne in the macroscopic sense), and hence the term ( f * - 2)/f * might not appear for real chains that do have entanglements. 9.10.6 Small-Angle Neutron Scattering

In the preceding text the chains were assumed to deform in an afne manner when the networks were swelled or stretched. Until recently there was no way to approach this problem experimentally. With the advent of small-angle neutron scattering (SANS), the conformation of the chains in the bulk state could be investigated (see Section 5.2.2.1). On swelling, chain dimensions increase, usually isotropically. When SANS studies are done on the stretched elastomers, the scattering pattern yields greater dimensions in one direction than the other, because the chains are anisotropic (see Section 5.2) (94106). The question arose whether polymer chains in a network had the same conformation as in the melt before cross-linking. Beltzung et al. (93) prepared well-dened poly(dimethyl siloxane) (PDMS) chains containing SiH linkages in the a and w positions. Blends of PDMS(H) and PDMS(D) were prepared, where H and D, of course, represent the protonated and deuterated analogues. These blends were end-linked by tetrafunctional or hexafunctional cross-linkers under stoichiometric conditions. Neutron scattering was carried out on both the PDMS melts and the corresponding networks. The principal results were that (a) the Gaussian character of the network chains in the undeformed state was conrmed, and (b) the chain dimensions were not changed by the cross-linking process.

466

CROSS-LINKED POLYMERS AND RUBBER ELASTICITY

Benoit and others (9498) prepared two types of tagged polystyrene networks: (a) type A networks containing labeled (deuterated) cross-link sites. This permitted a characterization of the spatial distribution of the cross-link points. (b) Type B networks containing a few percent of perdeuterated polystyrene chains (see Figure 9.23) (94). Cross-linking utilized divinyl benzene (DVB). Benoit et al. (94) studied these polystyrene networks as is, swollen in several solvents, and stretched. For the latter, stretching was done above Tg, followed by cooling in the stretched state. They found a maximum in the angular scattering curves of type A networks. The mean pair separation distance between chain ends, h, was found to be 0.5 both in the dry state and in the swollen state. On extenproportional of Mc sion, h|| and h^ values followed the expected afne deformation. The B network, on the other hand (94), appeared to deviate signicantly from the afne. As illustrated in Figure 9.24 (94) the chain radius of gyration

Figure 9.23 Schematic representation of labeled polystyrene networks. (A) Cross-linking points labeled. (B) Random labeled chains added (96).

Figure 9.24 Variation of the radius of gyration of type B networks of different functionalities. Dotted line, theoretical behavior for afne deformation; dashed line, theoretical behavior for the end-to-end pulling mechanism (96), for polystyrene.

9.10

SOME REFINEMENTS TO RUBBER ELASTICITY

467

increased on swelling far less than predicted by the afne deformation mechanism. The values of Rg were also less than predicted by afne deformation, leading to the end-to-end pulling mechanism (97), which accentuates the extension of the end portions of the chains rather than the central section. One might imagine that entanglements prevent the motion of the central portions of long chains. Another possible explanation (see below) is that the chains cross-link junction points rearrange to yield the system with the lowest free energy. This rearrangement minimizes the actual extension of the chain. More recent SANS experiments on stretched networks were performed by Hinkley et al. (99) and by Clough et al. (100,101). Hinkley et al. (99) prepared blends of polybutadiene and polybutadiene-d6. Both polymers were made by the living polymer technique, end-capped with ethylene oxide, and waterwashed to yield the dihydroxy liquid prepolymer. Uniform networks were prepared by reacting the prepolymers with stoichiometric amounts of triphenyl methane triisocyanate. The value of using polybutadiene over polystyrene, of course, is that the networks are elastomeric at ambient temperatures. These networks (99) were extended up to a = 1.6 and characterized by SANS. Owing to large experimental error, no denitive conclusion could be reached, although the data t the junction afne model better than either the chain afne model or the phantom network model (Section 9.10.5.2). Random types of cross-linking are of special interest for real systems. Clough et al. (100,101) blended anionically polymerized polystyrene with PSd8 and cross-linked the mutual solution with 60Co g-radiation. Bars of the crosslinked polystyrene were elongated at 145C and cooled. Specimens were cut in both the longitudinal and transverse directions and characterized by SANS. The quantity R= Rg (stretched) Rg (unstretched) (9.83)

