This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
= 90° entropy single chain scattering function total scattering function temperature threshold tearing strength mean energy of a chemical bond internal energy volume molar volume of solvent work strain energy function function in the Valanis-Landel equation number of monomer units in a statistical segment is at 90° to the direction of stretch R^ = R\. Theoretical values of R^ for deformed elastomers are easy to derive. In the phantom network model (0) (j)i
a1 a2 P y
parallel polarizability of a statistical chain segment perpendicular polarizability of a statistical chain segment quantity in the theory of highly stretched chains, p = ^'x (r/(n b)) shear gradient
N2 Pi p? pu pc P(r,r + dr, n) P P 2 (cos 9) q q qpeak r r r0 rtj r'ij
360
8 Rubber Elasticity
Q
angle between stretching axis and C - D bond scattering angle, equal to twice the Bragg angle polar angle polar angle of segment i parameter in the Flory-Erman theory neutron wavelength deformation ratio deformation ratios in x-, y-, and z-directions chemical potential of solvent splitting frequency in deuterium NMR spectrum osmotic pressure density stress principal stresses Flory-Huggins interaction parameter azimuthal angle function in the Ogden equation volume fraction of solvent volume fraction of polymer angle between stretching axis and magnetic field
GPC MEK NMR PB PS PTMI PTMO PTMT SANS TPE
gel permeation chromatography methylethylketone nuclear magnetic resonance polybutadiene polystyrene poly(tetramethylene isophthalate) poly(tetramethylene oxide) poly(tetramethylene terephthalate) small angle neutron scattering thermoplastic elastomers
y 9 9
Av n a
X cp (pi
cpx cp 2
8.1 Introduction and Historical Background
361
8.1 Introduction and Historical Background Rubber elasticity is the property of reversible deformability of large magnitude exhibited by a number of polymeric materials, chief among which is natural rubber, a substance prepared from the coagulated sap of a tropical tree, Hevea brasiliensis. Under certain conditions, noncrosslinked polymers will show rubbery behavior due to chain entanglements and/or some type of physical crosslinks such as crystalline microdomains (see Chapter 9). True rubber elasticity requires a permanent network (see Fig. 8-1). Upon cooling, rubbers become glassy, and some rubbers partially crystallize when cooled. Natural rubber becomes glassy at -70°C before vulcanization, and vulcanized natural rubber is a glass at -60°C. Conversely, crosslinked polymeric glasses become rubbery when heated. An example is polystyrene crosslinked by divinylbenzene, which is glassy below 80 °C, but is a rubber at 150°C. In the absence of deformation, natural rubber has an X-ray pattern at room temperature similar to that of a low molecular weight liquid. Katz (1925) showed by Xray diffraction that natural rubber crystallizes at ordinary temperatures when stretched. Crystallization of natural rubber occurs slowly at lower temperatures (0°C and less), in the undeformed rubber as well. The shear modulus of a typical rubber is of the order of 20 N/cm 2 , rubber has an elastic deformability of several hundred percent, and a bulk modulus thousands of times greater than the shear modulus. The molecular mechanism of rubber elasticity is distinctly different from the elasticity of metals and other hard solids. The relationship between molecular and macroscopic deformation in rubbers is central to the understanding of elastomeric materials.
Figure 8-1. (a) Schematic depicting entangled linear polymer chains. Such materials exhibit rubbery behavior for short times at temperatures above their glass transition, (b) Schematic depicting a crosslinked network with tetra-functional junctions. Such samples can show nearly ideal rubber elasticity.
The unique mechanical properties of elastomers account for their use in a wide variety of industrial products such as pneumatic tires, elastic foams and rubber bands. Much of the scientific interest in rubber is the consequence of the broad applicability
362
8 Rubber Elasticity
of elastomeric substances for commercial use. Michael Faraday (1826) worked on the chemical composition of rubber. He found a weight ratio of carbon to hydrogen of 6.8 (the correct value is 7.5). On this basis, he suggested that rubber contained a ratio of eight carbon atoms to 7 hydrogen atoms. Only a few years after this, Goodyear (1844) in America and Hancock (1844) in England, reported that heating rubber with sulfur improved its properties in several ways. Tackiness is eliminated, stiffness of the rubber can be increased if desired, and solubility in a variety of solvents is eliminated. In 1805, Gough (1805) discovered some highly important thermoelastic properties of rubber. He found that rubber which is rapidly stretched, increases in temperature. He also found that a strip of rubber from which a small weight is suspended shrinks on being heated. These results, obtained on unvulcanized rubber, were confirmed by Joule (1859) on the vulcanized product a half century later. Thomson (1857) (later Lord Kelvin) provided a thermodynamic analysis of rubber which yielded a clear explanation of the results found by Gough and by Joule. A number of researchers in the last decades of the nineteenth century and first decade of the twentieth century studied the addition of fillers in rubber. A report by Mote of the India Rubber, Gutta Percha and Telegraph Works of Silverton, England on the reinforcement of rubber by carbon black was never published, but laid the basis for the use of carbon black as a reinforcing filler in rubber tires. Several interesting synthetic rubbers were developed in the two decades after 1920. Butadiene-styrene copolymers known as Buna S were synthesized, and these were rubbery at room temperature if the butadiene content was sufficiently high. Rubbers
made from copolymers of styrene and acrylonitrile, known as Buna N, were resistant to attack by oil and grease, chloroprene (2-chlorobutadiene) could be made into a rubber with excellent properties. Thiokol rubber was prepared from sodium polysulfide and 1,2-dichloroethane. Polymerization of isobutene led to a tacky elastomer known as Vistanex, and copolymerization of isobutene with a small quantity of butadiene yielded the vulcanizable rubbery product known as butyl rubber. Today, a great variety of rubbery materials exist. All derive their favorable physical properties from their molecular network structure.
8.2 Rubber Processing 8.2.1 Vulcanization of Natural Rubber
An elastomeric material has a molecular structure of an interconnected network of polymeric chain molecules. The network is usually based on primary chemical bonds between the chains, but can also be formed by strong attraction between the chains (e.g., by functional groups on these chains or by crystallization). As mentioned above, the process of vulcanization was first developed by Goodyear (1844) and Hancock (1844) by heating natural rubber with sulfur. Chemical bonds are formed by chains of sulfur atoms which react with the unsaturated bonds of the primary macromolecules. Modern elastomers utilize a wide variety of chemicals to produce the permanent network. 8.2.2 Some Aspects of Rubber Chemistry
Natural rubber poly(ds-isoprene) contains one double bond per monomer before vulcanization, and still contains many unreacted double bonds after vulcaniza-
8.3 Theory of Rubber Elasticity
tion is completed. These unreacted double bonds react slowly with oxygen from the air, which can lead eventually to degradation of the rubber. The unsaturation of the rubber insures that rubber is chemically active with a large variety of substances both before and after vulcanization. Hydrogenation of rubber to saturate the residual double bonds brings about chemical stabilization. Rubbers can also be chlorinated, and these chlorinated products have many applications. Many other chemical reactions can be brought about with natural rubber. Network formation usually takes place by chemical bonding of the primary macromolecules of which the network is formed (for a detailed discussion, see Vol. 18 of this Series). Synthetic rubbers, prepared from hydrocarbons, contain unsaturated bonds in the chain which can be crosslinked by reaction with organic peroxides. A good example is the preparation of butyl rubber by reaction of an isobutenebutadiene copolymer with dicumyl peroxide. The mixture of polymer and peroxide is heated, the peroxide decomposes to form free radicals, and these react with the double bond in the polymer to form a stable molecular network. Polydimethylsiloxanes are usually prepared from the cyclic tetramer by ringopening polymerization. The product is a linear polymer chain which is liquid at room temperature. Addition of a small quantity of trifunctional or tetrafunctional monomer to the initial reactants leads to formation of a three dimensional rubbery network. Crosslinking can also be brought about by reacting a vinyl terminated linear polysiloxane with a source of free radicals which, for example, are generated by thermal decomposition of an organic peroxide. Silicone rubbers are more heat-resistant than hydrocarbon rubbers, and are often
363
used when stability at high temperatures is an important requirement and at very low temperatures where flexibility is important. Polyurethanes are condensation products of polyhydric alcohols and di- or triisocyanates. Linear polymers are formed by reaction of stoichiometric quantities between dihydric alcohols and diisocyanate. Addition of a small amount of tri- or tetrahydroxy alcohol or of a triisocyanate leads to the formation of a three-dimensional rubbery network. Fluorocarbon rubbers are prepared by crosslinking a copolymer of tetrafluoroethylene and hexafluoropropylene. These rubbers are thermally stable and have very low coefficients of friction. They are resistant to oxidative degradation and to swelling by hydrocarbons. They are excellent materials for a variety of specialized applications.
8.3 Theory of Rubber Elasticity 8.3.1 Thermodynamics of Rubber Elasticity
If the work done by the system is purely mechanical, either elastic or dilatational, dW = PdV — FdL, P being pressure, V9 volume, F, force, and L, length. In most experiments on elastomers, PdV is thousands of times smaller than FdL, since rubber volumes are largely unchanged when the rubber specimen is deformed. In such circumstances, the PdV term is neglected so that the derivative of the Gibbs free energy is
In applications in which pressure - volume changes are small, VdP and PdV are omitted in thermodynamic calculations. Then dU = dH and the Helmholtz free en-
364
8 Rubber Elasticity
ergy and Gibbs free energy are related as dA = dG. From Eq. (8-1), it follows that
2.0
if
(8-2) 1.6
//
and using Maxwell's relations PLT
'
(8-3)
d¥
Equations (8-2) and (8-3) are based on the constant volume assumption for rubber. Using these equations, changes in entropy, free energy, and internal energy are derived from uniaxial force elongation experiments as a function of temperature.
My//
1.2 ro Q_ _c b
0.8
0.4 /
&.A = J F dL;
T is constant
(8-4 a)
Li
(c)
0 .
\
(8-4 b) AU =
TAS
(8-4 c)
Studies of changes in free energy, entropy and internal energy of elastomers have been performed by many researchers. The actual experimental process may be difficult because of hysteresis in stretching and relaxation. Also, real rubbers may creep or exhibit stress relaxation if crosslinking is incomplete, or if the substance undergoes slow chemical degradation while under stress. In Fig. 8-2, experiments on changes in entropy, free energy, and inter* nal energy of natural rubber as a result of stretching (Anthony et al., 1942) are shown. It is clear from this study that changes in internal energy upon deformation of rubber are very small by comparison with changes in free energy and the entropy component of free energy. Changes in internal energy can become substantial at high deformations in some elastomers, but this is a result of crystallization. The small change in internal energy at moderate de-
Figure 8-2. Tensile stress-deformation measurements on natural rubber, (a) G versus A; (b) — T(dcr/dT)x versus A; (c) o— T(da/dT)^ versus X [curve (a) — curve (b)]. The areas under each of these curves yield A^4, - TAS, and A U, respectively. (Anthony et al., 1942.)
formation is akin to the small change of internal energy of a dilute gas under compression. Accordingly, an ideal rubber is similar to an ideal gas. Changes in internal energy upon isothermal deformation are generally small, though easily measurable. If a rubber specimen is rapidly stretched there is an increase in temperature. The specific heat is roughly constant during elongation. Data on rubber by Joule (1859) and by James and Guth (1943) on adiabatic heating of rubber are shown in Fig. 8-3. Note that AT is negative at very low elongation, which arises from an initial increase in entropy which takes place before the large decreases in entropy occur at higher deformations. The minimum in the
8.3 Theory of Rubber Elasticity
0.20
J
0.16
— C
o in
>
>
c
c
0.08
u
/
t_
L_
A/
A /
tic
0.12
o
sion
,o
/
(0
JQ
E 0.04
o
(0 TJ <^
/
A
(8-5)
A*
0.00
1.0
1.2
links is almost never fully extended, but is coiled to a considerable degree. In an assembly of such polymer chains, the ends of the chains are separated by a distance r which varies with time and differs from one chain to another. In the absence of external forces, the probability that the end-to-end vector lies between r and r + &r is given by
/ /
— In H
|
365
I
I
I
1.4
1.6
1.8
Figure 8-3. Temperature rise of rubber in adiabatic extension, A: Joule (1859); o: James and Guth (1947).
AT versus deformation plot is known as the isometric inversion point, and the point at which AT = 0 is known as the adiabatic inversion point.
8.3.2 The Statistical Theory of Rubber Elasticity
The fundamental ideas for a statistical mechanical analysis of rubber appear in works by Meyer et al. (1932) and also by Karrer (1933). The latter was aimed primarily on a theory of muscle relaxation. In 1934 a statistical theory of rubber elasticity was put forth (Guth and Mark, 1934) which developed the ideas of Meyer and co-workers by formal statistical mechanical methods. The Guth-Mark theory was based on what is known today as a Gaussian polymer chain. In this model, a rubber is considered as an assembly of polymer molecules, each composed of n links of length b connected to each other by covalent chemical bonds. These links rotate about each other in space. The chain of n
Since the mean square distance between the ends of such chains is given by (r2y = nb2, it is easy to verify that a = (3/2) nb2. This provides the background for calculations of molecular dimensions of a polymer chain in a rubbery polymer described by a Gaussian model. A rubbery material composed of an ensemble of Gaussian chain molecules undergoes a partial orientation and elongation when subjected to an applied external force. In the theory of rubber elasticity, the deformation of the macroscopic specimen is identified with the molecular deformation of the polymer chains of which the elastomer is composed. Two classical pictures play prominent roles in this description, one is known as thefixedjunction or junction-affine model, and the other is called the phantom network. In these models, three or more Gaussian polymer chains are joined at their ends by primary chemical bonds to form a three dimensional network with crosslink junction points (see Fig. 8-1 b). In the fixed junction model, these crosslink junctions are immobile with respect to the deformed Cartesian coordinates, which transform affinely with the macroscopic deformation. If the macroscopic deformation is given by XY, X2, and A3 in the x-, y-, and z-directions, it follows from the fixed junction model that the mean square Cartesian components of the
366
8 Rubber Elasticity
deform according to
end-to-end vector deform as
nb:
f
nb2^
(8-6 a) (8-6 b) <2 2 >
nb: =
(8-6 c)
The subscript '0' signifies the value of the quantity before deformation of the polymer chain. The angle brackets indicate an ensemble average. In the phantom network, the name currently in use for the model of James and Guth (James, 1947; James and Guth, 1947), the crosslink junctions fluctuate about mean positions. When the rubber is deformed and the mean values of the components of the end-to-end vector of each chain are deformed by factors /l 1? A2, and A3, r fluctuates about that mean owing to Brownian motion. These fluctuations are calculated using the theory of Brownian motion of harmonically bound particles. It is assumed that the fluctuations are not influenced by the presence of neighboring chains in the rubber, and are independent of r, the mean value of r. r = r + 5r 2
2
(8-7 a) 2
r = r + (Sr)
(8-7 b)
The network functionality, /, is defined as the number of polymer chains joined at a crosslink. In the phantom network 2 <(5r) 2 >=y
(8-8)
One of the interesting characteristics of the theory is that the magnitude of the fluctuations is independent of the deformation of the network. Since
(2/f)nb2, and
Upon
deformation, the Cartesian components
7
(8-9)
2
Similar equations hold for
(8-10)
Here, kB is Boltzmann's constant. If the specimen is deformed by factors Al9 A2, and X3 in the three principal directions, the contribution of the deformation along the x-direction to the entropy is ix)
=
const — kBaAf2x2
(8-11)
where Xf = X1 for the fixed junction model, and k%2 = X\ (1 - 2//) + 2 / / for the phantom network. For the entire network, the entropy is calculated by averaging over all arrangements of the unperturbed chain multiplied by the number of chains. The entropy change is given by
8.3 Theory of Rubber Elasticity
Equation (8-12) for the phantom network is just obtained by multiplying the fixed network result by a factor of (1—2//). The point of view taken in developing these equations is that the internal energy of the elastomer is unchanged by specimen deformation, at constant temperature, and therefore AA = - TAS (8-13) The number of network chains N equal the number of chains per mole, NA, which is Avogadro's number, multiplied by density and sample volume and divided by the molecular weight between crosslinks, M c . Equation (8-13) may therefore be written (8-14) In uniaxial extension, kx = k and A2 = A3 = X~1/2. In this case, the free energy change of the fixed junction model is given respectively by (8-15) The free energy change is the reversible work of stretching. Therefore, the force exerted on a specimen in uniaxial deformation is given by
RTQA0
(8-16)
Ao is the area of cross section of the specimen before deformation. Using the constancy of rubber volume upon deformation, the stress on the sample in the deformation direction, a = F/A9 is given by
-('-?)(*--
(8-17)
In many experimental studies, researchers work with rubbers swollen by
367
solvents. The free energy changes calculated above yield the elastic part of the free energy only. In addition, there is a term in the free energy arising from solvent-polymer interaction, and this is part of the total free energy change of swelling. The calculation of the total free energy of swelling is usually performed by treating the elastic and solution contributions as independent additive components. The free energy of swelling is based on the Flory-Huggins theory of polymer solutions (Flory, 1942; Huggins, 1942, 1943). In the Flory-Huggins theory, the free energy of mixing of Nx solvent molecules with N2 polymer molecules is given by (8-18) In this theory, the volume of a solvent molecule is taken to be equal to that of one monomer unit on a polymer molecule, and cp1 and q>2 are the volume fractions of solvent and polymer respectively. % is known as the Flory-Huggins interaction parameter, and its magnitude characterizes the strength of the polymer-solvent interaction energy (see Chap. 7, Vol. 5 of this Series). In applying Eq. (8-18) to a swollen elastomer, there is only one polymer molecule, that is, N2 equals one, and the term JV2 lncp2 is negligibly small. However, q>2 is a number usually between 0.1 and 1.0, and, accordingly, the final term in Eq. (8-18) is substantial in magnitude. The solvation part of the free energy of swelling of a network is then given by A^ s = /cBT(iV1ln(p1+ZiV1(p2)
(8-19)
The elastic free energy of swelling, A^4e, is given by Eq. (8-13) with the linear swelling ratios equal in all directions A =— A-£ -— A 2 ~~~ A 3
(8-20)
368
8 Rubber Elasticity
The total free energy of swelling becomes AA = (8-21) The number of polymer chains in the rubber specimen is AT, and these are interconnected by crosslinks to form a single gigantic macromolecule. The factor 1 — 2 / / appears in the phantom network result, but is omitted in the formula for the fixed junction model. Many studies on swollen elastomers are conducted by deforming the isotropic gel by stretching or compressing. The additional change in free energy arising from uniaxial deformation is given by - - 3
(8-22)
for the phantom network. In Eq. (8-22), X is the linear deformation ratio of the already swollen elastomer. In general, the chemical potential of the solvent in a swollen elastomer differs from that of the pure solvent, and this accounts for changes in vapor pressure of the solvent and accounts for an osmotic pressure of the swollen gel. The chemical potential of the solvent is calculated by differentiating AA with respect to N1. In doing this, it is essential that the interrelations between
(8-23)
To the extent that the solvent in the vapor phase may be treated as an ideal gas (8-24)
where px is the vapor pressure of the solvent above the swollen elastomer and p\ is the vapor pressure of the pure solvent. By measurement of these vapor pressures, it becomes possible to ascertain whether Eq. (8-23) correctly describes gel swelling. The osmotic pressure, n, of the solvent in the swollen gel is linked to the chemical potential of the solvent by An1 = -nvl
(8-25)
where v1 is the partial molecular volume of the solvent. Measurement of osmotic pressure provides another useful method for checking the theoretical equations. The equilibrium swelling ratio of an elastomer occurs at a point where the chemical potential of the solvent in the gel equals that of the pure solvent, that is, A / ^ ^ 0 . The Flory interaction parameter, X, can be extracted from Eq. (8-23) by setting Afix to zero, by use of measurements of the maximum swelling of the elastomer. This procedure has been adopted in a number of investigations. The thermodynamics of network deformation by applied mechanical forces or by solvent interaction as described above treats polymer chains as ideal random flight species taking no account of attractive and repulsive interactions between chain segments or of entanglements of the polymer chains. This has been long recognized as a serious defect of the elementary models, can in some recent theories, corrections for these omissions have been introduced. There is no general agreement among specialists how to make these corrections, and some current approaches will be considered next. One major effort to modify the elementary model was initiated by Flory (1976, 1977), and was further developed by Erman and Flory (Flory and Erman, 1982; Erman and Flory, 1982). In their model,
8.3 Theory of Rubber Elasticity
the effect of entanglement and other interchain interactions is to change the mean position of the crosslink junctions and to reduce the magnitude of the junction fluctuations. Flory and Erman introduced new parameters in the theory to account for the effects. The principal quantity, x, is the ratio of the mean square junction fluctuation in the phantom network to that of the real network. The fluctuations in this model are smallest in the undeformed network, but become larger as the network is stretched. The phantom network limit is approached as the deformation becomes very large. Networks tend to be more like the fixed junction model in the undeformed state. Another approach to a corrected theory of rubber elasticity is based on tube models. This idea, introduced by de Gennes (1971), has been extended by Edwards and co-workers (Ball, 1981; Edwards and Vilgis, 1987). In the Edwards-type model, a polymer chain in an elastomeric network can be treated as a long worm encased in a tube composed of neighboring molecules, and the presence of this tube constrains the motion of the chain and limits the number of conformations available to the chain. The equations for chain deformation and chain thermodynamics take into account the tube structure, and provide predictions of elastomeric behavior of a stretched or swollen network. Some specialists in rubber elasticity have used two-network models to represent the behavior of real rubbers. Chief among the proponents of this point of view have been Ferry and Graessley. In a two network model, contributions to the modulus arise from retractive forces imposed by the primary chemical network plus retractive forces originating in the entanglement structure. These contributions are assumed to be additive, with the actual elastic modulus taken as the sum of the moduli of the
369
two networks. Experiments by Ferry and associates and by Graessley and co-workers tend to be in good agreement with this view. Details of some of these studies will be discussed later. 8.3.3 The Mooney-Rivlin Model of Rubber Deformation The molecular theory of rubber elasticity based on the unmodified Gaussian structure of a polymer chain leads to the result that the energy stored in deformation of an elastomer may be written as W=C(kl + kl + kl-3)
(8-26)
where C is a constant derived from the molecular theory and W is known as the strain energy function. Equation (8-26) is clearly inadequate for explaining many experiments on rubber deformation. In 1940, Mooney (1940) proposed a more general equation which satisfies Hooke's law for a sheared elastomer. Mooney's equation for the strain energy function is
(8-27) A9
Aa
This result is consistent with a variety of experiments on rubber over a fairly wide range of deformation. A few years later, Rivlin (1948) developed this model much further. According to Rivlin, the stored energy could be written as a general function of the simplest symmetric invariants of the deformation. These invariants are (8-28 a) 12 ~~ A^ A 2 T
_
;2 ;2
1 3 — Ax
1 A^ A 3 ;2
A2A3
1 A 2 A3
(8-28 b) (8-28 c)
and since, for rubber at ordinary pressures, / 3 is approximately equal to unity
370
8 Rubber Elasticity
(due to its small compressibility), Mooney's equation is the linearized general expansion of the free energy in these invariants,
strain proposed that W=1£ci
(8-31 a)
i
W=I.aij(Ii-3)i(I2-3y
(8-29)
In fact, Rivlin and his associates often use the Mooney equation, Eq. (8-27), in interpreting experiments, though they recognized that higher terms are often required to provide a fit to experiments over a wide range of deformation. The Mooney-Rivlin analysis is a purely phenomenological model, and is not based on molecular theory. The lack of a molecular basis does limit the usefulness of the analysis, but nevertheless, the equations provide an interesting framework for treating elastomeric deformation. It is implicit in the model that a particular set of constants used for a given material is expected to apply regardless of the mode of deformation. Experiments, for example, in uniaxial deformation should yield the same Cx and C 2 constants in Eq. (8-27) as those obtained in shear, torsion, or biaxial stretching. 8.3.4 Other Phenomenological Models of Rubber Deformation Valanis and Landel (1967) proposed that the strain energy function could always be decomposed into a sum of functions of the deformation in each of the three principal directions according to W= w ^ J + w(A2) + w(X3)
(8-30)
This is consistent with the Mooney equation, Eq. (8-27), though it is not in accord with the Rivlin generalization, Eq. (8-29). The Valanis-Landel hypothesis is also consistent with other proposals of a number of different researchers. Ogden (1972), in an effort to account for rubber deformation over a broad range of
(Pi^A) = A1 + A 2 + A3
(o-31 D)
In this formulation, the quantities at are constants but are not necessarily integral, and are not necessarily the same for different materials. A two-network model of the Ogden type was adopted by Tschoegl and Gurer (Tschoegl and Gurer, 1985; Gurer and Tschoegl, 1985). It was applied to bulk elastomers and also to swollen networks. Agreement between experiment and theory was achieved by determining the parameters at at low strains, and applying the result to data collected over a much wider range of strain. Tschoegl and Gurer found ax equal to two and a2 equal to 0.34 for the experiments they analyzed.
8.4 Experimental Procedures in Rubber Deformation Rubber deformation can be studied in many ways. The procedure most commonly used is that of uniaxial stretching. In that case, the strain energy function of a Gaussian network reduces to (8-32) The strain energy function for uniaxial deformation in the Mooney equation is
(8-33) The stress is given by k
TV
- x
(8-34)
8.4 Experimental Procedures in Rubber Deformation
It follows from Eq. (8-34) that a plot of a/(A2-l/X) versus 1/1 yields both C1 and C 2 from which the modulus is obtained as a function of deformation. One common experimental technique is to measure the response of an elastomer to homogeneous biaxial deformation. In such experiments, it is customary to vary the deformation ratios in the two principal directions. Two cases are of particular importance. One of these is the pure shear experiment in which Ax is varied, k2 is kept equal to unity, and A3 adjusts to the constant volume requirement, /L3 = l/^i. The second important case is that of equal biaxial deformation where kl = X2 and A3 = l/Af. In the case of pure shear, the Mooney equation yields (8-35 a) (8-35 b)
= 2(C1-C2)
where u1 and <x2 a r e the principal stresses. The equal biaxial stretch result is
A? '
i = 1,2
(8-36)
Simple shear is not a pure strain. In simple shear X2 equals unity, X3 = l/X1 and Mooney's equation gives (8-37) The shear gradient y, is Xx — 1/Al5 from which it follows that W = (C1 + C2)y2. The corresponding shear stress, ak equals The strain energy function has been used in other deformation modes for calculation of stresses, which are worked out as needed. The stresses given in the formulae presented above are the stresses induced by
371
applied tensions or pressures. In reality, the principal stress in any direction is equal to that imposed by the applied forces minus the pressure. In experiments carried out at high pressures or at extremely low applied forces, the pressure should be included in the stress equations. 8.4.1 Hysteresis The analysis of elastomeric deformation discussed above is based on relations between stress and strain of a material in equilibrium. In fact, in many experiments in which applied force and sample deformation are measured, the forces required to produce a particular deformation change with time, and depend on whether the successive increments in stress are positive or negative. In order to obtain equilibrium results, it is necessary that experiments be carried out slowly or by raising and lowering temperature successively. The lag in attaining equilibrium is more common with bulk elastomers than with swollen systems, and it is often difficult to establish when a stretched rubber is at equilibrium. One of the problems is that some rubbers crystallize slowly during deformation. In some materials, oxidation is hastened by applied stress, and permanent chemical changes may obscure stressstrain equilibrium of the specimen. In fact, many experimental studies have been reported where there is doubt that equilibrium was attained, and this problem often fuels controversies over theoretical models of rubber deformation. 8.4.2 Some Experimental Investigations of Elastomeric Deformation
Changes of internal energy of rubber are small at low deformation, but may become large when deformations are large as is
372
8 Rubber Elasticity
commonly observed for natural rubber and other crystallizable elastomers. Large changes of internal energy in stress-strain behavior of rubber at high deformation were observed by Wood and Roth (1944) as shown in Fig. 8-4, while experiments on GR-S, a polybutadienepolystyrene rubber which does not crystallize, exhibit relatively modes changes in internal energy up to high deformations (Roth and Wood, 1944). In these investigations, both hysteresis and permanent distortion were evident in some examples. The elementary Gaussian statistical theory of rubber elasticity provides a basis for analysis of rubber mechanics at small deformations, but when deformations become substantial, the corrections to that model become significant. Experiments
8 Figure 8-4. Stress-deformation measurements on natural rubber, (a) o versus X\ (b) —T(da/dT)k versus L (Wood and Roth, 1944.)
with both bulk and swollen rubber by Gumbrell, Mullins and Rivlin (Gumbrell etal., 1953) provide an excellent example for the need for higher terms in the stored energy function. The Gaussian theory leads to a stored energy function given by Eq. (8-26) with the constant C equal to Q VR T/(2MC) for the fixed junction model and [QVRT/ (2MC)] [1 -(2//)] for the phantom network. In real rubbers dangling ends of chain molecules are not crosslinked, and thus are not part of the network. In addition, chains which are not crosslinked may be incorporated in the rubber matrix. Owing to these effects, the constant C obtained experimentally would be expected to be lower than the theoretical results of the models. Contrary to this expectation, it is often found that C or the front factor, as it is known, is larger than theoretical predictions. In the investigation of Dossin and Graessley (1979) on polybutadiene crosslinked by radiation, experiments were conducted on both bulk and swollen specimens. The data were analyzed by the Mooney equation, and the authors found that the constant Cx of Eq. (8-27) accounts for the entire chemical network and half of the topological entanglement contribution to the modulus. The constant C 2 accounts for the other half of the contribution from entangled chains. According to Dossin and Graessley, fluctuations of crosslink junction points are largely suppressed and the entanglement contribution to the modulus is predominant for those networks crosslinked in bulk. Ferry and co-workers (Carpenter et al., 1977; Ferry and Kan, 1978) carried out stress-strain studies on elastomers crosslinked by radiation while in a state of uniaxial extension. The crosslinking was performed in most cases below the glass transition temperature. After crosslinking and
373
8.4 Experimental Procedures in Rubber Deformation
rewarming to room temperature, the specimen length of the material in this so-called state of ease is greater than that of the undeformed material before crosslinking, but less than that of the stretched material during crosslinking. The authors explain this result as a balance between forces of the chemical network with those of the entanglement network. The model treats the chemical network with simple Gaussian statistics, and the entanglement network as one of the Mooney type. In the study of Carpenter etal. (1977), crosslinking was performed by y-radiation, and the initial elongation during crosslinking was varied from 1.2 to 2.8. The elongation in the state of ease after rewarming varied from 1.05 to 1.48. While most crosslinking experiments were carried out below Tg, those carried out above Tg sometimes showed a smaller fraction of trapped entanglements. Radiation times were varied. Some data are assembled in Table 8-1. A considerable body of experimental studies of deformation of rubbery networks has been carried out by Mark and associates, and they have analyzed the results of these experiments generally by application of the Flory-Erman theory. One good example is given in Llorente and Mark's investigation of end-linked polydimethylsiloxane (Llorente and Mark, 1980). Experiments on uniaxial extension and compression were carried out, and data plotted to yield constants of the Mooney equation. The modulus as a function of elongation is obtained, and therefore, the parameters of the theory. In this study, the authors estimate the degree to which the networks approach the fixed junction model in the unstretched network, and how closely the system approaches the phantom network at high elongation. Table 8-2 contains some of those results.
Table 8-1. Data collected on poly(butadiene) crosslinked in a state of strain. i 0 is the initial elongation at which the specimen was crosslinked. Xe is the elongation in a state of ease after crosslinking and rewarming. Tc is the temperature at which the specimen was crosslinked. (Results of Carpenter et al., 1977.)
1.400 1.565 1.620 1.749 1.242 1.275 1.306 1.309 1.508 1.826 1.839
K
TC(°C)
1.108 1.202 1.185 1.251 1.078 1.106 1.107 1.127 1.193 1.231 1.240
-15 -15 -15 -15 -10 -10
-io -10 -10 -10 -10
11.888 11.993 :1.782 1L.345 11.702 11.478 :>.499 11.330 11.553 \1.774
K
Tc (°C)
1.355 1.346 1.482 1.148 1.296 1.187 1.527 1.138 1.254 1.366
-10 -10 -10 -5 ^
-5 -5 0 0 0
Table 8-2. Stress-strain behavior of end-linked polydimethylsiloxane networks. / is network functionality. Ay is obtained from the constant 2Ct of the Mooney equation applied to experiment. If the Erman-Flory theory applies, Av should equal 1 — 2//. A'y is obtained from 2C 1 +2C 2 using the Mooney equation, and according to Erman and Flory should equal unity. (Data from Llorente and Mark, 1980.) /
1-2//
A.
A'
3.0 3.0 3.0 4.0 4.0 4.0 4.6 5 5 5 6 6 8 8 11 11 37 37
0.33 0.33 0.33 0.50 0.50 0.50 0.56 0.60 0.60 0.60 0.67 0.67 0.75 0.75 0.91 0.91 0.97 0.97
0.88 0.85 0.79 0.94
1.32 1.32 1.37 1.34 1.37 1.32 1.42 1.39 1.09 1.29 1.43 1.39 1.43 1.43 1.44 1.48 1.17 1.16
0.82 0.96 1.07 1.07 0.87 1.09 1.16 1.15 1.33 1.30 1.39 1.42 1.14 1.69
374
8 Rubber Elasticity
Thirion and Weil (1983) examined equilibrium stresses on a rubber prepared by crosslinking ds-polyisoprene with dicumyl peroxide. The data were analyzed in two ways, one by use of the Mooney equation, and the other by a variant of the tube model developed by Ball et al. (1981), in which slip-links characterize the restrictions of the polymer chains in the tube. A slipping coefficient needed in the theory was taken from earlier studies on deformation in pure shear. The authors come to the conclusion that the analysis of Ball et al. works well for rubber elasticity. The experiments in this study are in good accord with earlier work of Ferry and Kan (1978) and also with that found by Dossin and Graessley (1979). Tschoegl and Gurer (Tschoegl and Gurer, 1985; Gurer and Tschoegl, 1985) analyzed elastic deformation of a variety of elastomers using a two-term expression for the strain energy function. The first term was appropriate for a network of Gaussian chains, and the second was of the Ogden type, C2/[rn2(A™ + ^ + ^)]
with m = 034.
The coefficients were obtained using uniaxial extension measurements at small deformation, and then applied to uniaxial measurements over a much broader range of deformation. Elastomers studied included natural rubber, styrene-butadiene rubber, polydimethylsiloxane, poly-(ethyl acrylate) rubber swollen in bis(2-ethoxyethyl) ether, polybutadiene, and butyl rubber. Some experiments were carried out on the rubber in bulk, and others on specimens prepared by swelling in a solvent. The term quadratic in the quantities kt in the strain energy, according to Tschoegl and Gurer, arises from chemical crosslinks, the other term contains the contribution from the entanglement network.
8.4.3 Some Swelling Experiments Gee (1946) studied the maximum swelling of rubber in certain swelling agents, and also studied the vapor pressure of some of these swollen rubbers. From this investigation and earlier work, he obtained a set of X values for several liquids. Results are shown in Table 8-3. The numbers obtained from measurement of the maximum swelling ratio agreed well with that obtained from vapor pressure measurements in most cases. Brotzmann and Eichinger (1983) investigated the vapor pressure of cyclohexane above swollen crosslinked polymer and uncrosslinked solutions of polydimethylsiloxanes. The data are presented in a graph of kln(pc/pu) versus A2, p c , being the vapor pressure of the swollen crosslinked polymer, and pu, the vapor pressure above the uncrosslinked material at the same solvent concentration. Data are found in Fig. 8-5. The maximum in the figure is not found for the unmodified Gaussian model of rubber, but a maximum, less steep than that in the measurement, is given by the Flory-Erman analysis. No reasonable choice of parameters in the Flory-Erman
Table 8-3. Values of /, the Flory interaction parameter for natural rubber. (Data from Gee, 1946.) Swelling agent
CC14 CHC13
cs2
Benzene Toluene rc-Propyl acetate Ethyl acetate MEK Acetone
X Swelling
Vapor pressure
0.29 0.34 0.425 0.395 0.36 0.62 0.78 0.94 1.37
0.28 0.37 0.49 0.43 0.43-0.44
8.5 Orientation of Chain Segments
6.0 -
(a)
4.0 2.0
1
00
1
1
1
6.0 o Q.
(b)
4.0 2.0
1
00
1
1
1
6.0
1 (c)
4.0 2.0
1 1.0
1 2.0
1^-—1—-t— 3.0
375
where chain extension is limited such that the end-to-end distance of a chain is less than 30% of the length of the fully extended chain. For greater chain extensions, deviations from the Gaussian model become considerable. A careful investigation of the statistics of extended chains was put forth by Kuhn and Griin (1942). The polymer chain contains n segments, each of length b, with a mean square end-to-end distance, (j2y = nb2 in the unstretched state. The angle made by the zth segment initially with a particular direction, later to be identified with the stretch direction, is 9f. The fraction of segments at a polar angle 9 and azimuthal angle cp measured with respect to the stretch direction is given by
4.0
Figure8-5. ^ln(p c /p u ) versus X2 for swollen polydimethylsiloxane gels and solutions. Three specimens: Specimen (a): Mc = 11 300, / = 3; specimen (b): Mc = 11 300, / = 4; specimen (c): M c = 18 500, / = 4. (Brotzmann and Eichinger, 1983.)
dn n
g.38 a )
4 7i sinh P
where r/(n b) = 5£ (j5), and equivalently p = <£~x{rl(nb)\ Z£{p) is the Langevin function defined by if (j8) = coth(jS) - l/p
theory, however, can provide a quantitative fit to the experiment. Brotzmann and Eichinger have stated their opinion that part of the discrepancy between theory and experiment arises from the assumption in the theoretical development that the elastic and the swelling contributions to the free energy are additive, an assumption which has been widely adopted by theorists, but which remains unproved.
8.5 Orientation of Chain Segments 8.5.1 The Statistics of Highly Stretched Chains
A Gaussian model of polymer chain statistics may be applicable to polymer molecules or polymer chains in rubber
(8-38 b)
and ^?~1(r/(nb)) is the inverse function. The probability density for r is given by
-Ol. ,
cexp 1 - p- , .
b
\smh P) J
(8-39)
which reduces to the Gaussian result [Eq. (8-5)] for small r/(n b) by using the leading terms of the expansion of /} in terms of r/(nb). Roe and Krigbaum (1964) working with the Kuhn-Griin equations, and a closely related model of Treloar (1954), calculated values of w(cos0), which is the fraction of segments between 9 and (9 + dO), for chains of 50 and 150 segments (3-10 monomers), for several values of L Their results are shown in Fig. 8-6. Roe and Krigbaum de-
376
8 Rubber Elasticity
veloped a general formula for
n=50 1.0
(8-40 a)
^s:li-a where F2 is given by
0.8
" X-4 0.6
\
36
"" X-2 ^ ^ ^ O ^ :ir -^
0.4
-
0.2
-
2
5n 875 n 108 6125 n3
X 3
(8-40 b) (a)
I 60°
30°
90°
The orientation of polymer segments in a stretched rubber can be calculated from this analysis and is of interest in studies of optical birefringence, fluorescence, and NMR quadrupole splitting.
n = 150
8.5.2 Optical Birefringence of Elastomers
1.0 ^ v .
51=10
0.8
0.6
k^
0.4
0.2 I 30°
I 60°
90°
0 Figure 8-6. Plots of w (cos 6) versus 0 for (a) n = 50, A = 2, 4, 6; and (b) n = 150, A = 2, 4, 6, 8, 10. (Roe and Krigbaum, 1964.)
When a transparent rubber is uniaxially stretched, it becomes birefringent.The polarizability of a chain segment is normally different along the segment axis from the polarizability perpendicular to the segment axis. After stretching, the segmental orientation becomes anisotropic, and the overall polarization perpendicular to the stretching direction differs from that parallel to the stretch. This is manifested in a difference in refractive index in these two principal directions. The birefringence is calculated from the polarizability difference by use of a suitable equation. The Lorentz formula is normally adopted for this purpose. The optical birefringence ni — n2 for a Gaussian network takes the form 2 71 ( * '
— n7 =-45 (8-41)
8.5 Orientation of Chain Segments
<xl5 and a 2 are the polarizabilities of the polymer segment in the parallel and perpendicular directions respectively, and iVv is the number of polymer chains per unit volume; n is the mean refractive index of the rubber plus swelling agent. The uniaxial true stress in a swollen Gaussian network is also proportional to A2 —A"1, and
Co = n1—n2
(8-42 a)
where C, the stress-optical coefficient is given by
°" 45
(8-42 b)
HkBT
This simple stress-birefringence relation does not apply at extremely high orientations for which the Kuhn-Grun model of chain deformation is required. The optical birefringence for a highly stretched polymer depends on higher terms in the deformation, and is given by _2TT Hl
~n2~j5~
Stress optical coefficients have been determined for a number of different elastomers. It is generally noted that the values obtained experimentally for the constant C in Eq. (8-42 a) are higher than expected for the unswollen rubber on theoretical grounds, but the magnitude is very much reduced by swelling. The reduction is greatest if the solvent molecule is not very anisotropic. In Table 8-4, data collected by Ishikawa and Nagai (1969) show that C decreases upon swelling. Similar results were obtained by Gent (1969), who analyzed data collected earlier by Saunders (1956, 1957). As part of their analysis of their own birefringence studies of poly(l,4butadiene), Fukuda, Wilkes and Stein (Fukuda et al., 1971) discuss the possibility that the Lorentz-Lorenz equation is an inaccurate representation of the relation between birefringence and polarizability. 8.5.3 Fluorescence of Labeled Elastomers
(f
30/?
70 nz
377
(8-43)
In many experiments on birefringence of elastomers, deformation is not very great, and the Gaussian model is adequate. Simultaneous measurements of stress and birefringence lead to values of the stressoptical coefficient. The instrument of choice for the measurement of birefringence is a Babinet compensator. The data are often plotted with Mooney type equations for both stress and birefringence (8-44 a) (8-44 b)
In an elastomer, the orientation of chain segments can be measured by putting a fluorescent group in the polymer chain, and by observing the polarization of fluorescence in a stretched sample. The theoretical point of departure is taken from the Roe-Krigbaum analysis (Roe and Krigbaum, 1964) which yields the orientation of chain segments with respect to the stretchTable 8-4. Polarizabilities calculated from stress-optical coefficients of 1,4-ds-polybutadiene radiation crosslinked by 12 Mrad electrons. (From Ishikawa and Nagai, 1969.) OL± - a 2 ( 1 0 ~ 3 0 m 3 ) 7.5 6.1 5.5 5.9 4.9
Vi
1 0.15 0.12 0.14 0.21
(benzene) (CC14) (tetralin) (decalin)
378
8 Rubber Elasticity
ing direction for model chains, as a function of the number of segments per chain and the degree of deformation. In the experiment, the polarization of fluorescence is measured as a function of elongation. The mean second moment of the cosine of the angle between the transition moment of the fluorescent group and stretching direction is given by
(8-45 a) which, if the Gaussian model for stress is valid, yields (5QkBT(p2)
(8-45 b)
At high elongations, it is not expected that Eq. (8-45 b) would apply. In order to obtain the mean orientation of polymer segments with respect to the stretching direction, it is necessary to know the orientation of the fluorescent group with respect to the segments, and the direction of the fluorescent excitation relative to the molecular axis of the fluorophore. This is more of a problem if the fluorescent group can vary in conformation relative to the chain. Jarry and Monnerie (1978,1980) studied orientation in polyisoprene labeled with dimethyl anthracene in the backbone of the polymer chain. Some measurements were performed on dry networks and others on networks slightly swollen (
tends to disappear. Examples of results from the studies of Jarry and Monnerie (1979) are presented in Fig. 8-7. 8.5.4 Quadrupole Splitting in NMR Spectra of Swelling Agents by Oriented Rubber
An elastomer swollen in a deuterated solvent of high symmetry and stretched in an NMR spectrometer at right angles to the steady magnetic field will exhibit deuterium NMR spectral lines which are split. The stretched elastomer creates an anisotropic electric field gradient in the swelling agent which causes the splitting. The governing equation is Av =
:— P2 (cos Q) P2 (cos 9{) P2 (cos y) (8-46) where P2 is the second Legendre polynomial, Av is the difference in frequency arising from quadrupole splitting, e2 q Q/h, the gradient factor, is of the order of 200 Hz for a C-D bond, Q is the angle between the 2
0
5
10 15 o in MPa Figure 8-7. Second moment,
8.6 Small Angle Neutron Scattering (SANS) of Elastomers
379
1.0
0.8
lay N
0.6
/
>
0.4
0.2
^*TI
1
I
1
1
1
8
10
12
14
16
10 0.3
0.5 0.6 0.70.80.91.0
Figure 8-8. Quadrupole splitting of deuterobenzene in oriented rubber versus X2 — X~x. (a) (p2 = 0.93; (b) cp2 = 0.80; (c) cp2 = 0.52; (d) (?2 = 0.32. (DeLoche and Samulski, 1981.)
Figure 8-9. Quadrupole splitting of deuterobenzene in rubber. Av/(A2 — X~ *) versus q>2. Slope of line = 4/3. (DeLoche and Samulski, 1981.)
stretching axis and the magnetic field, 0t is the angle between the ith C - D bond and the molecular axis of the elastomer, and y is the angle between the stretching axis and the C-D bond. Experimental splittings of the order of 100 Hz can be measured, thus allowing determination of orientations with P2 less than 0.01. An example, found in the splitting of hexadeuterobenzene in swollen stretched sulfur-crosslinked polyisoprene, was obtained by DeLoche and Samulski (1981). Some results are exhibited in Fig. 8-8. Figure 8-9 shows how Av/{l2~rl) varies with polymer concentration in a swollen specimen.
8.6 Small Angle Neutron Scattering (SANS) of Elastomers The SANS method (see Chap. 20, Vol. 2B of this Series) for studying the shape and size of polymer chains has been applied to measurement of polymer chains which make up elastomeric networks. The coherent elastic scattering is proportional to the Fourier transform of the segmental distribution of labeled chains in the network. The labeling is achieved by replacing ordinary hydrogen with deuterium in some of the chains which make up the network. The intensity of coherent elastic scattering is given by q)
(8-47)
380
8 Rubber Elasticity
where A and B are known constants depending on polymer molecular weight, concentration, and scattering contrast factors. Ss(q) is a single chain scattering function, ST(q) is known as the total scattering function and contains both single chain and interchain contributions to the scattering. The interchain contributions are often referred to as intermolecular, though they arise from pairs of monomer units on different chains on the same molecular network. For a polymer chain, Ss(q) may be written as (8-48) UJ
Here r[j is a vector connecting monomers i and j on the same polymer chain, q is the scattering wave vector, defined by q = (2 n/X) (k — k0), where k0 is a unit vector in the direction of the incident neutron beam, k, a unit vector in the direction of the scattered beam, and X is the wavelength of the neutrons, normally between 0.3 nm and 1.5 nm in this sort of experiment. The magnitude of q is given by \Q\=
(8-49)
Here rtj is the vector connecting monomer i with monomer j but now, these monomers
may be on the same chain or on different chains. The average is taken over all possible arrangements of the chains in the specimen. In a bulk network containing no solvent, and in which labeled and unlabeled chains are chemically identical, the constant B of Eq. (8-47) equals zero. In a swollen network, B is, in general, not equal to zero. B is also nonzero if the labeled and unlabeled chains are chemically different, or if they differ in molecular weight. It is possible to cause B to vanish by adjusting the fraction of labeled chains, or by adjusting the deuterium-hydrogen ratio in the solvent. These are procedures occasionally adopted to insure that the coherent elastic scattering arises from single chain scattering only. For an isotropic sample, Ss(q) may be written as a power series in q (8-50) Rg is the radius of gyration of a polymer chain, and can be determined from a low q plot of [/(g)]"1 versus q2 whenever ST(q) = 0. The measured scattering intensity is given by (8-51) K is an instrumental constant which can be determined independently. Rg can be obtained by plotting I(q)-KBST(q)~1 versus q2 if ST(q) is nonvanishing. By determining how Rg varies with solvent concentration and molecular weight of the polymer chains, it becomes possible to determine changes in chain geometry under different conditions of network preparation and swelling. In a stretched network, Ss(q) may be written as
Ss(«) = l - V * +
(8-52)
381
8.6 Small Angle Neutron Scattering (SANS) of Elastomers
where R^ is the apparent radius of gyration measured at an azimuthal angle
&/
70 c
y^ 60 a
(8-53a)
\--A + =,
cr 50
'/<
^ "
(8-53 b)
a
*''
c
,-
*"*** — —Cr**"
40
a
(8-53 c) l
In the nonstretched swollen network [Eq. (8-53 a)], X is the linear swelling ratio, while in Eqs. (8-53 b) and (8-53 c) k is the uniaxial stretching ratio. In the fixed junction model, the calculated values of R2 are obtained by dropping the quantity 2 / / in Eqs. (8-53 a, b,c). There have been many SANS studies of elastomeric networks, and these have been directed, in part, to discovering which of the molecular models, if any, of an elastomeric network applies to the material. The earliest published experiments on SANS of rubbery networks are those of Benoit et al. (1976) on anionic polystyrene networks crosslinked with divinylbenzene and swollen in selected solvents. The polymerization procedure was such that the functionality of these networks was not determinable. The polymerization was carried out in 10% solution, the radius of gyration of the uncrosslinked polymer was measured as well as the radius of gyration of the chains in the unswollen network. Values of Rg versus degree of swelling for three networks are shown in Fig. 8-10. On the same figure, calculated values of Rg from a phantom network model with / = 4 is also given.
I
1
2.5 3.0 2.0 X Figure 8-10. SANS measurement of Rg of labeled chains in swollen polystyrene networks. Rg versus X for three networks, all with M c = 26000 in different solvents: (a) cyclohexane, (b) hydrofurfuryl alcohol, and (c) benzene. The theoretical prediction for a phantom network with / = 4 is shown in curve (d). (Benoit etal., 1976.) 1.0
1.5
Calculations of the scattering function of labeled polymer chains in a deformed network were obtained by Pearson (1977), by Warner and Edwards (1978) and by Ullman (1979, 1982). These calculations have been used for interpretation of experimental studies by many researchers. Duplessix (1975) and Bastide, Duplessix, Picot, and Candau (Bastide etal., 1984), have measured SANS spectra of crosslinked polystyrene networks swollen in benzene and subsequently partially deswollen osmotically. Some of their results are shown in Fig. 8-11. Beltzung and co-workers (1982, 1983, 1984) have applied SANS to an investigation of polydimethylsiloxane networks. These networks were prepared by crosslinking linear deuterated chains with chains of the protonated species using vi-
8 Rubber Elasticity
382
(b)
90 ^- '—
80 —
•E DC
^~~
1.3
_
Fixed junction/ •'if Phantom ° o °
1.2
70
1.1
60
1.0
US
o
o
X Phantom
0.9
50
**''''
1.0
I
I
1.2
1.4
I 1.6
o
1.8
-o—
rx,
2.0
.o
2.2
40 30 20 10 I
I
I
10
15
Figure 8-12. Ratio a of the radius of gyration parallel to (l?(|) or perpendicular to (R±) the stretch direction to the unperturbed radius of gyration {Rg) in stretched polydimethylsiloxane networks with various M c . •: M c = 3 x l 0 3 ; +: Mc = 6 x 103; x: Mc = 104; o: Mc = 2.5 x 104. (Beltzung et al., 1984.)
X3 Figure 8-11. Radius of gyration of chains in osmotically deswollen polystyrene networks versus X3. Curve (a) phantom network; curve (b) fixed junction model; data: n, •, •. experiments of Duplessix (1975).
nyl or allyl crosslinking agents designed to create either tetrafunctional or hexafunctional crosslink junctions. Radii of gyration of the uncrosslinked chains as well as radii of gyration of the chains in the crosslinked but unswollen network were determined. Crosslinking caused no change in the scattering of the unswollen rubber. i?t, values of the stretched network were equal to or less than that predicted by the phantom network, this being a function of the molecular weight of the chains. R± was close to that predicted for the phantom network. Figure 8-12 shows some results. These same networks, swollen in cyclohexane showed that Rg increased with molecular weight between crosslinks in accord with phantom network predictions. It also turned out that Rg values in these swollen networks varied with molecular weight in the same way as do Rg values of dilute solutions of uncrosslinked polymer in the same solvent. Some data are shown in Fig. 8-13.
4.0
4.5
5.0
logM z Figure 8-13. Radius of gyration of chains of polydimethylsiloxane networks swollen in cyclohexane (*), and Rg in dilute solutions of polydimethyl siloxane in cyclohexane (•), versus Mz (Z: average molecular weight measured by GPC). (Beltzung et al., 1983.)
Tsay and Ullman (1988) studied SANS spectra from polystyrene crosslinked by reaction with benzene meta disulfonylazide. Measurements were performed on both stretched and swollen networks. In these experiments, the molecular weight of the chains before crosslinking was 105, and networks were synthesized to obtain an average of 3, 4, or 5 crosslinks per chain. The results were compared with calculations
8.7 Rubber-Glass and Rubber-Crystal Transitions
made by Ullman (1982) on SANS from polymers with several crosslinks per chain. Plots of R\\/R% and RJR% for stretched networks lie between the fixed junction and phantom models, and are closer to the phantom network at higher elongation (2 = 1.6 to 1.9). The chain dimensions were changed much more by swelling in toluene than by swelling in cyclohexane. It was found that Rg(X)/Rsg in toluene was approximately equal to Rg(X)/R\ in cyclohexane where #g stands for radius of gyration in dilute solution. The radius of gyration of a polymer chain is determined from scattering at very low q values. SANS measurements at higher q yield further information on chain shape. These data are often examined by plotting q2 Ss(q) versus q, a technique originally used by Kratky for studies of small angle X-ray scattering of polymers. Bastide, Herz, and Boue (Bastide et al., 1985) investigated SANS from high molecular weight (>10 6 ) deuteropolystyrene crosslinked by y-radiation in cyclopentane solution with lower molecular weight protonated polystyrene. The Kratky plot of the deswollen network with a molecular weight between crosslinks of 35000 with a labeled chain of 2.6 x 106 molecular weight exhibited a broad peak at q&0.15 nm'1 (see Fig. 8-14). The peaks calculated for the crosslinked phantom and fixed junction networks were greater than that found in the experiment. The authors suggest that this peak could be accounted for in part by interchanges in spatial positions of geometric near neighbors with chemical near neighbors upon deswelling. In the Tsay and Ullman experiments, Kratky plots on stretched samples showed a broad peak at q « 0.5 nm" 1 for scattering in the perpendicular direction. Scattering in the parallel direction showed a slight maximum in the same q range.
i
I
383
I
I I I i i 0.4 0.6 1 q in nrrf Figure 8-14. Kratky plot comparing data obtained from a dry gel (A) and an isotropic melt (•) of deuterium labeled polystyrene to calculated form factors for the same deformation ratio: isotropic Gaussian (plain line below); junction affine model, Mmesh = 35000 (plain line above); phantom network model, Mmesh = 35OOO and Mmesh = 50000 (below), dashed lines. (Bastide et al., 1985.)
There have been many other SANS studies on rubbery networks which have not been discussed here. It has not yet been possible to determine from these studies which of the various models of rubber elasticity is most realistic. The Flory-Erman model and model of Edwards and associates have been applied to the SANS problem, but it is not possible at this stage to tell which view of rubber elasticity is most nearly in agreement with SANS measurements.
8.7 Rubber-Glass and Rubber Crystal Transitions When a rubber is cooled, it becomes a hard solid. If the rubber has an irregular chemical structure such as polystyrenebutadiene), the solid is a glass. If the elas-
384
8 Rubber Elasticity
tomer has a regular chemical structure, it may partially crystallize upon cooling. Lightly crosslinked natural rubber is in the latter category. The crystalline structure which is formed on cooling can also be induced by stretching the rubber above the crystalline melting point. Strain induced crystallization of rubber was first shown by Katz using X-ray diffraction (Katz, 1925) on natural rubber. One of the consequences of strain-induced crystallization is an increase in modulus and also in ultimate strength. Crystallinity in an oriented polymer is accompanied by a crystalline X-ray pattern at wide angles and, often, a two-point pattern in the small angle domain. When a polymer crystallizes, its density usually increases substantially since the chains pack more efficiently in the crystal than in the liquid or glassy states. In addition, a sharp endotherm is evidenced during heating since the polymer crystals melt. If a rubber is crystallized in part by stretching, the crystalline fraction can be melted by heating, in which case the modulus decreases and the density drops. The assumption that the volume of a rubber is largely unchanged by stretching, used widely here and in other studies of rubber elasticity, is only valid under circumstances where no crystallization takes place. Rubbers which are normally crystallizable can be prevented from crystallizing by increasing the extent of crosslinking. An increased density of crosslinks reduces the regularity of the chemical structure of the rubber and places severe constraints on the chains, resulting in a serious impediment to the formation of a crystalline structure. Crystallization of rubber without strain has been investigated by many research groups. In one interesting case, crystalline morphology of fractionated, uncrosslinked
natural rubber was studied by Phillips and Vatansever (1987).
8.8 Thermoplastic Elastomers Thermoplastic elastomers (TPEs) are multiblock copolymers, comprising hard segments and soft segments, which microphase separate into hard domains in a soft segment matrix (Legge et al., 1987). The crosslinks in TPE materials are physical in nature arising from aggregation of the hard segments into domains, which physically constrain the rubbery soft segments between them. The hard segment domains can be both glassy or crystalline, depending on the chemical structure of the hard segment. TPE materials such as polystyrene - polybutadiene - polystyrene triblock copolymers, now in large volume commercial use, depend for their elastomeric properties in the anchoring of the rubber chains by the hard segment domains. Polyurethanes, polyesteramides, polyetheramides, and oc-olefin TPE materials all exhibit interesting stress-strain behavior. The anionically prepared PS-P.BPS triblock materials are a useful model system since domain size, shape, and spacing are very well defined. Keller and coworkers (Folkes et al., 1973; Keller and Odell, 1985) have examined the deformation behavior at both low and high strain. For strains below about 50%, affine deformation takes place, but at higher strain levels, complications arise from hard segment domain reorientation and domain breakup. Stevenson and Cooper (1988) studied the changes which occur with deformation of the elastomeric block copolymer of poly(tetramethylene oxide) (PTMO) with poly(tetramethylene isophthalate) (PTMI) or with mixtures of PTMI and poly (tetra-
8.9 Ultimate Strength
methylene terephthalate) (PTMT) as the hard segment. The PTMO block had a number average molecular weight of 1000. The material prepared by transesterification of poly(tetramethylene ether glycol), dimethyl isophthalate, dimethyl terephthalate, and 1,4-butanediol had a molecular weight of 25000 to 30000. The ester block forms crystalline hard segment domains, while the ether block imparts flexibility to the polymer. The weight fraction of the hard segment (PTMI + PTMT) was approximately 60%. Samples were films prepared in a press by heating 30 °C above the melting point and then cooling. The crosslinks in this system are physical in nature, are localized in the hard segment aggregates, and change from glassy to partially crystalline as the system is annealed. Stress-strain measurements during crystallization at room temperature exhibit stresses which decrease markedly with time at a given elongation.
8.9 Ultimate Strength Elastomers tear when subjected to high stresses. The stress required to bring about this mechanical failure depends on a number of factors. The result may depend on viscoelastic response of the rubber, molecular weight between crosslinks, the temperature at which the stress was applied, the chemical makeup of the material, and the existence of flaws in the specimen. Rubbers which crystallize tend to be stronger than materials which remain amorphous upon stretching. In 1967, Lake and Thomas (1967) provided an analysis of strength of elastomers based on stored energy, and how this energy is released when the elastomer is torn. This work was extended by Gent and Tobias (1982) who performed many experi-
385
ments and incorporated results of earlier researchers in their studies. They derived a formula for the threshold tearing strength, a quantity which represents the smallest value of the tensile failure stress that would be measured in a real experiment. In many circumstances, the energy stored in a rubber on the verge of tearing is composed in part of stored energy which would relax viscoelastically if given sufficient time. As a result, the measured strength is greater than that expected in a threshold model. The Gent and Lake-Thomas equations take the form T0 = KM^/2c
(8-54 a)
To is the threshold tearing strength, M c , the molecular weight between crosslinks, and c is the length of a flaw perpendicular to the tearing direction. K is given by K=
3V /2 QU ~
71/2
M1'2
(8-54 b)
Here Q is density, u is the mean energy of a main chain bond, z is the number of main chain monomers per statistical segment, M is the molecular weight of a polymer chain, M o is the monomer molecular weight. K is predicted to be about 0.3 J/m2, measured values tend to be a factor of 3 higher. Threshold tearing strengths for polybutadiene and polyisoprene rubbers lie between 40 and 100 J/m 3 for a series of materials where Mc is bounded by 3 x 103 and 104. Polydimethylsiloxane networks have slightly lower threshold tear strengths. The data on siloxanes include both end-linked and randomly crosslinked materials. A number of measurements were performed on swollen specimens of the siloxanes, and, as expected, X2 To for the swollen specimens equalled To for the bulk network.
386
8 Rubber Elasticity
8.10 References Anthony, R. L., Caston, R. H., Guth, E. (1942), /. Phys. Chem. 46, 826. Ball, R. C , Doi, M., Edwards, S. E, Warner, M. (1981), Polymer 22, 1110. Bastide, X, Duplessix, R., Picot, C , Candau, S. J. (1984), Macromolecules 17, 83. Bastide, X, Herz, X, Boue, F. (1985), /. Phys. 46, 1967. Beltzung, M., Picot, C , Rempp, P., Herz, X (1982), Macromolecules 15, 1594. Beltzung, M., Herz, X, Picot, C. (1983), Macromolecules 16, 580. Beltzung, M., Picot, C , Herz, X (1984), Macromolecules 17, 663. Benoit, H., Decker, D., Duplessix, R., Picot, C , Rempp, P., Cotton, X P., Farnoux, B., Jannink, G., Ober, R. (1976), J. Polymer Sci., Polymer Phys. Ed. 14, 2119. Brotzmann, R. W. Jr., Eichinger, B. E. (1983), Macromolecules 16, 1131. Carpenter, R. L., Kramer, O., Ferry, X D. (1977), Macromolecules 10, 111. de Gennes, P. G. (1971), J. Chem. Phys. 55, 572. De la Condamine, C. M. (1751), Mem. Acad. Roy. Sci. 17, 319. DeLoche, B., Samulski, E. T. (1981), Macromolecules 14, 575. Dossin, L. M., Graessley, W. W. (1979), Macromolecules 12, 123. Duplessix, R. (1975), Thesis, Doctor d'Etat, L. Pasteur University, Strasbourg. Edwards, S. R, Vilgis, T. A. (1987), Reports on Progress in Physics 51, 243. 'Erman, B., Flory, P. X (1982), Macromolecules 15, 806. Faraday, M. (1826), Quart. J. Sci. 21, 19. Ferry, X D., Kan, H. C. (1978), Rubber Chem. Tech. 51, 731. Flory, P. X (1942), J. Chem. Phys. 10, 51. Flory, P. X (1977), /. Chem. Phys. 66, 5720. Flory, P. X, Erman, B. (1982), Macromolecules 15, 800. Folkes, M. X, Keller, A., Scalisi, F. P. (1973), Colloid and Polymer Sci. 251, 1. Fukuda, M., Wilkes, G. L., Stein, R. S. (1971), /. Polymer Sci. A2 9, 1417'. Gee, G. (1946), Trans. Faraday Soc. 42B, 33. Gent, A. (1969), Macromolecules 2, 262. Gent, A. N., Tobias, R. H. (1982), /. Polymer Sci., Polymer Phys. Ed. 20, 2051. Goodyear, C. (1844), U.S. Patent 3633. Gough, X (1805), Mem. Lit. Phil. Soc. Manchester I, 288. Granick, S., Ferry, X D. (1983), Macromolecules 16, 39. Gumbrell, S. M., Mullins, L., Rivlin, R. S. (1953), Trans. Faraday Soc. 49, 1495.
Gurer, C, Tschoegl, N. W. (1985), Macromolecules 18, 687. Guth, E., Mark, H. (1934), Monatsh. Chem. 65, 93. Hancock, T. (1844), British Patent. Huggins, M. L. (1942), /. Amer. Chem. Soc. 64,1712. Huggins, M. L. (1943), /. Phys. Chem. 46, 151. Ishikawa, T, Nagai, K. (1969), /. Polymer Sci. A2 7, 1123. James, H. M. (1947), /. Chem. Phys. 17, 651. James, H. M., Guth, E. (1943), J. Chem. Phys. 11, 455. James, H. M., Guth, E. (1947), /. Chem. Phys. 17, 669. Jarry, X P., Monnerie, L. (1978), /. Polymer Sci., Polymer Phys. Ed. 16, 443. Jarry, X P., Monnerie, L. (1979), Macromolecules 12, 316. Jarry, X P., Monnerie, L. (1980), J. Polymer Sci., Polymer Phys. Ed. 18, 1879. Joule, X P. (1859), Phil. Trans. Roy. Soc. 149, 91. Karrer, E. (1933), Protoplasma 18, 475. Katz, X R. (1925), Chem. Z. 49, 353. Keller, A., Odell, X A. (1985), in: Processing, Structure and Properties of Block Copolymers: Folkes, M. (Ed.). New York: Elsevier. Kramer, O., Carpenter, R. L., Ty, V., Ferry, X D. (1974), Macromolecules 7, 79. Kuhn, W., Grim, F. (1942), Kolloid Z. 101, 248. Lake, G. X, Thomas, A. G. (1967), Proc. Roy. Soc. A 300, 108. Legge, N. R., Holden, G., Schroeder, H. E. (1987), Thermoplastic Elastomers — A Comprehensive Review. Munich: Hanser Publishers. Llorente, M. A., Mark, X E. (1980), Macromolecules 13, 681. Meyer, K. H., von Susich, G., Valko, E. (1932), Kolloid Z. 59, 208. Mooney, M. (1940), /. Appl. Phys. 11, 582. Ogden, R. W. (1972), Proc. Roy. Soc. A 326, 565. Pearson, D. S. (1977), Macromolecules 10, 696. Phillips, P. X, Vatansever, N. (1987), Macromolecules 20, 2138. Rivlin, R. S. (1948), Phil. Trans. Roy. Soc. A 241, 379. Roe, R. X, Krigbaum, W. R. (1964), /. Appl. Phys. 35, 2215. Roth, F. L., Wood, L. A. (1944), J. Appl. Phys. 15, 749. Saunders, D. M. (1956), Trans. Faraday Soc. 52, 1414, 1425. Saunders, D. M. (1957), Trans. Faraday Soc. 53, 860. Stevenson, X C , Cooper, S. L. (1988), Macromolecules 21, 1309. Thirion, P., Weil, T. (1983), Polymer 25, 609. Thomson, W. (1857), Quart. J. Appl. Math. 1, 57. Treloar, L. R. G. (1954), Trans. Faraday Soc. 50, 881. Tsay, H. M., Ullman, R. (1988), Macromolecules 21, 2963. Tschoegl, N. W, Gurer, C. (1985), Macromolecules 18, 680.
8.10 References
unman, R. (1979), J. chem. phys. 7i, 436. Ullman, R. (1982), Macromolecules 15, 1395. Valanis, K. C , Landel, R. F. (1967), /. Appl. Phys. 38, FF
2197
Warner, M., Edwards, S. F. (1978), /. Phys. A. 11, 1649. Wood, L. A., Roth, F. L. (1944), J. Appl. Phys. 15, 781.
387
General Reading fe
Treloar, L. R. G. (1975), The Physics of Rubber Elasticit y- O x f o r d : Clarendon Press.
9 Viscoelastic and Rheological Properties Masao Doi Department of Applied Physics, Faculty of Engineering, Nagoya University, Nagoya, Japan
List of 9.1 9.2 9.2.1 9.2.2 9.2.3 9.2.4 9.2.5 9.2.5.1 9.2.5.2 9.2.5.3 9.3 9.3.1 9.3.2 9.3.3 9.3.4 9.4 9.4.1 9.4.2 9.4.3 9.4.3.1 9.4.3.2 9.4.4 9.4.4.1 9.4.4.2 9.4.4.3 9.4.4.4 9.4.4.5 9.5 9.5.1 9.5.2 9.5.3 9.5.4 9.5.5 9.5.6
Symbols and Abbreviations Introduction Linear Viscoelasticity Significance of Linear Viscoelasticity Shear Flow Stress Relaxation Boltzmann's Superposition Principle Parameters and Functions Characterizing Linear Viscoelasticity Steady Shear Flow Oscillatory Experiments Creep Experiments Characteristics of the Viscoelasticity of Polymeric Liquids Time-Temperature Superposition Rule Behavior of Relaxation Modulus Linear Polymers with Narrow Molecular Weight Distribution Effect of Molecular Weight Distribution and Branching Non-Linear Viscoelasticity Stress Tensor Basic Principles in Continuum Mechanics Non-Linear Viscoelasticity in Rheometrical Flows Shear Flow Uniaxial Elongational Flow Constitutive Equation Second Order Fluid Convective Maxwell Model Integral Form of the Convective Maxwell Model Non-Linear Convective Maxwell Model Single-Integral Constitutive Equation Molecular Theory I: The Rouse Model Polymer Motion in Concentrated Systems Basic Equation of the Rouse Model Normal Coordinates Self-Diffusion and Segmental Motion in Equilibrium Molecular Expression for the Stress Tensor Constitutive Equation for the Rouse Model
Materials Science and Technology Copyright © WILEY-VCH Verlag GmbH & Co KGaA. Allrightsreserved.
391 393 393 393 394 395 396 397 397 397 398 399 399 399 401 403 404 404 405 406 406 408 408 409 409 409 410 410 410 410 411 413 413 414 414
390
9.6 9.6.1 9.6.2 9.6.3 9.6.4 9.6.5 9.6.6 9.6.7 9.6.8 9.7
9 Viscoelastic and Rheological Properties
Molecular Theory II: The Tube Model Characteristics of the Polymer Motion in Strongly Entangled Systems . . . . Reptation and Fluctuations Reptation Model Self-Diffusion Constant Stress Relaxation Fluctuations Constitutive Equation Effect of Branching and Molecular Weight Distribution References
415 415 416 418 418 419 421 422 423 424
List of Symbols and Abbreviations
391
List of Symbols and Abbreviations B (t, t') E (£, f) f{P)
Finger-strain tensor deformation gradient tensor force exerted on the material below the plane j8 by the material above the plane /? G (t) relaxation shear modulus G' (co), G" (co) storage a n d loss modulus GN plateau modulus Go instantaneous relaxation modulus H (T) relaxation spectrum J(t) creep compliance Je steady state compliance kB Boltzmann constant M molecular weight Mc characteristic molecular weight Me molecular weight between the entanglement ML, M s molecular weight of long and short polymers Mx molecular weight of a polymer segment between the cross link Nx (y) first normal stress difference Qf^l tensor relating the stress and the formation tensor E R gas constant Rg center of mass of the molecule Sap orientational order parameter of a tube segment T torque T temperature Tg glass transition temperature t time U (t) orthogonal tensor representing the rotation of a frame u(s, t) unit vector tangent to the tube axis v velocity W elastic potential Z number of entanglement points per chain M/Me y (t) y (t) e & C rj rj (y) rj0 %(&) Y\ + (t, y) rj* (co), rj' (co)
shear strain shear rate elongation strain (Hencky strain) elongational rate friction constant of a Rouse bedd viscosity steady state viscosity zero shear rate viscosity elongational viscosity shear stress growth function complex viscosity
392
9 x xittP) k Q, QS a aE oa(i T rd Te TR Q co
9 Viscoelastic and Rheological Properties
gap angle velocity gradient tensor velocity gradient tensor evaluated at time t, at a material point P dimensionless material constant density at temperature T and Ts stress tensor elongational stress a,/? component of the stress tensor relaxation time reptation time relaxation time Rouse relaxation time angular velocity of the cone angular frequency
9.2 Linear Viscoelasticity
9.1 Introduction This chapter deals with mechanical properties of polymers in the liquid state. When a plastic is heated or when it is dissolved in a solvent, it becomes a liquid. The liquid state of polymers, generally called polymeric liquids, have unusual mechanical properties. Firstly, polymeric liquids have elasticity. This is demonstrated by chewing gum. Chewing gum is a liquid since it does not have any macroscopic equilibrium shape and can be molded into any shape. For slow deformation, chewing gum flows like usual liquids and shows viscosity. On the other hand, for quick deformation, it responds like an elastic material: for example, if we stretch the gum and release quickly the gum shrinks as if it remember the original shape. Thus chewing gum has both viscosity and elasticity. Such a material is called viscoelastic. Polymeric liquids are a typical example of viscoelastic material. Secondly, polymeric liquids have nonlinear mechanical properties. For example, the viscosity of polymeric liquids is not constant but decreases with the shear rate. This phenomena, called shear thinning, is quite dramatic: the viscosity changes over several orders of magnitude. The nonlinearity also appears as various curious flow phenomena: for example when a polymeric liquid is sheared between concentric cylinders, the liquid crawls up the inner cylinder. This effect is called the normal stress effect. The complex flow properties of polymeric liquids are important in various practical situations. Industrial processes of polymer technology involves flow of polymers: injection molding, fiber spinning, film blowing, surface coating are a few examples. In such processes, understanding
393
the flow properties is essential in designing and controlling the processes. The unique flow properties of polymers are also utilized in the industries of paints, glues, surface coating etc. The science of studying the flow and deformation of materials is called rheology. Although these problems have been discussed in the two well established sciences of fluid mechanics and elasticity theory, there are many materials which cannot be dealt with by classical continuum mechanics. Polymeric material is a typical example. Since it has both elasticity and viscosity, it can be dealt with neither by fluid mechanics nor elasticity theory. Rheology is concerned with such materials. Because of their importance in industry, polymeric liquids have been most intensively studied in rheology. Many examples of their unusual flow behaviors are given in the textbooks on rheology (Bird et al., 1987; Barnes et al., 1989). In this chapter, an introductory description of the rheology of polymeric liquids is given: in particular, it will be discussed (i) how the complex flow properties can be characterized, and (ii) how they are related to molecular structures.
9.2 Linear Viscoelasticity 9.2.1 Significance of Linear Viscoelasticity
The rheological properties of a material are characterized by the relationship between stress and strain. The stress represents the force applied to the material and the strain represents the deformation caused by it. The full description of such relations becomes rather complex for polymeric liquids, and will be postponed to a later section. Here we shall start with a simple case, called linear viscoelasticity.
394
9 Viscoelastic and Rheological Properties
When the stress level is low, a linear superposition principle, the precise meaning of which will be given later, holds in the relationship between the stress and strain. Such regime is called linear. In this regime, the viscoelasticity can be characterized by a single material function. In many practical applications, the stress level is not so small for linear viscoelasticity to be directly applied. However, linear viscoelasticity is important because of the following reasons: (i) Much experimental work has been done in the linear regime since material characterization is easy in this regime. Indeed most information on the relation between molecular structure and viscoelastic properties has been obtained through the study of linear viscoelasticity. (ii) Linear viscoelasticity is useful in predicting the behavior in the nonlinear regime. There are many empirical relations which predict the nonlinear behavior from linear viscoelasticity.
In shear flow, a material is sandwiched between two parallel plates, and the top plate is moved parallel to the bottom (see Fig. 9-1). If the material is homogeneous, and if the end effect is negligible, the velocity v (r, t) of the material at position r and time t is given by
9.2.2 Shear Flow
y (t, t') = J dt" y (t'f)
vv = vz = 0
(9-1)
The constant (9-2) is called the shear rate. To induce such a flow, a force has to be applied to the top plate. The shear stress a is the x component of this force per unit area. Actually, a is the x-y component of the stress tensor, and should be written as oxy. However, as we discuss only the shear stress in the next first two sections, we shall for now abbreviate it as o. The integral of y from time t' to t t
(9-3)
t'
The material function characterizing linear viscoelasticity can be obtained by various types of flow, but most commonly, shear flow is used. Therefore we shall describe linear viscoelasticity taking shear flow as an example.
Figure 9-1. Shear flow between parallel plates.
is called the shear strain applied between time t' and t. In general, if the velocity field is described by Eq. (9-1) in a certain reference frame, which may be moving with the fluid element, it is called a shear flow. Thus the flows shown in Fig. 9-2 are also shear flows: (i) Flow between concentric cylinders (see Fig. 9-2 a). Here two concentric cylinders with the radii rl9 and r2 (r1 < r2) are rotated relative to each other. The flow can be regarded as a shear flow if the gap distance (r2 ~ rt) is much less than rx. If the inner cylinder is rotated with the angular velocity Q relative to the outer cylinder, the shear rate is given by y = '1
r, Q ^ —
' 1
(9-4)
9.2 Linear Viscoelasticity
Figure 9-2. Examples of shear flow. Side and top views of (a) concentric cylinders and (b) cone and plate.
(a)
The shear stress is related to the torque T on the inner cylinder: T G = (9-5)
2nr\L
where 2nr1L is the surface area of the inner cylinder dipped in the liquid. (ii) Flow between the cone and plate (see Fig. 9-2 b). Here the liquid is sandwiched between a fixed plate and a rotating cone. The shear rate is constant throughout the fluid, and is given by y = Q/9
2nr3
the shear stress G (t). Alternatively, we can control the shear stress G(t) and measure the shear rate y (t). A particularly simple experiment is the stress relaxation, in which we apply a shear strain y0 in a very short time, and measure the stress while keeping the shear strain constant (see Fig. 9-3).
y(t-)
(9-6)
where Q is the angular velocity of the cone, and 6 is the gap angle. The shear stress is related to the torque T acting on the plate by G =
395
(a)
(9-7)
time, /•
a(t)
where r is the radius of the plate. 9.2.3 Stress Relaxation To characterize the viscoelasticity of a material, we apply a certain shear flow controlling the shear rate y {t\ and measure
(b)
time, t
Figure 9-3. The shear strain (a) and the shear stress (b) in stress relaxation.
396
9 Viscoelastic and Rheological Properties
If the material is an elastic material, the deformation creates a shear stress proportional to the applied shear strain y0, and the ratio G = a/y0 is called the shear modulus. If the material is viscoelastic, the flow creates a shear stress which decays with time. The stress eventually becomes zero if the material does not have an intrinsic shape. Such a material is called a viscoelastic liquid. Polymer melts and solutions belong to this class. On the other hand, for gels or rubbers, the stress approaches a constant non-zero value. Such a material is called a viscoelastic solid. In this chapter, we shall deal only with viscoelastic liquids. Chapter 7 deals extensively with the viscoelastic solids. If the shear strain y0 is small, the shear stress is proportional to y0, and is written as a(t) = G(t)y0
(9-8)
called Boltzmann's superposition principle. The principle states that provided the stress level is low, the effect of strain on stress can be added. Let us consider a shear flow characterized by the shear strain: (9-12) Suppose that when a shear strain y1(t) is applied to a material, the stress cr1(t) appears and that for a shear strain y2(0> a stress a2 (t) appears. Then the Boltzmann's principle states that when the shear strain is yi(t) + y2(i), the shear stress will be
MO + MOIf this principle holds, the stress for any strain history y (t) can be expressed by the relaxation modulus G (t). To see this, notice that any time-dependent shear strain y(i) can be regarded as a sum of step strains (see Fig. 9-4). The step strain applied be-
G (t) is called the relaxation shear modulus. For polymer melts with narrow molecular weight distribution, G (t) can be approximately written as G(0«Goexp(-t/T)
(9-9)
A/ (t')
Go is called the instantaneous relaxation modulus, and T the relaxation time. In the general case, G(t) has many relaxation times, and can be written as t/zp)
> t
(9-10)
or G(t)=
j dlnTH(T)exp(-r/T)
(9-11)
— oo
The function H (T) is called the relaxation spectrum. 9.2.4 Boltzmann's Superposition Principle We shall now explain the meaning of the linear superposition principle, which is also
r
t
Figure 9-4. The principle of calculating the stress for arbitrary strain.
9.2 Linear Viscoelasticity
397
tween t' and t' + At' is
9.2.5.2 Oscillatory Experiments
Ay (t<) = y(t' + At1) - y ( 0 = y (?) At1 (9-13)
In this experiment an oscillatory shear strain
If this strain is applied independently, it will create the following shear stress at a later time t: (9-14)
= G(t-t')Ay(t')
According to the Boltzmann's principle, the effect of these strains can be added. Thus the stress at time t is the sum of the contributions of all the previous step strains. Thus (9-15)
a(t)= } dt'G(t-t')y(t')
(9-16)
Equation (9-16) is the basic equation in the linear viscoelasticity. 9.2.5 Parameters and Functions Characterizing Linear Viscoelasticity
9.2.5.1 Steady Shear Flow A steady shear flow is realized when the shear rate is kept constant. The steady state viscosity rj0 is the ratio between the shear stress and the shear rate. For constant f, Eq. (9-16) gives
oo
is applied to a material. If the material is an ideal solid, the stress is proportional to the shear strain, and is written as a(t) = Gy(t) = G y0 cos (co t)
(9-20)
On the other hand, if the material is a Newtonian fluid, the stress is proportional to the shear rate, and is given by a(t) = rjy(t)= -riy0sin(cot)
(9-21)
<* (t) = Jo [G> M cos (co t) - G" (co) sin (co t)\ (9-22) G'(co) and G" (co) are called the storage modulus and loss modulus respectively. Often they are combined to give G* (co) = G' (co) + i G" (co)
According to Eq. (9-16), the shear stress in various strain histories can be calculated by G (t). Thus, in principle, all the parameters and the functions characterizing the linear viscoelasticity can be expressed by G (t). Here we shall give a few examples.
-
(9-19)
If the material is viscoelastic, the behavior is intermediate between solid and fluid:
or in the continuous limit
= y0cos(cot)
= y $dt'G(t') 0
ri*(co) =
(9-23) (9-24)
ICO
which are called the complex modulus and the complex viscosity respectively. From Eqs. (9-16) and (9-19), it can be easily shown that G' (co), G" (co) and G* (co) are related to G(t) as G'(co) =co] dtG(t)sin(cot) o
(9-25)
00
G" (co) =co\dtG o
(t) cos (co t)
G*(co) = ico JdtG(Oexp(-icoO o
(9-26) (9-27)
(9-17)
From Eqs. (9-18) and (9-26), rj0 can be expressed by G" (co) or rj' (co) also:
(9-18)
nQ = lim G-{^
Hence rj0 is given by
= lim rj'(co)
(9-28)
398
9 Viscoelastic and Rheological Properties
If G (0 is given by Eq. (9-9), G' (co), G" (co) and G* (co) are given by G'(co) =
G n CO 2 T 2
measured. In the linear regime, the shear strain is proportional to
G 0 COT 2
2
1+CO T ' 1COT
'1 + icor
(9-29)
(9-30)
The function J {t) is called the creep compliance. For a viscoelastic liquid, the material will eventually flow with constant shear rate (J0/rj0, so that J (t) approaches to 07 0 for large t. From Eq. (9-16), it is easily shown
When COT §> 1, G*(co) = Go, which means that the polymeric material behaves as an (9-31) elastic material for quick deformation. On the other hand, for co T <^ 1, G* (co) = i co Go T, The intercept of the asymptotic line of J (t) which indicates that the material behaves with the vertical axis gives the steady state as a viscous liquid of viscosity rj0 = Go T for compliance J e . slow flow. t
(9-32)
9.2.5.3 Creep Experiments
Je = lim
The creep experiment is shown in Fig. 9-5. Here a constant shear stress a0 is applied for a sample at equilibrium, and the time evolution of the shear strain y(t) is
It can be shown by Eq. (9-16) (Bird et al., 1987, p. 270) that J e is related to G(t) as
no
\<\tG{i)t o
(9-33)
(a
or by Eqs. (9-25) and (9-26), = lim
G"(co)2
(9-34)
Je can also be obtained by setting the stress to zero after the steady shear flow is attained (see Fig. 9-5). When the stress is set to zero at a time t l 5 the elasticity of the material causes a reverse flow, and J e is related to the recovered strain = -lim 0".0 t-> oo
Figure 9-5. The shear stress (a) and the shear strain (b) in the creep experiments.
(9-35)
399
9.3 Characteristics of the Viscoelasticity of Polymeric Liquids
9.3 Characteristics of the Viscoelasticity of Polymeric Liquids 9.3.1 Time-Temperature Superposition Rule
The mechanical properties of a polymeric material change dramatically with temperature. For example, polystyrene appears as a glassy solid at low temperature, but as the temperature is raised, it becomes soft and finally becomes a fluid. The effect of temperature on the rheological behavior can be well described by a useful empirical relation called the time temperature superposition rule. The time-temperature-superposition rule states that well above the glass transition temperature Tg, changing the temperature is approximately equivalent to changing the time scale. For example, let G (t, Ts) and G (t, T) be the relaxation moduli measured at two different temperatures Ts and T. The time-temperature superposition rule states that they can be related to each other as =
bTG(t/aT,Ts)
(9-36)
where aT and bT are constants independent of t. From Eqs. (9-18), (9-33) and (9-36), it follows
Empirical equations are available for aT and bT\ aT =exp[
-
(9-38) l
s
"T" ^ 2
and (9-39) where Q and QS are the densities of the material at temperature T and Ts respectively, and Cx and C 2 are constants. If Ts is taken as the glass transition temperature Tg, Q
and C 2 become independent of the material (Ct « 17 and C 2 ~ 52). Examples of C1,C2 and Ts can be found in the literature (Ferry, 1980). For the other viscoelastic functions, the time temperature superposition rule gives the following relations: = bTG' (coaT,Ts); (9-40) and =
l/bTJ(t/aT9TJ
(9-41)
A similar superposition rule exists for the effect of pressure = fcPG(t/flF,Ps)
(9-42)
These superposition rules indicate that there is only a single fundamental time scale which governs the dynamics of polymers, and that changing the temperature or the pressure only affects this fundamental time scale. 93.2 Behavior of Relaxation Modulus The relaxation modulus G(t) shown in the literature covers an extremely wide range of time scale; typically of the order of 1010. Data for such a wide time scale is obtained by doing the experiments at different temperatures and then synthesizing a master curve by using the time temperature superposition rule. Figure 9-6 shows a sketch of the relaxation modulus of polymer melts. The time dependence of the relaxation modulus can be broadly divided into four regions, the glassy region, transition region, rubbery region and flow region. The glassy region corresponds to the behavior at low temperature, where polymeric materials have a rather high modulus of the order of 1010 Pa. The relaxation modulus decreases in the transition region, and then becomes flat
400
9 Viscoelastic and Rheological Properties
flow
Figure 9-6. Sketch of the relaxation modulus of a polymeric liquid. 10-8
10-6
10-4
10-2
1
in the rubbery region, where the polymeric liquids behave like a soft rubber (G « 105 - 106 Pa for polymer melts). In the flow region, which corresponds to the time scale larger than the longest relaxation time, the polymeric materials behave as a viscous fluid. Generally speaking, the relaxation behavior at longer time scale reflects the molecular motion on larger length scale.
The behavior in the glassy and transition region is determined by the local structure of polymers, such as the chemical structure of monomeric units or the short side branches etc. On the other hand, the behavior in the rubbery and flow regions is determined by the global characteristics of polymers, such as molecular weight, molecular weight distribution, and the structure of any long branches. As the molecular weight is increased, the rubbery region extends towards longer times, but the glassy and the transition regions remain unchanged. Figure 9-7 shows a sketch of the storage and loss moduli. The above characteristics can also be seen in this graph. The flow region corresponds to the region of low frequency where G' (co) and G" (co) have a linear portion on the log-log plot with respective slopes of 2 and 1 (see Eqs. (9-24) and (9-34)) Gf(co) = rjlJeco2
Figure 9-7. Sketch of the storage modulus G' (co) and the loss modulus G" (co) of a polymeric liquid.
and Gff (co) = rj0co (9-43)
The rubbery region corresponds to the plateau of G' (co), and the transition region corresponds to the increase of G' (co) and G" (co) at higher frequency.
9.3 Characteristics of the Viscoelasticity of Polymeric Liquids
9.3.3 Linear Polymers with Narrow Molecular Weight Distribution
For linear polymer melts with narrow molecular weight distribution, the relaxation modulus in the rubbery and flow regions can be approximated to a single exponential curve. This is shown in Figs. 9-8
-6
-5
-4
-3
-2
401
and 9-9 in the curves of G'(co) and G"(co). According to Eq. (9-29), G' (co) becomes constant for COT ^ 1 and G"(co) shows a peak at COT = 1. Such a plateau and peak are indeed seen in experimental data for polystyrene in Figs. 9-8 and 9-9. The flat region of G' (co) is called the rubbery plateau, and its height GN, called the plateau
Figure 9-8. Storage modulus of polystyrene with narrow molecular weight distribution. (From Onogi, S., Masuda, T, Kitagawa, K. (1970), Macromolecules 3, 109; reproduced by permission of American Chemical Society.)
-1
PS 160 °C
x
6
1w yr-
J
7/.
V
-6
-5
-k
-3
Y>
-2
-1 0 log UQ T ,sec"
y
F
Figure 9-9. Loss modulus of polystyrene with narrow molecular weight distribution. (From Onogi, S., Masuda, X, Kitagawa, K. (1970), Macromolecules 3, 109; reproduced by permission of Americal Chemical Society.)
402
9 Viscoelastic and Rheological Properties
modulus, is independent of the molecular weight, GN oc M°
(9-44)
while the relaxation time is strongly dependent on the molecular weight; T oc M 3 - 4 (9-45) The plateau modulus GN is used to estimate M e the molecular weight between the entanglement junctions. As it is described in Chap. 8, the rigidity modulus G of a rubber can be related to the molecular weight Mx of a polymer segment between the crosslinks: M =
QRT
(9-46)
where Q is the density of the polymer, R the gas constant and Tthe absolute temperature. Likewise, from the height of the rubbery plateau, one can define a molecular weight, M =
QRT
M 3 4
M -
(M < MJ (M > Mc)
(9-48)
The characteristic molecular weight M c is about two or three times larger than M e . Table 9-1. Characteristic molecular weights for selected polymers. Polymer species Polystyrene Poly(a-methyl styrene) 1,4-Polybutadiene Polyvinyl acetate Poly(dimethyl siloxane) Polyethylene 1,4-Polyisoprene Polyisobutylene
Me 19 500 13 500 1900 12000 10000 1300 6 300 8 900
Mc
M'c
35 000 130000 28000 104000 5000 13 800 24 500 86000 24400 61000 3 800 12000 10000 35 000 15 200 —
101*-
1012-
(9-47) 1010-
which is supposed to characterize the spacing between the entanglement junctions. In the early theory, the physical entity of the entanglement junction was not clarified. However M e turns out to be an important quantity in the modern theory of entanglement (see Sec. 9.7). Values of M e for some typical polymers are listed in Table 9-1. In the region where G(co) shows a plateau, the steady state viscosity depends on M as rj0 oc M 3 4 . As the molecular weight is decreased, the plateau disappears, and the molecular weight dependence of rj0 changes (see Fig. 9-10). When the correction for the chain end effect on the monomeric friction constant is accounted for, the viscosity is proportional to M. Thus the molecular weight dependence of the viscosity can be approximately expressed by
a 10* -
106 -
10* -
102 -
100 -
107
Figure 9-10. Molecular weight dependence of the viscosity. The filled circles indicate data corrected for the chain end effects on the monomeric friction constant. The open circles indicate unadjusted result. (From Colby, R. H., Fetters, L. J., Graessley, W. W. (1987), Macromolecules, 20, 2226; reproduced by permission of American Chemical Society.)
9.3 Characteristics of the Viscoelasticity of Polymeric Liquids
/ DO
°V /
-o
9.3.4 Effect of Molecular Weight Distribution and Branching
oo o
?•
8-7-
/ o
ft-
LOG
5 Mw
Figure 9-11. Molecular weight dependence of the steady state compliance of polyisoprene with narrow molecular weight distribution. (From Odani, H., Nemoto, N., Kurata, M. (1972), Bull Inst. Chem. Res. 50, 117.)
Figure 9-11 shows the molecular weight dependence of the steady state compliance. The change in the molecular weight dependence is also seen in J c . For M < M'c, Je is proportional to M, while for M > M'c, Je becomes independent of M. Values of M c and M' are shown in Table 9-1.
PS 160
7
^ fit\^&T
5-
A?
o L 15 •
/
r
/ /
G
2-
1 -
-5
-4
-3
If the polymer has broad molecular weight distribution, or long side chains, the relaxation spectrum becomes broad, and the clear plateau region disappears. An example for polystyrenes is shown in Fig. 9-12. As a result of the broad relaxation spectra, the steady state compliance becomes much larger than that of narrow distribution polymers. On the other hand, the viscosity is known to be described by the empirical Eq. (9-48) even for a system with broad molecular weight distribution provided the weight averaged molecular weight M w is used for M. The effect of branching has been studied in some detail for star shaped polymers which consists of / arms of equal length connected at the central node. Figure 9-13 compares the viscosity of such polymers with that of linear polymers. With fixed molecular weight, the viscosity decreases with the increase in the arm number /,
°C
6-
3-
403
-1
-2
log
sec- 1
PS 7
s>
Figure 9-12. G'(co) and G" (co) of polystyrene of narrow molecular weight distribution (unfilled circles) and broad molecular weight distribution (filled circles). (From Masuda, T., Kitagawa, K., Inoue, T, Onogi, S. (1970), Macromolecules 3, 116; reproduced by permission of the American Chemical Society.)
404
9 Viscoelastic and Rheological Properties
tion. To do that, it is essential to introduce the stress tensor. The stress tensor describes the force acting within the material. The a-j? component of the stress tensor at a point P in the material is defined as follows (see Fig. 9-14). Consider a small region of area AS on a plane which includes the point P and is normal to the jff-axis. Let / (/?) be the force which the material above the plane exerts on the material below the plane through the region AS. For small AS, the force is proportional to AS. The stress tensor aap is defined by the ratio (9-49) Figure 9-13. Viscosity of linear and star branched polyisoprene in normal tetradecane: circle: linear, square: 4 arm star and hexagone: 6 arm star. (From Masuda, X, Onogi, S. (1973), Annual Rep. Res. Inst. Chem. Fibers. 33 9.)
since the molecule is becoming more compact. However with the increase of the molecular weight at fixed arm number, the viscosity of star polymers increases more rapidly than linear polymers, and eventually exceeds it at high molecular weight. This tendency is seen for other types of branched polymers. This is because at high molecular weight molecular disentanglement in branched polymers takes place much more slowly than in linear polymers.
If there is no electric or magnetic field, the stress tensor is symmetric. Gap = Gpa
(9-50)
Also since the polymeric liquids may be regarded as incompressible, the stress tensor is written as °ap =
(9-51)
where p is an arbitrary function which is determined by external conditions (such as the external pressure). Therefore as to the diagonal components, GXX9 Gyy, GZZ, only
9.4 Non-Linear Viscoelasticity 9.4.1 Stress Tensor
So far, we have restricted the discussion to a special case, i.e. the linear viscoelasticity in shear deformation. We shall now discuss the general case, i.e., the nonlinear viscoelasticity for general types of deforma-
Figure 9-14. Definition of the stress tensor. The case of p = z is shown.
405
9.4 Non-Linear Viscoelasticity
their difference oxx — oyy, oyy-azz need to be characterized by rheological experiments. 9.4.2 Basic Principles in Continuum Mechanics
Rheological properties of a material are described by the constitutive equation. The constitutive equation relates the stress tensor at each material element with the past history of the deformation of the material. There are two basic principles which the constitutive equation has to obey. (1) Principle of local action: The stress tensor at a material element depends only on the previous history of the material element and not on the state of the neighboring material elements. Let us consider a stress tensor at a material element P. Let r(t,P) be the position vector of P at time t. This is determined by the flow field v (r, t) of the material: d
dV
= v(r(t,P),t)
(9-54)
where we have dropped the dependence on P assuming that both the stress and the velocity gradient refer to the same material element. (2) Principle of frame indifference: The constitutive equation must be independent of absolute motion of the material element. Any superposed rigid body motion (i.e., translation and rotation) does not affect the mechanical response of the material. That the constitutive equation is independent of the translation of the material element has already been accounted for in Eq. (9-54) since a is independent of v. The independence of the rigid rotation imposes a restriction on the functional form for the constitutive equation. Suppose that the material is rotated around the material element P. The rotation changes the vector dr connecting P and its neighboring material element to (9-55)
(9-52)
where U is an orthogonal tensor satisfying:
Due to the principle of local action, the stress tensor of the material element P is determined by the deformation experienced by the element P. The principle of local action states that the deformation at the material element P is described by the history of the velocity gradient tensor at P, i.e., %a
o(t)=
(t\p) = —^
(9-53)
^ P at point P, time t'
and that the higher order derivatives of the velocity field are irrelevant. Given the previous values of x (?, P) for ? < t, the stress tensor a (t, P) is determined uniquely. Thus the general form of the constitutive equation is that a (t, P) is a functional of x (?, P) for ? < t. In a symbolic form,
U r (t')U(O = l
(9-56)
Then the velocity gradient tensor changes as
(9-57) T
T
W (?) = U (?) * (?) U (?) + U (?) U (?)
The principle of frame indifference states that the stress tensor & (t) in the rotating material must be equal to the geometrical rotation of a (t), i.e., (t) = UT(t)<j(t)U(t)
(9-58)
Thus the constitutive equation must satisfy:
= U r (t)#[x(t')]U(0
(9-59)
for arbitrary orthogonal tensor U (t).
406
9 Viscoelastic and Rheological Properties
Due to the principle of frame indifference, we may choose an arbitrary frame in the study of the constitutive equation. Also due to the principle of local action, we may disregard the spatial dependence of the velocity gradient. Thus without loss of generality, we can assume that the flow field is given by (9-60)
v(r9t) = x(
In this case, each material point moves according to (9-61)
dt
Since this is a linear equation for r(t\ the general solution is written as r(t) = E(Ut')r(t')
9.4.3.1 Shear Flow The shear flow is shown in Fig. 9-1. It has a reflection symmetry with respect to the xy plane. Therefore the components °xz = Gzx a n d °yz = Gzy a r e always zero. Thus there are three stress components to be characterized, axy = ayx, N1 = axx — ayy, N2 = Vyy — °zz- Ni anc * N2 a r e called the first and second normal stress differences respectively. In the steady shear flow, the shear stress, and the normal stress differences depend on the shear rate y only. Due to the symmetry of the flow, the shear stress is an odd function of y, while the normal stress differences are an even function of y. Thus for small f, these functions must behave as
(9-62)
and
NuN2ccf
f
(9-66)
where E (t, t ) is a solution of the equation 8 dt
(9-63)
The ratio between the shear stress and the shear rate
with the initial condition E(t\t') = \
(9-64) r
The tensor E (t, t ) is called the deformation gradient tensor. The deformation of the material is expressed either by the velocity gradient tensor or the deformation gradient tensor. The constitutive Eq. (9-59) may be written as (9-65)
a(t) =
(9-67)
y
r
9A3 Non-Linear Viscoelasticity in Rheometrical Flows We shall now briefly describe the characteristic nonlinear viscoelasticity of polymeric liquids in simple flows. Two types of flows are often employed in the characterization of the rheological properties.
is called the steady state viscosity, and the ratios
f
f2
(9-68)
are called the first and second normal stress coefficients. An example of the shear stress <Jxy(y) and the first normal stress difference N± (y) are shown by the solid lines in Fig. 9-15. Notice that the normal stress difference is quite large: indeed at high shear rate, the normal stress difference is much larger than the shear stress. In Fig. 9-15 2G'(co) and G" (co) are also plotted for comparison. The curves axy(y) and G" (co) coincide with each other on the left side of the graph since they are proportional to rj0 y or rj0 co respectively for small y or co. The curves Nx (y) and 2 G (co) also agree with each
9.4 Non-Linear Viscoelasticity
y/s"\ Figure 9-15. Shear stress
other on the left side. This agreement is not fortuitous. By the retarded motion expansion (Bird etal, 1987, p. 329), it can be shown that for any isotropic material, lim Nt (y)y~2 = 2 lim G'(co)co - 2 co-*
407
Fig. 9-16, where the solid line is \rj*(co)\, and the circles denote rj (y). In this case, the Cox-Mertz law works very well. Although there is no theoretical justification, the Cox-Mertz law is known to work well for many polymeric liquids. The second normal stress difference is negative and its magnitude is small compared to the first normal stress difference (usually 10% to 30% smaller than the first normal stress difference). Figure 9-17 shows the growth of the shear stress when a shear flow of constant shear rate y is started. Here rj+ (t, y) is defined by
n+(t,y) =
(9-72) y It is seen that for large y the shear stress reaches a maximum before reaching the steady state value. This phenomenon is
(9-69)
0
or (9-70) Notice that the curves start to deviate from the straight lines at about the same y and co, which indicates that the nonlinearity and the viscoelasticity become important at the same characteristic time. This is an example of the correlation between the linear and nonlinear viscoelasticity. The correlation between the linear and nonlinear viscoelasticity is even more clearly demonstrated by an empirical relation
which is called the Cox-Mertz rule. An example of the Cox-Mertz law is shown in
JO"' 102 y/s~\
Figure 9-16. Steady state viscosity r\ (y) and the first normal stress difference coefficient ^ ( y ) of 10% polystyrene solution in Kaneclor is plotted against the shear rate. Je (y) is defined by ^ (y)/2 rj (y)2. (Data reproduced from Takahashi, M., Masuda, T., Onogi, S. (1977), Rheol. Soc. Jpn. 5, 72; reproduced by permission of Rheological Society of Japan.)
408
9 Viscoelastic and Rheological Properties
isotropic material
105H
lim rjE (e) = 3 lim rj (y)
(9-75)
10*-
103-
102 10-2
10-2
100
101
102
Msec)
Figure 9-17. Shear stress growth function n + (t,y) = (Txy(t, y)/y of poly butadiene is plotted for several shear rates. (From Menezes, E. V., Graessley, W. W. (1982), J. Polym. Sci. Phys. 20, 1817; reproduced by permission of American Physical Society.)
In Fig. 9-18, the steady elongational viscosity is plotted together with the steady shear viscosity. As the elongational ratio is increased, rjE (e) increases slightly and then decreases. For other polymers, it is sometimes difficult to achieve the steady elongational flow since the material ruptures before it reaches the steady state.
called the stress overshoot. Sometimes the overshoot is observed for the first normal stress difference at high shear rate. 9.4.3.2 Uniaxial Elongational Flow
The velocity field of the uniaxial elongational flow is given by vy=-\ey\
VX=-\E\
vz = az (9-73)
where s is called the elongational rate. Such a flow is realized when a cylindrical specimen is stretched in the z direction. Due to the symmetry of the flow, axx is equal to ayy, and all the off-diagonal components of the stress tensor
G
zz ~
G
y
which is called the elongational stress. When s is constant, aE approaches a steady state value. The ratio (9-74) is called the elongational viscosity. By symmetry it can be shown that for any
-2 -1 log / ; logf( log s~1)
0
Figure 9-18. Shear viscosity rj(y) and elongational viscosity nE (e) of near monodisperse polystyrene are shown as a function of y and s respectively. (From Takahashi, M, Masuda, T., Onogi, S. (1984), Polymer Preprints, Japan 33, 871.)
9.4.4 Constitutive Equation
The constitutive equation is a mathematical model for the rheological properties of a material. Since the nonlinear viscoelasticity is a complex phenomena, it is very difficult to have a simple mathematical equation that describes all features of the material property. A variety of constitutive equations have been proposed, but none of them are complete. If excessive accuracy is pursued, many parameters and a complicated set of equations are needed, and it will make the model difficult to use. Thus which type of constitutive equations one should use depends on one's purpose,
9.4 Non-Linear Viscoelasticity
and the types of the flow to be analysed (see the recent literatures, Tanner, 1985; Crochet et al., 1984; Tucker et al., 1989). Here I shall describe a few constitutive equations which frequently appear in the literature. Detailed discussions on the constitutive equation for the polymeric liquids are given in the books by Bird et al. (1986) and Larson (1988). 9.4.4.1 Second Order Fluid The constitutive equation for the Newtonian fluid is (9-76)
ear effect, and holds rigorously for any fluid if the acceleration and the velocity gradient are small. However, the model cannot describe the viscoelastic effect. 9.4.4.2 Convective Maxwell Model The simplest model which accounts for the viscoelastic effect is the convective Maxwell model, whose constitutive equation is 5a_ . (9-81) bt This gives a relaxation modulus G{t) for the viscoelasticity
where
(9-82)
(9-77) is the rate of strain tensor. Equation (9-76) does not describe the shear rate dependence of the viscosity. An obvious way to account for the shear rate dependence is to add a higher order term of the strain rate tensor. From the principle of frame indifference, it can be shown that such an expansion must have the form *,/, = M , / . + & 2 ^ + & 3 W , ,
(9-78)
where, bl, b2 and b3 are material constants and the operator 5/81 is defined by dt
409
The distribution of the relaxation f;mes can be accounted for by writing v = H°k
(9-83)
k
where each ak obeys Eq. (9-81) with material constants xk and rjk. The convective Maxwell model predicts that in the steady shear flow, the first normal stress difference is positive. However, the model does not predict the shear thinning behavior: the calculated steady shear viscosity is independent of the shear rate. 9.4.4.3 Integral Form of the Convective Maxwell Model
at (9-79)
for any tensor Aafi. This model is called the second order fluid. By considering the shear flow, the coefficients bl9 b2 and b3 can be related to the first and the second normal stress difference:
The differential equation for the convective Maxwell model can be put into an integral form by using the deformation gradient tensor E(t, f). The solution of Eq. (9-81) is written as (9-84)
h = rj0; b2 = - \ W± (0); b3 = W2 (0) (9-80) The second order fluid model describes the lowest order correction for the nonlin-
where = E(t,t')-E(t,t')
+
(9-85)
410
9 Viscoelastic and Rheological Properties
is called the Finger-strain tensor. For the system with a relaxation spectrum, the integral form becomes: = \
(9-86)
This constitutive equation has the same limitation as the original Maxwell model; it fails to describe the shear thinning.
pends on B (t, t'). (9-89)
a(t)=
f d
where the tensor O (B) is a certain non-linear function of B. Often it is furthermore assumed that this function is derived from a potential function W which is a scalar function of B. Such a scalar function is generally written using the two invariants of the tensor B:
9.4.4.4 Non-Linear Convective Maxwell Model
The shear thinning effect can be accounted for by assuming that the coefficients x ox r] are the functions of V or a. A well known example is the combination of the convective Maxwell model and the second order fluid model 5a
(9-87)
where x' is a material constant with the dimension of time. This model is called the Oldroyd fluid B. Another model proposed by Gieskus is Xx c 6 + —a
5a hi
—
(9-88)
where X is a dimensionless material constant. These models can describe many characteristic features such as the shear thinning, the normal stress effect and the viscoelasticity. However, for quantitative description, one has to introduce the relaxation spectra as has been done for the convective Maxwell model. 9.4.4.5 Single-Integral Constitutive Equation
A simple generalization of Eq. (9-86) is to assume that the kernel in Eq. (9-86) de-
(9-90) Then
B
(9-91)
The form of the single integral equation (9-89) and (9-91) is called the factorized BKZ model. With an appropriate choice of the function W{Il9I2), the BKZ model describes the rather complex rheological behaviors of polymeric liquids. Examples of the potential function W are ,Q ^ Woe log [1 + | ( / i - 3)] or
WocTr(Ba)
where a is a constant between 0 and 1.
9.5 Molecular Theory I: The Rouse Model 9.5.1 Polymer Motion in Concentrated Systems
We shall now discuss the rheological properties from a molecular viewpoint. In polymer melts or concentrated solutions, polymers entangle with each other as is shown in Fig. 9-19. At equilibrium, the configurations of polymers are quite random; there is essentially no correlation in the position of the centers of mass, and in the direction of polymer segments.
9.5 Molecular Theory I: The Rouse Model
411
9.5.2 Basic Equation of the Rouse Model
Figure 9-19. Polymers in concentrated system.
The molecular origin of the viscoelasticity of such a system can be qualitatively understood as follows. Consider the stress relaxation experiment. When the sample is stretched, the polymers are stretched and create an elastic force which tends to recover the original shape. The molecular origin of the restoring force is the same as for the rubber elasticity, i.e., the loss of the configurational entropy. In the case of rubbers, the polymers cannot recover the equilibrium configuration as long as the sample is stretched since the polymer chains are cross-linked. In the case of polymer melts, however, the polymers can recover the equilibrium configuration even if the macroscopic shape is held fixed since the polymers in liquid can diffuse freely. To describe such a process, we need a dynamical model of polymer chains. There are two models for the molecular dynamics, the Rouse model and the tube model. The Rouse model describes the motion of unentangled polymers, or the motion of entangled polymers over a short time. On the other hand, the tube model describes the slow motion of strongly entangled polymers. These models are for the general dynamics of polymers, and can be used, for example, for the calculation of the diffusion constant.
The Rouse model represents a polymer by a set of beads connected by springs (see Fig. 9-20). The idea behind the model is as follows. Suppose we divide the backbone atoms of a polymer into N — 1 groups, each consisting of ATS consecutive backbone atoms. Each group is called a submolecule. Let Rt (i= 1,2,... N) be the position vector of the backbone atom at the boundary of the submolecules. If Ns is taken to be large, the distribution of the end-to-end vector of each submolecule at equilibrium is Gaussian (993) Weq({R{}) oc
This distribution is equivalent to the Boltzmann distribution exp (— U/kB T) of the bead-spring model whose potential energy is given by (9-94) Thus the bead-spring model represents the equilibrium configuration of the polymer accurately. To introduce dynamics for the beadspring system, we assume that each bead experiences a drag as it moves through the surrounding polymers and that the drag is described by Stokes law. Let us consider a macroscopic flow v(r,t) = x(t)r
(9-95)
Figure 9-20. The Rouse model (left) and a polymer (right).
412
9 Viscoelastic and Rheological Properties
then the drag is proportional to the difference between the segment velocity Rt and the average velocity at the point of the segment x i?,.. Thus the equation of motion is (9-96)
m Rt =
where £ is the friction constant of the bead, and the last term represents the random force due to the thermal motion. The time correlation function of ft is given by the fluctuation dissipation theorem (see Doi and Edwards, 1986, Chap. 3): a ( W ) > = 2UkBT5ijd{t - t')
(9-97)
Usually in the time scale relevant for rheological properties, the inertia term is very small compared to the frictional term, and may be neglected. Thus from Eq. (9-94) and (9-96), we have
i_1
- 2/?,) + jft
Equations (9-98) to (9-102) are the basic equations for the Rouse model. Historically, the Rouse model was originally proposed for a polymer in dilute solution. However, the model turned out to be inappropriate for the dilute solution since it neglects the hydrodynamic interaction. On the other hand, the model has turned out to be useful for polymer melts or concentrated solutions if the molecular weight is smaller than the entanglement molecular weight M e . Why the model works so well in polymer melts is not clearly understood, but many experimental results can be fitted by the Rouse model. For convenience of calculation, a continuous version is employed for the Rouse model. This is obtained by regarding i as a continuous variable ranging from 0 to JV, and using the following transformation rule from a discrete model to a continuous one:
dR
where Cb2 3kBT
(9-99)
corresponds to the characteristic relaxation time of a submolecule. For the end bead, the equation of motion is slightly different: for i = 1, the equation becomes (9-100) This equation can be regarded as a special case of Eq. (9-98) if we define Ro by (9-101)
Ro =
Similarly, the equation for RN is obtained from Eq. (9-98) if RN + 1 is defined by = RN
(9-102)
1
di2
(9-103)
Equation of motion for the continuous Rouse model is (9-104) 8 1 d2R 1 — R(i, t) = — - ^ r + -f(i91) + x J?(i, t) 2 dt TS di C where the time correlation function for the random force is given by ,g . ^ <J{i, 0/0', t)> = 2t;kBTId (i -j) 5 (t - f) The differential equation (9-104) is supplemented by the boundary conditions which are derived from Eqs. (9-101 and 9-102): — = 0 at i = 0 and i = N
(9-106)
9.5 Molecular Theory I: The Rouse Model
9.5.3 Normal Coordinates The Rouse model gives a linear equation for R (i, t). A standard way of treating such a system is to use normal coordinates each capable of independent motion. In the present problem, the normal coordinates are given by
9.5.4 Self-Diffusion and Segmentai Motion in Equilibrium At equilibrium, a polymer molecule moves around by thermal motion and its speed is characterized by the self-diffusion constant defined by - RG(0))2}
= lim
(p = 0,1,2,...)
(9-107)
It then follows from Eq. (9-104), 8 dt
P
Cp
P
P
(9-114)
iv o
P
(for p = l,2,...)
(9-108)
(9-113)
where RG(t) is the center of mass of the molecule. For the Rouse chain, RG (t) is given by
1 p
413
F r o m
where
(9-115) and CP =
for p = l,2,...
and 6n2kBT Nb2
J
= 0,1,2,...)
(9-109)
Substituting this into Eq. (9-96) and using Eq. (9-110), we have (t) - RG(0))2} =
(9-116)
and/ p 's are the random variables satisfying = 0 p = 0,1,2,...;
(9-110)
Comparing this with Eq. (9-113), we have (9-117)
According to Eq. (9-108), the characteristic relaxation time of Xp is Cp/kp = TR/p2, where 3 n2 kn T
(9-111)
TR corresponds to the longest relaxation time of the Rouse model, and is called the Rouse relaxation time. Since N is proportional to the molecular weight M, Eq. (9-111) indicates that TROCM2
(9-112)
Thus the self-diffusion constant is proportional to the inverse of the molecular weight: Z)Goc Af"1
(9-118)
Although the motion of the center of mass obeys the simple diffusion law, the motion of a Rouse bead does not follow this law. It can be shown that the mean square displacement of a Rouse bead is given by (de Gennes, 1979) 6DGt
for
tp
IOr
t ^ T
TR
1/2
(9-119)
414
9 Viscoelastic and Rheological Properties
Notice that in a short time scale the mean square displacement increases at t1/2. 9.5.5 Molecular Expression for the Stress Tensor Whe shall now discuss the viscoelastic properties of the Rouse model. To do this, first we have to know the molecular expression for the stress tensor. This is obtained as follows: As it is explained in Sec. 9.4.1 the stress tensor aafi represents the a component of the force acting through the plane normal to the /? axis. In polymeric materials, the force has two origins: one is the molecular potential (such as the van der Waals potential) which acts between the atomic groups in proximity, and the other is the chemical bond which connects the backbone atoms. The stress arising from the molecular potential is essentially the same as for usual liquids, and will not be important for viscoelastic properties. Thus we shall consider the stress arising from the forces acting through the backbone atoms. Now if a chain is passing through the plane at a polymer segment i, the upper plane exerts a force ,Q . ^ ™..x
3fc
B^
„
,
The isotropic term represents the contribution from the molecular potential. In terms of the normal coordinates, the stress tensor is written as (9-122)
*.,(') = v £
kp<Xpa(t)Xpfi(t)y-P8afi
9.5.6 Constitutive Equation for the Rouse Model Given the molecular expression for the stress tensor, we can obtain the constitutive equation for the Rouse model. First from Eq. (9-108), we can show (see Doi and Edwards, 1986, p. 112) (f
(t)X
(t)y = k TS
(9-123)
Using the relation and Eq. (9-108), we have
±<x,MXpf(t» =
Kll(t)Xpfl(t)
Xpx(t)[
\Xpf(t))-
_ J
3kBTdR(Ut)
on the lower plane (see Fig. 9-21). If there are v chains per unit volume, the probability that the segment i of any polymer passes through the plane is
xXfl(Xpfl(t)Xpll(t)) (9-124) where
where AS is the area of the plane. Thus the stress tensor is given by
= v-
(9-125)
2K
Notice that Eq. (9-124) is similar to the convective Maxwell model. Indeed if we define (9_12g) = vkp(^Xpa(t)Xpfi(t)>
-
9.6 Molecular Theory II: The Tube Model
415
Since the Rouse relaxation time TR is proportional to M 2 , Eq. (9-132) indicates that rj0 oc M;
(9-133)
Je oc M
This is in agreement with the experimental relation Eq. (9-48) behavior for M < Mc or M<M'C (see Figs. 9-10 and 9-11). For t < TR? the sum in Eq. (9-131) may be replaced by an integral for p. Thus
AS,
Figure 9-21. Molecular origin of the stress tensor for the Rouse model.
G{t)&vkBT J dpexp
/
2n2 —t
1/2
(9-134) we can rewrite Eq. (9-124) in the form of the convective Maxwell model
B(afi
fi
(9-127)
The complex modulus is obtained from the Fourier transform of G{t). Especially for CDTRP>
1, G*(co) is given by T
\l/2
- )
<JP + *P^t°P
=
vk
(9-128)
BTTpy/
The total stress is written as oo
(9-129)
* = Z Op
As an example, let us consider the linear viscoelasticity in the shear flow. From the above constitutive equation, we obtain the relaxation modulus:
expf-
(9-130)
p=i
The factor vkBT can be rewritten as Q R T/M (Q being the weight of polymers per unit volume). Hence the relaxation modulus is written as (W31) The viscosity and the steady state compliance are then calculated by Eqs. (9-18) and (9-33). The result is K2QRT
1
"
2M J
(
tj
ll2
KvNkBT(toTs) (l
+0
(9-135)
Thus G' (to) = G" (to) at high frequency and increases as co1/2. These features are seen in Figs. 9-8 and 9-9 at high frequency period.
9.6 Molecular Theory II: The Tube Model 9.6.1 Characteristics of the Polymer Motion in Strongly Entangled Systems
The basic idea of the tube model is as follows (de Gennes, 1979; Doi and Edwards, 1986). Consider the motion of a certain test polymer in a strongly entangled system (see Fig. 9-22 a). If the test polymer attempts to move perpendicularly to its own contour, it will create a large scale motion of many surrounding polymers and encounter large resistance. On the other hand, if the polymer tends to move along its own contour, it will feel much less resistance. Thus dynamically the polymer
416
9 Viscoelastic and Rheological Properties
Figure 9-22. Entangled polymer system (left) and the tube model (right).
is considered to be confined in a tube (Fig. 9-22 b). The tube represents a mean field potential created by the other polymers, and gives a constraint to the polymer in it. In a simple version of the model, the tube is assumed to be fixed in the material. The motion of the polymer perpendicular to the tube is restricted by the tube wall, but its motion along the tube is free. This situation is represented by the Rouse chain confined in a tube (Fig. 9-23). At equilibrium, the tube takes a random configuration. The persistence length of the tube would be of the order of the tube diameter. To specify the model, it is usually assumed that the central axis of the tube consists of straight segments (called the tube segments) of equal length a, connected in random directions. In this model, only one new parameter is introduced. This is the length a, which characterizes the entanglement effect of the surrounding polymers. The rest of the parameters are the same as for the Rouse
model, i.e., N, b and £. All quantities in the tube model can be expressed by these parameters. For example consider the contour length L of the tube axis. At equilibrium, the tube axis is a randomly connected tube segment of length a. Since there are Z = L/a tube segments, the mean square end-to-end distance of the tube axis is Za2. This must be equal to the mean square end-to-end distance of the polymer Nb2. Thus b2 N Z = N-^ = —
and
L = Za
(9-136)
where we have introduced N =
(9-137)
As it will be shown later, JVe corresponds to the entanglement molecular weight M e . The above model is consistent with the time temperature superposition rule. Among the parameters a, N9 b and £, only C is considered to depend on the temperature or pressure sensitively. If the other parameters are assumed to be independent of temperature, the time-temperature rule is derived from the model by dimensional analysis. 9.6.2 Reptation and Fluctuations
Figure 9-23. Rouse chain in a tube.
From the model shown in Fig. 9-23, one would realize that there are three characteristic types of motion.
9.6 Molecular Theory II: The Tube Model
417
1) The lateral fluctuation
(3) Reptation
In a very short time scale, each Rouse bead can move around freely without feeling the constraint of the tube wall. This picture would be valid provided the mean square displacement of a Rouse bead is less than a, or from Eq. (9-119)
In the time scale longer than TR, only the mode of p = 0 is effective. Thus the Rouse chain may be regarded as moving along the tube with a diffusion constant
~
1/2
— I
b
t
(9-138)
where 'a
T
N2
(9-139)
This time characterizes the fluctuations of the Rouse bead in the length scale a. (2) Longitudinal fluctuation On the time scale longer than Te, the polymer motion perpendicular to the tube axis is hindered, but the motion along the tube is free. The motion along the tube can be studied by using the normal coordinates taken along the tube. The modes corresponding to Xp with p > 0 represent the fluctuations of the segment density along the tube. The characteristic time of this longitudinal fluctuation is the Rouse relaxation time :JV 2 T C
c
T
Nt;
As the polymer moves along the tube, one end moves out of the tube, then a new part of the tube is created in a random direction such that the length of the new tube segment is a. When a tube segment is created on one end, the tube segment on the other end becomes empty of the polymer. Such a tube segment will not impose any constraint on the polymer, and can be regarded as destroyed. Thus the one-dimensional diffusion of the polymer is accompanied by the creation and destruction of the tube (see Fig. 9-24). Such a model was first considered by de Gennes, who called the motion reptation.
(9-140)
As a result of the longitudinal fluctuation, the contour length of the polymer along the tube fluctuate around L. The magnitude of the fluctuation has been estimated (see Doi and Edwards, 1986, p. 206) AL « N1'2 b
Z l/2
(9-141)
Thus for a very long polymer, the relative fluctuation AL/L becomes small, and the tube may be regarded as having a fixed contour length.
(9-142)
Figure 9-24. Reptation motion.
418
9 Viscoelastic and Rheological Properties
The characteristic time of reptation can be estimated by the time needed for the polymer to disengage from a certain tube, or the time needed for the polymer to move over the distance L along the tube. Thus L2
(9-143)
This is called the reptation time. Using Eqs. (9-136) and (9-142), this can be written as (9-144) Notice that the ratio of these characteristic times is written as (9-145) Thus for a very long polymer of Z > 1, the time scales of these modes are quite different from each other.
moves out of the tube, and can go in a random direction, whence:
r{L,t + At) = r(L9t) + v{t)
(9-148)
where v (t) is a random vector, whose mean and variance are given by <» = 0; (v2(t)}=aAs
(9-149)
Similarly if As < 0, r (0, t) changes as r (0, t + At) = r(0,t)-v (t)
(9-150)
Equations (9-147) to (9-150) describe the reptation motion. 9.6.4 Self-Diffusion Constant As an application of the kinetic equation, let us consider the mean square displacement of the center of mass: (9-151)
9.6.3 Reptation Model
From Eq. (9-147), we have
The analysis of the reptation model becomes simple if the fluctuation modes are neglected. In this case, the polymer is regarded as having a fixed contour length L and moving only along itself with the diffusion constant Dc. The equation of motion for reptation is obtained as follows. Let r (s, t) be the position of a polymer at the contour length 5 (0 < s < L) and time t. Consider that the polymer moves a distance As along itself in a time interval At. Clearly, As takes positive or negative values randomly, its mean and variance are given by
(9-146)
Since the polymer moves along itself, the time evolution of r (s, t) is given by r(s,t + At) = r(s + As,*)
(9-147)
Special consideration is needed for the chain end. If As > 0, the segment at s = L,
(9-152) 1 r G (t + At) - vG (t) = - (r (L, t) - r (0, t)) As where we have neglected the terms of order v/L. Taking the average of the square of Eq. (9-152) and using the fact that As is independent of r (s, t\ we have <(rG(t + At)-r G (t)) 2 > = = - ^ « v { L , t) - v(0, t))2}
Af b J1
3L2
7
(9-155)
9.6 Molecular Theory II: The Tube Model -6-
According to Eq. (9-155), DG is proportional to M~ 2 . Figure 9-25 shows the self-diffusion constant of monodispersive polystyrene in melts. The results can be described as
DGx
2
M~
for for
M<MC M>MC
419
monodisperse PS Tn +125 °C
-8-
(9-156) ^ -10-
This is in agreement with Eqs. (9-118) and (9-155).
© Bueche 6 Bachus+Kimmich
en o
-o Fleischer
-12-
9.6.5 Stress Relaxation Now we shall discuss the viscoelastic properties using the reptation model. First we have to know the expression for the stress tensor. This is obtained from Eq. (9-121) if we assume that the polymer is stretched homogeneously along the tube. According to this assumption, the Rouse segment i is located at the point s = L(i/N). Hence dR/di is evaluated as
dR(i,t)
(9-157)
where u(s, t) is the unit vector tangent to the tube axis, i.e.,
u(s,t) =
(9-158)
8s
Substituting this into Eq. (9-121), we have 3k T
L
^ L j d ^ f c O - P S . , (9-159)
J
-14-
I log Afw
Figure 9-25. Self-diffusion constant of linear polystyrene. (From Watanabe, H., Kotaka, T. (1987), Macromolecules 20, 530; reproduced by permission of American Chemical Society.)
but since the contour length of the polymer is relaxed, the polymer occupies the central part of the tube of contour length L (see Fig. 9-26). The initial value of Sa)8 can be calculated by geometrical considerations. If a tube segment is in the direction of a unit vector u before the deformation, it will be in the direction E //
(9-162)
u = IE-ill
where (9-160) Let us consider the stress relaxation experiment. Suppose that at t = 0, a step deformation is applied for a system in equilibrium. Let E be the deformation gradient. Under the deformation, the shape of the central axis of the tube deforms as rf =
(9-161)
B'
Figure 9-26. Deformation of a tube in shear. When the shear is applied, the points A and B are transformed to A' and B', but since the tube length is fixed, the ends of the tube come to A" and B".
420
9 Viscoelastic and Rheological Properties
after the deformation. The length of the tube segments becomes a|Eif|. Now the probability that an arbitrary chosen polymer segment is located on the tube axis parallel to u is proportional to the step length of the tube segment a|E«|. Thus SaP(s,+0) =
u (s, t) is isotropic at s = 0 and L. Hence Safi vanishes at s = 0 and L: Safi (s, t) = 0 for s = 0 and s = L
(9-167)
Equation (9-166) can be solved under the initial condition (9-163) and the boundary condition (9-167). The result is: (9-168)
Safi(s,t) =
sin where
(E-»),
= (ME)
(9-163)
where <...>„ denotes the average of u for the isotropic distribution of u\ <•••>« = J-$du--
(9-164)
For t > 0, w(s, t) satisfies u(s,t + At) = = u(s + As,t) (see Eq. (9-147)). Hence Sajg (s, t + At) is written as SaP(s,t + At) = (Safi(s + As,t)>
(9-169) = 3ZTR n2D, Substituting this into Eq. (9-159), we have 3vL 2 . Nb2 (9-170) where (9-171)
SaP{s,t + At) = Sap + < As >^ s «/> +
/
(2p-l)2 \
For shear strain, the deformation gradient E isj given by
(9-165)
where the average is taken for the random variable As. Expanding the right hand side of Eq. (9-165) with respect to As, and using Eq. (9-146), we have
l
8 y n2 p=
E=
1 y 0 0 l 0 0 0 1
(9-172)
For small shearing strain y, the tensor Q is expanded with respect to y as 4
~ ' "
(9-173)
The shear stress is then calculated as 4 vL2 ;kBTy
'a?
Thus the relaxation modulus is given by
or
^1
(9-166)
6s 8fL The equations at the chain end (Eqs. 9-148 and 9-150) indicate that the distribution of :
2
G(t) = Go0(t) where 4 vL2 5 Nb2
B
(9-175) 4viV 5 AT
B
(9-176)
421
9.6 Molecular Theory II: The Tube Model
Notice that
4vN Me J i\ or M 4N
4 QRTN
MNe
e
The molecular weight dependence of the viscosity rj0ccM3 is not completely in agreement with the experimental results (Eq. 9-48). The discrepancy, however, can be removed if the fluctuations are accounted for (see Sec. 9.6.6). On the other hand, the molecular weight in dependence of J e is in agreement with the experimental result for polymer melts with narrow molecular weight distribution (see Fig. 9-11).
(9-177) 9.6.6 Fluctuations
(9-178)
Thus ATe corresponds to Me, and a2 = Ne b2 corresponds to the mean square end-toend distance of the polymer with the molecular weight M e . The viscosity and the steady state compliance are calculated from G (t) by the formula (9-18) and (9-33). The result is (9-179)
The analysis in the previous section is valid for a very long polymer with Z > 1. In reality, however, Z is not very large, typically less than 100. Therefore, the coupling between the reptation and longitudinal fluctuations can be important. An important effect is in the reptation time r d . If the contour length is fluctuating rapidly, the polymer can move out of a tube when it moves the distance L — AL along the tube (see Fig. 9-27). Hence xd * (L - AL)2/DC
and
6Me
(9-180)
5QRT
(a)
L2
D
*
(9-181) 2
(l-Z^ocM3 1-
^
(b)
Figure 9-27. The Brownian motion of a polymer with (a) fixed contour length, and (b) fluctuating contour length. The oblique lines denote the region that has not been reached by the end of the polymer. The length of this region decreases faster in (b) than in (a). (From Doi, M. (1981), J. Polym. Sci. Lett. 19, 265.)
422
9 Viscoelastic and Rheological Properties
For the region of the molecular weight 10 < M/Mc < 100, the molecular weight dependence of Eq. (9-181) is close to the empirical relation r d oc M3<4. If this correction is introduced in Eq. (9-179), it explains the experimental relation for the viscosity rj0 oc M3-4. The effect of the lateral and the longitudinal fluctuations can be seen in the rheological properties of short time scales. Let us consider the stress relaxation. For t < xe, the behavior of the relaxation modulus of the tube model should be the same as that of the Rouse model. Thus G(t)
'T\
vNknT[-\
1 / 2
(for t
Therefore at t « Te N G(xe)xv — kBT
(9-183)
This is the same order of magnitude as Go, the initial value in the reptation model (Eq. (9-176)). Hence G(t) for the Rouse model crosses over smoothly to the reptation regime at t « r e . Thus the relaxation due to the longitudinal motion does not appear in G (t). This is because the contour length is unaffected by small shear strain. The longitudinal mode starts to contribute to the stress relaxation as the shear strain becomes large. Figure 9-28 shows the non-linear stress relaxation modulus defined by G (t, y) = oxy (t, y)/y. As the shear strain increases, a new relaxation process appears. The characteristic time of the process zk is proportional to M 2 , which indicates that zk corresponds to TR and that the process corresponds to the relaxation of the contour length. For t > TR, the stress is given by Eq. (9-170). Hence the nonlinear relaxation modulus is given by (9-184) ) = —Qxy(y)G
102-
a.
1{
100-
10-1-
100
102 103 t/s Figure 9-28. Nonlinear relaxation modulus G (t, y) for solution of polystyrene in chlorinated biphenyl. Magnitude of shear y are < 0.57, 1.25, 2.06, 3.04, 4.0, 5.3 and 6.1 from top to bottom. (From Osaki, K., Nishizawa, K., Kurata, M. (1982), Macromolecules 15, 1068; reproduced by permission of American Chemical Society.) 101
where (9-185) Thus for t > TR, G{t,y) can be factorized into two parts; one depends on shear strain only, and the other on time. This prediction has been confirmed by experiments (Fig. 9-29). The function h(y) is called the damping function. Figure 9-30 shows a comparison between the theoretical damping function and the experimental one. The agreement is very good considering that h (y) includes no adjustable parameter. 9.6.7 Constitutive Equation
It is not possible to get a simple analytic constitutive equation for the reptation model described above. However, there is a
423
9.6 Molecular Theory II: The Tube Model
100 -
\
ifl-1 -
\8"
V
\ \
10-2-
100
10-1
101
Figure 9-29. Reduced nonlinear relaxation modulus G{t,y)/h(y). Each curve for y > 1.25 in Fig. 9-27 is shifted vertically by an amount — log [h (y)] so that it superposes on the top curve in the long time region. TJ indicates the longest relaxation time, and xk the characteristic time below which the superposition is not possible. (From Osaki, K., Nishizawa, K., Kurata, M. (1982), Macromolecules 15, 1068; reproduced by permission of American Chemical Society.)
Figure 9-30. The damping function h (y) obtained by the procedure explained in Fig. 9-29. Filled circles represent polystyrene of molecular weight 8.42 x 106 and the unfilled circles of 4.48 x 106. Directions of pips indicate concentrations which range from 0.02 gem" 3 to 0.08 gem" 3 . The solid line represents the theoretical value. (From Osaki, K., Nishizawa, K., Kurata, M. (1982), Macromolecules 15, 1068; reproduced by permission of American Chemical Society.)
useful approximation, called the independent alignment approximation which enables us to derive a simple constitutive equation. In the presence of a flow, the equation of motion for u (s, t) becomes
culate the stress tensor for general flows. The result is (9-188)
u (s,t + At) = u{s + As, t) + Awflow (9-186) The first term represents the reptation motion, and the second term the effect of the macroscopic flow. The independent alignment approximation assumes that Anflow is given by the change of a unit vector embedded in the material. Atffiow = (x(t) • u - (*:uu)u)At
(9-187)
This is not consistent with the deformation model shown in Fig. 9-26, but the error of the approximation has been shown be small as long as the direction of the flow is not changed. If the independent alignment approximation is used, it is possible to cal-
°*,(t) = G'o \ — 00
dtd(P{t~tf)QW(E(t,t>)) or
where G'o = 5/4 Go, &(t) is given by Eq. (9-171), and Q^(E) is given by
Equation (9-188) is the BKZ type constitutive equation explained in Sec. 9.4.4.5. It reproduces major characteristic aspects of the non-linear viscoelasticity. Detailed comparison of the theory with experiments has been done by Osaki and Doi (1984). 9.6.8 Effect of Branching and Molecular Weight Distribution
The crucial assumption of the model described above is that the tube can be re-
424
9 Viscoelastic and Rheological Properties
garded as fixed in the material. For linear polymers with narrow molecular weight distribution, this simple picture appears to hold: at least many experimental results can be accounted for qualitatively or sometimes quantitatively by this simple model. However, the simple picture breaks down for polymers with broad molecular weight distribution. The failure of the fixed tube model can be demonstrated by a simple example. Consider a mixture of short (S) and long (L) polymers. If the molecular weight of the short polymers M s , is close to that of the long polymers M L , the fixed tube assumption will be valid. However as M s decreases, the constraints imposed by the short polymers become weaker. In the extreme case of M L > M s , the constraint imposed by the short polymers will be negligible for the long polymers. Thus in a system with broad molecular weight distribution, the assumption of the fixed tube will be invalid. A mechanism for the motion of the tube is illustrated in Fig. 9-31. The topological constraints imposed on the polymer A is
Figure 9-31. Constraint release process.
released and recreated if the polymer C moves as shown in Fig. 9-31. Such a process, called the constraint release, causes the motion of the tube. The constraint release is considered to be important also for branched polymers. If the polymer has long side branches, the reptation motion is severely suppressed. In such a case, the contour length fluctuation, and the constraint release are considered to be the dominant mechanism of stress relaxation. However, it is difficult to account for the coupling between reptation, constraint release and contour-length-fluctuation, and quantitative theory is not yet available. The constraint release is discussed in detail by Graessley (1982) and Marrucci (1985).
9.7 References Barnes, H. A., Hutton, X E, Walters, K. (1989), An Introduction to Rheology. Amsterdam: Elsevier. Bird, R. B., Armstrong, R. C. Hassager, O. (1987), Dynamics of Polymeric Liquids, 2nd ed. volume 1. New York: John Wiley. Crochet, M. I, Davies, A. R., Walters, K. (1984), Numerical Simulation of Non-Newtonian Flow. Amsterdam: Elsevier. de Gennes, P. G. (1979), Scaling Concepts in Polymer Physics. New York: Cornell Univ. Press. Doi, M., Edwards, S. F. (1986), The Theory of Polymer Dynamics. Oxford: Oxford University Press. Ferry, I D. (1980), Viscoelastic Properties of Polymers, 3rd ed. New York: Wiley. Graessley, W. W (1982), Adv. Polym. Sci. 47, 67. Larson, R. G. (1988), Constitutive Equations for Polymer Melts and Solutions. Stoneham: Butterworths Publishers. Marrucci, G. (1985), Adv. Transport Processes 5: Mujumdar, A. S., Mashelkar, R. A. (Eds.), Wiley Eastern Ltd., New Delhi. Osaki, K., Doi, M. (1984), Polym. Eng. Rev. 4, 35-72. Tanner, R. I. (1985) Engineering Rheology. Oxford: Clarendon Press. Tucker III, C. L., (Ed.) (1989), Computational Modeling for Polymer Processing. Munich: Hanser Publishers. Walters, K. (1975), Rheometry. London: Chapman & Hall.
9.7 References
General Reading Astarita, G., Marrucci, G. (1974), Principles of NonNewtonian Fluid Mechanics. London: McGraw Hill. de Gennes, P. G. (1975), The Physics of Liquid Crystals. Oxford: Oxford Univ. Press. Edwards, S. F. (1976), The Configurations and Dynamics of Polymer Chains in Molecular Fluids: Balian, R., Weill, G. (Eds.). London: Gordon and Breach. Flory, P. F. (1953), Principles of Polymer Chemistry. Ithaca, N.Y.: Cornell Univ. Press.
425
Janeschitz-Kriegl, H. (1964), "Polymer Melt Rheology and Flow Birefringence", Elastic Liquids: Springer, A. S., Lodge, T. P. (Eds.). London: Academic Press. Lodge, T. P., Rotstein, N. A., Prager, S. (1990), "Dynamics of Entangled Polymer Liquids: Do Linear Chains Reptate?", Adv. Chem. Phys. 79, 1-132. Osaki, K., Doi, M. (1984), "Nonlinear Viscoelasticity of Concentrated Polymer Systems", Polymer Engineering Reviews 4, 35-72. Tirrell, M. (1984), Rubber Chem. Tech. 57, 523. Tobolsky, A. V. (1960), Properties and Structure of Polymers. New York: John Wiley.
10 Plastic Deformation of Polymers Buckley Crist Department of Materials Science and Engineering, Northwestern University, Evanston, IL, U.S.A.
List of 10.1 10.1.1 10.2 10.3 10.4 10.4.1 10.4.2 10.4.3 10.5 10.5.1 10.5.2 10.5.3 10.6 10.6.1 10.6.2 10.6.3 10.7 10.8 10.9
Symbols and Abbreviations Introduction Basic Concepts of Plastic Deformation Mechanical Testing and Definitions Criteria for Yielding and Crazing Yielding and Deformation Behavior Glassy Polymers Semicrystalline Polymers Blends and Block Copolymers Fundamental Nature of Polymer Yielding Glassy Polymers Semicrystalline Polymers Crystalline and Liquid Crystalline Polymers Post-Yield Deformation and Modeling Viscoelastic Models Molecular Models Continuum Mechanics of Necking Summary Acknowledgements References
Materials Science and Technology Copyright © WILEY-VCH Verlag GmbH & Co KGaA. All rights reserved.
428 430 430 432 436 437 438 440 444 445 446 451 457 460 460 461 464 465 467 467
428
10 Plastic Deformation of Polymers
List of Symbols and Abbreviations a A, Ao b C E (t) Eo F G H AH k /, l0 Za lc L Lk m n P r0 R* t t' T Td 7^ Tg Tm TR w* A I/* v* w z*
radius of molecule or bundle in a double kink instantaneous loaded area, initial loaded area modulus of the Burger's vector of a dislocation exponential factor in nonlinear viscoelastic model tensile stress relaxation modulus elastic (unrelaxed) tensile modulus force elastic shear modulus hardness activation energy in absence of stress Boltzmann's constant instantaneous length, initial length thickness of amorphous layer thickness of lamellar crystal long period of semicrystalline polymer length of a kink band number of cooperatively rearranging conformers width parameter of distribution of relaxation times pressure radius of dislocation core critical radius of dislocation loop time relaxation time or average relaxation time absolute temperature temperature of tensile drawing temperature at which yield stress vanishes glass-liquid transition temperature melting temperature reference temperature for stress-strain behavior critical size of region sheared by screw dislocations activation barrier for forming nucleus of critical size activation volume height of kink band in loading direction critical size of kink pair
a 7, y 7y f a, g n , £t e, £ n , £t £p £y
fractional crystallinity shear strain, shear strain rate shear yield strain preexponential frequency factor in Arrhenius equation uniaxial strain, nominal strain, true or logarithmic strain uniaxial-strain rate, nominal-strain rate, true- or logarithmic-strain rate plastic strain uniaxial yield-strain
List of Symbols and Abbreviations
e* critical strain for onset of plastic flow r\ viscosity 9 fictive temperature A uniaxial stretch ratio fi pressure coefficient of shear yield-stress v Poisson's ratio
Williams-Landel-Ferry equation small-angle neutron scattering Kohlrausch-Williams-Watts function
429
430
10 Plastic Deformation of Polymers
10.1 Introduction Plastic deformation is closely associated with the technology of polymers. Thermoplastics and thermosets in engineering applications ultimately fail by yielding, that is, by the onset of plastic deformation. This may be caused by large scale ductile failure, or by local crazing which precedes "brittle" fracture. Yielding thus limits the conditions under which polymer materials can be used in load bearing situations, but at the same time provides for toughness by ductile rather than truly brittle fracture. The study of this process is essential for understanding the strength of solid polymers. From a different viewpoint, processing to impart macroscopic anisotropy to chain-like macromolecules is routinely employed to modify the behavior of polymers. A good example of property enhancement by plastic deformation is fiber drawing, which increases tensile strength and elastic modulus in the draw direction by factors of 5 or more, compared to the isotropic material. Films are also subjected to biaxial stretching to modify mechanical, optical, and permeation properties. Despite the importance of yielding and post-yield deformation in design, processing, and utilization of plastics, films, or fibers, our fundamental understanding of these phenomena in polymers is not so advanced as in other classes of materials. There are many reasons for this situation. The principal one is that physical microstructures of polymer solids are more irregular and heterogeneous than those of polycrystalline metals or ceramics. This complicates the establishment of structureproperty relationships. Mechanical behavior of polymers is furthermore sensitive to such variables as temperature, strain rate, hydrostatic pressure, and environment. Presented here is an overview of plastic
deformation and the theoretical approaches that have been developed to account for observed behavior. Emphasis is on shear yielding, which occurs at essentially constant volume. Crazing, very local plastic deformation with considerable void volume dispersed among microscopic fibrils, has been reviewed by Kramer and Berger (1990). The relation of crazing to deformation and fracture is treated in Chap. 15 of this Volume. Further articles treating mechanical properties of solid polymers are those on elastic behavior of crystalline polymers (Chap. 7, this Volume) and elastomeric networks (Chap. 8, this Volume). 10.1.1 Basic Concepts of Plastic Deformation A solid body responds to an applied stress field by changing shape and, in most cases, volume. Simple elastic deformation depends only on the instantaneous stress field; elastic strains are established immediately upon imposition of stress and are independent of the history of the material. In practical terms, the dimensions of an elastic body are time-independent in a constant stress field and revert to their original values when the stress is removed. For an ideal elastic-plastic material, permanent or irreversible shape changes occur at a particular stress, which is defined as the yield stress; there is an associated (elastic) yield strain imposed before the onset of plastic deformation. The yield stress and yield strain define the yield point. Within a solid that has yielded, the local packing of atoms, ions, or molecules has been perturbed to a new configuration which is stable in the absence of the stress. Perhaps the clearest example is the shear deformation of adjacent planes of spherical atoms depicted in Fig. 10-1. Two features
10.1 Introduction
oooo
DQOD (a)
431
(b) y = 0 . 5
of this simple atomic model are quite general and apply to polymers: incremental plastic deformation occurs by local shear strains on the order of unity, and there is no appreciable change in volume or density. It is obvious that this model does not describe the local arrangement of matter in polymers, particularly in glassy ones. Computer simulation of the deformation of polymer glasses indicates that smaller shear strains may be distributed over relatively large volumes (Chap. 10, Vol. 6; Mott etal., 1993). Nevertheless, the basic idea that structural elements (repeat units of polymer chains) are mutually displaced by local shear to new, perhaps metastable, equilibrium positions is appropriate. At this point one feature which is unique to plastic deformation of chain-like macromolecules can be identified. The conformations of polymer chains are altered by the deformation process, that is, the chains are stretched. This leads to the well-known phenomenon of "reversion" or essentially complete recovery of the undeformed macroscopic shape when the material is heated beyond an appropriate temperature
(c)
Figure 10-1. Model for elementary shear-displacement; rows of circles represent planes of atoms or polymer chains. Energies of undeformed (a) and plastically deformed (c) states are the same. The maximum resistance i y (derivative of energy with respect to shear) occurs at yy = 0.25, midway between the states of lowest (a) and highest (b) strain energy.
(the glass transition temperature Tg for amorphous polymers, or the melting temperature Tm for semicrystalline polymers). The significance of plastic deformation in polymer processing derives from chain orientation and its effect on properties. Interatomic bonding is conspicuously anisotropic for polymers in either glassy or crystalline states. Local shear is accomplished by disruption and re-establishment of weak intermolecular bonds by a process analogous to the sketch in Fig. 10-1. Strong covalent bonds along the polymer chain are unaffected (aside from their orientation), and the collective effect is distortion of the chain conformation as shown in Fig. 10-2. Macroscopic chain orientation leads to anisotropic macroscopic properties within the plastically deformed polymer (Ward, 1975). If, on the other hand, a polymer is deformed in the liquid state at a temperature so high that microbrownian motion can rapidly relax the chain elongation, shear flow occurs by mutual displacement of essentially undistorted coils; macroscopic shape change is achieved without substantial chain orientation or
432
10 Plastic Deformation of Polymers
Figure 10-2. An isotropic polymer chain (a) is stretched and partially oriented by local shear (b). This conformation is retained in the unloaded glass (or crystal) (c) because new intermolecular bonds have been established as in Fig. 10-1. Entropic forces drive recovery of original conformation (a) when the sample is melted.
attendant anisotropy of properties. Melt processing of polymers by viscous flow is in this sense similar to plastic deformation of metals or other atomic solids. Viscoelastic and rheological properties of melts and solutions are discussed elsewhere (Chap. 9, this Volume). The model in Fig. 10-1 can be used to estimate the conditions at which plastic deformation will commence (see e.g., Bowden, 1973). The result is a yield point defined by a critical shear stress i y « G/6, where G is the elastic shear modulus, and a critical shear strain yy « 0.25. Experimental shear yield strengths at low temperature are about a factor of 2 below this estimate (Bowden, 1973; Brown, 1986). Glassy and semicrystalline polymers thus have strengths which are reasonably close to the theoretical limit established by the simple model; this subject is treated in more detail in Sec. 10.5. Characterization of plastic deformation requires definition of the yield point (the stress and strain at which irreversible strains are first observed). Another quan-
tity of interest is the maximum plastic stretch that can be imposed before fracture; this is a measure of the ductility or toughness of the material. Related to postyield deformation are strain softening and strain hardening, that is, changes in the material's susceptibility to additional plastic deformation. Since polymers are viscoelastic, one expects mechanical responses to be rate and temperature dependent. There are also pronounced effects from pressure and certain chemical environments. Experimental and theoretical aspects of plastic deformation are covered in this chapter. Sections 10.2 and 10.3 outline procedures used for mechanical testing and the relation between yield under different stress fields. Descriptions of yield and deformation behavior of isotropic polymers, both glassy and semicrystalline, are presented in Sec. 10.4. The emphasis is on homopolymers, though copolymers and blends are treated briefly. Theoretical developments for yield in glassy, semicrystalline, and crystalline polymers are discussed in Sec. 10.5. More phenomenological models are used to describe the flow of polymers after the yield point; these are reviewed in Sec. 10.6.
10.2 Mechanical Testing and Definitions Mechanical properties are derived from observation of dimension changes in a solid body subjected to uniaxial or multiaxial stress fields. Yielding does not occur under uniform hydrostatic stress; loading geometries of lower symmetry are needed to achieve plastic deformation. This statement does not apply to anisotropic or porous materials, which are not considered here. The most common procedure is uni-
10.2 Mechanical Testing and Definitions
axial tensile deformation of a specimen having a gage section of reduced area Ao and length l0 between two clamping regions of larger area as sketched in Fig. 10-3 a. The clamped ends are separated at a constant rate and the force F is recorded as a function of this separation. Samples are usually rectangular or circular in cross section. The true tensile stress is force F divided by instantaneous area A; at=F/A. One frequently employs the nominal stress on = F/Ao, where ^40 is the area before deformation. Strain is the displacement per unit length. Most commonly used is the engineering or nominal strain: A/
L
(10-1)
where / is the instantaneous gage length, generally calculated from displacement of the clamps or loading points. Alternatively, strain gages may be used to define nominal strain over a small section of the gage length (Rutherford and Brown, 1980). True uniaxial strain is the integral of infinitesimal nominal strains: d/
(10-2)
Since properties of viscoelastic polymers are rate dependent, the nominal strain rate sn is reported as the quotient of separation rate (dl/dt) over Zo. The true strain rate et can be controlled or determined with procedures discussed at the end of this section. Uniaxial compression (Fig. 10-3 b) employs relatively short samples with uniform circular or rectangular cross sections. Stress and strain are defined as above. Difficulties can be encountered with compressive buckling (alignment problems) and with frictional end effects. An alternative arrangement for compressive yield testing
(c)
433
(d)
Figure 10-3. Mechanical tests used to study yield in polymers: (a) tension; (b) uniaxial compression; (c) plane strain compression; (d) simple shear. (After Bowden, 1973.)
is plane strain compression, one arrangement for which is sketched in Fig. 10-3 c (Bowden, 1973). An obvious advantage is that the area Ao between the loading surfaces is constant, though friction between the dies and the polymer must still be reduced by a lubricant. Plane strain compression can be thought of as pure shear, as only two of the three sample dimensions are changed. Simple shear is more difficult to achieve and may be complicated by coordinate rotations at large strains. One scheme for imposing shear deformation is shown in Fig. 10-3 d. Shear stress is given by x — FjA, where the force F is applied in the plane of area A. Pure shear deformation does not affect A, so no distinction is required between true and nominal shear stress. Shear strain is defined as y = Ax/y, where Ax is the displacement of planes separated by a distance y; Ax is in the direction of the applied force and is perpendicular to y. Note that, as with plane strain compres-
434
10 Plastic Deformation of Polymers
sion, there is no change of sample dimension in the third (z) direction. The shear strain rate (time derivative of y) is written as y. In most instances tensile or compressive plastic deformation in polymers is macroscopically inhomogeneous, complicating evaluation of stress and strain from overall sample dimensions or changes in load point separation. For the most commonly encountered tensile experiment, multiple shear bands inclined ~ 45° to the loading direction coalesce to form the familiar "neck". Compression testing of rectangular samples frequently results in a single large shear band, while higher symmetry cylindrical samples bulge or "barrel" along the gauge section. Strains within these yielded regions are obviously larger than those calculated from overall sample dimensions. Uniaxial stress is larger in the necked down portion of a tensile specimen (a t >(7 n ). There is local thickening in compressive (shear) yield, leading to a lower true stress in the deformed section (at < an). True stress and true strain are required, however, to describe the material response in post-yield flow. There is no way to avoid inhomogeneous yielding and neck formation in tensile deformation; intrinsic material response dictates this behavior as described below. Before or up to the yield point the relative difference between true and nominal tensile-stress is of the order of the strain 2 < sy « 0.1. Thus the yield point can be determined with reasonable accuracy from force-displacement data; see for example Fig. 10-6 below. By careful control of sample and loading geometry to minimize stress concentrations, it is possible to achieve homogeneous uniaxial compression to / « 0.4 /0 (Boyce and Arruda, 1990). In this range it is straightforward to calculate true stress (and true strain) from overall sample dimensions. With properly
designed pure shear experiments, the stress T and strain y are uniform throughout the deformed section of the sample (G'Sell etal, 1983; G'Sell, 1988). Representative tensile stress-strain behaviors are given in Fig. 10-4. Curve A shows the case where yielding is followed quickly by failure; the yielded volume element has a reduced area A
Figure 10-4. Representative tensile stress-strain curves. (A) neck formation and fracture; (B) stable neck propagation; (C) uniform drawing without a neck. Dashed lines indicate magnitude of yield stress oy. Plastic strain ep is not recovered after load is removed.
10.2 Mechanical Testing and Definitions
yield and undergoes a reasonable amount of plastic strain sp before failure. This type of behavior is rare for reasons which will be considered below. Neck formation obscures the intrinsic response of the material because of the abrupt decrease in loaded area A of the deforming volume element. Considered construction (Fig. 10-5) can be used to understand some essential features (Vincent, 1960; Haward, 1973; Ward, 1983). True tensile stress at is plotted against nominal tensile strain en for cases sketched in Fig. 10-4. Material A has a yield point but no strain softening (decrease in true stress) and little strain hardening (increase in true stress). Throughout this chapter "strain softening" refers to a decrease in true stress with increasing strain (negative tangent modulus). Material Bx yields without strain softening but strain hardens thereafter, while B 2 displays both strain softening and strain hardening. Materials A, Bx and B 2 stretch inhomogeneously after neck formation at the maximum nominal stress on which is conventionally named yield stress crv. The condition for a maximum in
-1
o
Figure 10-5. Considered construction for tensile deformation. Conditions for maxima or minima in on are given by tangent lines from en = — 1. Material A forms unstable neck, Bt and B2 form stable necks, C deforms without a neck.
G
435
n = 0tO + O is easily shown to be represented by the first point of tangency of at with a line from e n =—1; this extrinsic yield point corresponds to the maximum load F or stress
436
10 Plastic Deformation of Polymers
are seldom applied. Some consequences of variable strain rate on neck formation are discussed in Sec. 10.6.3. Strain softening can be studied by yield behavior in compression or shear, where loaded area A cannot decrease. A load maximum in either of these geometries is sufficient to establish strain softening and intrinsic yield behavior.
300 -
10.3 Criteria for Yielding and Crazing Large scale or macroscopic plastic deformation occurs in polymers by shear displacements which are usually in the direction of maximum resolved shear stress (45° from a principal stress axis). A good example is afforded by polycarbonate at 77 K, well below 7g = 418 K (145 °C). As shown in Fig. 10-6, this glassy material yields in either tension or uniaxial compression at a yield stress a y « 220 MPa and at a yield strain sy« 0.1. Yield strength is somewhat higher in compression than in tension because of the pressure dependence of shear strength discussed below. Fracture occurs more readily in tension than in compression, as suggested by the data in Fig. 10-6. Tensile failure almost invariably involves craze formation and the growth of a crack through the craze (Chap. 15, this Volume). Crazes cannot be formed in pure compression or shear fields, so shear yielding always occurs in those geometries. A polymer may yield in uniaxial or biaxial tension by shear yielding or crazing, or both mechanisms may coexist. Conditions for shear yielding in isotropic polymers are best summarized by the pressure modified von Mises criterion (Raghava etal., 1973; Argon, 1975): a2)2
{a2 - <73)2
= 6T2
(10-3)
Figure 10-6. Stress-strain curves for polycarbonate at T = 11K in tension and uniaxial compression. Nominal stress on is given by dashed lines; true stress at by solid lines. The material fractures soon after yielding in tension. From Imai and Brown (1976); reproduced with permission. © John Wiley & Sons, Inc.
Here o-1? a2, and o3 are the principal stresses and ry is the stress required for yield in pure shear. Equation (10-3) corresponds to yield occurring when the elastic shear strain energy density in the stressed material achieves a critical value. The absolute value of Ty depends, of course, on temperature, strain rate, and pressure. Given the fact that polymers are for the most part van der Waals solids, it is not surprising that their mechanical properties are subject to hydrostatic pressure effects. The yield stress ry in Eq. (10-3) is observed to be pressure dependent: (10-4)
10.4 Yielding and Deformation Behavior
Here xy is the (strain rate dependent) yield stress at zero pressure and P = — (GX + G2 + <73)/3 is the negative of the hydrostatic part of the stress tensor. In a compressive-stress field (hydrostatic pressure P > 0), the strain energy terms on the left side of Eq. (10-3) must increase; yield is more difficult at higher pressures. The pressure coefficient /x has been determined by Bowden and Jukes (1972) to be of the order of 0.1-0.2 for most glassy polymers. One often employs a uniaxial stress field (G± = Gy, G2 = G3 = 0) for which Gy = ^3zy
(10-5)
It should be recalled that ry (and hence Gy) is strain rate and pressure dependent. For a representative pressure coefficient /i = 0.15, the uniaxial yield stress Gy will be 19% larger in compression than in tension. An example of this is seen in Fig. 10-6 for polycarbonate where the ratio of true yield stresses is, in fact, 1.19. For macroscopically isotropic polymers, yielding or plastic deformation of any sort cannot occur in purely dilatometric or hydrostatic stress fields. The shear yield surface is thus a tapering cylinder, centered on the GX = G2 = G3 axis, which broadens as applied pressure P becomes larger (increasing hydrostatic compression). A two-dimensional cut through this surface produces the elliptical line in Fig. 10-7. It should be kept in mind that the shear strengths depicted by such a line are for a set of pressures P (determined in part by
437
\ \ \
/A \
Figure 10-7. Section of the yield surface for the von Mises criterion, indicating yield stresses for uniaxial tension (A), biaxial tension (B), simple shear (C), uniaxial compression (D), and biaxial compression (E). Crazing will not occur below or to the left of the dashed line al = — o2.
discussed by Sternstein (1975) and reviewed by Kinloch and Young (1983) and Yee and Narisawa in Chap. 15 of this Volume. For the purposes of the present discussion, we note that crazing may occur in uniaxial or biaxial tension; this is the prime reason for limited tensile ductility of many polymers. Crazing cannot occur in pure shear (the line corresponding to G1 = — G2 in Fig. 10-7) or in any stress field to the lower left of that line.
10.4 Yielding and Deformation Behavior In this section we describe some important phenomenological aspects of plastic deformation of polymers. More complete reviews have been presented by Bowden (1973), Haward (1973), Ward (1983), Matsuoka (1986), and Brown (1986). Throughout this article uniaxial yield strength Gy and shear yield strength T are assumed to
438
10 Plastic Deformation of Polymers
be related by Eq. (10-5) for isotropic materials. This condition does not apply for the oriented systems treated in Sec. 10.5.3. 10.4.1 Glassy Polymers Uniaxial stress-strain behavior of polycarbonate depicted in Fig. 10-6 is typical of polymer glasses. Yielding in tension is followed almost immediately by fracture at this low temperature, while considerable post-yield flow occurs in compression. Other typical aspects are negative temperature dependence of yield strength (Fig. 10-8) and positive strain-rate dependence of ay (Fig. 10-9). Note that the tensile modulus E is not sensitive to either temperature or strain rate en for testing well below Tg. For polycarbonate the tensile yield stress drops to 45 MPa at 352 K (80°C), less than
one quarter of its value at liquid nitrogen temperature. Increasing 8n by four orders of magnitude increases oy by 35%, and the yield strain £y is increased by less than 20%. While the sensitivity of oy to temperature and strain rate varies from material to material, these features are common to all glassy thermoplastics and thermosets well below Tg. Bulk viscoelastic properties become more evident when the glass transition temperature is approached, as demonstrated by results in Fig. 10-10 for plane strain compression of polystyrene. The elastic modulus [E/(l — v2), where v is Poisson's ratio] is conspicuously temperature and strain rate dependent at 75-80°C because the material is fairly close to Tg = 100°C. The drop in true stress at after yield shows that polystyrene yields intrinsically
100 r-
80
60
5% per minute
20
12 Strain (%)
15
18
Figure 10-8. Tensile stress-strain curves for polycarbonate as a function of temperature, sn = 0.083/s. Lines are calculated from nonlinear viscoelastic models described in Sec. 10.6.1. From Matsuoka (1986); reproduced with permission of Carl Hanser Verlag.
10.4 Yielding and Deformation Behavior
5 min"
12 Strain (%)
Figure 10-9. Tensile stress-strain curves for polycarbonate as a function of strain rate at room temperature. Lines are calculated from nonlinear viscoelastic models described in Sec. 10.6.1. From Matsuoka (1986); reproduced with permission of Carl Hanser Verlag.
Strain rate Is'1)
60-
1.5 > 50
•£30
75 °C
20 80 °C 10
0.05
0.10 0.15 True strain
0.20
Figure 10-10. True stress-true strain curves for polystyrene determined in plane strain compression. From Bowden and Raha (1970); reproduced with permission of Taylor and Francis Ltd.
439
with strain softening, though the amount or intensity of strain softening is moderated at higher temperatures and lower strain rates. Note that no strain softening is observed at 80 °C for the lowest et. Virtually all glassy polymers strain soften after yielding in tension, compression, or shear (Haward, 1973; G'Sell et al., 1983). This lowered resistance to further plastic deformation leads to inhomogeneous stretching in regions where stress concentrations, etc., trigger the first yield events. The more compliant, strain softened regions are thus subjected to larger (local) strains and strain rates than the unyielded portions, resulting in the formation of microshear bands or shear patches (Bowden, 1973; Argon, 1975; Kinloch and Young, 1983). The true stress in Fig. 10-10 is seen to increase after a certain amount of strain softening. This strain hardening is essential to stabilize a region of plastic instability, particularly in tensile deformation; see Considered construction, Fig. 10-5. If the material did not strain harden, those regions that yielded first would continue to deform until a stretched zone failed by one of many possible mechanisms. It should not be surprising that strain hardening is seen in most polymers of commercial importance. Those without this characteristic would be difficult to process and unreliable in service. Haward (1973) has discussed strain hardening in detail, associating it with uncoiling polymer chains or sections of chains in the stretch direction. The chain orientation process is reversible, leading to recovery of macroscopic sample shape when the plastically deformed glass is heated above Tg (see Fig. 10-2). It should be remembered that the glassy polymer is still globally isotropic at, and just beyond, the yield point. Chain stretch and macroscopic
440
10 Plastic Deformation of Polymers
anisotropy are generated when post-yield flow is continued. Strain or orientation hardening of this sort can be treated in terms of the finite extensibility of chains, by analogy with rubber elasticity theory, and is considered in Sec. 10.6.2. Qualitative support for such an approach is given by the strain hardening characteristics of glassy epoxies crosslinked to different extents (Lohse et al., 1969). Increased crosslink density leads to more pronounced strain hardening at lower strains, the expectation if "network" strands in the glass were defined by crosslinks. The major mechanical distinction between glassy thermoplastics and thermosets is in strain-hardening characteristics. Crosslinking by itself has little or no fundamental influence on yielding or strain softening, though ay at a fixed temperature may be affected if the glass transition temperature and elastic modulus (E or G) are changed. This effect has been demonstrated nicely by tensile-yield behavior of (uncrosslinked) poly (methyl methacrylate) reported by Hope et al., 1980. The presence of 6% monomer plasticized the glass, lowering Tg from 110°C to 90°C and decreasing
amorphous regions between lamellar crystals. The crystals, particularly those for chains preferring extended or all trans conformations such as polyethylene, polyamides (nylons) and poly(ethylene terephthalate), are very anisotropic. Melt crystallized polymers generally exhibit a spherulitic morphology; ribbon-like lamellae are disposed radially in the polycrystalline aggregate (Peterlin, 1987; Chap. 4, this Volume). Spherulite diameters are normally in the range of 2-20 jim. Other relevant sizes are 10-25 nm for crystal thickness /c in the chain axis direction, ~ 0.1 - 1 |im for the transverse dimensions of the lamellae and 5-10 nm for amorphous layer thickness /a. A deviatoric-stress field causes the spherulite to change shape in an affine manner at low (elastic) strains. Yielding is inhomogeneous within a spherulite (Samuels, 1974; Schultz, 1974) and requires irreversible shear deformations in both the amorphous and crystalline regions. Lamellar bundles oriented at 45° to the principal stress direction experience the largest resolved shear stress, deforming by interlamellar shear (stretching of the amorphous regions between crystals) or intracrystalline shear (sliding of polymer chains along planes containing the covalently bonded molecular axis). Deformation continues until the original spherulites are transformed to well oriented microfibrils with transverse dimensions of the order of 100 A (Peterlin, 1971,1987). Recent studies on nylon 6 (Galenski et al., 1991) and polyethylene (Galenski et al., 1992) provide good summaries of techniques and models used to evaluate morphology changes which occur during deformation of semicrystalline polymers. As with glassy polymers, plastic deformation causes chain axes to become aligned in the stretch direction. Besides this orientation effect, there are a number of
10.4 Yielding and Deformation Behavior
other morphological changes. We concentrate here on the "cold drawing" regime above Tg of the amorphous regions and usually not too far below the melting temperature Tm of the crystals. Here tensile deformation proceeds with a stable neck giving a "natural draw ratio" A = exp(at), where et is the true tensile strain after necking. Structural changes resulting from this type of deformation have been studied and described many times (Samuels, 1974; Heisse et al., 1977; Peterlin, 1987); here we review the most salient features. The isotropic, usually spherulitic, microstructure is converted to microfibrils, generally with no significant change in crystallinity. The crystalline-amorphous long period L = lc + /a changes to a value determined by the draw temperature, independent of the original L in the undrawn polymer. Chain orientation in the stretch direction is greater for the crystalline segments, though amorphous orientation continues to improve as X is increased at a fixed temperature. Transverse crystal dimensions (perpendicular to the chain axis) are decreased. Macroscopic sample dimensions as well as most aspects of the original isotropic structure can be recovered by heating the drawn polymer above Tm. The significance of these effects is considered in Sec. 10.5.2. There are two important temperatures for semicrystalline polymers, Tg for the amorphous fraction and Tm for the crystals. A major distinction applies above and below the glass transition temperature; below Tg both components are solids, while above Tg the amorphous component is liquid-like. For Tg
441
sponse of the liquid-like phase is sensitive to temperature and strain rate, particularly in the range just above Tg. When T < Tg the mechanical properties of the two phases (glassy and crystalline) are relatively similar. Here properties are less dependent on degree of crystallinity and, to a first approximation, are comparable to those of glasses. An example of the latter type of crystalline-glassy polymer is poly(butylene terephthalate), for which T g =87°C and Tm=33O°C. The temperature dependent tensile behavior in Fig. 10-11 is similar to that seen for polycarbonate in Fig. 10-9; oy and sy both increase as the temperature is lowered, whereas the tensile elastic modulus is virtually independent of temperature. Note the "glassy" yield strain values of sy « 0.05-0.10. The load drop after yield is more pronounced at lower temperatures. All these features are generally consistent with the behavior of polymer glasses (Bowden, 1973; Haward, 1973; Matsuoka, 1986). As a rule, the amount of tensile ductility in semicrystalline polymers below Tg is small, though considerable post yield deformation can be achieved in compression or shear. It is important to keep in mind that chains in both glassy and crystalline regions are involved in the yield process. Semicrystalline polymers above Tg behave quite differently. Tensile results for high-density polyethylene are given in Fig. 10-12. One of the most conspicuous features is the temperature dependence of the tensile modulus, a result of linear viscoelastic processes in both phases (Boyd, 1985). Note that the yield stress is rather small; oy« 30 MPa at room temperature and drops to zero as Tm is approached. For comparison the (compressive) yield stress of poly(methyl methacrylate) is 50 MPa at 100°C, only 10° below Tg (Bowden, 1973). Yield strain ey, defined at the maximum of
442
10 Plastic Deformation of Polymers
120
100 -
80 -
Figure 10-11. Tensile stress strain curves for poly(butylene terephthalate) as a function of temperature, sn = 0.086/s. Lines are calculated from nonlinear viscoelastic models described in Sec. 10.6.1. From Matsuoka (1986); reproduced with permission of Carl Hanser Verlag. 10 Strain (%)
load or <jn, can be as large as 0.2 at sufficiently high temperature or low strain rate. Lowering the temperature causes the (extrinsic) yield stress ay to increase, the sharpness of the load drop to increase, and ey to decrease. The first two of these effects are common to glassy polymers, and the "negative" shift in yield strain is seen in glasses near Tg (see for instance Fig. 10-10). Tensile 160
1
1
1
1
I
1
1
77
140
1 -
120
2
IOO
u 80
• A~
-
172 "^^^.^^ ^215
CD
£ 60
= ^ >
40 20
-
294
^ 3TT-
wr^K-'^v-—r-
~0
2
4
6
81 10 1 12 14 Strain (%)
16
18 20
Figure 10-12. Tensile stress-strain curves for polyethylene as a function of temperature, en = 3.3 x 10~ 4 /s. Material fractures before yield at T<100K. From Kamei and Brown (1984); reproduced with permission. © John Wiley & Sons, Inc.
ductility vanishes as Tg (about 160 K for linear polyethylene) is approached from above. From the curves in Fig. 10-12, "cold drawing" with stable neck propagation is achieved at temperatures above 240 K. Strain-rate effects (Fig. 10-13) generally mimic temperature effects, with a higher en corresponding to a lower temperature. In semicrystalline polymers there is the opportunity to vary crystallinity by either changing cooling rate or incorporating comonomer or other irregularities in the chains. Crystallinity effects have been studied extensively in polyethylene and its random copolymers, for the most part at room temperature. Increasing crystallinity increases the yield stress (Fig. 10-14) and the sharpness of the load drop which accompanies neck formation (Popli and Mandelkern, 1987; Crist et al., 1989). Tensile yielding in polyethylene is extrinsic; in contrast to glassy polymers, there is no strain softening or drop in true stress (Meinel and Peterlin, 1971; G'Sell and Jonas, 1979; Coates and Ward, 1980). Strain hardening is obviously present to stabilize the neck.
10.4 Yielding and Deformation Behavior
J
I
L_
443
Figure 10-13. Tensile stressstrain curves for polyethylene as a function of strain rate, T = 25 °C. Lines are calculated from nonlinear viscoelastic models described in Sec. 10.6.1. From Matsuoka (1986); reproduced with permission of Carl Hanser Verlag.
10 Strain (%)
One interesting observation is that the strain-hardening rate is greatest in copolymers of low crystallinity, as shown in Fig. 10-15. A similar effect is seen in ultrahigh
0.90
0.92
0.94
0.96
p (g/cm3)
Figure 10-14. Tensile yield stress as a function of density Q of polyethylene and random copolymers of polyethylene. Crystallinity a varies from 0.35 and 0.70 and crystal thickness lc varies from 4 to 27.5 nm over the reported density range. •, o linear polyethylene; o, • polyethylenes with short chain branching; A, A, • polyethylenes with short long chain branching. From Crist etal. (1989).
molecular weight polyethylene, which also has a crystallinity less than 50% (Meinel and Peterlin, 1971). Strain hardening in semicrystalline polymers is discussed in Sec. 10.6.2. Strain hardening in semicrystalline polymers also depends on the type of deformation, again in contrast to glasses. G'Sell etal. (1983) have measured shear stressshear strain relations for a large number of polymers shown in Fig. 10-16. Each glassy polymer yields in shear with a stress drop (strain softening) followed by strain hardening. Semicrystalline polymers have no strain softening in shear, which is consistent with tensile behavior (see Fig. 10-15). Note, however, that no strain hardening is observed in post-yield shear deformation for either polyethylene or polypropylene. Since both of these materials draw in tension with stable neck propagation, there is considerable strain hardening in tension; this is shown explicitly for polyethylene in Fig. 10-15. Different strain hardening characteristics in tension and shear are discussed in Sec. 10.6.2.
444
10 Plastic Deformation of Polymers r
LU 0.5
(r r = 2.6) (r r = <
150 LU 7
100
50
5
10
2 3 Shear strain
Figure 10-15. True stress-true strain curves for polyethylene and copolymers at room temperature. Sample designations give number of short chain branches/1000 carbon atoms; density ranges from 0.918 g/cm 3 (LU34) to 0.961 g/cm 3 (LU0.5). From Meinel and Peterlin (1971); reproduced with permission of Pergamon Press pic.
Figure 10-16. Shear stress-shear strain curves for various polymers at room temperature. (1) Polyoxymethylene, (2) poly (methyl methacrylate), (3) nylon 66, (4) polycarbonate, (5) poly (vinyl chloride), (6) polybutene-1, (7) polypropylene, (8) polyethylene. From G'Sell et al. (1983); reproduced with permission of Chapman and Hall.