was plotted versus a, as illustrated in Figure 9.25 (100,101). The transverse measurements are divided into anisotropic and end-on. The anisotropic measurements refer to the case where the beam was perpendicular to the direction of orientation (101), and the end-on measurements refer to the case where the beam was parallel to the stretch direction. In neither of these cases was afne chain deformation followed. Both of these experiments, however, yielded identical results within experimental error, conrming important macromolecular hypotheses. In Figure 9.25 the quantity R = [(a 2 + 3)/4]0.5 was obtained from the dependence predicted for a tetrafunctional phantom network (101). The quantity R = [(a 2 + 1)/2]0.5 represents the afne junction case. The constant value of R^ up to a = 2 suggests that the chains are deforming far less than the junctions. These results follow neither the R = a (parallel) nor the R = a-0.5 (perpendicular) prediction but rather support Benoit et al. (94) that afne chain behavior is not followed.

468

CROSS-LINKED POLYMERS AND RUBBER ELASTICITY

Figure 9.25 The ratio R = Rg (stretched)/Rg (unstretched) as a function of sample elongation (100): , longitudinal; , longitudinal; , transverse, anisotropic; , transverse, end-on; , transverse, anisotropic; , transverse, end-on.

In a series of theoretical papers, Ullman (103105) reexamined the phantom network theory of rubber elasticity, especially in the light of the new SANS experiments. He developed a semiempirical equation for expressing the lower than expected chain deformation on extension: l*2 = l2 (1 - a ) + a (9.84)

as the basis of a network unfolding model. The quantity l* is dened as the ensemble average of the deformation of junction pairs connected by a single submolecule in the network, and l is the corresponding quantity calculated for the phantom network model. The quantity a expresses the fractional deviation from ideality. The phantom network corresponds to a = 0. If a = 1, the chain does not deform at all on network stretching or swelling. From the data of Clough et al. (100,101), Ullman (103,104) concluded that a was in the range of 0.36 to 0.53. In the above, Hinkley et al. (99), Clough et al. (100,101), and Benoit et al. (94,95) utilized end-linked networks. Ullman (103,104) delineated the differences between the two types of network. He pointed out that randomly cross-linked chains deform to a greater extent than end-linked chains, that sensitivity to network functionality is much greater for end-linked chains, and that for high cross-linking levels, the randomly cross-linked chain approaches the macroscopic deformation of the sample. Ullman (105) recently reviewed these and other SANS experiments on the deformation of polymer networks. Recently Hadziioannou et al. (106) prepared amorphous polystyrene with extrusion ratios up to 10, using a solid-state coextrusion technique. Their poly-

9.11 INTERNAL ENERGY EFFECTS

469

Figure 9.26 Model of the end-pulling mechanism, showing how R|| increases, while R^ remains nearly constant: (A) relaxed and (B ) stretched.

styrene had a molecular weight of 5 105 g/mol. The anisotropy of the Rg values agreed with those predicted on the basis of a chain afne model. While general conclusions appear to be premature, it appears that the crosslink sites rearrange themselves during deformation to achieve their lowest free-energy states; thus the chains deform less than the afne mechanisms predicts. A modied end-pulling mechanism is also possible. A possible molecular mechanism, which results in minimal changes in R^, is illustrated in Figure 9.26 (107). The debate over the exact molecular mechanism of deformation is sure to continue.

9.11 INTERNAL ENERGY EFFECTS 9.11.1 Thermoelastic Behavior of Rubber

In Section 9.5 some of the basic classical thermodynamic relationships for rubber elasticity were examined. Now the classical and statistical formulations are combined (108,109). Rearranging equation (9.8), fe = f - T f T L ,V (9.85)

Dividing through by f and rearranging, fe ln f = 1 ln T L ,V f (9.86)

Rewriting equation (9.79) in terms of force, and substituting equation (9.38), we nd that

Anda mungkin juga menyukai