Plasticizers may have an appreciable effect on the deformation of semicrystalline polymers if the system is tested near the glass transition region of the amorphous component. A good example is nylon 6, for which Tg = 75 °C when dry and Tg = 20°C when equilibrated at 50% relative humidity (water content is about 3%). Tensile yield-stress at room temperature is decreased from 80 MPa to 44 MPa by the addition of this amount of plasticizing water (Bonner et al., 1973). Adding a plasticizer to such a polymer is thus equivalent to raising the deformation temperature. (See the strong influence of temperature on yield strength in polyethylene for ^ r g , Fig. 10-12.)
as extensively as the single polymer systems described above. A good review of earlier work can be found in Manson and Sperling (1974). The classic example of a multiphase polymer-polymer system is rubber toughened polystyrene. Incorporation of 0.1 volume fraction of ~ 1 jim droplets of elastomer (e.g., polybutadiene) decreases the yield stress and increases toughness remarkably. Unmodified polystyrene fractures in tension by crazing at very small strains, while rubber modified polystyrene can be elongated by ~ 50%. Argon and Cohen (1989) have shown that this macroscopic strain comes from a large number of crazes originating at interfaces between the glassy matrix and the more compliant inclusions; crazing efficiency is calculated in terms of particle size and relative elastic moduli. The crazing strain acts as a stress relief mechanism, preventing or retarding fracture. Another important aspect of the polystyrene-polybutadiene system is plas-
10.4.3 Blends and Block Copolymers
Multicomponent polymer systems, most of which have more than one phase, are of great technological importance. These complex materials have not been studied
10.5 Fundamental Nature of Polymer Yielding
ticization of the matrix by the rubber molecules, which can dissolve in polystyrene under hydrostatic tension. The behavior of other binary systems is not so well understood. Most polymerpolymer blends are macroscopically phase separated with limited mutual solubility at the interface. Generally the yield (or fracture) stress of such blends is below that predicted by linear combination of component properties. The critical role of the interface in such blends can be modeled in terms of thermodynamic parameters. Pukansky and Tiidos (1990) treat the ability of stress to be transmitted from phase to phase in terms of interfacial surface energy. This approach appears to work well for the composition dependence of strength in immiscible systems, for which it was designed. It accounts for the properties of a single phase (miscible) blend as well, probably because the macroscopic sample is implicitly treated as an "interface". Chow (1990) has adopted a different method for treating the composition dependence of ay in glasses of miscible polymers. A nonequilibrium interaction between chains is invoked for the glassy state; this excess attraction is manifested as a minimum in specific volume or a maximum in yield stress at intermediate compositions. Chow's scheme effectively correlates nonadditivity of volumes with enhanced yield strength. Interface strength is not an issue in triblock copolymers with long enough block lengths, wherein covalently bonded chains traverse microdomains having dimensions ~10nm, governed by the size of blocks or subchains. It is possible to combine "glassy" and "rubbery" blocks over a large range of volume fractions and morphologies (Chaps. 1 and 6 of this Volume). Of most interest are thermoplastic elastomers which contain a large amount of rubbery material. These have some unique mechan-
445
ical properties which are exemplified by the styrene-butadiene triblock system studied by Kawai and Hashimoto (1979). Solvent-cast block copolymer has well formed alternating lamellae of polystyrene (Tg = 100°C) and polybutadiene (Tg = — 90 °C), each of thickness ~ 25 nm. This material yields in tension, oy = 5 MPa, sy = 0.05, with neck formation qualitatively similar to glasses or semicrystalline polymers. Strain is largely reversible, however, recovering from en = 3.5 to sn = 0.2 (94% strain recovery) with significant hysteresis. This retraction is attributed to the elastomeric component. The yield mechanism was not identified, but post-yield stretch involved rotating and shearing the "rigid" polystyrene lamellae into smaller blocks. Original morphology (and presumably mechanical properties) was recovered by annealing at 60°C for a few minutes. The tensile behavior of block copolymers approximates that of nonlinear elastic models used to treat inhomogeneous plastic deformation (Sec. 10.6.3).
10.5 Fundamental Nature of Polymer Yielding Macroscopic yielding requires local shear displacements and establishment of new intersegmental bonds as sketched in Fig. 10-1. These occur in both glassy and semicrystalline polymers in times as short as seconds, as seen from the stress-strain curves presented in Sec. 10.4. There are two broad categories of explanations for rearrangements at temperatures where (stress free) chains are effectively immobile. One proposes that the stressed material is converted to the liquid state, after which the molten chains respond quickly to the applied stress, then solidify. The second invokes various molecular mechanisms to
446
10 Plastic Deformation of Polymers
account for rapid chain rearrangement in the stressed solid. 10.5.1 Glassy Polymers Yielding mechanisms in glassy polymers have been reviewed by Bowden (1973), Ward (1983) and Kinloch and Young (1983). The earliest ideas were based on mechanical work raising the sample temperature to Tg, or on free volume from dilatometric strain bringing Tg down to the test temperature. In either case chain stretching would occur quickly in the stressed "glass". While both these effects may be important under certain conditions, they cannot be the fundamental reason for yield in polymer glasses. The mechanical work done on a glass is generally insufficient to raise the sample temperature to Tg, even in the adiabatic limit (Haward, 1983). Yielding furthermore occurs at low strain rates or with thin samples under conditions which are isothermal (T
solute rate theory (Bowden, 1973; Argon, 1975; Ward, 1983). Yielding is described as viscous flow in which the activation barrier for local shear displacements (analogous to those in Fig. 10-1) is decreased by the applied stress. For an experiment performed at a shear rate y under a shear stress T, which corresponds to the flow stress or yield stress i y in this case, the strain rate is given by (,0-6) Here f is a fundamental rate factor, AH is the activation energy for the flow process (in the limit of T = 0), v* is the "activation volume" and k is Boltzmann's constant. It is assumed in Eq. (10-6) that D * T > / C T , meaning that reversal of shear steps (back reaction) is neglected. This analytical scheme describes the temperature and strain-rate dependence of yield stress and the (phenomenological) shear strength T 0 at 0 K: • = Tn +
d In y
2kT
2kT v*
2 AH tn
=
In
(10-7 a)
V
(10-7 b) (10-7 c)
The activation volume v* is readily evaluated from the experimental strain-rate dependence of Ty through Eq. (10-7 b). The remaining parameters f and AH are obtained from the temperature dependence of Ty. Note that athermal strength T 0 is proportional to AH, the barrier for intersegmental rearrangements. The Eyring model accounts for many features of yielding in glassy polymers, notably the negative temperature dependence [Eq.(10-7a); ln(y/T)<0] and positive strain rate dependence [Eq. (10-7 b)] of the
447
10.5 Fundamental Nature of Polymer Yielding
yield stress. The basic concept is that yielding and plastic deformation of the glass occur by viscous flow which is activated by the combination of applied stress and thermal fluctuations. Interpretation of the kinetic parameters in physical terms is difficult, though the values of v* are found to be the size of a few repeat units. These and other aspects of the Eyring model have been discussed by Bowden (1973) and Ward (1983). A more specific model for yield of glassy polymers, also based on applied stress and fluctuations of thermal energy, has been developed by Argon (1973, 1975). It is proposed that local shear displacements for glassy, interpenetrating polymer molecules can be represented by the rotation of a chain segment of length z into the direction of principal stretch. Rotation is accomplished by nucleation of a pair of kinks, each characterized by a bending angle co « 2 radians, in a chain segment. This value of w was chosen to correspond to valence bond angles of about 115°, but rotation about single bonds in the backbone is a more feasible mechanism for changing the shape of a chain. In this sense the kink pair is analogous to a "crankshaft" defect. The resultant strain is proportional to 71 a2 CD2 z, where a is the radius of one chain or a cooperatively deforming bundle of chains. Argon calculates the net energy A U of a kink pair of size z, formed under a stress T in an elastic medium having shear modulus G and Poisson's ratio v. Resistance to kink formation is intermolecular, as is the case with the Eyring model. The expression for A U includes the self-energy of the kink pair and the kink-kink interaction energy, reduced by the (plastic) work done on the system by the stress T in forming a kink pair of size z. AC/ goes through a maximum with respect to z as the kink pair increases in size. The maximum value
AC/* is the activation barrier for nucleation of a stable kink pair, i.e., one which grows beyond the critical value z*. Yielding occurs when thermal fluctuations over the stress dependent activation barrier AC/*(T) create kink pairs at a rate which provides the macroscopic strain rate y: y=Texp
H^]
(10-8)
In this expression the activation energy AC/*(ry) is made smaller by the applied stress (here equated to the yield stress) in a manner similar to the Eyring formulation in Eq. (10-6). The shear stress required to achieve a particular strain rate is 16/5
(10-9) where the athermal shear strength T0 is 0077 G u
1-v
,10-lOJ
Note that Argon's kink pair nucleation model [Eq.(10-9)] and the Eyring treatment [Eq. (10-7 a)] have similar dependencies of i y on both temperature T and strain rate y; the exponents are 1.2 and 1.0, respectively. The kink pair nucleation model has been used by Argon (1973, 1975), Argon and Bessonov (1977), and Yamini and Young (1980) to interpret the temperature dependence of Ty in a number of glassy polymers. Equations (10-9) and (10-10) require only two adjustable parameters, the frequency factor f and the product co2 a3 (or a3, since co = 2), provided that G and v are known as functions of temperature. Fits of the model to experimental data are generally excellent. Derived parameters are reasonable, with f & 10 13 s" 1 and the effective chain
448
10 Plastic Deformation of Polymers
radius a ranging from about 0.8 to 3 times the single chain radius a0. The model also predicts absolute shear strength in terms of the elastic shear modulus G. Evaluating Eq. (10-10) with a representative value of Poisson's ratio v « 0.38 (Brown, 1983), the athermal strength at T = 0K becomes 0.077 G ;0.12G 1-v
(10-11)
This affords another check on the kink nucleation model, provided that low temperature values of ry and G are available. A third approach to stress and thermal activation of yield was proposed by Bowden and Raha (1974). The energetics of elementary shear displacements are treated in terms of a dislocation loop of radius R forming and growing under the combined effects of shear stress T and thermal energy kT The (unknown) radius of the dislocation core, r0, is replaced by y/3 b/e (b is the modulus of the Burger's vector of the dislocation and e is the base of natural logarithms) to conform to the maximum shear strength: = 0.184 G
(10-12)
predicted for a close packed atomic crystal. The analysis proceeds by evaluating the nucleation barrier
.__, Gb2R* (2R* AU*= In bJ3
(10-13a)
for creating a dislocation loop of the critical size R*: R* =
Gb 4 TIT,
HSH
(10-13 b)
Equations (10-13 a) and (10-13 b) cannot be rearranged to solve for ry analytically. The model is fitted by adjusting the Burger's vector b and the ratio AI7*/(feT) to con-
form to the experimental temperature dependence of i y at a particular strain rate y. The critical activation barrier is frequently chosen as 50 kT (Bowden and Raha, 1974), which leaves only one disposable parameter, b. This choice of At/* establishes the strain rate dependence of r y , as well. It should be emphasized that Bowden and Raha (1974) do not contend that dislocations of the classical sort exist in polymer glasses. Implicit, however, is that elementary shear displacements occur in a spatially correlated linear domain which can close on itself to form a loop. Description of the energy of this configuration by a dislocation model is not based on micromechanics which are unique to crystals. Dislocation theory is quite amenable to continuum treatments; only far field elastic effects are considered in terms of modulus G and Burger's vector b, the nature of the dislocation core being ignored aside from its radius r0. In crystals the Burger's vector b has a well defined length imposed by structural periodicity. The Burger's vector in Eq. (10-13) should be thought of as an average step size for shear displacements in the glass. The derived values of b are consistent with the size of the repeat unit in the chains (Bowden and Raha, 1974; Yamini and Young, 1980). The three thermally activated yield models are compared in Fig. 10-17. Calculated strength i y is normalized by the athermal yield stress r 0 in Eq. (10-10). This ratio exceeds 1 for the Bowden-Raha model because the normalization factor is smaller than that defined in Eq. (10-12). Parameters are based on compressive yielding of poly (methyl methacrylate) at e «10~ 3 /s, f = 1013/s, v* = 1.75 x 10" 2 7 m 3 (Bowden, 1973), and T 0 = 0.43 GPa (Brown, 1983) in Eq. (10-7 a); r = 10 13 /s, co2 a3 = 1.95xlO" 2 8 m 3 (Argon, 1973), G = 3.0 GPa, and v = 0.34 (Brown, 1983) in Eq.
10.5 Fundamental Nature of Polymer Yielding
(10-9); b = 2Jx 10" 1 0 m, At/* = 50 fcT(Yamini and Young, 1980), and G = 3.0GPa (Brown, 1983) in Eqs. (10-13 a) and (10-13 b). The format of Fig. 10-17 compares only the intrinsic temperature dependence associated with thermal fluctuations over activation barriers. Actual temperature dependence of Ty will be more pronounced for the kink pair and dislocation loop models because the shear modulus G decreases at larger T. Similarly, the activation volume y* in the Eyring analysis appears to become smaller at higher temperatures, giving rise to a larger decrease in ry than indicated in Fig. 10-17. As mentioned above, experimental distinctions between the Eyring and Argon treatments are too subtle to permit discrimination between them. It should be appreciated, however, that the kink pair model establishes both the absolute value of Ty and its temperature dependence with measured elastic constants and explicit molecular parameters. The fitting parameter a = 3.7 x 10 " 1 0 m is identical to the ra-
o
200 300 Temperature (K)
Figure 10-17. Temperature dependence of yield stress calculated by the Eyring model ( ), the kink pair nucleation model (—) and the dislocation loop nucleation model (--). Parameters used are appropriate for compressive yield of poly (methyl methacrylate).
449
dius of a single poly (methyl methacrylate) chain, though other relative sizes are seen with different polymer glasses (Argon and Bessonov, 1975). The Eyring model accounts for a nearly identical temperature dependence of ry in terms of an "activation volume" which corresponds to the size of about 12 monomer units in poly (methyl methacrylate); the physical significance of v* has not been established. It is clear that the temperature effect is much different for the dislocation model, which has a large, negative slope of the plot at low T. This characteristic feature has not been observed, though it is difficult to obtain data at low temperatures where the effect is most pronounced. Yield stress ry near 300 K is appreciably lower than that of the other models, but this results from the temperature dependence of G having been ignored when establishing b = 2 . 7 x l 0 ~ l o m (Bowden and Raha, 1974). The high temperature value of Ty/T0 can be adjusted to ~ 0.6 if the Burger's vector b is increased by 40%, which is still about the size of a monomer unit in poly (methyl methacrylate). An additional comparison can be made on the basis of strain rate dependence expressed as dT y /dlny. The Eyring analysis gives 4.6 MPa [which is the experimental value, see Eq. (10-7 b)], the kink pair model 3.2 MPa, and the dislocation loop model 2.0 MPa, all at 300 K. Agreement between Eyring and Argon is improved if the shear modulus G is dropped to the room temperature value. The Eyring treatment is strictly phenomenological and it predicts that the deformation is homogeneous; there is no mechanism for intrinsic yield or strain softening. Argon's model is molecularly based. As presented here, deformation is also globally homogeneous and incapable of strain softening. Once a segment of critical size z* has rotated, no further strain results
450
10 Plastic Deformation of Polymers
from separation of kinks (z>z*); additional strain increments are achieved by nucleating additional kink pairs with the same energy penalty. The dislocation approach, on the other hand, does predict that plastic strain is inhomogeneous, being confined to dislocation loops which can grow to micrometers in diameter. The amount of strain per nucleus is not determined explicitly. Argon (1973, 1975) and Bowden and Raha (1974) both consider how strain softening may occur in the context of their respective models, though these concepts have not been used to account for experimental data. The three treatments of yielding in glasses discussed above are based on the applied stress modifying intermolecular interactions. An entirely different approach has been offered by Robertson (1966). This is a uniform deformation model, in the sense of the Eyring treatment, but the applied stress is considered to increase the intramolecular energy of chains, which have high energy (flexed) and low energy (unflexed) rotational states. The ensemble of stressed chains with this additional energy can flow at a rate which would be achieved by stress-free chains at a fictive temperature 9 > Tg. At this fictive temperature the solid behaves as a Newtonian fluid: T=rj(6)y
(10-14)
The 9 dependence of viscosity rj is calculated from the empirical WLF equation (Williams et al., 1955) with no adjustable parameters; universal constants Cx = 17.44 and C2 = 51.6° are employed together with the characteristic viscosity rj(Tg) = 1013'6 Pas at the glass transition temperature. This model is remarkably successful, accounting for the absolute value of Ty within a factor of 2, as well as its temperature and strain rate dependencies with even better
accuracy. Such agreement is seen within about 80 °C of the glass transition temperature Tg; Argon and Bessonov (1977) have discussed the shortcomings of Robertson's treatment at lower temperatures. From the comparisons in Fig. 10-18, Robertson's model seems superior to the double kink model as Tg is approached, while Argon's kink model captures the behavior at lower temperatures where proximity to Tg is unimportant. These behaviors stem from the "reference states" used in the two treatments. Argon's model is based on shear strength T 0 at T = 0 K, while Robertson's model uses the flow stress at Tg. Brown (1983) has considered the maximum shear strength T 0 at 0 K of polymer glasses having three elementary displacement mechanisms operating in parallel. The dominant one is concluded to be conventional intersegmental shear as represented schematically by Fig. 10-1. The critical stress to achieve flow or yield is estimated to be T 0 = (0.064-0.092) G, which compares favorably to the experimental average of T 0 = (0.074 ± 0.03) G for 6 polymers for which ry was extrapolated to OK. It can be concluded that polymer glasses yield at stress levels which are remarkably close to theoretical estimates of strength. Experimental athermal yield-stress T 0 « 0.074 G agrees with Brown's estimate and is only 40% below the prediction from the more comprehensive kink pair model, Eq. (10-11). The dislocation model is based on a rather arbitrary definition of T 0 [Eq. (10-12)] which is about 2.5 times larger than experiment; this discrepancy could be moderated, without effect on other aspects of the analysis, by assigning a larger value to the radius r0 of the dislocation core. It is believed, however, that the extreme temperature dependence of that model (see Fig. 10-17) makes it an unlikely candidate
10.5 Fundamental Nature of Polymer Yielding
for correctly describing yield in polymer glasses. Confining our attention to the Argon and Eyring treatments in Fig. 10-17, thermal activation decreases i y by only about 45% at T = 350 K, thus yield occurs at more than half the maximum theoretical strength. Larger drops in Ty, seen for instance in Fig. 10-18, result from the temperature dependence of material parameters G or AH and v*. Another indication that glasses yield near theoretical values is the magnitude of the yield strain e y ^0.1 (Sec. 10.4.1). Converting this to an equivalent shear strain one obtains yy = ^/3 8 y ^0.17, which is of the order of yy = 0.25 for the basic shear model sketched in Fig. 10-1. The discussion in the preceding paragraph is based on stress (and thermal energy) overcoming intersegmental interactions to cause shear deformation. Elastic energy barriers drop very rapidly as Tg is approached, leading to underestimation of ry. In this region Robertson's intramolecular model accounts better for observed behavior.
451
10.5.2 Semicrystalline Polymers The complexity of the deforming system - a representative region contains an anisotropic lamellar crystal, amorphous chains, and two crystalline-amorphous interfaces - has frustrated efforts to define the basic yield mechanism or mechanisms in semicrystalline polymers. Most experiments have been done on polyethylene and polypropylene for which the amorphous regions lie above Tg at room temperature; in this range one achieves very large stretch ratios {sn > 4) which cause useful increases in stiffness and strength. To a first approximation there is no yielding in the liquidlike amorphous component of such composite materials; stresses there are accommodated by elastic or viscoelastic strains. This simplification is consistent with the observation that tensile and shear elastic moduli of polyethylene are dominated by the compliance of the amorphous component, meaning that the mechanical response at small strains is approximated by a series model (Boyd, 1983; Crist et al., 1989, Chap. 7, this Volume).
0.12
fc 0.10
-
Figure 10-18. Normalized shear yield strength for poly (methyl methacrylate) and polystyrene as a function of temperature. The shear modulus G at 0 K is represented by the symbol (x. Lines are calculated from kink-pair (Argon, 1973) and intramolecularflex-energy (Robertson, 1966) models. Fom Argon and Bessonov (1977); reproduced with permission of Taylor and Francis Ltd.
."5 0.08
0.06
0.04 PMMA o Argon o
PMMA • Robertson 0.02
PS O Argon PS a Robertson i
1
100
200 Temperature (K)
300
400
452
10 Plastic Deformation of Polymers
It has long been recognized that the yield stress increases with crystalline fraction a when deformation is done at Tg < T
Evidence supporting the partial melting and recrystallization mechanism for yielding has been reviewed thoroughly by Popli and Mandelkern (1987). Here we emphasize one of the most compelling facts, the observation that the long period L=lc+la is a function of deformation temperature Td, essentially independent of L in the isotropic polymer before drawing (Peterlin, 1971). This is observed for plane strain compression (Bessell and Young, 1974) and solid state extrusion (Chuah et al., 1986) as well as uniaxial tensile drawing. The decrease in long period L seen in such experiments is a particularly strong argument for the deformation stress mobilizing chain segments in a manner very similar to melting. Chuah et al. (1986) term this "quasimelting", and note that it does not occur in polyethylene which was crystallized in extended chain (as opposed to lamellar) morphology. Something akin to quasi-melting can be seen in the classic model of Peterlin (1965) for the transformation to oriented fibrils (Fig. 10-19). Phillips and Philpot (1986) have reported electron microscopy evidence for a molten zone some 5-10 |im in extent for polyethylene drawn at room temperature. This is consistent with the partial or local melting hypothesis of Popli and Mandelkern (1987). Additional support for quasi-melting is provided by small-angle neutron scattering (SANS) from mixtures of conventional and deuterated polyethylene. These experiments rely on the tendency of the two isotopic forms of polyethylene to segregate during crystallization, though statistical cocrystallization can be achieved by rapid quenching from the melt. Wu and Wignall (1985) and Wu et al. (1992) have examined intentionally clustered mixtures before and after plane strain compression at room temperature. They find that the amount of isotopic segregation, determined from the
453
10.5 Fundamental Nature of Polymer Yielding
absolute intensity scattered at zero angle, is reduced by as much as a factor of five after plastic deformation. This intensity drop commences at yield and increases continuously with further deformation. From considerations of intensity and "cluster size" it is argued that chain mixing of this sort must result from melting during the deformation process. Sadler and Barham (1990 a, b) have done related SANS experiments with isotopic mixtures of polyethylene subjected to tensile drawing at or above room temperature. Their analyses consider changes in the radius of gyration or cluster size, not the absolute scattered intensity. For k^l they find evidence for melting and recrystallization when drawing is done above a fairly well defined temperature in the range from 65°C to 95°C; this critical temperature increases with molecular weight (64000-350000 g/mol). It is concluded that melting and recrystallization do not occur during tensile drawing to X K, 7 at room temperature or on further stretching to X « 40 at Td < 118°C. Thus two sets of SANS experiments each support somewhat disparate concepts of melting and recrystallization. It is not known whether these differences result from unlike deformation modes, or from peculiarities of the SANS experiments. Alternatives to the melting or phase transformation model are based on more conventional approaches to crystal plasticity. These include stress induced crystalcrystal transformations, twinning and slip which have been reviewed by Bowden and Young (1974), Young (1979), and Saraf and Porter (1988). Slip of covalently bonded chains on (h k 0) planes in either the chain axis direction [001] or transverse directions [u v 0] have received the most attention. It is assumed that the crystal structure has been defined conventionally with the chain axis in the crystallographic [001] direction.
(hkO) [001] slip is the most likely mechanism for achieving large macroscopic strains with conversion to a fibrillar morphology as suggested in the Peterlin model, Fig. 10-19. Furthermore, this shear mechanism can be used to provide quantitative estimates of the yield stress t y , which can be measured with considerable precision for various slip systems using macroscopically textured samples (Bartczak et al., 1992 b; Lin and Argon, 1992). The relation between (hkO) [001] slip and the experimental yield stress of about 5-30 MPa in polyethylene can be established with a dislocation nucleation model. Shadrake and Guiu (1976) calculated the energy of a [001] screw dislocation in a lamellar crystal of thickness lc interacting with a free (hkO) surface or another [001] screw dislocation. A sketch of the system is shown in Fig. 10-20. Under the influence of a shear stress T, the net energy AU for creating a shear region of size u is written in terms of T, lc, elastic modulus G for shear on (h k 0) planes, Burger's vector b (equal to the crystallographic c axis in the chain direction), and dislocation core radius r0. This model is developed similarly to the
(a)
(b)
Figure 10-19. Model for transformation from lamellar (a) to fibrillar (b) morphology on drawing a semicrystalline polymer. From Peterlin (1965); reproduced with permission. © John Wiley & Sons, Inc.
454
10 Plastic Deformation of Polymers
Figure 10-20. Sketch of [001] screw dislocation nucleus of size u in a lamellar crystal of thickness /c. Magnitude of the Burger's vector b is exaggerated. From Crist et al. (1989).
Bowden-Raha analysis of the dislocation loop in Sec. 10.5.1. The nucleation barrier for creating a shear region of critical size u* is AU* =
2TT
10 d (nm)
Figure 10-21. Dependence of yield stress of polyethylene crystals on crystal thickness lc (here represented by d) at room temperature. • linear polyethylene; • branched polyethylene. Line is calculated from Eqs. (10-15 a) and (10-15 b). From Young (1988); reproduced with permission. © Institute of Metals and Materials Australasia.
(10-15 a)
The yield stress at which a stable nucleus of size w* is formed by thermal fluctuations is Ty
Gb 2nu*
(10-15 b)
where w* is defined through Eq. (10-15 a). Young (1988) and Crist et al. (1989) have applied this dislocation nucleation model to the experimental yield stress of polyethylene. The only assumption is that the nucleation barrier A I/* is set at 50 kT to 60 kT so laboratory scale strain rates of ~ 10~ 3 / s c a n be achieved when the fundamental frequency is ~ 10 13 Hz. All structural and elastic parameters are known with reasonable accuracy; in neither study were model parameters adjusted to fit observed results. One sees immediately that the absolute magnitude of ry is quite good, though the model has too large a dependence of Ty on crystal thickness (Fig. 10-21) and too small a dependence on temperature (Fig. 10-22). The steep rise in the experimental strength ry for T < 175 K is due to strengthening from the vitrifying amor-
200
300
400
7"(K)
Figure 10-22. Temperature dependence of resolved yield stress i y for polyethylene under uniaxial tension (o) and plane strain compression (•). Lines are calculated from Eqs. (10-15 a) and (10-15 b) for lamellae of indicated thicknesses (nm). From Crist et al. (1989).
phous fraction as Tg is approached (Brown, 1986); the purely crystalline model is not applicable at such low temperatures. Some additional features can be obtained from the dislocation model. The
10.5 Fundamental Nature of Polymer Yielding
athermal strength at T = 0 K for a crystal yielding by (hkO) [001] slip via screw dislocations is Gb Tn =
= 0.015 G
(10-16)
where the core radius r0 has been replaced by 4 b. Note that this theoretical strength for a crystal (which is independent of thickness /c) is about one tenth the value for glasses which yield by either kink pairs [Eq. (10-11)] or dislocation loops [Eq. (10-12)]. The screw dislocation [Eqs. (1015 a) and (10-15 b)] can be compared in detail to the dislocation loop [Eqs. (10-13 a) and (10-13 b)] for glasses. A factor of 2 in T0 arises from different strain energies, and the balance comes from evaluation of the dislocation core radius r0. Bowden and Raha (1974) somewhat arbitrarily set r 0 = v / 3 b / e = 0.64 b for glasses, whereas computer simulation of the [001] screw dislocation core in polyethylene (Bacon and Tharmalingam, 1983) gives a more reliable result r o = l n m = 4fc. Using G = 3.1 GPa for the low temperature shear modulus of the polyethylene crystal, T 0 = 47 MPa, which is appreciably lower than experimental strength at low temperature where the vitrified amorphous regions have reinforced the composite (see Fig. 10-22). This confirms the idea of Brown (1986) that yield below Tg is governed by the stronger glassy component. The strain rate dependence of ry for yield by nucleation of [001] screw dislocations can also be calculated from Eqs. (10-15 a) and(10-15b): dlnT y _ 2nkT dlny ~~Gb*Tc
(10-17)
Evaluating this with parameters for polyethylene at room temperature, i y is predicted to increase by about 10% for a 103
455
increase in strain rate. This is much smaller than the - 100% increase in Fig. 10-13 or the more modest 50% increase reported by G'Sell and Jonas (1979) from a series of experiments done at constant true strain rates et which conform to the theoretical framework. This shortcoming of the model is related to underestimation of the temperature dependence of Ty mentioned above. It is probable that the amorphous regions, particularly the fold surfaces or "interphases", influence the energy of the dislocation nucleus. Boyd (1985) describes the a-relaxation process in polyethylene in terms of coupled displacements of fold surface segments and crystalline stems. A crystal-amorphous coupling of this sort would make the effective shear modulus G more temperature and strain-rate dependent, leading to better agreement between model predictions and experiment. The dislocation model was motivated primarily by the positive dependence of ay (or Ty) on crystal thickness /c. Balta Calleja (1985) has used microhardness measurements on polyethylene (hardness H « 3 ay) to consider the same issue. Those experiments led to cry~ v // c , which was interpreted as a decrease in resistance to chain slip resulting from less dense packing of chains in thinner crystals. In a different analysis of essentially the same data (Balta Calleja and Kilian, 1985), it was concluded that the increase in oy is caused by a reduction in the number density of (hkO) shear planes in lamellar aggregates with a large lc. Neither of these schemes accounts for the absolute value of
456
10 Plastic Deformation of Polymers
dence of xy. Flory and Yoon (1978), among others, contend that the irregular fold surface in melt crystallized polymers precludes substantial shear displacements of the sort sketched in Fig. 10-19, which is based on adjacent reentry or a regular fold surface. Nevertheless, X-ray diffraction shows clearly that [001] slip is the mechanism by which yielding and plastic deformation occur in plane strain compression of melt crystallized polyethylene (Young et al., 1973). Detailed pole figure studies by Krause and Hosford (1989) reach the same conclusion for uniaxial extension, uniaxial compression, and plane strain compression. These findings have been confirmed for plane strain compression at different temperatures by Bartczak et al. (1992 a, b). Elegant electron microscopy experiments on specially textured melt crystallized polyethylene deformed in tension have been done by Adams et al. (1986). Here [001] slip is directly seen to occur on both (100) and (010) shear planes. These experiments all demonstrate that melt crystallized polyethylene deforms by [001] or caxis slip, regardless of constraints imposed by chain reentry at the lamellar surfaces. Most of the experimental evidence, together with the reasonable success of dislocation nucleation as a quantitative model, lead to the conclusion that intracrystalline shear by [001] slip is the dominant yieldmechanism. This statement applies to linear polyethylene deformed at or near room temperature, though it is thought to be more general. The change of the semicrystalline long period L is inconsistent with this mechanism. Increases in the long period at high draw temperatures can, however, be ascribed to normal lamellar thickening by annealing at Td, perhaps accelerated by the applied stress. Indeed, the increase of L when drawing polypropylene at 20°C < Td < 150°C can be duplicated by
annealing at the respective draw temperatures (Peterlin, 1971). The more problematic decrease of L on drawing has been addressed in the electron microscopy studies of polyethylene deformation by Brady and Thomas (1989), where well developed lamellae are seen to be transformed into fibrils in a manner generally consistent with Fig. 10-19. After the lamellae have sheared by [001] slip, block-like crystals are observed to "decrystallize" (become less ordered as judged by electron diffraction) by the generation of a large number of unspecified internal defects. These structurally irregular regions become thinner in the chain axis direction. This process can be thought of as "melting" of crystals containing sufficient defects to lower Tm to the deformation temperature Td (room temperature in this case). But destruction of the original crystals is done by mechanical work, not by the increase of thermal energy from viscoelastic effects. The mechanically mobilized chains respond to applied stress and are reordered into crystals with a thickness characteristic of Td. Galenski et al. (1992) have proposed a related explanation. They envision (amorphous) defects to rearrange, permitting a "virtual shape change" of highly sheared crystal fragments to establish a new structural periodicity L. For the best studied case of polyethylene, it appears that yield (the onset of plastic deformation) at or near room temperature is controlled by crystal plasticity via dislocation nucleation. Post-yield flow, involving neck formation and stabilization in tension, involves "quasi-melting" or "decrystallization" at subsequent stages in the deformation process.
10.5 Fundamental Nature of Polymer Yielding
457
10.5.3 Crystalline and Liquid Crystalline Polymers
Highly oriented polymers display a unique yielding mechanism when deformed in compression along the chain axis direction. The strain is localized in bands a few nanometers thick which are inclined by an angle ^ « 5 0 ° - 7 0 ° to the compression direction. These may be isolated or occur in bundles as large as 100 jrni in the loading direction. This yield process is important because it limits the compressive strength of polymer preparations which are, by virtue of good chain axis orientation, very strong in tension. DeTeresa et al. (1988) have considered a number of highly oriented fibers of liquid crystalline polymers. These have (brittle) tensile strengths of about 2.5-3.2 GPa but fail by yielding at compressive stresses only — 0.1 as large. The macroscopic strain for this compressive yielding is quite small, typically less than 0.5% (|ey| < 5 x 10~3). Continued deformation increases the number and size of the bands. Axially oriented cracks form at the boundaries between sheared and unsheared regions, ultimately leading to fracture. Examples are shown for a liquid crystalline polymer fiber (Fig. 10-23) and a single crystal of substituted polydiacetylene (Fig. 10-24) deformed in compression. The applied stress first causes the chain axes to shorten by elastic strain. Conventional plastic deformation by [001] slip cannot be activated because the resolved shear stress is zero on (h k 0) slip planes. At some structural defect, frequently on the surface, local stress concentrations induce cooperative shear displacements of chain sections which evolve into the shortened "steps" in Figs. 10-23 and 10-24. The driving force for this local shear is the reduction in overall strain energy in the
Figure 10-23. Dark field electron micrograph of a poly(para-phenylene benzobisoxazole) fiber fragment. A single kink of length Lk « 500 nm and height w « 40 nm is toward the top of the figure. A large group of kinks is in the center of the figure. From Martin and Thomas (1991); reproduced with permission of Chapman and Hall.
compressed body which now has a smaller length in the loading direction. For polymer single crystals the boundary between sheared and unsheared regions is often a low index crystallographic plane [e.g., (012) in Fig. 10-24], and the strain is described as chain axis rotation deformation twinning (Bevis, 1978; Young et al., 1979). In liquid crystals and in oriented polycrystalline materials such as extended chain polyethylene one sees noncrystallographic boundaries indicative of kink bands (Martin and Thomas, 1991). A common feature is that the chain axes are bent severely and cooperatively at two well defined bound-
458
10 Plastic Deformation of Polymers (b)
la) <
ao 0=10.9° SDjun
Figure 10-24. (a) Scanning electron micrograph of a deformation twin in a polydiacetylene single crystal. Bending of the chain molecules at the (012) twin boundary is illustrated schematically in (b). From Young et al. (1978); reproduced with permission of Chapman and Hall.
aries which may extend entirely across a 10 jam diameter crystal or fiber. Each boundary is a mirror plane relating the chain directions in sheared and unsheared regions. A sketch of the atomic arrangements is presented in Fig. 10-24 where the angle of the twin boundary <> / = 90° — x//. The chains are sheared by a strain y = 2 tan 0 = 2 cot \jj in creating the band. Note that this shear strain is neither parallel nor perpendicular to the [001] direction of the chain axes. For deformation twinning these boundaries are low index crystallographic (hkl) planes which are true mirror planes, while there is no such geometric restriction for kinking. Hence the angle x// or 0 must have fixed values for deformation twins, while it may vary continuously for kinks. Twin or kink bands have some characteristics of crazes formed under tensile stress fields; plastic strain is localized in planar regions and fracture originates from these regions in both cases. Crazing, however, transforms originally isotropic material into oriented microfibrils with appreciable void formation (Chap. 15, this
Volume). This is quite distinct from the bands here, in which only the local orientation direction of the polymer chains is changed. Some theoretical work has been done on these deformation modes. Possible (hkl) twin planes (deformation twin boundaries) for bands of the type illustrated in Fig. 10-24 are established from crystallographic symmetry (Bevis, 1978). This analysis ignores energy penalties at the twin boundary, aside from the requirement that chains are not broken or excessively distorted. It is usually assumed that those twin deformation modes having the smallest shear displacements (smallest rotations of chain segments) are favored, though exceptions have been noted (Young et al., 1979). The precondition of no excess energy at twin boundaries is an oversimplification when chains at the boundary are "bent" by rotation about single bonds or by deformation of intramolecular bonds (see Fig. 10-24). A more realistic description would include energy required to deform the chains. This has been done by Pertsev et al. (1981), though their rather cumbersome theoreti-
10.5 Fundamental Nature of Polymer Yielding
cal results have not been applied to experiments. The concept of a twin plane describing the relation between sheared and unsheared sections of chains is less appropriate in liquid crystalline polymers within which the chains lack strict crystallographic registration. Martin and Thomas (1991) have applied the treatment of kink bands developed by Frank and Stroh (1952). The model uses arrays of dislocations of opposite sign to define the two boundaries between kinked and unkinked regions. The strain fields of these dislocation arrays provide a formal method for calculating the energy of the boundary between sheared and unsheared material. The energy At/ of a kink band of length L k and height w is evaluated as a function of the shear strain y in the kink band, the applied shear stress T, elastic constants G and v, Burger's vector b and dislocation core radius r0. AU goes through a maximum with respect to L k , permitting definition of the critical kink nucleus of size L\ which can grow by generation of additional dislocation pairs, thus lowering the energy of the system by increasing the length L k of the kink band. With some simplifications the result is
2L*J
(10-18)
This nucleation model differs from those in Sees. 10.5.1 and 10.5.2 in that the critical size L\ is not evaluated explicitly in terms of At/*, meaning that absolute shear strength xy (and its temperature and rate dependence) is not obtained. Furthermore, the relation between shear stress T and compressive stress a is determined by the unspecified angle \j/ (related to local strain by y = 2 cot \j/) between the boundaries and the axial loading direction.
459
Martin and Thomas (1991) have used the kink nucleation model to interpret the kink bands seen in poly(para-phenylene benzobisoxazole) fibers. They find that the experimental height of a single kink (w « 30 nm), is in good accord with the predicted value [w = L k T/(Gy) « 24 nm]. Tensile and compressive hydrostatic stress fields caused by the dislocation arrays are suggested to account for the relative sharpness of the top boundary and the slight downward curvature of the kink band as it grows in from the surface towards the center of the fiber (see Fig. 10-23). Thus the Frank-Stroh model accounts for the approximate size as well as certain morphological features of kink bands seen in liquid crystalline polymer fibers. Absolute strength i y is not given by the Frank-Stroh model, though this could be obtained in a fairly straightforward manner by invoking thermal activation, as was done for the kink pair and dislocation models described in Sees. 10.5.1 and 10.5.2. Mechanical response of single poly (paraphenylene benzobisoxazole) and other rigid chains has been modeled with semiempirical molecular orbital calculations by Wierschke et al. (1992). That work shows the chain "kinks" under compression by distortion of covalent bond angles in one of the heterocyclic rings. Such molecular yielding bears a possible relation to kink deformation in fibers, though the calculated (athermal) stress for this process is about —10 GPa, considerably larger than experimental strengths, which are about - 0 . 3 GPa. DeTeresa et al. (1985) have presented a rather phenomenological model of fiber buckling by a mechanism similar to kink-band formation. Their model predicts that the compressive stress required to cause buckling is oy = G (the shear modulus of the fiber parallel to the chain axes), while experiments give ay « G/3 (DeTeresa
460
10 Plastic Deformation of Polymers
et al., 1988). This discrepancy is attributed to various defects, though no consideration is given to thermal activation of the local yielding process.
10.6 Post-Yield Deformation and Modeling Glassy or semicrystalline polymers generally exhibit considerable ductility or flow before fracture. Stress-strain curves can be represented by various mechanical models which serve to correlate data taken under different conditions and, in some cases, provide insight on the nature of the deformation process. 10.6.1 Viscoelastic Models Much work deals with tensile deformation at a constant strain rate e, for which the simplest linear viscoelastic model is (10-19) Parameters are the unrelaxed modulus Eo and the relaxation time t'. The stressstrain curve has an initial slope Eo. The slope decreases conspicuously at a strain s « s t\ after which o approaches the limiting (Newtonian) flow stress Eo 81'. No distinction is usually made between nominal stress and true stress because strains are modest in the region of interest. This model can be modified to mimic the behavior of plastics deforming in tension. Matsuoka (1986) has considered nonlinear viscoelastic expressions of the following type: a{8) = £ o e e x p ( - Ce) exp( - — 1 (10-20) The first exponential factor involving C 8 is invoked to describe yielding, i.e., the presence of a maximum in o versus 8. Recall,
however, that a maximum in the nominal tensile-stress always results from extrinsic yield, not from strain softening as implied in Eq. (10-20) (see Sec. 10.2). The yield stress has a strain-rate dependence dominated by the second exponential factor. The single relaxation time has been replaced by an empirical distribution of relaxation times having the average value t' and a width parameter n. The response function for this distribution [sometimes called the Kohlrausch-Williams-Watts or KWW function (Matsuoka, 1992)] is a "stretched exponential" R{t) = exp[-(t/t')n], which is seen to reduce to a simple exponential for n = 1. The width parameter n is obtained directly from the time dependence of the elastic modulus in a relaxation experiment; n approaches 1 as the width of the distribution vanishes. This model predicts that ay ~ £n, where n is of the order of 0.05 for glassy polymers (Matsuoka, 1986). Examples of fits to experiments on polycarbonate are given by the solid lines in Fig. 10-9. The model does not include strain hardening, so the constant flow stress implied in Fig. 10-9 is arbitrary. A similar treatment, with one additional parameter to account for the strain dependence of t\ will account for the positive shift of yield strain £y with strain rate. The two factors relating stiffness and relaxation time to strain can be associated with modified entropy and free volume in the deforming solid (Matsuoka, 1986). A related analysis of tensile deformation which accounts for yield with intrinsic strain softening has been developed by Knauss and Emri (1987). The interesting idea here is that a tensile strain, assumed for the moment to be constant, causes free volume and chain mobility to increase in a time dependent manner. This is because the stress-volume response function, i.e., the bulk compliance, changes with time from a
10.6 Post-Yield Deformation and Modeling
low glassy value to a higher liquid-like value with a time constant which is itself decreased by the evolving free volume. When this concept is incorporated in an otherwise linear viscoelastic model for tensile stretching, a stress maximum followed by strain softening appears due to the time delayed, autocatalytic increase in free volume and mobility. Strain softening is more pronounced at high strain-rates, in agreement with experimental observations. The same model handles very nicely nonlinear viscoelastic stress relaxation, and accounts for time-dependent effects of hydrostatic pressure, for example, physical aging. Shay and Caruthers (1990) have developed a related treatment in which yield occurs when the entropy of the deformed body is increased to that of the unstretched material at Tg. Entropy also governs the time evolution of the system relaxation time t\ largely through the volume increase in tensile stretching. A basic feature is that nonlinear effects are predicted in terms of measurable linear viscoelastic properties. With input data for poly (vinyl acetate) the model predicts extrinsic yield (no decrease in true stress) in tension. While there is no stress drop, strain hardening is observed after the yield point. This interesting behavior comes from a decrease in volume after yield, though the origin of this densification is not understood. Temperature effects can be treated by time-temperature superposition in viscoelastic models. Matsuoka (1986) has developed an analysis based on Arrhenius activation for the average relaxation time t' which reduces to a particularly simple relation. All stresses and strains obtained at a temperature T are related to those at a reference temperature TR by the coefficient (Tc - T)/{TC-TR\ where Tc is the temperature at which the yield stress of the glass extrapolates to zero. Application of this
461
scheme to polycarbonate is illustrated in Fig. 10-8 for temperatures above and below 7 R =23°C. The analysis requires that the basic shapes of the curves do not change with temperature, a feature which is observed sufficiently below Tg. Semicrystalline polymers below Tg can be treated as if they were glassy; see Fig. 10-11. Those above Tg have initial moduli which are temperature and strainrate dependent. A fairly simple empirical scheme for correlating data has been developed by Matsuoka (1986). The strain rate dependence for polyethylene is fitted in Fig. 10-13. These treatments of tensile stress-strain relations ignore the fact that nominal stress on will drop after the extrinsic yield point because a neck is formed. A simple linear viscoelastic model [Eq. (10-19)] will generally result in a maximum load or an if treated according to Considered construction; see curve A in Fig. 10-5. Nevertheless, virtually all glasses yield with intrinsic strain softening as accounted for correctly, if phenomenologically, by nonlinear viscoelasticity. To account for strain softening, these analyses rely explicitly or implicitly on a free volume increase during tensile deformation. That assumption is necessary, in this context, for yield and plastic strain. The free volume approach is conceptually difficult to implement for shear or compression where macroscopic volume is constant or decreasing. 10.6.2 Molecular Models Kinematics of deformation, appropriate for describing large plastic stretches, has been developed by Boyce et al. (1988). In that framework they analyze the response of an elastic-plastic constitutive model which is based on molecular parameters for polymers. The yield stress is described
462
10 Plastic Deformation of Polymers
close to T g =110°C, in Fig. 10-25. The model captures most elements of the observed behavior, though the calculated yield strains are about 0.2 of the observed values, for reasons which are not clarified. Especially impressive is the agreement in the strain hardening region. Matsuoka (1991,1992) has recently proposed a model for viscoelastic-plastic behavior which eliminates the need for free volume to activate flow in a glass. It is postulated that yield or plastic strain occurs by the cooperative rearrangement of domains composed of m "conformers"; a conformer is a repeat unit containing a carbon-carbon single bond in the polymer backbone. Conformational entropy of a domain having m units is utilized in a transition-rate theory to establish the relation between the average relaxation time t' and applied stress a. According to this model, linear viscoelasticity prevails until the
by Argon's kink pair model [Eq. (10-9)], which predicts the temperature and strain rate dependence of ry without recourse to free volume. Pressure effects are described by the von Mises criterion [Eq. (10-4)] and strain softening is introduced by an empirical modification of T 0 [Eq. (10-10)] after the yield point. Strain hardening is treated by back stresses generated as Langevin chains with finite extensibility are stretched. The idea that strain hardening in polymers results from increased entropic resistance had been proposed by Haward (1973). Back stress or strain hardening rate depends linearly on temperature T, but is assumed to be independent of strain rate. All model parameters except those for pressure effects and strain softening are based on molecular theories. Calculated curves for tensile deformation at constant (true) strain rates are compared to experiment for poly (methyl methacrylate) at T= 90 °C,
i
•
i
i
|
•
i
i
i
|.;,
; • / /
:[!
100.0
/
50.0 s"1
- Exp. i-\
Mod. f = 1 s-1 %
>
Jlt\ * -%? '* M ^^. " * - - - : - - * ^
- •• Exp. fr0.1 S"1
<** f^
- Mod. fr0.1 s"1 - Exp. f=0.01 s-1 - Mod. fr0.01 s-1
n
r
0.50
,
i
i
1.00
•
1
I
I
1
1.50
Figure 10-25. Tensile true stress-true strain curves for poly (methyl methacrylate), Td = 90 °C, at constant strain rates indicated. The molecularly based constitutive model accounts quite well for observed behavior. Exp.: experimental; Mod.: model. From Boyce et al. (1988); reproduced with permission of Elsevier Science Publishers B.V.
10.6 Post-Yield Deformation and Modeling
strain energy reaches a critical value at which point the mechanical behavior shifts to classical plasticity. This viscoplasticity approach is distinct from linear (or nonlinear) viscoelastic models in which elastic deformation evolves into viscous flow after a certain time t&t' [see Eq. (10-19)]. Matsuoka simplifies the energy criterion for yield in terms of a critical strain 8* = ay/E0. A spectrum of relaxation times is required for the model to have the observed dependence of <7y (or e*) on strain rate. The distribution of relaxation times is described by a width parameter n and an average t' as discussed for Eq. (10-20). Equating 8* with the yield strain ey, the result is
= —{e*y-n{ety
(10-21)
Here E (t) is the relaxation modulus, written for uniaxial deformation, either tensile or compressive. A comparable expression can be obtained for shear deformation, as volume changes have not been invoked. This model again predicts that the yield stress is proportional to 8n, but without recourse to free volume effects. Plastic deformation of the glass results from stress enhanced rate of intramolecular rearrangement in a manner related in some ways to the approach of Robertson (1966). As the viscoelastic-plastic model represented in Eq. (10-21) extends only to the yield point, the issue of strain softening is not addressed. Valuable insight into the plastic deformation of isotropic crystalline polymers has been provided by the analysis of Parks and Ahzi (1990). They calculate stressplastic strain relations for an initially isotropic array of polyethylene crystals which deform by shear. Earlier crystal plasticity theories rely on the presence of at least 5 independent (hkl) [uvw] slip systems to permit plastic flow of arbitrarily oriented
463
crystals without stress singularities. Polymer crystals are deficient in the number of slip systems because of constraints imposed by unbreakable covalent bonds. The authors develop the kinematics of deformation of these deficient systems with a "constrained hybrid" model for an incompressible rigid-viscoplastic material; no volume change occurs, small elastic strains are ignored, and the critical resolved shear stress for slip has a positive strain rate dependence. The model provides numerical solutions for stress, strain and texture [orientation of (hkl) planes] for an initially isotropic set of 248 crystals. This analysis was applied to polyethylene having 4 slip systems ([001], [100], [010], and [110] slip directions on appropriate (hkO) planes). The model is constructed for 100% crystalline polyethylene; no account is taken of the amorphous component. Calculated results for tensile and shear deformations are presented in Fig. 10-26. The equivalent macroscopic stress <req = ert = T y/3 is normalized by the critical resolved shear stress for [001] slip and equivalent (true) strain is 8eq = st = y/y/3. One sees immediately that plastic deformation commences at a common equivalent stress after which strain hardening develops much more rapidly in tension than in shear. This calculated response is very similar to the constant true strain rate results of G'Sell et al. (1991) for polyethylene (crystallinity a « 0.7) in Fig. 10-27. The difference in strain hardening rates can be attributed to development of a c axis or [001] texture and rotation of the (hkO) slip planes into the stretch direction (Parks and Ahzi, 1990). For tensile deformation this leads to "textural hardening" as the resolved shear stress on active slip planes becomes smaller, requiring ever increasing tensile stress at for additional strain. In shear, rotation of [001] and slip planes to-
464
10 Plastic Deformation of Polymers
0.5
o
1.0
1.5
Figure 10-26. Calculated stress-strain curves for isotropic crystalline polyethylene. Strain hardening is pronounced in tension, negligible in shear (cf. Fig. 10-27). From Parks and Ahzi (1990); reproduced with permission of Pergamon Press pic.
200 HDPE
150
and polyoxymethylene are qualitatively similar to polyethylene, yielding abruptly at a shear strain of y y « 0.3 followed by essentially flat flow curves with very modest shear strain hardening. Each of these polymers strain hardens in tension with stable neck propagation. The other two semicrystalline polymers, nylon 66 and polybutylene-1, have rather diffuse yield points followed by strain hardening in shear. The reason for this different behavior is not understood. In summary, glassy polymers display strain hardening in tension and shear, which is accounted for by finite extensibility of chain segments in the entangled network of interpenetrating chains (see Fig. 10-25). No fundamental reason for ubiquitous strain softening in polymer glasses has been proposed. Semicrystalline polymers behave differently, yielding with no strain softening in either tension (Fig. 10-15) or shear (Fig. 10-16). For these materials the presence of strain hardening in tension and the absence of strain hardening in shear is attributed to texture evolution.
100
10.6.3 Continuum Mechanics of Necking 50
Simple shear
0
0.5
1.0 Equivalent strain
1.5
2.0
Figure 10-27. True stress-true strain curves for polyethylene deformed at constant strain rate in tension and shear. From G'Sell et al. (1991); reproduced with permission of the author.
ward the stretch direction still allows most of the deformation to occur by easy chain slip without pronounced textural hardening. This type of behavior seems fairly general, as seen in Fig. 10-16. Polypropylene
Knowledge of intrinsic mechanical response of the material, that is, the true stress - true strain behavior, is essential for understanding the response of a polymer to an arbitrary stress field. Considerable attention has been paid to the plastic deformation of polymers in uniaxial tension when, as described in Sec. 10.2, plastic instability often leads to neck formation. Interest derives from two factors. From an engineering viewpoint, "cold drawing" by neck propagation is used to achieve high orientation and increased stiffness and strength (Ward, 1975). Description of the necking process is, furthermore, a challenge to those interested in the continuum
10.7 Summary
and molecular approaches to polymer deformation. Necking phenomena are based on postyield deformation, as introduced in Sees. 10.2 and 10.4. Intrinsic mechanical response (at versus et) is needed as a function of strain rate and pressure because of the strain rate gradient and hydrostatic tension associated with a nonuniform cross section. These combine to determine the profile of the neck and the stress at which it propagates. Coates and Ward (1980) have described the relations between true stress, true strain, strain rate, and neck profiles in polyethylene. More pronounced strain hardening stabilizes the neck, leading to a smaller stretch ratio in the necked portion and a more gradual thinning of the sample in the neck region. Materials with small strain hardening rates undergo large stretches with sharp necks. Most viscoelastic polymers have a positive dependence of yield stress and post-yield flow stress on strain rate. This strain rate sensitivity also stabilizes the neck, leading to less deformation and more gradual neck profiles. Hutchinson and Neale (1983) provided the first kinematic description of steady state neck propagation in materials like semicrystalline polymers with strain hardening but without intrinsic strain softening. They first modeled the true stresstrue strain behavior by a nonlinear elastic solid (this is permissible if the true stress does not decrease in the process). A "jump" analysis based on initial and final states permits a Maxwell line solution for the nominal or engineering draw stress and the "natural" draw ratio X in the necked portion. Elastic-plastic constitutive models were analyzed with J2-flow theory for a more complete description of the propagating neck with multiaxial local stresses in a cylindrical rod. The effects of strain-hardening rate and strain-rate sensitivity of
465
flow resistance were considered. Increasing either of these material characteristics stabilizes the neck, i.e., decreases both the draw ratio and the sharpness of the neck. Necking phenomena can be examined in somewhat more detail with finite-element analyses of nonlinear-elastic models, as first done by Neale and Tugcu (1985). They found that the analytical approach of Hutchinson and Neale (1983) was generally adequate for neck propagation in axisymmetric fibers. A similar study of plane strain neck propagation was performed by Fagler and Bassani (1986) and later extended to include plastic anisotropy by Batterman and Bassani (1990). Boyce and Arruda (1990) have done a revealing finite element study of necking in polycarbonate using the molecularlybased constitutive model described in Sec. 10.6.2. The model reproduces the observation that the strain hardening rate is greater in tension than in uniaxial compression because the amount of chain stretching, which leads to back stresses, is smaller in uniaxial compression (or biaxial extension). The same model accounts for the shape of the neck which evolves in tension and the variable strain rate of the material in the neck.
10.7 Summary The yield strength of polymer glasses absolute value and temperature and strainrate dependence - can be accounted for by nucleation models of either kink pairs (Argon, 1973) or dislocation loops (Bowden and Raha, 1974) with reasonable molecular parameters. Within about 100° of Tg, the intramolecular energy model (Robertson, 1966) is a viable, even preferable, alternative. The shear strength of polymer
466
10 Plastic Deformation of Polymers
glasses is described quantitatively by two or three distinct theories, which is an unusual situation. If one accepts the kink-pair model, it follows that glassy polymers yield at, or very near, their theoretical strength. Most of the drop in ry (or oy) that occurs when the temperature is raised toward Tg comes from the decrease of shear modulus G. Fundamental understanding of shear strength and its temperature dependence hence reverts to the temperature dependence of G. Theodorou and Suter (1986) have developed a semiempirical method to calculate the structure and elastic constants of an isotropic polymer glass, though it has not been applied to temperature effects. This simulation technique has been extended to the study of plastic strain in a polymer glass (Mott et al., 1993); an account is presented in Chap. 10, Vol. 6 of this Series. Structural information from such simulations may permit discrimination between existing theories or provide insight to generate better ones. Semicrystalline polymers at temperatures below Tg of the amorphous component yield similarly to glasses. However, the yield stress ay decreases much more rapidly in the range just above Tg. This "transition region" over which the amorphous regions become liquid-like had not been studied in detail, but it is probable that most of the effects are associated with easier shear in the noncrystalline portions of the composite (Brown, 1986). At still higher temperatures, where amorphous strength is negligible, yield stress is found to correlate with thickness lc of lamellar crystals. This has been modeled successfully with nucleation of screw dislocations for the case of polyethylene (Young, 1988; Crist etal., 1989), taking advantage of known structures of the crystal and the postulated defect. Calculations of equilibrium structures, elastic constants, and their
temperature dependencies are well established for crystals (McCullough and Peterson, 1973; Sorensen etal., 1988; Chap. 7, this Volume), being much more straightforward than those for glasses. Static defects such as dislocations have been modeled by Bacon and Tharmalingam (1983). It is hoped that these methods will be extended to other polymer crystals. Complementing the mechanical studies are investigations of morphology, generally before and after considerable plastic deformation. Some of these have led to the suggestion that yielding and plastic deformation involve melting of the crystallites (Popli and Mandelkern, 1987), though even semiquantitative predictions have yet to be made. While a large number of models have been proposed to account for yield strength, relatively little consideration has been given to descriptions of plastic deformation beyond yield. The important issue of strain softening, observed in glasses under almost all conditions, has been treated phenomenologically without much attention to its cause. One exception is "delayed dilatancy", which can account for strain softening in tension with a nonlinear viscoelastic model (Knauss and Emri, 1987). An interesting development in this area has been reported by Hasan et al. (1993). Positron annihilation lifetime measurements indicate that the size of free-volume packets increases irreversibly after compressive yielding of poly (methyl methacrylate). This implies that free volume ideas may be more general than thought previously; additional experiments to probe local free-volume (as opposed to macroscopic volume changes) should be quite important. It would be desirable to have structural or dynamic measurements of polymers under stress to confirm these ideas. Semicrystalline polymers do not display strain softening in tension or shear,
10.9 References
perhaps because they are usually tested above Tg of the amorphous component. Virtually all polymer solids, glassy or semicrystalline, display conspicuous strainhardening in tension. Glasses strain harden in compression and shear, as well. This has been modeled successfully by entropic resistance to flow as chain segments of finite extensibility are stretched. Reasonable molecular parameters describe observed behavior, and the model accounts for the strain-hardening rate being greater in tension than in compression (Boyce and Arruda, 1990). Quite different explanations apply to semicrystalline polymers, at least between Tg and Tm. Texture development by [001] slip leads to pronounced strain hardening in tension, and very little strain hardening in shear (Parks and Ahzi, 1990). Computer simulations of large strain behavior have been extended by Lee et al. (1993) to include deformations of amorphous and crystalline regions. This approach accounts for many features of postyield flow of polyethylene, and should become a powerful method for unraveling the complex response of other semicrystalline polymers. The past decade has seen a number of significant advances in our understanding of yielding and plastic deformation in polymers. Experimental innovations and refinements have provided intrinsic responses which are essential input for theories and models. Sophisticated theories and powerful numerical methods have been applied with success. Future developments will certainly involve advanced microscopies, spectroscopies and simulations to describe how structure and dynamics on the molecular scale respond to stresses beyond the elastic limit.
467
10.8 Acknowledgements The author is indebted to B. Moran and many other colleagues who assisted in the preparation of this article through discussions and helpful comments. Partial funding was provided by the Gas Research Institute, Physical Sciences Department, Contract No. 5090-260-2066.
10.9 References Adams, W. W, Yang, D., Thomas, E. L. (1986), /. Mater. Sci. 21, 2239. Argon, A. S. (1973), Phil. Mag. 28, 839. Argon, A. S. (1975), in: Polymeric Materials: Baer, E., Radcliffe, V. S. (Eds.). Metals Park, OH: Am. Soc. Met., pp. 411-486. Argon, A. S., Bessonov, M. I. (1977), Phil. Mag. 35, 917. Argon, A. S., Cohen, R. E. (1990), Adv. Polym. Sci. 90/91, 301. Bacon, D. I , Tharmalingam, K. (1983), / Mater. Sci. 18, 884. Balta Calleja, F. J. (1985), Adv. Polym. Sci. 66, 111. Balta Calleja, F. J., Kilian, H. G. (1985), Coll. Polym. Sci. 263, 697. Bartczak, Z., Cohen, R. E., Argon, A. S. (1992a), Macromolecules 25, 4692. Bartczak, Z., Argon, A. S., Cohen, R. E. (1992b), Macromolecules 25, 5036. Batterman, S. D., Bassani, X L. (1990), J. Polym. Sci. 30, 1281. Bessell, T. J., Young, R. J. (1974), J. Polym. Sci., Polym. Lett. Ed. 12, 629. Bevis, M. (1978), Colloid Poly. Sci. 256, 234. Bonner, R. M., Kohan, M. I., Lacey, E. M., Richardson, P. N., Roder, T. M., Sherwood, L. T. (1973), in: Nylon Plastics: Kohan, M. I. (Ed.). New York: Wiley-Interscience, Chap. 10. Bowden, P. B. (1973), in: The Physics of Glassy Polymers: Haward, R. N. (Ed.). New York: Wiley, Chap. 5. Bowden, P. B., Raha, S. (1970), Phil. Mag. 22, 463. Bowden, P. B., Raha, S. (1974), Phil. Mag. 29, 149. Bowden, P. B., Jukes, J. A. (1972), /. Mater. Sci. 7, 52. Bowden, P. B., Young, R. J. (1974), J. Mater. Sci. 9, 2034. Boyce, M. C , Parks, D. M., Argon, A. (1988), Mech. Mater. 7, 15. Boyce, M. C , Arruda, E. M. (1990), Polym. Sci. Eng. 30, 1288. Boyd, R. H. (1983), J. Polym. Sci., Polym. Phys. Ed. 21, 493.
468
10 Plastic Deformation of Polymers
Boyd, R. H. (1985), Polymer 26, 323, 1123. Brady, J. M., Thomas, E. L. (1989), /. Mater. Sci. 24, 3311. Brown, N. (1983), Polymer 18, 2241. Brown, N. (1986), in: Failure of Plastics: Brostow, W., Corneliussen, R. D. (Eds.). New York: Hanser, Chap. 6. Chow, T. S. (1990), Macromolecules 23, 4648. Chuah, H. H., Lin, J. S., Porter, R. S. (1986), Macromolecules 19, 2732. Coates, P. D., Ward, I. M. (1980), J. Mater. Sci. 15, 2897. Crist, B., Fisher, C. X, Howard, P. R. (1989), Macromolecules 22, 1709. DeTeresa, S. X, Porter, R. S., Farris, R. J. (1985), /. Mater. Sci. 20, 1645. DeTeresa, S. X, Porter, R. S., Farris, R. X (1988), J. Mater. Sci. 23, 1886. Fagler, L. O., Bassani, X L. (1986), Int. J. Solids Struct. 22, 1243. Flory, P. X, Yoon, D. Y. (1978), Nature 272, 226. Frank, F. C , Stroh, A. N. (1952), Proc. Phys. Soc. B65, 811. Galenski, A., Argon, A. S., Cohen, R. E. (1991), Macromolecules 24, 3953. Galenski, A., Bartczak, A., Argon, A. S., Cohen, R. E. (1992), Macromolecules 25, 5705. G'Sell, C. (1988), Rev. Phys. Appl. 23, 1085. G'Sell, C , Jonas, X X (1979), /. Mater. Sci. 14, 583. G'Sell, C , Boni, S., Shrivastava, S. (1983), J. Mater. Sci. 18, 903. G'Sell, C , Hiver, X M., Dahoun, A., Phillipe, M. X, Esling, C. (1991), in: Abstracts of Plasticity 91 Symposium, Grenoble.. Hasan, O. A., Boyce, M. C , Li, X. S., Berko, S. (1992), /. Polym. Sci., Polym. Phys. Ed. 31, 189. Haward, R. N. (1973), The Physics of Glassy Polymers. New York: Wiley, Chap. 6. Heisse, B., Kilian, H. G., Peitrella, M. (1977), Progr. Colloid Polym. Sci. 62, 16. Hope, P. S., Duckett, R. A., Ward, I. M. (1980), /. Appl. Polym. Sci. 25, 1373. Hutchinson, X W, Neale, K. W. (1983), J. Mech. Phys. Solids 25, 1373. Imai, Y, Brown, N. (1976), J. Polym. Sci., Polym. Phys. Ed. 14, 723. Juska, T, Harrison, I. R. (1982), Polym. Eng. Sci. 22, 166. Kinloch, A. X, Young, R. X (1983), Fracture Behavior of Polymers. New York: Applied Science, Chaps. 4,5. Kamei, E., Brown, N. (1984), /. Polym. Sci., Polym. Phys. Ed. 22, 543. Kawai, H., Hashimoto, T. (1979), in: Contemporary Topics in Polymer Science: Shen, M. (Ed.). New York: Plenum Press, pp. 245-266. Knauss, W G., Emri, I. (1987), Polym. Eng. Sci. 27, 86. Kramer, E. X, Berger, L. L. (1990), Adv. Polym. Sci. 91/92, 1.
Krause, S. X, Hosford, W. G. (1989), J. Polym. Sci. B, Polym. Phys. 27, 1853, 1867. Lee, B. X, Argon, A. S., Parks, D. M., Ahzi, S., Bartczak, Z. (1993), Polymer, in press. Lin, L., Argon, A. S. (1992), Macromolecules 25, 4011. Lohse, G., Schmid, R., Batzer, H., Fisch, W. (1969), Br. Polym. J. 1, 110. Manson, X A., Sperling, L. H. (1974), Polymer Blends and Composites. New York: Plenum. Martin, D. C , Thomas, E. L. (1991), J. Mater. Sci. 26, 5171. Matsuoka, S. (1978), Polym. Sci. Eng. 21, 907. Matsuoka, S. (1986), in: Failure of Plastics: Brostow, W, Corneliussen, R. (Eds.). Munich: Carl Hanser, Chap. 3. Matsuoka, S. (1991), in: Proc. MRSSymp., Vol. 215: Roe, R. X, O'Reilly, X M. (Eds.). Pittsburgh, PA: Mater. Res. Soc, pp. 71-80 Matsuoka, S. (1992), Relaxation Phenomena in Polymers. New York: Carl Hanser, pp. 111-120. McCullough, R. L., Peterson, X M. (1973), / Appl. Phys. 44, 1224. Meinel, G., Peterlin, A. (1971), Eur. Polym. J. 7, 657. Mott, P., Argon, A. S., Suter, U. W. (1993), Phil. Mag., in press. Neale, K. W, Tugcu, P. (1985), /. Mech. Phys. Solids 33, 323. Parks, D. M., Ahzi, S. (1990), J. Mech. Phys. Solids 38, 701. Peterlin, A. (1965), /. Polym. Sci. C9, 61. Peterlin, A. (1971), /. Mater. Sci. 6, 490. Peterlin, A. (1987), in: Encyclopedia of Polymer Science and Engineering, Vol. 10: Mark, H. F., Bikales, N. M., Overberger, C. G., Menges, G., Kroschwitz X I. (Eds.). New York: Wiley, pp. 72-94. Pertsev, N. A., Romanov, A. E., Vladimirov, V. I. (1981), /. Mater. Sci. 16, 2084. Phillips, P. X, Philpot, R. X (1986), Polymer Commun. 307. Popli, R., Mandelkern, L. (1987), /. Polym. Sci. B, Polym. Phys. 25, 441. Pukansky, B., Tudos, F. (1990), Makromol. Chem., Macromol. Symp. 38, 221. Raghava, R., Caddell, R. M., Yeh, G. S. Y. (1973), J. Mater. Sci. 8, 225. Robertson, R. E. (1966), J. Chem. Phys. 44, 3950. Rutherford, X L., Brown, N. (1980), in: Methods of Experimental Physics: Polymers: Physical Properties, Vol. 16C, Fava, R. A. (Ed.). New York: Academic, pp. 117-135. Sadler, D. M., Barham, P. X (1990 a), Polymer 31, 36. Sadler, D. M., Barham, P. X (1990 b), Polymer 31, 46. Samuels, R. X (1974), Structured Polymer Properties. New York: Wiley. Saraf, R. F., Porter, R. S. (1988), J. Polym. Sci. B, Polym. Phys. 26, 1049. Schultz, X (1974), Polymer Materials Science. Englewood Cliffs, NJ: Prentice-Hall; Chaps. 3, 11. Shadrake, L. G., Guiu, F. (1976), Phil. Mag. 34, 565.
10.9 References
Shay, R. M., Caruthers, J. M. (1990), Polym. Eng. Sci. 30, 1266. Sorensen, R. A., Liau, W. B., Boyd, R. H. (1988), Macromolecules 21, 194. Sternstein, S. S. (1975), in: Polymeric Materials: Baer, E., Radcliffe, S. V. (Eds.). Metals Park, OH: Am. Soc. Met., pp. 369-410. Theodorou, D. N., Suter, U. W. (1986), Macromolecules 19, 139. Vincent, P. I. (1960), Polymer 1, 7. Ward, I. M. (1975), Structure and Properties of Oriented Polymers. New York: Wiley. Ward, I. M. (1983), Mechanical Properties of Solid Polymers, 2nd ed. New York: Wiley, Chap. 11. Wierschke, S. G., Shoemaker, J. R., Haaland, P. D., Pachter, R., Adams, W. W. (1992), Polymer 33, 3357. Williams, M. L., Landel, R. R, Ferry, J. D. (1955), /. Am. Chem. Soc. 77, 3701. Wu, W, Wignall, G. D. (1985), Polymer 26, 661. Wu, W, Wignall, G. D., Mandelkern, L. (1992), Polymer 33,4131. Yamini, S., Young, R. J. (1980), J. Mater. Sci. 15, 1814.
469
Young, R. J. (1979), in: Developments in Polymer Fracture-1, Andrews, E. H. (Ed.). London: Applied Science, pp. 223-261. Young, R. J. (1988), Mater. Forum 11, 210. Young, R. I, Bowden, P. W, Ritchie, J. M., Rider, J. G. (1973), J. Mater. Sci. 8, 23. Young, R. I, Bloor, D., Batchelder, D. N., Hubble, C. L. (1978), /. Mater. Sci. 13, 62. Young, R. I, Dulniak, R., Batchelder, D. N., Bloor, D. (1979), J. Polym. Sci., Polym. Phys. Ed. 17, 1325.
General Reading Haward, R. N. (Ed.) (1973), The Physics of Glassy Polymers. New York: Wiley. Peterlin, A. (Ed.) (1971), Plastic Deformation of Polymers. New York: Marcel Dekker. Schultz, J. (1974), Polymer Materials Science. Englewood Cliffs, NJ: Prentice-Hall. Ward, I. M. (1983), Mechanical Properties of Solid Polymers. New York: Wiley.
11 Dielectric Properties of Polymers Graham Williams Department of Chemistry, University College of Swansea, Swansea, U.K.
List of Symbols and Abbreviations 472 11.1 Introduction 476 11.2 Phenomenological Theory 478 11.2.1 General Frequency-Time Relations 478 11.2.2 Empirical Relations for Permittivity 479 11.2.3 Temperature and Pressure Dependence of the Average Relaxation Time . 481 11.3 Molecular Theory 482 11.3.1 Equilibrium Theory 482 11.3.2 Dynamic Theory 485 11.4 Experimental Methods 487 11.5 Experimental Data for Polymer Systems 488 11.5.1 Introduction 488 11.5.2 Experimental Data for Liquids and Solutions 489 11.5.2.1 Flexible Chains 489 11.5.2.2 Rod-Like Chains 492 11.5.3 Experimental Data for Amorphous Solid Polymers 494 11.5.4 Experimental Results for Crystalline Polymers 501 11.5.4.1 Introduction 501 11.5.4.2 Polymers of High Degree of Crystallinity 501 11.5.4.3 Polymers of Medium Degree of Crystallinity 503 11.5.5 Experimental Data for Liquid Crystalline Polymers 507 11.5.5.1 Introduction 507 11.5.5.2 Non-Chiral Liquid Crystalline Side-Chain Polymers 509 11.5.5.3 Chiral Liquid Crystalline Polymers 517 11.6 Concluding Remarks 521 11.7 Acknowledgements 523 11.8 References 523
Materials Science and Technology Copyright © WILEY-VCH Verlag GmbH & Co KGaA. All rights reserved.
472
11 Dielectric Properties of Polymers
List of Symbols and Abbreviations A a, b a, b A, B, To, T^ aktk±j
material constant (Arrhenius equation) Landau coefficients Havriliak and Negami relaxation parameters parameters of Vogel equation coefficients which add the contributions to
List of Symbols and Abbreviations
473
h H, H o k £ M M(0),M(t)
relaxation function Hamiltonian, unperturbed Hamiltonian Boltzmann constant Liouville operator molecular weight spontaneous polarization of macroscopic volume V at times 0 and t, respectively Mc critical molecular weight for entanglements Mz (t) average dipole moment along laboratory z-axis n director axis in L C phase n refractive index Nr number of dipole groups per unit volume JV V concentration of dipolar molecules p (co) internal field factor [ = x (co)/x (0)] P pressure V\,CL\ conjugate position and momentum of particle i Pz (t) macroscopic polarization along laboratory z-direction °p r probability of obtaining environment V Q apparent activation energy for relaxation (Arrhenius equation) <2P,<2v constant pressure and constant volume apparent activation energies, respectively gpr strength factor for r'th component of P relaxation process R G a s constant r• (0), r• (t) end to end vector for chain i at times 0 and t, respectively S molecular order parameter for a uniaxial liquid crystal Sd macroscopic director order parameter S ^ S J j S ^ S f ^ S x chiral-smectic A, C, F, J, and X phases, respectively t time T temperature Tc clearing temperature Tg glass transition temperature Tt transition temperature v,v[9vb volume, free volume, bound volume, respectively V volume W(t) time-dependent transition probability X degree of crystallinity (0 < X < 1) /? p (0), /? (t) y1 yLC £(/) S(CD) sn, sY
K W W relaxation parameter polar angle between molecular dipole moment fi{ and laboratory z-axis in a LC material effective rotational viscosity of a liquid crystal rotational viscosity of a ferroelectric liquid crystal complex dielectric permittivity relative to free space dielectric permittivity tensor defined by Eqs. (11-48 a, b)
474
11 Dielectric Properties of Polymers
sl],8h 2V e el e 1 ,e 2 SQ^S^ s\ e" e^ efT9 e" £|l (co), £±(co) e
O||>
fi
°o||
ds As, i ( A£ p
n
9C 9kk> 90 x fi, fi0
fi0 lik,\ik> ju1?iut [ir, fik /i s ,ju b /A-J (0), \xA (t) <JUJ|>,
Q o T r b ,T s
upper and lower bounds of measured permittivity permittivity of free space complex permittivity due to electrode polarization permittivities of the amorphous a n d crystalline phases, respectively limiting low and high frequency permittivities, respectively real a n d imaginary parts of dielectric permittivity, respectively maximum loss factor real and imaginary parts, respectively, of relaxation part of total permittivity s principal permittivities parallel, perpendicular to the uniaxial liquid crystal director axis n principal low frequency permittivities of a uniaxial liquid crystal principal high frequency permittivities of a uniaxial liquid crystal dielectric anisotropy (= fiy — £±) dielectric relaxation strength, for individual processes relaxation strengths of the a and (3 relaxation processes, respectively internal rotation parameter (=
List of Symbols and Abbreviations
(t) (t)
CO
time-correlation functions for the oc and Pr relaxation processes total dipole moment correlation function for macroscopic volume V time correlation function of order J and indices "Pm" for dipole motion in a liquid crystal anisotropy of magnetic susceptibility
coc
angular irequency (= znj) orientation of body in Euler space cut-off angular frequency
DGEBA DOBAMBC DRS ITO KWW LC LCD LS MW NM NMR PEMA PET PHIC POM POE PVC SRT VDCN/VAc VHF, UHF WLF WVWP
particular epoxy resin (Sec. 11.6) particular liquid crystal mixture (Sec. 11.5.5.3) dielectric relaxation spectroscopy indium/tin oxide Kohlrausch-Williams-Watts liquid crystal liquid crystal display (high frequency) local segmental mode M ax well - Wagner (low frequency) normal mode nuclear magnetic resonance poly(ethyl methacrylate) poly(ethylene terephthalate) poly(hexyl isocyanate) polyoxymethylene polyoxyethylene poly(vinyl chloride) single relaxation time vinylidene cyanide/vinyl acetate very, ultra high frequency Williams, Landel and Ferry Warchol, Vaughan, Wang and Pecora
Q,Q0
475
476
11 Dielectric Properties of Polymers
11.1 Introduction Synthetic polymer materials as amorphous or crystalline thermoplastics, as elastomers, thermosetting resins or composites find diverse applications in modern society. The commercial products take the form of moulded or extruded articles, spun fibres, foams, films, tapes, laminates or adhesive layers. The practical uses of such materials are determined mainly by their mechanical and/or optical properties but in many important applications, e.g., for electrical cable insulation, film capacitors, components in electrical or electronic circuitry, the electrical properties are of great importance and interest. For such applications the polymer material must be a good electrical insulator at the frequencies and voltages of operation, i.e., the in-phase electric current should be as small as possible, thus avoiding dielectric heating or electrical breakdown or minimizing the attenuation of signals in submarine telephone cables. During World War II, Professor Arthur von Hippel and his group at M.I.T. and Professor Willis Jackson and his group at London University developed experimental methods for measuring the dielectric properties of different materials in the range extending from power frequencies (10-10 3 Hz), to audio and radio frequencies (10 2 -10 5 Hz), UHF and VHF frequencies (10 5 -10 8 Hz) and microwave frequencies (K^-IO 1 1 Hz). Data were obtained for the complex dielectric permittivity e ( / ) measured at different frequencies, /, and different temperatures, T, where the dielectric properties are expressed in terms of the complex relative dielectric permittivity £(/), where
(11-1) and e'(/) is the real permittivity and c" (/)
is the dielectric loss factor, both measured at a frequency of / = co/2 n. The additional quantity D = tan£ = s" (f)/sr (f) is also commonly used as a measure of the energy dissipation in a dielectric material. Extensive data for a number of materials, including synthetic polymers were published in von Hippel's 1954 monograph. Knowledge of the dielectric properties at different frequencies and different applied voltages allowed the applications of the materials to be assessed without any need for molecular interpretations of the origins of the particular values of the permittivity and loss factor and their dependencies on temperature and frequency. However, the molecular origins of dielectric relaxation spectroscopy had been established much earlier for dipolar molecular liquids and rotator-phase solids by Debye (1927) who had shown that the applied electric field perturbs the orientational distribution function for the dipolar molecules, leading to a static relative permittivity s0 greater than n2, where n is the optical refractive index, and a dispersion for £'(/) accompanied by a peak in e"(/). Dielectric relaxation due to classical molecular reorientational motions is a form of pure absorption spectroscopy whose frequency range of interest for materials, including synthetic polymers, is 10~6 to 10 11 Hz. Experimental data for different dipolar systems, mainly of low molar mass, were reviewed by Smyth (1955), Frohlich (1958) and later by Hill et al. (1969). It is fair to say that the static permittivities of simple dipolar liquids and rotator-phase solids are well-understood in terms of the molecular structure and organization of the phase while the dielectric relaxation behaviour of such systems is well-understood in terms of the classical reorientation motions by rotational diffusion or motion in barrier systems (for solids).
11.1 Introduction
In parallel with the studies of von Hippel and Willis Jackson, the American scientists, R. M. Fuoss and J. G. Kirkwood, carried out some of the first systematic studies of the dielectric behavior of polymer materials, e.g., poly(vinyl chloride) in its unplasticized and plasticized forms. They showed that extensive motions of the dipolar chain segments occurred above the apparent glass transition temperature Tg and that these motions, whose average rate was strongly dependent on sample temperature, could be investigated by dielectric relaxation spectroscopy. They documented the dielectric properties of several polymers as a function of frequency and temperature and gave a simple empirical function for the loss factor which fitted, approximately, the broad experimental curves. They also proposed molecular theories for £0 and s(f) for linear polymers (Fuoss and Kirkwood, 1941; Kirkwood and Fuoss, 1941). e0 was related to the dipole moment of a chain segment and a g-factor, the Kirkwood correlation factor, which expresses the angular correlations between dipolar groups along a chain and which arise from chain connectivity and chain conformation. Their studies laid the foundation of our understanding of the dielectric properties of polymers. As new polymers were prepared during and after the war, their dielectric properties were investigated and the literature expanded at a remarkable rate. Of the many publications, that by Reddish (1950) for polyethylene terephthalate, which featured three-dimensional contour diagrams for permittivity and loss, and that by Scott and coworkers (1962) for polychlorotrifluoroethylene, which covered a range of sample crystallinity and particularly wide frequency and temperature ranges, were of special importance in establishing molecular structure-dielectric property relations
477
for solid polymers. Multiple dielectric relaxation processes were observed including one associated with the dynamic glasstransition process (a relaxation) and another relating to limited motions of groups in the glassy state (P relaxation). Further systematic studies were made by Ishida and his group for several amorphous and crystalline polymers (see Ishida, 1969, for a review). By the early 1960's many of the conventional polymers used today were available and their dielectric properties had been investigated. The dielectric studies complemented those made using the techniques of dynamic mechanical relaxation and nuclear magnetic resonance. Comparing the results of such studies it was found possible to assign the origins of the multiple relaxation regions and to suggest specific mechanisms for motion. In 1967 McCrum, Read and Williams published a research text in which the dynamic mechanical and dielectric relaxation properties of solid polymers were compiled, were described phenomenologically, and were discussed, in a comparative way. The origins of the different relaxations were discussed in terms of the underlying molecular relaxation (or motional) processes (McCrum et al., 1967). Since that time several reviews and texts have been published which are concerned with the dielectric properties of polymers (Williams, 1972; van Turnhout, 1975; Hedvig, 1977; Williams, 1979; Jonscher, 1983; Williams, 1989). Thus, there already exists a substantial literature and review of both the experimental and theoretical aspects of the dielectric properties of polymers. The present account seeks to outline certain essential aspects of the phenomenological theory and molecular theory of the dielectric properties of polymers, to indicate some of the modern methods used to measure the
478
11 Dielectric Properties of Polymers
dielectric permittivity and loss factor and to describe representative results for different systems. It is worth noting that although the dielectric properties of polymers have been studied for over fifty years, (i) the general theory relating the complex permittivity to molecular behaviour in terms of the time-correlation functions of the motion is fairly recent (Cole, 1965; Cook et al., 1970; Williams, 1972); (ii) that there has been a marked improvement in dielectric measurement techniques during the past ten years as a result of improved instrumentation and computer control; (iii) the precise nature of the different relaxation processes seen for amorphous, crystalline and liquid crystalline polymers is still a matter of debate. A similar comment applies to the equivalent processes observed using dynamic mechanical relaxation, NMR relaxation, dynamic light scattering and the time-dependence of fluorescence depolarization.
11.2 Phenomenological Theory
where the subscript "r" refers to the complex admittance (relaxation component) and a is the dc conductivity (in Q~x cm" 1 ) of the sample. £,(/) and sj.'(/) depend on the frequency of measurement, £,(/) showing dispersion generally (a falling-off of its value with increasing frequency) and e" (/) showing an absorption peak (see Fig. 11-1 below). Experimentally, s(f) is the steadystate complex permittivity for the sample in the presence of alternating field of frequency / An alternative experiment is to apply a step-on or step-off field to the sample and to measure the effective permittivity as a function of time. The normalized decay function
(11-3)
— fin
where e0 and are the values of sr (t) at t = 0 and t -+ oo, respectively. These are also known, respectively, as the relaxed and unrelaxed permittivities (McCrum 1.2-1
11.2.1 General Frequency-Time Relations
The basic definitions of the permittivity and complex permittivity of a dielectric material are well known and will not be developed here (see, e.g., Smyth, 1955; Hill et al., 1969; Bottcher and Bordewijk, 1978; McCrum et al., 1967). It is usual to regard a dielectric sample as being a complex admittance shunted by a purely resistive element - which represents the steady dc conductivity component of the sample. The complex relative permittivity s(f) with respect to the free-space permittivity ev is given by 1.8xlO12<7
7
(11-2)
0.2-
Figure 11-1. Plots of the normalized permittivity (£r-eoo)/(eo-eoo) a n d l o s s factor £?/(£()-£«,) against log 10COT for a single relaxation time process, indicated as points o and • respectively.
11.2 Phenomenological Theory
479
etal, 1967). The connection between dielectric measurements made in the frequency domain and in the time domain is given by the Fourier transform relation below (McCrum etal., 1967; Williams, 1972).
curves in the frequency domain means that the relaxation function
-"^—^=l-io)j4>(t)-exp[-io)t]dt S £ °- " ° (11-4) where co = 2nf and where s0 and s^ now take the meaning of the limiting low and high frequency permittivities, respectively. Separating real and imaginary parts of Eq. (11-4) we may write
Several empirical functions have been proposed which give broadened loss curves in the frequency domain. The usual approach is to modify the SRT equation empirically, as was done by Cole and Cole (1941), Davidson and Cole (1950) and Havriliak and Negami (1966). The general equation which includes as special cases all three empirical functions is as follows
£r ( f) —
(11-6) where J^ and #~c indicate the one-sided Fourier sine and cosine transforms, respectively. For the special case where # (t) follows the single exponential decay function 0 (t) = exp (— t/x)
(11-7)
then Eqs. (11-5) and (11-6) give the familiar single relaxation time expressions
£
0 ~~ COT
b0
fc^
1 - f CO T
Figure (11-1) shows plots of the normalized permittivity and loss against log COT for this single relaxation time (SRT) process. The dispersion and absorption curves are complementary. The loss curve has a half width of 1.14 units in log! 0 COT and is symmetrical about log COT = 0. For solid polymers the dispersion and absorption curves are generally far broader, and often unsymmetrical, compared with that for a SRT process. Such broadening of loss
11.2.2 Empirical Relations for Permittivity
1 [l+(icoT)a]b
(11-9)
For a = b = 1 we have the SRT relation. For b = 1, 0 < a < l , Eq. (11-9) becomes the Cole-Cole function, for a = l, 0 < b < l , Eq. (11-9) becomes the Davidson-Cole function while for 0 < b < 1, 0 < a < 1, Eq. (11-9) becomes the Havriliak-Negami equation. The Cole-Cole function gives a loss curve which is broad but symmetrical about log COT = 1 while both the Davidson-Cole and Havriliak and Negami equations give unsymmetrical loss peaks which are skewed to the high-frequency side. For solid polymers in the amorphous state two dominant dielectric relaxation processes are observed. The a process (dynamic glass transition process) is well-fitted by the Havriliak-Negami equation (see, e.g., data for polyvinyl acetate by Mashimo et ah, 1982 b) while the p process (local motions of chain segments in the glassy state) is well-fitted in many cases by the Cole-Cole equation (see, e.g., data for polyethylene terephthalate in McCrum, Read and Williams, 1967). Other functions which represent empirical modifications of the single relaxation time function in the
480
11 Dielectric Properties of Polymers
where 0 < /? < 1. This function was used originally by R. Kohlrausch in 1854 to describe charge decay in the Leiden jar (Kohlrausch, 1854) but has subsequently found wide applications for creep and stress relaxation in glass-forming solids (see, e.g., Williams, 1985, for a review). The frequency-dependence of the dielectric permittivity corresponding to Eq. (11-12) was determined by Williams and Watts (1970) and Williams et al. (1971), by use of the (11-10) = J 0(r)exp(transform relation Eq. (11-4). It was shown o that broad, asymmetric dispersion and abwhich on insertion into Eq. (11-4) gives sorption curves were obtained for different values of /?, and these were very similar to those exhibited by the oc-process in amorphous polymers (see Williams et al., 1971; Experimental dielectric data may always Williams and Watts, 1971 a, b,c). It is also be fitted approximately by a suitable choice found that a wide range of organic and of the distribution function
frequency domain are due to Fuoss and Kirkwood (1941), Kirkwood and Fuoss (1941) and Jonscher, Hill and Dissado (see Jonscher, 1983). An alternative approach to empirical representations of dielectric relaxation is to express the relaxation function as a discrete or continuous distribution of single exponential decay functions. For the latter case we write
11.2 Phenomenological Theory
or loss factor. An alternative method of presentation of dielectric data is the Argand diagram in which s" is plotted against e!t which, for a single relaxation time process, is a semi-circle. For the Cole-Cole function a depressed semi-circle is obtained while for the Havriliak-Negami, Davidson-Cole and KWW functions a skewed arc is obtained which resembles the semi-circular arc only at the lowest frequencies. Such diagrams are a convenient and compact way of representing dielectric data for polymers, but plots of permittivity or loss against log frequency are to be preferred since they indicate clearly all three items of information which may be obtained from the dielectric experiments, namely (i) the relaxation strength(s), Asa, A^p, etc., (ii) the average relaxation time
The temperature dependence of
= A - exp
B T-Tr
In
applications of the WLF equation to dielectric, dynamic-mechanical and NMR relaxation data in the a relaxation region for amorphous polymers are widespread (Ferry, 1970; McCrum et al, 1967; McKenna, 1989). The common feature of the a relaxation in amorphous polymers is that the plot of log/ m vs. [T(K)]" 1 is strongly curved as Tg is approached from higher temperatures. Several theories have been proposed to account for Vogel-type (or WLF) behaviour, and those involving free-volume concepts are prominent (Ferry, 1970; McCrum et al., 1967; McKenna, 1989). The secondary (or P) relaxations for amorphous polymers give extremely broad loss curves but
where A and Q are material constants and Q is known as the apparent activation energy for relaxation (McCrum et al., 1967). The temperature dependence of
and are related according to
-RT'< C 2 + (T2 -
(11-14)
where subscripts 1 and 2 refer to temperatures Tx and T2, respectively, and where C1 = B/(T1 - To) and C2 = TY-
(11-15)
= . 4 - e x p ( ^
(11-13)
where A, B and To are material constants. The familiar Williams-Landel-Ferry equation (WLF equation) follows from Eq. (11-13) as
481
T o . The
/81nr\
pTjv{
dP )T
(11-16)
(11-17)
(see Williams, 1964a,b). Here (8P/8T)K is the thermal pressure coefficient and is given by the ratio of the thermal expansion coefficient and the isothermal compress-
482
11 Dielectric Properties of Polymers
ibility of the material. It follows from Eq. (11-17) that Qv < QP in practice. According to the simplest free-volume theories for relaxation in the Tg range (oc process) the value of
For polymers in solution and for unoriented amorphous polymers the theory for Ae is essentially that due to Kirkwood and to Frohlich (see McCrum et al., 1967, for a historical account). For an ensemble having N polymer molecules in a macroscopic spherical volume V the desired relation between Ae and molecular quantities is (Cook etal, 1970) (11-18) 4TI 3e o (2e <M(0) • M(0)> 3kT (2e0
11.3 Molecular Theory
where M(0) is the instantaneous dipole moment of the macroscopic sphere at the arbitrary time t = 0, and where M(0) may be expressed as the sum of the dipole moments, suitably defined, for the polymer molecules which are contained in the sphere. Through thermal motions the dipole moment vectors change their direction continually in time leading to <M(0)> = 0 but to a finite value of the macroscopic mean square moment <M(0)-M(0)>. Here "<>" indicates a statistical average over all configurations of the dipole moments in the sphere. For the special case of rigid dipolar rodlike molecules each having a dipole moment fi then (11-19)
11.3.1 Equilibrium Theory
The theory of the static permittivity of a polymer material is different for amorphous, crystalline and liquid crystalline systems. In each case a mean square dipole moment, suitably defined, is determined from the incremental dielectric permittivity Ae = e0 — e^ where s0 and e^ are the limiting low frequency and high frequency permittivities so that Ae contains the dispersion magnitudes of all dielectric relaxation processes for the material at the given sample temperature and pressure.
where iVv is the number of dipolar molecules per unit volume "//' is a "liquid" dipole moment which is related to the vacuum dipole moment fi0 according to the relation (11-20)
3(2e o
Combination ofEqs. (11-18), (11-19) and (11-20) gives the familiar Onsager equation _4nNv ~ 1kT
3e 0 1 P -UP
fe^ + 2^2
483
11.3 Molecular Theory
which allows JH0, the molecular dipole moment of the rod-like molecule to be determined from a knowledge of ATV, e0 and 8^. For flexible polymer chains in bulk or in solution it is convenient to consider as the representative reference unit any dipole group, k say, in any chain where the group is sufficiently removed from the chain ends. It follows that (Cook et al, 1970) <M(0) • M(0)>
(11-22)
and where
Nr and \ir are the number of dipole groups per unit volume and the dipole moment of a repeat unit, respectively. The terms
lowing Kirkwood) (11-24)
g = \+ E
is composed for the example of a model chain. A particularly convenient case is that of the polyether series [(CH 2 ) p _ 1 -O] n where for p = 2 we have polyoxymethylene and for p = 3 we have polyoxyethylene. Cook et al. (1970) adapted the theory of the dipole moments of these polyethers given by Read (1965) to the form equivalent to Eq. (11-24). Table 11-1 shows the results of these model calculations. Here <j> = 0 for the trans conformation and 4> = ± 120° for the gauche conformations. The coefficients akk±j add the contributions to
Table 11-1. Values of the coefficients akJi±j for polyoxymethylene (p = 3) for a tetrahedral chain for chosen values of the internal rotation parameter ?7 = <(cos(/>>. Polyoxyethylene
Polyoxymethylene
n -0.5 -0.25 0 + 0.25 + 0.50 + 0.75
a
k,k±i
-1.00 -0.639 -0.222 + 0.250 + 0.778 + 1.361
a
k,k±2
+ 0.250 + 0.084 -0.025 + 0.016 + 0.300 + 0.926
a
k,k±3
+ 0.062 + 0.013 -0.003 -0.007 + 0.117 + 0.630
a
k,k±i
-0.50 -0.183 -0.074 -0.187 -0.537 -1.137
a
k,k±2
+ 0.062 + 0.013 -0.003 -0.007 + 0.113 + 0.630
a
k,k±3
+ 0.024 0.000 0.000 -0.004 -0.038 -0.361
484
11 Dielectric Properties of Polymers
bor correlated dipoles. For POE all cross correlation terms are small for rj = 0 so in this case the dielectric permittivity increment is well-approximated by the contribution from the autocorrelation terms alone. As the ether groups are more-widely spaced along the chain [POM -> POE -> polytrimethylene oxide etc. (see Cook et al., 1970)] then the higher order terms are of decreasing importance if the chains have r\ around zero. However, when the chains take up preferred conformations leading to \r\\ quite different from zero then the cross correlation terms retain their importance over large separations of the groups, e.g., for POM for rj = 0.75 (trans-preferred conformation) the cross-correlation terms are large and positive while for POE for rj = 0.75 the cross-correlation terms are large and alternate in sign. These calculations demonstrate that for the special case of chains with free internal rotation and with well-separated dipolar groups, we may approximate the dielectric behavior as arising from the motions of the individual equivalent dipolar groups. However, for chains with strongly preferred internal conformational states, as is the case at low temperatures for most polymer chains, the value of g is an algebraic sum of autocorrelation (1) and cross-correlation terms «cos0 kk ,(O)». The theory of the dipole moment of polymer chains is well-developed through the work of Volkenstein (1963) and Flory (1969), and many examples have been treated quantitatively. Although it is possible to determine the dipole-correlation parameters using chain conformation models for bulk amorphous polymers together with experimental dielectric data, this is seldom done (McCrum et al., 1967). An interesting example is that of polyacetaldehyde (Williams, 1963) where the energy difference AE between trans and gauche conformations was found
to be ~ 3.9 kJ mol 1 using static dielectric permittivity data. The theory of the static permittivity of crystalline dipolar polymers is less well-developed. The anisotropy of the crystalline phase necessarily means the static permittivity becomes a tensor, having different values along different crystallographic directions. Since single crystals of polymers are extremely difficult to prepare, little interest has been attached to measurements of their dielectric properties. However, the average dielectric permittivity of unoriented and oriented specimens of lightly-oxidized polyethylene has been discussed by Boyd (1983) taking into account the anistropy of the permittivity of the individual crystals. Liquid crystalline polymers are a recent addition to the dielectrics literature and the theory of the static dielectric permittivity in this case has received scant attention. However, the theory of the static permittivity of low molar mass liquid crystals is well-developed through the works of Meier, Martin and Saupe, de Jeu and Bordewijk (see Bottcher and Bordewijk, 1978; de Jeu, 1980). For the case of a nematic (axially-symmetric) liquid crystal phase the principal components of the static dielectric permittivity tensor are eO|| and £ 0 1 where "||" and " ± " indicate measurements made parallel with or perpendicular to, respectively, the director n of the liquid crystalline phase. The generalization of the Onsager-Kirk wood-Frohlich equation for isotropic media to the nematic mesophase is outlined by Bordewijk (Bottcher and Bordewijk, 1978) who writes To
_ o
L
£
1
0y J
[8 Oy
3&r
(11-25)
485
11.3 Molecular Theory
where y is || or _L and hence gy is g^ or g ±, which correspond to the longitudinal and transverse Kirkwood g-factors and which contain terms relating to the (anisotropic) angular correlations between molecular dipole groups in the LC phase. Ny is the dipole group concentration, {eO||, s0 _,_} and {£oo||>£oo±} a r e the limiting permittivities measured parallel and perpendicular to n. The factors A^ and A ± are geometric quantities and (11-26) - S) • /j
S/2) • ft (11-27)
where fx and / t are local field factors and /j.l9 Ht are the dipole moment components of the molecular (or group) dipole moment parallel with or perpendicular to, respectively, the (assumed) principal axis of the molecule (or group). S is the local order parameter of the principal molecular axis with respect to the LC director axis and is typically in the range 0.4 < S < 0.8 for low molar mass liquid crystals. Although Eqs. (11-25) to (11-27) have not been applied, as yet, to liquid crystalline polymers, they have been used for low molar mass liquid crystals with some success (Bottcher and Bordewijk, 1978; Dunmur and Miller, 1980). Parallel correlations (gy > 1) and antiparallel correlations (gy < 1) between dipoles in the semi-ordered LC phase may be deduced from static permittivity data. These equations, Eqs. (11-25) to (11-27), may be applied to LC polymers having the dipole groups in the side chain. It is clear from the form of the above equations that experimentally-determined values of eO|| and 80 ± reflect, in part, differences in the terms g ^ and g ± (jU^)- In the absence of orientational correlations between dipole groups (i.e., g{l = g ± = 1) then e0 ,| and
a0 _L are seen to differ due to the presence of the order parameter S 4= 0 in Eqs. (11-26) and (11-27) and the dielectric anisotropy term (s0^ — s0 ±) is proportional to S. 11.3.2 Dynamic Theory The theory of the dielectric relaxation of dipolar media is well-documented (Hill et al., 1969; Bottcher and Bordewijk, 1978), but is made complicated by local field factors. For a detailed assessment the reader is referred to the accounts of Deutch and coworkers (Titulaer and Deutch, 1974; Sullivan and Deutch, 1975). The modern approach to dielectric relaxation theory, in which the complex permittivity is related to molecular dipole moment quantities independent of any assumed model for relaxation, is the linear-response theory of Kubo as developed for dielectric relaxation by Glarum (1960) and Cole (1965). Williams (1978) has outlined their theory for an ensemble of N equivalent dipolar molecules for the simple case of a medium of low permittivity, i.e., e0 ~ e^, as follows. The phase space distribution function f(p,q) obeys the Liouville equation of motion (11-28)
dt
where £ is the Liouville operator and
_ " [ 6 H 8/
8H 8/"
and H is the Hamiltonian of the system, pt and qt are the conjugate position and momentum of molecule (or group) i. For an electric field Ez (t) applied in the laboratory z-direction
H(p9q9t) = H0(p9q)-MM' EM (U-30) where Mz(q) - Ez(t) = - Z / i ^ ) • Ez(t) and is the energy of interaction between the
486
11 Dielectric Properties of Polymers
dipole moments and the electric field. After considerable manipulation, the field-induced average macroscopic dipole moment <MZ (t)} is obtained as the superposition relation (Williams, 1978) <M(0) • M(0)> =
JkT
• } EAn-^t-ndf
(n-31)
— oo
where <M (0) • M(0)> has its previous significance [see Eq. (11-18) above] and the total dipole moment correlation function <^(T) = <M(0)
• M ( T ) > / < M ( 0 ) • M(0)>
has
Eq. (11-34) we write
Le
(a)) = l - i
(11-36) ° where p (co) = x (co)/x (0). Comparing Eqs. (11-4) and (11-36) we see that for the special case p (co) = 1 that (^ (t) =
appeared. Equation (11-31) is the general result of the linear response theory for the arbitrary form for Ez (t). Three cases are of special interest. (i) Eo applied as a step at t' = 0
i k l
(11-32)
(ii) Eo applied at t -+ — oo, removed as a step at t' = 0
<M, W ) =
J
(11-33)
(iii) Steady state: E (t') = Eo exp (i co t') 3kT (11-34) where #" indicates the Fourier transform. Since the macroscopic electric polarization along the z-direction is given by 'Ez(t)'x-8y
(11-35)
where x is an internal field factor and av is the permittivity of free space, it follows that Eqs. (11-30) to (11-35) give the time-dependent or frequency-dependent permittivities for prescribed E(t). For example, from
(11-37) where dipoles /c and k' belong to the same chain and the strengths of the cross-correlation terms decay with increasing separation of dipoles along the chain, as discussed above (see also Cook et al., 1970). For flexible polymer chains it is convenient to think of dielectric relaxation as arising from motions of a reference dipole group k, giving the autocorrelation term *fc (0)' h (0) to which is added cross-correlation terms for dipoles (k, k') along the chain. Since Tg. Note that the linear response theory predicts that the transient rise and decay functions for stepon and step-off electric fields, respectively, are given by [1 - #„(£)] and (p^t) and that
11.4 Experimental Methods
the time-domain and frequency-domain behaviors are related through a Fourier transform [see Eqs. (1 l-32)-(l 1-34)]. Note also that the theory relates the dielectric properties to ^ ( t ) , the molecular property, and that several relaxation mechanisms may contribute to its decay from unity to zero. The theory of the dielectric relaxation of crystalline and liquid-crystalline systems will not be detailed here but will be described below when particular systems are considered.
11.4 Experimental Methods The conventional transient, lumped circuit and distributed circuit methods for measuring the dielectric properties of polymer systems are well documented, and cover, in total, the frequency range 10 ~6 to 1010 Hz (McCrum et al., 1967; Hill et al, 1969). During the past five years or so, modern instrumentation together with computer control of data-acquisition and processing has made a marked improvement in the speed with which accurate data for polymer systems may be obtained. We summarize briefly some of the techniques which are available for particular frequency bands. In the range 10 ~6 to 1 Hz, time-domain or frequency domain methods are used conveniently. Mopsik (1984) has described a particularly wide-ranging instrument for measuring polymer samples. Both permittivity and loss factor data have been obtained in the range 10 ~4 to 104 Hz by using measurements in the time-domain and carrying out a numerical Fourier transform to determine s(co). One inherent difficulty with the time-domain method is the enforced truncation of the Fourier integral since the data are band limited, but Mop-
487
sik (1984) has described methods for overcoming this difficulty. An alternative transform procedure to numerical integration is that due to Brather (1979) which has been applied to data for solid polymers by Ribelles and Calleja (1985) and by Attard etal. (1987 c). An advantage of the transient method is its speed but truncation and electrode-polarization effects make it difficult to use in practice. A low frequency bridge, the Solartron 1250 Frequency-Response Analyzer with a special high impedance input attachment allows measurements to be made in the range 10" 3 to 65 kHz at selected frequencies, and is particularly attractive for studies of polymers in the Tg range. Perhaps the most-used frequency range for studying the dielectric properties of solid polymers is 10-10 5 Hz. This may be covered with high precision and accuracy using commercial apparata such as the DETA instrument (Polymer Laboratories), the Wayne Kerr-B221 meter, the HewlettPackard 4272 LCR meter or the Gen Rad 1689 Digibridge. With computer control such instruments may be arranged to acquire accurate dielectric permittivity and loss spectra over this frequency range and over a wide temperature range and the data may be displayed on the screen and as hard copy. The Hewlett-Packard Model 4192A LF Impedance Analyzer provides a convenient method for measuring permittivity and loss over the range 5 Hz to 13 MHz. The range 10 5 -10 8 Hz connects lumped circuit and distributed circuit methods and has always been a difficult range experimentally. The Hewlett-Packard Model 4191A RF Impedance Analyzer with suitably-designed cells may be used for polymer systems in the range 10 6 -10 9 Hz. The method of time-domain-reflectometry, which involves measuring the time-de-
488
11 Dielectric Properties of Polymers
pendent reflexion coefficient of a sample contained in a coaxial line, has been developed in recent years by Cole and coworkers (Cole, 1977; Mashimo et al., 1983), to operate in the equivalent frequency range 10 7 -10 9 Hz. The method uses commercially-available equipment and has been used to obtain accurate dielectric data for polymer solutions (Mashimo et al., 1983). However, considerable expertise is required to obtain reproducible accurate data (Cole, 1977; Clarkson et al., 1977). Studies of the dielectric properties of insulating polymers (e.g., polyethylene, polytetrafluoroethylene) in the microwave range K^-IO 1 1 Hz continue to be made using conventional microwave methods operated at single frequencies, using methods described briefly by McCrum et al. (1967). However, the measurements are time-consuming and it is difficult to obtain data points over the entire range for a given material and to cover a range of sample temperatures. Consequently, few studies have been made for polymer systems in recent years. With regard to the geometry of samples, parallel-plate or coaxial configurations are frequently-used for lumped circuit measurements in the range 10" 6 to 108 Hz. In recent years, the parallel plate samples used would be typically 1 cm diameter, 100 jum to 3 mm in thickness and would be contained in a three-terminal electrode system (see, e.g., McCrum et al., 1967). A particularly useful, but inexpensive, cell for solids has been described by Jones et al. (1976).
11.5 Experimental Data for Polymer Systems 11.5.1 Introduction
As indicated above, the dielectric permittivity and loss data for dipolar polymer systems give information on chain conformation and structure through the mean square dipole moment (Volkenstein, 1963; Flory, 1969) and on the reorientational-dynamics of chains (McCrum et al., 1967; Williams, 1979). The existing literature for the dielectric behavior of different polymer systems is far too extensive to be reviewed in detail in the present account. In consequence, we shall seek to summarize briefly the kind of information on polymer behavior which can be deduced from studies of selected systems, and we shall attempt to give interpretations which reflect current models for chain dynamics. In a typical dielectric experiment the permittivity and loss curves for a polymer system are measured over a set frequency range (say 10-10 6 Hz) for different sample temperatures and, possibly, applied pressures. For polymer solutions a range of solute concentration c would also be studied. The information obtained experimentally is in three parts (i) relaxation strengths for individual processes, As{(T,P, c\ (ii) average relaxation time
11.5 Experimental Data for Polymer Systems
liquid, bulk amorphous solid, crystalline solid, liquid crystalline solid, we give brief accounts of selected studies in all such systems. 11.5.2 Experimental Data for Liquids and Solutions 11.5.2.1 Flexible Chains
The segmental motions of flexible polymer chains in solution or in the bulk liquid state at elevated temperatures would normally occur on a time-scale < 1 jas thus leading to dielectric absorption in the frequency region above 1 MHz. Such studies are difficult to perform experimentally so there is not a substantial body of dielectrics literature in this range. However, in recent years studies have been made up to 3 x 109 Hz for poly(methyl vinyl ketone) in dioxane (Mashimo et al., 1983), copoly(methyl methacrylate-methyl acrylate) in benzene (Mashimo et al., 1982 a) and for polypropylene oxides in benzene (Mashimoto et al., 1984). The results for poly(methyl vinyl ketone) are interesting because two distinct dielectric absorption peaks are observed, centred around 60 MHz and 1 GHz and are due to skeletal motions of chain segments in the backbone (low frequency process) and the independent motions of the side chains (high frequency process). This is a case where the dipole moment /i of the chain segment has components fih and /is associated with backbone and side group motions and the autocorrelation function for dipole relaxation may be written as
as + ah = l
and Cb (t) and Cs (t) are the normalized autocorrelation functions for the relaxation of \i\ and /x2, respectively. Thus,
489
decays in time with two component relaxations, giving two fairly-well-resolved loss peaks in the frequency domain. In physical terms this means that the motions of the side groups relaxes fi2 at short times (high frequencies) and the segmental motions of the chain relax the remainder of \x2, i.e., /i b . The correlation times for each process [Ti = (2nfmi)~1, z' = b,s] are, in this case, sufficiently different to allow each process to be observed independently as a loss peak. In many polymer systems we would expect Tb ~ TS, leading to a single broad loss peak which may not be deconvolved into the component relaxations since with at least six adjustable parameters (three for each process) the uncertainties in their determination become large. Note that if the side group motions are far slower than the backbone motions, fi2 will be relaxed along with [i\ so Eq. (11-38 a) would become, in this instance
+ Z Z £ (0) • lii) (0 tube-disengagement process, due to Klein i j (1978) was unable to rationalize the high value of m obtained in these experiments. where /i is the parallel dipole moment comThese studies by Adachi and coworkers ponent of a chain unit and rt (t) is the vector demonstrate how dielectric relaxation sum of the bond vectors along a chain so spectroscopy can give detailed information (r{ (0) • ri (t)y is the end-to-end vector coron the long-range motions of whole polyrelation function in times for a reference chain i. The terms , and the mechanism of this process changes on going from 11.5.2.2 Rod-Like Chains M < Mc (where the Rouse-Zimm modes apply) to M > Mc, where entanglements Synthetic polypeptides, such as poly-ylead to the M 3 7 power law for the relaxbenzyl L-glutamate, and certain polyalkyl ation time. isocyanates are rod-like molecules in soluRecently Adachi and coworkers have tion. They also possess cumulative dipole observed both LS and NM processes for moments along the chain contour, thus it is solutions of ds-polyisoprene in the good expected that they would give a low fresolvents benzene and toluene (Adachi and quency dielectric relaxation process in soKotaka, 1988; Adachi et al, 1989a, b, c) lution due to the small-step rotational dif-
493
11.5 Experimental Data for Polymer Systems
fusion of the molecules. The dielectric studies of polypeptides in solution were made by A. Wada several years ago, while similar studies of polyalkyl isocyanates were made some years ago by Yu and coworkers (Yu et al., 1966) and by Bur and Roberts (1969). We shall only discuss, briefly, the work with polyalkyl isocyanates. In the early studies of Bur and coworkers it was shown that poly(n-butyl isocyanate) in solution gave a well-defined dielectric absorption at low frequencies due to the reorientation of the molecules. It was shown that the molecules are rod-like up to M ~ 105 but a higher molecular weights the chains tend to form a rigidworm-like configuration. Subsequently, several dielectric and dynamic Kerr-effect studies were made for poly-n-butyl and poly-rc-hexyl isocyanate in solution and these are discussed in detail by Beevers et al. (1977). The essential results support the original conclusions of Bur and coworkers. Of special interest is the formation of lyotropic liquid crystalline (LC) solutions at high concentrations of the rod-like solute molecules. Aharoni had shown that polymers and copolymers of the higherpoly-alkyl isocyanates gave LC solutions at concentrations exceeding about 30% solute. In a series of papers Moscicki, Aharoni and Williams showed that solutions of copoly(n-butyl/rc-nonyl) isocyanate and poly(rc-hexyl isocyanate) exhibited a complex pattern of dielectric behavior on going from isotropic solution to lyotropicnematic LC solution via a bi-phasic range (Moscicki et al., 1981, 1982; Moscicki and Williams, 1983 a, b). As one example, Fig. 11-4 shows dielectric absorption data for poly(rc-hexyl isocyanate) (PHIC) (molecular weight 7.6 x 104) in toluene at 292 K and for concentrations ranging from 4.9% to 40% solute. In the initial range 4.9%-
24-
A
2016-
128-
//
/
1
/
^vK
k-
0-
i
1
i
2
i
i
i
3 log [MHz)]
4
5
Figure 11-4. e" against log [/(Hz)] for PHIC in toluene at 292.2 K. Curves 1-8 correspond to c equal to 4.9, 10.0, 14.8, 21.4, 25.0, 30.6, 35.3 and 40.4% polymer (w/w) respectively. (After Moscicki et al., 1982, reproduced with permission.)
21.4% a single broad dielectric absorption peak is observed and is due to the reorientational motions of the rod-like molecules in isotropic solution. The peak height is proportional to concentration in this range, as expected, and the frequency of maximum loss decreases due to the increasing viscosity of the solution. Above c = 21.4%, after a slight increase in height (at 25%), the loss peak magnitude falls markedly, broadens and moves to higher frequencies. Figure 11-5 shows plots of log / m , the static permittivity s0 and the loss-peak height e^ as concentration is varied throughout the entire range studied. It is apparent from the data of Figs. 11-4 and 11-5 that the bi-phasic range, which is a mixture of isotropic and lyotropic LC phases, forms for c > 25% and that the homogeneous LC phase exists beyond c ~ 35%. In the LC phase the rod-like molecules are unable to reorientate totally,
494
11 Dielectric Properties of Polymers
the rod-like molecule. Equation (11-40) shows that \x2 D\ is not relaxed, since the motions are limited to the cone (flip-flop motion is excluded) and that the strength of the dielectric absorption is proportional to D\. The effective relaxation time is
\oq[fm/[Hz))
T = [vJ(vi + l)D r ]- 1
c(wt.%) Figure 11-5. (a) Log [/m(Hz)], (b) s0 and (c) s'^ against solute concentration (% w/w) for PHIC in toluene at 292.2 K. (After Moscicki et al, 1982, reproduced with permission.)
their motions are restricted to angular fluctuations up to several degrees in an effective cone imposed by the LC nematic potential field. A convenient model for such motions is due to Warchol and Vaughan (1978) and Wang and Pecora (1980) (see also Williams, 1983 b) who deduced the dipole moment correlation function for a rod-like molecule undergoing free, smallstep rotational diffusion in a cone WVWP show that if the cone angle 90 < TT/3 then v{(vi + l)-D r t]}
(11-40)
where D°, D\ and v\ depend upon Qo, and Dr is the rotational diffusion coefficient of
(11-41)
For isotropic rotation v} = 1 so T = (2!),.)"1, but for motion in a cone v\ increases rapidly with decreasing 90 (Wang and Pecora, 1980), giving a marked decrease in T. Hence on going from isotropic solution to the LC phase, the loss peak is predicted to decrease in magnitude and move to higher frequencies and thus provides an explanation of the behavior shown in Figs. 11-4 and 11-5. Moscicki and Williams also calculated the phase behaviour of polydisperse rods in solution using the Flory model (see Moscicki and Williams, 1983 a and references therein) and, in combination with the WVWP model for dielectric relaxation for rods in a cone and the Perrin-Doty model for isotropic diffusion of rods in isotropic solution, were able to rationalize most aspects of the dielectric behavior of PHIC in its isotropic, biphasic and LC phases (Moscicki and Williams, 1983 b). 11.5.3 Experimental Data for Amorphous Solid Polymers
The dielectric properties of amorphous polymers have been reviewed extensively and in great detail (McCrum et al., 1967; Ishida, 1969; Williams, 1979,1985; Pethrick and Richards, 1982; Williams, 1989 and references therein). All amorphous polymers exhibit two relaxation processes, one, the a process, is due to the large scale microbrownian motions of chains and becomes prohibitively slow for temperatures below the operational glass transition Tg
11.5 Experimental Data for Polymer Systems
and the other, the (3 process, is due to the limited motions of chains and is normally observed in the glassy state. Certain general features of the dielectric relaxation of all amorphous solid polymers are apparent, irrespectively of chemical structure, distribution of molecular weight, tacticity or whether the material is a homopolymer or a random copolymer. The relaxation map which gives the loci of the different relaxation processes in all amorphous polymers is indicated schematically in Fig. 11-6. The oc-process follows the Vogel (or WLF) equations described in Sec. 11.2.3 above and as Tg is approached its frequency moves rapidly to very low values. The (3 process, which is usually, but not always, a small broad process is normally observed below Tg and its locus follows the Arrhenius equation also described earlier. As the temperature is raised through Tg the p process is continued to be observed until coalescence with the oc process leads to the formation of the oc (3 process which relaxes all of the mean square dipole moment of the chain.
10—,
5N X
' 0-
-5-
495
For many polymers, e.g., poly(vinyl chloride), poly(vinyl acetate), isotactic poly(methyl methacrylate), aromatic polycarbonates, polyethylene terephthalate) and high molecular weight poly(ethylene terephthalate) and high molecular weight poly(propylene oxide), Aea > AB^. However, there are notable exceptions for which Asp > A£a, as is the case for poly(ethyl methacrylate) and atactic and syndiotactic poly(methyl methacrylate). In the latter polymers the relative magnitudes of a and P processes may be reversed by applying pressure to the polymer. This leads to a marked decrease in Aep accompanied by a corresponding increase in Asa showing that oc and (3 processes are inter-related although the mechanisms for motion are entirely different. [For pressure studies see Williams (1979), for a review, Williams (1966 a, b), Sasabe and Saito (1968).] Without the need to discuss data for particular polymers, it appears that most of the features of the dielectric relaxation behavior of amorphous polymers may be understood in terms of the partial reorientation (P process) and total reorientation (a process) of a representative dipolar repeat unit of a chain (Williams and Watts, 1971 a, b, c; Williams, 1979). The group is supposed to have the possibility of a range of local environments (given by the probability °pr) and undergoes, at short times, partial relaxation in a given environment V , say, characterized by a relaxation strength \x2 • q^r and a correlation function >Pr(0- Summing over all partial relaxations, the remaining relaxation strength Aa = ^2YJ°pr(i — q$r) is relaxed by microbrownian motions, with the assoiated correlation function >a(t)The overall correlation function for group motion is given by
1/r Figure 11-6. Log [/m(Hz)] against reciprocal temperature for the a, p and (a P) absorptions in an amorphous polymer (schematic).
(H-42)
496
11 Dielectric Properties of Polymers
The relaxation map, Fig. 11-6 is consistent with this picture. For T < Ty9 >a(£) is so slow that only the P process can be observed. For T ~ Tg, a and P processes coexist in plots of e" vs. log / while for higher temperatures 0a(f) decays faster than the Pr relaxations, giving Aga for certain polyalkyl methacrylates merits further ex-
planation. In these polymers the ester group pendant to the chain undergoes motion fairly independently of the backbone, in the local environments prescribed by neighbouring chains in the solid state, hence most of \x2 is relaxed by this process under normal conditions. On raising pressure, the extent of this motion, but not the average rate of motion, is changed appreciably (Williams, 1966 b) thus As^ decreases and Aea increases. The fact that isotactic polymethyl methacrylate exhibits Aea > As^ indicates that the local motions of the ester group are extensively blocked in this polymer. Pressure is found to have a large effect on the frequency location of the oc process in amorphous polymers. As one example, Fig. 11-7 shows the permittivity and loss
Figure 11-7. e' and e" as a function of log (/// m ) for poly(ethyl methacrylate) at 273.2 K and different applied pressures. Curves 1, 2, 3, 4 correspond to 1, 340, 690 and 1020 bar respectively. (After Williams and Watts, 1971a, reproduced with permission.)
e"xio
log
[f/fj
11.5 Experimental Data for Polymer Systems
curves for the a relaxation in polyethyl acrylate (Williams and Watts, 1971a). As pressure is increased the loss curve moves rapidly to lower frequencies with a slight change in height and no change in shape. The quantity (81og/ m /8P) T lies in the range l - 4 k b a r - 1 , depending on sample temperature and pressure. The constant volume and constant pressure apparent activation energies, Qy and QP lie in the range 100-160 and 160-240 kJ mol" 1 , respectively, giving Qy/Qp ^ 0.70 and raising doubts regarding the applicability of simple free volume theories to such data (see Sec. 11.2.3 above). The master curves for permittivity and loss factor are shown in Fig. 11-8. The loss curve is broad and asymmetrical and is typical of that observed for the dielectric a relaxation in
497
amorphous polymers (see Williams et al., 1972, for a discussion). Although such a master curve may be fitted approximately by Davidson-Cole or Havriliak-Negami functions, a reasonable fit was obtained using the transform of the KWW function Eq. (11-12) above. The continuous curves in Fig. 11-8 were calculated for P = 0.38 and a satisfactory fit is noted. Williams and coworkers found that the a relation in a range of amorphous polymers including poly(vinyl acetate), poly(methyl acrylate), poly(propylene oxide), polyethylene terephthalate), poly(vinyl octanoate), acrylonitrile-butadiene copolymer (40:60), styrene-p-chlorostyrene copolymers and styrene-acrylonitrile copolymers could all be fitted using the frequency transform of the KWW function, with /? values in the
Figure 11-8. Normalized master curves for permittivity and loss for poly(ethyl acrylate) as obtained from the data of Fig. 11-7. •, • and • correspond to 1, 340 and 690 bar respectively. The continuous curves were calculated using the KWW function with 0 = 0.38. (After Williams and Watts, 1971a, reproduced with permission.)
498
11 Dielectric Properties of Polymers
range 0.4-0.7 but with most around 0.5 (Williams et al., 1971; Williams and Watts, 1971 a, b; Williams et al., 1972; Cook et al., 1975). The fact that the KWW function gave a reasonable representation of the dielectric a relaxation in polymers of such different chemical structures was intriguing and became the object of much attention (see, e.g., Williams, 1982 a, b, 1985; Ngai et al., 1986; Ngai et al., 1987 a, b; Rendell and Ngai, 1985). In addition, it was found that the KWW function also gave a reasonable representation of other relaxation phenomena for amorphous polymers in the glass-transition region including mechanical, NMR and volume relaxations, quasi-elastic light scattering and fluorescence depolarization (see, e.g., Williams, 1982a, b, 1985, 1989; Ngai et al., 1986; Matsuoka et al., 1985; Rendell et al. 1987; and references therein). The common occurrence of the KWW form for different relaxation processes studied in amorphous polymers also extends to studies of smallmolecule glass-forming systems (see, e.g., Williams, 1975; Williams and Hains, 1971, 1972; Williams and Crossley, 1977) and to ionic systems (see, e.g., Moynihan et al., 1976). The ubiquity of the KWW (or stretched exponential) function in molecular dynamics is presently a major challenge for the theorist. The cooperative nature of motions in polymeric and nonpolymeric systems which are glass-forming provides a common basis for all models of the dynamics of low frequency processes but the precise formulations which lead to KWW behavior is still a matter of discussion [see, e.g., Williams (1985); Blumen (1987), for a consideration of different models including the defect-diffusion models, kinetic Isinglattice models and the continuous timerandom walk models of Shlesinger and Montroll (1984)]. An interesting approach is taken by Ngai and Rendell (Ngai et al.,
1986, 1987 a, b; Rendell and Ngai, 1985) who take as the starting point the following rate equation for a relaxing quantity h9 say, (11-43) For W(t) equal to a constant, a single relaxation time process occurs. For W(t) having the following form W(t)=
Wo(coct)-n
(11-44)
i.e., a time-dependent rate coefficient with parameters coc and n, then for the isothermal case integration of Eq. (11-43) using Eq. (11-44) gives the KWW form (f/r)1""]
(11-45)
where (11-46) and T 0 = Wo *. According to Ngai and coworkers as n decreases away from unity, the degree of complexity of the relaxation increases. The application of Eqs. (11-43) to (11-46) to isothermal and non-isothermal relaxations in glass-forming polymers has met with success. Ngai and coworkers have reasoned that Eq. (11-43) should be regarded as the starting point for their "coupling model", and they have outlined different models which may lead to this equation (Ngai et al., 1986, 1987a). We note that Ngai and Rendell (1987) have used the KWW approach to fit dielectric data obtained by Adachi, Okazaki and Kotaka for poly(2,6-dichloro-l,4-phenylene oxide) in chlorobenzene. As the polymer concentration is raised from 2.5 to 56.9% the loss peak moved from ~ 107 Hz to ~10 3 - 5 Hz and broadened in the KWW sense. The polymer concentrations were below the entanglement condition for all solutions studied. Using a deconvolution method the KWW parameter /? was deter-
11.5 Experimental Data for Polymer Systems
mined for each solution and was found to vary systematically from 1.0 in dilute solution to 0.38 for the most concentrated solution. Of the many studies of the dielectric oc process in amorphous polymers that of Mashimo et al. (1982 b) for poly(vinyl acetate), which covered the range 10" 6 to 10 + 6 Hz, is perhaps the most comprehensive. In that case the loss curves were wellfitted by the Havriliak-Negami equation.
499
The permittivity values of dipolar amorphous polymers lie in the range 2-20. Recently, it was shown (Furukawa et al., 1986) that amorphous copoly(vinylidene cyanide/vinyl acetate) (VDCN/VAc copolymers) have exceptionally high permittivities due to the high polarity of the paired nitrile groups in the chain. Figures 11-9 and 11-10 show the permittivity and loss spectra for the alternating 1:1 copolymer. At the lowest frequencies the rising
190° C \ \ 195°C 3-
185°C \ 180
Iog6' \
\
\
•A\\
V
2-
175°C V
Figure 11-9. Log e' against log [/(Hz)] for the VDCN/ VAc copolymer at different temperatures. (After Furukawa et al., 1986, reproduced with permission.)
170°C 1-
-2
-1 log [MHz)
log 6
Figure 11-10. Log e" against log [/(Hz)] for the VDCN/ VAc copolymer at different temperatures. After Furukawa et al., 1986, reproduced with permission.)
01
1
I
CNI
-1
0
I
1 log If (Hz)]
I
2
I
3
500
11 Dielectric Properties of Polymers
loss factor indicates the presence of a conductivity process (the slope of the plot is — 1) while the rising permittivity indicates electrode polarization due to charge migration. Treating the sample as a complex capacitance shunted by a resistance which is in series with an equivalent capacitance Cel due to electrode polarization, with £e! oc (i co)~m where m is an empirical exponent, Furukawa et al. were able to resolve the different contributions to the observed permittivity and loss. Figure 11-11 shows, as one example, the result at 190 °C and the fit is extremely satisfactory. From Fig. 1111 we see that the static permittivity at this temperature is ~ 120 which is far larger than the values obtained for other dipolar polymers [poly(vinyl acetate) ~ 9.5, poly(acrylonitrile) ~43]. The loss curve may be fitted using the Havriliak-Negami equation with a = 0.8, b = 0.6 [see Eq. (11-9) above]. The frequency-temperature location of this process (a process) followed the WLF equation described above, with Tg = 170°C, C± = n.09 C2 = 5i. These data for a copolymer serve to illustrate how a dipole relaxation process which is partly obscured by conductivity and electrode processes may be extracted from the
experimental permittivity and loss curves. We note that Furukawa et al. (1986) comment that this copolymer exhibits a strong piezoelectric activity which is proportional to the remanent polarization induced by a poling field when the latter is very large due to the large static permittivity of the material. Up to this point we have discussed, mainly, the a relaxation in amorphous polymers. The (3 relaxation is always extremely broad with a half-width in the frequency plot of the loss-factor of 4 - 6 decades. It seems almost certain that this unusual broadness arises from individual local motions occurring with very different rates in a variety of local environments [see Eq. (11-42)]. A surprising feature of the P relaxation is that it follows the Arrhenius law, Eq. (11-15), in all polymers. Pressure is found to have only a small effect on log fm for the P process in all cases studied (see, e.g., Williams, 1966 a, b, 1979, Williams and Watts, 1971c). The effect of pressure on the strength Aep is more interesting. For polyethylene terephthalate (Williams, 1966 a) a 20% decrease was observed in e^ for 2kbar applied pressure while for poly(vinyl chloride) (Williams and Watts,
Figure 11-11. Comparison of the observed data (o, •) with the fitted curves for the VDCN/ VAc copolymer at 190 °C. The line - • - • - • - indicates the curve for the dipole relaxation process. (After Furukawa et al., 1986, reproduced with permission.) 1
2 log [//(Hz)]
3
11.5 Experimental Data for Polymer Systems
1971c) a 50% decrease was observed for 2 kbar applied pressure. For poly(ethyl methacrylate) (Williams, 1966 b) a decrease similar to that for PVC was noted for T < Tg but for T ~ Tg, Aep decreases by a factor of three for 2 kbar applied pressure. In all cases, the extent of motion is being decreased by application of pressure. The qualitative difference between the response for PEMA and PVC reflects the fact in PEMA the motion of the ester side group is being suppressed by application of pressure. With regard to a detailed mechanism for motion for the P process, simple models are precluded since they invariably lead to single relaxation time processes. The (3 process is also observed in non-polymeric glass-forming liquids, as was first made clear by Johari and Goldstein (Johari and Goldstein, 1970; Johari, 1987, and references therein) so its mechanism is not unique to polymer glasses. It seems likely that the dielectric (3 process in all glassforming systems arises due to limited reorientational motions in a variety of temporary local environments (Williams and Watts, 1971a) as expressed by Eq. (11-42) above. 11.5.4 Experimental Results for Crystalline Polymers 11.5.4.1 Introduction
Partially-crystalline polymers may broadly be classified into two categories, (A) those which have a degree of crystalline, X say, which does not exceed 50% and, (B) those which have X above about 80%. Melt crystallized polyethylene terephthalate, nylons 6-6 and 6-10, fall into category A while high density polyethylene, isotactic polypropylene and polyoxymethylene fall into category B. For materials of type A the amorphous phase may be glassy or mobile, depending on sample temperature,
501
and would be expected to exhibit multiple relaxations of the kind observed for amorphous polymers. For materials of type B, the amorphous phase is of a different kind from that in bulk amorphous polymers and is sometimes regarded as being disordered regions on the surface of crystals and within crystals. It is therefore expected that the dielectric relaxation behavior of these two categories of crystalline polymer will be quite different. The dielectric properties of several crystalline polymers have been reviewed and discussed in detail (McCrum et al., 1967; Williams, 1979,1982 a, b, 1989) and the reader is referred to those accounts for comprehensive documentation and analysis. The interest in the dielectric properties of crystalline polymers stems from their applications as high quality electrical insulation [poly(ethylene), poly(tetrafluoroethylene)], as piezo and pyroelectric materials [poly(vinylidene difluoride) and related copolymers] and as the plastic base for magnetic tapes (polyethylene terephthalate). We summarize briefly the essential dielectric properties of some of these polymers. 11.5.4.2 Polymers of High Degree of Crystallinity
Polyethylene occupies a special position among crystalline polymers owing to the simplicity of its chemical structure and its wide application for cable insulation (e.g., submarine telephone cables). Since the CH 2 group has only a very small dipole moment, pure polyethylene samples may have loss factors down to the practical limit of measurement (£"<10~ 5 ). However, practical materials are always slightly oxidized, with carbonyl groups decorating the chain and revealing intrinsic motions which occur in the disordered and crystalline regions. These motions may be studied
502
11 Dielectric Properties of Polymers
using dielectric relaxation spectroscopy and extensive accounts of the early experimental work are available (McCrum et al., 1967; Ishida, 1969; Wada, 1977; Hoffman etal., 1966; Ashcraft and Boyd, 1976). Chlorinated poly(ethylenes) (Matsuoka et al., 1972; Ashcraft and Boyd, 1976) and copolymers of ethylene and carbon monoxide, containing 0.5-1% carbon monoxide comonomer have also been studied. In all cases, three relaxation regions, a, P and y in decreasing order of temperature are observed. The definitive studies are those of Ashcraft and Boyd (1976) which gave extensive data for lightly oxidized or chlorinated low density and high density polyethylenes. In all cases, the carbonyl or C-Cl groups act as a probe on the motions which occur even in the pure polymer. The essential findings are as follows. (i) The occ process is assigned to a rotation-translation of chains in the crystalline lamellar. Support for this mechanism is provided by the observation that the a process disappears on melting oxidized linear polyethylene. A rotation-translation model has been modelled by Frohlich (1958) and by Hoffman et al. (1966) and it is shown that for short alkane chains, up to ~ 40 A long, the motion is that of a rigid rod while at higher chain lengths, corresponding to those for lamellar crystals, the chain reorientation is assisted by chain twisting, leading to a constant apparent activation energy for the process. Subsequently, Mansfield and Boyd (1978) made conformational energy calculations for the twistassisted reorientation process. They find that reorientation is accomplished by a twisted region ~ 1 2 C H 2 units long which propagates smoothly along the chain and hence across a crystal lamellar. In view of the detailed experimental data available for the a relaxation in the different polyethylenes, and the success in representing
the molecular-length dependence of the relaxation parameters (Hoffman et al., 1966; Ashcraft and Boyd, 1976; Mansfield and Boyd, 1978), the mechanism for this process is perhaps the best understood of all those studied in crystalline polymers. (ii) The P process is assigned to the motions of chain segments in the "amorphous" regions of the polymer, and is analogous to the a process observed in amorphous polymers, its activation energy being characteristically large (220 kJ mol" 1 ). The p process is more intense in low density polyethylenes. (iii) The y process is thought to be a composite process containing contributions arising from the amorphous and crystalline regions (Hoffman et al., 1966). Support for this possibility is provided by Kakizaki et al. (1985) whose studies of the dielectric, mechanical and NMR relaxations in linear polyethylene indicate that the y process arises from motions in an interlamellar amorphous region and at the lamellar surface. However, Ashcraft and Boyd (1976) conclude that the dielectric y process originates in the amorphous phase. As is the case for the P-process in amorphous polymers, the y process in oxidized and chlorinated polyethylenes is extremely broad in the frequency domain. Ashcraft and Boyd (1976) suggest that this arises because the chain segments in the amorphous phase find themselves in a variety of local environments, each with its local barrier system to reorientation. As a result a broad distribution of activation energies is presented for the y relaxation leading to a very broad dielectric absorption. This interpretation is very similar to that given above for the P relaxation in amorphous polymers (see Eq. (11-42). The dielectric properties of polymers having high degrees of crystallinity, such as polyoxymethylene, polytrimethylene oxide
11.5 Experimental Data for Polymer Systems
and polytetramethylene oxide, have been studied extensively (see McCrum et al., 1967; Wetton and Williams, 1965). Two relaxation processes, termed p and y in descending order of temperature, are observed. In a classic study, Ishida et al. (1965) showed that a single crystal mat laminate of polyoxymethylene exhibited only the y process while the same specimen, on melting and recooling to form a melt-crystallized bulk sample, exhibited both y and (3 processes. This may be taken as evidence that the y process arises from motions of the ether dipoles in defect regions associated with the crystals and in the disordered regions on the lamellar surface. The P process is due to motions of the amorphous regions beyond the crystal surfaces; but it is likely that such motions are strongly hindered by the presence of the crystals, as is the case for polyethylene terephthalate (see below). As the sample temperature was lowered from 35 °C to — 75 °C the P-loss peak in bulk polyoxymethylene broadens markedly, indicating that the motions in the amorphous regions are increasingly constrained by the crystalline regions. These studies demonstrate that different motional processes in crystalline polymers may be detected using dielectric relaxation spectroscopy and that the influence of the crystalline regions on the motions in the amorphous regions are manifested as a broadening of the loss process as sample temperature is decreased. The effect of pressure on the dielectric a and y relaxations has been described by Sayre et al. (1978), while Gilchrist (1978) and Yano et al. (1977) have given dielectric data for poly(ethylenes) at liquid helium temperatures. The important feature of the latter results is that facile motions of hydroxyl groups present as additives in the pure polyethylene are detected as low fre-
503
quency-low temperature dielectric absorption peaks. 11.5.4.3 Polymers of Medium Degree of Crystallinity Of the many polymer materials which may be crystallized from the melt to give samples of only medium degree of crystallinity, the nylons and polyethylene terephthalate are of particular interest due to their continued commercial importance. The dielectric properties of these polymers have been described by McCrum et al. (1967). For the nylons, multiple relaxation regions are observed, the a process being due to motions of the amide groups in the amorphous regions, and the shape of this process broadens markedly as sample temperature is decreased which indicates that the crystalline regions hinder the motions in the amorphous regions, as we have discussed above. In an important study Boyd and Porter (1972) found that the dielectric oc process in nylon 6-10 was continuous on going from the crystalline solid into the melt, demonstrating that the process is occurring in the amorphous regions in the solid. The strength of the dielectric a process increased by a factor of two on melting the polymer which is predicted since more amorphous phase is formed at the expense of the dielectrically-inactive crystalline phase. Further work by Yemni and Boyd (1979) for the odd-numbered nylons, 77 and 11, showed that the dielectric strengths of the a processes in these materials were larger than those for the even-numbered nylons, which is consistent with the fact that the dipole moment components of the chain perpendicular to the chain-axis are parallel to each other in the odd nylons and oppose each other in the even nylons. Such studies demonstrate that the dielectric relaxation behavior is sensitive to
504
11 Dielectric Properties of Polymers
chain structure and conformation in a way which is not found for dynamic mechanical relaxation. The motions of the amide groups in the nylons are made complicated by the inter-molecular effects of hydrogen bonding. One linear polymer which is of medium crystallinity but does not have hydrogen bonding is polyethylene terephthalate (PET). It was first studied over wide ranges of frequency and temperature by Reddish (1950) who found that a melt crystallized sample gave well-defined a and (3 relaxations where the a process is due to the large-scale microbrownian motions (dynamic Tg process) of the chains in the amorphous regions and the P process is due to local segmental motions of the chain in the glassy amorphous phase and is observed below the Tg of the polymer. A full account of the early studies of PET, for different degrees of crystallinity are given in McCrum et al. (1967). Essentially, it is found that the degree of crystallinity (X) may be varied in the range 0-0.5 by quenching the melt to room temperature, giving amorphous material, then heating to different temperatures to induce different degrees of crystallinity (quench-anneal method). Alternatively, by cooling from the melt slowly into the crystalline state, melt-crystallized materials is obtained. The dielectric data show that as X is raised from 0 to 0.5, the dielectric oc process broadens, moves to lower frequencies and decreases in its relaxation strength (due to removal of amorphous phase). However, the a process in the wholly-amorphous material and the a process in the crystallized material (X ~ 0.5) have been shown to be qualitatively different as a result of the following experiment. Tidy and Williams (1978) (see Williams, 1979) prepared an amorphous sample of PET by quenching a sample from the melt. Rapid heating to 106.7 °C gave an amorphous material
above its Tg whose loss spectrum could be measured in the a relaxation range. Keeping the sample at this temperature resulted in its crystallization over a period of time. At this temperature approximately six hours were required to crystallize the sample to the extent that the material appeared fully-spherulitic in the optical microscope. The dielectric spectra were recorded for different times of crystallization. It was found that as the normal a peak reduced in magnitude, due to crystallization, its frequency, location and shape did not change. In parallel with the disappearance of the a peak, a broad, lower frequency process appeared (a' say) and grew as crystallization proceeded. In the later stages of crystallization, the magnitude of this peak decreased with time, indicating secondary crystallization. The peak of the a' process occurred at ~ 2 units of log 10 frequency lower than that for the normal a peak. This experiment suggests that the motions of the chains in the crystallized specimen (a' process) are hindered by the presence of the crystallites, giving a lower frequency (larger average relaxation time) for the motion and a marked increase in the breadth of the loss peak where the latter indicates that the motions occur in a variety of local amorphous states influenced by the crystallites. It is interesting to note that the crystallized sample appears to be all spherulites, as observed in the optical microscope, so the "abnormal" amorphous phase is contained entirely within the spherulites. The dielectric data show there is an intimate connection between the disordered and ordered regions within spherulites. More recently, Coburn and Boyd (1986) studied nine specimens, ranging from X = 0 to X = 0.62, over wide ranges of frequency and temperature. The partially crystalline samples were prepared by heat-
11.5 Experimental Data for Polymer Systems 1
4.7
1
1
1
1
I
I
I
I
I
I
I
I
I
I
I
505
I
O 1hz 0 10hz a 100hz ^ 1khz O 10khz a 100khz
-
4.5 ~
~~4.3 a
Jillrr
o
£3.9 o .« 3.7 Q 3.5 3.3 I
-200
I
-160
I
I
-120
[
I
-80
I
-40
I
I
I
I
0
I
I
80
I
I
120
I
I
160
Temperature (°C)
(a)
0.12
I
-
0.08 =r
Figure 11-12. (a) Curves of permittivity e' and (b) curves of loss factor e" against temperature at given frequencies for a partially crystalline sample of polyethylene terephthalate (X = 0.38). (After Coburn and Boyd, 1986, reproduced with permission.)
0.04 -
0.00 -200
(b)
-160
-120
-80
-40 0 40 Temperature (°C)
ing the glassy amorphous film to different temperatures above Tg and allowing crystallization to go to completion. The samples of the highest degree of crystallinity were prepared by isobaric crystallization from the melt at high applied pressure, followed by annealing at 100 °C stabilise their density. Plots of e" vs. T at chosen frequencies clearly revealed the a and (J relaxations in each specimen. Figure 11-12 shows, as
120
160
one example, the permittivity and loss data as a function of temperature for a sample having X = 0.38. The entire set of data for all samples was fitted numerically using a Havriliak-Negami function for each process. Such data-analysis provides an example of what can be presently achieved using modern numerical methods. The strength factor Ae, the relaxation frequency fm and the Havriliak-Negami parameters, a, b,
506
11 Dielectric Properties of Polymers
[Eq. (11-9) above] were obtained for a and (J processes for all nine samples over a wide range of temperature, thus allowing conclusions to be drawn based on quantitative data. The essential observations are as follows. The relaxation strengths Asa and Aep were shown to decrease approximately linearly with increase in sample density (Q in the range 1.34 to 1.4 gem" 3 ). Zero relaxation strengths for both processes would appear to occur at crystallinities, X, comparable with that for the single crystal, hence both processes are considered to arise from the amorphous regions of the materials (Coburn and Boyd, 1986). Using a composite model for the dielectric permittivity of a partially-crystalline polymer, as developed by Boyd (1983), it was possible to calculate the permittivity of the amorphous phase in each sample, and hence use the Onsager-Kirkwood equation to calculate the correlation factor [see Eq. (11-24) above]. The composite model leads to upper and lower bounds for the measured permittivity according to the equations By + 2 8H
1 1 / 1 eL 3 \ e v
(11-47 a)
2 ,
(ll-47b)
where s^ and eL are the upper and lower bounds of the measured permittivities defined by — v)&2 1
v
1-v
—=—+——
(11-48 a) (11-48 b)
where &x and s2 are the dielectric permittivities of the amorphous and crystalline phases, respectively, and v is the volume fraction of the amorphous phase.
Proceeding in this way, the experimental values for the g-factor for the (3 relaxation were in the range 0.05-0.07 for all nine samples while those for the combined a and (3 processes were in the range 0.230.29, depending upon the sample and its temperature. The low value for g for the (3 process seems not to be due to orientational correlations between dipoles, but is due to the possible inapplicability of the Onsager-Kirkwood equation to the glassy state. The derivation of this equation assumes that individual dipoles may access 471-solid angle of orientation space, whereas in the glassy state the dipole motions are spatially limited - see Sec. 11.5.3 above for a discussion of oc and (3 processes in amorphous polymers. The consistency of the g-values for the combined oc and (3 processes is impressive. According to Coburn and Boyd, the relatively low value of the g-factor (-0.31) for the a and (3 processes in all samples arises from intramolecular correlations where the dipole moments of the two ester groups on each ring partially cancel, the degree of cancellation depending on the conformation (cis or trans) of the two groups with respect to each other. The g-factors for the semicrystalline specimens were all significantly lower than that for the amorphous sample (# a ~ 0.33-0.29 in the range 60-100°C). This property seems to arise from the restriction on conformations and spatial freedom for the chains due to the presence of crystals to which they are directly connected. Thus, the static permittivity data of Coburn and Boyd indicates that the amorphous phase in the partially crystalline PET is influenced in its equilibrium properties by the presence of the crystallites. The a relaxation in the amorphous material is far narrower, and more skewed, in the Davidson-Cole or KWW sense, than the fairly symmetrical and broad a' process
11.5 Experimental Data for Polymer Systems
observed in the partially crystalline specimens. There is a marked difference in the relaxation map for the a process between that for the amorphous sample (a) and those for the crystalline samples (a'). However, according to Coburn and Boyd, as X is raised, the locus is displaced to lower temperatures, indicating significant changes in the nature, and hence the dynamics, of the amorphous regions within the spherulites as the degree of crystallinity is raised. The samples prepared by crystallization under pressure, which have the highest crystallinities of all samples measured, give relaxation maps for the a' process which are rather different from those for all the other partially-crystalline samples in that the slope of the plot log/ m vs. 1/T is lower than that for the others. It seems possible that the amorphous regions in these pressure-crystallized samples are less constrained by the crystals than are the corresponding amorphous regions in the isothermally-crystallized specimens (Coburn and Boyd, 1986). Further support for this suggestion is obtained from the earlier mechanical and structural work of Illers and Breuer (1963). These detailed studies of polyethylene terephthalate show that the dielectric method enables information to be obtained on the nature, structure and the dynamics of the amorphous regions within a partially crystalline polymer. It is evident that the amorphous phase is abnormal in the sense that the crystallites limit and slow-down the motions in the amorphous regions and also broaden the loss process, indicating that there are a wide range of environments in which the cooperative motions of the chains, which lead to the a' process (and hence the Tg process), occur. The dielectric data may be complemented by NMR data, which also may give information on different ranges of mobility in
507
the amorphous phase in polyethylene terephthalate, as has been described by English (1984) and discussed by Coburn and Boyd (1986). 11.5.5 Experimental Data for Liquid Crystalline Polymers 11.5.5.1 Introduction
Thermotropic liquid crystalline (LC) polymers having the mesogenic (liquidcrystal-forming) groups in the main chain and/or in the side chain of a linear polymer have been the object of intense interest during recent years. The interest in main chain LC polymers arises from the fact that they provide anisotropic solid materials with a high modulus in the direction of orientation and which have a low coefficient of expansion. Their applications are in the form of sheet material, or film or fibre material, with good mechanical, electrical and thermal properties. From the dielectrics point of view, the interest would be to study the different thermal transitions, but only a small amount of work has been done (Laupretre et al., 1985) for the main chain polyester. On the other hand, the interest in side chain LC polymers concerns their possible use as materials, in the form of films (1-100 jam thick) for optical information storage (Shibaev et al., 1983; Coles and Simon, 1983, 1985; Eich etal., 1987; Kaempf, 1987; Schmidt, 1989; McArdle, 1989), for nonlinear optics (Williams, 1983 a; McArdle, 1989; Chemla and Zyss, 1987) and for optical elements such as diffraction gratings, optical waveguides and Fresnel zone plates (Eich et al., 1987; McArdle, 1989). All applications of LC side chain polymers require a material to be aligned macroscopically in the homeotropic (H), homogeneous (Hs) or planar (P) forms. Figure 11-13 illustrates the H- and Hsaligned materials by showing the orienta-
508
11 Dielectric Properties of Polymers
(a) Homogeneous Alignment
(b) Homeotropic Alignment Figure 11-13. Schematic illustration of the organization of the mesogenic groups in homeotropic and homogeneous alignments of a nematic LC material, n indicates the director axis for the LC phase. The main chains of the polymer are not shown.
tion of the director n, for a non-chiral liquid crystal, with respect to the geometry of the film. Both H and Hs samples are monodomains of the LC phase (shown as nematic phase in the figure) and differ in that n || x y-plane for the Hs sample and nixy-plane for the H sample. The planar arrangement (not shown) has n\\xyplane, like the Hs form, but the local directors are random in the x y-plane, unlike the Hs form. All three forms may be achieved for a film of a LC side chain polymer using methods which have been developed for low molar mass liquid crystals (see Clark etal, 1980; Clark, 1981; Pranoto and Haase, 1983; McArdle, 1989). These involve either surface treatment of the substrate to which the LC material is attached or application of dc or ac electric fields or magnetic fields. Having achieved a film having a particular orientation, full advantage may be taken of its anisotropic optical properties. Thus, for example, data storage may be achieved with H- or Hs-aligned films using a focussed laser beam which
heats the material locally, producing a scattering texture. This is known as the thermorecording method (see Shibaev et al., 1983; Coles and Simon, 1985; Simon and Coles, 1986; McArdle etal., 1987; McArdle, 1989; Schmidt, 1989). An alternative method is to include photoactive groups in the side chain (as mesogenic groups) or as guest molecules in the LC polymer host. In these cases, a polarized focussed laser beam acting on an aligned film will produce a birefringent, dichroic image on the film through the process of photobleaching. This optical recording method for LC side chain polymers has been described by Eich et al. (1987) (Eich and Wendorff, 1987; and by Anderle et al., 1989). The optical recording process appears to involve photoisomerization of groups via angular dependent photoselection and is applicable to both LC polymer films and glassy polymer films. The precise details of the mechanism of this optical recording method are still to be established for LC films but the mechanism of angular dependent photoselection has been demonstrated for a photochromic glassy polymer (Jones etal., 1989; Kozak and Williams, 1989). A feature of both the thermorecording and optical recording methods, as applied to LC polymers, is that reversible digital, analogue or holographic optical storage at high densities may be achieved (see Schmidt, 1989). It is important to remember that the electric-field induced switching of a LC material, of low molar mass or polymeric, is caused by the dielectric torque, which is proportional to 5s = s^ — s±9 where £|l and e ± are the dielectric permittivities measured parallel with and perpendicular to, respectively, the director axis w, of the liquid crystal (assumed to be axially-symmetric for simplicity). e(| and £ ± are frequency dependent for a LC polymer mate-
11.5 Experimental Data for Polymer Systems
rial and a dielectrically-positive material will give dispersions and absorptions as indicated schematically in Fig. 11-14. For such a material the crossover frequency / c for the permittivity will determine the macroscopic alignment achieved by the mesophase when a strong ac electric field of frequency fd is applied. H-alignment is obtained if fd
log [MHz)]
Figure 11-14. Schematic illustration of the frequency dependence of the real and imaginary components of £|l (co) and s ± (co) for a dielectrically-positive LC material. Note the crossover frequencies fc and / c ' for permittivity and loss factor, respectively.
509
chiral groups. Such films show promise as fast optical switches for display technology. In the account below we summarize briefly some of the features of the dielectric properties of selected LC polymer materials. 11.5.5.2 Non-Chiral Liquid Crystalline Side-Chain Polymers
The earliest studies of the dielectric properties of LC side-chain polymers were made by Kresse and co-workers (Kresse and Talrose, 1981; Kresse, 1982; Kresse and Shibaev, 1983, 1984) for polymers of the structure
(1)
where ,R = H (acrylate polymer), R = (methacrylate polymer). For ,R = H the polymer formed a nematic (N) LC phase and had the transitions gN323N360I, where gN indicates glassy nematic. For R = CH3 a smectic (S) phase was formed. The transition temperatures were gS333S397I, where gS and I indicate glassy smectic and isotropic phases, respectively. All transition temperatures are in Kelvin. A well-defined relaxation process was observed for unaligned or partially aligned samples of both polymers, with relaxation frequencies in the range 1 to 105 Hz and, for a given temperature, the motion in the (smectic) methacrylate polymer is substantially slower (~ 3 decades) than that for the (nematic) acrylate polymer. A further feature of their data is that log fm shows only a slight increase on going from N -> I for the acrylate polymer, whereas Alog/ m = 0.85 for the acrylate polymer at the S -> I transition. Subsequently, Zentel et al. (1985) studied a num-
510
11 Dielectric Properties of Polymers
ber of unaligned acrylate LC polymers and observed, 5, a, p, yx and y2 processes in their dielectric experiments. They have assigned each process to particular motions of the dipolar groups. Further studies were made by Attard et al. (1986) for a siloxanechain polymer which had the following structure
(2)
gN 273 N 3261. Similar results to those obtained by Kresse and Talroze (1981) and Kresse et al. (1982) were obtained in that one broad process was observed in the LC state which narrowed on going into the isotropic state, and the process showed a slight increase in relaxation frequency at the transition. However, it is not possible to analyze such data to reveal the underlying relaxation processes. That may only be achieved by studying the dielectric properties of materials prepared in different states of macroscopic alignment. Systematic studies of the dielectric properties of LC side-chain polymers in different states of alignment have been made by Haase and co-workers (see Pranoto et al., 1984; Haase and Pranoto, 1985; Haase et al., 1985; Pranoto et al., 1986; Bormuth and Haase, 1987, 1988; Bormuth et al., 1987; Haase and Pfeiffer, 1990, and references therein) and by Williams, Attard and co-workers (see Attard et al., 1986; Attart and Williams, 1986a-e; Attard et al., 1987a-d; Attard, 1986; Araki et al., 1988; Kozak et al., 1989 b; Attard et al., 1988, 1989; Williams et al., 1990; Attard and Araki, 1986; Araki and Attard, 1986; and references therein). It is not possible to give a complete description of all these studies here. Instead, we consider representative examples and
initially describe briefly the study by Attard and Williams (1986d) and Attard et al. (1987 c, 1988) of the following smectic LC side-chain polymer CH 3
si-of O-(CH 2 ) 8 -O-<^
(3)
where n ~ 35. The glass transition temperature Tg = 274 K and the clearing temperature Tc = 360 K. A 100 jim thick sample was aligned homeotropically using 300 V (root mean square) at 3 kHz applied to the melt sample followed by cooling into the LC phase with the voltage maintained. The H-aligned material was optically transparent, the unaligned (U) sample was turbid. Figure 11-15 shows plots of G/co(= &" Co) against log 10 [/(Hz)] for the U and the Haligned samples. Here C o is the geometrical capacitance of the sample. Several features of the relaxation curves are apparent: (i) The U-sample spectra are broad with evidence of a bimodal feature while the Haligned sample spectra are close to those for a single relaxation time process (halfwidth ~1.4 compared with 1.14 for the SRT process). (ii) a^ax is increased by a factor of ~ 2.3 on going from U -• H-aligned sample but the integrated areas below both curves, at a given temperature, are not very different. (iii) log/ max (observed) for each sample increases rapidly with increasing temperature at a rate comparable with that for the oc process in amorphous polymers (see Sec. 11.5.3 above). The apparent activation energy depends on temperature but is near 130kJmol - 1 for both samples in the higher temperature region just below Tc. The variation in loss spectra and the origins of the loss processes may be under-
11.5 Experimental Data for Polymer Systems 15-
10GT
/
2
^
/
Q_
30-{
u 1
I
I,
•Q
/
T
511
3 ,--—\ 20-
343.2K
t
328.2K
3r
320.2K
j |.
309.2K
02
3 log 10 lf(Hz)]
log 10 lf(Hz)]
Figure 11-15. E" - C0( = G/co) plotted against log[/'(Hz)] for unaligned (U) and homeotropically-aligned (H) samples of LC polymer (3) of the text, for different sample temperatures. Curves 1 - 4 correspond to 309.2, 320.2, 328.2 and 343.2 K, respectively. (After Attard and Williams, 1986d, reproduced with permission.)
stood in terms of the results of earlier work with low molar mass liquid crystals (Maier and Meier, 1961; Nordio et ah, 1973; Luckhurst and Zannoni, 1975; Bottcher and Bordewijk, 1978; de Jeu, 1980; Kelker and Hatz, 1980). The dielectric permittivity for a uniaxial liquid crystal (e.g., nematic, smectic A, which are also the simplest forms of thermotropic liquid crystals) is a tensorial quantity s(co) whose principal permittivities are s^{co) and s±(co) where subscripts "||" and "_L" mean measured parallel to or perpendicular to, respectively, the LC director axis n. From Fig. 11-13 if s(co) is measured in the z-direction then the H-aligned sample gives s^(co) and the Hs-sample gives £L(co). A generalization of the theory of the static permittivity of a nematic LC material by Maier and Meier (1961) to the dynamic situation has been made, independent of any assumed model for motions of the mesogenic groups, by Attard et ah (1987d) and by Araki et ah (1988). They obtain the following relations for the principal permittivities
+ (1 - S)
(ll-49a)
G
3kT
[(!-.
(ll-49b)
where e^y and 8 ^ are the limiting high frequency permittivities measured parallel or perpendicular, respectively, to the local director n. G is a constant, S is the local order parameter for the LC phase, /JLX and jnt are the longitudinal and transverse components of the dipole moment fi of the mesogenic group and Fj(w) = l-ico^
[Fj (t)]
(11-50)
where ^ indicates a one-sided Fourier transform [see Eq. (11-4)] and the Fj(t) are
512
11 Dielectric Properties of Polymers
linear combinations of certain time-correlation functions #p m (0 for the orientational motions of the mesogenic groups and are given by
(11-51)
where the Dpm (Q) are Wigner rotation matrix elements expressed in the laboratory frame. Equations (11-49) indicate that four orthogonal dielectric relaxation modes will occur, two for e^ and two for s ±. We denote the modes as 00, 01, 10 and 11 and their relative strengths are given (1 + S/2) • /it2. Pictorial representations of the motions of the axes in Euler space (a, /?, y) of the mesogenic grouping have been given by several workers (see, e.g., Attard, 1986). The 00 mode has a relaxation function given by (\\ S7\ where /? is the polar angle in between the molecular z-axis and the laboratory Z-axis. The well-defined sharp loss peak seen for the H-aligned sample in Fig. 11-15 arises from the 00 relaxation mode, i.e., it is due to the motions of fi{ about the short axis of the mesogenic group. The high-frequency tail to the overall absorption is assigned to the 01 mode. The loss curve for the U-sample is due to a superposition of the four component relaxation modes, suitably weighted. The permittivity of the U-sample corresponds to the average permittivity E(co) which is given by the relation
So from Eqs. (11-49) and (11-53) we see how the four modes contribute to &(co). The broad loss curves for the U-sample (Fig. 11-15) cannot be resolved into its four components without making further assumptions. Attard and Williams (1986d) assumed that the 01, 10 and 11 relaxation
modes all had the same relaxation functions, allowing the loss peak for the Usample to be decomposed into a low frequency process (01, 10 plus 11). The fits of the observed loss data for H-aligned, partially-aligned and U-samples of polymer (3) two component loss curves were satisfactory, and allowed calculations to be made for (S,/iJiid giving the values (0.63, 1.43). Thus the dielectric loss curves could be analyzed using a molecular theory for the anisotropic motions of the dipolar mesogenic groups, without specifying the precise nature of those motions, e.g., small-step diffusion vs. large step-diffusion of mesogenic group reorientations in the LC local potential. Note that the main source of the dipole moment (and hence incremental permittivity) in polymer (3) resides in the mesogenic group, the siloxane backbone having a much smaller dipole moment per repeat unit. However, for the majority of LC side chain polymers this is not the case, e.g., acrylate and methacrylate polymers have ester dipole moments which relax with the motions of the chain backbone and may give an important contribution to dielectric relaxation in addition to the four relaxation modes associated with the anisotropic motions of the mesogenic group. In such cases, an additional term, isotropic in nature, is added to the equations for £|l (co) and s ± (co) as has been explained by Haase and co-workers. Further words of caution are appropriate at this point. Any analysis of the dielectric data for aligned LC polymers should take into account the Kirkwood g-factors, 011 and g ± [see Eq. (11-25)] for the orientational correlations between mesogenic groups, but this has not yet been done. Furthermore, the molecular theories for dipolar group motion in nematic and smectic A LC phases assume that these groups undergo 180° rotations (or "flip-
11.5 Experimental Data for Polymer Systems
flop" motions) in the local LC potential allowing /xx to reverse, and contributing to the 00 relaxation mode. However, it seems physically unlikely that the dipolar mesogenic groups in LC polymers [see, e.g., structures (l)-(3) above] can achieve such a reorientation since the tail of the alkyl group is attached to the polymer chain and, furthermore, it would be necessary for all mesogenic groups to undergo such motions while preserving the director axis. It seems difficult to imagine that such largescale motions, which would of necessity be co-operative, would not lead to a disalignment of the director and hence to the disalignment of a macroscopically-aligned sample. The precise molecular origin of the different relaxation modes observed in dielectric experiments for LC side chain polymers is still a matter of discussion and, in the writer's view, also raises similar questions for the analogous dielectric relaxation processes observed for low molar mass liquid crystals. We note that for the latter materials large static dielectric per-
513
mittivities may be obtained for the nematic phase and for the different smectic phases and their dielectric relaxations occur at low frequencies (de Jeu, 1980; Bat a and Buka, 1981; Kresse, 1982). The molecular structures of the smectic phases indicate well-ordered arrangements of the molecules, especially for the higherorder smectics (Goodby and Gray, 1984) so it seems unlikely that large scale motions such as "flip-flop" motions can occur. It is possible that angular fluctuations of the mesogenic head groups are responsible for the observed anisotropic dielectric permittivities and relaxations. While such motions would not relax all of the available <^2>, the relaxation strengths would be enhanced if the Kirkwood g-factors were greater than unity. Attard et al. (1987c) showed that the plots of log fm vs. 1/T for the resolved 8 and oc processes for polymer (3) were curved, Fig. 11-16, and each obeyed, approximately, the Vogel equation with (B, To) values of (1800 K, 28 K) and (946 K, 238 K) for the
Figure 11-16. Log [fm (Hz)] against 103 T ^ K ) for LC polymer (3). Unaligned material: observed peak (£<•), resolved 8-peak (o), resolved oc-peak (A, £). Homeotropically-aligned material: observed peak (•, |). (After Attard et al., 1987 c, reproduced with permission.) 2.7
2.9
i 1 3.1 i o 3 r -1 IK)
I
3.3
1
I
3.5
3.7
514
11 Dielectric Properties of Polymers
8 and a processes, respectively. The loci tend to merge as T -> Tg, i.e., at the lowest frequencies 1 0 ~ 2 t o l 0 ~ 4 Hz, which means that the complex anisotropic motions of the mesogenic head groups, which couple to motions of main chain and spacer groups, undergo a critical slowing-down as Tg is approached. For T < Tg the relaxational modes are effectively frozen in this the glassy LC phase. Similar information has been obtained from studies of related siloxane-chain LC polymers. It was shown (Araki and Attard, 1986; Attard and Araki, 1986; Araki et al., 1988) that the following polymer CH3 (CH 2 ) 6 -O-/
(4)
[n~35, Tg = 275K, TC(N - • / ) = 314 K] could be aligned homeotropically (H) or planarly (P) by cooling from the melt in the presence of a strong low frequency or high frequency electric field, respectively. Figure 11-17 shows the loss curves for the three forms of the same sample in the dielectric cell. Several points may be noted. (i) The loss curve for the U-sample is broad and featureless with no indication of a high frequency shoulder, in contrast to that obtained for polymer (3) in its unaligned state. (ii) The 5 process obtained for the Haligned sample is similar in shape to that obtained for the equivalent process in polymer (3). This is assigned to the 00-relaxation mode. (iii) The loss curves cross at a common frequency (~ 900 Hz) which is an isosbestic frequency analogous to the crossover frequency for the real permittivity (Fig. 11-14). Assuming that each loss component has a Fuoss-Kirkwood form, the data for the H-
and P-aligned samples were analyzed and the loss curve for the U-sample was calculated using Eq. (11-53). Excellent agreement between the experimental and calculated curves was obtained (Araki et al., 1988, Fig. 9 therein). Attard et al. (1987d) showed that the complex permittivity for a partiallyaligned LC sample of uniaxial symmetry was given by
+ |(1 -Sd)'8±(co)
(11-54)
where Sd is the macroscopic director order parameter which describes the average orientation of the local directors with respect to the measuring field (z) direction S = |<3cos20nz-1>
(11-55)
S = 0,1 and -(1/2) for unaligned, Haligned and P-aligned samples, respectively. For Sd = 0 Eq. (11-54) reduces to Eq. (11-53). Taking the real part of Eq. (11-54) we see that the condition £|( (co) = s\(co) defines the crossover frequency / c and that the permittivity curves for samples of different alignment — 1/2 < Sd < 1 all cross at fc for a given sample temperature. The crossover frequency is an isosbestic frequency. Similarly the condition s'l'l (co) = e"± (co) defines the crossover frequency / c ' for the loss curves for samples of different alignment [e.g., Fig. 11-17 for polymer (4)]. The existence of isosbestic frequencies have been demonstrated by Attard et al. (1987 a), Araki and Attard (1986), Attard and Araki (1986) for LC polymers, and dielectric measurements with the use of Eq. (11-54) have been shown to be a simple and direct means of determining the extent of macroscopic ordering in a partially-aligned LC polymer. We note that fc depends upon sample temperature in a manner similar to that for log fm for the 5
11.5 Experimental Data for Polymer Systems
515
Figure 11-17. e" • Co plotted against log [/ (Hz)] for a H-aligned, P-aligned and unaligned (U) sample of LC polymer (4) at 309.2 K. (After Araki et al., 1988, reproduced with permission.) log [f (Hz)]
and a processes. The ability to achieve Hand/or P-alignments using directing electric fields of chosen frequencies is determined by the value of the crossover frequency at a given sample temperature. In practice the crossover frequency for a LC polymer may be too high to be accessible experimentally from conventional power sources so that while H-alignment may be obtained, P-alignment may not be obtained in cooling from the melt in the presence of an ac field. Further difficulties in achieving H- or P-alignment stem from (i) dielectric heating due to dielectric dipolar losses or conduction losses at the frequency of the ac directing field, (ii) electrohydrodynamic instabilities and flow of material in the directing ac field, (iii) the high viscosity of the polymeric LC phase which slow-down the alignment process. As a result of (ii), homopolymers (3) and (4) may not be aligned in the LC state by directing ac electric fields. However, they may be aligned by cooling from the melt in the presence of a strong ac electric field. Copolymers having siloxane backbones
(Kozak et al., 1989 a; Williams et al., 1990) and a malonate polymer (Kozak et al., 1989 b) may be aligned in the LC state using ac electric fields. The rate of alignment decreases markedly as the temperature is reduced below Tc (Kozak et al., 1989 a). The dielectric relaxation behavior of a siloxane polymer
(5) and its mixtures with the low molar mass liquid crystal. C 6 H 1 3 O-
(6)
have been described recently by Seiberle et al. (1990). The samples were aligned in a directing magnetic field and s^(co) and 8 ± (co) were determined over a wide range of temperature in the frequency range 102 to 107 Hz. A well-defined 8 process was observed in all the mixtures with no evidence for two peaks despite the fact that
516
11 Dielectric Properties of Polymers
log/m((5) values are ~ 104 different for the pure materials at a given temperature. This demonstrates that each kind of mesogenic group [see structures (5) and (6)] moves, on average, at the same relaxation rate in the mixtures. The a process was also observed and, in accord with the result for the siloxane LC polymer, structure (3) above, the 5 and a processes tend to coalesce on lowering the temperature. The dynamics of realignment of the aligned LC from one orientation to another, 10° distant, was also studied using the dielectric method, and was found to follow the relation As(t)
As(oo)
x
/
= tan
t
(11-56)
where
constants of the LC phase. Similarly, Araki etal. (1989b) used DRS to demonstrate the recovery of H-alignment in a partiallyhomeotropic sample of polymer (4) when it was cooled, in the absence of an electric field from the biphasic region back into the nematic LC state. The theory of alignment recovery, and the loss of its efficiency by repeated cycling, was outlined by Attard (1989) and involves the template recrystallization of LC material on an aligned substrate. Williams and coworkers (1990) have studied the realignment behavior of a copolymer having longitudinally-attached and transverselyattached mesogenic groups in the side chains. In this case, a sample is realigned H -> P using a high frequency field, the field is removed and the sample is found to realign P -> H as judged by the change in the dielectric loss spectrum with time. Such memory effects (Araki etal., 1989b; Attard, 1989; Williams etal., 1990) are readily studied by DRS, in a quantitative manner and are not easily studied using other methods, e.g., polarized spectroscopy. While dielectric studies of aligned LC polymers give more information on component relaxations than those for unaligned polymers, considerable information is obtained from the latter studies, e.g., for polymers of the structure (7)
:OO(CH2)m-C
-COCH
(7)
with (Rl9 m, R2) being (H, 2, CH3), (H, 6, CH 3 ) and (CH 3 , 6, C 4 H 9 ) (Vallerien et al., 1989 a). These authors observed 5, a, p and y dielectric processes in these polymers where (i) the 5 process is assigned to motions of the mesogenic group around the polymer main chain, and (ii) the a process is assigned to the dynamic Tg process (cf. oca
11.5 Experimental Data for Polymer Systems
process) of main chain segments, including the resolved dipole moment component of the ester group attached to the main chain, (iii) The |3 process is assigned to motions of the mesogenic group about its long axis and, (iv) the y process is assigned to motions of the terminal butoxy group since it was only observed in the polymer containing that group. Both p and y processes were only observed for T
Broad-band dielectric relaxation spectroscopy has been used to study the molecular dynamics of combined main-chain and side-chain LC polymers (Kremer et al., 1989). One material studied was structure (8) below:
(CH2)6-O
517
€H 2 .CH.C 2 H 5 CH3
(8) The unaligned materials (nine different materials were studied) all exhibited an a process, which was assigned to the dynamic Tg process, and one or more |3 processes, labelled |3m or |3S and which were assigned to rotational motions of the mesogenic groups in the main chain (m) or side chain (s) about their individual long axes. For polymer (8) the (3m and Ps processes are clearly seen in the plots of e" vs. T(K), with peaks evident at - 250 K and 180 K, respectively, i.e., the local motions of main chain mesogenic groups are, on average, slower than those of the side chain mesogenic groups. A feature of these studies is that there is no apparent discontinuity or change in the dielectric loss peak as the oc process moves to higher frequencies through the transformation range from LC phase to isotropic phase (see Fig. 4 of Kremer et al., 1989). If samples were aligned by electric fields or magnetic fields, as we have discussed above in Sec. 11.5.5.2, then marked changes in the dielectric spectra would be anticipated and a 5 process, presently not observed for their materials, might be detected. The polymers studied by Kremer et al. (1989) have a chiral mesogenic head group in the side chain. The interest in chiral LCforming molecules of low or high molar mass stems from the observation that ferroelectric films may be prepared in thin surface-stabilized LCD cells which are subjected to strong electric fields. The LC materials are aligned macroscopically in this way and being a chiral-smectic phase, fast optical switching may be induced by ap-
518
11 Dielectric Properties of Polymers
plied electric fields, leading to many applications in display technology and optical data storage devices. A full description of the dielectric properties of ferroelectric LC materials is beyond the scope of this account, so reference is made to the following recent accounts by Martinot-Legarde and Durand (1981), Osipov and Piken (1983), Osipov (1984), Beresnev et al. (1984, 1988), Levstik et al. (1987), Carlsson et al. (1988), Filipic et al. (1988), and Gouda et al. (1989). These experimental and theoretical studies demonstrate that ferroelectric LC materials are formed by alignment of a chiral smectic phase (S£, S£) in a "bookshelf" geometry [see Vallerien et al. (1989 c) for an illustration of this arrangement] and that their dielectric properties are characterized by: (i) a large spontaneous polarization for the sample, (ii) large static permittivities and two large low frequency dielectric loss processes due to director motions which are described as the "soft" mode and the "Goldstone" mode. These relaxation modes, which are properties of a macroscopicallyaligned film ( ~ l - 1 0 jum thick), have a complicated, but well-defined dependence on temperature and applied external biasing field. As examples of studies of the dielectric behavior of low molar mass ferroelectric liquid crystals we may refer to the works of Gouda et al. (1989), Biradar et al. (1989 a c), Wrobel et al. (1989), Biradar and Haase (1989, 1990), and Vallerien etal. (1989c). The latter authors studied the material C 8 H 1 7 O-
VooC.CH-CH-C2H5
(9)
Cl CH*
which has two chiral centres (indicated as "*") and exhibits chiral smectic A (S*) and chiral-smectic C (S*) phases (crystal
322KSg328KS£338KI). The ferroelectric film was formed between ITO conducting glass plates (separation, 10 |im) coated with polyimide. Parallel rubbing of the glass surface led to the desired bookshelf geometry. Figure 11-18 shows the dielectric loss data, in the vicinity of the S* <-> S* transition, for the sample subjected to a biasing dc voltage of 5 kVcm" 1 . Two loss processes are observed, the high frequency process being the soft mode, the low frequency process being the Goldstone mode. On lowering temperature towards T(S% - Sg) the relaxation strength of the Goldstone mode increases dramatically. Below the transition the soft mode and Goldstone mode both contribute to the total loss curve. The Goldstone mode may be effectively removed by application of a strong biasing dc field, as is shown in Fig. 11-19. According to theory the biasing field unwinds the helix of the chiral smectic phase [see Fig. 1 in Vallerien et al. (1989 c), for an illustration of this effect] and thus quenches the strength of the Goldstone mode and allows the soft mode to be clearly observed. Figure 11-20 shows how the mean relaxation frequency and reciprocal relaxation strength depend upon temperature for the soft mode as the transition region S*<->S£ is traversed. According to continuum theory the soft mode corresponds to the fluctuations of the tilt angle
11.5 Experimental Data for Polymer Systems
519
20 • 328.9 K • 327.9 K © 326.9 K
16
Soft mode
Figure 11-18. e" plotted against log [/ (Hz)] for a low molar mass ferroelectric LC material, structure (9), at three different temperatures; sample subjected to a dc biasing field of 5 kVcm" 1 . (After Vallerien et al., 1989 c, reproduced with permission.) U 5 log [f (Hz)]
200 160-
\
\ 0
o
120-
o oo
o
o oQ
80-
o
o
40-
o (
0
1
i
1
i
2
4
6
€
E(kV cm" 1
Figure 11-19. e" plotted against log [/(Hz)] for a ferroelectric LC material, structure (9) at 326.4 K; sample subjected to dc biasing electric fields from 0 to 8 kVcm" 1 . The insert shows the field dependence of e" at 2.51 kHz. (After Vallerien et al., 1989 c, reproduced with permission.)
520
11 Dielectric Properties of Polymers 400320(a)
fm (Hz)
240-
160-
Figure 11-20. (a) Mean relaxation frequency vs. temperature for the soft mode and Goldstone mode, for different values of the external biasing field, for the ferroelectric LC material, structure (9). The Goldstone mode (x) is only present for the bias field <5 kVcm" 1 , o, n, • and
8 0 - Goldstone mode \
0.16-
(b)
0.12-
0.08-
0.04-
320
324
I 332
328
336
340
T{K)
product fm • Ae is expected to be approximately independent of temperature on both sides of the transition, and this has been shown to be the case for the mixture "DOBAMBC" (see Gouda etal, 1989, Fig. 3 therein). Also it may be shown that for the soft mode 1 f
(11-58)
where a and b are Landau coefficients and yLC is a rotational viscosity coefficient. Hence yLC may be obtained from the experimental data for fm{T) and l/Ae(T). We have discussed the behavior of low molar mass LC ferroelectric materials in some detail since they form the basis of our understanding of the behavior of poly-
meric LC ferroelectric materials. It is difficult to make ferroelectric films of low molar mass chiral LC materials and it may be more difficult to prepare ferroelectric films of their polymeric analogues. Despite the many difficulties of a chemical and physical nature, films of ferroelectric polymers have been prepared and studied (Vallerien et al., 1989d; Pfeiffer et al., 1990). Vallerien et al. (1989d) have shown that ferroelectric films of the chiral main chain-side chain LC polymer, structure (8) above may be prepared between parallel-rubbed polyimide-coated ITO glass plates (separation 10 jam) to which an electric field of 35 kV cm" 1 was applied for 14 hours in the Sg phase directly below the S% -> Sg transition temperature. The dielectric loss data
11.6 Concluding Remarks
(Fig. 3 in Vallerien et al., 1989d) show that the soft mode is observed in the S*, S* and Sf phases and that the Goldstone mode is observed in the Sg phase. Further dielectric work with ferroelectric LC polymers has been conducted with the following material, (10) (Vallerien et al., 1989e).
l CH3
(10) It was shown that the Goldstone mode could be observed in the aligned S£ phase, and the soft mode in the S£ of this material and that a strong biasing dc electric field decreases the strength of the Goldstone mode by at least a factor of four. In a further study Vallerien et al. (1990) measured the dielectric properties of a ferroelectric material whose structure was similar to (10) except the biphenyl group was replaced by an azoxy group in the side chain [cf. structure (8) above]. In this material the Goldstone mode was sufficiently suppressed to allow a soft mode to be observed in both the S% and S$ phases, and it was also shown that fm for this mode exhibited Curie-Weiss behavior around T(S% — S$) (cf. Fig. 11-20 for a low molar mass LC). However, whereas the range of critical slowing-down of the soft mode is ~ ± 5 K about the transition for the low molar mass LC, it is ~ + 40 K for the ferroelectric polymeric LC. Also, whereas the relaxation strength was strongly-dependent on temperature in the transition region for the low molar mass LC, only a small dependence of this quantity on temperature was observed for the polymeric material. These first studies of ferroelectric LC polymers are of great interest and in addi-
521
tion to the use of such materials in storage and display systems, it is envisaged that new applications may emerge, such as piezomaterials from cross-linked elastomers, pressure sensors and an optical molecular gyroscope, as have been mentioned by Kremer et al. (1990). Very recently, Dumon et al. (1990) showed that aligned siloxane-chain LC copolymers which have a mesogenic group in the sidechain terminated by two chiral centres, as in structure (9) above exhibit ferroelectric behavior. They report the fastest switching times yet reported for ferroelectric LC polymer films (300 JIS using 2 V/jim at 73 °C). It is apparent that such materials, despite their chemical complexity and the difficulties associated with the preparation of well-aligned ferroelectric films, show considerable promise, especially as multifunctional materials which may exhibit different dielectric, piezoelectric, pyroelectric and electro-optical effects [see e.g., Buckley et al. (1990) for accounts of multifunctional polymer materials].
11.6 Concluding Remarks It is apparent from the above account that dielectric relaxation spectroscopy (10" 3 to 107 Hz) provides detailed information on the molecular dynamics and physical structure of polymers in solution and in the bulk amorphous, crystalline and liquid crystalline states. Such information complements that obtained using other spectroscopic, scattering and relaxation methods. The modern instrumentation allows accurate measurements to be made quickly, as has been indicated above and is further illustrated by the work of Liu et al. (1990) for trans- 1,4-polyisoprene in the ranges 100-320 K and 10~2 to 104 Hz. We have not been able to discuss some of the
522
11 Dielectric Properties of Polymers
most recent work in detail, e.g., the dielectric study of Boese et al. (1990) of multiarmed star polymers of bulk ds-l,4-polyisoprene of narrow molecular weight in which both normal mode and segmental mode relaxations were observed (cf. Adachi and Kotacka studies described above) and which provides a test of models for chain motions both below and above the entanglement condition. Similarly, we have not discussed the recent dielectric study of the plasticization of poly(ethylmethacrylate) by the non-polar gas carbon dioxide at pressures up to 60 atm by Kamiya et al. (1990). In this study is was shown that remarkable changes in loss behavior occur as the gas pressure is raised, whose interpretation may relate to the earlier studies of this polymer in which oc, p and (a |3) relaxations were demonstrated (Williams, 1966 b). It is evident from these studies that dielectric relaxation spectroscopy is able to provide important and detailed information on the behavior of novel polymer systems. It is also important to emphasize that modern instrumentation allows dielectric studies to be undertaken which were not possible previously. We have indicated above (Sec. 11.5.4.3) that the crystallization process of amorphous polyethylene terephthalate can be followed by dielectric spectroscopy. In those studies, manual balancing of the transformer bridge, and noting the times of measurement, allowed a cross plot to be made which gave loss factor against frequency at different times during the isothermal crystallization of the sample. Such measurements could now be made automatically thus allowing accurate and comprehensive dielectric data to be obtained. The speed of measurement would allow more comprehensive experiments to be conducted (i.e., at different temperatures of crystallization). Thus, it is envisaged that dielectric studies of the crystallization, in
time, will be a new application of the dielectric method. Similarly, it is expected that the dielectric method will be extensively used to study polymerization and cross-linking reactions as they proceed in real time. For example, it is known that bisphenol-A (DGEBA) epoxy resins exhibit dielectric relaxation properties which depend on the structure and molecular weight of the resin (Sheppard and Senturia, 1989). Mangion and Johari (1990 a, b) have shown that the Tg relaxations and sub-Tg relaxations of bisphenol-A-based thermosets cured with diaminodiphenyl methane or diaminodiphenyl sulfone vary markedly as the materials are cured. In a particular important paper, Mangion and Johari (1990 b) have shown how the permittivity and loss factor vary with time as the curing reaction proceeds. sf(co) decreased from a large value, characteristic of a liquid dipolar medium, to a smaller value (ef ~ 4) as the material transformed from a liquid, through a viscoelastic range to a glassy solid. In parallel, sff(co) decreased from a plateau, through to a minimum and then gave a peak at the time when &'(<JO) exhibits an inflexion. These data show the course of the curing reaction can be monitored in real time using dielectric relaxation spectroscopy. Mangion and Johari note that space-charge losses (MW losses) do not appear in their data for their materials, in contrast to studies of other systems where the short-time behavior of s' and e" is mainly concerned with the MW process (Sheppard and Senturia, 1986). Mangion and Johari show how the curves of s' (co) and s"(co) vs. curing time vary systematically with the measuring frequency / = co/2 71 and they discuss the observed time-temperature variations of s'(co) and e" (co) in terms of the relaxation behavior of a system which changes from a mobile liquid to a glass through a viscoelastic range
11.8 References
523
as the curing reaction proceeds. One obsersimilar to that used for integrated circuit vation is that the complex-plane (Colechips. In these connections polyimides feaCole) diagram of e" (CO) VS. e' (co), where co is ture strongly as a result of their low dielecfixed and s' and s" vary as the reaction tric permittivity and loss, their processiproceeds, has a contour which is well-fitted bility into accurately-defined geometrical using the KWW function with exponent in structures and their excellent thermal and the range 0.3-0.35, i.e., the scaling of the mechanical properties. Again the polymer relaxation behavior of a chemically-changhas a passive role in the final product. ing system with time rather than with freTechnical problems with the polyimides inquency gives behavior of the type found for clude the mismatch between the thermal chemically-stable systems in their glassexpansion of polymer with ceramics or transition regions. These studies show that metals, which affect adhesion of elements dielectric relaxation spectroscopy provides in an assembly, and the permeation of a powerful direct method of following the moisture into the polymer, which affects its curing reaction of a system composed of insulation characteristics. These and other dipolar molecules. aspects of polymers as electrical insulation materials are discussed in the articles by Finally we note that a major use of synTummala, Keyes, Grobman and Kapur in thetic polymers is to provide electrical inTummala and Rymaszewski (1989) and in sulation in electrical cables, electrical and Barfknecht et al. (1989). electronic components (e.g., capacitors), circuit boards and integrated circuits. It is fair to say that synthetic polymers are in11.7 Acknowledgements dispensible in such applications, the principal materials being the different polyethylThe author acknowledges the support of enes, poly(vinyl chloride), poly-(styrene), the Air Force Office of Scientific Research polytetrafluoroethylene and different poly(research programme sponsored by SDIO/ imides. Ultra-low loss polyethylene, for 1ST and managed by AFOSR under Conexample, is used in submarine telephone tract F49620-87-C-0111). The United States cables while polyimides may be used for Government is authorized to reproduce circuit boards. In all such applications the and distribute reprints for governmental role of the polymer is a passive one, the aim purposes notwithstanding any copyright being to provide a material of low dielecnotation hereon. tric loss, of high dielectric strength (for high field situations), of accurately defined dimensions and having good thermal and mechanical properties. The traditional needs of the electrical and electronic industries are well-met by the polymers indicated above. In recent years there has been an increasing demand for polymers to be used in circuit boards of all descriptions and for thin-film packaging of electronic circuits in which the conductors and insulators are fabricated together in a manner
11.8 References Adachi, K., Ishida, Y. (1976), J. Polym. ScL, Polym. Phys. Ed. 14, 2219. Adachi, K., Kotaka, T. (1983), Macromolecules 16, 1936. Adachi, K., Kotaka, T. (1984), Macromolecules 17, 20. Adachi, K., Kotaka, T. (1985), Macromolecules 18, 466. Adachi, K., Kotaka, T. (1988), Macromolecules 21, 157. Adachi, K., Fujihara, I., Ishida, Y (1975), J. Polym. ScL, Polym. Phys. Ed. 13, 2155.
524
11 Dielectric Properties of Polymers
Adachi, K., Hattori, M., Ishida, Y. (1977), / Polym. Sci., Polym. Phys. Ed. 15, 693. Adachi, K., Imanishi, Y, Kotaka, T. (1989a), /. Chem. Soc, Faraday Trans. 1, 85, 1065. Adachi, K., Imanishi, Y, Kotaka, T. (1989b). /. Chem. Soc, Faraday Trans. 1, 85, 1075. Adachi, K., Imanishi, Y, Kotaka, T. (1989c), /. Chem. Soc, Faraday Trans. 1, 85, 1083. Alper, T., Barlow, A. I , Gray, R. W. (1976), Polymer 17, 665. Anderle, K., Birenheide, R., Eich, M., Wendorff, J. H. (1989), Makromol Chem. Rapid Commun. 10, 477. Araki, K., Attard, G. S. (1986), Liq. Cryst. 1, 301. Araki, K., Attard, G. S., Kozak, A., Williams, G., Gray, G. W, Lacey, D., Nestor, G. (1988), J. Chem. Soc. Faraday Trans. II, 84, 1067. Araki, K., Aoshima, M., Namiki, N., Ujiie, S., Koide, N., Iimura, K., Imamura, Y, Williams, G. (1989a), Macromol. Chem. Rapid Commun. 10, 265. Araki, K., Attard, G. S., Williams, G. (1989b), Polymer 30, 432. Ashcraft, C. R., Boyd, R. H. (1976), / Polym. Sci., Polym. Phys. Ed. 14, 2153. Attard, G. S. (1986), Molec Phys. 58, 1087. Attard, G. S. (1989), Polymer 30, 438. Attard, G. S., Araki, K. (1986), Mol. Cryst. Liq. Cryst. 141, 69. Attard, G. S., Williams, G. (1986a), Chem. Britain 22, 919. Attard, G. S., Williams, G. (1986b), Polymer 27, 2. Attard, G. S., Williams, G. (1986c), Polymer 27, 66. Attard, G. S., Williams, G. (1986d), Liq. Cryst. 1, 253. Attard, G. S., Williams, G. (1986e), /. Molec Electronics 2, 107. Attard, G. S., Williams, G., Gray, G. W, Lacey, D., Gemmel, P. A. (1986), Polymer 27, 185. Attard, G. S., Araki, K., Williams, G. (1987a), /. Molec. Electronics 3, 1. Attard, G. S., Moura-Ramos, J. J., Gray, G. W, Lacey, D., White, M., Nestor, G., Toyne, K . I (1987b), Makromol. Chem. RapidPubl. 188, 2769. Attard, G. S., Moura-Ramos, J. J., Williams, G. (1987c), /. Polym. Sci., Polym. Phys. Ed. 25, 1099. Attard, G. S., Araki, K., Williams, G. (1987d), Brit. Polym. J. 19, 119. Attard, G. S., Araki, K., Moura-Ramos, J. I, Williams, G. (1988), Liq. Cryst. 3, 861. Attard, G. S., Araki, K., Moura-Ramos, J. J., Williams, G., Griffin, A. C , Bhatti, A. M., Hung, R. S. L. (1989), in: Polymer Associated Structures, Microemulsions and Liquid Crystals: El Nokaly, M. A. (Ed.). Washington: ACS Symp. Series, No. 384. Barfknecht, A. T., Partridge, J. P., Chen, C. X, Li, C. J. (1989), Advanced Packaging Materials, Mat. Res. Soc Symp. Proc, Vol. 167. Bata, L., Buka, A. (1981), Mol. Cryst. Liq. Cryst. 63, 307.
Baur, M., Stockmayer, W. H. (1965), J. Chem. Phys. 43, 4319. Beevers, M. S., Garrington, D. C , Williams, G. (1977), Polymer 18, 540. Beevers, M. S., Elliott, D. A., Williams, G. (1979), Polymer 20, 787. Beevers, M. S., Elliott, D. A., Williams, G. (1980), Polymer 21, 13. Beresnev, L. A., Pozhidayev, E. P., Blinov, L. M. (1984), Ferroelectrics 59, 321. Beresnev, L. A., Blinov, L. M., Osipov, M. A., Piken, S. A. (1988), Mol. Cryst. Liq. Cryst. 158A, 3. Biradar, A. M., Haase, W. (1989), 18. Freiburger Arbeitstagung Flussigkristalle, Freiburg. Biradar, A. M., Haase, W (1990), Liq. Cryst. 7, 143. Biradar, A. M., Wrobel, S., Haase, W. (1989 a), Ferroelectrics 99, 149. Biradar, A. M., Wrobel, S., Haase, W. (1989 b), Liq. Cryst. 5, 1227. Biradar, A. M., Wrobel, S., Haase, W. (1989c), Phys. Rev. A. 39, 2693. Blumen, A. (1987), in: Molecular Dynamics and Relaxation Phenomena in Glasses, Springer Lect. Notes in Physics, Vol. 277: Dorfmuller, Th., Williams, G. (Eds.). Berlin, Heidelberg: Springer Verlag, p. 1. Boese, D., Kremer, R, Fetters, L. J. (1990), Macromolecules 23, 1826. Bormuth, F. I, Haase, W. (1987), Mol. Cryst. Liq. Cryst. 153, 207. Bormuth, F. X, Haase, W (1988), Liq. Cryst. 3, 881. Bormuth, F. X, Haase, W, Zentel, R. (1987), Mol. Cryst. Liq. Cryst. 148, 1. Bottcher, C. X F., Bordewijk, P. (1978), Theory of Electric Polarization, 2nd. ed., Vol. 2. Amsterdam: Elsevier. Boyd, R. H. (1983), J. Polym. Sci., Polym. Phys., Ed. 21, 505. Boyd, R. H., Porter, C. H. (1972), J. Polym. Sci. A-2, 10, 647. Brather, A. (1979), Coll. Polym. Sci. 257, 785. Buckley, A., Gallagher-Daggitt, G., Katasz, F. E., Ulrich, D. R. (Eds.) (1990), Multifunctional Materials, Mat. Res. Soc Symp. Proc, Vol. 175. Pittsburgh (Pa.): Mat. Res. Soc. Bur, A. X, Roberts, D. E. (1969), J. Chem. Phys. 51, 406. Carlsson, T, Zeks, B., Filipic, C , Levstik, A., Blinc, R. (1988), Mol. Cryst. Liq. Cryst. 163, 11. Chemla, D. S., Zyss, X (1987), Nonlinear Optical Properties of Organic Molecules and Crystals, Vol. 1 and 2. Orlando: Academic Press. Clark, M. G. (1981), Displays, 169. Clark, M. G., Harrison, K. X, Raynes, E. P. (1980), Phys. Technol. 11, 232. Clarkson, T. S., Glasser, L., Tuxworth, R. W, Williams, G. (1977), Adv. Mol. Relax. Interact. Proc 10, 173. Coburn, X C , Boyd, R. H. (1986), Macromolecules 19, 2238.
11.8 References
Cole, K. S., Cole, R. H. (1941), J Chem. Phys. 9, 341. Cole, R. H. (1965), J. Chem. Phys. 42, 637. Cole, R. H. (1977), Ann. Rev. Phys. Chem. 28, 283. Coles, H. J., Simon, R. (1983), U.K. Patent GB2146 787A. Coles, H. X, Simon, R. (1985), Polymer 26, 1801. Cook, M., Watts, D. C , Williams, G. (1970), Trans. Faraday Soc. 66, 2503. Cook, M., Williams, G., Jones, T. T. (1975), Polymer 16, 835. Davidson, D. W, Cole, R. H. (1950), J. Chem. Phys. 18, 1417. Debye, P. W. (1927), Polar Molecules, New York: Chemical Catalog Co. de Jeu, W. H. (1980), Physical Properties of Liquid Crystalline Materials. New York: Gordon and Breach. de Gennes, P. G. (1971), /. Chem. Phys. 55, 572. de Gennes, P. G. (1979), Scaling Concepts in Polymer Physics. Ithaca (N.Y.): Cornell Univ. Press. Doi, M., Edwards, S. F. (1978), J. Chem. Soc, Faraday Trans. 2, 74, 1789, 1802, 1818. Dorfmuller, T., Williams, G. (Eds.) (1987), Molecular Dynamics and Relaxation Phenomena in Glasses, in: Springer Lect. Notes in Physics, Vol. 277. Berlin, Heidelberg: Springer Verlag. Dumon, M., Nguyen, H. T., Mauzac, M., Destrade, C , Achard, M. K, Gasparoux, H. (1990), Macromolecules 23, 355. Dunmur, D. A., Miller, W. H. (1980), Mol. Cryst. Liq. Cryst. 60, 281. Eich, M., Wendorff, J. H. (1987), Makromol Chem. Rapid Commun. 8, 467. Eich, M., Wendorff, J. H., Reck, B., Ringsdorf, H. (1987), Makromol. Chem. Rapid Commun. 8, 59. English, A. D. (1984), Macromolecules 17, 2182. Ferry, J. D. (1970), Viscoelastic Properties of Polymers, 2nd. ed. New York: Wiley. Filipic, C , Carlsson, X, Levstik, A., Zeks, B., Blinc, R., Gouda, F., Lagerwall, S. T., Skarp, K. (1988), Phys. Rev. A. 38, 5833. Flory, P. J. (1969), Configurational Statistics of Polymer Chains. New York: Interscience. Frohlich, H. (1958), Theory of Dielectrics. Oxford: Oxford, U.P. Fuoss, R. M., Kirkwood, J. G. (1941), J. Amer. Chem. Soc. 63, 385. Furukawa, T., Date, M., Nakajima, K., Kosaka, T., Seo, I. (1986), Jap. J. Appl. Phys. 25, 1178. Gilchrist, J. Le G. (1978), J. Polym. Sci., Polym. Phys. Ed. 16, 1773. Glarum, S. H. (1960), /. Chem. Phys. 33, 1371. Goodby, J. W, Gray, G. W. (1984), Smectic Liquid Crystals. Glasgow and London: Leonard Hill. Gouda, F , Andersson, G., Carlsson, T., Lagerwall, S. T., Skarp, K., Stebler, B. (1989), Mol. Cryst. Liq. Cryst., Lett. 6, 151. Graessley, W. W (1974), Adv. Polym. Sci. 16, 1. Haase, W, Pfeiffer, M. (1990), in: Multifunctional Materials: Buckley, A., Gallagher-Daggitt, G.,
525
Karasz, F. E., Ulrich, D. R. (Eds.). Pittsburgh (Pa.): Materials Research Society, p. 257. Haase, W, Pranoto, H. (1985), in: Polymeric Liquid Crystals, Blumstein, A. (Ed.). New York: Plenum, p. 313. Haase, W, Pranoto, H., Bormuth, F. J. (1985), Ber. Bunsenges 89, 1229. Havriliak, S., Negami, S. (1966), /. Polym. Sci. C14, 99. Hedvig, P. (1977), Dielectric Properties of Polymers. Bristol: Hilger. Hill, N., Vaughan, W E., Price, A. H., Davies, M. (1969), Dielectric Properties and Molecular Behaviour. New York: Van Nostrand Reinhold. Hoffman, J. D., Williams, G., Passaglia, E. (1966), /. Polym. Sci. C14, 173. Illers, K. H., Breuer, H. (1963), J Coll. Sci. 18, 1. Ishida, Y. (1969), /. Polymer Sci. 7, 1835. Ishida, Y, Matsuo, M., Takayanagi, M. (1965), / Polym. Sci. B3, 321. Johari, G. P. (1987), in: Molecular Dynamics and Relaxation Phenomena in Glasses, Springer Lect. Notes in Physics, Vol. 277: Dorfmuller, Th., Williams, G. (Eds.). Berlin, Heidelberg: Springer Verlag, p. 90. Johari, G. P., Goldstein, M. (1970), /. Chem. Phys. 53, 2372. Johari, G. P., Monnerie, L. (1986), J. Polym. Sci., Polym. Phys. Ed. 24, 2049. Jones, B., Tidy, D., Williams, G. (1976), J. Phys. E9, 693. Jones, P., Darcy, P., Attard, G. S., Jones, W J., Williams, G. (1989), Mol. Phys. 67, 1053. Jonscher, A. K. (1983), Dielectric Relaxation in Solids. London: Chelsea Dielectrics Press. Kaempf, P. (1987), Polym. J. (Japan) 19, 257. Kakizaki, M., Kakudate, T., Hideshima, T. (1985), /. Polym. Sci., Polym. Phys. Ed. 23, 787, 809. Kamiya, Y, Mizoguchi, K., Naito, Y. (1990), J. Polym. Sci., Polym. Phys. 28, 1955. Kelker, H., Hatz, R. (1980), Handbook of Liquid Crystals. Weinheim: VCH publishers. Kirkwood, J. G., Fuoss, R. M. (1941), /. Chem. Phys. 9, 329. Klein, J. (1978), Macromolecules 11, 852. Kohlrausch, R. (1854), Pogg. Ann. Physik 91, 198. Kozak, A., Williams, G. (1989), Mol. Phys. 67, 1053. Kozak, A., Moura-Ramos, J. X, Simon, G. P., Williams, G. (1989a), Makromol. Chem. 190, 2463. Kozak, A., Simon, G. P., Williams, G. (1989b), Polym. Commun 30, 102. Kremer, F , Vallerien, S. U., Zentel, R., Kapitza, H. (1989), Macromolecules 22, 4040. Kremer, F , Vallerien, S. U., Zentel, R. (1990), Adv. Mat. 2, 145. Kresse, H. (1982), Fortsch. der Physik 30, 507. Kresse, H., Shibaev, V. P. (1983), Z. Phys. Chem. (Leipzig) 264, 161. Kresse, H., Shibaev, V. P. (1984), Makromol. Chem. Rapid Commun. 5, 63.
526
11 Dielectric Properties of Polymers
Kresse, H., Talroze, R. V. (1981), Makromol. Chem. Rapid. Commun. 2, 369. Kresse, H., Kostromin, S., Shibaev, V. P. (1982), Makromol. Chem. Rapid Commun. 3, 509. Kresse, H., Rotz, U., Lindau, J., Kuschel, F. (1989), Makromol. Chem. 190, 2953. Laupretre, K, Noel, C , Jenkins, W. N., Williams, G. (1985), J. Chem. Soc, Faraday Discuss. 79, 191. Levstik, A., Carlsson, T., Filipic, I., Levstik, I., Zeks, B. (1987), Phys. Rev. A. 35, 3527. Lindsey, C. P., Patterson, G. D. (1980), J. Chem. Phys. 73, 3348. Liu, F., Boyd, R. H., Woodward, A. E., Zemel, I. S. (1990), Macromolecules 23, 219. Luckhurst, G., Zannoni, C. (1975), Proc. Roy. Soc. A 343, 380. Maier, W, Meier, G. (1961), Z. Naturforsch. 16a, 262. Mangion, M. B. M., Johari, G. P. (1990a), J. Polym. Sci., Polym. Phys., Ed. 28, 71. Mangion, M. B. M, Johari, G. P. (1990b), J. Polym. Sci., Polym. Phys., Ed. 28, 1621. Mansfield, M., Boyd, R. H. (1978), /. Polym. Set, Polym. Phys. Ed. 16, 1227. Martinot-Legarde, Ph., Durand, G. (1981), J. Phys. 42, 269. Martins, A. E, Esnault, P., Volino, E (1986), Phys. Rev. Lett. 57, 1745. Mashimo, S., Nakamura, H., Chiba, A. (1982a), /. Chem. Phys. 76, 6342. Mashimo, S., Nozaki, R., Yagihara, S., Takeishi, S. (1982 b), J. Chem. Phys. 77, 6259. Mashimo, S., Winsor, P., Cole, R. H., Matsuo, K., Stockmayer, W H. (1983), Macromolecules 16, 965. Mashimo, S., Yagihara, S., Chika, A. (1984), Macromolecules 17, 630. Matsuoka, S., Roe, R. X, Cole, H. E (1972), in: Dielectric Properties of Polymers: Karasz, F. E. (Ed.). New York: Plenum Press, p. 255. Matsuoka, S., Williams, G., Johnson, G. E., Anderson, E. W, Furukawa, T. (1985), Macromolecules 18, 2652. McArdle, C. B. (Ed.) (1989), Side Chain Liquid Crystal Polymers. Glasgow: Blackie. McArdle, C. B., Clark, M. G., Haws, C. M., Wiltshire, M. C. K., Parker, A., Nestor, G., Gray, G. W, Lacey, D., Toyne, K. J. (1987), Liq. Cryst. 2, 573. McCrum, N. G., Read, B. E., Williams, G. (1967), Anelastic and Dielectric Effects in Polymeric Solids. London and New York: Wiley. Dover Ed. (1991), New York: Dover Publ. McKenna, G. B. (1989), in: Comprehensive Polymer Science: Allen, G., Bevington, J. C. (Eds.), Vol. 2: Booth, C , Price, C. (Eds.). Oxford: Pergamon Press, Ch. 10, p. 311. Mopsik, E I. (1984), Rev. Sci. Instrum. 55, 79. Moscicki, J. K., Williams, G. (1983 a), J. Polymer Sci., Polym. Phys. Ed. 21, 197.
Moscicki, J. K., Williams, G. (1983 b), /. Polymer Sci., Polym. Phys. Ed. 21, 213. Moscicki, J. K., Williams, G., Aharoni, S. M. (1981), Polymer 22, 571, 1361. Moscicki, J. K., Williams, G., Aharoni, S. M. (1982), Macromolecules 15, 642. Moynihan, C. T., Macedo, P. B., Montrose, C. X, Gupta, P. K., de Bolt, M. A., Dill, X E, Dom, B. E., Drake, P. W, Easteal, A. X, Elterman, P. B., Moeller, R. P., Sasabe, H., Wilder, X A. (1976), Ann. N.Y. Acad. Sci. Vol. 279, 15. Ngai, K. L., Rendell, R. W (1987), Macromolecules 20, 1066. Ngai, K. L., Rajagopal, A. K., Rendell, R. W (1986), IEEE Trans. Electrical Insul. EI-21, 313. Ngai, K. L., Rendell, R. W, Rajagopal, A. K., Teitler, S. (1987a), Ann. N.Y. Acad. Sci. 484, 150. Ngai, K. L., Wang, C. H., Fytas, G., Plazek, D. L., Plazek, D. T. (1987b), J. Chem. Phys. 86, 4788. Nordio, P. L., Rigatti, G., Segre, U. (1973), Mol. Phys. 25, 129. Osipov, M. A. (1984), Ferroelectrics 58, 305. Osipov, M. A., Piken, S. A. (1983), Mol. Cryst. Liq. Cryst. 103, 57. Pethrick, R., Richards, R. W (Eds.) (1982), Dynamic Properties of Solid Polymers, NATO ASI. Dordrecht: Reidel. Pfeiffer, M., Scherowsky, G., Beresnev, L. A., Kuhnpast, K., Hannischfeger, R., Haase, W (1990), Adv. Mat., submitted, private communication. Pranoto, H., Haase, W (1983), Mol. Cryst. Liq. Cryst. 98, 299. Pranoto, H., Haase, W, Finkelmann, H. Kiechle, U. (1984), 14. Freiburger Arbeitstagung Flussigkristalle Vortragsreferat 12. Freiburg: p. 1. Pranoto, H., Bormuth, X, Haase, W, Kiechle, U., Finkelmann, H. (1986), Makromol. Chem. 187, 2453. Read, B. E. (1965), Trans. Faraday Soc. 61, 2140. Reddish, W. (1950), Trans. Faraday Soc. 46, 459. Rendell, R. W, Ngai, K. L. (1985), in: Relaxations in Complex Systems. Washington (D.C.): U.S. Govt. Printing Office, p. 309. Rendell, R. W, Ngai, K. L., Fong, G. R., Aklonis, X J. (1987), Macromolecules 20, 1070. Ribelles, X L. G., Calleja, R. D. (1985), /. Polym. Sci., Polym. Phys. Ed. 23, 1505. Sasabe, H., Saito, S. (1968), /. Polym. Sci. A-2 6, 1401. Sayre, X A., Swanson, S. R., Boyd, R. H. (1978), J. Polym. Sci. Polym. Phys. Ed. 16, 1739. Schmidt, H. W. (1989), Angew. Chem. Int. Ed. Engl. Adv. Mater. 28, 940. Scott, A. H., Scheiber, D. X, Curtis, A. X, Lauritzen, XL, Hoffman, X D. (1962), /. Res. Natl. Bur. Stand. Sect. A 66, 269. Seiberle, H., Stille, W, Strobl, G. (1990), Macromolecules 23, 2008. Sheppard, N. E, Senturia, S. D. (1986), Adv. Polym. Sci. 80, 1.
11.8 References
Sheppard, N. R, Senturia, S. D. (1989), J. Polym. ScL, Polym. Phys., Ed. 27, 753. Shibaev, V. P., Kostromin, S. G., Plate, N. A., Ivanov, S. A., Petrov, V. Y, Yakolev, I. A. (1983), Polym. Commun. 24, 364. Shlesinger, M. R, Montroll, E. W. (1984), Proc. Nat. Acad. ScL 81, 1280. Simon, R., Coles, H. J. (1986), Liq. Cryst. 1, 281. Smyth, C. P. (1955), Dielectric Behaviour and Structure. New York: McGraw-Hill. Stockmayer, W. H. (1967), Pure Appl. Chem. 15, 539. Sullivan, D. E., Deutch, J. M. (1975), J. Chem. Phys. 62, 2130. Tidy, D., Williams, G. (1978), unpublished. Titulaer, U. M., Deutch, J. M. (1974), /. Chem. Phys. 60, 1502. Tummala, R. R., Rymaszewski, E. T. (Eds.) (1989), Microelectronics Packaging Handbook. New York: Van Nostrand Reinhold. Vallerien, S. U., Kremer, R, Boeffel, C. (1989a), Liquid Crystals 4, 79. Vallerien, S. U., Kremer, R, Hiiser, B., Spiess, H. W. (1989 b), Colloid Polym. Sci. 267, 583. Vallerien, S. U., Kremer, R, Kapitza, H., Zentel, R., Frank, W. (1989 c), Phys. Lett. A 138, 219. Vallerien, S. U., Zentel, R., Kremer, R, Kapitza, H., Fischer, E. W. (1989d), Makromol. Chem. Rapid. Commun. 10, 333. Vallerien, S. U., Zentel, R., Kremer, R, Kapitza, H., Fischer, E. W. (1989e), Proc. 2nd Ferroelectric Liquid Cryst. Conf. Goteborg: Ferroelectrics, in press. Vallerien, S. U., Kremer, R, Kapitza, H., Zentel, R., Fischer, E. W. (1990), Proc. 7th. Int. Mtg. Ferroelectricity (IMF.7). Saarbriicken (1989): Ferroelectrics, in press, personal communication. Van Turnhout, J. (1975), Thermally Stimulated Discharge of Polymer Electrets. Amsterdam: Elsevier. Volkenstein, M. V (1963), Configurational Statistics of Polymeric Chains. New York: Interscience. Von Hippel, A. R. (1954), Dielectric Materials and Applications. New York, John Wiley. Wada, A. (1977), in: Dielectric and Related Molecular Processes, Spec. Period Report, Vol. 3: Davies, M. (Ed.). London: The Chemical Society, p. 143. Wang, C. C , Pecora, R. (1980), J. Chem. Phys. 72, 5333. Warchol, M. P., Vaughan, W. (1978), Adv. Mol Relax. Proc. 13, 317. Wetton, R. E., Williams, G. (1965), Trans. Faraday Soc. 61, 2132. Williams, D. J. (Ed.) (1983 a), Nonlinear Optical Properties of Organic and Polymeric Materials, ACS Symp. Series, No. 233. Washington (DC): Am. Chem. Soc. Williams, G. (1963), Trans. Faraday Soc. 59, 1397. Williams, G. (1964 a), Trans. Faraday Soc. 60, 1548. Williams, G. (1964 b), Trans. Faraday Soc. 60, 1556. Williams, G. (1966a), Trans. Faraday Soc. 62, 1321. Williams, G. (1966 b), Trans. Faraday Soc. 62, 2091. Williams, G. (1972), Chem. Rev. 72, 55.
527
Williams, G. (1975), in: Dielectric and Related Molecular Processes, Spec. Period Reports, Vol. 2: Davies, M. (Ed.). London: The Chemical Society, p. 151. Williams, G. (1978), Chem. Soc. Rev. 7, 89. Williams, G. (1979), Adv. Polym. Sci. 33, 60. Williams, G. (1982 a), in: Dynamic Properties of Solid Polymers, NATO ASI: Pethrick, R., Richards, R. W. (Eds.). Dordrecht: Reidel. Williams, G. (1982b), IEEE Trans. Electr. Insul. EI-17, 469. Williams, G. (1983b), J. Polym. ScL, Polym. Phys. Ed. 21, 2037. Williams, G. (1985), IEEE Transactions on Electrical Insulation El-20, 843. Williams, G. (1989), in: Comprehensive Polymer Science: Allen, G., Bevington, J. C. (Eds.); Vol. 2: Booth, C , Price, C. (Eds.). Oxford: Pergamon Press, Chap. 18, p. 601. Williams, G., Crossley, J. (1977), Ann. Report Chem. Soc. (Phys. Chem. A), p. 77. Williams, G., Hains, P. J. (1971), Chem. Phys. Lett. 10, 585. Williams, G., Hains, P. J. (1972), /. Chem. Soc, Faraday Symp. 6, 14. Williams, G., Watts, D. C. (1970), Trans. Faraday Soc. 66, 80. Williams, G., Watts, D. C. (1971 a), in: Nuclear Magnetic Resonance, Basic Principles and Progress, Vol. 4, NMR of Polymers. Berlin, Heidelberg: Springer Verlag, p. 271. Williams, G., Watts, D. C. (1971b), in: Dielectric Properties of Polymers: Karasz, R E. (Ed.). New York: Plenum, p. 17. Williams, G., Watts, D. C. (1971c), Trans. Faraday Soc. 67, 1971. Williams, G., Watts, D. C , Dev, S. B., North, A. M. (1971), Trans. Faraday Soc. 67, 1323. Williams, G., Cook, M., Hains, P. J. (1972), J. Chem. Soc, Faraday Trans. II 68, 1045. Williams, G., Nazemi, A., Karasz, R E . (1990), in: Multifunctional Materials: Buckley, A., Gallagher-Daggitt, G., Karasz, R E., Ulrich, D. R. (Eds.). Pittsburgh (PA): Materials Research Society, p. 227. Wrobel, S., Biradar, A. M., Haase, W (1989), Ferroelectrics 100, 271. Yano, O., Saiki, K., Tarucha, S., Wada, Y, (1977), /. Polym. ScL, Polym. Phys. Ed. 15, 43. Yemni, T, Boyd, R. H. (1979), J. Polym. Sci. Polym. Phys. Ed. 17, 741. Yu, H., Bur, A. X, Fetters, L. J. (1966), /. Chem. Phys. 4, 2568. Zentel, R., Strobl, G. R., Ringsdorf, H. (1985), Macromolecules 18, 960.
528
11 Dielectric Properties of Polymers
General Reading
Dorfmuller, T., Williams, G. (Eds.) (1987), Molecular Dynamics and Relaxation Phenomena in Glasses, in: Springer Led. Notes in Physics, Vol. 277. Berlin, Heidelberg: Springer Verlag.
McCmm, N. G., Read, B. E., Williams, G. (1991), Anelastic and Dielectric Effects in Polymeric Solids. New York: Dover. Williams, G. (1979), Adv. Polym. Sci. 33, 60. Williams, G. (1989), in: Comprehensive Polymer Science: Allen, G., Bevington, J. C. (Eds.); Vol.2: Booth, C , Price, C. (Eds.). Oxford: Pergamon Press, Chap. 18.
12 Optical Properties of Polymers Wolfgang Knoll
Frontier Research Program, The Institute of Physical and Chemical Research (RIKEN), Wako, Saitama, Japan and Max-Planck-Institut fur Polymerforschung, Mainz, Federal Republic of Germany
List of Symbols and Abbreviations 12.1 Introduction 12.2 Theoretical Concepts for the Description of the Interaction of Light with Matter 12.2.1 Wave Propagation in Free Space 12.2.2 Polarization of Light 12.2.3 Maxwell's Equations in a Medium 12.2.4 Linear Dielectric Media 122 A A Linear, Nondispersive, Homogeneous, and Isotropic Media 12.2.4.2 Inhomogeneous and Anisotropic Media 12.2.4.3 Absorption 12.2.4.4 Dispersion 12.2.4.5 Dielectrics with Resonances 12.2.5 Nonlinear Optics 12.2.5.1 Second-Order Nonlinear Optics 12.2.5.2 Third-Order Nonlinear Optics 12.3 Polymers with Specifically Tuned Optical Properties 12.3.1 Polymers as Transparent Substrates for Optical Discs 12.3.2 Solid Polyelectrolytes as Glassy Polymers with Tunable Refractive Indices and Dispersion Characteristics 12.3.3 Magneto-Optical Polymers 12.4 Polymers in Special Configurations 12.4.1 Ultrathin Films 12.4.1.1 Langmuir Monolayers 12.4.1.2 Langmuir-Blodgett-Kuhn Multilayer Assemblies 12.4.1.3 Self-Assembly Mono- and Multilayers on Solid Supports 12.4.2 Integrated Optics 12.4.2.1 Planar Waveguide Structures 12.4.2.2 Channel Waveguides 12.4.2.3 Polymer Optical Fibers (POFs) 12.4.3 Polymers for Optical Recording 12.4.4 Polymers for Molecular Opto-Electronics Materials Science and Technology Copyright © WILEY-VCH Verlag GmbH & Co KGaA. Allrightsreserved.
531 535 536 536 537 538 538 539 539 540 540 541 542 543 544 545 545 546 547 549 549 550 553 560 561 562 569 570 572 575
530
12.5 12.5.1 12.5.2 12.5.2.1 12.5.2.2 12.5.2.3 12.5.3 12.5.3.1 12.5.3.2 12.5.3.3 12.5.4 12.6 12.7 12.8
12 Optical Properties of Polymers
Nonlinear Optical Properties of Polymers General Concepts for the Functionalization of Polymers for Nonlinear Optics Second-Order Nonlinear Optics Control of Orientational Order Second-Harmonic Generation (SHG) Electro-Optic (EO) Modulation Third-Order Nonlinear Optics Third-Harmonic Generation (THG) Degenerate Four Wave Mixing (DFWM) All-Optical Switching Photorefractive Polymers Outlook Acknowledgements References
576 576 581 581 585 587 588 590 591 592 592 593 593 593
List of Symbols and Abbreviations
List of Symbols and Abbreviations A A, B c Co
D d d0 e E E(0) Eo Ely
f F FB
FR G H Ho h I k k~ k0 k0 kiz
*fP k° k°
^sp /
L Lc Ld m m! M n N ND nD, nF, n c
area per molecule amplitude of Eiy in medium i magnetic induction speed of light in a medium speed of light in free space electric displacement film thickness thickness per monolayer magnitude of electron charge electric field static electric field electric field amplitude y-component of E in medium i local field factor force electric force restoring force grating vector magnetic field magnetic field amplitude Planck constant intensity wavenumber complex wavenumber wavenumber in free space wavevector in free space z-component of k in medium i fc^p including a shift caused by thin dielectric coating parallel component of photon wavevector wavevector of plasmon surface polariton at metal-air interface waveguide propagation vector component penetration depth; sample length device length coherence length conjugation length mass of electron; order of waveguide mode order of second harmonic mode magnetization density refractive index number of charges per unit volume density of (active) dye molecules refractive indices at three standard wavelengths (blue, yellow, red) refractive index of medium i
531
532
12 Optical Properties of Polymers
np nz n0 n2 H|l? n± An p P PNL r R rp rs R, 5 t T Tg Tm u V JC X x y
refractive index of prism refractive index of a multilayer normal to its plane background refractive index optical Kerr coefficient refractive index parallel, perpendicular to film plane refractive index change pressure polarization density nonlinear component of the polarizability spatial coordinate (position vector) reflectivity reflectivity coefficient for p-polarized light reflectivity coefficient for s-polarized light chain segments, i = 1, 2, 3, 4 variable (Kramers-Kronig relations) time absolute temperature glass transition temperature figure of merit variable representing Ex y z or Hx y z volume; Verdet constant displacement vector functional group unit vector in the x-direction unit vector in the y-direction
a a0 a± /? PD P2 y yD 7xxxx y0 Th F{ A g s 8tj s0 3 6
absorption (extinction) coefficient; exponential factor extinction coefficient prior to poling extinction coefficient normal to film after poling propagation constant molecular second-order nonlinear optical polarizability nonlinear absorption coefficient surface tension molecular third-order nonlinear optical polarizability c o m p o n e n t of y D surface tension of w a t e r - a i r interface homogeneous linewidth inhomogeneous linewidth ellipsometric phase angle electric permittivity of a medium complex electric permittivity of a medium electric permittivity tensor element (second rank) electric permittivity of free space distribution angle incident angle
List of Symbols and Abbreviations
K
X A
K
x0
Av 77 T
z z Xu
i z" z(2) Xo
00
CO
ATR a.u. BAM CCD CD DFWM DMPE DRAW DUV EO
critical angle for total internal reflection resonance angle resonance angle including a shift caused by thin dielectric coating elastic constant of a spring wavelength of light groove spacing laser light wavelength wavelength in free space spectral position of linear absorption magnetic permeability frequency of light Abbe number central frequency spectral width or line width lateral surface pressure switching time Faraday rotation angle electric susceptibility complex electric susceptibility electric susceptibility tensor element (second rank) real part of (complex) susceptibility imaginary part of (complex) susceptibility coefficient of second-order optical nonlinearity coefficient of third-order optical nonlinearity macroscopic averaged nonlinear susceptibility susceptibility constant in dielectrics with resonances phase phase factor (internal) resonance angle order parameter nonlinear phase shift ellipsometnc amplitude angle photon frequency frequency of laser light attenuated total internal reflection arbitrary unit Brewster angle microscopy charge-coupled device compact disc degenerate four wave mixing dimyristoylphosphatidylethanolamine direct read after writing deep ultraviolet electro-optic(al)
533
534
12 Optical Properties of Polymers
esu GI IOC IR LBK LC NIR NLO OWM PC PM-co-OLG PPA PPV ROM s, p SAM SH(G) SI SPM THG UV
electrostatic unit graded index integrated optical circuit infrared Langmuir-Blodgett-Kuhn liquid crystalline near infrared nonlinear optics optical waveguide microscopy poly(carbonate) poly(y-methyl-L-glutamate-co-y-octadecyl-L-glutamate) poly(phenylacetylene) poly(p-phenylenevinylene) read-only memory polarization of light (s: perpendicular, p: parallel) self-assembly monolayer second-harmonic (generation) step index surface plasmon microscopy third-harmonic generation ultraviolet
12.1 Introduction
12.1 Introduction Without polymers and their optical properties our daily life certainly would be by far less colorful: just think about the pictures in a magazine, all the paint in and around our buildings, or our clothing (be it a Japanese kimono or a western T-shirt). However, in most of these cases it is not the polymer itself that shows the desired coloring effect through the interaction with light but rather some pigments and dyes that are dispersed into the material, more or less loosely connected to the macromolecular matrix. These obvious optical properties are not within the scope of this contribution. We will rather concentrate on the intrinsic optical properties of polymers, that is, focus on the presentation of those phenomena and properties of the materials that describe their response to visible light. A rough classification of these properties could start, for example, with transparency and birefringence, which determine the intensity decrease and change of polarization, respectively, of light being just transmitted through the sample. The response to external fields is also an intrinsic property, for example, the change of the optical properties by applying a (strong) magnetic field is described by the magneto-optical effect. Some polymers are designed so as to strongly interact with light through specific electronic excitations, giving rise to a broad range of phenomena: spectral changes due to the resonance character of absorption, or secondary processes (following the excitation) like luminescence phenomena (fluorescence, phosphorescence, etc.) are examples. Many systems show changes in other properties when photoexcited, for example, become photoconductive. The general term photoresponsive was coined to summarize the nearly unlim-
535
ited list of property changes that can be induced by a purposeful design of the material, typically by the proper functionalization: this list includes photoisomerization and photochemical reactions resulting in a change of optical, mechanical, structural, morphological, rheological, or (meso-)phase properties. Finally, nonlinear optical properties, relevant for the description of the response of the material to ultrastrong optical fields, that is, to a giant laser pulse, define the most active field of research in recent years related to optical properties of polymers. In what follows, we will be able to cover only a few selected examples to try to highlight some of the improvements that have been achieved by extensive research and development that has mostly been focused on two nearly opposite objectives: one goal was to minimize the interaction with light, that is, to reduce the losses in a polymer optical fiber (POF) used in (local) data communication networks or to decrease the noise for optical recording and reading in information storage on optical discs. The other strong effort was to enhance the interaction with light, for example, to increase the light intensity-dependent refractive index change n2 in selected polymers needed to switch "light by light" in a future photonics device application. This limitation to some examples from the above list also means that many other aspects of the interaction of light with polymers will be ignored. These include, in particular, the interaction with light of wavelengths other than in the visible [X = 400 to 800 nm) with the only exception to this being some considerations related to the near infrared (NIR) losses of POFs. No IR vibrational spectroscopy will be considered. Another area important at least for diagnostic purposes in polymer research is the scattering of light (either
536
12 Optical Properties of Polymers
elastic, quasi-elastic, or inelastic, i.e., Brillouin- or Raman-spectroscopies) that we have to ignore. (But see Chap. 14 and Chap. 7 of this Volume.) And finally, all versions of optical labeling techniques, including absorption or fluorescence labels in steady state or time-resolved experimental configurations will not be covered. However, for some of the optical phenomena and properties described we will also give some consideration to the experimental techniques that had to be developed in order to be able to quantitatively determine the relevant material parameters. Finally, some of the discussed improvements of materials properties need to be seen in the context of their present and future application in optical device configurations. Here, the emerging technology of the 21st century, photonics, is a major driving force for research and development in polymer materials science and technology. This chapter is organized as follows: In Sec. 12.2 we give a brief summary of the basic theoretical concepts, mostly based on the classical electrodynamic description of the interaction of light with matter given by Maxwell's theory, also including a short introduction to nonlinear optics. Section 12.3 concentrates on some "passive" materials properties related, for example, to its transparency or its dispersion properties. Section 12.4 is called "polymers in special configurations" and describes polymers in ultrathin films from monomolecular layers to multilayer assemblies, integrated optics with planar and channel waveguide structures and fibers, and optical recording principles. Section 12.5 finally is devoted to nonlinear materials and device concepts based on second (x(2)) and third (#(3)) order nonlinear optical effects.
12.2 Theoretical Concepts for the Description of the Interaction of Light with Matter Light is an electromagnetic wave phenomenon which is described by the same theoretical concepts that govern all forms of electromagnetic radiation. Figure 12-1 displays the (unlimited) spectrum of electromagnetic waves. Optical frequencies occupy a band of frequencies, v, and wavelengths, k, from the infrared (v = 3x 1011 Hz, l = lmm) to the ultraviolet (v = 3 x 1016 Hz, X = 10 nm) with the visible spectrum being limited to the narrow range between X = 400 and 800 nm. The appropriate theory, which describes the interaction of light with matter and hence is also the background for the understanding of the optical properties of polymers, is based on Maxwell's equations. We will give in the following only a very brief outline of the concepts that are needed to understand some of these properties and to see the specific requirements of the material for some functional applications discussed in this contribution. 12.2.1 Wave Propagation in Free Space We start with the basic principles describing the propagation of light in free space (Born and Wolf, 1980). These are summarized in the following partial differential equations, known as Maxwell's equations: V x H(r, t) = 80
\xE(r,t) =-
dE(r,t) dt dt
(12-1) (12-2)
V-E(r9t) =0
(12-3)
V-H(r,t) =0
(12-4)
with the electric permittivity e0 « 1/(36 TC) x l 0 ~ 9 and the magnetic permeability
12.2 Theoretical Concepts for the Description of the Interaction of Light with Matter
and
Frequency / Hz 1018
1015
109
1Q12
1Q6
X
1mm
1m
(12-9)
1Q3
CD
1nm
537
1km Wavelength
Figure 12-1. The electromagnetic spectrum and the range of "light" (after Saleh and Teich, 1991).
k0 is called the wavevector with its magnitude the wavenumber k0 = 2 n/l0 = 2n v/c0. From Maxwell's equation it follows that k0, E and H are mutually orthogonal. Since E and H lie in a plane normal to the propagation direction given by k0, the wave is called a transverse electromagnetic (TEM) wave. 12.2.2 Polarization of Light
fi0 = 4TC x 10 7 (in MKS units, in SI units: 80 = 8.8542 x 1(T 12 As V" 1 m" 1 , fi0 = 1.2566 x 10 ~6 Vs A ^ m " 1 ) of free space. Vx and V- are the usual curl and divergence operations, respectively. E(r, t) is the electric field and H(r,t) is the magnetic field that describe the electromagnetic wave as a function of position r and time t. A direct consequence of Maxwell's equations, (12-l)-(12-4), is the so-called wave equation r2
(12-5)
c)t2
where u represents any of the three components (£ x , E y , Ez) of E or the three components (ifx, ify, Hz) of H, V2 is the Laplace operator given by 8 8x
8 2• + •
' 8y
6 2
' 8z2
(12-6)
and
« 3xlO 8 m/s
(12-7)
is the speed of light in vacuum. One particular solution to the above wave equation is a monochromatic, electromagnetic wave whose electric and magnetic field components are plane waves: E{r,t) =
(12-8)
Consider a monochromatic plane wave of frequency v travelling in the z-direction with velocity c0. The electric field lies in the x-y plane and can be described by (12-10)
with ^ x = ^ x O COS
(12-11)
and Ey = Ey0 cos
(12-12)
x and j> are unit vectors along the x- and y-directions, Ex and Ey are the £-field components in x- and y-directions, respectively, and cpx and cpy take into account that the two orthogonal partial waves, in general, can have different phases. If the components also differ in £ x 0 and E y0 , one is dealing with elliptically polarized light because at a fixed value of z, the tip of the electric field vector E rotates periodically in the x-y plane, tracing out an ellipse. The shape of this ellipse depends on the ratio of the two amplitudes Ex0/Ey0 and on the phase difference (p = cpx — cpy. This is schematically depicted in Fig. 12-2 a. If one of the components vanishes (Ex0 = 0, for example) the light is linearly polarized in the direction of the other component (the y-direction). The wave is also
538
12 Optical Properties of Polymers
This wave is called right circularly polarized. For cp = — n/2 the rotation is counter clockwise and the light is said to be left circularly polarized. These two cases are sketched in Fig. 12-2c. 12.2.3 Maxwell's Equations in a Medium
In a dielectric medium with no free charges or currents Maxwell's equations read —^
V xE(r,t)
(c) right
\4
8 / > M ) J± \l , I)
r! I /////// K\ ! /\! z
WVJWl y/'
(12-13)
dt
dB(r,t)
(12-14)
dt
V •D(r,t) = 0
(12-15)
V •B(r,t) = 0
(12-16)
The electric displacement D (r, t) is related to the electric field by
left
f \>¥\/\i\!Pr\!\: n
where P(r,t) is the polarization density of the dielectric medium. Similarly, the magnetic induction B(r, t) is related to the magnetic field by , t) -
Figure 12-2. Trajectories of the endpoint of the electric-field vector at a fixed position z = const, (right side) and a snapshot at a fixed time t = const, (left side) for an elliptically polarized wave (a), a linearly polarized wave (b), and for left and right ciruclarly polarized light (c) (after Saleh and Teich, 1991).
r, t)
M(r9
(12-18)
with M(r,t) being the magnetization density. Since we have to deal only with nonmagnetic materials Eq. (12-18) is replaced by (12-19)
linearly polarized if the phase difference
12.2.4 Linear Dielectric Media
The nature of the dielectric medium is exhibited in the relation between the polarization density P(r, t) and the electric field E(r,t\ called the medium equation. This relation then also determines the optical properties of polymers.
12.2 Theoretical Concepts for the Description of the Interaction of Light with Matter
12.2.4.1 Linear, Nondispersive, Homogeneous, and Isotropic Media
with a speed of light in the medium c =
Let us first recall a few definitions (Saleh and Teich, 1991). A medium is said to be linear if the polarization density P(r,t) is linearly related to the electric field E(r,i). Then the principle of superposition holds. The medium is said to be nondispersive if its response is instantaneous. Clearly, this is an idealization, because any system, no matter how fast it may be, has a finite response time. It is said to be homogeneous if the relation between P and E is independent of the position r. The medium is called isotropic if the relation between P and E is independent of the direction of E, that is, the material looks the same from all directions. Then, P and E must be parallel. And finally a medium is said to be spatially nondispersive if P at each position is induced only by E at the same position. For such a medium the relation between P and E is given by p(v A — p v F(v i\ A If , Li — OQ /
JL* It , Li
(\ ?-7(W \±Z*Z*\Jj
a
where x is scalar constant called the electric susceptibility. Substituting Eq. (12-20) into Eq. (12-17) gives /)(r, t) = eo(l +x) E(r, t)
(12-22)
Hence it follows that also D and E are parallel and proportional. The constant is the electric permittivity of the medium. The ratio s/e0 = (1 +x) *s the dielectric constant. The wave equation (12-5) then is changed to
y2u
-^w=°
The ratio of the speed of light in free space to that in the medium is called the refractive index n: (12-24)
n=—
and is related to the dielectric constant by
« = (—Y=(l+Z)*
(12-25)
12.2.4.2 Inhomogeneous and Anisotropic Media In an inhomogeneous dielectric, for example, a graded index polymer (cf. Sec. 12.4..2.3), which otherwise is linear, nondispersive, and isotropic, the relations P = soxE and D = &E remain valid but the proportionality constants x a n d s a r e now functions of the position x = x(r) a n d £ = £ (r). Likewise, the index of refraction n = n(r) is position dependent. For materials with slowly varying inhomogeneities (on the length scale of the wavelength of light) one can show that the wave equation still holds, however, with a locally varying speed of light c (r): d2E
\2E-
=0
(12-26)
(12-21)
which we write as D(r,t) = eE(r,t)
539
(12 23)
"
where c(r) = co/n(r). A material with a certain profile of the index of refraction, therefore, has a correspondingly varying speed of light. In an anisotropic medium the relation between the vectors P and E depends on the direction of E and the two vectors are not necessarily parallel. If the medium is linear, homogeneous, and nondispersive, each component of P is a linear combination of the three components of E p _ y^ F
y
F
(\ 2-21)
540
12 Optical Properties of Polymers
where ij = x, y, z. Thus, the polarization response of an anisotropic material to an electric field is described by the secondrank susceptibility tensor Xij- This is schematically characterized in Fig. 12-3. Similarly D and E are related by the electric permittivity tensor stj n — vP U^
— 2^ ^ij j
l
(12-31)
(\i ?R^
F *-*\
\VL-LQ))
By a suitable rotation of the coordinate system one can always make the off-diagonal elements to go to zero, a situation which then defines the principal axes of the medium. 12.2.4.3 Absorption Dielectric materials that absorb light can be phenomenologically represented by a complex susceptibility X = X' +
A plane wave propagating in such a medium in the z-direction is described by a complex amplitude proportional to Q~^Z. Since £ is complex it is useful to write k in terms of its real and imaginary parts
'n"
(12-29)
corresponding to a complex permittivity £ = £O(1+X). The wavenumber then also becomes complex: (12-30)
Then the intensity of the wave is seen to be attenuated by the factor |e^ z | 2 = e~ az so that the coefficient a represents the absorption coefficient (or extinction coefficient). /? is the propagation constant P = n-k0
(12-32)
with n the effective refractive index of the medium. Substituting Eq. (12-32) into Eqs. (12-31) and (12-30) we obtain an equation which relates the refractive index n and the absorption coefficient a to the real and imaginary parts of the susceptibility / and x", n —l
2fc0
'
*
•
A. ,
For weakly absorbing media (/, obtain
(12-33) <^1) we (12-34)
- Kf
(12-35)
12.2.4.4 Dispersion
Figure 12-3. For an anisotropic medium the relation between the E-field and the polarization P is characterized by nine elements of the susceptibility tensor
Dispersive media are characterized by a frequency-dependent (and therefore wavelength-dependent) susceptibility x(v), refractive index n(v), and speed of light c(v). This phenomenon, well-known from inorganic materials like glasses, can be either the basic concept for the function of an optical component, for example, the desired dispersion of white light into its different colors by a prism, or can be the source of a limited performance, for example, the chromatic aberration of a lens.
541
12.2 Theoretical Concepts for the Description of the Interaction of Light with Matter (a)
(b)
blue
red
Figure 12-4. The dispersion of light in a dielectric medium with a wavelength dependent refractive index (a) is responsible for the separation of the light's "colors" by a prism (b), but also for chromatic aberration of a lens (c) (after Saleh and Teich, 1991).
This is schematically shown in Fig. 12-4. A quantitative measure of the dispersion of the material is the Abbe number vD, defined by Vn =
(12-36)
- nc)
with nD, nF, and nc being the refractive indices at three standard wavelengths (blue: A = 486.1nm, yellow: 589.2 nm, and red: 656.3 nm, respectively). A direct consequence of the relation between the polarization density P(t) and the applied electric field E(i) given by theory for a linear system are the so-called Kramers-Kronig relations:
n I s2 - v2
ds
(12-37)
niv2-s2
ds
(12-38)
A practical consequence of these equations which relate the real and the imaginary part of the susceptibility and hence relate absorption and dispersion is that once the absorption spectrum of a given material is known (measured) over a sufficiently broad wavelength range (including all relevant excitations) the index of refraction of the material can be calculated.
12.2.4.5 Dielectrics with Resonances
An instructive way to visualize the optical properties of a medium near an electronic transition (excitation) is to consider it as an assembly of forced harmonic oscillators according to the model due to Lorentz (Zernicke and Midwinter, 1973). The electrons are treated as charges, e, attached to the nuclei by a spring of elastic constant K. In an isotropic approximation the force FE exerted by the electric field on the electron FE = eE
(12-39)
then induces a displacement x from equilibrium position which in turn induces a restoring force FK on the electron by the spring FR=
(12-40)
-KX
The resulting force produces an acceleration of the mass m of the electron and the equation of motion reads — F — OITV
)2 x —
—
(\ 2-41 ^
where 2nv0 = KJm is the resonance frequency of the (free) oscillator electronnucleus. Introducing a damping term, — 2 7i Av • dxjdt, then gives the familiar differential equation describing the forced,
542
12 Optical Properties of Polymers
a (v)
•n 0
Av
damped harmonic oscillator: — ^ + 2n Av - ^ + (2TT V 0 )22JCJC== — dt d m ' (12-42) From the electron displacement x results a polarization density P = N - e - x, N is the number of charges per unit volume. A monochromatic wave with an £-field oscillating with frequency v,E = Eo ei27CVt, satisfies Eq. (12-42) if =
(2nvo)2soxoE
(12-43)
with 2
eN mso(2nvo)2 Comparing Eq. (12-43) with Eq. (12-20) and Eq. (12-29) then gives the real and imaginary parts of x (v), v2o(v2o-v2) (12-45) V AV
= ~Xo
Figure 12-5. Absorption coefficient a (v) and refractive index n (v) of a dielectric medium (background index n0) near an electronic resonance (after Saleh and Teich, 1991).
This frequency dependence is illustrated in Fig. 12-5. Typical dielectric media contain multiple resonances. The overall susceptibility is the sum of contributions from these resonances. An example for a material that exhibits resonance absorptions in the ultraviolet and in the infrared is given in Fig. 12-6. Shown are the absorption coefficient and the refractive index as a function of the wavelength. 12.2.5 Nonlinear Optics In nonlinear media the relation between P and E is nonlinear (Shen, 1984). However, Maxwell's equations still can be used to derive a nonlinear partial differential equation that describes the behavior of .Visible
(12-46)
If these resonant oscillators are sufficiently dilute in a host medium of (background) refractive index n09 then the overall refractive index and absorption coefficient are 2nc
«(v)«-('^V(v) noco
(12-47) (12-48)
0.01
0.1
10
100
X//xm
Figure 12-6. Typical wavelength dependence of the absorption coefficient a and the refractive index n of a material with several resonant transitions (after Saleh and Teich, 1991).
543
12.2 Theoretical Concepts for the Description of the Interaction of Light with Matter
electromagnetic waves in nonlinear media. For a homogeneous and isotropic medium one obtains d2E
d2P
(12-49)
It is convenient to write P as the sum of linear and nonlinear parts as they may be obtained by an expression in a Taylor series about E=0: (12-50)
der terms are needed to describe the polarization induced by an E-field. One has to bear in mind, however, that these deviations are typically extremely small and it was not until the invention of the laser in 1960 that optical field intensities could be generated that could compete with interatomic electric fields (typically 105 to 108 V/m) so that the experimental investigation of nonlinear optical effects could begin.
with (12-51)
12.2.5.1 Second-Order Nonlinear Optics
(12-52)
We examine briefly the optical properties of a nonlinear material that exhibits a second-order nonlinear susceptibility
Equation (12-49) then becomes
V2E l 2 v h 2 dt ~ c
dt2
which is a wave equation in which the term 92 P L — /x0 ^ acts as a (nonlinear) source of radiation from the material. Equation (1252) is the basic equation that underlies the theory of nonlinear optics. Figure 12-7 may illustrate graphically the situation. Figure 12-7 a gives the polarization for a material responding to an electric field in a linear way: doubling the £-field doubles the polarization. Figure 127 b shows the situation (highly exaggerated) for a nonlinear material. Higher or-
(a)
(b)
]
\E\2
(12-53)
If a sinusoidal E-field E(z,t) = E o COS(2TTV£ — kz) induces a nonlinear polarization in such a medium we find
= y Xi2) E2
COS(4TIV£- 2kz)]
=
(12-54) This process is illustrated graphically in Fig. 12-8 a. PNL(0) describes the generation of a DC voltage, a process called optical rectification. The term PNL(2 v) is the source for radiation at twice the input frequency, a process called second-harmonic generation (SHG). If we apply an optical field E(v) together with a DC (or low frequency AC) electric field E (0) then the sum of the two fields E = E(0) + Eo cos(2nvt-
kz)
(12-55)
induces a polarization Figure 12-7. The relation between the polarization P and the electric field E for a linear (a) and a nonlinear medium (b).
E0 cos(2nvt-
kz)]2 (12-56)
544
12 Optical Properties of Polymers
dotted line. The polarization at frequency v
(a)
PNLW
PNL(0)
= 2s0X(2> £(0) Eo coS(2nvt-
kz) (12-57)
can be rewritten in the form PNL(v) = e 0 A Z E(v)
(12-58)
where AX = 2X(2)E(0)
(12-59)
represents a change in the susceptibility proportional to the electric field E(0). The nonlinear nature of the medium creates a coupling between the electric field £(0) and the optical field E(v), causing one to control the other. This effect is called the linear electro-optic effect (Pockels effect).
time t
(b)
12.2.5.2 Third-Order Nonlinear Optics PNi_(t)
In media with centrosymmetry the second-order term in the expansion of PNL is absent. The dominant nonlinearity is then of third order E(v)
time t
Figure 12-8. (a) Scheme demonstrating the effect of a sinusoidal input field exciting a nonlinear optical medium: a DC-polarization PNL (0) ("optical rectification") and a polarization component at twice the input frequency (giving rise to second-harmonic generation) are obtained, (b) Linearization of the secondorder relation |P NL | = s0x(2)\E\2 in the presence of a strong electric field E (0) and a weak optical field E (v). (After Saleh and Teich, 1991.)
which contains components at frequencies 0, v, and 2 v. If the optical field is substantially smaller than the DC field \E(v)\2 <^ \E(0)\2, the 2 v-polarization component can be neglected. This is equivalent to the linearization of PNL as a function of E at E = E(0) as depicted in Fig. 12-8b, dash-
ID I _ p v (3) I J7|3 I NLI — 0 /C I I
(IJ-fiCW v W)
and the material is called a Kerr medium. A monochromatic optical wave at frequency v induces a nonlinear polarization which contains contributions at the frequencies v and 3v. The latter process is called third-harmonic generation (THG). The polarization component at v corresponds to an incremental change of the susceptibility A% at frequency v given by A* = 4
# = TX ( 3 ) |£(V)I 2
(12-61)
This change in susceptibility is equivalent to a change in refractive index An. Both are proportional to the square of the field amplitude. The overall refractive index is, therefore, a linear function of the optical intensity /, n{l) = n + n2l
(12-62)
12.3 Polymers with Specifically Tuned Optical Properties
where v<3)
(12-63)
This effect, called the optical Kerr effect, is a self-induced effect in which the phase velocity of the wave depends on the wave's own intensity. This intensity-dependent refractive index change n2 is the basis for a broad range of nonlinear optical phenomena observed also in polymers, the most important being the degenerate four wave mixing. This form of real-time holography is particularly helpful for the determination of ultrafast nonlinear response times needed for future photonic device configurations. This technique will be discussed in Sec. 12.5. Other phenomena based on the optical Kerr effect are, for example, self-phase modulation, self-focusing, optical phase conjugation, and spatial solitons.
12.3 Polymers with Specifically Tuned Optical Properties Following the outline given in the introduction we will concentrate in this section on the discussion of the remarkable progress that has been made in recent years in the development of polymeric materials with a specific, tailor-made property profile. Depending on the various applications, these synthesis- and processing-optimizations were aimed at reducing as much as possible any interaction with light so as not to change its intensity, state of polarization, etc. Clearly, these are the applications where polymers are to replace inorganic glasses as windows, shields, eye glasses, substrates etc. A more recent example, the use of polymers as substrate materials for optical storage discs, will be briefly discussed below. A second rapidly growing area for the use of highly transpar-
545
ent polymers as polymer optical fibers will be presented in Sec. 12A.23. Another important optical property of polymers is their dispersion behavior: The isotropic or anisotropic refractive index and its dependence on the wavelength of light are key parameters that need to be tunable to certain values in all areas of polymers for integrated optics. As an example, solid polyelectrolytes with a broad range of refractive index values and dispersion characteristics are discussed below. Next, a short introduction to some recent developments of magneto-optical applications of polymers is given. Probably the most important application of polymers as resist materials in microelectronics is treated in Vol. 18 of this Series. These resist materials are optimized with respect to their absorption of photons (and, of course, with respect to other property changes such as solubility following the excitation by the exposure to light). 12.3.1 Polymers as Transparent Substrates for Optical Discs
The availability of low cost lasers has made optical information storage an attractive alternative to existing mass storage concepts. The largest volume use of polymers for optical recording, however, is still in the disc substrates. The material used for compact discs (CD) and CDROMs (read-only memory) and video discs has to fulfill extreme requirements (Kampf, 1985): • optical homogeneity (low birefringence); • high transparency in the visible, in some cases also in the UV (2 = 280-380 nm); • high shape stability (low thermal expansion, water uptake less than 0.3%); • isotropic expansion upon temperature or humidity increase; • good mechanical properties;
546
12 Optical Properties of Polymers
• constant thickness, planarity; • constant rheological properties for the molding process of the "pits" (cf. Fig. 12-41). Currently, mostly poly(methylmethacrylate) (PMMA) and poly(4-methyl-l-pentene) as well as bisdiallylpolycarbonates (PCs) are used for the fabrication of substrate discs. These polymers offer the optical quality of glass. Some selected physical properties of PMMA and PC are summarized in Table 12-1 (after Pearson, 1988). 12.3.2 Solid Polyelectrolytes as Glassy Polymers with Tunable Refractive Indices and Dispersion Characteristics
The various available polymers can have very different refractive indices and dispersion properties. However, Fig. 12-9 shows that these materials exhibit typically much lower refractive indices than inorganic glasses due to the absence of highly polarizable units like, for example, the heavy metals with their many electrons in flint
s
"D
1.80
s
/
1.60-
p^
uo-.
^ ^
^
m
\
organic polymers
A 80
60
Physical property
PC
PMMA
Transmittance (%) at 830 nm Refractive index at 830 nm Tg (°C) Thermal expansion coefficient
90 1.58 150 6-7
92-93 1.49 100 7.6
Rockwell hardness (Mscale) Izod impact strength (kg cm/cm) H 2 O absorption (%)
75 1-2 0.25
90 1.6 0.54
glasses. One extreme case, for example, is poly(tetrafluorethylene) with nD = 1.345 (data point A in Fig. 12-9). A material with a substantially higher refractive index such as poly(iV-vinylcarbazole) with nD = 1.675 has a very high dispersion vD = 19 (data point B in Fig. 12-9) [note that the higher vD the lower the dispersion, cf. Eq. (12-36)]. One successful attempt to extend the available polymeric materials towards higher refractive indices was recently reported for solid polyelectrolytes, also called ionenes (Simmrock et al., 1989). Given their general structure [R1-N + (R3)2-R2-N+(R4)2]X2-
/
s
- inorganic glasses
Table 12-1. Selected physical properties of PMMA and PC substrates (after Pearson, 1988).
20
Figure 12-9. Refractive indices nD and optical dispersion, vD, of inorganic glasses (dotted area), organic polymers (dashed area) and a series of selected ionenes (numbered I through VII according to Table 12-2). A: poly(tetrafluorethylene), B: poly(JV-vinylcarbazole).
with the main chain segments Rx and R2, and R3 and R4 being organic moieties, for example aliphatic groups, there is an almost unlimited range for tuning their solid state properties, for example, their crystallinity or their ability to solidify as glasses, while on the other hand their optical properties can be largely varied by the introduction of suitable counterions X 2 ~, both monovalent and divalent. In particular, by the introduction of structural irregularity into the main chain (R± / R2) glassy materials with very different polymeric properties, e.g. the glass transition temperature, can be synthesized. Table 12-2 indicates the progress in extending the selec-
"•^
CO
oo
oo O
m
o
1.62*
m
L6U l.58f
Q
vo
in
1.591
0
5.0 9.7 1.0 7.6 4.6
12.3 Polymers with Specifically Tuned Optical Properties
o- vo (N
PQ 3
PH
PQ
-a PH
PQ
PQ
3 3sX
><
PQ
N
X E U
m
u
u
CJ u U
u u u u
3
a?
G
u
X
X
u u
o o
str cture
a?
2 tn
u1)
X
m
ci
a?
_o
73 o
ON
o
\
X
X)
6
jo
73 73
5"
o
o
cJT
o
o
o
OT
y^
CJ
(N
X
X
uu u
u
XI X PC
u u11 u 11
u
X
U
1
X
1
1
ii 5C
11
uuu
'OT
X
1
efi]
X
I(N
D
stru
cycl
y,
eye]
6 6
own
d
CJ
a?
I
IN
X X
X
u u u
u
oC
547
tion of polymers towards higher refractive index with a wide range of Abbe numbers (cf. Fig. 12-9). The interrelation between high refractive index and high dispersion is demonstrated in Fig. 12-10 for two different ionenes (Mathy et al., 1991). The first one, polyelectrolyte I (see Table 12-2), with ZnBr4 as counterions has its strongest absorption in the DUV range not exceeding l = 220nm (see Fig. 12-10 a). The Kramers-Kronig analysis [cf. Eqs. (12-37) and (12-38)] with a background refractive index of no = 1.516 shows that as a result the dispersion in the visible is very low despite a relatively high refractive index (Fig. 12-10 b). On the other hand, the introduction of Hgl 4 ions into the polyelectrolyte VI extends the absorption of the material into the near UV range (Fig. 12-10 c). This causes a consequently much higher (resonance enhanced) refractive index but also a very strong wavelength dependence, that is, low Abbe number (Fig. 12-10d). It is clear that this flexibility for tuning the linear optical properties of polymeric materials is a very promising feature for future device configurations, for example, in integrated optics. We should also mention that the introduced counterions, of course, may also have high nonlinear polarizabilities, allowing the preparation of highly functionalized materials, for example, for x(3)-applications (Meyer et al., 1991).
bi)
i
i
i
OH
i
i
on
ss i
PQ
(N
IN
i
E o
1
nding
-Do.lC
Ion
s
I
PQ
-6.10-6.Do-6.10- e-Znl
I
PQ
PH
Do.Pi -Me-' -Do.Pi -Me-
-o G N u
e-Cd] e-Cd]
PQ
^*'
-6.10-
S-i
PQ
w
o
i
i
o o o o o o o
OT OH
O OT
|
d
H-H
^" ^>
^ >
12.3.3 Magneto-Optical Polymers Certain materials act naturally as polarization rotators, a property known as optical activity: waves with right- and left-circular polarization (cf. Fig. 12-2) propagate with different velocities. This property is found in media in which the molecular units have an inherently helical character
548
12 Optical Properties of Polymers
1.70
1.62
180
260
340
420
1.54
250
350
450
550
650
X/nm
X/nm
n
A
1.9-
1.8-
1.7-
(d)
Figure 12-10. Absorption spectra of ionene I (cf. Table 12-2) (a), and VI (c) and the corresponding Kramers-Kronig analysis (b), and (d), respectively, with a background refractive index, n0 = 1.515.
\
A
V
1.6-
180
260
3Z.0 Z.20 X/nm
250
350 £50 550 X/nm
(quartz, selenium, tellurium, but also many organic/polymeric materials). Materials that have different absorption coefficients for right- and left-circularly polarized light are said to exhibit circular dichroism. An effect related to this, the Faraday effect, describes the polarization rotation of light passing through a material placed in a static magnetic field. The Faraday rotation angle £ is given by Z = V-B-l
(12-64)
where B is the magnetic induction in the direction of the light wave propagation, / is the length of the sample, and V is a material parameter, called the Verdet constant. This constant can be relatively high in polymers containing phenyl groups [e.g., V « 0.1 min/(Oe • cm) (« 7.5 s/A) for poly(styrene), Muto and Ito (1992)]. In a Faraday rotator the sense of the rotation is governed by the direction of the magnetic
650
field and, therefore, does not change with the reversal of the direction of propagation of the wave. This is the basic concept for optical isolators based on this phenomenon. Its principle is outlined in Fig. 12-11: A linearly polarized wave passing the Faraday rotator may experience a polarization rotation of, say, 45°. It passes another polarizer B tuned to this angle. After reflection it can again go through the polarizer, but is then rotated by the rotator by another 45° in the same direction. This wave is now orthogonal to the incident polarizer A and hence is blocked. Muto and coworkers (1991) have used a poly(amethylstyrene) rod of I = 19 mm to construct a 45° rotator for A = 488 nm light at 4 kOe (« 3.2 x 105 A/m) magnetic field. Its transmission characteristic is shown in Fig. 12-12. Inserting the isolator into the laser beam causes a loss of about 5 dB, but the isolation obtained is approximately
12.4 Polymers in Special Configurations
549
A 45°
(a)
Polarizer B
y
Faraday rotator
Polarizer A
Figure 12-11. An optical isolator based on the Faraday effect transmits light in one direction (a) but blocks it in the opposite direction (b) (after Saleh and Teich, 1991).
15 dB. Further improvements of the device performance are expected for materials with lower birefringence.
12.4 Polymers in Special Configurations The foregoing discussion was focused on polymers where only the bulk optical properties were of interest, specifically tailored for selected applications. In this section we concentrate on optical properties of polymers prepared in a very special configuration, where this configuration and the fabrication protocol are an integral aspect of their optical properties. 12.4.1 Ultrathin Films Ultrathin polymer films, in the extreme case only one monomolecular layer thick, are of special interest from an applied point of view for all kinds of surface modifica-
tions, be it for lubrication, isolation, adhesion, or any other possible purpose, and provide a special challenge for the experimentalist who has to develop methods that are sensitive enough to allow for an optical characterization of coatings only a few
insertion loss
-90 -Lb 0 LS 90 Faraday rotation angle/deg
Figure 12-12. Transmitted light intensity through a polymeric Faraday rotator measured as a function of the analyzer angle (polarizer B in Fig. 12-11) (after Muto etal., 1991).
550
12 Optical Properties of Polymers
nanometers thick. In the following we will summarize just a few of the techniques developed in recent years for that purpose concentrating on a few special types of monolayers: Langmuir films prepared at the water-air interface, Langmuir-Blodgett-Kuhn multilayer assemblies prepared by the multiple deposition of monolayers onto solid substrates, and self-assembled monolayers obtained by controlled adsorption from a solution to a solid support. 12.4.1.1 Langmuir Monolayers
All surface-active molecules like soaps, lipids, etc. can be prepared as monomolecular layers at a water-air interface. Driven by the reduction of the surface free energy of water these molecules spread when applied to the surface, for example, from a volatile solvent (Fig. 12-13). The physical properties of these monolayers were first investigated in the 1940s by I. Langmuir. A so-called Langmuir trough (cf. also Fig. 12-14) filled with water defines an exactly known area for the spread molecules (Fig.
12-13). At low lateral density these molecules behave like a quasi-two-dimensional gas. If the area for the molecules is reduced by a movable barrier, this lateral compression will eventually lead to a measurable lateral pressure 77 (force F per unit length of barrier). It can be measured by a socalled Wilhelmy balance (Kuhn et al., 1972), which is the difference between the surface tension of the free, y0, and the layer-covered water surface, y: =
yo-y
(12-65)
If the lateral pressure is plotted as a function of the area per molecule A (cf. Fig. 12-13), one obtains a U-A isotherm, which is the equivalent of the three-dimensional pV diagrams of real gases: Here, too, any change in the compressibility indicates changes in the 2D phase behavior of such a monolayer at a given temperature. What is sketched for a low molecular mass amphiphile in Fig. 12-13 was extended in recent years also to the characterization of polymeric amphiphiles.
area/(molecule • nm 2 )
Figure 12-13. Pressure-area (77 — A) isotherm for a hypothetical amphiphile organized as a monomolecular layer at the water-air interface showing a phase transition from a fluid-expanded to a solid-condensed state upon compression. The insets show cross-sectional diagrams of the Langmuir trough with the compressed monolayer in different phase states.
12.4 Polymers in Special Configurations
It is clear that for the investigation of the optical properties of these highly organized polymer films extremely sensitive techniques are necessary. Ellipsometry has proven to be also applicable for the characterization of monolayers at a water-air interface. Figure 12-14 shows schematically the set-up used in a so-called compensating configuration: laser light is passed through a polarizer and a compensator to become elliptically polarized. After the reflection from the water surface typically at an angle of incidence close to the Brewster angle for maximum sensitivity, the then linearly polarized light passes through an analyzer at orthogonal polarization and the minimum intensity is detected (null-ellipsometry). Polarizer and analyzer angles at compensation can then be converted to the two ellipsometric angles A and V which describe the complex ratio of the reflectivities for s- and p-polarized light, that is, for light with its plane of polarization being perpendicular (s) and parallel (p) to the plane of incidence, respectively, = eiA tan
551
Wilhelmy-system
Teflon-trough
barrier
subphase
Figure 12-14. Ellipsometric set-up integrated into a Langmuir trough.
186 -
(12-66)
One difficulty associated with ellipsometric measurements of polymer monolayers at the water surface is the limited sensitivity of V. This is exemplified in Fig. 12-15, where we have plotted A and W for increasing monolayer thicknesses calculated with different refractive indices as indicated. Given the experimental accuracy of the ellipsometric angle determination of about 0.005°, it becomes evident that for monolayer thicknesses below approximately 2 nm only A can be reliably measured. This means then, however, that no independent determination of the (geometrical) thickness of the polymer layer and its refractive index (not even in an isotropic approximation) can be obtained.
5.08 r| M M | I i i i | I i i i | i i i i | i i i I | I I
0
0.5
1.0
1.5
2.0
2.5
thickness / nm
Figure 12-15. A and "Pasa function of n1, dx for a monolayer at the air-water interface, calculated with A = 594nm, 0 = 50.00°, no = 1.333. Different curves correspond to different values of n1 as indicated in the figure. Above the Brewster angle, 6 « 53°, an increase in A indicates a thicker layer; below the Brewster angle, it is the reverse.
552
12 Optical Properties of Polymers
Nevertheless, even by recording only A as a function of the molecular area (ellipsometric isotherm) helpful information about polymer monolayers can be obtained. One example measured with poly(y-methylL - glutamate -co-y- octadecyl - L - glutamate) (PM-co-OLG) is given in Fig. 12-16. The n~A isotherms during compression and decompression are compared with the A-A isotherms from ellipsometry. One can see, for example, that the first break in the compressibility at approximately 0.20 nm 2 per repeat unit is caused by a change in the intermolecular interaction whereas the overall film thickness still increases monotonically. On the other hand, the plateau region beyond the second break in the TI-A isotherm seen for areas per repeat unit smaller than about 0.14 nm 2 is not associated with a collapse of the monolayer everywhere on the trough. It is rather an instability just in front of the barrier: Only if the material piled up by this "snowplow" effect reaches the laser spot of the ellipsometric measurement do the A -values increase strongly, indicating the increasing film thickness. These ellipsometer configurations typically average monolayer properties over approximately 1 mm2. In some cases, however, for example during first-order phase transitions with coexisting domains, higher lateral resolution of the optical data is needed. Two microscopic techniques were recently introduced that give qualitative information about optical properties on the jLim2 scale. One is the Brewster angle microscope (BAM), which operates in a reflection mode at the Brewster angle (of the pure water surface) with p-polarized light (Honig and Mobius, 1991; Henon and Meunier, 1991). The dark image of the surface changes its intensity when covered by a monolayer (Fig. 12-17). In this way, details of the lateral structure formation of
0.05
0.10
0.15
0.20
0.25
area / repeat unit (nm2)
Figure 12-16. II — A diagram and simultaneously measured \A — AU2O\ isotherm of PM-co-OLG during compression and decompression. Measurements were carried out at T = 20 °C and I = 633 nm at an angle of incidence of 0 = 54.07°.
-V >
*
i
«t
++ •••
4#+:++*++,+.*• Figure 12-17. Brewster angle microscopic image of an azobenzene-derivatized liquid-crystalline Langmuir monolayer at the water-air interface.
12.4 Polymers in Special Configurations
polymer monolayers during compression can be resolved that were previously observable only by fluorescence microscopy after adding a suitable dye. Clearly, the BAM has the advantage that no label is necessary. In passing we note that also the ellipsometric set-up can be converted easily into a microscope that allows one to take pictures of laterally heterogeneous monolayers. Figure 12-18 shows an example obtained from a phospholipid monolayer in the coexistence region of fluid and condensed domains. By changing the polarizer settings contrast inversion can be induced and analyzed. This gives information about the optical thickness of the different phases just as ellipsometry does but with a high lateral resolution (Reiter et al., 1992).
553
Figure 12-18. (a) Image of domains of the phospholipid DMPE at the air-water interface, compressed at a temperature of 20 °C into the coexistence region. The dark parts correspond to the crystalline analogous domains and the bright ones to the fluid matrix, (b) Contrast inversion by changing the extinction settings with respect to the fluid matrix.
12.4.1.2 Langmuir-Blodgett-Kuhn Multilayer Assemblies The above-mentioned monolayers can be transferred to a solid support by dipping and withdrawing a suitable substrate through this highly organized film at the water-air interface. Thus, multilayer assemblies with a precise control of their thickness can be obtained, as is schematically sketched in Fig. 12-19. First reported by K. Blodgett (1935), this technique has regained worldwide interest as an excellent tool for the build-up of supramolecular architectures (Roberts, 1990) with tailormade molecules ever since the pioneering work of H. Kuhn in the early 1970s. These systems are therefore called LangmuirBlodgett-Kuhn (LBK) layers. Various techniques have been developed in recent years sensitive enough to allow a quantitative evaluation of the optical properties of these ultrathin coatings (Swalen, 1986). One group of techniques uses evanescent waves as highly specific
Langmuir
—
Blodgett
—
Kuhn
Figure 12-19. Schematic representation of the Langmuir-Blodgett-Kuhn technique: The Langmuir monolayer prepared at the water-air interface at a selected surface pressure is transferred during the withdrawal of the solid substrate. Supramolecular architectures can be built up by choosing different functional units in the individual layers.
sensitive interfacial light probes (Knoll, 1991 a, b). The simplest case for the existence of an evanescent wave is the well-known total internal reflection of a plane electromagnetic wave at the base of a glass prism (index of refraction nx) in contact with an optically less dense medium (with n2
12 Optical Properties of Polymers (b)
etector
Glass Prism
Detector
Glass Prism •Metal
Evanescent Field
0
Figure 12-20. Two configurations for evanescent wave optics, (a) Top: total internal reflection of a plane wave at the base of a glass prism; bottom: the reflectivity R recorded by a detector as a function of the angle of incidence shows the increase to unity at 6C, the critical angle for total reflection, (b) ATR setup for the excitation of surface plasmons (PSPs) in Kretschmann geometry. Top: a thin metal film (d « 50 nm) is evaporated onto the base of the prism and acts as a resonator driven by the photon field; bottom: the resonant excitation of the PSP wave is seen in the reflectivity curve as a sharp dip at coupling angle 60.
proaches the critical angle, 0C, for total reflection (Fig. 12-20 a, bottom). Closer inspection of the E-field distribution in the immediate vicinity of the interface shows that above 9C the light intensity does not fall abruptly to zero in air, but there is instead a harmonic wave traveling parallel to the surface with an amplitude decaying exponentially normal to the surface. The depth of penetration / defined by the 1/e attenuation is given by \
...
.
(12-67)
and is found to be on the order of the wavelength of light. This type of wave is called an evanescent wave.
One way of introducing surface plasmons [plasmon surface polaritons or PSPs (Burstein et al., 1974; Raether, 1977)], especially in the experimental set-up called the Kretschmann configuration (Kretschmann, 1972) (see Fig. 12-20b, top), is to note that the nearly free electron gas in the thin (^50nm) metal film evaporated onto the base of the prism acts as an oscillator that can be driven by the electromagnetic wave impinging upon that interface. Therefore, we are dealing with the resonant excitation of a coupled state between the plasma oscillations and the photons, that is, the "plasmon surface polaritons" (Raether, 1988). This resonance phenomenon can be clearly seen in the ATR scan (attenuated
12.4 Polymers in Special Configurations
total reflection; see Fig. 12-20b, bottom): Below 6C the reflectivity is very high because the metal film acts as a mirror with little transmission. Above 6C for total internal reflection, however, a relatively narrow dip in the reflectivity curve at 90 indicates the resonant excitation of such a PSP wave at the metal-air interface. The coupling angle is given by the energy and momentum matching condition between photons
PSP U
PSP1
555
and surface plasmons: CO
{
.
= nx — sm< c
(12-68)
with fc°p being the magnitude of the PSP wavevector, k°h the parallel component of the photon wavevector, co the photon energy, c the speed of light, and (p0 the (internal) coupling angle (see Fig. 12-20b). Again, we are dealing with an evanescent wave propagating along the interface with a penetration depth into air of the order of the wavelength. The resonance character of this excitation gives rise to an enhancement of the £-field at the interface by more than a factor of 10, which is the origin of the remarkable sensitivity enhancements obtainable, for example, in Raman spectroscopy when working with PSP light (Ushioda and Sasaki, 1983; Knobloch etal., 1989). Since we are dealing with well-defined modes obeying a known dispersion relation, co versus fc°p, each photon of energy h coL allows the excitation of exactly one PSP mode. This is schematically depicted in Fig. 12-21 a. The solid curve represents the dispersion of the surface plasmons at a
8i
50 nm Ag or Au Langmuir-Blodgett-Kuhn Multilayer Assemblies
Figure 12-21. (a) Dispersion relation, co vs. ksp, of plasmon surface polaritons at an Ag-air interface (PSP0, solid line) and at an Ag-dielectric coating-air interface (PSP1, broken line). Laser light of energy hwL couples at angles 60 and 61, respectively, given by the energy and momentum matching condition (see the intersection of the horizontal line at coL with the two dispersion curves), (b) Schematic of the experimental set-up for surface plasmon spectroscopy.
556
12 Optical Properties of Polymers
(a) Hydrocarbon Polyglutamate Hydrocarbon Silver Glass
(b)
PSPs at the base of the prism which is first covered with the thin metal layer by evaporation and then coated, for example, by LBK layers of increasing thickness. The reflected intensity is monitored as a function of the angle of incidence by a photodiode. The thin dielectric coating causes a shift of the dispersion curve (PSP) to higher momentum klp = k°p + Aksp
Z.0
55
6/deg
Figure 12-22. (a) Schematic drawing of the architecture of the thin film: 50 nm Ag is evaporated onto the glass-prism and then a double layer of polyglutamate is laid down by the Langmuir-Blodgett-Kuhn technique. The long, oriented rods are the polypeptide oc-helices. Hydrocarbon represents the disordered alkyl sidechains. (b) Reflected intensity (AL = 633 nm) as a function of the (external) angle 6 for various film architectures: Solid circles are bare Ag (eAg = —16.35 + i0.6, thickness 5 = 50.8 nm), solid triangles are two layers of polyglutamate on top, solid squares: 4; crosses: 6; open triangles: 10, and diamonds: 20 layers. Solid lines are Fresnel calculations.
metal (e.g., Ag)-air interface (PSP0). The horizontal line at coL intersects the dispersion curve at /c°p and thus defines the coupling angle 60. A typical experimental setup is schematically given in Fig. 12-21 b. p-Polarized photons from a laser excite
(12-69)
which, according to Eq. (12-68), shifts the resonance to a higher angle 81. From this shift and Fresnel's equations, one can calculate the optical thickness of the coating (Gordon and Swalen, 1977). This is demonstrated quantitatively for LBK multilayer assemblies prepared from polyglutamate monolayers. A schematic drawing of a transferred double layer is presented in Fig. 12-22 a. It shows the quasi-two-dimensional nematic structure of these "hairy rods" - called stiff polyglutamate-helices with long, flexible alkyl sidechains (Duda et al., 1988). Figure 12-22b gives the results of ATR scans obtained from the bare metal and 2, 4, 6, 10, and 20 layers of polyglutamate, respectively (Hickel et al., 1990). The symbols are the experimental data points, the solid lines Fresnel calculations. All these layer preparations can be described by a constant index of refraction, nz = 1.486, and the multiple of a constant thickness per monolayer of d0 = 1.75 nm.
12.4 Polymers in Special Configurations
The sensitivity of this technique to extremely small refractive index changes is also demonstrated in Fig. 12-23. Figure 12-23 a gives the angular scan (full circles) of the PSP resonance found for a sample consisting of Ag/8 layers of polyglutamate/ 6 layers of an azobenzene derivatized (and hence photoreactive) liquid-crystalline polyacrylate (structural formula is given in the inset) which was processed by the LBK technique (Sawodny et al., 1992). The spacer layers were necessary in order to decouple (electronically) the azobenzene chromophores from the metallic acceptor states in the substrate. Illuminating the sample with UV light [1 = (360 ± 30) nm] shifts the dark-adapted all-trans chromophores through trans-cis isomerization to a new photostationary equilibrium with a
557
high ds-isomer content. As a result the optical refractive index anisotropy is changed and the index normal to the layers is reduced. This shifts the surface plasmon resonance to smaller angles - the LBK multilayer is now optically thinner (open triangles in Fig. 12-23 a). If this change of the optical thickness is followed on-line (during illumination) by recording the reflected intensity at a fixed angle of incidence (e.g., 0 = 47.2°, cf. Fig. 12-23 a) a kinetic analysis of the refractive index changes can give information about the reaction rates, the equilibrium changes, the reversibility, etc. The latter important aspect is shown in Fig. 12-23b. The first rapid decrease of nz upon UV illumination can be partly restored by switching to visible light [X = (450 ± 30) nm], which isomer-
-VIS-
4?*
1.57 (b)
100
1.56 •
1.55
46
48 6 /deg.
1.54
0
10 20 exposure time / min
Figure 12-23. (a) Reflected intensity as a function of the angle of incidence for a sample consisting of Ag 8 layers PM-co-OLG/6 layers of the photoreactive liquid-crystalline polymer given in the inset. Full circles: layers as prepared (in the dark-adapted trarcs-state); open triangles: after illumination of the layers with U V light (A = 360 nm) for 10 min. Full curves are Fresnel fit calculations, (b) Refractive index change as obtained by recording the reflected intensity of the same sample at a fixed angle of incidence [0 = 47.2°, cf. (a)] while first illuminating with UV light, then after 6 min switching to visible light (VIS), etc.
558
12 Optical Properties of Polymers
izes the azobenzenes back into the transstate. These reaction cycles can be conducted many times. It is important to note that, given the signal-to-noise level of these data, index changes as small as Anz = 0.0001 can be monitored in reactive layers that are only 15 nm thick! The above described angular shift of the resonance condition for different coating thicknesses is also the basic mechanism that generates the high contrast achievable in surface plasmon microscopy (SPM) of heterogeneous thin films (Rothenhausler and Knoll, 1988). Here, the reflected, scattered, and diffracted plasmonic light is Fourier-backconverted by a lens to form an image of the interface in real space on a TV camera in our case (see Fig. 12-24 a). Only those areas that are at resonance appear dark, whereas all other domains according to their relative shift of the coupling condition are more or less bright. If the laser light is coupled in at 90 the bare metal surface is at resonance for PSP excitation, and hence almost no light is reflected. Any coated region, however, is still offresonance, and hence reflects the full laser intensity. The different areas, therefore, show maximum contrast in reflection. In previous work we demonstrated that variations of the thickness of different areas as small as a few tenths of a nanometer are enough to generate sufficient contrast for an image (Hickel et al., 1989). Illumination of the sample is typically done with a He-Ne taser beam of approximately 1 mm2 spot size. The lateral resolution of this microscopic technique has been shown to be better than 5 jim (Hickel and Knoll, 1990 a). If SPM pictures are taken (and stored on magnetic tape) as a function of the angle of incidence, image analyzing computer routines allow quantitative evaluation of the reflected intensity from areas as small
as 5 x 5 }im2. The information obtained, again reflected intensity versus angle of incidence as in the usual ATR scan, can be fit to Fresnel's formulas as discussed above to yield the optical thickness, but this time with lateral resolution. The example given concerns the SPM of laterally structured subtle thickness changes caused by UV-photon-induced scission of Si-Si bonds in polysilane films, which leads to a volatilization of the illuminated material (Sawodny et al., 1991). This class of substances recently attracted considerable interest as self-developing photoresists, materials which do not require additional wet etching after exposure to convert the illumination pattern into a surface relief structure (Miller and Michl, 1989). Figure 12-24b shows a series of SPM pictures taken at various angles of incidence from approximately 40 nm thick poly(methylphenylsilane) film spin-coated onto a Cr/Au substrate and then illuminated for 15 min in an Ar atmosphere (10~ 3 MPa) through an electron microscope copper grid. The exposed areas tune into resonance first - a clear indication that they are optically thinner. A quantitative analysis is possible if one plots the mean gray values of preselected areas obtained by an image analyzing program that takes the intensity histograms from the pixels in each of these local frames as a function of the angle of incidence, 6. This is given in Fig. 12-24c for both illuminated and unexposed areas. Assuming that the index of refraction for both areas is the same (n = 1.635), we can obtain from a Fresnel fit to the data the two respective thicknesses, as indicated in Fig. 12-24c. Given the shift of the resonance curve upon ablation of only 2 nm, one can estimate that the thickness resolution of this quantitative microscopic technique is a few angstroms.
12.4 Polymers in Special Configurations
559
(a)
Laser, wavelength X
Screen/Camera
Evanescent Field of PSP
(b)
185°
66
68 9/deg
Figure 12-24. (a) Scheme of SPM set-up in the Kretschmann configuration. The scattered and out-coupled plasmon light is Fourier transformed by a lens to give an image of the interface in real space, (b) Series of SPM pictures taken at various angles of incidence from a structured polysilane film partly ablated by irradiation for 15 min in an Ar atmosphere through an electron microscopy grid (a line grating in this case). First, the thinner areas are tuned into the PSP coupling resonance: Between 66° and 67° the contrast inversion is reached, and then the unexposed areas are darker, (c) Quantitative analysis of the average gray value (a measure of the reflected intensity) taken from the two different areas (irradiated: open circles, unexposed: solid circles) pictured in (b) as a function of the respective angle of incidence. Solid and broken lines are Fresnel fits to the data points. From their relative shift, one calculates an ablated thickness of Ad = 2 nm.
560
12 Optical Properties of Polymers
12.4.1.3 Self-Assembly Mono- and Multilayers on Solid Supports
An extremely attractive alternative to transferred Langmuir monolayers are monomolecular films prepared by a selforganization process: Suitable substrates are just immersed into a solution that contains the surface-active molecules. These then self-assemble at the solid-solution interface to a well-organized monomolecular layer with a structural (positional and orientational) order absolutely comparable to Langmuir films. This is schematically shown in Fig. 12-25. The first examples, introduced by Sagiv (1980) were based on the specific interaction between the silyl-headgroup of the amphiphiles (e.g., of octadecyltrichlorosilane) and the polar silanol groups on glass- or quartz-substrates. More recently, a major research activity has focused on the physics and chemistry of self-assembly monolayers (SAMs) based on the interaction of thiol-, disulfide-, and sulfide-groups with Au and Ag surfaces. The possibility for chemical manipulations at the co-position of long chain thiols [X-(CH 2 ) n -SH with X: functional group] opened a wide field of tailor-made surface functionaliza-
Figure 12-25. Scheme of the self-assembly process that leads to a (polymer) monolayer at a solid substrate.
tions. This molecular engineering of surface properties has wide implications for lubrication, patterning, (bio-)sensing, cellsurface interactions, etc. One example of a polymeric SAM with multifunctionalities is presented in Fig. 1226. The structural formula of this copolymer (Fig. 12-26 a) shows the disulfide-containing "stitching"-units that accomplish the covalent and hence stable binding to the (noble) metal surface (Spinke et al., 1992). The long alkyl-chains in another comonomer play the role of the hydrophobic tails of "classical" amphiphiles while the short alcohol moieties are introduced to balance the polar and apolar parts of the monolayers. The self-assembled monolayers can be analyzed again very sensitively by PSP spectroscopy. This is demonstrated in Fig. 12-26b. The resonance angle recorded for the bare Au substrate (full circles and full curve) is shifted to higher angles by A9 = 0.5° (open circles). From the Fresnelfit calculations (dashed curve) one obtains an (optical) thickness of the monolayer of d = 1.9 nm (assuming an index of refraction of ft = 1.55). From this result (and from other measurements) the structural model that is depicted in Fig. 12-26c has emerged: a well-ordered monomolecular layer of amphiphilic moieties are coupled to the rough and irregular substrate by a polymeric buffer layer. This decoupling of the perfectly organized monolayer from the imperfect solid support not only results in a polymer coating with interesting viscoelastic properties but is also believed to allow for the build-up of fluid membrane systems, an extremely relevant aspect for many membrane biophysical problems. In passing we note that recently the formation of multilamellar structures by the self-assembly technique was also reported (Maoz etal, 1988; Tillmann et al., 1989):
561
12.4 Polymers in Special Configurations
(a) CH2 I ^CH2(CH2)16CH3 CH,-C-COO-CH, CH,-NH-CO-N V •" -CH2(CH2)16CH3 CH2 H-OCOO-CH 2 CH 2 -OH
CH 3 -S-S-CH 2 CH 2 -OOC-C-CH 3
Figure 12-26. (a) Structural formula of a ter-polymer with different functionalities that leads to a well-ordered self-assembly monolayer, (b) PSP-resonance of the bare gold substrate (full circles) and after coating with a self-assembly monolayer of the ter-polymer (open circles), (c) An artist's view of the polymer-supported monolayer prepared by the self-assembly process on a rough substrate.
(b)
(c) Monolayer
Polymer Support
Substrate 6/deg
After the formation of the first monolayer, suitable processing of the endgroups of the molecule (e.g., the oxidation of terminal double bonds to carboxyl-groups) transforms the SAM layer into a substrate for further self-assembling of the next monolayer. Even thick multilayers of more than 100 layers have been demonstrated. 12.4.2 Integrated Optics
Since the concept of "integrated optics" emerged in the late 1960s a significant growth of the areas of optical communication, signal processing, etc. has stimulated
a major research and development effort in the field of organic and polymeric materials aimed at optimizing their functional properties by molecular design and synthesis (Stegeman et al., 1987; Prasad and Williams, 1991; Messier et al., 1989). This concerns not only their nonlinear optical properties (i.e., primarily their second- and third-order hyperpolarizabilities, see Sec. 12.5) but also their linear optical response to light and, equally important, also other chemical and physical properties of the material and processing parameters like stability (against chemical attack, temperature, light, aging, delamination, etc.), ease
562
12 Optical Properties of Polymers
of preparation (compatible with technologies developed for microelectronic device fabrication), cost-effectiveness and others. One very crucial requirement of a material for the envisaged integrated optics application is that it can be prepared as a structure capable of guiding light waves with acceptable losses. The three corresponding basic configurations (Lee, 1986; Chang, 1991) are schematically depicted in Fig. 12-27: The planar waveguide structure (Fig. 12-27 a) given in the so-called asymmetric slab configuration with a solid substrate and air as superstrate, the various channel waveguides (Fig. 12-27 b, strip, embedded strip, and buried strip, from top to bottom) prepared by starting in some cases with a planar waveguide and then further structuring and etching the sample, and the optical fiber (Fig. 12-27c) with a
Planar Waveguide Substrate
high index core and a lower index cladding. Such waveguide structures are discussed in the literature not only with respect to their role as passive interconnects between different integrated optical circuit (IOC) components but also in the context of various design concepts for active IOC modules such as directional couplers, phase and amplitude modulators, and second harmonic generators for laser diodes (Stegeman and Stolen, 1989). In this section we first describe very briefly the basic theoretical approaches (within Maxwell's theory) needed to understand the phenomenon of guiding light in confined geometries. We limit the discussion to the case of a planar waveguide structure but the general ideas allow for a straightforward extrapolation to more complicated designs. After a short presentation of various schemes for coupling plane electromagnetic waves, for example, photons from a laser, into such structures we give a few examples of waveguides fabricated from novel polymeric materials with promising optical properties that can be tailored to some extent for future device applications.
Channel Waveguides
12.4.2.1 Planar Waveguide Structures
Guiding Light in an Asymmetric Slab
c) Fiber Cladding Core
Figure 12-27. Basic waveguide structures, (a) Planar waveguide (shaded) on a substrate, (b) Channel waveguides: strip, embedded strip, buried strip (from top to bottom), (c) Optical fiber with the core and the cladding of lower refractive index.
We consider the asymmetric slab configuration sketched in Fig. 12-28. The thin waveguide layer (film with index of refraction n2 and thickness d) is bound by the substrate (n3) and the superstrate (e.g., air n1 = 1). In order to obtain total internal reflection at each interface the waveguide material must fulfill the requirement n1
12.4 Polymers in Special Configurations
563
Air Film Substrate n3
Figure 12-28. The geometry of a planar waveguide with substrate (index of refractionrc3),film (thickness d, n2), superstrate, e.g., air (n^). Given is also the coordinate system used to derive the mode equation. For illustration only note the ray optic description of the guided light.
schematically in Fig. 12-28). Note the evanescent wave extension giving rise to the Goos-Hanchen shift (Goos and Hanchen, 1947). Within Maxwell's theory it is straightforward to show that for the threelayer system depicted in Fig. 12-28 one can write the optical field in each of the media (omitting the time dependence and the propagation in the x-direction). z>d 0
^3y —
(12-70) (12-71) (12-72)
Matching both electric and magnetic fields at each interface we obtain at z = 0 (12-73)
A3 = A2 and fc
3z^3 = fc2z^2 s i n ( £
(12-74)
Therefore, 0 = tan" 1 p23 with p23 = k3jk2z
(12-75)
and (12-76)
= A3 At z = d one finds A, = A2 cos (k2zd + >)
(12-77)
and — klzA1 =
-k2zA2 sin(k2zd + 4>) (12-78)
(12-80) The phase factor
Then (j) = —k2zd -f tan"ip21
and
with jg 21 = ' C l z / ' C 2 z (12-79)
The three most important techniques for coupling light into a waveguide are
564
12 Optical Properties of Polymers Air
Film
Substrate Figure 12-29. Optical field distribution in a planar waveguide structure of substrate/film/air configuration for the first three s-polarized modes (m = 0,l, 2). The higher index of the substrate (n3 = 1.46) compared to the superstrate (air, n1 = \.O) is the reason for the much stronger evanescent field penetrating into the substrate material.
schematically sketched in Fig. 12-30. The grating (Fig. 12-30 a) and prism coupling method (Fig. 12-30b) are commonly used for planar waveguides. Efficient coupling can be achieved only under wavevector matching conditions. For grating coupling this requires that the photon wavevector component parallel to the surface plus or minus a multiple of the grating vector magnitude G = 2 n/A with A being the groove spacing of the grating matches to the waveguide propagation vector
= kon1sin9
±mG
(12-82)
a) Grating
Prism
c)
End-Fire
Figure 12-30. The three basic coupling geometries, (a) Grating coupling, (b) prism coupling. Both methods require the tuning of the angle of incidence, 9. (c) End-fire coupling, e.g., by a microscope objective.
In the case of prism coupling the projection of the photon wavevector at the base of the prism must be equal to the guided wavevector: /cx?m = /corcpsin6>
(12-83)
with np being the index of refraction of the prism. This requires a prism with an index higher than the waveguide material. The end-fire coupling (e.g., from a microscope objective to a channel waveguide) matches the incident field spatial distribution to that of the guided wave (Fig. 1230 c). For a given waveguide, for example, a planar structure of fixed thickness, the number of modes and their angular positions when excited by a grating or prism [see Eqs. (12-82) and (12-83)] depends on the wavelength employed and on the index of refraction of the various layers. For a thin film of an unknown waveguide material this allows for the determination of the corresponding materials parameters, as we will show next. Planar Waveguides from Novel Polymeric Materials For the examples presented in the following, a modified prism coupling technique as schematically depicted in Fig. 1231a was used. The waveguide modes were launched with the help of a high index
12.4 Polymers in Special Configurations
,7) Detector
Glass Prism Metal Waveguide
(b) CH 3
CH 3
100
:
\
2
i
50
t
_J
20
1
i *
1
1
0/deg
Vv
f
1; 80
Figure 12-31. (a) Attenuated total reflection set-up for coupling surface plasmon - as well as waveguide modes into a planar film structure. Coupling is achieved through a thin metal gap at the base of the prism, (b) Reflected intensity recorded as a function of the angle of incidence for a thin film of a solid polyelectrolyte (structure formula given on top) spincoated onto the Ag-coated (s = 50 nm) base of a high index prism (SF57) [cf. (a)]. The dots are the experimental data points; the full curve is the theoretical fit. Excitation with p-polarized light from a He-Ne laser (A = 633nm).
prism by coupling through a metal (Au or Ag) gap of 50 nm thickness excited with p-polarized light. In this geometry the m = 0 mode is the surface mode, the surface plasmon. Its optical intensity peaks at the metal-waveguide interface and decays (for
565
films of typical waveguide thicknesses of approximately 1 jim or more) completely within the film. Its polarization is predominantly parallel to the film normal. Its angular position, therefore, depends only on nz, not on the waveguide thickness. The other waveguide modes can be excited with either s- or p-polarized light and will, in general, depend on all three indices, that is, HX, tty, and nz as well as on the film thickness [cf. Eq. (12-81)]. Figure 12-31 b shows a series of waveguide modes obtained with p-polarized excitation of X = 633 nm in a spin-coated film of a solid polyelectrolyte whose structure formula is given in the inset. The various modes were measured in reflection. The broad dip in the reflected intensity at the high angle side (6 « 73°) is the surface plasmon. The crosses are the experimental data points, the full line (for better clarity somewhat shifted) is a theoretical fit based on a Fresnel calculation (Wolter, 1956). From this one obtains the thickness of the waveguide film d = 1.85 jwm as well as its index of refraction n = 1.63. By recording the reflectivity as a function of the angle of incidence for s-polarized light (not shown) it was found that an isotropic approximation of the refractive index was sufficient in order to characterize the whole waveguide mode pattern. Nevertheless, a closer comparison between the experimental and the theoretical curve of the angular reflectivity shows some distinct differences. First of all, the experimental modes are somewhat broader than calculated with an ideal slab geometry for the waveguiding film. This indicates some thickness and lateral refractive index heterogeneities within the spin-coated film. In addition, not all calculated angular positions for the various modes fully coincide with the experimental ones. Given the different field intensity distributions of the different modes (cf. Figs. 12-29 and 12-31)
566
12 Optical Properties of Polymers
this discrepancy points to a certain perpendicular refractive index variation (an index profile) within the waveguide. Both deviations from ideality within the plane and normal to the plane need to be carefully avoided in an IOC structure because such heterogeneities clearly would limit the optimum performance of the device.
Screen/Camera
Waveguide
Optical Waveguide Microscopy (OWM) One way of optically characterizing the quality of planar waveguide structures, for example, spin-coated films, is the recently introduced optical waveguide microscopy (OWM) (Hickel and Knoll, 1990 b) which was developed in analogy to surface plasmon microscopy. Both techniques employ bound optical waves to illuminate the sample instead of using normal photons. The basic principle of OWM is schematically depicted in Fig. 12-32 a. Laser light is coupled to the waveguide at an angle of incidence 9. The nonideal planar structure of the film leads to an angular broadening of the resonance condition. Now, the spe-
Figure 12-32. (a) Scheme of the experimental configuration for taking microscopic pictures of the local resonances in planar waveguide structures. Laser light is coupled to the waveguide at an angle of incidence 6. The scattered and out-coupled light is Fourier-backconverted by a lens to form an image of the waveguide on a TV camera, (b) Series of optical waveguide microscopic pictures taken from a planar film of a solid polyelectrolyte spin-coated onto the base of a Ag-coated prism [cf. (a)]. The images taken at different angles of incidence (as indicated) show the local variation of the resonance by the changing pattern of dark areas. Each picture represents 570 urn x 420 jam on the waveguide, (c) Topographical map of the "planar" waveguide structure investigated by OWM. The lines were obtained from pictures like those shown in (b) by simple image analyzing routines indicating the local waveguide resonances. The numbers give the corresponding local thickness in nanometers of the waveguide as derived from Fresnel calculations.
345.0
12.4 Polymers in Special Configurations
cial prism coupling configuration used for these experiments ensures that the propagation lengths of the guided modes do not exceed circa 10 jim because guided light intensity constantly couples out through the metal gap and the prism. This light reflected and scattered and out-coupled is Fourier backconverted by a simple lens to form an image of the thin film structure on a screen or on a TV camera (CCD) in our case, just as discussed above for surface plasmon microscopy. As a result we are dealing with a microscope that uses the guided light itself as the illumination light source for the imaging of the sample. Figure 12-32b shows a series of OWM pictures taken with p-polarized light (k = 633 nm) at different angles of incidence as indicated. The sample was a spun-on film of the ionene polymer shown in Fig. 12-31 b. The dark regions correspond to the local resonant coupling to waveguide modes. As one varies the angle of incidence, different areas tune into resonance corresponding to the respective local fulfillment of the eigenvalue equation [see Eq. (12-81)]. The reduced propagation length of the modes in this experimental configuration allows for the rather local analysis of this resonance pattern. If the index of refraction of this thin film material is known, a Fresnel calculation allows for the conversion of the angular position of the resonance into a local thickness. A simple image analyzing computer routine gives for each frame, the line(s) of lowest intensity which - for a series of pictures taken at different angles of incidence - can be used to construct a topographical map of the thickness variations of the waveguide slab. This is shown in Fig. 1232 c for our ionene sample. That we are dealing, in fact, with thickness and not refractive index variations was checked independently by performing the same analysis
567
with s-polarized light (not shown), which gave essentially the same thickness data. It is very obvious that such a film would not meet the requirements for device applications. This example demonstrates, however, that experimental techniques are available that allow a comprehensive testing of polymer optical properties and processing steps. Waveguide Loss Measurements Typically, waveguides are designed to propagate light over long distances. This requires, first of all, the absence of any absorptive losses, either by molecular, electronic or vibronic transitions or by dissipation, for example, in metallic materials used as electrodes. Equally important is the control of any scattering losses, which would also result in an enhanced damping of the guided modes. This includes not only the minimization of bulk (volume) fluctuations of the optical density but requires also the ultimate control of surface scattering processes, for example, at interfacial roughness caused by processing steps. One way of characterizing waveguide structures along these lines is to measure the losses of the guided waves directly. Figure 12-33 a shows a scheme of an experimental set-up designed for loss measurements in planar structures. Laser light is coupled to the waveguide by a grating (or prism). A coherent fiber bundle is used to transmit the light scattered off the guided streak in the planar structure which is a measure of the local mode intensity. A photomultiplier tube with a slit diaphragm reads this scattered intensity with high spatial resolution and hence allows the quantitative analysis of the (exponentially) attenuated optical field in the waveguide. A data acquisition and analysis system gives the
568
12 Optical Properties of Polymers
translation stage PMT
waveguide substrate
aspects are noteworthy: (i) For most of the wavelength range investigated the data seem to follow a l ~ 4 power law (straight line in Fig. 12-33 b) which suggests Rayleigh scattering as the predominant attenuation mechanism, (ii) For wavelengths X > 650 nm losses in the range of 1 dB/cm or less seem to be feasible with these materials, which makes them also technologically interesting. Supramolecular Design of Waveguide Films by the Langmuir-Blodgett-Kuhn Technique
500 600 Wavelength X/nm
700
Figure 12-33. (a) Scheme of the set-up used for the determination of mode losses in waveguide structures. Laser light is coupled to the planar waveguide by means of a grating or prism. A fiber bundle records the local level of stray light, which is a measure of the field intensity along the guided streak, (b) Loss data for s-polarized (m = 0) modes (given as attenuation in dB/cm) for various wavelengths as obtained for a planar waveguide structure fabricated by spin-coating from a solid polyelectrolyte. The straight line is a X ~ 4 power law. Data from Mathy et al. (1991).
loss-values typically in dB/cm (1 dB corresponds to an intensity attenuation to 79%). Figure 12-33 b gives the results obtained from a polyelectrolyte thin film prepared by spin-coating onto a low-index glass substrate measured in air with prism coupling. Data points taken from Mathy et al. (1991) were measured with different laser lines and plotted correspondingly as a function of the wavelength employed. Two
Finally we should briefly discuss possibilities for the ultimate control of a planar waveguide structure at the molecular level as it is in principle possible by the LBK method. Here, multilayer assemblies capable of carrying guided modes are built-up, layer by layer, by dipping and withdrawing a substrate many times through a monomolecular layer prepared at the water-air interface (Swalen et al., 1978). The number of possible molecular structures of the individual layers is near infinity and so are the functionalities that can be tailored into these systems. Even if it is unlikely that the tedious preparation process will ever make it into a large-scale production line, these LBK films nevertheless are extremely helpful as model systems to develop and test molecular structural concepts, functional units (nonlinear chromophores, etc.), supramolecular assemblies, and device concepts. As an example of a simple passive planar waveguide structure prepared by this LBK technique we present in Fig. 12-34 results obtained with a multilayer assembly of the polyglutamates previously presented. Figure 12-34 demonstrates the good optical quality of a thick multilayer assembly (150 layers) capable of guiding modes with different polarizations: the upper frame was
12.4 Polymers in Special Configurations
569
n
\\ ~ni = 0.01 and demonstrates the sensitivity of the coupling conditions to subtle refractive index variations. 12.4.2.2 Channel Waveguides
8/deg
Figure 12-34. Waveguide mode pattern as obtained from a polyglutamate (PM-co-OLG) multilayer assembly of 150 layers for p- (upper frame) and s-polarized light (lower frame). Full circles are obtained for modes propagating perpendicular to the dipping direction, which is the preferred axis of the hairy rods, crosses are from modes propagating parallel to the rods. The full curves are Fresnel calculations showing the corresponding in-plane anisotropy of the index of refraction.
taken with p-polarized light, the lower frame shows data with s-polarization. The full circles are experimental points measured with waveguide modes propagating perpendicular to the dipping direction, the crosses are for parallel propagation. The full lines are Fresnel fits and show the excellent theoretical description of these structures as planar slabs. The optical inplane anisotropy that can be analyzed quantitatively amounts to only An =
The theoretical description of guiding light in a two-dimensional waveguide structure (optical confinement in two dimensions as in the channel waveguides shown in Fig. 12-27b) adds nothing new, and just requires more complex numerical techniques for the calculation of the exact dispersion relation and hence the various modal structures. These channel waveguides are perhaps the ultimate structures needed for any IOC. These include, in particular, active devices based on nonlinear optical effects. Two basic structures are shown in Fi^. 1235: the (electro-optical) coupler, in which two channel waveguides are sufficiently close so that they can interact like two weakly coupled mechanical oscillators, is shown in Fig. 12-35 a. The strength of the interaction which determines, for example, the flow of light intensity from one channel (a)
Electrode / \Channel Waveguides
(b)
Figure 12-35. Basic structures for integrated optical circuits (with special emphasis on electro-optical control): (a) Channel waveguide coupler, (b) (integrated) Mach-Zehnder interferometer.
570
12 Optical Properties of Polymers
to the other can be controlled by the application of a DC electric field through the Pockels effect provided the polymeric waveguide material has a correspondingly large second order susceptibility x(2)- The integrated Mach-Zehnder interferometer shown in Fig. 12-35 b allows for a phase shift of the light in one branch relative to that propagating in the other channel. This then leads to constructive or destructive interference at the output. Again, this interaction can be controlled by either #(2)-active materials through the application of an electrical field (as shown in Fig. 12-35 b) or by a material with an intensity-dependent refractive index n2 [cf. Eq. (12-62)]. The fabrication of these channel waveguides in many cases starts with a polymeric thin film. In addition to the mechanical embossing of the channel structures, a technique called photolocking, schematically sketched in Fig. 12-36, has been widely used to generate a local confinement for guided light: a photoreactive polymer film is exposed to light through a suitable mask in a way similar to photoresist patterning. Heating the film leads to an evaporation of the unreacted dopant whereas the cross-linked areas define the channel waveguide. Refractive index changes of approximately 1%, sufficient for most configurations, can thus be obtained. Another technique has recently been demonstrated as an easy way to also generate passive waveguide structures. A polymeric multilayer composed of buffer layer, a nonlinear core layer and a top buffer layer was used as a matrix for the fabrication of a 1 x 2 channel waveguide splitter by a simple bleaching of the chromophores in the core layer. A 6 jim wide channel was split in a Y-configuration of two channels, each 6 jim wide and 50 |im apart (Mohlmann et al, 1990). The intensity output
(a)
I1 i
11 ! (b)
(c)
Figure 12-36. The photolocking process: (a) Doped polymer film, spin cast onto a substrate; (b) polymerization reaction following (local) irradiation with light; (c) heating the film evaporates the unreacted dopant.
profile of the 1 x 2 splitter shown in Fig. 12-37 demonstrates the good optical isolation between the two channels. There is certainly still a long way to go before a more complex integrated optical chip is a commercial reality but the feasibility in principle of all basic structures has been demonstrated. 12.4.2.3 Polymer Optical Fibers (POFs) Single-mode glass optical fibers are widely used as long-distance and high-
6|im guides
Figure 12-37. Intensity distribution at the output of a 1 x 2 passive channel waveguide splitter (see insert). Data from Mohlmann et al. (1990).
12.4 Polymers in Special Configurations
bandwidth communication media because of their excellent transparency. In shortdistance networks, however, where many junctions between different fibers are necessary, already slight displacements of two coupling cores each with a diameter of only about 5 jim would cause substantial coupling losses. Polymer optical fibers (POFs) with their considerably larger core diameter (ca. 1 mm) are in this respect much easier to handle. In addition, they offer other advantages, such as low cost, easy processing and flexibility. However, all commercially available POFs have been of the step-index (SI) type (Fig. 12-38 a). For these multimode fibers the different guided waves all propagate with the same phase velocity but along different paths within the core, a phenomenon which is called modal dispersion. This is schematically depicted in Fig. 12-38 a. Due to internal mode coupling, even a single input pulse (cf. Fig. 12-39 a) is greatly distorted and considerably broadened. An example is given in Fig. 12-39 b. This means that two consecutive input pulses need to be separated in time in order to be resolved at the output as two individual pulses. The transmission bandwidth is thereby greatly reduced and currently amounts for SI-POF to about 5 MHz km. A solution to this problem is the graded index (GI) fiber. Here the refractive index profile with the highest values in the center of the core results in a higher phase velocity of those "rays" propagating in the outer periphery of the core. Their longer pathways are, therefore, compensated by a faster propagation speed. As a consequence the input pulse is far less broadened, which results in a higher bandwidth. An example that was recently reported for GI-POFs prepared by a special copolymerization process (Koike, 1991) is given in Fig. 12-39: Fig. 12-39 c shows the refractive
571 n=1
(a)
(b)
Figure 12-38. Typical optical rays and refractive index profile of a multimode step-index (SI) fiber (a), and a graded-index (GI) fiber (b) (after Saleh and Teich, 1991).
1ns/div (a)
Input pulse
SI POF , ^ %
(b)
Put pulse
Out
GI POF (d)
ft Output
pulse
(c)
-0.005 "
-0.010 " -0.015 0.0
Figure 12-39. (a) Short input pulse in a POF; (b) output pulse after traveling through a step-index fiber; (c) index profile of the graded-index fiber that transmits the input pulses nearly unbroadened as shown in (d) (after Koike, 1991).
572
12 Optical Properties of Polymers
index profile, Fig. 12-39 d compares the input pulse with the output pulse. The bandwidth obtained was shown to be 2 GHz • km (Koike, 1991). In the development of polymer optical fibers another improvement concerns the intrinsic losses arising from absorption processes. This is indicated in Fig. 12-40 for a fiber made of poly(methylmethacrylate) (Groh et al, 1989). The higher overtone vibrations of the C - H groups in the material are responsible for attenuation losses at A^680nm of more than about 200dB/km. Perdeuteration results in a substantial reduction of the band strengths but also in a shift of their spectral position. Both effects reduce the loss in these fibers considerably (Fig. 12-40). Other attempts
to reduce the absorption coefficient at relevant communication wavelengths successfully introduced fluorocarbon moieties into the material with the additional improvement of the maximum operating temperature. This is an important issue, for example, for the implementation of POFs in automobiles (Groh et al., 1989). 12.4.3 Polymers for Optical Recording
With the development of low-cost solidstate laser diodes, the optical storage of information has become an attractive alternative to, for example, magnetic storage concepts. The basic principle is schematically depicted in Fig. 12-41. The information is stored on the disc as a succession of
200 -|
J
150
X
OSS
DQ 100
50
PMMA-d8 structural imperfections
Figure 12-40. Spectral loss data for a regular PMMA POF (dashed curve) and for a fiber made from perdeuterated PMMA (full curve). Some vibrational band assignments are given (after Groh et al., 1989).
I
500
600
700
800
900
X (nm)
Pre-Embossed Data Pits Reflecting Layer Disk Substrate Laser Read Beam
Figure 12-41. Some details of the pit sequence in a CD-ROM. Schematically given is also the laser read beam.
12.4 Polymers in Special Configurations
small pits of varying length and repetition frequency (Bowden, 1988). Read-out is accomplished by focusing a laser beam (typical diameter ~ 1 |Lim) onto the surface of the rotating disc and monitoring the reflected light by a photodiode. The reflectivity of the pits differs from that of the surface or land area, so that the intensity (or polarization) of the reflected light and hence the electrical output of the photodiode will be modulated according to the pattern of the pits. Various schemes have been developed for optical data storage: the read-only memories (ROMs), the "write-once-readmany times" (DRAW: direct read after write) memories, and the erasable memories. The first category follows the consumer products such as the optical video disk and the digital audio disc (CD: compact disc) where the pit sequence is produced already at the disc fabrication step, for example, by injection molding into a master mold containing all the pit structures (CD-ROM). Some requirements of the polymer material for the disc substrates were discussed already in Sec. 12.3.1. For the write(-once) process different mechanisms have been introduced. They can be categorized according to the schematic picture given in Fig. 12-42. Ablative processes produce marks in the form of shallow pits (Fig. 12-42 a) or deep pits (Fig. 12-42b). Another thermal process may lead to bubble formation (Fig. 12-42c) or to any other change of the optical proper-
573
ties of the thin film recording medium (Fig. 12-42d). In all cases, light typically from a near IR laser is (locally) absorbed by the film and converted into heat. This requires a functionalization of the polymer by a suitable absorber which can be an inorganic additive (carbon black, metal powders, or y-Fe 2 O 3 pigments) or an organic dye either mixed or copolymerized with the polymer matrix. The low thermal conductivity of polymers leads to very high local temperatures within the irradiated spot, causing the polymer material to degrade and volatilize. For bubble formation this degradation of the material occurs at the interface to the (reflecting) substrate so that the trapped vapor deforms the hot absorber layer to reproduce the desired microstructure. Some of the mechanisms suggested for phase-change storage in polymer films (Fig. 12-42d) include photobleaching and photochromism of dyes or the thermal conversion of amorphous into crystalline domains. By far less developed are erasable optical storage concepts. Examples of organic polymer based reversible media are few. One class consists of liquid crystalline (LC) side group polymers (see also Chap. 5 of this Volume). The LC polymer in a homogeneous, transparent meso-phase, for example, in a smectic state (Fig. 12-43 a), frozen below the glass transition temperature is locally distorted by the heat generated by the absorption of a laser pulse (Fig. 12-43b). This distortion can be read,
Figure 12-42. "Marks" produced during different laser writing processes: (a) Shallow pit, (b) deep pit, (c) bubble deformation, (d) optical property change.
574
12 Optical Properties of Polymers
(a) mmwmmm
t Figure 12-43. Schemes of the data storage process in liquid-crystalline polymer matrices doped (by mixture or copolymerization) with photo-addressable dyes, (a) In the smectic mesophase the matrix is highly ordered and hence transparent; (b) a focused laser pulse locally heats the sample into a disorded state; the bit is written; (c) the disorder remains frozen in and scatters the light if the reading beam is scanned across the polymer; (d) erasure of the stored information is accomplished by heating the sample into the nematic, which allows the disorder to anneal.
for example, by its ability to scatter light (Fig. 12-43 c). Since the matrix is in a glassy state back-relaxations are extremely slow and the data point is sufficiently long-lived. Erasure of the stored information can be accomplished by heating the film to a temperature above the glass transition temperature (but below the transition temperature to the isotropic phase) (Fig. 12-43 d). Another reversible storage concept is based on the bubble formation mentioned above. Here, a double layer structure, called the expansion and the retention layers, are the basic unit (Fig. 12-44). The expansion layer is an elastomer containing a dye absorbing the writing laser light at 840 nm. Its expansion deforms the retention layer - a surface bump that can be detected during the read cycle. Erasure is effected by exposure to a second laser beam at 780 nm, the wavelength which can be absorbed by the dyes in the retention layer but not in the expansion layer. The thermal softening of the retention material then allows the relaxation of the elastic energy stored in the deformation, and planarity is restored. The aforementioned two-dimensional storage concepts are limited in storage
density by the spot-size of the focused laser beam and reach currently about 108 bit/ cm2. Future availability of blue laser diodes emitting in the wavelength range of X = 400 nm will allow a further enhancement of the bit density. However, other
- Retention layer - Expansion layer - Substrate
-Write beam
- Read beam
Erase beam
Figure 12-44. Erasable dye-polymer double layer storage scheme based on bubble formation (for details see text).
12.4 Polymers in Special Configurations
concepts involving three- (or even four-) dimensional storage principles will compete with the "classical" 2D versions. It was suggested that the use of volume holograms would be one way to enhance the storage density by including the third spatial coordinate of the storage medium as well. One possibility is the use of photopolymers which change their refractive index irreversibly (An < 10 ~2) upon irradiation. This way, the interference pattern between a reference beam and an object beam both from a coherent (laser) light source can be stored persistently in the material. Read-out is performed just with the reference beam. Suitable polymers with a residual monomer content of approximately 10% are polyacrylates (Franke et al., 1984), polyacrylamides, polystyrenes, and polycarbonates. Theoretical storage densities of up to 10 14 bits/cm3 should be possible. A completely different principle for introducing a third dimension is based on photochemical hole-burning. The schematic representation of this storage process in the frequency domain is shown in Fig. 12-45. The absorption spectrum of a dye in a polymer matrix shows a rather broad band, reflecting the broad distribution of sites, each of them defining a slightly different microenvironment for the incorporated dye molecules. As a result each dye molecule with a narrow homogeneous
absorption band Fh appears in the absorption spectrum with a slightly different solvent shift thus generating the broad inhomogenous absorption spectrum (linewidth /]) of the dye ensemble (Fig. 12-45 a). A laser with a narrow line profile (narrower than the homogeneous line width of the dye molecule) at a (stabilized) wavelength, i L , within the absorption band is able to interact only with those dye molecules whose absorption profile overlaps with that of the laser. If this absorption leads to a photochemical process with a photoproduct absorbing at a distinctly different wavelength, a narrow hole is burned into the broad absorption spectrum (Fig. 1245 b). If the laser wavelength is tuned across the inhomogeneously broadened line, different subsets of molecules can thus be addressed within the same illuminated laser spot. With rjrh = 103 to 104 a storage density of 1011 bits/cm2 is then possible. The main drawback of this storage technology is the required low temperature: Only at T < 20 K are the holes sufficiently narrow and stable. 12.4.4 Polymers for Molecular Opto-Electronics
In this last section we just touch on a research area with still rather fuzzy contours: molecular electronics. The ultimate goal is the realization of logic operations at
(b)
(a)
575
Laser
O ) p LO
Figure 12-45. Photochemical hole-burning into an inhomogeneous absorption band composed of many individual homogeneous absorptions (a). Irradiating with a narrow laser line at coL burns a population hole which may appear as a photoproduct at a different spectral position &)p (b).
576
12 Optical Properties of Polymers
the molecular level. This requires molecular wires, bistable molecules that can store information, etc. Although it is completely unsolved how, in general, to address such molecular devices from the outside macroworld, some concepts include special optical properties of selected polymers. An example is given in Fig. 12-46. The photochromic isomerization of a salicylideneaniline unit incorporated into a polyacetylene chain could be used as a switch in the molecular wire from an on-state to an offstate. It is well conceivable that with the availability of tunnel microscopy tips and their ability to control and manipulate systems at the molecular level we will see a rapidly growing research activity towards molecular electronics and bio-chips.
12.5 Nonlinear Optical Properties of Polymers Polymers with large nonlinear optical susceptibilities are likely to play a major role in the rapidly growing area of photonics. This can be deduced from the many conferences and workshops in this area, from a tremendous increase in the number of publications both in journals and as books and, finally, from the fact that all major funding agencies have substantial programs to stimulate research in this interdisciplinary field of physics, chemistry, and materials science. Many excellent reviews have been devoted recently to the
off - s t a t e
topic so that we can concentrate in the following on some general fundamental device concepts, materials design considerations, and on some specific polymer issues relevant for nonlinear optics. We first give a few remarks on the functionalization of polymers for their use in nonlinear optical device configurations (Zyss, 1985). We then focus on some applications related to the second-order nonlinear optical susceptibility x(2) and finally summarize some third-order / ( 3 ) effects and how they could be implemented in future technologies for information processing. 12.5.1 General Concepts for the Functionalization of Polymers for Nonlinear Optics
The nonlinear optical response of polymers developed for future IOCs needs to be optimized through a suitable functionalization of the matrix by chromophores with correspondingly large second- or third-order susceptibilities. Both processes need, however, different strategies for an optimized material (Kajzar, 1992). For the second-order processes (cf. Sec. 12.2.5.1) the relation between the macroscopic {average) nonlinear susceptibility X(2) and the molecular nonlinear polarizability pD is determined by the following relation (Williams, 1987; Singer et al, 1987):
(12-84)
where ND is the density of the active molecules, / is a local field factor of the
Figure 12-46. Switching unit (salycilideneaniline) incorporated into a molecular wire (polyacetylene) that can be switched optically from an on-state to an off-state (after Kampf, 1985).
12.5 Nonlinear Optical Properties of Polymers
optical field, and
j5D at 1.9 urn (xl(T 3 O esu) 5.7
N(CH
2
)
2
^C^
C
NC'"
_
N
21.4
.CN 41.8
20.1
50.7
23.4
61.6
111.2
577
strates that by a systematic variation of the donor- and acceptor-strengths and of the conjugation lengths substantial improvements in /?D can be achieved. The next step, the maximal level of functionalization, concerns the question of how to dope the polymer matrix. Three strategies have been suggested. The first is the mere mixing of a suitable dye as the functional guest in a polymeric host material. One example of a suitable dye is presented in Fig. 12-47 (Levenson et al., 1991). The dye's structure formula is given in the inset. The addition of 15wt.% dye to the PMMA matrix changes the linear optical properties of the polymer as discussed already in Sec. 12.2.4.5 (cf. also Fig. 12-5): The relatively flat background index of refraction of PMMA (full curve in Fig. 1247 a) shows for the mixture the resonant dispersion behavior (broken curve) of a strong chromophore. The corresponding behavior is seen for the extinction coefficient OL (Fig. 12-47b). This material showed (after poling, see next section) decent x(2) values and could be successfully implemented into an electro-optic phase modulator. A major drawback of the host-guest systems is the limited solubility of the dyes in the polymer: typically only 15-20 wt.% can be mixed without phase segregation and crystallization. Another problem, in particular in view of device applications, arises from the typically low temporal stability of the noncentrosymmetric orien, tation distribution induced by the poling process (see below). Depending, of course, on the storage temperature, decay times of a few hours to days have been reported (Page et al., 1990). However, certain hostguest systems based on polyimide matrices have been demonstrated with stable %(2) values for hours even at temperatures as high as T = 300°C (Wu et al., 1991).
578
12 Optical Properties of Polymers 0
0 II II 2 -0-C-CH 3
1
0
16
(a) 1 ^
doped PMMA
-
\ \
PMMA ^
/
\
1.6 -
/
/
/
|
.075-
>^
^'
i
v/ — —
(b)
Figure 12-47. Example of a host (PMMA)-guest (TP, structural formula given in the inset) system, (a) The wavelength dependence of the refractive index n of pure PMMA (full line) and after addition of 15 wt.% TP (dashed line), (b) The corresponding absorption coefficient. (After Levenson et al., 1991.)
doped PMMA /
/// V\ v\ /
.05 -
PMMA
025s
0
300
400
500
600
700
800
900
X / nm
The aforementioned problems have led to the development of polymers where the chromophores were either an integral part of the main chain (Kohler et al., 1990) or, more widely reported, were attached to the host polymer as a functional side chain. A small selection of the many systems presented in the literature is given in Fig. 1248, only examples based on acrylates or methacrylate polymers are added. All these attempts were aimed at fabricating materials that can be easily processed, for example in the form of a glassy film or channel waveguide, that show a high /?D-value of the employed chromophores, that allow a high NLO moiety content, and that give a stable configuration when poled (Man and Yoon, 1992). In fact, these copolymers allowed a much higher functionalization of the material than the host-guest systems. Yet another approach recently reported (Eich etal., 1989; Jungbauer et al., 1991; Swalen et al., 1991; Chen et al., 1991) currently seems to be the ultimate solution for
a stable chromophore orientation distribution: The incorporation of the NLO moieties into the polymer network as an integral part of the cross-linked matrix. One of the systems is shown in Fig. 12-49 (Swalen et al., 1992): An epoxy monomer (Fig. 1249 a) is reacted with £-field oriented NLO active amine monomers (Fig. 12-49b) to give a highly functionalized cross-linked and hence stable x(2)-material (Fig. 12-49 c). Very good long-term stabilities have been achieved (see below). For the third-order nonlinear optical processes polymeric materials with a high degree of delocalization of the conjugated electron cloud seem to be the most promising candidates. This leads to a large component yxxxx of the molecular hyperpolarizability tensor yD along the polymer chain (conjugation) direction. Some examples, currently being studied in many laboratories, are given in Fig. 12-50 (see, e.g., Jenekhe etal., 1992, and references therein), other systems are summarized in Table
12.5 Nonlinear Optical Properties of Polymers
579
(a)
(b)
Q CH3-C-C-O-CH3 CH2_ CH3-C—C-O—CH2—CH2N -N=NCH3— CH 2 7
CH2
(c) CH 3 -C—C—0— CH2 o I II Ho C C 0
—CH3
(CHo)ov
CHo
> CH3
(d) CH2 0 HC-C-O-(CH2)2-OH CH20 HC-( I !-(OCH 2 CH 2 )3
o
Figure 12-48. Examples of different dye-substituted sidechain polymers. After: (a) Herman etal. (1991), (b) Shuto et al. (1991) and Sasaki (1992), (c) Verbiest etal. (1991), (d) Penner et al. (1991), and (e)Tomono etal. (1992).
CH3-C-C-O-CH3 CH2 CH 3 -C-C-O-CH 2 -CH 2 ^
CH2
I
"tm
CH3CH2
12-4. According to (12-85) this is one prerequisite for a high value of the macroscopic nonlinear susceptibility X(3). Another important issue also for thirdorder materials is the orientational order
of the chromophores described by
580
12 Optical Properties of Polymers
1
(a)
H
trans-Polyacetylene
H
Polydiacetylene NH2
(b) INK*1
R1
o
/"V
Poly(para)phenylene
NO 2 Amine Monomer
M
Polythiophene
n
(c)
"Rigid rod polymer"
p, Figure 12-49. Examples of a crosslinkable system composed of bifunctional matrix monomer (a) and a trifunctional /(2)-active chromophore (b). (c) Scheme of the network obtained by curing these two constituents (after Swalen et al., 1992).
Poly(para)phenylenevinylene from precursor
Figure 12-50. Structural formulas of some selected X(3)-active polymers (cf. also Table 12-4).
Table 12-4. Some conjugated polymers and their linear and nonlinear optical properties.
PPV
460
1.2 + 0.6x10" 1 0
PPA-1
R: Si(CH3)3
525
1.9 + 0.4x10" i i
PPA-2
R: CH 2 CH 3
450
8.4+1.7x10"
PPA-3
R: CH,
450
1.5 + 0.4x10" 1 1
order whereas
12
mophores which determines ND [cf. Eq. (12-85)] is another factor that needs to be optimized for a large #(3) but has to be balanced against other materials and processing parameters. For example, most of the chromophores presented in Fig. 12-50
12.5 Nonlinear Optical Properties of Polymers
are nearly intractable, that is, they are largely insoluble in most solvents and cannot be prepared, for example, in thin film form. This has led to many more or less successful routes to a final product via some soluble and hence processable precursor materials. One example is given in Fig. 12-50: here, a tetramethylene sulfonium chloride precursor polymer is converted as a spin-cast thin film by a thermal treatment to the final poly(p-phenylenevinylene) system (Lenz et al., 1988) which showed decent yD-values (Bubeck et al., 1989). Another trend in current x(3)-materials research concerns the incorporation of molecular structures with a two-dimensional 7i-electron delocalization. Some examples are given in Fig. 12-51 (Sasabe
581
et al., 1990; Schrader et al., 1991). The obtained nonresonant #(3)-values, however, are not too promising as yet (Schrader etal., 1991). 12.5.2 Second-Order Nonlinear Optics
Among the numerous device configurations based on #(2)-processes that have been realized and demonstrated in various laboratories we will only focus on two very basic concepts, that is, a polymeric waveguide structure designed for second harmonic generation (SHG) and an electrooptic modulator which is the key unit for many switching devices. Both modules require that the NLO active chromophores are highly oriented in a noncentrosymmetric narrow distribution. 12.5.2.1 Control of Orientational Order
(a)
T
CH3XCK
CH3 CH3
Annulene
Figure 12-51. Some examples of systems with a near two-dimensional 7r-configuration. (a) After Schrader et al. (1991), (b) and (c) after Sasabe et al. (1990).
For polymeric materials two strategies have been developed for obtaining oriented samples. The first is based on the LBK technique. For the usual Y-type deposition where one monolayer (A) is transferred on the down-stroke and the other (B) on the up-stroke this requires a double trough arrangement with the /(2)-active layer being transferred, for example, only on the upstroke whereas an optically inert material is deposited on the down-stroke (Beckerbauer, 1990; Penner etal., 1991). A different approach to the problem of generating noncentrosymmetric structures by the LBK method has recently been reported for molecules (2-docosylamino-5-nitropyridine) whose chromophores are arranged in a plane parallel to both dipping directions with a net polar orientation giving rise to a large x(2)-activity (Bosshard et al., 1991). If two different types of chromophores with opposite pD components are used for the A- and for the B-layer, respectively, then it is possible to prepare
582
12 Optical Properties of Polymers
LBK multilayer assemblies in the Y-mode with a very high chromophore density (Cresswell et al., 1990). In passing we should note that the self-assembly multilayers with their intrinsically noncentrosymmetric arrangement have also been successfully introduced for x (2) ~ act i ve structures (Allan et al., 1991). All these systems, however, have in common that in addition to the NLO active chromophores highly specialized molecular moieties need to be incorporated in order to tune the molecule's properties to the desired character (e.g., to allow for a self-assembly process) (Bubeck et al., 1991). From a technical point of view, it is also highly unlikely that such a time-consuming process as the LBK deposition of multilayers will ever make it to a commercially competitive device level. Therefore, the alternate strategy, that is, starting from a functionalized polymer with random chromophore orientation spin-coated onto a solid substrate and then subsequently oriented by a poling process, is certainly the more promising and hence more widely applied approach. In the following we will briefly review some aspects of the poling process and its stability. Some of the basic experimental configurations used for the static E-field induced noncentrosymmetric orientational order of the chromophores in functionalized polymer thin films are summarized in Fig. 12-52. The complete EO waveguide structure shown in Fig. 12-52 a includes in addition to the NLO active core the two buffer layers that isolate the guided mode from any absorption losses in the (metallic) electrodes. The latter will be used in the final device as the control electrodes but can also be used to apply a DC field of the order of some 10 MV/cm during the poling process at elevated temperatures (see below). A simplified version optimized for mi-
croscopic characterization of planar waveguide structures (cf. Sec. 12.4.2.1 and Fig. 12-31 a) is given in Fig. 12-52b. These two versions for typically out-of-plane orientation of the chromophores are complemented by an in-plane poling set-up schematically depicted in Fig. 12-52c. And finally, the top electrode free poling process based on a corona discharge set-up is sketched in Fig. 12-52d. At present, it is not yet clear which approach - electrode or corona poling - will yield the better results, that is, produce a higher degree of orientation, better homogeneity, tolerance against defects, etc. (Herminghaus et al., 1991). In any case, the poling process requires an orientational (rotational) flexibility of the chromophores in the polymer matrix. This is achieved by heating the sample to a temperature close to the glass transition temperature Tg of the material. Details of the actual temperature-electric field (applied voltage) protocol differ from laboratory to laboratory but certainly depend also on whether one is dealing with a hostguest, main chain or side chain polymer or a crosslinkable system that can be cured by a thermal process. For the latter case, a typical protocol is given in Fig. 12-53 (Swalen et al., 1991). Without any voltage applied, the system is heated to a temperature where crosslinking is initiated leading to a slight increase of Tg. This precuring period is followed by the actual poling process which in the presence of the orienting field further crosslinks the polymer-chromophore network with a concomitant increase of the glass transition temperature. The poling process can be followed online by recording the SHG or electro-optic efficiency or, very simply, by monitoring the linear absorption data. This is exemplified in Fig. 12-54. The extinction coefficient a is plotted as a function of the wavelength measured for a thin polymer film in trans-
12.5 Nonlinear Optical Properties of Polymers
583
(a)
Top Electrode
- Buffer Layer EO-active Polymer -Buffer Layer -Electrode -Substrate
Detector
Electrode EO -active Waveguide Electrode
(c)
top view
side view
Electrodes Polymer Film Substrate
(d)
Needle Electrode
Metallic Grid Ugrid
Polymer Film Conducting Layer • Substrate Heated Copper Block
Figure 12-52. (a) Schematic cross-section of a polymer layer with electro-optic activity sandwiched between two buffer layers and two electrodes. This configuration can be used for (out-of-plane) poling and for (simultaneous) waveguide spectroscopy. (b) Simplified version of (a), sufficient for many characterizations of the EO active polymer, (c) In-plane poling, (d) Scheme of a set-up used for corona (discharge) poling.
584
12 Optical Properties of Polymers
Tg CJ
cure
Temperai
CD
1 *
Vol tage
precure
*cure
•
*
-
•
Figure 12-53. The temperature-voltage protocol for a dye-polymer system that undergoes a thermally induced cross-linking reaction with a simultaneous increase of the glass transition temperature Tg (after Swalen et al., 1991).
time
mission prior to (Fig. 12-54 a) and after the poling process (Fig. 12-54b). The decrease of the optical density normal to the film is caused by the out-of-plane poling that orients the chromophores preferentially normal to the film. This optical anisotropy can be used to derive information on the orientational distribution function characterized by an order parameter 0S (Page et al., 1990) & = 1-—
(12-86)
0.4"
a 0.30.2" 0.1 " 0.0 400
500
600
700
X / nm Figure 12-54. Change of the absorption of a thin dyedoped polymer film measured in transmission before (a) and after (b) out-of-plane poling (after Chen et al., 1991).
with a± being the extinction normal to the film after poling and a0 the extinction prior to the poling assuming an isotropic orientation distribution of the chromophores. Once the polymer sample is cooled down again to room temperature, this polar anisotropy and hence the noncentrosymmetric x(2)-activity is frozen in, however, not completely, depending on the system. For device purposes this is, of course, an issue of utmost importance and hence has been the subject of many studies. With some exceptions (discussed above already), the general behavior can be classified as shown in Fig. 12-55: the host-guest systems show the fastest back-relaxation to a more or less isotropic distribution. Much better results are obtained with chromophores covalently attached to the polymer backbone although clearly this reduced flexibility also may result in a reduced net orientation obtainable through poling. The best stability is found for the crosslinked system. It should be pointed out that most of these results are obtained, at least to some extent, by a trial-and-error approach.
585
12.5 Nonlinear Optical Properties of Polymers
(a) 100 poled + crosslinked
o i
0)
copolymers
50-
host - guest
Poled Polymer time
Figure 12-55. Qualitative stability behavior of poled polymer films. Typically, the host-guest mixed system randomizes very fast; the copolymer system has a high residual orientation even after long times; the cross-linked system has a long-term stability.
(b)
7.9 5.9-
ZJ CD
O
3.9-
CO
1.9-
Many of the details are not understood yet and in many cases only very crude models have been treated. What is certainly lacking is a better understanding of the correlation of these linear and nonlinear optical properties with other polymer-specific characteristics given by the structural and dynamical features of these functionalized materials. 12.5.2.2 Second-Harmonic Generation (SHG) One of the key device applications of X(2)-active polymers is second-harmonic generation. The driving force is the search for efficient blue light sources, in this case through frequency doubling of red laser diode emission, for example, for the use in optical storage concepts (cf. Sec. 12.4.3). The basic configuration is given in Fig. 1256 a. Laser light of frequency co, incident on the sample of thickness d at an angle 9 with a fixed polarization, is refracted at the surface of the nonlinear medium and propagates through the film. The second harmonic wave follows the fundamental beam and as such is called the bound wave (Jer-
-0.1 -70
-40
-10
20
50
80
0/deg
(c) ZJ CD
O CO
-70
-50
-30
-10 0 10
30
50
70
0/deg Figure 12-56. (a) Schematic representation of the propagation vectors whose interference leads to the Maker fringe pattern, (b) Maker fringes for a thick polymer film, (c) One Maker fringe for a thin poled polymer film. (After Swalen et al., 1992.)
phagnon and Kurtz, 1970). In addition, a second harmonic wave is generated at the surface of the film which is, however, refracted at a different angle because the refractive index n(2co) of the NLO material is different from the value at the fundamental n(co). This wave is called the free wave.
12 Optical Properties of Polymers
Both waves interfere to give the overall SH intensity with an oscillatory pattern as one rotates the sample called the Maker fringes (Ledoux et al., 1987; Verbiest et al., 1991). Depending on the thickness of the sample several minima and maxima can be observed [Fig. 12-56 b for a thick film, Fig. 12-56c for a thin film, after Swalen et al. (1992)]. Now, the required high laser light intensity for efficient nonlinear conversion is best realized in a waveguide configuration with strong optical field confinement. The ultimate structure, therefore, will be the second harmonic generating channel waveguide (Norwood and Khanarian, 1990). However, the dispersion of the material and of the waveguide structure (modal dispersion) would result in a useless device because after a certain propagation length of the fundamental and the second harmonic beam (called the coherence length Lc), their relative phase difference would lead to destructive interference of the second harmonic light. Several concepts have been put forward to overcome this problem called the phase-matching problem. The first approach is based on the modal dispersion of waveguide structures: This term accounts for the fact that different modes (modes of different order) propagate with different phase velocities, hence can be described by different effective refractive indices. Phase-matching in such a concept is achieved by tuning the dimensions of the waveguide, for example, the thickness of a planar structure, in such a way that the dispersion of the material (cf. Fig. 12-6) is compensated by the modal dispersion. This then results in a configuration in which the guided fundamental mode of order m propagates with the same velocity as the second harmonic mode of order m'. Problems arising from the need to optimize the field intensity overlap integrals
have so far hampered a major breakthrough of this approach. The second concept called quasi-phase matching is schematically depicted in Fig. 12-57. After the first coherence length in the poled polymer waveguide the x(2)-activity is either destroyed (by bleaching or randomization of the chromophores) or reversed in its polar direction (by the use of correspondingly patterned and poled electrodes) (Fig. 12-57 a). The first configuration is called a unidirectional poled waveguide, the second, bidirectional; both overcome the problem of destructive interference of the SHG light, with the bidirectionally poled structure showing a twice as efficient build-up of the 2co signal (Fig. 1257 b). Finally, we should mention a third version called the Cerenkov frequency doubler. Here, the refractive indices of waveguide material and substrate are such that the fundamental light is guided in the film, but the SH light is constantly coupled out through the substrate (because the condi(a)
Poled Waveguide, unidirectional
\J|f
\\f\^\l\\H\\&^\lV\/^J.\\if bidirectional
n\\l\\/[tHVJlff\ f\JV/\\ffi\Urt\\f f (b)
bidirectional SHG Build-up
586
/ ^ /*
/ ^
unidirectional
Figure 12-57. (a) Scheme of unidirectional and bidirectional poling of waveguides for quasi-phasematching and (b) the resulting build-up of secondharmonic light in the waveguide.
12.5 Nonlinear Optical Properties of Polymers
587
Detector
Transparent Substrate Transparent Electrode Polymer Film Reflecting Electrode
tion for total internal reflection at the filmsubstrate interface is not fulfilled at 2co). In this way, also no destructive interference limits the SHG efficiency. 12.5.2.3 Electro-Optic (EO) Modulation The second major class of x(2)-based devices use the Pockels effect, that is, the manipulation of the refractive index of the material by an applied electric (DC) field. The ultimate structure, again, is certainly a channel waveguide (Shuto et al., 1991) with an architecture schematically shown in Fig. 12-52 a. If only materials parameters are to be measured various simplified architectures are sufficient. One set-up operating in a reflection mode similar to ellipsometry is given in Fig. 12-58 (Teng and Man, 1990). The highly reflecting Au or Al electrode on the rearside of the EO active thin film is used to measure differences of the reflectivities for s- and p-polarized light propagating through the poled polymer in the absence and presence of an applied E-field. The differences of the reflected light intensities can be quantitatively analyzed in terms of the two / ( 2 ) tensor components, x?3 anc * X33> needed to describe the nonlinear (Pockels) response of a uniaxial film (Gadret et al., 1991). The advantage of such a simple set-up for NLO materials characterization relative to a SHG set-up are obvious: No high power
Figure 12-58. Set-up for EO coefficient measurement in reflection.
laser is needed, and, therefore, wavelengthdependent measurements are much easier. A more detailed analysis of the linear and nonlinear optical properties of planar thin film samples is possible by surface plasmon (Cross et al., 1988; cf. also Sec. 12.4.1) and by waveguide spectroscopy outlined in Sec. 12.4.2.1. This has been demonstrated for many samples of poled polymers. In particular, if electro-strictive effects are also present in the investigated material only the waveguide technique allows a separation of the different processes (Dumont and Levy, 1989). It is straightforward to demonstrate the potential of the method for a waveguide structure prepared by the LBK technique. The sample consisted of 240 buffer layers of an inert material on each side of the EO active A-B layer system of 20 layers each. The structural formula of the chromophore-containing LBK systems is given in the inset. The experimental set-up was as in Fig. 12-52b. Figure 12-59 a shows the three waveguide modes (m = 0, 1, 2 as indicated) measured with s-polarized light {X = 593 nm). Figure 12-59b gives the electro-optic response of the sample under the influence of an applied voltage of 30 V peak-to-peak. Most remarkable is the alternating magnitude of the effect: The m = 1 mode shows only a very small modulation. The interpretation is obvious on the basis of the field intensity distribution given in Fig. 12-29.
588
12 Optical Properties of Polymers
according to yDocLad ~
0.5
(a)
CH3-(CH2)17-0-<5>-CH=N-NH-<5>-N02
Different theories give for ID systems oc = 5 to 7 (Rustagi and Ducuing, 1974; Agraval et al., 1978; Chopra et al., 1989; Grossman et al., 1989). For example, treating the conjugated polymers as one-dimensional systems with a 7i-electron delocalization length L d , Agrawal et al. (1978) and Flytzanis (1987) derived a scaling law for the third-order susceptibility X(3)ocL6d
8/deg
60
Figure 12-59. Waveguide mode pattern (a) and electro-optic response by applying a 30 V peak-to-peak voltage (b) across a LBK multilayer assembly composed of 240 buffer layers/40 A-B layers (with B being the #(2)-active material)/240 buffer layers. Structure of the NLO active dye is given in the inset (Aust et al., 1993).
The m = 1 mode has a node in the middle of the waveguide right where the EO active material is located. Any refractive index modulation by the applied electric field influences, therefore, the overall index configuration for this mode only very little. The even-indexed modes, however, with their intensity maximum in the middle, are influenced much more strongly. The quantitative analysis gives for all three modes the same electro-optic coefficient x{il = 14pm/V. 12.5.3 Third-Order Nonlinear Optics The molecular hyperpolarizability yD of one-dimensional (ID) conjugated polymers depends on the conjugation length L d
(12-87)
(12-88)
If L d is taken to be inversely proportional to the optical gap then x{3) should scale with A^ax with Amax being the spectral position of the linear absorption. This is qualitatively found for a wide range of polymers (Fig. 12-60) although for certain systems true nonresonant #(3)-values would be required for a rigorous testing of this theoretical approach (Zhao et al., 1988; Neher et al., 1990). In view of possible applications of these materials for all-optical information processing where (nonresonant) X(3)-values in the range of 10" 9 esu or larger are estimated to be required, it is, however, highly doubtful whether linearly polyconjugated molecules allow this nonresonant #(3)-value to be reached at all. At present, it is an open question whether materials with an improved nonlinear response and/or an optimized figure of merit will be available in the future. These figures of merit are defined such as to scale the enhancement of a desired nonlinear susceptibility, for example, the intensity dependent refractive index change in the sample n2 [cf. Eq. (12-62)], against other linear or nonlinear response parameters. For example, in a nonlinear directional coupler (cf. Fig. 12-35 a) the (desired) maxi-
12.5 Nonlinear Optical Properties of Polymers
589
10-12 700 /nm
Figure 12-60. Survey of #(3)-values as determined from THG experiments. In addition to values taken from Table 12-4, data from the literature are plotted: polydiacetylene (PDA) and polyacetylene (PA) from Neher et al. (1990); poly(3-decyl thiophene)s (PTs) are from Salcedo et al. (1987). x(3)-values are plotted as a function of the spectral position Xmax of the linear absorptions. The higher %(3)-values found for PPV are explained as a consequence of the higher 7i-electron density due to the lack of bulky substituents (Neher et al., 1990).
mum phase shift A>NL is given by (12-89) with L being the interaction (device) length. For a given n2 of the material, the required phase shift of the device could be obtained simply by choosing L correspondingly (high). However, the maximum value of L is limited by the absorption in the polymer: max {L} < a" 1 . Since a = a 0 + /J2 • /, that is, a contains not only the linear absorption term a0 but also the nonlinear, for example two-photon, absorption term P2 -1 which is also intensity dependent, one may define (even in the absence of linear absorption oc0 < /?2 • I) a figure of merit Tm = 2k^
(12-90)
which must be smaller than 1 for a material to be used in a nonlinear directional coupler. On the other hand, many of these parameters are wavelength dependent. This
defines the need for a rigorous materials characterization and optimization program for those research activities committed to the development of photonic devices based on polymeric systems. Firstly, of course, X(3) has to be determined. One experimental technique for this purpose is third-harmonic generation (THG), which will be briefly presented in the next section. Of interest are not only the contributions from electronic or thermal processes or from excited states but, in particular, the spectral dispersion behavior of #(3). This must also be known for the linear and, equally important, for the nonlinear absorption coefficients. In fact, it is currently one of the promising strategies to find spectral windows where the device performance is not limited by the two-photon (or three-photon) absorption in the materials. Since one is interested in fast data processing devices, with switching times T < 0.1 ps, the nonlinear recovery time is of vital importance. One experimental technique for its determination is degenerate
590
12 Optical Properties of Polymers
four wave mixing (DFWM) briefly outlined below. We should point-out, once again, that in addition to these (linear and nonlinear) optical parameters, other material properties, for example stability and processibility into device configurations such as waveguides, are equally important for a successful application. In the last section we then give an example of a nonlinear directional coupler fabricated from polydiacetylene as a basic unit for the photonic concept of all-optical switching. 12.5.3.1 Third-Harmonic Generation (THG) Third-harmonic generation describes the process in which an incident photon field of frequency co generates, through nonlinear polarization in the medium, a coherent optical field at 3 co. It gives directly the electronic hyperpolarizability with ultrafast response time. The typical experiment performed with thin film polymer samples prepared on suitable substrates either in transmission or reflection geometry monitors contributions from the substrate as well as from the film. Rotating the sample, as outlined above for the SHG experiments, leads to a coherent superposition of the free and bound waves, giving rise to a fringe pattern similar to the one shown in Fig. 12-56 for the SHG signal. If one knows the substrate's cubic susceptibility, the experimental data can yield both the real and imaginary parts (or the modulus and the phase) of the third-order susceptibility of the polymer material. If suitable laser systems are available these data can be obtained covering a wide range of frequencies, allowing the search for suitable spectral windows off the one-, two-, and three-photon resonances. One example is given in Fig. 12-61 for a polydiacetylene system (after Guo et al., 1992). The experi-
1000
1500 X/nm
2000
Figure 12-61. Modulus (a) and phase (b) of x(3) measured for polydiacetylene (cf. Fig. 12-50, with Rx = R2 = {(CH2)4OCONHCH2COO(CH2)3CH3}) as a function of the fundamental wavelength. The full curves are theoretical calculations based on a model with 4 essential states (after Guo et al, 1992).
mental data show the strong spectral dependence of both the modulus and the phase of #(3). The full curve is a theoretical description of the nonlinear susceptibility of the polymer on the basis of an "essential states" picture with four levels (Guo et al., 1992) which gives for these and other data satisfying agreement between experiment and theory. 12.5.3.2 Degenerate Four Wave Mixing (DFWM) Degenerate four wave mixing is a holographic technique in which the output of a laser system is split into three beams. Two of them coherently interfere within the thin film sample, thus generating a phase grating (refractive index grating) through the intensity-dependent refractive index change n2 in the Kerr medium. The third
12.5 Nonlinear Optical Properties of Polymers (a)
sample
3
©
LBK - film (50 layers) of polymer
*)
®
LBK - film (80 layers) of monomer
Cl
©
Polystyrene - film doped with monomer
Cl
®
Poly (styrene - co - phthalocyanine) 0 - C H 3
R
R2
R
3
591
R' R
3
Ri = O-CH3, R 2 = 0 - CsHi7 in all permutations
CH 3
CH 3
Si ~
(c)
A.
A.
3
•—^_
2
A. 600
700
Intens ity/a.u.
(b)
CH 3
r ..K
2
1 . 800
0
1 10
20
Delay Time/ps X/nm Figure 12-62. (a) Chemical structure of phthalocyanine derivatives and sample configurations used for DFWM experiments, (b) Absorption spectra and (c) decay-curves of transient grating of some phthalocyanine thin films: (1) LBK film of 50 layers of a polymeric sample; (2) LBK film of monomeric dye; (3) polystyrene, doped with monomer; (4) styrenephthalocyanine-copolymer.
beam probes this grating: Its diffraction efficiency (the 4th wave!) is monitored during the experiment. If short laser pulses are used, time-resolved experiments can be performed: If the third pulse is slightly delayed relative to the two interfering pulses the diffracted intensity is a measure of the grating decay. Thus, information about the important nonlinear recovery time can be obtained. (By comparison with a material
of known #(3) also the magnitude of the third-order nonlinear susceptibility can be obtained.) Examples are given in the literature of DFWM experiments performed with 60 fs pulse duration, showing the ultrafast response times possible with polymeric systems (Prasad, 1991). Figure 12-62 summarizes some data aimed at demonstrating the influence of electronic coupling between different chromophores. The close interrelation between the detailed molecular structure of the chromophore, the supramolecular architecture of the matrix into which it is incorporated, and the linear and nonlinear optical properties of the resulting functionalized material is demonstrated for various thin film samples of phthalocyanines. These dyes are of general interest because of their excellent light and temperature stability. The chemical structure and a few details of the different systems used in these studies are given in Fig. 12-62 a. In system (4) the phthalocyanine moieties were incorporated into a copolymer with styrene. Figure 12-62b shows the linear absorption spectra and (Fig. 12-62c) the decay curves as obtained from DFWM experiments at X « 650 nm. Remarkable differences for the various systems are found (Kaltbeitzel et al., 1989). An isolated phthalocyanine
592
12 Optical Properties of Polymers
molecule as, for example, in system (4) gives a sharp absorption spectrum. This leads to a very slow decay time of the transient grating beyond the resolution of the set-up. Increasing aggregation eventually leads to an inhomogeneous broadening of the absorption band and a concomitant reduction of the transient time (by orders of magnitude!). Obviously, the electronic coupling between individual resonators affects both the linear and the nonlinear response of a chromophore system. Various contributions to the observed phenomena may include (i) fluorescence life time reduction by aggregational quenching, (ii) energy migration (Salcedo et al., 1987), for example exciton diffusion into the nonilluminated area, (iii) energy trapping at lower states, (iv) bimolecular exciton quenching (Ho and Peyghambarian, 1988), etc. The response times of the conjugated polymers described above are presumably also related to the relaxation processes (iii) and (iv). The extended 7i-conjugation with a tight electronic coupling via intra- and interchain energy migration leads to the ultrafast transient times of these systems, even near resonances. 12.5.3.3 All-Optical Switching
In this last section we just touch on a still very speculative subject, namely controlling light by light (through x(3)-active materials). Various device concepts have been proposed, some of which have also been experimentally tested. A very basic unit is the nonlinear directional coupler. Its realization was reported by Townsend et al. (1989). However, it turned out that the intensity-dependent switching of the input laser pulse from one output channel to the other was based on a thermal effect attributed to a three-photon absorption. Recently, a similar structure was presented by
Kaneko et al. (1992) who claimed on the basis of their experiments with Ti: A12O3 laser pulses of 100 fs duration at X = 825 nm an electronic mechanism for the observed switching behavior. The high pulse repetition rate still means a substantial power deposition in the sample and it will take single-pulse experiments to show unambiguously the fast electronic nature of the all-optical switching phenomena seen in these polymer waveguide samples. 12.5.4 Photorefractive Polymers
The photorefractive effect was first established in LiNbO 3 (Chen, 1967). It arises when charge carriers, photogenerated by a spatially modulated illumination, separate and become trapped to produce a nonuniform space charge distribution. The resulting internal field then modulates the local refractive index via the linear electro-optic effect, which results in a phase image or hologram of the illumination pattern. A photorefractive material, therefore, needs to combine some functional units that act as photoionizable charge carriers, some that transport the photogenerated charges until they are trapped at certain sites, and it has to show a nonlinear electro-optic change of its refractive index upon a spacecharge field. Many possible applications have been suggested, including high-density optical data storage, various imageprocessing techniques, phase conjugation, and programmable interconnection (Giinter and Huignard, 1988). Polymers again would be ideal materials for photorefractive devices because of their excellent processability and potentially high electro-optic coefficients (see above). Also, photoconductivity in functionalized polymers is a well-established phenomenon because of its importance for the photocopy process (Burland and Schein, 1986)
12.8 References
and for laser-printing. It was nevertheless not until very recently that the first reports on photorefractive polymers were published (Schildkraut, 1991; Ducharme et al., 1991; Tamura et al., 1993) showing indeed very promising features which certainly will stimulate a very dynamic research activity in this field.
12.6 Outlook Polymers with their broad variety of optical properties have already their established place in modern products and technologies. The needs of our information society and the emerging technology of the next century, photonics, means a rapidly growing demand, chance, and challenge for ever better performing polymers with specifically tuned optical (and other) properties. Optical data storage for consumer products and in information technologies will be a growing area for the use of more and more sophisticated materials. Closely linked to this development is the requirement for a cheap yet powerful blue laser light source. At present it is an open race whether the blue solid state laser diode or a device based on a red laser diode with an integrated frequency doubling (polymeric) material will be available first. Integrated optics, in general, will have a growing share in data transmission, distribution, and processing. Here, a great deal will depend on whether improved nonlinear optical materials will be available. These ends will be met only if, by an interdisciplinary effort, a better understanding of the various materials parameters at the molecular level can be obtained. The close relation between macroscopic properties and functions on the one hand and the microscopic aspects of structure, dynamics, and order on the other, will re-
593
main a continuous challenge for modern materials science and technology.
12.7 Acknowledgements It is my pleasure to thank many colleagues and friends for their support and for many helpful discussions. In particular, Y. Koike, H. Kuhn, Y. Levy, W. H. Meyer, V. Mizrahi, M. R. Philpott, H. Ringsdorf, H. Sasabe, K. Sasaki, G. I. Stegeman, X Stumpe, I D. Swalen, C. Urban, G. Wegner, T. Wada, and J. Yang contributed also to this chapter through their stimulating comments. As to our own results, I am particularly grateful to my coworkers E. F. Aust, C. Duschl, W. Hickel, H. Motschmann, M. Sawodny, R. Reiter, and B. Rothenhausler. Financial support came from the Deutsche Forschungsgemeinschaft, the Bundesministerium fur Forschung und Technologie, the Volkswagen Stiftung and the Leonhard-Lorenz-Stiftung.
12.8 References Agrawal, G. P., Cojan, C , Flytzanis, C. (1978), Phys. Rev. B17, 116. Allan, D. S., Kubota, K, Marks, T. I, Zhang, T. X, Liu, W. P., Wong, G. K. (1991), Nonlinear Optical Properties of Organic Materials IV, Proc. SPIE 1560, 362. Aust, E., Hickel, W., Knobloch, H., Knoll, W. (1993), /. Appl. Physics, in press. Blodgett, K. B. (1935), J. Am. Chem. Soc. 57, 1007. Born, M., Wolf, E. (1980), Principles of Optics. Oxford: Pergamon Press. Bosshard, C , Kiipfer, M., Florsheimer, M., Giinter, P. (1991), Nonlinear Optical Properties of Organic Materials IV, Proc. SPIE 1560, 344. Bowden, M. J. (1988), in: Adv. Chem. Ser. 218: Electronic and Photonic Applications of Polymers: Bowden, M. J., Turner, S. R. (Eds.). Washington, DC: Am. Chem. Soc. Bubeck, C , Kaltbeitzel, A., Lenz, R. W, Neher, D., Stenger-Smith, J. D., Wegner, G. (1989), in: Nonlinear Optical Effects in Organic Polymers. Dordrecht, Netherlands: Kluwer Academic Publ., p. 143.
594
12 Optical Properties of Polymers
Bubeck, C , Laschewsky, A., Lupo, D., Neher, D., Ottenbrei, P., Paulus, W, Prass, W, Ringsdorf, H., Wegner, G. (1991), Adv. Mater. 3, 54. Burland, D. M., Schein, L. B. (1986), Phys. Today, May, 46. Burland, D. M., Rice, J. E., Downing, X, Michl, J. (1991), Nonlinear Optical Properties of Organic Materials IV, Proc. SPIE 1560, 184. Burstein, E., Chen, W. P., Chen, Y. X, Hartstein, A. (1974), /. Vac. Sci. Technol. 11, 1004. Chang, K. (Ed.) (1991), Handbook of Microwave and Optical Components, Vol. 4: Fiber and Electro-optical Components. New York: John Wiley and Sons. Chen, F. S. (1967), J. Appl Phys. 38, 3418. Chen, Y M., Mandal, B. K., Lee, J. Y, Miller, P., Kumar, J., Tripathy, S. (1991), Mater. Res. Soc. Symp. Proc. 214. Boston, MA: Mater. Res. Soc, p. 35. Chopra, P., Carlacci, L., King, H. K, Prasad, P. N. (1989), /. Phys. Chem. 93, 7120. Cresswell, J. P., Tsibouklis, I, Petty, M. C , Feast, W. J., Carr, N., Goodwin, M., Lvov, Y M. (1990), Nonlinear Optical Properties of Organic Materials III, Proc. SPIE 1337, 358. Cross, G. H., Peterson, I. R., Girling, I. R., Cade, N. A., Goodwin, M. J., Carr, N., Sethi, R. S., Marsden, R., Gray, G. W, Lacey, D., McRoberts, A. M., Scrowston, R. M., Toyne, K. J. (1988), Thin Solid Films 156, 39. Ducharme, S., Scatt, J. C , Twieg, R. J., Moerner, W. E. (1991), Phys. Rev. Lett. 66, 1846. Duda, G., Schouten, A. I, Arndt, T, Lieser, G., Schmidt, G. F., Bubeck, C , Wegner, G. (1988), Thin Solid Films 159, 221. Dumont, M., Levy, Y (1989), in: Springer Proc. Phys., Vol. 36: Nonlinear Optics of Organics and Semiconductors: Kobayashi, T. (Ed.). Berlin: Springer-Verlag, p. 256. Eich, M., Reck, B., Yoon, D. Y, Willson, C. G., Bjorklund, G. C. (1989), J. Appl. Phys. 66, 3241. Flytzanis, C. (1987), in: Nonlinear Optical Properties of Organic Molecules and Crystals, Vol. 2: Chemla, D. S., Zyss, X (Eds.). Orlando, FL: Academic Press. Franke, H., Festl, H. G., Kratzig, E. (1984), Colloid Polym. Sci. 262, 213. Gadret, G., Kajzar, F., Raimond, P. (1991). Nonlinear Optical Properties of Organic Materials IV, Proc. SPIE 1560, 266. Goos, F , Hanchen, H. (1947), Ann. Physik 1, 333. Gordon II, J. G., Swalen, J. D. (1977), Opt. Commun. 22, 374. Groh, W, Lupo, D., Sixl, H. (1989), Adv. Mater. 101, 1580. Grossman, C , Heflin, J. R., Wong, K. Y, ZamaniKhamiri, O., Garito, A. F. (1989), in: NATO ASI Ser: Appl. Sci., Vol. 162: Nonlinear Optical Effects in Organic Polymers: Messier, X, Kajzar, F , Prasad, P., Ulrich, D. (Eds.). Dordrecht, Netherlands: Kluwer Academic Publ., pp. 225-245.
Gunter, P., Huignard, X-P. (Eds.) (1988), Photorefractive Materials and Their Applications. Berlin: Springer-Verlag. Guo, D., Mazumdar, S., Stegeman, G. I., Cha, M., Neher, D., Aramaki, S., Touellas, W, Zanoni, R. (1992), Mater. Res. Soc. Symp. Proc. 247. Boston, MA: Mater. Res. Soc, p. 151. Henon, S., Meunier, X (1991), Rev. Sci. Instrum. 62, 936. Herman, W. N., Rosen, W A., Sperling, L. H., Murphy, C. X, Jain, H. (1991), Nonlinear Optical Properties of Organic Materials IV, Proc. SPIE 1560, 206. Herminghaus, S., Smith, B. A., Swalen, X D. (1991), /. Opt. Soc. Am. B8, 2311. Hickel, W, Knoll, W (1990a), J. Appl Phys. 67, 3572. Hickel, W, Knoll, W (1990 b), Appl. Phys. Lett. 57, 1286. Hickel, W, Kamp, D., Knoll, W. (1989), Nature 339, 186. Hickel, W, Duda, G., Jurich, M., Krohl, T, Rochford, K., Stegeman, G. I., Swalen, X D., Wegner, G., Knoll, W (1990), Langmuir 6, 1403. Ho, Z. Z., Peyghambarian, N. (1988), Chem. Phys. Lett. 148, 107. Honig, D., Mobius, D. (1991), /. Phys. Chem. 95, 4590. Hsiung, H., Rodriguez-Parada, X, Beckerbauer, R. (1991), Chem. Phys. Letters 182, 88. Hunsperger, R. G. (1984), Integrated Optics: Theory and Technology. Berlin: Springer-Verlag. Jenekhe, S. A., Osaheni, X A., Meth, X S., Vanherzeele, H. (1992), Chem. Mater. 4, 683. Jerphagnon, X, Kurtz, S. K. (1970), J. Appl. Phys. 41, 1667. Jungbauer, D., Reck, B., Twieg, R., Yoon, D. Y, Willson, C. G., Swalen, X D. (1991), Appl. Phys. Lett. 56, 2610. Kajzar, F (1992), in: Guided Wave Nonlinear Optics: Ostrowsky, D. B., Reinisch, R. (Eds.). Dordrecht, Netherlands: Kluwer Academic Publ. Kaltbeitzel, A., Neher, D., Bubeck, C , Sauer, T., Wegner, G., Caseri, W. (1989), in: Electronic Properties in Conjugated Polymers HI: Kuzmany, H., Mehring, M., Roth, S. (Eds.). Berlin: Springer-Verlag. Kampf, G. (1985), Ber. Bunsenges. Phys. Chem. 89, 1179. Kaneko, A., Ito, A., Kuwabara, T, Sasaki, S., Sasaki, K., Sokoloff, X P., Wada, T, Sasabe, H. (1992), Sen-i Gakkai Symp. Preprints B118. Knobloch, H., Duschl, C , Knoll, W (1989), J. Chem. Phys. 91, 3810. Knoll, W (1991 a), MRS Bulletin XVI, 29. Knoll, W (1991b), Makromol. Chem. 192, 2827. Kohler, W, Robello, D. R., Dao, P. T., Willand, C. S., Williams, D. X (1990), /. Chem. Phys. 93, 9157. Koike, Y (1991), Polymer 32, 1131. Kretschmann, E. (1972), Opt. Commun. 6, 185.
12.8 References
Kuhn, H., Mobius, D., Biicher, H. (1972), in: Physical Methods of Chemistry: Weissberger, A., Rossiter, B. W. (Eds.). New York: John Wiley and Sons. Ledoux, I., Josse, D., Vidakovic, P., Zyss, J., Hann, R. A., Gordon, P. R, Bothwell, B. D., Gupta, S. K., Allen, S., Robin, P., Chustaing,, E., Dubois, J. C. (1987), Europhys. Lett. 3, 803. Lee, D. L. (1986), Electromagnetic Principles of Integrated Optics. New York: John Wiley and Sons. Lenz, R. W, Han, C. C , Stenger-Smith, J. D., Karasz, F. E. (1988), J. Polymer Sci. A 26, 3241. Levenson, R., Liang, X, Toussaere, E., Zyss, J. (1991), Nonlinear Optical Properties of Organic Materials IV, Proc. SPIE 1560, 251. Man, H.-T., Yoon, H. N. (1992), Adv. Mater. 4, 159. Maoz, R., Netzer, L., Gun, X, Sagiv, X (1988), /. Chem. Phys. 85, 1059. Mathy, A., Simmrock, H.-U., Bubeck, C. (1991), /. Phys. D: Appl. Phys. 24, 1003. Messier, X, Kajzar, R, Prasad, P., Ulrich, D. (Eds.) (1989), NATO ASI Series E: Appl. Sci., Vol. 162: Nonlinear Optical Effects in Organic Polymers. Dordrecht, Netherlands: Kluwer Academic Publ. Meyer, W. H., Pecherz, X, Mathy, A., Wegner, G. (1991), Adv. Mat. 3, 153. Miller, R. D., Michl, X (1989), Chem. Rev. 89, 1359. Mohlmann, G. R., Horsthuis, W. H. G., McDonach, A., Copeland, M. X, Duchat, G., Fabre, P., Diemer, M. B. X, Trommel, E. S., Suyten, F. M. M., Van Tomme, E., Baquero, P., Van Daele, P. (1990), Nonlinear Optical Properties of Organic Materials III, Proc. SPIE 1337, 215. Muto, S., Ito, H. (1992), Sen-i Gakkai Symp. Preprints, B-17. Muto, S., Seki, N., Ichikawa, S., Ito, H. (1991), Opt. Commun. 81, 273. Neher, D., Kaltbeitzel, A., Wolf, A., Bubeck, C , Wegner, G. (1990), in: Conjugated Polymeric Materials: Opportunities in Electronics, Optoelectronics, and Molecular Electronics: Bredas, X L., Chance, R. R. (Eds.). Dordrecht, Netherlands: Kluwer Academic Publ. Norwood, R. A., Khanarian, G. (1990), Electron. Lett. 26, 2105. Page, R. H., Jurich, M. C , Reck, B., Sen, A., Twieg, R. X, Swalen, X D., Bjorklund, G. C , Willson, C. G. (1990), /. Opt. Soc. Am. B7, 1239. Pearson, X M. (1988), in: Adv. Chem. Ser. 218: Electronic and Photonic Applications of Polymers: Bowden, M. X, Turner, S. R. (Eds.). Washington, DC: Am. Chem. Soc. Penner, T. L., Armstrong, N. X, Willand, C. S., Schildkraut, X S., Robello, D. R. (1991), Nonlinear Optical Properties of Organic Materials IV, Proc. SPIE 1560, 377. Prasad, P. N. (1991), Polymer 32, 1746. Prasad, P. N., Williams, D. X (1991), Introduction to Nonlinear Optical Effects in Molecules and Polymers. New York: John Wiley and Sons.
595
Raether, H. (1977), in: Physics of Thin Films, Vol. 9: Hass, G., Francome, M. H,, Hoffmann, R. W (Eds.). New York: John Wiley and Sons. Raether, H. (1988), in: Springer Tracts in Modern Physics, Vol. I l l : Surface Plasmons on Smooth and Rough Surfaces and on Gratings. Berlin: SpringerVerlag. Reiter, R., Motschmann, H., Orendi, H., Nemetz, A., Knoll, W. (1992), Langmuir 8, 1784. Roberts, G. (Ed.) (1990), Langmuir-Blodgett Films. New York: Plenum Press. Rothenhausler, B., Knoll, W. (1988), Nature 332, 615. Rustagi, K. C , Ducuing, X (1974), Opt. Commun. 10, 258-261. Salcedo, X R., Siegmann, A. E., Dlott, D. D., Fayer, M. D. (1987), Phys. Rev. Lett. 41, 131. Sagiv, X (1980), J. Am. Chem. Soc. 102, 92. Saleh, B. E. A., Teich, M. C. (1991), Fundamentals of Photonics. New York: John Wiley and Sons. Sasabe, H., Wada, T., Hosoda, M., Ohkawa, H., Hara, M., Yamada, A., Garito, A. F. (1990), Nonlinear Optical Properties of Organic Materials III, Proc. SPIE 1337, 62. Sasaki, K. (1992), Sen-i Gakkai Symp. Preprints B-92. Sawodny, M., Stumpe, X, Knoll, W. (1991), /. Appl. Phys. 69, 1927. Sawodny, M., Schmitt, A., Stamm, M., Knoll, W, Urban, C , Ringsdorf, H. (1992), Thin Solid Films 210/211, 500. Schildkraut, X (1991), Appl. Phys. Letters 58, 340. Schrader, S., Koch, K. H., Mathy, A., Bubeck, C , Mullen, K., Wegner, G. (1991), Prog. Colloid Polym. Sci. 85, 143. Shen, Y. R. (1984), The Principles of Nonlinear Optics. New York: John Wiley and Sons. Shuto, Y, Amano, M., Kaino, T. (1991), Nonlinear Optical Properties of Organic Materials IV, Proc. SPIE 1560, 111. Simmrock, H.-U., Mathy, A., Dominguez, L., Meyer, W H., Wegner, G. (1989), Angew. Chem. Adv. Mat. 101, 1148. Singer, K. D., Kuzyk, M. G., Sohn, X E. (1987), /. Opt. Soc. Am. B4, 968. Singer, K. D., Sohn, X E., King, L. A., Gordon, H. M., Katz, H. E., Dirk, C. W. (1989), /. Opt. Soc. Am. B6, 1339. Spinke, X, Yang, X, Wolf, H., Liley, M., Ringsdorf, H., Knoll, W (1992), Biophys. J. 63, 1667. Stegeman, G. I., Stolen, R. H. (1989), J. Opt. Soc. Am. B6, 652. Stegeman, G. I., Seaton, C. T, Zanoni, R. (1987), Thin Solid Films 152, 231. Swalen, X D. (1986), J. Mol. Electron. 2, 155. Swalen, X D., Tacke, M., Santo, R., Rieckhoff, K. E., Fischer, X (1978), Helv. Chim. Ada 61, 960. Swalen, X D., Bjorklund, G. C , Ducharme, S., Fleming, W, Herminghaus, S., Jungbauer, D., Jurich, M., Moerner, W. E., Reck, B., Smith, B. A., Twieg, R., Willson, C. G., Zentel, R. (1991), in:
596
12 Optical Properties of Polymers
Organic Molecules for Nonlinear Optics and Photonics: Messier, X, Kajzar, R, Prasad, P., Ulrich, D. (Eds.). Dordrecht, Netherlands: Kluwer Academic PubL, p. 433. Swalen, J. D., Fleming, W, Jurich, M., Moerner, W. E., Smith, B. A., Herminghaus, S., Bjorklund, G. C. (1992), Mater. Res. Soc. Symp. Proc. 228. Boston, MA: Mater. Res. Soc, p. 101. Tamir, T. (Ed.) (1979), Topics in Applied Optics, Vol. 7: Integrated Optics. Berlin: Springer-Verlag. Tamura, K., Padias, A. B., Hall, H. K., Jr., Peyghambarian, N. (1993), Appl Phys. Lett., in press. Teng, C. C , Man, H. T. (1990), Appl. Phys. Lett. 56, 1734. Tien, P. K. (1969), Rev. Mod. Phys. 49, 361. Tillmann, N., Ulman, A., Penner, T. L. (1989), Langmuir 5, 101. Tomono, T., Nishikata, Y, Pu, L. S. (1992), Sen-i Gakkai Symp. Preprints B-102. Townsend, P. D., Jackel, J. L., Baker, G. L., Shelburne III, J. A., Etemad, S. (1989), Appl. Phys. Lett. 55, 1829. Ushioda, S., Sasaki, Y. (1983), Phys. Rev. B27, 1401. Verbiest, T., Persoons, A., Samyn, C. (1991), Nonlinear Optical Properties of Organic Materials IV, Proc. SPIE 1560, 353. Williams, D. J. (1987), in: Nonlinear Optical Properties of Organic Molecules and Crystals: Chemla, D. S., Zyss, J. (Eds.). Orlando, FL: Academic Press. Wolter, H. (1956), in: Handbuch der Physik: Fliigge, S. (Ed.). Berlin: Springer-Verlag. Wu, J. W., Valley, J. R, Stiller, M., Ermer, S., Binkley, E. S., Kenney, J. T., Lipscomb, G. F , Lytel, R. (1991), Nonlinear Optical Properties of Organic Materials IV, Proc. SPIE 1560, 196. Zernicke, F , Midwinter, J. E. (1973), Applied Nonlinear Optics. New York: John Wiley and Sons.
Zhao, M. T., Singh, B. P., Prasad, P. N. (1988), /. Chem. Phys. 89, 5535-5541. Zyss, J. (1985), J. Mol. Electron. 1, 2.
General Reading Bowden, M. J., Turner, S. R. (Eds.) (1988), Electronic and Photonic Applications of Polymers, Advances in Chemistry Series, Vol. 218. Washington, DC: ACS. Emerson, J. A., Torkelson, J. M. (Eds.) (1991), Optical and Electrical Properties of Polymers, Materials Research Society Symposium Proceedings, Vol. 214. Pittsburgh, PA: MRS. Kai Chang (Ed.) (1991), Fiber and Electro-Optical Components, Vol. 4 of Handbook of Microwave and Optical Components. New York: Wiley. Kampf, G. (1985), Ber. Bunsenges. Phys. Chem. 89, 1179. Knoll, W. (1991), Mater. Res. Bull 16, 29. Prasad, P. N., Williams, D. J. (1991), Introduction to Nonlinear Optical Effects in Molecules and Polymers. New York: Wiley. Raether, H. (1988), Surface Plasmons on Smooth and Rough Surfaces and on Gratings, Springer Tracts in Modern Physics, Vol. 111. Berlin: Springer. Roberts, G. G. (Ed.) (1990), Langmuir-Blodgett Films. New York, Plenum. Saleh, B. E. A., Teich, M. C. (1991), Fundamentals of Photonics. New York: Wiley. Singer, K. D. (Ed.) (1991), Nonlinear Optical Properties of Organic Materials IV, Proc. SPIE 1560.
13 High Performance Polymer Fibers Hao Jiang Lawrence Associates Incorporated, Dayton, OH, U.S.A. W. Wade Adams Materials Directorate, Wright Laboratory, Wright-Patterson Air Force Base, OH, U.S.A. Ronald K. Eby Institute of Polymer Science, University of Akron, Akron, OH, U.S.A. List of 13.1 13.1.1 13.1.2 13.1.3
Symbols and Abbreviations Introduction Definition of High Performance Polymer Fibers Ideal Structure of High Strength and High Modulus Polymer Fibers Historical Review of the Development of High Performance Polymer Fibers 13.2 Fiber Processing 13.2.1 Gel Spinning of Flexible and Semi-flexible Polymers 13.2.2 Melt Spinning of Thermotropic Liquid Crystalline Polymers 13.2.3 Dry-Jet Wet Spinning of Lyotropic Liquid Crystalline Polymers 13.2.3.1 Polymer Synthesis 13.2.3.2 Lyotropic (Liquid Crystal) Solutions 13.2.3.3 Fiber Spinning 13.2.4 Solution Spinning 13.2.5 Solid State Extrusion 13.3 Fiber Morphological Structure 13.4 Fiber Properties 13.4.1 Tensile Properties 13.4.2 Compressive Properties 13.4.3 Thermal Properties 13.4.4 Miscellaneous Properties 13.4.5 End Uses of High Performance Polymer Fibers 13.5 Other High Performance Fibers 13.5.1 Fibers for Medical Uses 13.5.1.1 Surgical Fibers 13.5.1.2 Biological and Chemical Membrane Fibers 13.5.1.3 Fiber Carriers for Biologically Active Agents 13.5.2 Polymer Fiber Precursors to Superconductors 13.5.2.1 Fibers from Mixtures of Superconducting Oxides and PVA 13.5.2.2 Conducting Polymer Fibers 13.6 The Future 13.7 References Materials Science and Technology Copyright © WILEY-VCH Verlag GmbH & Co KGaA. All rights reserved.
598 600 600 602 603 605 605 607 609 610 611 613 615 616 617 626 626 631 635 636 638 639 639 640 642 644 644 645 645 646 647
598
13 High Performance Polymer Fibers
List of Symbols and Abbreviations / /0 m Mw P
Hermans-Stein orientation factor test length of the fiber Weibull modulus molecular weight cumulative failure probability function
Tg Tm TNI
glass transition temperature melting temperature isotropic-nematic transition temperature
M O
intrinsic viscosity stress normalization factor for Weibull model
ABA ABP AFM ANA AS BF CSA CTE DABDO DABTO DF DMSO DP ED FESEM GPC HMPA HPF HPOPF HREM HT IA LM MSA MTM NMP PAN PBZO PBZT PDA PE
p-acetoxy benzoic acid p,p-diacetoxy biphenyl atomic force microscopy acetoxy naphthoic acid as-spun bright field chlorosulfonic acid coefficient of thermal expansion 4,6-diamino-1,3-benzenediol dihydrochloride 2,5-diamino-l,4-benzenedithiol hydrochloride dark field dimethyl sulfoxide degree of polymerization electron diffraction field emission scanning electron microscopy gel permeation chromatography hexamethylphosphoramide high performance fiber high performance organic polymer fiber high resolution electron microscopy heat treated isophthalic acid light microscopy methane sulfonic acid micro-tensile testing machine rc-methylpyrrolidone polyacrylonitrile poly( p-phenylene benzobisoxazole) poly( p-phenylene benzobisthiazole) polydiacetylene polyethylene
List of Symbols and Abbreviations
PET PGA PMMA POM PP PPA PPD PPTA PSF PVA SAED SAXS SEM SSF STM SWAXS TA TCI TDT TEM TFA TGA THF TLCP TMA UHMW UV WAXS
poly(ethylene terephthalate) polyglycolic acid polymethyl methacrylate polyoxymethylene polypropylene polyphosphoric acid p-phenylene diamine poly( p-phenylene terephthalamide) polysulfone poly(vinyl alcohol) selected area electron diffraction small-angle X-ray scattering scanning electron microscopy spin stretch factor scanning tunneling microscopy synchrotron WAXS terephthalic acid terephthaloyl chloride thermal distortion temperature transmission electron microscopy trifluoroacetic acid thermogravimetric analysis tetrahydrofuran thermotropic liquid crystalline polymer thermal mechanical analysis ultra-high molecular weight ultraviolet wide-angle X-ray scattering
599
600
13 High Performance Polymer Fibers
13.1 Introduction "A civilization is both advanced and limited by the materials at its disposal" Sir George Padget Thompson 13.1.1 Definition of High Performance Polymer Fibers High performance fibers (HPFs) have excellent mechanical properties or other superior features, for example, thermal stability, chemical resistance, electrical properties, or unique biological behavior. They can be divided into two groups: the inorganic and the organic polymers. The representatives of the inorganic group are carbon, boron, silicon carbide (SiC), high strength glass, alumina (A12O3), aluminum and titanium alloy fibers. The typical high performance organic polymer fibers (HPOPFs) are aromatic aramid fibers such as poly(/7-phenylene terephthalamide) (PPTA) under the Kevlar trademark, including Kevlar 29, Kevlar 49 and Kevlar 149; aromatic heterocyclic polymer fibers such as poly(/>-phenylene benzobisthiazole) (PBZT) and poly(j?-phenylene benzobisoxazole) (PBZO); aromatic polyesters such as polyester-polyarylate fiber (Vectran); and ultra-high molecular weight (UHMW) extended chain polyethylene (PE) fibers such as Spectra, including Spectra 900 and Spectra 1000 (Table 13-1). These broad classes of high performance fibers make giving a unique definition difficult, if not impossible, and indeed many names have been used, such as "high-tech fibers", "strong fibers" and "super fibers". In this chapter, our discussion is confined to high performance organic polymer fibers. A detailed discussion of carbon and other inorganic high performance fibers is presented in Volume 13 of this Series.
High performance organic polymer fibers are unique because they possess many excellent properties that significantly exceed those of traditional textile fibers, such as nylon and polyester (Fig. 13-1). High stiffness and strength combined with low density make the HPOPFs very competitive among fibrous materials, Table 13-1. Properties of fibers. Materials (Fibers)
Tensile Tensile Compres- Density modulus strength sive strength (GPa) (GPa) (g/cm3) (GPa)
Steel 200 71 Al-Alloy Ti-Alloy 106 350-380 Alumina Boron 415 SiC 200 S-Glass 90 Carbon PI00 725 (pitch-based) Carbon M60J 585 (PAN-based) Kevlar 49 125 Kevlar 149 185 PBZT 325 PBZO 360 Spectra 1000 172 Vectran 65 Technora 70 Nylon 6 Textile PET 12
2.8 0.6 1.2 1.7 3.5 2.8 4.5 2.2
_ 6.9 5.9 3.1 >1.1 0.48
3.8
1.67
1.94
3.5 3.4 4.1 5.7 3 2.9 3.0 1.0 1.2
0.39-0.48 0.32-0.46 0.26-0.41 0.2-0.4 0.17 _ 0.1 0.09
1.45 1.47 1.58 1.58 1.0 1.4 1.39 1.14 1.39
7.8 2.7 4.5 3.7 2.5-2.6 2.8 2.46 2.15
'Glass
6
12
Strain (%) Figure 13-1. Stress-strain curves of various fibers.
18
601
13.1 Introduction 4.0 T"
T1000 O PBZO
O Spectra 1000
3.0 - -
Kevlar 49 QSpectra 900 Technora O Kevlar 149 OVectran OAS4
c 0)
2.0
Kevlar 29
"c/5 c
S-glass
O M60J
°T300
(D
O Boron
T50 ;
OsiC
O Nylon
O E130
OGY70
P100
Figure 13-2. Specific tensile strength versus specific tensile modulus of various fibers.
OP25
Q. CO
Steel Al-alloy
0.0
0
70
140
210
280
350
420
Specific Tensile Modulus (MJ/Kg)
especially when compared on the basis of specific properties, such as specific strength and specific modulus (Fig. 13-2). These specific properties are defined as strength or modulus divided by density. Table 13-2 lists the unit conversions. Other useful characteristics of HPOPFs vary somewhat from fiber to fiber and include good chemical, corrosion, moisture, and high-temperature resistance, dielectric
properties, easy processing and (sometimes) low cost. High performance can also refer to unique biological properties, such as resistance to biodegradation, ability to form hollow fibers with selective permeability, and ability to act as carriers for drugs. The development of HPOPFs has created a new generation of materials. They are an enabling technology for many present and future high-technology prod-
Table 13-2. Unit conversions.a
g/Tex
g/d
g/Tex
kg/mm2
dynes/cm2
1
0.111 1
QX 8.995
#x 8.818 xlO 8
QX 8.818 x 107 e x 1.281 xlO 4
^x 80.96
QX 7.936 x 109
QX 7.936 x 108 QX 1.152 x 105
1
9.803 xlO 8
9.803 xlO 7
1.281 x 104
xlO" 1 0 1.019 x l 0 ~ 8
1
0.1
6.889 X 104
xlO" 9
10
1
1.451 X 10" 4
6.889 x104
6.889 x103
1
9.000
dynes/cm2
xl0~9 Q
Pa
0.0124
0.111
kg/mm2
lb/in2
Q
xlO~
lb/in2 a
Pa
Q is the density of the fiber in g/cm3.
1.019xl0~ 8 7.013
602
13 High Performance Polymer Fibers
ucts, especially as the backbone materials of the advanced composites used in the aerospace industries. Composites with high performance polymer fibers provide advances in stiffness and weight savings (every kilogram saved in the structure of an airplane saves hundreds of dollars over its lifetime, and for a spacecraft, tens of thousands of dollars) (Adams and Eby, 1987). Therefore, not only performance but also cost effectiveness can benefit from their increased utilization. 13.1.2 Ideal Structure of High Strength and High Modulus Polymer Fibers
Research to date has concentrated on the one-dimensional architecture, i.e. linear molecular chain polymers, for high performance fibers. Staudinger predicted the structure of the ideally crystalline polymer in the early 1930s when he sketched the "continuous crystal" model (Fig. 13-3) (Staudinger, 1932). His crystal model corresponds to the current ideal structure of a high performance polymer fiber: all molecular chains fully extended to form crystals with no or very few chain-end defects. All the molecules are well-oriented and densely packed in good lateral order to create a structure with very low free energy. The modulus of the fiber is thus able to approach the modulus of the crystals, which in turn approaches the intrinsic molecular chain stiffness. Therefore, as the precursor materials of high performance fibers, polymers must meet the following four requirements: bonding in the molecular backbone should be very strong and stiff; molecular weight should be high to guarantee long molecular chains; molecular conformation should be linear (or close to linear); and the crosssection of the molecular chain should be small. Table 13-3 lists bonding energies for
Figure 13-3. Schematic diagram of Staudinger's "continuous crystal" model for polymers.
Table 13-3. Bonding energy of different bonds. Bond
Bonding energy (kcal/mol)
C-C
83 146 200 86 73 147 65 128
c=c c=c c-o C-N C=N
c-s c=s
different types of connections. Polyethylene, ^ra^-polyacetylene and polyene are three good examples, since their bonding angles are 112°, 122° and 180°, respectively (for polyene, the sp-sp connection is completely linear), and the bond length between carbon atoms decreases and the bonding energy (as seen in Table 13-3) increases in this sequence. Figure 13-4 is a plot of theoretical axial modulus [measured values in the cases of polydiacetylenes (PDAs), diamond and graphite] against the effective density. The effective density of a polymer has been defined as the density of the proportion of atomic mass being actually in the backbone (Donald and Windle, 1992). For example, the effective density of PE is taken as the actual density x 24/28, which is the ratio of the molecular weight of the backbone atoms to that of the total molecular weight per chain repeat, (CH 2 -CH 2 -) n . As shown in Fig. 13-4, the calculated values of ultimate stiffness for PE, PPTA and PBZO, as well as the measured values for PDAs,
13.1 Introduction 1500
Diamond < 1000 — -
Graphite
• PBO 500 -
PE • PPTA « • PDAs | 0
1
I 2
3
Effective chain density (g/cm3)
Figure 13-4. Plot of calculated axial modulus versus effective density (which only takes into account the mass of the backbone atoms) for different materials (Donald and Windle, 1992).
diamond and graphite (in the basal plane), lie approximately in a line, which implies that the backbone packing density affects the ultimate fiber stiffness (Donald and Windle, 1992). Thus, the goal in high strength/stiffness fibers is, from suitable precursors and through optimum processing, to manufacture highly ordered polymer fibers that have structure and properties close to those of the ideal model. 13.1.3 Historical Review of the Development of High Performance Polymer Fibers
Fiber technology is ancient, with the first fibers being from natural products which were abundantly available. These can be divided into three groups: vegetable (cellulose), e.g. linen and cotton; animal (protein), e.g. silk and wool; and mineral, e.g. asbestos and some metal filaments. Nature has provided man with abundant spinners of fine organic fibers, for example, spiders (silk). Earliest records of linen prepared from flax appeared during the early Egyptian dynasties nearly 6000 years ago. Silk, with the cultivation of bombyx
603
rnori, began to attract the attention of Chinese royalty during the 3rd millennium B.C. and gold threads were used in India to decorate saris thousands of years ago (Delmonte, 1985). As in the development of synthetic polymers, the first step toward the development of synthetic fibers was to modify natural materials. This step took place in the nineteenth century, when filaments were spun from solutions of cellulose derivatives, such as cellulose nitrate. In the early part of the twentieth century, viscose rayon, cellulose acetate, and others followed. Since the pioneering work of Carothers which led to the introduction of nylon fibers by DuPont in the 1930s, polymer scientists have pursued the development of high performance fibers to replace natural or metallic products with products of improved mechanical properties and reduced weight. Strong motivation for the invention of new polymer fibers has come from the aerospace industry, which seeks fibers to use in reinforced composite structural materials. Development was relatively slow and evenly paced until DuPont revolutionized the field with the release of Kevlar. In the 1970s and 1980s, the field boomed with new developments. Table 13-4 lists in chronological sequence of publication some authors and literature that reported either new fiber processing technologies or new high performance polymer fibers. This table is given only as an example since many new processing methods developed in companies have not been published yet, or were published much later. In the past 20 years, as shown in Fig. 13-5, polymer fibers have shown great improvements in strength, compared to some inorganic fibers, such as glass and boron fibers, whose properties have remained largely unchanged. Today, several high performance polymer fibers are com-
604
13 High Performance Polymer Fibers
Table 13-4. Literature survey of high performance polymer fibers. Fiber
Reference
PE Kevlar PE PE Nylon, PET PE PBZT POM PDAs POM PE
Southern and Porter (1970) Kwolek (1972) Pennings et al. (1972) Capaccio and Ward (1974) Kunugi etal. (1979) Smith and Lemstra (1980 b) Allen etal. (1981b) Nakagawa et al. (1983) Galiotis and Young (1983) Ishida (1987) Kanamoto et al. (1987)
Method Solid-state extrusion Liquid-crystal spinning Surface growing Super drawing Zone drawing Gel spinning Liquid crystal spinning Microwave heating-drawing Solid-state polymerization High-pressure thermo-drawing Reactor powder two-stage drawing
mercially available with properties that compete with the best inorganic fibers and are far superior to metal fibers. There are, at present, two primary technical routes to HPOPFs: (1) For the flexible or semiflexible polymers, new technologies are used, such as gel spinning, to make fibers with extended molecular chains that form crystals highly oriented along the fiber axis direction (Pennings, 1977; Smith and Lemstra, 1980a, b; Lemstra et al., 1986; Hoogsteen et al., 1988 a, b; Igachi and Kyotoni, 1990).
(2) For the rigid or semi-rigid polymers, because of the stiff aromatic or heterocyclic structural units in their molecular backbone, liquid crystal spinning solutions are used in dry-jet wet-spinning techniques to produce fibers with very high orientation and crystallinity. Some of these anisotropic polymers can also be melt or solution spun followed by high-draw and heat treatment to obtain high performance fibers (Watt and Perov, 1985; Adams and Eby, 1987; Ohta et al., 1988; Yang, 1989; Yamada, 1991; National Materials Advisory Board, 1992).
Extruded liquid crystal f i b e r s .
Melt-spun nylon
Wet-spun cellulose Natural fibers \ 1900
1910
1920
1930
Figure 13-5. Strength of typical commercial organic polymer fibers developed over a number of years (After National Materials Advisory Board, 1992).
1940
1950
Year
1960
1970
1980
1990
13.2 Fiber Processing
13.2 Fiber Processing Advances in the performance of polymer fibers have come about because of continuing progress in new fiber-forming precursors and fiber processing research and development. A more detailed account of fiber processing is presented in Volume 18 of this Series. Due to breakthroughs in processing techniques and new materials, especially since the 1970s, the mechanical properties of polymer fibers have improved dramatically (Fig. 13-5). The new, nearly perfectly oriented fibers can now be spun from liquid-crystal polymers through melt-spinning and dry-jet wet-spinning techniques, as well as from linear flexible chain polymers through gel spinning. The selection of a suitable process depends largely upon the solution or melt behavior of a polymer. For example, gel spinning may not be satisfactory for rigid polymers because of limited gel drawability. Similarly, semi-rigid polymers which do not form liquid crystal solutions or melts may be converted to high performance fibers from their isotropic solutions followed by drawing at elevated temperature. Therefore, such process selection is significantly dependent on polymer structure. Various processes are now reviewed. 13.2.1 Gel Spinning of Flexible and Semi-flexible Polymers
As detailed in Chapter 4, in the 1970s and 1980s, the pioneering work of Pennings (Pennings et al., 1972; Pennings, 1976) developed the idea of making high performance fiber from solution spinning of flexible chain PE, and later improvements made by Smith and Lemstra (Smith and Lemstra, 1980 a, b) involved the drawing of a gel. Because the linear molecular chains are arranged parallel to the fiber
605
axis, the tensile strength and modulus of gel-spun PE fiber can reach 7 GPa and 200 GPa, respectively (Ward, 1985 a, b). This technique has created a new route to produce high strength and high modulus fibers from flexible chain polymers. The decade of the 1980s was the primary stage for the industrialization of "gel spinning". Up to now, three commercialized PE extended-chain fibers have been introduced to the world market, Spectra manufactured by Allied-Signal, Dyneema by DSMToyo and Tekmilon by Mitsui (Chang and Weedon, 1986; C & E News, 1987). Interest in this area is now growing and many other flexible or semi-flexible chain polymers, such as nylon, polypropylene (PP), poly(vinyl alcohol) (PVA) and polyethylene terephthalate) (PET) are being studied in relation to this technique to produce high performance fibers (EP 83108258; EP 84114872; Peguy and Manley, 1984; Takahashi et al., 1991a, b). In the so-called "gel spinning" of polyethylene fibers, the key feature is to use ultra-high molecular weight PE (Mw > 1 000 000) dissolved in a solvent, such as paraffin, xylene, naphthalene or decalin at a low concentration. When the dilute solution of ultra-high molecular weight PE emerges from the spinneret, it is first quenched by an air gap and then enters a cooling bath. A thermodynamically reversible gel is created. Owing to the very long polymer chains in the dilute solution, entanglements between molecules are formed and in the procedure of making gel, the entanglement-network is maintained. It is very interesting that even after removal of the solvent the gel can still be highly drawn. The reason for this is that during cooling, the as-spun fiber forms small chain folded lamellae with a significant amount of solvent between them so that the as-spun fiber keeps the physical
606
13 High Performance Polymer Fibers
gel formation, namely, the swollen network in which crystallites form the junctions, as shown in Fig. 13-6 (Ward, 1987). In the drawing, entanglements act as temporary physical crosslinks and play an important role. This feature allows the asspun fiber to be drawn to very high ratios, even over 200 times. The maximum draw ratio is closely related to the average distance between the entanglements, i.e. the "crosslinking" density and the solution concentration (Fig. 13-7) (Smith and Lemstra, 1980 a, b; Iguchi and Kyotani, 1990). Upon hot drawing of the as-spun fiber, the lamellae are gradually transformed into smooth fibrils. If the draw ratio is high (typically over 20), "super-drawn" fiber is produced, as shown in Fig. 13-8 (Ohta etal., 1988). In the super-drawn fiber, there are long arrays of extended-chain crystals interrupted by disordered domains that originate from the presence of topo-
i i i i
I I i i I
i i i i i i i i i i i i
i i i i
I I I I I
i i i i i i i i i i i i
i i i i
I I I I I
i i i i i i i ii
f\
i i i
r\
i i
I I I
I I I I I I
I i
I I I
I I I I I I
i i
I I i
I II
i i
I I I
I I I I I I
w
I I i
M M I i
I i
I I I
I I I I I I I I I I I I
I I I
I I I I I I I I I I I I
I I I
I I I I I I I I M
I I
I I I
I I I I I I I I I I
I I I
I I I I I I I I '
i I I
II I II
Figure 13-6. Structural model of gel-like fibers (Ward, 1987).
100
10 Concentration of PE (%)
Figure 13-7. Relation between maximum draw ratio and UHMW PE solution concentration at 120 °C (circles) and room temperature (squares) (Iguchi and Kyotani, 1990).
Figure 13-8. Structural model of super-drawn UHMW PE fibers (Ohta et al., 1988).
logical defects, such as chain entanglements. Adjacent crystalline blocks are thought to be connected by taut tie molecules traversing the disordered domains (Penning et al., 1991; Agosti et al., 1992) (see Chap. 7). Therefore, this structure results in a high strength and high modulus fiber. The so-called gel-spinning process is in fact sol-spinning and gel-drawing. A typical spinning process employs a solution of 5-15 wt.% UHMW (generally 5-7 million) PE in paraffin oil which is extruded through an air gap into the nhexane bath. Figure 13-9 (Lemstra etal., 1986) is a schematic diagram of the gelspinning process. The as-spun fiber is dried under vacuum and then drawn in a tube
13.2 Fiber Processing
607
Suspension UHMW PE
Quenching/extraction bath
Oven
containing silicon oil with a temperature gradient of 100-150°C. The maximum draw-ratio is normally 30-100 times (Barham and Keller, 1985; Ohta et al., 1988; Hoogsteen et al., 1988 a, b; Iguchi and Kyotani, 1990). The cost of solvent recovery and the controlled low-speed spinning and drawing increase the cost of the fiber to a level comparable to other high performance fibers, despite the lower polymer cost (National Materials Advisory Board, 1992). 13.2.2 Melt Spinning of Thermotropic Liquid Crystalline Polymers
Incorporating aromatic cyclic or heterocyclic rings in the molecular backbone to build stiff chain fiber-forming polymers is another commercially successful way to improve fiber mechanical and thermal properties. Following this synthetic approach from the 1950s, many new polymers have been produced and processed into high performance fibers, typified especially by the aromatic polyesters. One attractive feature of these polymers is that they can form a liquid crystalline phase
o
Fiber
Figure 13-9. Schematic diagram of gel spinning (Lemstra et al., 1986).
over a certain temperature range in the melt, i.e. without the need for a solvent. Such polymers are called thermotropic liquid crystalline polymers (TLCP). Therefore, they can be spun with conventional melt-spinning techniques to achieve high performance properties. Generally, their melting points are around 300-400°C, near their isotropic-nematic transition temperature (TNI). Unfortunately, very few of the aromatic polyesters are suitable for fiber-making because of their tendency to degrade at high temperature and the poor tractability of the polymer melts (Yang, 1989; MacDonald, 1991). The commercialized thermotropic liquid crystal copolyesters can be divided into three general groups, all of which are made by condensation polymerization (Kyotani, 1990). The first group has low melting temperatures, and their thermal distortion temperatures (TDTs) are around 60120 °C. An example from this group is polyester X7G, manufactured by Kodak. X7G was the first thermotropic liquid crystalline fiber on the world market. Polyester X7G is a copolymer of polyethylene terephthalate (PET) and />-acet-
608
13 High Performance Polymer Fibers
oxy benzoic acid (ABA) with a melting point (Tm) of about 230 °C. The comonomer ratio of PET to ABA is in the range of 40:60 to 20:80. The chemical synthesis of X7G is as follows: O
O
O II -C-OH
o
--O-CH 2 -CH 2 -O-C-/(J)VC--
CH.-C-C
-CH.COOH
ABA
PET
O-CHo-CHo-O-CO-
-CO X7G
The second group of TLCPs has intermediate melting temperatures with TDTs of about 180-240°C, as represented by Vectran, marketed by Hoechst Celanese Co. Vectran is a polyester-polyarylate fiber with 7^ around 270 °C. Its synthesis, with a comonomer ratio of about 30:70 ANA (acetoxy naphthoic acid): ABA, is as follows:
COOH
O
O
II
/p^\
II
CH 3 -C-OVrjVC-OH
-CH3COOH
ABA
ANA
Vectran
The third group of TLCPs has high melting temperatures with TDTs around 250350 °C. Ekonol, which was developed at Carborundum and is now produced by Sumitomo Chemical Co. and Nippon Exlan Co., is representative of this group. Its melting temperature is above 400 °C. The general synthesis for the base polymer is o
o
CH3~CO-O VQWQVO-CO-CH3 + ABP O O II II HO-C-r^^rC-OH IA
TA
o o CH 3 -C-OVQVC-OH
-CH3COOH
ABA
Ekonol
Its comonomer ratio of (ABP: TA: IA: ABA) is 5:4:1:10, where ABP is />,//-diacetoxy biphenyl; TA, terephthalic acid; and IA, isophthalic acid.
13.2 Fiber Processing
The spinning process is similar to conventional PET melt spinning. The polymer is first washed with water, alcohol or acetone to remove impurities. Then it is heated to about 300-400°C, near its isotropic-nematic transition temperature, to form the thermotropic liquid crystalline spinning melt. The spinning melt with a typical intrinsic viscosity of 1.5-3 dl/g, is extruded at 280-350°C through spinneret holes to form the fiber. The spinning speed is in the range of several hundred to several thousand meters per minute. The as-spun fibers are heat treated at elevated temperatures (normally around 250- 300 °C) for several hours, which is much longer than in a conventional annealing process. During this time a solid state polymerization occurs which is an important factor in increasing the fiber strength. The spinning and heat treatment conditions are very critical to the fiber formation and quality. These processing conditions are largely selected from experience. The temperature of melt spinning must be above the crystalto-nematic transition temperature to guarantee formation of the liquid crystalline phase. The higher the processing temperature, the faster the melt degrades; but the lower the temperature, the higher the melt viscosity and the poorer the processing. Molecular weight control is of special importance. In the melt, increasing the molecular weight leads to a sharp increase in the melt viscosity, which can even be sufficient to prevent processing. Thus the upper limit of the molecular weight for the melt spinning of thermotropic liquid crystal polymers is relatively low, which results in lower fiber tensile strength. However, the possibility of these fibers undergoing a solid state polymerization during the heat treatment can cause a significant increase in the molecular weight, and thus the tensile properties. This process thus requires
609
much longer heat-treatment times than conventional ones. It was observed that after this type of heat treatment there were new connections between the original microfibrils, but the overall fibril structure and orientation did not change (Matsuda, 1985a, 1985b; Yang, 1989). Heat-treatment technologies with shorter times have been evolved. It was reported that by attaching substituents (such as the phenyl group) to the hydroquinone moieties of aromatic polyesters, even with short heattreatment times (30 min), the fibers had satisfactory tensile properties (Jackson, 1980). In general, these fibers have tensile strengths of 2 - 6 GPa, tensile moduli of 60-130 GPa and an elongation to break of 2 - 4 % (Ohta et al., 1988; Yang, 1989; Kyotani, 1990; Koide, 1990). Compared with the solution spinning of lyotropic liquid crystal polymers (which will be discussed in the next section), melt spinning of thermotropic liquid crystal polymers has the following benefits: there are no solvent disposal and recycling problems; it is easy to obtain small diameter fibers; and there is less difference between the outer layer of the fibers, "skin", and the center or "core". Many major chemical companies have been involved in research on thermotropic liquid crystal polymer fibers, driven by a potential market for new industrial applications and tempered by the realities of the high cost and poor availability of the TLCP precursors, as well as their processing difficulties. 13.2.3 Dry-Jet Wet Spinning of Lyotropic Liquid Crystalline Polymers
Most high performance polymers decompose at high temperatures without melting and can be dissolved in very few solvent systems due to their aromatic structure and to their rigid molecular
610
13 High Performance Polymer Fibers
backbone. Therefore, conventional meltspinning and solution-spinning technologies cannot be used. Fortunately, it has been found that certain rod-like polymers such as aromatic polyamides, when dissolved in an appropriate solvent, form a liquid crystalline solution. High performance fibers can be processed from these lyotropic (liquid crystal) systems by means of the dry-jet wet-spinning technique. Since the late 1960s, many lyotropic polymers have been reported and very active fiber research has occurred all over the world, including the U.S., Japan, the Netherlands, France, and the former Soviet Union. Examples are poly(ra-phenylene isophthalamide), known as Nomex from DuPont; poly(/?-benzamide), called Fiber B (DuPont); polyamide hydrazide, X-500 studied by Monsanto in the early 1970s;
-NH-r^rNH-CO-
13.2.3.1 Polymer Synthesis -CO-
Nomex
CO-NHH-CO Polyamide hydrazide, X-500
Kevlar
PBZT
the first commercial and the most successful high performance fiber, poly(/?-phenylene terephthalamide) (PPTA), DuPont's Kevlar; poly(/?-phenylene benzobisthiazole) (PBZT), and, poly(/?-phenylene benzobisoxazole) (PBZO), which are two of the strongest high performance fibers today. However, in spite of all these efforts to date, only Kevlar (introduced to the market in 1972 by DuPont) and Twaron (later by Akzo) are produced on a large scale (Adams and Eby, 1987). The reasons for this are difficult polymer processibility, few precursor sources, high manufacturing cost, small market size and the long time required to evaluate fiber behavior in industrial usage. In this chapter, we focus on the most successful aramid fiber, Kevlar, and the strongest polymer fibers, PBZT and PBZO, as models of HPOPFs.
The synthesis of rigid-rod polymers appears to be deceptively simple condensation chemistry. In practice, it is technologically difficult, to which the great time lag in commercialization testifies. PPTA is synthesized by the condensation of terephthaloyl chloride (C1CO-C 6 H 4 -COC1) (TCI) and /?-phenylene diamine (H 2 NC 6 H 4 -NH 2 ) (PPD) in a mixture of hexamethylphosphoramide (HMPA) and nmethylpyrrolidone (NMP) solvents (Matsuda, 1985a, 1985b; Yang, 1989). The polycondensation reaction is normally kept at a low temperature (preferably at 10-20°C) to avoid degradation and side reactions. The reactants are vigorously stirred and the products are subsequently ground in water, filtered, and the resulting solid polymer thoroughly washed. The degree of polymerization (DP) of PPTA is relatively low (about 7-264), with a Mw in the range of 1700-63 000 measured using
611
13.2 Fiber Processing
gel permeation chromatography (GPC) (Yang, 1989; Fitzgerald and Irwin, 1991). Among the aromatic heterocyclic polymers, the most attractive for potential use as high performance fibers at present are PBZT and PBZO. These two polymers offer excellent fiber tensile and thermal properties unsurpassed by other aromatic fibers. Polymer chemists at the U.S. Air Force's Materials Laboratory, working along directions parallel to the DuPont workers, were exploring aromatic heterocyclic polymers for high-temperature applications in the 1960s. These polymers have higher thermal stability than the polyesters or the aramides. In the 1970s1980s, an extensive development program at both Celanese and DuPont led to successful fiber preparation from these rigidrod polymers. Recently, Dow Chemical Co. has been considering the commercialization of PBZO fibers, because of their excellent mechanical properties and cheaper synthetic route compared to PBZT (Reisch, 1987; Arnold, 1989). trans-PBZT is polymerized from 2,5-diamino-l,4-benzenedithiol hydrochloride (DABDT) and terephthalic acid (TA), while cis-PBZO is based on 4,6-diamino-l,3-benzenediol dihydrochloride (DABDO) and TA, as shown in Fig. 13-10. Both reactions are carried out in polyphosphoric acid (PPA) solvent. Technological requirements include small particle size, crystalline TA and very pure PPA, DABDT and DABDO. The synthesis of DABDO and DABDT and subsequent polymerization were originally developed by Wolfe (Wolfe, 1985 a, b,c, 1988, 1989; Lysenko, 1988). 13.2.3.2 Lyotropic (Liquid Crystal) Solutions
In 1956 Flory predicted that concentrated solutions of rigid-rod polymer mol-
TA
DABDT
HCl-H2N
NH2-HCl HOOC^QVCOOH
HO
OH TA
DABDO
PBZO Figure 13-10. Syntheses of PBZT and PBZO.
ecules readily form liquid crystals (Flory, 1956). The development of liquid crystalline solutions depends on the molar mass, the length-to-diameter ratio and the flexibility of the molecules, the solvent and the temperature (also see Chapter 5 of this Volume). However, the most important parameter is the polymer structure, which must exhibit some rigidity. The necessary rigidity of many of the polymers usually makes them decompose before melting. The important consequence of solvent addition is that liquid crystallinity can be obtained with rigid-rod polymers which would not otherwise exhibit the mesophase at temperatures below that of thermal decomposition. Because they are much less soluble in common solvents than more flexible polymers, their dissolution often requires the use of strong, protonating acids (such as H 2 SO 4 , CF 3 SO 3 H and CH3SO3H) or aprotic solvents (such as
612
13 High Performance Polymer Fibers
dimethylacetamide or TV-methyl pyrrolidone in conjunction with LiCl or CaCl2 in small percentages). The persistence length of a molecule in solution is considered a useful measure of its rigidity, and is defined as the average projection of the end-to-end distance of an infinite chain in the direction of the first link. The longer the persistence length of the polymer, the lower the concentration necessary to form a liquid crystal (Ohta etal., 1988; Donald and Windle, 1992). Persistence lengths of a few polymers are listed in Table 13-5. In the stiff chain polymers there is significant uncertainty associated with the persistence length measurement, because of the difficulty of finding suitable good solvents. An additional problem is that the presence of solvent may change the preferred conformation of the molecule and thus its rigidity (Berry, 1989; Brelsford and Krigbaum, 1991). Therefore, the data in Table 13-5 should be considered only for relative comparison.
The results for PBZT and PBZO are in qualitative agreement with the results of molecular dynamics simulations (Zhang and Mattice, 1992; Farmer and Chapman, 1993). Such results can be subject to uncertainties resulting from questions of force fields, sampling of all of phase space, etc. As polymer is added to the solvent, the viscosity increases [Fig. 13-11 (Won Choe and Kim, 1981)] but the solution remains isotropic and clear. With increasing polymer concentration, the rod-like polymers tend to cluster together in bundles of quasi-parallel rods and to form "domains". These domains are anisotropic
2.1 AT 70 °C 90
80
70
Isotropic
Anisotropic
60 Table 13-5. Measured persistence lengths for various polymers. 50
Polymer trans-PBZT ds-PBZO PPTA
Experimental persistence length (A) >500 (Crosby etal., 1981) >640 (Wong etal., 1978; Aharoni, 1985) 100-175 (Wong etal., 1978) 435 (Erman etal., 1980) 200-400 (Arpin, 1976; Arpin and Strazielle, 1977)
PBA
410 (Erman etal., 1980) 200-400 (Arpin, 1976; Arpin and Strazielle, 1977)
X-500
75 (Ciferri, 1975)
Cellulose PE
42 (Kamide, 1983) 5.75 (Erman etal., 1980)
30
20
1
1
V
2
i
i
i
3
k
5
6
7
8
9
10
% PBZO in 100% H2SO4
Figure 13-11. Critical concentration curve for PBZO (Won Choe and Kim, 1981).
13.2 Fiber Processing
and within them there is nematic order of the chains. However, there is little or no directional correlation among these domains. At this stage, even though the solution changes from monophasic to biphasic, its viscosity still increases with concentration. At a critical concentration, the solution becomes opaque and anisotropic. There is a sharp decrease in the viscosity with further increase in polymer concentration. This decrease reflects the formation of the oriented liquid crystal domains in which the chains can be aligned parallel to the flow direction, thereby reducing the frictional drag on the molecules. The critical concentration is a function of the solvent and the temperature. It tends to decrease with increasing molar mass and with increasing length-to-diameter ratio of the polymer. Theoretical predictions of the behavior of viscosity around the critical concentration have been developed (Kawai, 1980; Doi, 1981). Another characteristic of the liquid crystal solutions is that their viscosity is a function of the alignment of the molecules. In other words, all external fields, such as flow, stress, and even electric and magnetic fields, which can influence the orientation of the molecules will affect their viscosity (Collings, 1990; Donald and Windle 1992). For instance, in the anisotropic liquid crystal solutions the directors readily align with the shear direction, and with increasing shear rate the viscosity drops (Hermans, 1962; Baird, 1978). Although the rheological behavior of the lyotropic solutions has been studied extensively, especially for those technologically important materials, it is still not well understood (Donald and Windle, 1992). For PPTA in concentrated H 2 SO 4 the intrinsic viscosity [rj] correlates with the weight average molecular weight (Mw) (Schaefgen et al., 1976) as
613
M = 7.9xlO- 5 Mi/ 06 (13-1 a) (for Mw> 12000) [rj] = 2.8 x
10~7Ml7
(for Mw< 12000) (13-1 b) The fiber strength increases with increasing viscosity (Matsuda, 1985 b). When the viscosity increases from 4 dl/g to 6 dl/g, corresponding to a molecular chain length of 220 nm, the strength improves linearly from 2.6 GPa to 3.7 GPa. Therefore, to make satisfactory fibers, the viscosity of the polymers should be greater than 4 dl/g (i.e., molecular weight > 27 000) and normally polymers with a viscosity around 5.1 dl/g are used (Ohta et al., 1988). PBZT and PBZO are soluble only in strong aprotic acids such as polyphosphoric acid (PPA), methanesulfonic acid (MSA), chlorosulfonic acid (CSA), trifluoroacetic acid (TFA), and 100% sulfuric acid. For PBZT and PBZO in MSA the intrinsic viscosity-weight average molecular weight relations are as follows (Wolfe, 1988): for PBZT (13-2) 7
[rj] = 2.77 x 10" x M
for PBZO (13-3)
By dissolving isolated trans-VBZT in MSA (about 10%) or using PBZT/PPA solution (5-6%), a nematic spinning dope is made for dry-jet wet spinning. PPA solutions offer significant advantages of improved properties and process simplicity, since one less step is involved, as the polymer does not have to be isolated and redissolved (Wolfe, 1988, 1989). 13.2.3.3 Fiber Spinning
The superior mechanical properties of Kevlar, PBZT and PBZO fibers have been achieved with an optimized combination of high molecular weight polymer, liquid crystal spinning dope, spinning process
614
13 High Performance Polymer Fibers
and final tensioned heat treatment at high temperature. It has been demonstrated that for lyotropic polymer solutions, dryjet wet spinning can produce excellent fibers (Won Choe and Kim, 1981; Allen etal., 1981a, b, 1985; Allen and Farris, 1989; Yang, 1989). As schematically shown in Fig. 13-12, this technique involves the extrusion of the liquid crystalline solution under heat and pressure through an air gap into a coagulation bath, followed by washing, drying, tensioning and heat-setting. As shown in the phase diagram [Fig. 13-13 (Kikuchi, 1982)] of the PPTA/H 2 SO 4 system, a solution with a concentration of about 20% reaches the nematic liquid crystal phase at ca. 80 °C and the isotropic
Air gap —
Winder Washing bath
Coagulation Bath
Figure 13-12. Schematic diagram of dry-jet wet spinning.
phase at ca. 140 °C. Thus the spin dope is extruded at 70-90°C into the air and then immediately passed through a coagulation bath. The air gap serves two purposes: to allow the spin dope to be at a considerably higher temperature, enabling more concentrated spinning solution and higher spinning rates to be used; and to provide a region of extensional flow in the fiber while it is still liquid crystalline in order to promote the development of good orientation of the molecular chains across the full cross-section of the fiber. Normally the spinning speed is around several hundred meters per minute. It is interesting that a spinning speed of 2000 m/min can be achieved (Yang, 1982). It has been reported that the spin stretch factor (the ratio of filament wind speed to extrusion speed) (SSF) is an important parameter for the spinning of lyotropic polymers (Kwolek et al., 1977). Fig. 13-14 shows the SSF effect on the tensile strength of Kevlar as-spun fibers. Below a SSF of 3, the fiber tenacity is considerably lower (Ohta et al., 1988). For Kevlar, cool water at about 5°C is used as coagulant, and for PBZT other coagulants such as dilute phosphoric acid, methanol, and N H 4 0 H have also been used, as well as a room temperature water coagulation (Yang, 1989; Rakas and Farris, 1989). During coagulation, the acid diffuses out of the fiber, which crystallizes
140 120 Isotropic
100
H 2 S0,
80 60
4.8
98%
40
1.3 1.3
96%
Solid
20 0
5
10
15
20
25
Polymer concentration (wt.%)
30
98%
Figure 13-13. Phase diagram of PPTA/H2SO4 system (Kikuchi, 1982).
615
13.2 Fiber Processing
3
Figure 13-14. Relation between the tensile strength of as-spun Kevlar fibers and the spin stretch factor (SSF) (Ohta et al., 1988).
in a highly oriented manner. The coagulation bath may not be a simple one-step process if strong acids are used (as in the production of Kevlar), which themselves need to be neutralized without hydrolyzing the fiber. The as-spun fiber is then washed to remove the residual solvent and dried. Subsequent heat treatment is often used to further improve fiber mechanical properties and it is commonly done under tension. For PBZT, the tensioned fiber passes through a tubular oven in a nitrogen atmosphere. Heat treatment temperatures of 500-700 °C with a residence time of a few seconds to several minutes are typically used (Pottick and Farris, 1985; Rakas and Farris, 1988, 1989; Allen and Farris, 1989; Chenevey and Timmons, 1989; Ledbetter et al., 1989). After tensioned heat treatment, the fibers have an improvement in their overall axial orientation, crystal perfection and lateral order, which results in an increase in their mechanical properties. Table 13-6, taking PBZO as an example,
shows the enhancement in orientation (measured from azimuthal scans of the Xray equatorial reflections), axial and lateral crystallite size (determined from the equatorial diffraction patterns), and the tensile modulus and strength through the tensioned heat treatment (Martin and Thomas, 1991a; Krause et al., 1988). Other novel lyotropic polymer fibers have also been spun by using the dry-jet wetspinning techniques (Kumar, 1990 a, 1991). However, several difficult hurdles are encountered during the development of the spinning process for the polymer lyotropes, such as environmental problems, waste disposal, toxicity and corrosiveness of ingredients, and handling of hazardous materials. Multi-disciplinary chemical and engineering skills are required to solve these problems and thus, costs can be high. 13.2.4 Solution Spinning
High performance polymer fibers can also be manufactured by solution spinning from isotropic spinning dopes of aromatic copolyamides. Technora fiber, commercialized by Teijin Ltd. in the late 1980s, is an example. Technora is based on copoly(/7-phenylene/3,4'-diphenyl ether terephTable 13-6. Changes in structure and property parameters of PBZO fibers through heat treatment at 600 °C (Krause et al., 1988; Martin and Thomas, 1991a). As-spun Heat treated PBZO PBZO Average lateral and (axial) crystallite size (nm) Misorientation angle Modulus (GPa) Tensile strength (GPa) Breaking elongation (%)
2.6
3.9
(3.0)
(3.9)
6°
4°
165
317
4.3
4.9
2.8
1.7
616
13 High Performance Polymer Fibers
thalamide). The polymer is synthesized by the low-temperature polymerization of pphenylenediamine, 3,4'-diaminodiphenyl ether, and terephthaloyl chloride in an amide solvent with a small amount of alkali salt (calcium chloride or lithium chloride) (Shimada et al., 1982). The polymerization reaction is carried out at 0-80°C in 15 hours at a polymer concentration of 6 12% and is terminated with the addition of neutralizing agents (calcium hydroxide, calcium oxide). The reaction mixture is filtered to give a pure isotropic polymer solution. The solution is then spun from spinnerets into an aqueous coagulating bath to form fibers. The coagulant contains 3 5 50% CaCl2 or MgCl2 to control the coagulation rate. The as-spun fiber is washed and extensively drawn (6-10 times) at a high temperature (485 °C). As a consequence, Technora exhibits properties that resemble those of conventional fibers in some aspects and those of rigid chain aramid fibers in other aspects. In addition to high tensile strength (3 GPa) and modulus (70 GPa), it offers some unique fiber properties, such as chemical and impact resistance (Yang, 1989). 13.2.5 Solid State Extrusion Extraordinary fibers have been produced from flexible polymers such as PE by modifying the method of solid state extrusion (Zachariades and Porter, 1983; Ciferri and Ward, 1979; Igushi and Kyotani, 1990). A bulk polymer in the solid state (i.e. at an elevated temperature but below its melting point) is forced through a narrowing die at very high pressures to form a transparent fiber. In the original process, melt crystallized polyethylene was extruded to draw ratios of up to 48, and the extrudate had a modulus of up to about 40 GPa with a degree of crystallinity of
about 85 % (Adams et al., 1985). However, by extruding (typical extrusion draw ratios of about 5) a sedimented mat of single crystals of ultra-high molecular weight PE (up to about 8 million Daltons) followed by post drawing the extrudate at temperatures near the melting point (typical draw ratio of 40-50), very high overall draw ratios, up to 280 times, were achieved (Kanamoto et al., 1983; Porter, 1992). This drawing process, starting with few chain entanglements, produced extremely efficient conversion of the lamellar crystals into fibrillar crystals, with a measured degree of crystallinity of up to 96 % in the fibers. The modulus of the fibers depends primarily on the draw ratio and increases with draw ratio. On the other hand, the tensile strength appears to depend much more on the polymer molecular weight (Yang, 1989). It has been reported that the highest achieved modulus for PE is 220 GPa, which is a significant fraction of the theoretical estimated modulus of about 320 GPa (see Chap. 7) (Kanamoto et al., 1983; Ward, 1985 a; Iguchi and Kyotani, 1990). Although this two-step method appears to be rather difficult to commercialize at present, it has set the current aim for a flexible chain polymer fiber tensile modulus. Many other polymers have been extruded from the solid state, and a great deal of interest currently centers around poly(vinyl alcohol) (PVA). PYA has theoretically high mechanical properties, comparable with those of UHMW PE (Sakurada, 1985; Kanamoto et al., 1990), and it may be possible for its intra- and interchain hydrogen bonds to act as junction-points in forming a highly crosslinked network which may be beneficial for enhanced extension (Sakurada, 1985; Shibayama et al., 1990) and for reducing the likelihood of fiber plastic deformation by creep (Conibeer et al., 1990).
13.3 Fiber Morphological Structure
13.3 Fiber Morphological Structure The morphological structure of high performance polymer fibers has been extensively investigated using wide-angle X-ray scattering (WAXS), small-angle X-ray scattering (SAXS), light microscopy (LM), scanning electron microscopy (SEM), and transmission electron microscopy (TEM), including bright-field (BF), dark-field (DF), and selected-area electron diffraction (SAED) techniques (Yabuki et al,, 1975, 1976; Northolt and Ven Aarsten, 1977; Dobb, 1979; Adams et al., 1979; Roche et al., 1980; Minter et al., 1981; Blumstein, 1984; Gutierrez and Blackwell, 1984; Tashiro et al., 1987, 1988; Krause et al., 1988; Odell et al., 1989; Martin and Thomas, 1989, 1991a, 1991b; Fratini etal., 1989; Shimamura and Hashimoto, 1991). It is very interesting that there are several common structural features in high performance polymer fibers. Most of the fibers have round cross-sections, except for a few which have specialized shapes for specific applications, for example, triangular cross-sections for higher surface area.
617
The typical fiber diameter varies from 10 to 25 microns. Most solution-spun high performance polymer fibers have a distinct skin-core structure. As seen in Fig. 13-15, the TEM micrographs and SAED patterns of Kevlar 149 exhibit an apparent difference in morphological structure between skin and core regions (Young et al., 1992). The existence and magnitude of the skin/ core differences are believed to be the result of processing conditions, such as solvent diffusion and fiber coagulation. There are also voids in these fibers. The SAXS pattern of heat-treated PBZT (Fig. 13-16) shows strong equatorial scattering attributed to voids elongated parallel to the extrusion direction. One of the most prominent structural characteristics of high performance polymer fibers is their oriented fibrillar microstructure. Fig. 13-17 displays LM and SEM surface micrographs of UHMW PE fibers. The fiber surfaces appear rather uniform and smooth but with a fibril-looking texture, which is the consequence of high-drawing. Sometimes the fibers show kinks, which are the result of axial com-
Figure 13-15. TEM micrographs and SAED patterns (inset) of skin (a) and core (b) regions of Kevlar 149 fibers (Young et al., 1992). (a)
(b)
618
13 High Performance Polymer Fibers
Figure 13-16. SAXS pattern of heat-treated PBZT showing strong equatorial scattering caused by voids (Minter, 1982).
pression. A hierarchical structure for high performance polymer fibers has been proposed (Sawyer et al., 1993). As shown in Fig. 13-18, a fiber is envisioned to be comprised of macrofibrils, fibrils, and microfibrils, the typical diameters for these being 5 microns, 0.5 micron and 0.05 micron, respectively. The microfibrils consist of crystallites in combination with amorphous or disordered regions. Although a fibril structure is the common feature of high performance polymer fibers, the dimensions cannot always be discerned from the ratios indicated in the above model. In other words, different high performance polymer fibers have different fibril structures because of the various molecular structures and processing conditions. The PPTA molecule has a semi-rigid and extended chain structure. This rigidity is mainly attributed to the structure of aromatic rings connecting with the planar amide groups O
H
50
Figure 13-17. SEM surface morphology of UHMW PE fibers showing fibril structure, (a) optical micrograph; (b) SEM micrograph (DeTeresa, 1985).
which is resistant to rotation. The crystal structure of PPTA fiber is a monoclinic unit cell (modification I) with pseudoorthorhombic symmetry and two chains per cell {a = 0.773-0.784 nm, b = 0.5150.522 nm, c (repeat distance) = 1.28 1.29 nm, and y = 90°), as depicted in Fig. 13-19 (Northolt and Sikkema, 1991). Interchain hydrogen bonding occurs only along the (100) planes whereas weak Van der Waal forces operate in other lateral directions. The fiber density observed is 1.44 g/cm3 and the calculated (crystal) density is 1.52-1.54 g/cm3. By changing the spinning and coagulation conditions, different crystal unit cell structures are obtained. Spinning from sulfuric acid solu-
13.3 Fiber Morphological Structure
619
Oriented fiber
^lacrofibri ^ — 5.0u.m
\
Fibril 0.5|im Microf ibril Enlarged view of microfibrils
Figure 13-18. A hierarchical structural model for high performance polymer fibers, with more detail of the micro fibril sizes, shapes and order (Sawyer et al., 1993).
(a)
(b)
Figure 13-19. Crystal unit cell structure (modification I) of PPTA, (a) the view is perpendicular to the c-axis and parallel to the a-axis, (b) the view is down the c-axis (Northolt and Ven Aartsen, 1973).
620
13 High Performance Polymer Fibers
tions with a concentration of less than 9 % yields crystal modification II (Haraguchi et al., 1979a, b). It has almost the same unit cell dimensions, except the hydrogenbonded plane through the center of the unit cell is shifted along the 6-axis by a distance of 6/2. For the spinning dope with a concentration of 9-15%, both crystal modifications I and II coexist. When the concentration > 15 %, only modification I forms (Northolt, 1981). It has also been found that if the coagulant changes from methanol to water, the PPTA crystals change from modification I to II (Northolt and Ven Aartsen, 1973; Haraguchi et al., 1979a, b; Shen et al., 1983). The broadening of the equatorial reflection in the PPTA X-ray diffraction pattern is mainly caused by the lateral crystallite size distribution. The apparent crystallite size in various directions for Kevlar 49 is 6.4, 5.3, and > 10.9 nm measured from the (110), (200), and (006) reflections, respectively (Kumar, 1991). TEM imaging of the PPTA crystal lattice exhibits a very high degree of parallel alignment of the chains inside the crystallites (Dobb et al., 1975). The electron diffraction (ED) patterns demonstrate that within a domain size of about 0.5 jim the crystallites are almost perfectly oriented (Dobb et al., 1977 a, b). A distinctive structural feature of PPTA and some other aramid fibers is the pleated sheet structure, which was observed using the techniques of dark field TEM imaging from meridional reflections and polarized light microscopy. As modeled in Fig. 13-20 (Dobb et al., 1977a, b), the pleated sheets consist of parallel-oriented chains, which are hydrogen-bonded along the direction of the crystallographic 6-axis. They are usually oriented perpendicular to the surface of the fibers, i.e. in the circular crosssection the 6-axis is directed preferentially along the radius. The angle between adja-
Figure 13-20. Model of radial pleated-sheet structure of PPTA (Dobb et al., 1977).
cent planes of the pleat is about 170° and the distance between two pleats is about 150-250 nm, but both values may vary with fiber processing conditions. However, the pleat structure can be diminished by heat treatment under tension, especially with a super-heated stream (Fitzgerald and Irwin, 1991), and in the new, high modulus Kevlar product, Kevlar 149 (which has a different heat treatment process from the other Kevlar products), the pleated structure has apparently been straightened (Krause et al., 1989). Several different structural models have been suggested for the morphology of PPTA fibers (Fig. 13-21). They all share common structural features. (1) As mentioned above, the fiber is characterized by a pleated structure along the fiber axis and the pleated sheets are composed of fibrils. However, the highly oriented heat-treated fibers do not exhibit any lateral banding (Krause et al., 1989; Sawyer et al., 1993). (2) By means of TEM, field emission scanning electron microscopy (FESEM) and scanning tunneling microscopy (STM), it has been found that fibrils about 1 jim across consist of microfibrils about 10 nm
621
13.3 Fiber Morphological Structure
fill I
Core
a
]_J200-250nmf
\
I1
/
f \ \ \ \
111 11 \11 •
Chain with length 220nm Fibril axis
a
—| |—60nm Skin 0.1^m thick random chain end distribution
Shish-Kebab
70A
70A
"Lr
H600h nm
Longitudinal cleavage plane Zigzag fibril
Transverse cleavage plane (Spacing- 0.3/jm) \ Tangential
Epidermis Cylindrical lamella (Spacing 0.2-0.3yL/m) 10-15/im — .
Figure 13-21. Different models for PPTA morphological structure: (a) Northolt (Northolt and Ven Aartsen, 1977); (b) Morgan (Morgan et al., 1983); (c) Panar (Panar et al., 1983); (d) Manabe (Manabe, 1980); (e) Yabuki (Yabuki et al., 1975); (f) Horio (Horio, 1984) and (g) Li (Li et al., 1983).
622
13 High Performance Polymer Fibers
across. The smallest microfibrils are 3 to 30 nm wide and 2 to 5 nm thick. Both fibrils and microfibrils appear to be flat or tape-like, and readily form a layered structure (Sawyer et al., 1992, 1993). Most of the microfibrils are well aligned in the fibrils, but some are twisted and/or slightly tilted in relation to each other (Weng et al., 1986; Sawyer et al., 1992, 1993). It is considered that the radially oriented crystallites form the main, highly ordered fibrillar structure which constitutes the basic load bearing element of a fiber. (3) There is a skin-core differentiation in PPTA fibers. The chain ends in the skin are arranged essentially randomly relative to one another, but become progressively more clustered in the fiber interior (Northolt, 1981; Northolt and Sikkema, 1991). Elongated voids with their long axis approximately parallel to the fiber axis are sometimes formed in the core. In addition, kink bands observed on fiber surfaces are also seen in the bulk (Sawyer et al., 1992). PBZO and PBZT are inherently stiff molecules. Measurement of the orientation distribution indicates a highly uniform orientation of the molecules parallel to the fiber axis, but certain instances of local misorientation have been observed. Figure 13-22 shows 3-4° misorientation between adjacent crystallites within 20 nm in the high resolution electron microscopy (HREM) image of a PBZO fiber (Martin and Thomas, 1991a). SEM and TEM show that the ordered molecular chains comprise crystallites, which form fibrils/ microfibrils, and that the fibrils/microfibrils are predominantly lath-shaped (Allen et al., 1985). For example, PBZO consists of ribbon-like fibrils with widths from about 30 nm up to a few micrometers. The crystallites are about 5-20 nm wide and 5-10 nm long. Occasionally, crystallites of up to 30 nm long and 40 nm in lateral ex-
Figure 13-22. HREM image of PBZO showing local misorientation between adjacent crystallites (Martin and Thomas, 1991a).
tent were observed (Krause et al., 1988). Comparison of the crystallite structures for Kevlar, PBZO and PBZT shows that all the molecules, crystallites and microfibrils/fibrils are well aligned along the fiber-axis. The monoclinic crystal unit cell and lattice parameters of PBZO are similar to those of PBZT [Fig. 13-23 (Fratini et al., 1989)]. For PBZO, a = 1.120nm, b = 0.354 nm, c = 1.205 nm and y = 101.3°. The Z?-axis length is roughly the perpendicular distance between the faces of heterocyclic rings between chains, the a-axis length is roughly the distance between equivalent edges of heterocyclic rings adjacent in the chains, and the oaxis is the distance along the chain of a single repeat
13.3 Fiber Morphological Structure
unit; for PBZT, a = 12.2 nra, b = 3.65 nm, c = 12.5 nm and y = 106.2°. The difference in lattice parameters between PBZO and PBZT is that in PBZT the a-axis is slightly larger than in PBZO, possibly due to the non-planarity of PBZT molecules and the larger diameter of the sulfur atoms. The calculated crystal densities for both the fibers are 1.66 g/cm3, while the measured heat-treated fiber densities are 1.58 g/cm3 (Krause et al., 1988; Martin and Thomas, 1991 a). Crystallites of PBZO fibers are 3dimensionally ordered and exhibit limited long range order, as determined from WAXS and SAED patterns. PPTA fibers have greater three-dimensional (3D) order of their crystallites and also greater long range order. In PBZO, the planar molecules have improved chain registry, thus promoting 3D crystallinity. However, the PBZO crystalline structure is not as well ordered as that of PPTA. In WAXS and SAED measurements, only first order (hkl) reflections are clearly present in PBZO (as shown in Fig. 13-24), whereas many higher order (hkl) reflections are observed in PPTA. In PBZT, the non-planar molecules appear to have only 2D crystal order with less long range periodicity. Be-
(b)
Figure 13.23. C-axis view of the molecular packing of PBZT (a) and PBZO (b), the a-axis is horizontal (Fratini et al., 1989).
623
Figure 13-24. SAED patterns of PBZO (Adams et al., 1986).
cause of the axial translational disorder, no off-axis (hkl) reflections are observed in WAXS measurements of PBZT. Another feature of the crystallites is that through heat treatment, the crystallites in PBZO grow in the (hkO) directions by a factor of about 2 while in PPTA and PBZT, the crystallites grow in the (00/) directions by a factor of roughly 2. Therefore, the transverse dimensions of heat-treated PBZO crystallites are generally equal to or greater than those of PPTA and PBZT, while the axial dimensions are less than those of PPTA and PBZT by a factor of one half to one eighth (Krause et al., 1988). A significant skin-core difference is also found for PBZT and PBZO fibers (Krause et al., 1988; Young et al., 1992), which is similar to that for the PPTA fibers. For PBZO, this effect is on a scale of 10-100 nm (Krause et al., 1988). By measuring the intensity spread of (hkO) electron diffraction reflections, the skin material is found
624
13 High Performance Polymer Fibers
to have a better defined structure and a higher degree of molecular orientation than the core material (Young et al., 1990, 1992). The special features of Technora fiber are that its fibrils are composed of ordered crystallites and highly oriented amorphous regions, and that it exhibits a uniform, dense fiber morphology. However, because of the presence of amorphous regions and the absence of lateral crystalline order, Technora cannot attain as high a modulus as Kevlar (Yang, 1989). Gel spun HPOPFs also fit the "standard" morphological model for high performance fibers, where the oriented fibrils/ microfibrils contain a number of crystallites in series. In Spectra 1000, the fibrils are about a thousand angstroms long and 5-20 nm wide, as measured by synchrotron X-ray diffraction (Grubb and Prasad, 1992 b). By means of atomic force microscopy (AFM), Sheiko suggested that the elemental fibril unit in UHMW PE has a diameter of about 20-90 nm and named it a "nanofibril". The fibril structure is formed by the nanofibrils. The diameter of the fibrils observed by electron microscopy and X-ray diffraction is strongly affected by the draw ratio and processing conditions, but the diameter of nanofibrils remains fairly constant (Sheiko et al., 1992). One of the unique structural characteristics is that a shish kebab morphology can be obtained in UHMW PE fibers, except in the case of extremely high draw ratios (> 200). The shish kebab structure consists of core microfibrils with nearly extended chains and a number of chain folded platelets (lamellae) attached to the cores [Fig. 13-25 (Pennings, 1977)]. At draw ratios of greater than about 200, UHMW PE fibers consist mainly of smooth microfibrils with nearly extended molecules, while the amount of lamellar platelets is
Figure 13-25. A schematic representation of a shish kebab model showing the intimate connections of crystal platelets with the central core (Pennings, 1977).
negligible (Hofmann and Schulz, 1989). No long period or only a weak long period was detected by X-ray diffraction of the fully drawn fibers, which indicates that chain folding would be rare in fibrillar units (Van Hutten et al., 1985; Grubb and Prasad, 1992 a). The microfibrils of UHMW PE fibers have a finite length of about 1000-2000 nm, with an aspect ratio of (length/diameter, L/£>) 250-500 (Schaper et al., 1989). In addition, the presence of a weak intensity, two-spot S AXS pattern indicates a weak long period of about 150-200 nm (Grubb and Prasad, 1992 b). The structure of microfibrils is thus not uniform in density, but contains alternating regions of different density attributed to perfect crystals and domains containing a high concentration of crystal defects. A WAXS study of Spectra 1000 PE fibers revealed that the crystal unit cell is essentially the same as that reported originally (Bunn, 1939): orthorhombic with a = 0.741 nm, b = 0.495 nm and c = 0.255 nm [Fig. 13-26 (Busing, 1990)]. With increasing draw ratio, the crystallite length increases from 30 to 70 nm and the lateral size of the crystallites is 5.2-5.9 nm (Murthy, 1990; Grubb and Prasad, 1992 a). Measurements of Raman spectra and synchrotron WAXS (SWAXS) on single fibers show that Spectra 1000 is not perfectly
13.3 Fiber Morphological Structure
625
(a)
Figure 13-26. Crystal unit cell structure of Spectra 1000 PE fiber, (a) the view down the fiber axis, (b) the view perpendicular to the c-axis and parallel to the 6-axis (Busing, 1990). (b)
crystalline, but generally contains crystallites and "amorphous" regions. The crystallinity, as normally measured, is in the range of 85-95%. The "amorphous" (or disordered) region is not detectable by WAXS but by the Raman band at 1063 cm" 1 . The order within the "disordered" region is similar to that of a nematic liquid crystal. This partially ordered region is generally associated with tie
molecules, which connect the crystallites in and between the microfibrils (Prasad and Grubb, 1990). It has been reported that only a minor portion of the intercrystalline tie molecules are taut (Hofmann and Schulz, 1989). Electron microscopy also indicates that these disordered domains, whose longitudinal dimension appears to be 4 - 5 nm, are "amorphous" domains that may contain a substantial amount of
626
13 High Performance Polymer Fibers
chain ends. This "amorphous" domain is covalently bonded to the adjacent, nearly perfect, needle-like crystalline domains whose L/D is about 40 (Prevorsek et al., 1992). Both crystallites and microfibrils show very high orientation along the fiber axis direction. A recent WAXS study of a bundle of Spectra 1000 fibers has given a variation for the axial misorientation of only 0.69° (Busing, 1990), the corresponding Hermans-Stein orientation factor, defined a s / = {3
13.4 Fiber Properties All high performance organic polymer fibers display outstanding ultimate tensile strength and elastic modulus, two of the most elementary fiber properties. Most high performance fibers also exhibit excellent hydrolytic and chemical resistance. In addition, aromatic fibers show good ther-
mal stability and are suitable for use in the temperature range of 300 to 450 °C, depending upon the environment and exposure time. Most of the high performance organic polymer fibers have high radiation resistance, but they are not as resistant as carbon fibers. A prolonged exposure to light and ultra violet (UV) radiation is detrimental to the mechanical properties of aramid fibers such as Kevlar. Normally, the HPOPFs have good dielectric properties, but poor electrical conductivity. Some conjugated fibers also exhibit interesting non-linear optical behavior. One of the weak points for most of the high performance polymer fibers is their comparatively poor axial compressive strength. If this weakness could be overcome, it would be a breakthrough in their market share compared to their primary competitors, the carbon fibers. 13.4.1 Tensile Properties
Table 13-1 shows the tensile properties of various fibers. Tensile moduli of the high performance polymer fibers are at least four times higher than for polyester, at least nine times higher than for high tenacity nylon, and at least two times higher than for steel and glass fibers. Figure 13-2 is a plot of specific tensile strength versus specific tensile modulus of various fiber materials. It is interesting that all the high performance polymer fibers rank high, especially UHMW PE due to its low density and PBZO due to its superior tensile properties. Typical stress-strain curves of some fibers are given in Fig. 13-1. In general, their elongation at break is relatively low, about 1-4%. Fig. 13-4 shows the predicted theoretical chain moduli of some fibers (Donald and Windle, 1992). Recent advances in computational materials science allow the full quantum me-
627
13.4 Fiber Properties
chanical calculation of the intrinsic axial molecular modulus of polymer chains (Wierschke, 1989, 1992), as well as its strain and temperature dependence (Klunzinger et al., 1991 a, b). These calculations, while still at early stages of refinement and validation, serve to establish upper limits for the achievable tensile modulus. To date, few fibers in practice exceed 50 % of these theoretical values. The theoretical tensile strengths of Kevlar, PBZT, PBZO and UHMW PE extended-chain fibers are all more than 20 GPa (Ohta et al., 1988). If the polymer chains in the test fiber were as long as the gauge length and perfectly parallel to the fiber axis, the value of the measured stress at break would be the same as the theoretical one. This stress, which is solely determined by the covalent bonds and lateral packing density of the main chains, is estimated to be 10-15% of the theoretical chain modulus (Kelly and MacMillan, 1986). However, calculation of the tensile strength of real fibers is rather complicated due to many factors such as chain length distribution, the misorientation of the chain with respect to the fiber axis, different connections between atoms (covalent bonds in the main chain and secondary forces between chains), the non-uniformity and anisotropy of the fiber structure and morphology, as well as the presence of impurities and voids. Different ones or combinations of these are important in different situations. We shall present here some of the considerations and hypotheses that have been put forward. The Weibull model, which is based on the "weakest link theory", has been widely used for the statistical description of the fiber tensile strength (Weibull, 1951). This model assumes that fiber tensile strength is governed by pre-existing strength-limiting defects, which are randomly distributed
throughout the fiber. Thus the cumulative failure probability function, P, which represents the fraction of fibers that fail at or below a stress a is given by (13-4) where l0 is the test length of the fiber,
65""o
4-
_j
c
—
3-
CL
i
2-
^
1-
jy
y
oL0=25mm -13
3.5 4 4.5 Strength (GNrrf2)
5
Figure 13-27. A Weibull model fitting of the fracture stress distribution of a PPTA (Twaron) fiber with m = 10.6 (Northolt and Sikkema, 1991).
628
13 High Performance Polymer Fibers
strength and lifetime under constant stress of single PPTA fibers using Weibull statistics discloses that after short creep times, the fiber failure mechanism is transverse crack propagation, but after long creep times, it changes to splitting and fibrillation (Wu et al., 1988). It is suggested that because of the fibril morphology, if the shear forces at a discontinuity exceed the shear strength between the fibrils, the fiber tensile strength should be proportional to the fiber shear strength (Knoff, 1987). There is another argument that for the high performance organic polymer fibers, flaws might have a relatively small detrimental effect on the tensile strength (Donald and Windle, 1992). Due to their highly oriented fibril structure and comparatively weak lateral interchain bonding, the transverse cracks are either effectively blunted by intermolecular shear, or deflected to run along the fiber axis direction. The study of fracture morphology plays an important role in understanding the factors controlling the fiber strength. Disregarding the effect of stress concentration due to impurities, it may be expected that brittle fracture, characterized by a small area fracture surface oriented normal to the fiber axis, probably results from normal stress on the crystallite (directed along the c-axis), while a fibrillated, large area fracture surface is caused by shear stress (Northolt and Sikkema, 1991). All the aromatic polymer fibers produced by the dryjet wet-spinning technique display a fibrillated fracture surface. The possible reason is that due to the pre-existing impurities and inhomogeneities and a strong anisotropy in the axial and lateral moduli, a large stress concentration arises near a crack tip. The maximum shear stress is parallel to the direction of the largest modulus value (Kelly and MacMillan, 1986), which would ultimately cause a fibrillated
fracture surface. Thus, for highly oriented fibers containing impurities or inhomogeneities the tensile strength might not depend on the fiber modulus (Northolt and Sikkema, 1991). It has also been found that no apparent differences in the fracture morphology are seen in PPTA fibers broken under simple tensile, fatigue and creep conditions (Yang, 1989). The tensile properties, especially the tensile strength, of the high performance polymer fibers generally depend on their molecular weight. For UHMW PE, tensile strength and modulus increase approximately with the logarithm of the molecular weight. When the molecular weight reaches 3 x 106, the experimental tensile strength is about 14% of the theoretical value and the experimental modulus is 65% of the theoretical one (Ohta et al., 1988). Comparison of experimental data published by different authors shows that the highest modulus achievable for UHMW PE fibers appears to depend mainly on the molecular weight (Wang et al., 1990). As mentioned in Sect. 13.2.3.2, the tensile strength of Kevlar fibers increases with increasing intrinsic viscosity of the spinning dope (increasing molecular weight). Molecular weight also affects the mechanical properties of PBZO fibers (Ledbetter et al., 1989). A PBZO/PPA solution with an intrinsic viscosity of approximately 30 dl/g gave fibers with the highest tensile modulus and strength. For heattreated Vectran fibers, there is a correlation between the tensile strength, the molecular weight and the molecular weight distribution (Yoon, 1990). Based on the theoretical study of the influence of the molecular weight (M w ) on the tensile properties of polymer fibers, it has been suggested that for Kevlar and UHMW PE fibers, the molecular weight drastically affects the tensile strength, but only moder-
13.4 Fiber Properties
ately affects the Young's modulus (Termonia et al., 1985 a, Termonia and Smith, 1986, 1987 a, b; Smith and Termonia, 1989). In particular, for the UHMW PE fibers the modulus tends to be constant for M w > 1 0 4 (Termonia et al., 1985b). It is also predicted that the nature of the fiber failure process would gradually change from chain slippage to chain fracture with increasing molecular weight (Termonia et al., 1985b). Figure 13-28 shows the significant effect of orientation with respect to the direction of applied stress on the effective Young's modulus for a PBZT single crystal (Jiang, 1989). It is clear that for small angles of orientation, a small change in angle can cause a large change in modulus. Thus the orientation produced irreversibly by the processing of a highly oriented crystalline polymer can have a strong effect on the tensile modulus. The orientation produced by processing can also be increased reversibly by the application of tensile stress
2
4
6
8
10
12
14
16
18
or(Degree)
Figure 13-28. A schematic diagram of the effect of misorientation angle a on the effective Young's modulus for a PBZT single crystal (Jiang, 1989).
400
800
1200
629
1600
Stress (MPa) Figure 13-29. Sonic modulus at room temperature as a function of static tensile stress for PBZT fibers: as-spun, heat-treated at 525 and 650 °C. The tested fibers were, under static tension, subjected to pulses from a Q-switched Nd: YAG laser. The resultant ultrasonic wave speed was used to calculate the sonic modulus (Jiang et al., 1989).
at room temperature. There is a concomitant reversible increase in the tensile modulus of the fiber (nonlinear elasticity). The effects of processing conditions and tensile stress applied at room temperature on the modulus are shown in Fig. 13-29 (Jiang et al., 1989,1990). The corresponding changes in the orientation are shown as the halfwidth at one-half of the maximum intensity of the orientation distribution. The change of orientation with applied stress at room temperature accounts for most, but not all, of the observed change in the modulus (Jiang, 1989). A stiffening of the
630
13 High Performance Polymer Fibers
molecules themselves with increasing stress also plays a role (Klungzinger et al., 1992 a). Possibly there are some other factors involved which are not well understood yet (Jiang et al., 1989, 1990). As will be discussed later, temperature also affects the nonlinear elasticity. Nearly all of the HPOPFs show creep and stress relaxation phenomena, which are indicators of viscoelastic behavior. However, the rates of creep and stress relaxation for HPOPFs are extremely low compared with the conventional textile fibers like nylon and polyester. For PPTA, the observed creep strains amount to less than 20 % of the initial elastic strain after several years under stress (Walton and Majumdar, 1983). Presumably this low creep is caused by the combination of rigid chains and high crystallinity. One of the unique characteristics of Vectran is its lack of creep under static loads of up to 50% of the breaking strength (Beers and Ramirez, 1990) (also see Chapter 5 of this Volume). Figure 13-30 (Beers and Ramirez, 1990) shows the results of stress-relaxation tests for Vectran HS (high strength Vectran fibers), Kevlar 29 and Spectra 900 ropes. It is clear that no relaxation is oberved for the Vectran rope. By comparison, UHMW PE fiber has the highest relaxation. The fatigue behavior of high performance fibers has been a major concern in light of their poor compression behavior. In the fatigue test of cyclic bending over a freerunning spindle, it is found that Kevlar 29 and 49 fibers exhibit rapid strength loss in the first few cycles, but a more gradual decline afterwards. Kevlar 49 has a greater strength loss than Kevlar 29 (Dobb, 1985). During cyclic axial compression, after 100 cycles, the strength loss of Kevlar 49 was only about 10%, even though the compressive strain of 1.2 % exceeded the strain required for kinking (DeTeresa et al.,
Load, (lbs.) 7000
6000 -
5000 -
4000 -
3000 -
2000 10 100 Log time, (hours)
1000
Figure 13-30. A stress relaxation test for different fibers. The tested samples were tensioned to a fixed load and the tension was recorded periodically from load cells at each end of the samples. As creep occurred, the load decreased and the samples were subsequently retensioned back to the original load (Beers and Ramirez, 1990).
1984) (see next section for discussion of compressive kinking). The tensile properties of the high performance polymer fibers are temperature dependent. At 250 °C, heat-treated PBZT fibers retained 80 % of their room temperature tensile modulus and about 70% of their room temperature tensile strength (Allen et al., 1985, Allen and Farris, 1989; Im et al., 1989; Kumar, 1990a, 1991). However, the fiber which was heated to 300 °C for 65 hours in air and then tested at room temperature did not indicate any loss of tensile strength and modulus (compared to the fiber at room temperature) (Kumar, 1990). For some fibers, such as PBZT, the nonlinearity increases with increasing temperature. In the 300-400°C range, PBZT exhibits a mechanical relaxation that is associated with a structural
13.4 Fiber Properties
change in the same range (Feldman et al., 1987; Jiang et al., 1989, 1990; Klunzinger et al., 1992a; Macturk et al., 1993).
631
5-
ID
13.4.2 Compressive Properties
Compared with the superior tensile properties, the axial compressive strength of high performance polymer fibers is disappointingly low. The ratio of compressive strength to tensile strength is only about 10 to 20%. This is one of the obstacles preventing high performance polymer fibers being even more widely used in structural composites. While composites can fail in compression by a variety of modes, an important one results from the weak axial compressive strength of the fiber itself. Compressive strengths are listed in Table 13-1. It is clear that the compressive strength of highly oriented polymer fibers is low, while isotropic boron, alumina and silicon carbide fibers are among the highest compressive strength fibers available. PAN-based carbon fibers exhibit compressive strengths superior to pitchbased carbon fibers. The compressive behavior of fibers has been extensively studied in the last decade (DeTeresa et al., 1982, 1984, 1985a, 1988; Drzal, 1986; Allen, 1987; Kovar et al., 1989; Kumar et al., 1988; Kumar and Helminiak, 1989b; Kumar, 1989, 1990a, b, 1991; Kumar and Adams, 1990; Martin and Thomas, 1989, 1991a, b; Huh et al., 1990; Jiang et al., 1991 a, b). There is also nonlinearity in compressive behavior (Macturk et al., 1991; McGarry and Moalli, 1991 a, b; Klunzinger et al., 1991 b, 1992 b). Contrary to nonlinear behavior in tension, with increasing axial compressive stress the compressive modulus decreases, the so-called "modulus softening" phenomenon. Figure 13-31 shows a typical stress-strain curve for the high perfor-
32-
-2
-1
3
4 5 Strain (%)
Figure 13-31. Typical tensile (a-c) and compressive (d) stress-strain curves of PBZO fibers: (a) as-coagulated (wet), (b) as-spun (dried), (c) and (d) heattreated (Martin and Thomas, 1989; Kumar, 1990 a).
mance polymer fibers under tension and compression (Martin and Thomas, 1989; Kumar, 1990 a). The compressive modulus (Ec) may be less than the tensile modulus (Et), and some theoretical estimates predict Ec to be 70-90% of Et (McGarry and Moalli, 1991a). However, it was also reported that the compressive modulus appears approximately equal to the tensile modulus (Fawaz et al., 1992; Klunzinger et al., 1992 b). Under compression, the fiber birefringence decreases corresponding to a crystal disorientation away from the fiber axis. This disorientation would contribute to the decreasing modulus and could play a role in compressive failure (Klunzinger et al., 1992b). There is no simple correlation between the axial compressive behavior and tensile properties. For most of the high performance polymer fibers, an improvement in their tensile modulus does not change their compressive strength sig-
632
13 High Performance Polymer Fibers
nificantly (Van der Zwaag and Kampschoer, 1988; Van der Zwaag et al., 1989). On the other hand, the formation of kink bands during compressive deformation causes only a small loss in tensile strength (DeTeresa et al., 1984,1985, 1988; Van der Zwaag and Kampschoer, 1988; Van der Zwaag et al., 1989). These results imply that the compressive failure process depends upon different structural factors from those controlling tensile strength. Several HPOPFs, such as PPTA, PBZT and UHMW PE show a linear relation between compressive strength and shear modulus. However, the compressive strength of most high performance polymer fibers is in the range of about V* to % of their shear modulus (Kumar, 1992). Compressive failure of high performance polymer fibers is always associated with the formation of kink bands (shown in Fig. 13-32). Kink bands are nucleated in a localized region, normally initiated somewhere near the fiber surface, at a certain critical stress, and then propagate away from the initiating point and con-
tinue to grow in towards the center with the other side of the fiber under compression. For different fibers, the strain at which the first kink band occurs is different (Van der Zwaag and Kampschoer, 1988; Van der Zwaag et al., 1989) due to their variation in structural order. For PBZT fibers observed by optical microscopy, there are no kinks formed up to about 0.1 % compressive strain. When the strain increases above 0.1%, kink bands start to appear and the number of kink bands increases suddenly. Most new kink bands occur near the initial kink bands, showing some periodicity in the concentration of kink bands (Huh et al., 1990). For a certain range of compressive strain, the kink band density is proportional to the applied compressive strain, but at a larger compressive strain, the kink band density tends to become saturated (Takahashi et al., 1983; Huh et al., 1990). Upon the application of axial tensile strain after compressive kinking, kink bands have been reported to disappear, at least on the scale observed by LM, but this "reversibility"
Figure 13-32. SEM image of kink bands on Kevlar 49 fiber surface (a) as-received; (b) after 3 % axial compression, kink bands form on the fiber surface; (c) after > 3 % axial compression, more kink bands form. Compression due to nylon matrix shrinkage (DeTeresa, 1985b).
13.4 Fiber Properties
phenomenon has not been confined yet by higher resolution techniques (Huh et al., 1990). It was reported that the kink band angle related closely to the fiber structure and properties (Van der Zwaag and Kampschoer, 1988; Van der Zwaag et al., 1989). However, it is easy to obtain incorrect information by using optical microscopy to observe the angle of the kink bands, because of insufficient resolution for clearly distinguishing individual kinks (Martin and Thomas, 1991; Vezie et al., 1992). Furthermore, only in the correct position, namely, the in-plane position, would one obtain the correct angles: a, the angle of change in molecular orientation across the kink boundary, and /}, the kink boundary angle [shown in Fig. 13-33 (Vezie et al., 1992)]. There is clearly a correlation between the structure and the two angles. If p = a/2, the boundary is symmetric and the kink is isovolumetric. This is the case for compressive kinks in the crystallites of PBZO fibers (Martin and Thomas, 1991 a, b). If p > a/2, the boundary is nonsymmetric and the volume inside the kink increases (tension between the structural units). Most cases, such as kinks in the fibrils and microfibrils of PBZO fibers, belong to this category (Martin and Thomas, 1991b). If /?
633
a = change in molecular orientation across kink boundary P = kink boundary angle = angle between normal to kink boundary and molecular axis x0 = distance between chains microfibrils or width of whole fiber in unkinked region x1 = length of kink boundary W = perpendicular distance between chains or microfibrils within the kink, or the transverse width of the kink ff=
cos/? P ^ 90°, a - p # 90° (a - 90°, p = 0 => not physical) Figure 13-33. Measurement of kink band angles (Vezie et al., 1992).
(Martin and Thomas, 1989, 1991b). A possible mechanism for this reorientation has been investigated by computational modeling (Wierschke, 1989; Klunzinger and Eby, 1992). There are several different arguments about the mechanisms of the kink formation and the compressive failure. Argon pointed out that for composite materials, kink band formation occurs when a compressive stress (a) along the orientation direction causes locally misaligned elements to experience a relative shear stress which depends on the misorientation angle, # 0 . The compressive stress would be given by o =
(13-5)
634
13 High Performance Polymer Fibers
where TS is the plastic shear strength of the matrix (Argon, 1972). DeTeresa proposed that the kink formation relates to an elastic buckling instability in high performance polymer fibers and there is a correlation between the compressive strength and the longitudinal shear modulus of the fiber (DeTeresa et al., 1982, 1984, 1985, 1988). Huh found that the compressive strength of high performance polymer fibers is not limited by fiber buckling, but by fibril buckling according to the Euler buckling model (Huh et al., 1990). Van der Zwaag argued that the kink bands are formed before elastic instability occurs, so they may therefore be attributed to a plastic deformation process (Van der Zwaag and Kampschoer, 1988; Van der Zwaag et al., 1989). Northolt suggested that compressive deformation and the glass transition have common origins in chain flexibility and intermolecular interactions; hence an empirical relation holds between the compressive strength and (Tg)2 (Northolt, 1981). This result can also be derived quantitatively by assuming that the work for compressive yielding is proportional to the activation energy of the glass transition (Northolt and Sikkema, 1991). Attenburrow and Basset investigated the kinking of chain-extended PE and found that the extended chain deformation facilitates "fine" shear (Attenburrow and Bassett, 1979). Shigematsu proposed that the uniform oaxis shear and the intercrystalline slip mechanisms are both involved in forming kink bands (Shigematsu et al., 1985). It is also noted that kink formation is related to crystalline slip (Takahashi et al., 1983; Miwa et al., 1991a, b). For UHMW PE, (110X001 > and (100)<001> slip systems (Takahashi et al., 1983) are associated with kink formation; for Kevlar, the (200)<001> (Takahashi et al., 1991 a b); for PBZT the (010)<001> (Roche
et al., 1980) and for the aromatic polyester fiber, the (110)<001> and the (100)<001> (Takahashi et al., 1991a, b; Xiao and Takahashi, 1990). However, many unanswered questions still remain in this field. Up to now an outstanding problem in the study of compressive behavior of a single fiber is that there is no satisfactory technique to measure the axial compressive strength of a single fiber owing to its easy buckling. Several methods, most being indirect ones, have been developed, such as the bending beam (DeTeresa, 1982, 1984, 1985), elastica loop (Sinclair, 1950; Greenwood and Rose, 1974), tensile recoil (Allen, 1987), single fiber embedded composite (Drzal, 1986), fiber broken pieces in composite (Ohsawa et al., 1990; Miwa etal., 1991a, b) and the composite test. Direct methods are single fiber compression using a micro-tensile testing machine (MTM) (Fawaz et al., 1992) and "nanocompressometer" (Macturk et al., 1991). Little has been done to establish the correlation between the data obtained on the same fiber and these different tests. Therefore, data from different methods must be compared with caution. Significant research efforts have focused on attempting to enhance the compressive properties of the high performance fibers. Changing process conditions, morphological structures and modifying synthetic routes, increasing shear strength by bonding molecular chains more strongly together, coating the fibers as well as infusing monomer into the fibers then curing, are all ways to improve compressive behavior. In some cases, compressive strengths of 500-700 MPa were reported (Bhattacharya et al., 1989; Chuah et al., 1989), but most were in the range of 2 5 400 MPa (Dang et al., 1989; Kumar and Helminiak, 1989 a). Due to lack of reliable test techniques for the single fiber com-
635
13.4 Fiber Properties
pressive strength and the nonlinearity of the compressive behavior, the validity of these results may be questionable. It was also reported that the presence of glass within PBZT/sol-gel microcomposite films could increase the resistance of the film to compression (Kovar et al., 1989). Similar treatment of PBZT and PBZO fibers may be expected to produce PBZT/sol-gel and PBZO/sol-gel microcomposite fibers with improved compressive strength (Kumar, 1990 a). Unfortunately, no significant success has yet been reported. 13.4.3 Thermal Properties
All the aromatic high performance fibers have demonstrated great thermal stability because of their aromatic chemistry, rigid molecular chains and high crystallinity. Kevlar does not undergo decomposition until 480 °C in N 2 and about 380 °C in air (Yang, 1989). For PBZT and PBZO, the thermal behavior is even better. The thermogravimetric analysis (TGA) measurements show that the degradation temperature for PBZO in air is 620 °C. At 371 °C in air for 200 h, the weight retention in PBZO is reported to be 78% and in
PBZT, 70%. Kevlar, PBZT and PBZO fibers do not melt below the decomposition temperature and do not support combustion, but will char at high temperature. They generate only low quantities of smoke when burning (Goldfarb et al., 1989). HPOPFs as a group have negative and small axial coefficients of thermal expansion (CTE). For example, for Kevlar, ac = - 2 x 10" 6 /°C and for PBZT, - 1 to - 2 . 5 x l O " 6 / ° C (Yang, 1989). For the latter, the coefficient extrapolates to — 7 x 10 ~ 6 at perfect orientation (Lusignea, 1989). The negative values arise from transverse vibration of the molecules, as has been demonstrated by molecular dynamics simulations and molecular orbital calculations (Klunzinger et al., 1991a, 1992 a; Klunzinger and Eby, 1993; Macturk et al., 1993). Therefore, these fibers have very low shrinkage at high temperature and good thermal dimensional stability. Table 13-7 lists the CTE data for wet and dry PBZO and Kevlar 49 fibers (at two loads) for various heat cycles. The CTEs for a given tension were approximately the same for all heat cycles. The results from wet PBZO at low tension were also similar to those of their dry counterpart. For wet
Table 13-7. Axial CTEs (K" 1 x 10" 6) of Kevlar 49 and PBZO fibers from thermal mechanical analysis3 (Im et al., 1989). Fiber
PBZO PBZO PBZO Kevlar 49 Kevlar 49 Kevlar 49 a
Condition
Dry Dry Wetb Dry Dry Wetb
Applied load
Thermal cycles
(MPa)
1
2
3
4
1.24 62.1 1.24 0.62 31.0 0.62
-10.8 -7.6 -10.3 -11.0 -7.8 -1.5
-10.5 -8.2 -10.8 -11.8 -7.8 -9.0
-10.6 -7.6 -11.3 -11.4 -7.7 -9.6
-10.8 -7.4 -11.9 -11.3 -7.7 -10.3
Heating rate of 5 °C/min under a nitrogen environment; Wet fiber specimens were prepared by soaking them in water for 1 min, followed by conditioning at 100% relative humidity at room temperature for 24 hours. b
636
13 High Performance Polymer Fibers
Kevlar 49 the first scan was significantly different from the subsequent scans, from - 1 . 5 to 9.0 x 10~6/°C. This difference between wet PBZO and Kevlar 49 is due to the difference in their moisture retention characteristics. When exposed to a temperature range of 25-250°C for 24 hours, Vectran HS has a strength retention of about 85%, better than Kevlar 29. However, when treated at elevated temperature Vectran does not retain its strength as well as the solutionspun aramid fibers because of its thermoplastic characteristics (Beers and Ramirez, 1990). Figure 13-34 shows the stressstrain curves of UHMW PE fiber (Dyneema SK60) at different temperatures. With decreasing temperature (down to — 30 °C), its tensile strength and modulus increase. At 130°C for 3 hours, the retention of tensile strength and modulus is about 80% (Ohta et al., 1988). 13.4.4 Miscellaneous Properties High performance polymer fibers are hydrolytically stable and resistant to most
solvents. For evaluation of hydrolytic stability, PBZT fibers exposed at 120°C to a pH 5 buffer solution were found to be extremely stable compared to other reinforcing organic fibers. After the fibers were dried at 250-300°C in a vacuum, the moisture regain was < 1 wt.%. These results are to be expected as PBZT will not undergo hydrolysis. While most organic solvents have no effect on Kevlar fiber, it is attacked by strong acids or bases at high temperature and at high concentrations. Aromatic polyester fibers are less resistant to solvents than UHMW PE. Table 13-8 shows that after being immersed in various chemical solutions for a period of six months, Spectra fiber retained almost all of its original strength (Allied Signal Fiber Data Sheet, 1990). The study of radiation stability for a number of polymers disclosed that high performance polymer fibers are the most electron radiation-resistant (Kumar and Adams, 1990b). However, Kevlar is degraded by UV radiation. The affected surfaces are self-screening so that thick fabrics
Table 13-8. Strength retention (%) after immersion for six months (Allied Signal Fiber Data Sheet, 1990).
-30°C
Agent
25°C 80°C
Strain (%)
Figure 13-34. Effect of temperature on stress-strain behavior of UHMW PE fiber (Dyneema SK60) (After Science and Technology Center of Osaka, 1987).
Sea water 10% Detergent solution Hydraulic fluid Kerosene Gasoline Toluene Perchlorethylene Glacial acetic acid 1 M Hydrochloric acid 5 M Sodium hydroxide Ammonium hydroxide (29%) Hypophosphite solution (10%) Clorox bleach (5.25% Sodium hypochlorite)
Spectra 900
Aramid fiber
100 100 100 100 100 100 100 100 100 100 100 100 91
100 100 100 100 93 72 75 82 40 42 70 79 0
637
13.4 Fiber Properties
retain most of their original strength, although thin fabrics may lose 50% of their strength on exposure to Florida sunlight for 5 weeks (Fitzgerald and Irwin, 1991). PBZT fiber has good tensile strength retention after exposure to UV radiation. Both as-spun (AS) and heat-treated (HT) PBZT fibers exposed to UV radiation in the carbon-arc weather-o-meter (GM test) exhibited little or no strength loss over the course of the experiment. Figure 13-35 is a comparison of strength retention of different fibers under light radiation. UHMW PE fiber exhibits excellent light radiation stability (Allied Signal Data Sheet, 1990). For most high performance polymer fibers except UHMW PE, dimensional stability is excellent. For example, Kevlar shows negligible dimensional change in dry air at 160°C, in boiling water or at high relative humidity, even though moisture pick-up may be as high as 7.5% (Fitzgerald and Irwin, 1991). High performance polymer fibers also have very good dielectric properties. The high dielectric strength coupled with elevated temperature resistance and outstanding moisture resistance provide electrical efficiency in prevention of current leakage. This combination along with dimensional stability and low CTE make the high performance polymer fibers useful for specialized electronic uses. The dielectric constant and loss tangent of composite materials are generally considered to be a measure of radar transparency. The type of fiber selected would dominate the resulting composite's dielectric properties because composites usually have a fiber-to-resin ratio of 60:40. In practice, most transmissivity problems encountered in radome materials are caused by reflection of the electromagnetic energy at the interface of two materials having different dielectric constants. The smaller
HP-PE
100 300 500 700
1000
1500
Time (hr)
Figure 13-35. Comparison of light stability for different fibers (Ward, 1987).
Table 13-9. Dielectric constants of some composite materials (DuPont Product Literature, 1988; Anderson and Nguyen, 1992; Cordova and Collier, 1992; Harper et al., 1992).a Fiber/matrix
Spectra 1000/polyester Kevlar 49/polyester Quartz/polyester Glass/polyester Polyester resin only Spectra 1000/epoxy Kevlar 49/epoxy Quartz/epoxy Glass/epoxy Epoxy resin only a
Dielectric constant
Loss tangent
2.4 3.5 3.5 4.3 2.7-3.2 2.7 4.1 3.5 4.2 3.0-3.4
0.007 0.025 0.017 0.015 0.005 0.02 0.017 0.017 0.017 0.018 0.01-0.03
Data for 10 GHz, at 23 °C.
the difference, the smaller the reflection. Table 13-9 lists the dielectric constants of some composite materials (Anderson and Nguyen, 1992; Cordova and Collier, 1992; Harper et al., 1992). Apparently, the advantages of high performance organic polymer fibers used in radome construction lie in their low dielectric constant and low loss tangent. Compared to carbon fiber composites, composites with high performance polymer fibers are better suited for low radar observability. Another interesting feature is that PBZT and PBZO
638
13 High Performance Polymer Fibers
polymers are good candidates for third-order nonlinear optical materials because of their conjugated 7r-structure. They have a very high optical damage threshold, higher than 10 GW/cm2 with picosecond and shorter pulses (Rao et al., 1986; Prasad, 1989; Goldfarb and Medrano, 1989; Goldfarb et al., 1989; Lee et al., 1991). 13.4.5 End Uses of High Performance Polymer Fibers
After about 20 years of production, Kevlar plays a key role in the high performance fiber market. Many applications have been developed, which include composites, anti-ballistics, fabrics, tires, ropes, cables, and pulps. Almost all of them characteristically utilize the great strength, modulus, toughness, thermal stability, and light weight of Kevlar fibers. Kevlar composites can generally retain a large amount of tensile strength under static and cyclic conditions. They also provide significant advantages in composite density, a good balance of lamination properties and impact damage tolerance. Although their compressive strength is relatively poor, Kevlar 49/epoxy composites do not suffer catastrophic failure at high bending, as do graphite and glass composites. The environmental stability of Kevlar composites is generally good except in extremely hostile conditions (Yang, 1989; Fitzgerald and Irwin, 1991). In light of their tensile properties, thermal-oxidative stability, chemical and hydrolytic resistance, dielectric and nonlinear optical behavior, PBZT and PBZO fibers have great potential to be widely used for composite materials and a variety of special uses, such as space and marine applications, ballistic and fire protection, athletic equipment, cables and multilayer circuits and non-linear optical applica-
tions. It was reported that PBZT and PBZO composites compared favorably in tensile behavior to other composites reinforced with Kevlar 49, S-glass as well as aluminum (Yang, 1989). For lower temperature usage, UHMW PE fiber offers good impact, cutting and abrasion resistance. The impact energy of Spectra is of the order of 20 times that of glass, aramid and graphite. Therefore, it is recognized as the most ballistic-resistant fiber and has been widely used in shield materials, cut-resistant clothing and gloves. Since it has light weight, high tensile strength and modulus, low moisture regain and superior chemical resistance, it is suitable for strings, ropes, cables, cordage, sails and composite materials in marine and aerospace applications (Ladizesky and Ward, 1986; Ohta et al., 1988; Allied Signal Fiber Data Sheet, 1990). Vectran and Technora fibers can be used in an intermediate temperature range as composite-enhanced materials, ropes, cables, chemical resistant packing and sporting equipment, because of high strength with very low creep, excellent abrasion, chemical and moisture resistance, and good property retention over a broad range of temperatures (Yang, 1989; Beers and Ramirez, 1990). An interesting application area for rigidrod polymers is their use in "molecular composites". A molecular composite is the molecular-level analog of fiber-reinforced composites. The goal is to form a moderately homogenous mixture of a rigid-rod polymer and a flexible chain matrix polymer. The concept of forming molecular composites originated in the early 1970s with scientific literature appearing in the early 1980s (Takayanagi et al., 1980; Hwang et al., 1983). There are several approaches: (1) Takayanagi finely pulverized and mixed rigid crystalline polymer mi-
13.5 Other High Performance Fibers
crodomains (such as PPTA) and various flexible polymer matrices (such as nitrilebutadiene rubber, nylon 6 and nylon 66) (Takayanagi et al., 1980; Takayanagi, 1984); (2) using guidance from solution thermodynamics [theoretical (Flory, 1978) and experimental (Hwang et al., 1983)], Hwang, Helminiak and coworkers codissolved the rigid-rod (such as PBZT) and various flexible polymer matrices (such as aromatic heterocyclic semiflexible polymers or coil-like nylon polymers) in strong acids (such as methane sulfonic), and then quickly coagulated the isotropic solution to "quench" in the dispersed state (Helminiak etal., 1980; Kozakiewicz, 1986; Hwang etal., 1986); (3) Lenke and Wiff polymerized rigid-rod molecules (such as polyazomethine) in a matrix precursor (such as for nylon 6), then polymerized the matrix precursor to form an "in-situ" molecular composite (USP 5,068,291; Wiff and Lenke, 1992). Fibers and films made from molecular composites display outstanding mechanical and thermal properties, toughness, chemical and environment resistance. In addition, some of the detrimental interface problems with conventional fiber-reinforced composites can be avoided, such as mismatch of CTE between fibers and matrix (Wiff et al., 1988; Adams and Kumar, 1992). Unless care is taken, the rigid-rod molecules are not completely dispersed, but form small domains (Maclachlan et al., 1987). Many new molecular composite systems are under development and research activity is concentrated on improving processability (Kovar et al., 1989; National Materials Advisory Board, 1992; Wiff and Lenke, 1992; Adams and Kumar, 1992).
639
13.5 Other High Performance Fibers As fiber science and technology have progressed, the definition of high performance has been gradually extended. Some specialty polymer fibers which have unique functions to fulfill specific requirements are also considered to be high performance fibers. In addition to their use in construction materials, sports equipment, auto and airplane parts, polymer fibers have found wide use in biomedical, pharmaceutical, food and other physics and chemistry related areas in the last decades. More and more novel fiber products have been developed, and at the same time, more and more research efforts have been made to meet the extreme competition. Since these fields make use of so many different kinds of fiber materials and a great variety of new fiber products continue to emerge, we confine ourselves to only the relatively mature and actively studied areas to give a concise introduction. 13.5.1 Fibers for Medical Uses The biomedical area is one of the newest and most promising fields for further HPF applications. Fiber materials used in these areas can be divided into three main groups: (1) materials in surgery, such as suture threads, artificial skin and tendon, gauze for alloplasty of thoracic and abdominal walls and man-made blood vessels; (2) bio- and chemical membranes with the functions of separation, condensation and purification (for example, artificial kidney and lung); (3) carriers of bioreactors for immobilization of enzymes and microorganisms (see Volume 14).
640
13 High Performance Polymer Fibers
13.5.1.1 Surgical Fibers Two kinds of sutures are currently available, the assimilated (digestible) type and the non-assimilated. An ideal suture should be a monofilament with the necessary mechanical properties and a smooth surface which can pass through the skin easily. It should be biocompatible, nontoxic, and be absorbed in the living body within a predetermined time. Compared to natural sutures, polymer fibers possess greater mechanical strength, uniformity, and resistance to microorganisms. Synthetic polymer fibers also allow the control of physical and chemical properties through fiber composition and structure (Lyman, 1991; Chu, 1991) (see relevant chapters in Volume 14). For a long time, catgut was the only choice for an assimilated suture. The first successful polymer suture designed to meet the above requirements was polyglycolic acid (PGA), Dexon, developed by the American Cyanamid Co. in the 1960s-1970s. Its molecular structure is [-O-CH 2 -CO-] n , and it is synthesized through a polycondensation of hydroxyacetic acid (HO-CH 2 -COOH). PGA fibers are made by a melt-spinning process. They can have high molecular weight (M w : 10 000-100 000) and high crystallinity with a Tm of 224-226 °C and a density of 1.501.64 g/cm3. PGA is insoluble in common solvents. Laboratory and clinical studies have demonstrated that Dexon has compatibility, predictable absorption time, flexibility, ease of handling, knot security and high tensile strength. It is one of the strongest suture materials available [see Table 13-10 (Privalova and Zaikov, 1990)]. It normally takes about 60-90 days to be absorbed by the body, but in some cases, it was not fully absorbed by the body after one to two years (Privalova and Zaikov,
Table 13-10. Knot-pull strength of different sutures (with standard USP 3-0 gaugea) (Privalova and Zaikov, 1990). Suture material Silk Catgut Cotton Dacron (PET) Nylon PE Polypropylene (PP) Dexon (PGA) Stainless steel
Knot-pull tensile strength (kg) 1.54 1.73 1.23 1.63 1.54 1.41 1.50 2.09 2.36
a In the standard of United States Pharmacopoeia (USP), the unit of knot-pull strength is lb.
1990; Hongu and Philips, 1990). The applications of PGA have expanded considerably recently. A very interesting application is bio-active fibers produced by the processes of adding drugs and/or bio-additives into the spinning dope. The bio-active fibers include antimicrobial, anesthetic, anti-inflammatory, homeostatic, anticoagulant, radioactive, enzyme-containing and X-ray contrast fibers. Along with suture threads, PGA has also been used in woven gauze and felt-like sponges for packing the surface of a bleeding organ. There are some shortcomings in using PGA sutures. Since the fibers are a little too stiff to be used as a monofilament suture, they are commonly used as a bundle, which is more prone to infection. In spite of great experience acquired in the last decade, the chemical, physical and medicobiological behavior during treatment in the living body is not yet clearly understood. PGA suture is not recommended in cases where the tissues should be kept stretched. Much research has gone into improving PGA-based sutures, including Maxon, the product of Johnson and Johnson. Most of the efforts are concentrated
13.5 Other High Performance Fibers
on copolymerization with lactones, which have a lower Tm and less crystallinity than the homopolymer. These attributes favor a decrease in the absorption time, an increase in the flexibility, and a lower processing temperature. The latter favors making bio-active fibers because of the low decomposition temperature of most drugs and bio-active additives. Artificial skin is used mostly in the case of burns. Normally, there are three choices for the skin treatment: cultured self-skin from the patient; natural materials; and polymer materials. Cultured human skin is a good choice but since grafting of other people's skin has not been successful in the case of large area burns, artificial skin is necessary. Nonwoven polymer materials made from polyurethanes and Kevlar fibers are used as artificial skins. The former have advantages of compatibility, elasticity and gas-permeability; the latter have good mechanical properties (Takakuda, 1991). One promising material for artificial skins is the nonwoven fiber product made
641
from chitin and chitosan compounds, which belong to the polysaccharides. Chitosan is a linear (1-4) linked 2-amino-2-deoxy-/?-Z)-glucan and chitin is its 7V-acetyl product. As shown in Fig. 13-36, both structures are similar to cellulose. Chitin and chitosan are widely distributed in nature and can be made into fibers with solution-spinning techniques. It was reported that the fiber tensile strength was 1.22.6 g/d when dried, and decreased to 0.151.2 g/d under wet conditions. Its knot strength was 0.1-2 g/d and its breaking elongation, 3-35%. Both chitin and chitosan are almost non-toxic, biodegradable, very compatible to and able to be absorbed by the body. Laboratory safety and clinic tests show that they are suitable as medical materials (Hirano and Tokura, 1982; Hirano e t a l , 1986; Shimahara, 1990; Sakurai, 1990; Iizuka, 1990; Mochizuki and Yamashita, 1990). Up to now, they have been used as powders and tablets for the controlled release of drugs and nutrients, dialysis membranes, digestible sutures and artificial skins for wound healing and
OH (a) 0 —
NHCOCH, (b)
H H
NHCOCH3
0—
H CH2OH
(0
Figure 13-36. Chemical structure of (a) cellulose, (b) chitin, and (c) chitosan.
642
13 High Performance Polymer Fibers
dressing. However, the barrier which prohibits their wide use is the absence of suitable solvents to dissolve them, due to the strong resistance of their micelle structure to chemical agents. In light of their affinity to the growth of new cells during the healing of human surface skin and other excellent bio-behavior, they could become one of the main sources of artificial skin. Furthermore, if the processing and mechanical properties of their fibers were improved, they would be ideal materials for suture and artificial tendons. Artificial tendons require strong fibers with good dynamic mechanical properties. In addition to bio-compatibility, they should have good anti-fatigue and antifriction behavior, good interfaces with bones, excellent adaptability to the geometry of their position and the ability to assist the growth of human organisms on their surface, in vivo. Commonly used polymer fiber materials such as polypropylene (PP), polyethylene terephthalate) (PET), Kevlar, Teflon, PE, fiber-reinforced polyurethane and carbon fiber composites have been used as artificial tendons. The bio-absorbed polymer fibers are also used, such as the PGA and chitin families (Snyder, 1991).
out each year all over the world and the number of the patients using artificial kidneys is increasing (Hongu and Philips, 1990). Some people have survived more than 25 years depending on an artificial kidney. The artificial kidneys are not yet able to perform all the renal functions, but the removal of waste, toxic metabolic products and excessive water from the blood, to avoid organ toxicity, can be carried out. It is well known that hollow fibers are used as purification membranes because of their great specific surface areas and optimum packing density (Lloyd, 1985; Ikada, 1991). Generally, the diameter of the fibers is about 200 \\m. They are arranged parallel to each other or crossed at a small angle. Blood passes through the center of the hollow fibers while the dialysate solution flows around the fibers. By using the pressure difference, excessive water is filtered out and by means of the concentration differences, the small solutes in the blood and in the dialysate solution, such as electrolytes and urea proteins, can permeate
blood in
13.5.1.2 Biological and Chemical Membrane Fibers
The kidney is one of the chief excretory organs of the human body, along with the respiratory system and the skin. As kidney function decreases, various substances which are normally excreted by it will accumulate, thus leading to morbidity and disability. The loss in excretory capacity has to be matched by an artificial kidney in order to keep the overall mass balance in the body. More than 20 million artificial kidney procedures are successfully carried
electrolytes
water, urea etc.
t out
in
dialysate solution
Figure 13-37. Schematic presentation of the function of an artificial kidney.
13.5 Other High Performance Fibers
through the membrane (as shown in Fig. 13-37). Among the hollow fiber membranes, cellulose and its derivatives have about 80% of the market share (Ikada, 1991). Cellulose nitrate (Collodion) was the first membrane used in hemodialysis. Cuprophane hollow fibers, a regenerated cellulose made by the cuoxan process, are the most widely used membranes today. Cellulose acetate hollow fibers, which can be made by melt-spinning or solutionspinning techniques, are also used in great amounts in commercial products. It appears that even though a number of new synthetic polymer membranes have been developed over the last few years, Cuprophane is still considered the hollow fiber of choice for an artificial kidney, due to its good bio-behavior, permeability, mechanical properties and low cost. Research is now mainly focused on socalled selective permeability. The blood contains proteins of molecular weight between 10000 and 30000, which can cause kidney trouble, and also beneficial substances, such as albumin, of molecular weight around 70 000. Therefore a suitable membrane must have the ability to pass proteins of molecular weight around 20000, but to keep those around 70000. Modification of cellulose hollow fibers is one way to make high performance fibers with more and bigger open pores in order to raise the ultrafiltration rates for middlesized molecules. Coating or grafting functional materials on the surface of the cellulose can also improve its properties (Ikada, 1991; Corretge et al., 1988). As an alternative, new synthetic polymer hollow fibers made from polysulfone (PSF), polyacrylonitrile (PAN), polycarbonate and poly(methyl methacrylate) (PMMA) have also been commercialized to provide adequate clearance for molecules of low hydrophilicity. In particular, PSF has been
643
extensively studied for membrane application because of its good chemical resistivity (Pusch and Walsch, 1982). The lung is a form of gas exchanger and supplies oxygen to the blood and removes carbon dioxide. The artificial lung was first developed to take over the lung's functions during heart surgery, and it is now extensively used for this purpose. It is even used as a supplementary respiratory device over a longer term, to assist the breathing of patients (about 250000 U.S.A. patients per year) (Hongu and Philips, 1990). A successful artificial lung must provide efficient gas-blood contact in order to oxygenate up to five liters of blood per minute to nearly 100% saturation. Simultaneously, it should remove a proper amount of CO 2 without causing respiratory acidosis (CO2 retention) or respiratory alkalosis (CO2 depletion). In addition, it should eliminate any harmful side-effects. The breakthrough in this field was the development of heterogeneous hydrophobic membranes, such as microporous PP hollow fibers and Teflon membranes. The gases can freely pass through the hydrophobic pores in the membranes, but blood cannot because blood cells and most components of blood are hydrophilic. The PP microporous membranes can offer gas transfer rates for both O 2 and CO 2 of about four times those of Cellophane. However, it was found that the microporous membranes damage blood plasma components and suffer a progressive reduction in the gas transport, because the gas channels in the membrane flood with fluid during long term use. Most artificial lung membranes only last about a week, due to a drop in their ability to remove CO 2 . In order to overcome these drawbacks, several interesting approaches have been proposed, such as treating the fiber surface with bioactive agents, e.g. albumin.
644
13 High Performance Polymer Fibers
Hollow fibers or woven tubes made from PTFE and polyester are two wellknown artificial blood vessels. The former exhibits both good biocompatibility and anticoagulant activity, but cannot be used for small diameter vessels. The latter is biocompatible and has good mechanical properties, but poor anticoagulant activity. A very fine polyester blood vessel has been successfully developed recently (Hongu and Philips, 1990; Ikada, 1991). 13.5.1.3 Fiber Carriers for Biologically Active Agents
A relatively new and developing field in biotechnology is to utilize enzymes or micro-organisms which are capable of producing a great variety of antibiotics, vitamins, amino acids and hormones for different applications. Over 2000 enzymes have been found suitable for use in a bioreactor, but only a few are now commercialized. It has been realized that there are some drawbacks when the biologically active entities, such as enzymes, are employed in free form. They are not sufficiently stable, especially in extremes of both temperature and pH. They are watersoluble and hence difficult to separate from substrates and products, which prevents them from being used repeatedly. In clinical applications, the free form would usually cause immunological reactions, rapid removal and inactivation. Immobilization is a way to solve the above problems. Fibers, especially porous, non-circular cross-section and hollow fibers, have already been employed in bioreactors for bio technological applications. They are generally used in two immobilization techniques: the enzymes are covalently bound to the fiber carriers; or the enzymes are adsorbed ionically or physically to the fiber support materials. The polymer fibers
can be used, not only as a carrier, but also as a barrier to separate one bulk solution from another. Bioactive agents may be attached to the surface of the fibers or entrapped within the fiber network. The substrate molecules can freely permeate through the membrane and react with enzymes or micro-organisms. The product molecules are transported freely in opposite directions, but the membrane is substantially impermeable in either direction to the much higher Mw species. For example, penicillinacylase can be linked onto a reduced porosity PAN fiber to produce 6amino penicillin, which serves as a raw material for reactive penicillin. Several thousand tons are manufactured annually. In another commercial application acetobacter are adsorbed onto cotton-like PP fibers. The bioreactor is used in acetic acid production, and the carrier fibers can be regenerated by replacing the inactive enzyme. The fibers can also be made in the form of filter paper or non-woven fabrics and used as biosensors. 13.5.2 Polymer Fiber Precursors to Superconductors
Since the development of new oxides which exhibit superconductive behavior at temperatures above that of liquid nitrogen, a new horizon has been opened in materials physics. Making superconductive fibers is of great interest to meet the many potential new applications requirements (Goto, 1987a, b, 1988a, b, c). Research into making superconducting fibers is now world-wide. There are several spinning processes for making superconducting fibers, such as suspension-solution spinning, sol-gel spinning and melt spinning. Among them, the suspension-solution spinning methods are associated with polymer fiber technology, because of its
13.5 Other High Performance Fibers
maturity and the great advantages in process versatility. 13.5.2.1 Fibers from Mixtures of Superconducting Oxides and PVA
The primary process at present is a suspension-solution spinning method (Goto, 1987a, 1987b, 1988a, 1988b, 1988c). Powders of the oxides and dispersers are put into a PVA solution to make a spinning dope, which is extruded through spinnerets into a coagulation bath. The asspun fibers are heat-treated to produce superconducting fibers. The spinning dopes can be of two types, aqueous and nonaqueous solutions. For the former, the powders of Y 2 O 3 , BaCO 3 , and CuO are used as precursors to make superconducting Y 1 Ba 2 Cu 3 O Jc . Then the powders of the superconductor and the dispersers (nonionic type + anionic type) are placed in a water solution, and added to the PVA solution to form the spinning dope. The spinning solution is extruded into the coagulation bath (Na 2 SO 4 + NaOH) to produce fibers. After washing and drying, the asspun fibers are heated in O 2 atmosphere at 980 °C for five minutes, then gradually cooled to room temperature at 100°C/h. In this process, the PVA decomposes and the superconducting fibers are formed. For the latter type (non-aqueous solution), the solvent for PVA is dimethyl sulfoxide (DMSO), the coagulant is methanol and the superconductor is Ln 1 Ba 2 Cu 3 O :c (where Ln = Er, Ho, Dy). The degree of polymerization (DP) of PVA has a great effect on the fiber spinning and the fiber properties. If the DP < 2450 or > 16 000, spinning cannot be carried out. If the DP is kept in the range of 3300 to 12100, the superconducting properties are improved as the DP is increased. Unfortunately, the superconducting properties of these fibers have not yet met practical requirements.
645
13.5.2.2 Conducting Polymer Fibers
As early as the 1930s, theoretical studies suggested that it was possible to have conductive polymeric materials (Jones, 1937). In the last two decades, research on conducting polymers has become gradually more and more active. Now several polymers are of great interest, including polyacetylene, polythiophene, polypyrrole, and poly(/?-phenylene vinylene). The common characteristic of all these conducting polymers is conjugation along the backbone of the polymer chain, which appears to be a prerequisite for a conducting polymer. However, due to this high conjugation, most of the conducting polymers are very difficult to process. Several methods have been proposed for making conducting polymer fibers. The so-called "Durham" route is a three-step method (Feast and Winter, 1985). The first step is to use a soluble polymer precursor to produce fibers, then the second step is to convert the fiber into the conjugated polymer fiber, and the third step is to dope for conductivity. For example, 3,6-bis(trifluoromethyl) pentacyclo-[6.2.0.0]dec-9-ene was used as a monomer to synthesize an intermediate polymer that is relatively stable at room temperature and is conveniently processed. Then the intermediate polymer is converted to polyacetylene by heating it to 75 °C to eliminate hexafluoro-6>-xylene. Poly(2,5-thienylene vinylene) and poly(2,5-furylene vinylene) can also be prepared by this route (Jen, 1987; Yamada, 1987). The main problem with this method is that all the precursor routes lead to a polymer which cannot be reprocessed, and this technique is, consequently, of limited use. Sato used 3-alkyl thiophene which can be dissolved in common organic solvents (such as CHC13 and tetrahydrofuran,
646
13 High Performance Polymer Fibers
THF) for polymerization. The polymer can be spun into fibers by means of the melt-spinning technique (Sato, 1987). Polyheterocyclics synthesized within the mesoscopic pores of a host membrane is another way of making conductive polymer fibers. The membrane contains linear and cylindrical pores and separates a solution of the monomer from a solution of a chemical oxidizing agent. The monomer and oxidizing agent diffuse toward each other through the pores to accomplish the synthesis. Polypyrrole and poly(3-methylthiophene) fibers were made according to this method (Cai and Martin, 1989). PBZT and PPTA have been spun from dopes containing metallophthalocyanine for making conducting fibers. The mechanical properties are adversely affected as the concentration of metallophthalocyanine increases, but process optimizations have not yet been carried out. Some ladder polymers are also possible candidates as conducting polymer fibers produced by solution spinning (Little, 1964; Blythe, 1979; Wynne et al., 1985; Polis et al., 1989). All these activities are pushing research in this field to higher levels although it appears now that there is still a long way to go to reach the target of conductive polymer fibers for practical applications. Conducting polymeric materials are very attractive high performance materials for science and technology development in the 21st century and we can reasonably predict breakthroughs in the next decade.
13.6 The Future This cursory glance at a few representative high performance polymer fibers is not intended to exhaustively survey this field, or even to delineate all the areas of current
interest in these materials. Instead a few areas of real or potential new levels of performance have been identified in order to interest the materials scientist in the vast scope of performance of polymers. A tenfold growth in consumption of high performance fibers is predicted by the year 2000, and the real potential for replacement of higher density materials has probably not yet been fully identified. Certainly the aerospace industry has driven the search for new, lighter, stronger, stiffer, durable, tough, processable, and even radar-invisible fibers to be used in reinforced composites for airplanes and spacecraft. Achievements such as the non-stop, round-the-world, unrefueled flight of the Voyager and massive research projects such as the "Orient Express", the Mach 15 National Aerospace Plane to travel from New York to Tokyo in two hours, are glowing testament to progress in composites. There remain tremendous scientific problems to be solved in the understanding of fibers. It is only by understanding the mechanisms of strength, stiffness, compressive failure, and transport in these high performance materials that we can design new molecules or new processes to overcome current limitations. The rewards for success in these endeavors should be considerable, both scientific and technological. Projections of world demand for advanced composites indicate that by the early decades of the 21st century industrial and other applications will have grown to 55% of the market share while the aircraft/aerospace market share will drop to 45 % (National Materials Advisory Board, 1992). Therefore, when the fiber quality meets the requirements, reducing manufacturing and processing costs and even making new low-cost high performance
13.7 References
polymer fibers are critical. It is forecast that if the cost of these fibers could be reduced to a few dollars per pound, the demand would be a factor of 10 higher (National Materials Advisory Board, 1992). In addition, the tailorability of polymers, with resultant enormous variations in properties make them very attractive for research programs aimed at specific achievement of property targets. The specialty polymers used in the high efficiency separation and conductivity fields are new materials in the expanding catalog of high performance materials. With the ongoing rapid progress in the science and technology of materials and processing, more new polymer fibers will emerge to advance the available technology of high performance fibers. "A civilization is both advanced and limited by the materials at its disposal" Sir George Padget Thompson.
13.7 References Adams, W W, Eby, R. K. (1987), MRS Bull. Nov./ Dec, p. 22. Adams, W. W, Kumar, S. (1992), in: Ultrastructure Processing of Advanced Materials'. Uhlmann, D. R., Ulrich, D. R. (Eds.). New York: Wiley, p. 343. Adams, W W, Azaroff, L. V., Kulshreshtha, A. K. (1979), Z. Kristallogr. 150, 321. Adams, W. W, Briber, R. M., Sherman, E. S., Porter, R. S., Thomas, E. L. (1985), Polymer 26, 17. Adams, W. W., Grieshop, T., Helminiak, T., Hunsaker, M., O'Brien, J. R, Altieri, M., Bai, S. I , Brandt, M., Fratini, A. V., Hwang, W.-R, Price, G. E., Soloski, E., Haddock, T, Krause, S. I , Lenhert, P. G. (1986), AFWAL-TR-86-4011. Agosti, E., Zerbi, G., Ward, I. M. (1992), Polymer 33, 5219. Aharoni, S. M. (1985), Macromolecules 18, 2624. Allen, S. R. (1987), /. Mater. Sci 22, 853. Allen, S. R., Farris, R. I (1989), in: The Materials Science and Engineering of Rigid-Rod Polymers. Adams, W W, Eby, R. K., McLemore, D. E. (Eds.). Pittsburgh, PA: MRS Symp. Proc. Vol. 134, p. 297.
647
Allen, S. R., Filippov, A. G., Farris, R. J., Thomas, E. L., Chenevey, E. C. (1981a), Macromolecules 14, 1138. Allen, S. R., Filippov, A. G., Farris, R. I, Thomas, E. L. (1981 b), J. Appl. Polym. Sci. 26, 291. Allen, S. R., Farris, R. I, Thomas, E. L. (1985), /. Mater Sci. 20, 2727. Allied Signal (1990), Fiber Data Sheet. Anderson, B. E., Nguyen, H. X. (1992), in: 37th Int. SAMPE Symp. andExhib.: Grimes, G. C , Turpin, R., Forsberg, G., Rasmussan, B. M., Whitney, J. (Eds.). Anaheim, CA: Vol. 37, p. 850. Argon, A. S. (1972), in: Treatise on Materials Science and Technology. New York: Academic Press, p. 79. Arnold, R E. (1989), in: The Materials Science and Engineering of Rigid-Rod Polymers: Adams, W. W, Eby, R. K., McLemore, D. E. (Eds.). Pittsburgh, PA: MRS Symp. Proc. Vol. 134, p. 75. Arpin, M. (1976), Macromolecules 9, 585. Arpin, M., Strazielle, C. (1977), Polymer 18, 591. Attenburrow, G. E., Bassett, D. C. (1979), J. Mater. Sci. 14, 2679. Baird, D. G. (1978), in: Liquid Crystalline Order in Polymers. Blunstein, A. (Ed.). New York: Academic, Chap. 7. Barham, P. I , Keller, A. (1985), /. Mater. Sci. 20, 2281. Beers, D. E., Ramirez, J. E. (1990), / Text. Inst. 81, 561. Berry, G. C. (1989), in: The Materials Science and Engineering of Rigid-Rod Polymers: Adams, W. W., Eby, R. K., McLemore, D. E. (Eds.). Pittsburgh, PA: MRS Symp. Proc. Vol. 134, p. 181. Bhattacharya, S., Chuah, H. H., Dotrong, M., Wei, K. H., Wang, C. S., Vezie, D., Day, A., Adams, W. W. (1989), Proc. Polym. Mater. Sci. Eng. (ACS) 60, 512. Blumstein, A. (1984), in: Polymeric Liquid Crystals: Blunstein, A. (Ed.). New York: Plenum, p. 154. Blythe, A. R. (1979), Electrical Properties of Polymers. New York: Cambridge Univ. Press. Brelsford, G. L., Krigbaum, W. R. (1991), in: Liquid Crystallinity in Polymers: Ciferri, A. (Ed.). New York: VCH, p. 61. Bunn, C. W. (1939), Trans. Faraday Soc. 35, 482. Busing, W. R. (1990), Macromolecules 23, 4608. C &E News (1987), 65 (29), p. 34. Cai, Z., Martin, C. R. (1989), J. Am. Chem. Soc. Ill, 4138. Capaccio, G., Ward, I. M. (1974), Polymer 15, 233. Chang, H. W, Weedon, G. C. (1986), Polym. News 12, 102. Chenevey, E. C , Timmons, W. D. (1989), in: The Materials Science and Engineering of Rigid-Rod Polymers: Adams, W. W, Eby, R. K., McLemore, D. E. (Eds.)- Pittsburgh, PA: MRS Symp. Proc. Vol. 134, p. 245. Chu, C. C (1991), in: High-Tech Fibrous Materials, ACS Symp. Series 457: Vigo, T. L., Turbak, A. F (Eds.). Washington, D.C.: ACS, p. 167.
648
13 High Performance Polymer Fibers
Chuah, H. H., Tsai, T. T., Wei, K. H., Wang, C. S., Arnold, F. E. (1989), Polym. Mater. Sci. Eng. 60, 517. Ciferri, A. (1975), Polym. Eng. Sci. 15, 191. Ciferri, A., Ward, I. M. (1979), Ultrahigh Modulus Polymers. London: Applied Science. Collings, P. J. (1990), Liquid Crystals. NJ: Princeton Univ. Press. Conibeer, C. X, Lin, K. Z., Edward, G. H. (1990), Mater. Forum 14, 224. Cordova, D. S., Collier, L. S. (1992), in: 37th Int. SAMPESymp. and Exhib.: Grimes, G. C , Turpin, R., Forsberg, G., Rasmussan, B. M., Whitney, J. (Eds.). Anaheim, CA: Vol. 37, p. 1406. Corretge, E., Kishida, A., Konish, H., Ikada, Y. (1988), in: Polymers in Medicine III: Migliaresi, C. et al. (Eds.). Amsterdam: Elsevier, p. 61. Crosby, C. R., Ford, N. C , Karasz, F. E., Langley, K. H. (1981), J. Chem. Phys. 75(9), 4298. Dang, T. D., Tan, L. S., Wei, K. H., Chuah, H. H., Arnold, F. E. (1989), Polym. Mater. Sci. Eng. 60, 424. Delmonte, J. (1985), Origins of Materials and Processes. Lancaster, PA: Technomic Publishing Co., Inc., pp. 295. DeTeresa, S. X, Farris, R. X, Porter, R. S. (1982), Polym. Compos. 3, 57. DeTeresa, S. X, Allen, S. R., Farris, R. X, Porter, R. S. (1984), J. Mater. Sci. 19, 57. DeTeresa, S. X, Porter, R. S., Farris, R. X (1985), /. Mater. Sci. 20, 1645. DeTeresa, S. X, Porter, R. S., Farris, R. X (1988), /. Mater. Sci. 23, 1886. Dobb, M. G. (1979), Polymer 20, 1284. Dobb, M. G. (1985), in: Strong Fibers: Watt, W, Perov, B. V. (Eds.). North-Holland: Elsevier Science Publishers, Part 3, Chapter XVII, p. 673. Dobb, M. G., Hindeleh, A. M., Johnson, D. X, Saville, B. P. (1975), Nature 253, 189. Dobb, M. G., Johnson, D. X, Saville, B. P. (1977a), /. Polym. Sci., Polym. Symp. 58, 237. Dobb, M. G., Johnson, D. X, Saville, B. P. (1977 b), J. Polym. Sci., Polym. Phys. Ed. 15, 2201. Doi, M. (1981), J. Polym. ScL, Polym. Phys. Ed. 19, 229. Donald, A. M., Windle, A. H. (1992), Liquid Crystalline Polymers. Cambridge: University Press. Drzal, L. T, AFWAL-TR-86-4003. DuPont (1988), Product Literature. Erman, B., Flory, P. X, Hammel, X P. (1980), Macromolecules 13, 484. Farmer, B. L., Chapman, B. R. (1993), Polymer, in press. Fawaz, S. A., Palazotto, A. N., Wang, C. S. (1992), Polymer 33, 100. Feast, W. X, Winter, X (1985), J. Chem. Soc, Chem. Commun. 202. Feldman, L., Zihlif, A., Farris, R. X, Thomas, E. L. (1987), J Mater. Sci. 22, 1199.
Fitzgerald, X A., Irwin, R. S. (1991), in: High Value Polymers: Fawcett, A. H. (Ed.). London: Royal Society of Chemistry; p. 392. Flory, P. X (1956), Proc. R. Soc. A234, 73. Flory, P. X (1978), Macromolecules 11, 1138. Fratini, A. V, Lenhert, P. G., Resch, T. X, Adams, W. W. (1989), in: The Materials Science and Engineering of Rigid-Rod Polymers: Adams, W W, Eby, R. K., McLemore, D. E. (Eds.). Pittsburgh, PA: MRS Symp. Proc. Vol. 134, p. 431. Galiotis, C , Young, R. X (1983), Polymer 24, 1023. Goldfarb, I. X, Medrano, X (1989), in: Nonlinear Optical Effects in Organic Polymers: 2: London: Kluwer, p. 93. Goldfarb, I. X, Reale, H., Wierschke, S., Medrano, X (1989), in: The Materials Science and Engineering of Rigid-Rod Polymers: Adams, W. W, Eby, R. K., McLemore, D. E. (Eds.). Pittsburgh, PA: MRS Symp. Proc. Vol. 134, p. 609. Goto, T. (1987 a), Jpn. J. Appl. Phys. 26, LI 527. Goto, T. (1987 b), Jpn. J. Appl. Phys. 26, L1970. Goto, T. (1988 a), Jpn. J. Appl. Phys. 27, L680. Goto, T. (1988 b), J. Mater. Sci. Lett 57, 283. Goto, T. (1988 c), Sen-i Gakkaishi 44(7), P-249 (in Japanese). Greenwood, X H., Rose, P. G. (1974), J. Mater. Sci. 9, 1809. Grubb, D. T, Prasad, K. (1992 a), ACS Polym. Symp. 33, 302. Grubb, D. T, Prasad, K. (1992b), Macromolecules 25, 4575. Gutierrez, G. A., Blackwell, X (1984), Macromolecules 17, 2744. Haraguchi, K., Kajiyama, T, Takayanagi, M. (1979 a), J. Appl. Polym. Sci 23, 903. Haraguchi, K., Kajiyama, T, Takayanagi, M. (1979b), J. Appl. Polym. Sci. 23, 915. Harper, X D., Downs, X W, Glennon, A. E. (1992), in: 37th Int. SAMPESymp. and Exhib.: Grimes, G. C , Turpin, R., Forsberg, G., Rasmussan, B. M., Whitney, X (Eds.). Anaheim, CA: Vol. 37, p. 760. Helminiak, T. E., Benner, C. L., Arnold, F. E., Husman, G. E. (1980), U.S.P. 4207407. Hermans, X X, Hermans, P. H., Vermaas, D., Weidinger, A. (1946), Reel. Trav. Chim. Pays-Bas 65, All. Hermans, X (1962), /. Colloid Sci 17, 638. Hirano, S., Noishiki, Y, Kinugawa, X (1986), in: Advances in Biomedical Polymers: Gebelein, C. G. (Ed.). New York: Plenum Press, p. 285. Hirano, S., Tokura, S., Proc. 2nd Int. Conf. on Chitin and Chitosan, July 12-14, Japan. Hofmann, D., Schulz, E. (1989), Polymer 30, 1964. Hongu, T, Philips, G. (1990), New Fibers. New York: Ellis Horwood. Hoogsteen, W, van der Hooft, R. X, Postema, A. R., ten Brinke, G., Pennings, A. X (1988a) J. Mater. Sci. 23, 343.
13.7 References
Hoogsteen, W., Kormelink, H., Eshuis, G., ten Brinke, G., Pennings, A. J., (1988 b), /. Mater. Sci. 23, 3467. Horio, M. (1984), Sen-i Gakkaishi 40, T-285. Huh, W., Kumar, S., Helminiak, T. E., Adams, W. W. (1990), in: SPE Annual Tech. Conf. Proc, 1245. Hwang, W.-R, Wiff, D. R., Verschoore, C , Price, G., Helminiak, T. E., Adams, W. W. (1983), Polym. Eng. Sci. 23, 784. Hwang, W.-R, Wiff, D. R., Helminiak, T. E. (1986), U.S.P. 4631318. Iguchi, M., Kyotani, H. (1990), Sen-i Gakkaishi 46, P-471 (in Japanese). Iizuka, E. (1990), Sen-i Gakkaishi 46(11), P-528 (in Japanese). Ikada, Y. (1991), Sen-i Gakkaishi 47(3), P-120 (in Japanese). Im, I, Percha, P. A., Yeakle, D. S. (1989), in: The Materials Science and Engineering of Rigid-Rod Polymers: Adams, W. W, Eby, R. K., McLemore, D. E. (Eds.), Pittsburgh, PA: MRS Symp. Proc. Vol. 134, p. 307. Ishida, S. (1987), Polym. Prepr. Jpn. 36, 1115. Jackson, W. J. (1980), Br. Polym. J. 12, 154. Jen, K. (1987), J. Chem. Soc. Chem. Commun. 309, 1113. Jiang, H. (1989), Ph.D. thesis, The Johns Hopkins University; Baltimore, MD. Jiang, H., Eby, R. K., Adams, W. W, Lenhert, G. (1989), in: The Materials Science and Engineering of Rigid-Rod Polymers'. Adams, W W, Eby, R. K., McLemore, D. E. (Eds.), Pittsburgh, PA: MRS Symp. Proc. Vol. 134, p. 341. Jiang, H., Eby, R. K., Adams, W. W. (1990), Polym. News 15, 11. Jiang, H., Desai, P., Kumar, S., Abhiraman, A. S. (1991 a), Carbon 29, 635. Jiang, H., Damodaran, S., Desai, P., Kumar, S., Abhiraman, A. S. (1991b). Proc. Polym. Mater. Sci. Eng. 64, 383. Jones, J. L. (1937), Proc. R. Soc. A158, 280. Kamide, K. (1983), Makromol. Chem., Rapid Commun. 4, 33. Kanamoto, T., Ohama, T., Tanaka, K., Takeda, M., Porter, R. S. (1987), Polymer 28, 1517. Kanamoto, T., Kiyooka, S., Tovmasyan, Y, Sano, H., Narukawa, H. (1990), Polymer 31(11), 2039. Kawai, T. (1980), Makromol. Chem., Rapid Commun. 1, 187. Kelly, A., MacMillan, N. H. (1986), in: Strong Solids, Oxford: Oxford Science Publications, p. 21. Kikuchi, T. (1982), Surface 20, 270 (in Japanese). Klunzinger, P. E., Green, K. A., Eby, R. K., Parmer, B. L., Adams, W. W, Czornyj, G. (1991 a), in: Tech. Paper, XXXVII, Soc. Plastics Eng., p. 1532. Klunzinger, P. E., Macturk, K. S., Eby, R. K., Adams, W. W. (1991b), Polym. Prepr. (ACS) 32(2), 187.
649
Klunzinger, P. E., Eby, R. K., Adams, W. W. (1992a), MRS Symp. Proc. Vol. 252, p. 119. Klunzinger, P. E., Ramirez, R. G., Thomas, D. A., Eby, R. K. (1992 b), APS Bull. 37(1), 508. Klunzinger, P. E., Eby, R. K. (1993), Polymer, in press. Knoff, W. P. (1987), J. Mater. Sci. 6, 1392. Koide, N. (1990), Sen-i Gakkaishi 46, 496. Kovar, R. R, Haghighat, R. R., Lusignea, R. W. (1989), in: Mater. Sci. Eng. Rigid-Rod Polym.: Adams, W. W, Eby, R. K., McLemore, D. E. (Eds.). MRS Symp. Proc. Vol. 134, p. 389. Kozakiewicz, J. J. (1986), U.S.P. 4614784. Krause, S. X, Haddock, T. B., Vezie, D. L., Lenhert, P. G., Hwang, W.-R, Price, G. E., Helminiak, T. E., O'Brien, J. R, Adams, W W. (1988), Polymer 29, 1354. Krause, S. J., Vezie, D. L., Adams, W. W. (1989), Polym. Commun. 30, 10. Kumar, S., Adams, W. W, Helminiak, T. E. (1988), /. Reinf Plast. Compos. 7, 108. Kumar, S., Helminiak, T. E. (1989 a), in: The Materials Science and Engineering of Rigid-Rod Polymers: Adams, W. W, Eby, R. K., McLemore, D. E. (Eds.). Pittsburg, PA: MRS Symp. Proc. Vol. 134, p. 580. Kumar, S. (1989), SAMPE Q. 20, 3. Kumar, S., Helminiak, T. E. (1989), in: The Materials Science and Engineering of Rigid-Rod Polymers: Adams, W. W, Eby, R. K., McLemore, D. E. (Eds.). Pittsburgh, PA: MRS Symp. Proc. Vol. 134, p. 363. Kumar, S. (1990 a), in: International Encyclopedia of Composites 4: Lee, S. (Ed.). Weinheim: VCH, p. 51. Kumar, S., Adams, W. W (1990), Polymer 31, 15. Kumar, S. (1990 b), 35th SAMPE Int. Symp., p. 2224. Kumar, S. (1991), Indian J. Fiber Test Res. 16, 52. Kumar, S. (1992), private communication. Kunugi, T., Suzuki, A., Akiyama, I., Hashimoto, M. (1979), Polym. Prepr. (Am. Chem. Soc, Div. Polym. Chem.) 20, 778. Kwolek, S. L., Morgan, P. W, Schaefgan, J. R., Gulrich, J. R. (1977), Macromolecules 10, 1390. Kyotani, M. (1990), Sen-i Gakkaishi 46, P-519 (in Japanese). Ladizesky, N. H., Ward, I. M. (1986), Comput. Sci. Technol. 26, 129. Ledbetter, H. D., Rosenberg, S., Hurtig, C. W. (1989), in: The Materials Science and Engineering of RigidRod Polymers: Adams, W W, Eby, R. K., McLemore, D. E. (Eds.). Pittsburgh, PA: MRS Symp. Proc. Vol. 134, p. 253. Lee, C. Y-C, Swiatkiewicz, J., Prasad, P. N., Mehta, R., Bai, S. J. (1991), Polymer 32(7), 1195. Lemstra, P. X, Bastiaansen, C. W. M., Meijer, H. E. H. (1986), Angew. Makromol. Chem. 145/146, 343. Li, L.-S., Allard, L. R, Bigelow, W. C. (1983), /. Macromol. Sci. Phys. B22, 269.
650
13 High Performance Polymer Fibers
Little, W. A. (1964), Phys. Rev. 134. Lloyd, D. R. (1985), in: Materials Science of Synthetic Membranes, ACS Symp. Ser., p. 1. Lusignea, R. W. (1989), in: The Materials Science and Engineering of Rigid-Rod Polymers'. Adams, W. W., Eby, R. K., McLemore, D. E. (Eds.). Pittsburgh, PA: MRS Symp. Proc. Vol. 134, p. 265. Lyman, D. J. (1991), in: High-Tech Fibrous Materials, ACS Symp. Ser. 457: Vigo, T. L., Turbak, A. F. (Eds.). Washington D.C.: ACS, pp. 116. Lysenko, Z. (1988), U.S.P. 4766244. MacDonald, W. A. (1991), in: High Value Polymers: Fawcett, A. H. (Ed.). London: Royal Society of Chemistry, p. 428. Maclachlan, J., Madey, M., Eby, R. K., Adams, W. W. (1987), Polym. Commun. 28, 326. Macturk, K. S., Eby, R. K., Adams, W. (1991), Polymer 32, 1782. Macturk, K. S., Eby, R. K., Farmer, B. L. (1993), Polymer, in press. Manabe, S. (1980), Sen-i Gakkaishi 33, T-93. Martin, D. C , Thomas, E. L. (1989), in: The Materials Science and Engineering of Rigid-Rod Polymers: Adams, W. W, Eby, R. K., McLemore, D. E. (Eds.). Pittsburgh, PA: MRS Symp. Proc. Vol. 134, p. 415. Martin, D. C , Thomas, E. L. (1991a), Macromolecules 24, 2224. Martin, D. C , Thomas, E. L. (1991 b), J. Mater. Sci. 26, 5171. Matsuda, K. (1985 a), Sen-i Gakkaishi 38, P-9 (in Japanese). Matsuda, K. (1985 b), Sen-i Kikai Gakkaishi 38, P-225 (in Japanese). McGarry, F J., Moalli, J. E. (1991a), Polymer. 32, 1811. McGarry, F. J., Moalli, J. E. (1991b), Polymer. 32, 1816. Minter, J. R., Shimamura, K., Thomas, E. L. (1981), J. Mater. Sci. 16, 3303. Minter, J. R. (1982), AFWAL-TR-82-4097. Miwa, M., Tsushima, E., Takayasu, J. (1991a), /. Appl. Polym. Sci. 43, 1467. Miwa, M., Tsushima, E., Takayasu, J. (1991b), Sen-i Gakkaishi 47, P-171 (in Japanese). Mochizuki, A., Yamashita, S. (1990), Sen-i Gakkaishi 46(12), P-570 (in Japanese). Morgan, R. X, Pruneda, C. O., Steele, W. J. (1983), /. Polym. Sci. Phys. Ed. 21, 1757. Murthy, N. S. (1990), Polym. Commun. 31, 30. Nakagawa, K., Maeda, O., Yamakawa, S. (1983), /. Polym. Sci. Polym. Lett. Ed. 21, 933. National Materials Advisory Board (1992), High Performance Synthetic Fibers for Composites. Washington, D.C.: National Academy Press. Northolt, M. G., Ven Aarsten, J. J. (1973), J. Polym. Sci. Polym. Lett. 11, 333. Northolt, M. G., Ven Aarsten, J. J. (1977), J. Polym. Sci. Polym. Symp. 58, 283.
Northolt, M. G. (1981), Br. Polym. J. 13, 64. Northolt, M. G., Sikkema, D. J. (1991), in: Advances in Polymer Science 98. Berlin: Springer-Verlag, p. 115. Odell, J. A., Keller, A., Atkins, E. D. T, Nagy, M. J., Feijoo, J. L., Ungar, G. (1989), in: The Materials Science and Engineering of Rigid-Rod Polymers: Adams, W. W, Eby, R. K., McLemore, D. E. (Eds.). Pittsburgh, PA: MRS Symp. Proc. Vol. 134, p. 223. Ohsawa, T, Miwa, M., Kawade, M. (1990), J. Appl. Polym. Sci. 39, 1733. Ohta, T, Kunugi, T, Yobuki, K. (1988), High Tenacity and High Modulus Fibers: Society of Polymer Science, Japan (Ed.), Kubunshi Shinsozai One Point-9. Tokyo: Kyoritsu Shuppan Co., Ltd. Panar, M., Avakian, P., Blume, R. C , Gardner, K. H., Gierke, T. D., Yang, H. H. (1983), J. Polym. Sci., Polym. Phys. Ed. 21, 1955. Peguy, A., Manley, R. S. J. (1984), Polym. Commun. 25, 39. Penning, J. P., Dijkstra, D. I, Pennings, A. J. (1991), J. Mater. Sci. 26,4121. Pennings, A. J., Schouteten, C. J. H., Kiel, A. M. (1972), J. Polym. Sci. C38, 167. Pennings, A. J. (1976), Colloid Polym. Sci. 253, 452. Pennings, A. J. (1977), /. Polym. Sci., Polym. Symp. 59, 55. Polis, D. W, Dalton, L. R., Vachon, D. J. (1989), in: The Materials Science and Engineering of RigidRod Polymers: Adams, W. W, Eby, R. K., McLemore, D. E. (Eds.). Pittsburgh, PA: MRS Symp. Proc. Vol. 134, p. 679. Porter, R. S. (1992), private communication. Pottick, L. A., Farris, R. J., Thomas, E. L. (1985), Polym. Eng. Sci. 25, 284. Prasad, K., Grubb, D. T. (1990), /. Polym. Sci., Pt. B, Polym. Phys. 28, 2199. Prasad, P. N. (1989), in: The Materials Science and Engineering of Rigid-Rod Polymers: Adams, W. W, Eby, R. K., McLemore, D. E. (Eds.). Pittsburgh, PA: MRS Symp. Proc. Vol. 134, p. 635. Prevorsek, D. C , Chin, H. B., Kwon, Y D., Kavesh, S. (1992), Polym. Prepr. 33, 368. Privalova, L. G., Zaikov, G. E. (1990), Polym.-Plast. Technol. Eng. 29(5&6), 455. Pusch, W, Walsch, A. (1982), Angew. Chem. Int. Ed. Engl. 231, 660. Rakas, M. A., Farris, R. J. (1988), Polym. Mater. Sci. Eng. 58, 1024. Rakas, M. A., Farris, R. J. (1989), in: The Materials Science and Engineering of Rigid-Rod Polymers: Adams, W W, Eby, R. K., McLemore, D. E. (Eds.). Pittsburgh, PA: MRS Symp. Proc. Vol. 134, p. 277. Rao, D. N., Swiatkiewicz, I, Chopra, P., Ghoshal, S. K., Prasad, P. N. (1986), Appl. Phys. Lett. 48(18), 1187. Reisch, M. S. (1987), C & E News 65(5), p. 9.
13.7 References
Roche, R. I , Takahashi, T., Thomas, E. L. (1980), in: ACS Symp. Ser. 141, 303. Sakurada, I. (1985), Polyvinyl Alcohol Fibers. New York: Marcel Dekker. Sakurai, K. (1990), Sen-i Gakkaishi 46(12), P-553 (in Japanese). Sato, M. (1987), Synth. Met. 18, 229. Sawyer, L. C , Jaffe, M. (1992), Mater. Res. Soc. Symp. Proc. 255, 75. Sawyer, L. C , Chen, R. T., Jamieson, M. G., Musselman, I. H., Russell, P. E. (1992), /. Mater. Sci. Lett. 11, 69. Sawyer, L. C , Chen, R. T., Jamieson, M. G., Musselman, I. H., Russell, P. E. (1993), /. Mat. Sci. 28, 225. Schaefgen, J. R., Foldi, V. S., Logillo, F. M., Good, V. H., Gulrich, L. W, Killian, F. L. (1976), Polym. Prepr. (Am. Chem. Soc, Div. Polym. Chem.) 17, 69. Schaper, A. (1989), Phys. Stat. Sol. (a)116, 179. Science and Technology Center of Osaka (1987), in: Report on High Strength Polymers', p. 108 (in Japanese). Sheiko, S. S., Magonov, S. N., Moller, M. (1992), Polym. Prepr. (Am. Chem. Soc, Div. Polym. Chem.) 33, 788. Shen, D. Y, Molis, S. E., Hsu, S. L. (1983), Polym. Eng. Sci. 23(10), 543. Shibayama, M., Kurokawa, H., Nomura, S., Roy, S., Stein, R. S., Wu, W.-L. (1990), Macromolecules 23(5), 1438. Shigematsu, K., Imada, K., Takayanagi, M. (1985), /. Polym. Sci., Polym. Phys. Ed. 13, 73. Shimada, K., Mera, H., Sasaki, N., Aoki, A. (1982), U.S.P. 4355151. Shimahara, K. (1990), Sen-i Gakkaishi 46(12), P-547 (in Japanese). Shimamura, K., Hashimoto, A. (1991), Sen-i Gakkaishi 47, P-325 (in Japanese). Sinclair, D. (1950), /. Appl. Phys. 21, 380. Smith, P., Lemstra, P. J. (1980 a), Colloid Polym. Sci. 7, 258. Smith, P., Lemstra, P. J. (1980b), J. Mater. Sci. 15, 505. Smith, P., Termonia, Y (1989), Polym. Commun. 30, 66. Southern, J. H., Porter, R. S. (1970), J. Macromol. Sci., Phys. B4, 541. Staudinger, H. (1932), in: Die Hochmolekularen Organischen Verbindungen: Berlin: Springer, p. 111. Stein, R. S. (1958), J. Polym. Sci. 31, 327. Snyder, R. W. (1991), in: High-Tech Fibrous Materials, ACS Symp. Ser. 457: Vigo, T. L., Turbak, A. F. (Eds.). Washington, D.C.: ACS, p. 124. Takahashi, T, Miwa, M., Sukurai, S. (1983), J. Appl. Polym. Sci. 28, Takahashi, T, Xiao, C. R, Sakurai, K. (1991 a), Sen-i Gakkaishi 47, P-397. Takahashi, T, Suzuki, K., Aoki, T, Sakurai, K. (1991b), J. Macromol. Sci., Phys. B30(l&2), 101.
651
Takakuda, K. (1991), Sen-i Gakkaishi 47(3), P-142 (in Japanese). Takayanagi, M., Ogato, T, Morikawa, Kai, T (1980), J. Macromol. Sci., Phys. B17, 591. Takayanagi, M. (1984), in: Interrelations between Processing Structures and Properties of Polymeric Materials: Seferis, J. C , Theocavis, P. S. (Eds.). Amsterdam: Elsevier. Tashiro, K., Nakata, Y, Li, T, Kobayashi, M., Chatani, Y, Tadokora, H. (1987), Sen-i Gakkaishi 43, P-627 (in Japanese). Tashiro, K., Nakata, Y, Li, T, Kobayashi, M., Chatani, Y, Tadokora, H. (1988), Sen-i Gakkaishi 44, P-7 (in Japanese). Termonia, Y, Meakin, P., Smith, P. (1985a), Macromolecules 18, 2246. Termonia, Y, Meakin, P., Smith, P. (1985b), Polym. Prepr. (Am. Chem. Soc, Div. Polym. Chem) 26 (2), 158. Termonia, Y, Smith, P. (1986), Polymer 27, 1845. Termonia, Y, Smith, P. (1987a), Macromolecules 20, 835. Termonia, Y, Smith, P. (1987b), Polymer Commun. 28, 60. Van der Zwaag, S., Kampschoer, G. (1988), in: Integration of Fundamental Polymer Science and Technology, 2nd Ed. London: Elsevier Chap. 2, p. 545. Van der Zwaag, S. (1989), J. Test. Eval. 17, 292. Van der Zwaag, S., Picken, S. J., Van Sluijs, C. P. (1989), in: Integration of Fundamental Polymer Science and Technology, 2nd Ed. Lemstra, P. J., Kleintjes, L. A. (Eds.). London: Elsevier Appl. Sci., Vol. 3, p. 199. Van Hutten, P. R, Koning, C. E., Pennings, A. J. (1985), /. Mater. Sci. 20, 1556. Vezie, D., Speck, J. S., Thomas, E. L. (1992), APS Symp. Bull. 37(1), 516. Walton, P. L., Majumdar, A. J. (1983), /. Mater. Sci. 18, 2939. Wang, L.-H., Porter, R. S., Kanamoto, T. (1990), Polym. Commun. 31, 457. Ward, I. M. (1985 a), in: Key Polymers: Properties and Performance, Advances in Polymer Science. New York: Wiley-Interscience, p. 1. Ward, I. M. (1985b), Adv. in Polym. Sci. 70, 1. Ward, I. M. (1987), Development in Oriented Polymers - 2. Amsterdam: Elsevier, p. 39. Watt, W, Perov, B. V. (1985), Strong Fibers, Handbook of Composites Series: 1. Amsterdam: Elsevier Science Publishers. Weibull, W. (1951), /. Appl. Mech., 18, 293. Weng, T, Hiltner, A., Baer, E. (1986), /. Mater. Sci. 21, 744. Wierschke, S. G. (1989), in: The Materials Science and Engineering of Rigid-Rod Polymers: Adams, W. W, Eby, R. K., McLemore, D. E. (Eds.). Pittsburgh, PA: MRS Symp. Proc. Vol. 134, p. 313. Wierschke, S. G., Shoemaker, J. R., Haaland, P. D., Pachter, R., Adams, W. W. (1992), Polymer 33(16), 3357.
652
13 High Performance Polymer Fibers
Wiff, D. R., Helminiak, T. E., Hwang, W.-F. (1988), in: High Modulus Polymers: Zachariddes, A. E., Porter, R. S. (Eds.). New York: Marcel Dekker, Chap. 8, p. 225. Wiff, D. R., Lenke, G. M. (1992), in: 37th Int. SAMPE Symp., p. 991. Wolfe, I (1985 a), U.S.P. 4533692. Wolfe, J. (1985b), U.S.P. 4533693. Wolfe, J. (1985c), U.S.P. 4533724. Wolfe, J. (1988), in: Encyclopedia of Polymer Science and Engineering: Kroschwitz, J. I. (Ed.), //: 11. New York: Wiley, p. 601. Wolfe, J. F. (1989), in: The Materials Science and Engineering of Rigid-Rod Polymers: Adams, W W, Eby, R. K., McLemore, D. E. (Eds.). Pittsburgh, PA: MRS Symp. Proc. Vol. 134, p. 83. Won Choe, E. W, Kim, S. N. (1981), Macromolecules 14, 920 Wong, C. P., Ohnuma, H., Berry, G. C. (1978), /. Polym. Set 65, 173. Wu, H. E, Phoenix, S. L., Schwartz, P. (1988), /. Mater. Sci. 23, 1851. Wynne, K. I, Zachariades, A. E., Inabe, T., Marks, T. J. (1985), Polym. Commun. 26, 162. Xiao, C. E, Takahashi, T. (1990), Sen-i Gakkaishi 46, T-49. Yabuki, K., Ito, H., Oda, T. (1975), Sen-i Gakkaishi 31, T-524. Yabuki, K., Ito, H., Oda, T. (1976), Sen-i Gakkaishi 32, T-55. Yamada, S. (1987), /. Chem. Soc, Chem. Commun. 309, 1448. Yamada, S. (1991), Sen-i Gakkaishi 47, P-24 (in Japanese). Yang, H. H. (1982), U.S.P. 4340559. Yang, H. H. (1989), Aromatic High-strength Fibers. New York: Wiley. Yoon, H. N. (1990), Colloid Polym. Sci. 268, 230. Young, R. X, Day, R. X, Zakihhani, M. (1990), J. Mater. Sci. 25, 127.
Young, R. X, Lu, D., Day, R. X, Knoff, W. E, Davis, H. A. (1992), /. Mater. Sci. 27, 5431. Zachariades, A. E., Porter, R. S. (1983), The Strength and Stiffness of Polymers. New York: Marcel Dekker. Zhang, R., Mattice, W L. (1992), Macromolecules 25, 4937. European Patents: 83108258; 84114872.
General Reading Adams, W W, Eby, R. K., McLemore, D. E. (Eds.) (1989), The Materials Science and Engineering of Rigid-rod Polymers. MRS Symp. Proc, 134. Baer, E., Moet, A. (1991), High Performance Polymers. New York: Hanser (distributed by Oxford University Press). Donald, A. M., Windle, A. H. (1992), Liquid Crystalline Polymers. Cambridge: University Press. Hongu, T, Philips, G. (1990), New Fibers. New York: Ellis Horwood. Kelly, A., Macmillan, N. H. (1986), Strong Solids. Oxford: Oxford Science Press. Northolt, M. G., Sikkema, D. X (1991), in: Advances in Polymer Science 98. Berlin: Springer-Verlag. Ohta, T., Kunugi, T, Yobuki, K. (Eds.) (1988), High Tenacity and High Modulus Fibers. Society of Polymer Sciences, Japan, Tokyo: Kyoritsu Shuppan Co. Ltd. Vigo, T. L., Turbak, A. F. (Eds.) (1991), High-Tech Fibrous Materials. Washington, DC: ACS Symp. Series 457. Watt, W, Perov, B. V. (Eds.) (1985), Strong Fibers. Amsterdam: Elsevier Science Publishers. Yang, H. H. (1989), Aromatic High-Strength Fibers. New York: Wiley.
14 Polymer Surfaces and Interfaces with Other Materials Matthew Tirrell and Edward E. Parsonage Department of Chemical Engineering and Materials Science, University of Minnesota, Minneapolis, MN, U.S.A.
List of Symbols and Abbreviations 14.1 Introduction and General Considerations 14.1.1 Motivation 14.1.2 Scope and Aims 14.2 Computer Simulation 14.2.1 Monte Carlo and Molecular-Dynamics Results 14.2.2 Implications and Interpretations 14.3 Surface Tension of Polymeric Materials 14.3.1 Introduction 14.3.2 Experimental Techniques for Measuring Surface Tension or Surface Energy 14.3.2.1 Polymer-Liquid Surface Tension 14.3.2.2 Polymer-Solid Surface Energy 14.3.3 Experimental Results for the Surface Tension of Polymeric Materials 143 A Theoretical Aspects of Polymer Surface Tension 14.4 Techniques to Examine Polymer Surface Properties 14.4.1 General Considerations 14.4.2 Imaging Methods 14.4.3 Reflection Methods 14.4.4 Spectroscopic Methods 14.4.5 Ion-Beam Methods 14.5 Materials Science Issues Concerning Polymer Surfaces 14.5.1 Overview 14.5.2 Surface Composition 14.5.3 Surface Ordering 14.5.4 Thin Films on Solids: Wetting, Spreading, Interactions and Dynamics 14.6 Conclusions and Remarks on Future Directions 14.7 References
Materials Science and Technology Copyright © WILEY-VCH Verlag GmbH & Co KGaA. Allrightsreserved.
654 657 657 658 659 659 663 667 667 668 668 670 673 675 679 679 680 682 683 685 686 686 686 689 692 693 694
654
14 Polymer Surfaces and Interfaces with Other Materials
List of Symbols and Abbreviations a
K b D E F AF M AF C T / k K L M
Mn n "i
N P P q R,R\,R2
s s SB
Sc T Tg T* U Vs
z Y 7 7C A £ bb 0 K X /Hi
radius of contact area (between two solids) area (surface region) site size (in a lattice) thickness of surface region internal energy external force, applied load tensile force to separate two solids Flory-Huggins free energy of mixing surface free energy per segment intensity Boltzmann constant bulk elastic modulus period repeat-unit molecular weight number-average molecular weight Macleod exponent number of moles of /th component number of segments in a polymer chain length, or degree of polymerization pressure "parachor" wave vector radii of elastic spheres (JKR theory) radius of gyration entropy surface area segmental-order parameter chain-order parameter temperature glass-transition temperature reduced temperature energy volume (surface region) number of chain segments surface tension (excess free energy per unit area) solid-surface energy in vacuum critical surface tension phase difference of components of an electric field bead-bead (or segment-segment) interaction parameter contact angle, incident angle, or take-off angle (in XPS) compressibility wavelength, or mean free path (in XPS) chemical potential of /th component
List of Symbols and Abbreviations
655
E, ;re p a 0 0e 0 X W
correlation length spreading pressure density, or density of chain ends segment diameter average monomer density surface-excess composition volume fraction of segments in a lattice Flory-Huggins parameter used in tan*F, where this term represents the amplitude ratio of components of an electric field
AFM B-PE d-PS DSIMS ELLI ESCA FRES HREELS IETS IR-ATR IR-GIR JKR L-PE NR NRA PCP PDMS PEO PEP-PEE PET PIB PMDA-ODA PMIM PMMA PnBMA PP PPO PS PS-PMMA PYAC PVME PVP RBS rms
atomic-force microscopy branched poly(ethylene) deuterated poly(styrene) dynamic secondary-ion mass spectroscopy (spectrometry) ellipsometry electron spectroscopy for chemical analysis forward-recoil spectroscopy (spectrometry) high-resolution electron-energy-loss spectroscopy (spectrometry) inelastic electron-tunneling spectroscopy (spectrometry) infrared-attenuated total-reflection spectroscopy (spectrometry) infrared grazing-incidence reflection spectroscopy (spectrometry) Johnson-Kendall-Roberts (theory) linear poly(ethylene) neutron reflectometry (reflectivity) nuclear-reaction analysis poly(chloroprene) poly(dimethylsiloxane) poly(ethylene oxide) poly(ethylene propylene)-poly(ethyl ethylene) block copolymer poly(ethylene terephthalate) poly(isobutylene) poly(pyromellitic dianhydride oxydianiline) phase-measurement interference microscopy poly(methyl methacrylate) poly(n-butyl methacrylate) poly(propylene) poly(propylene oxide) poly(styrene) poly(styrene)-poly(methyl methacrylate) diblock copolymer poly(vinyl acetate) poly(vinyl methyl ether) poly(vinyl pyridine) Rutherford backscattering root mean square
656
SCMF SEM SIMS SSIMS STM TEM VASE XPS XR
14 Polymer Surfaces and Interfaces with Other Materials
self-consistent mean field scanning electron microscopy secondary-ion mass spectroscopy (spectrometry) static secondary-ion mass spectroscopy (spectrometry) scanning tunneling microscopy transmission electron microscopy variable-angle spectroscopic ellipsometry X-ray photoemission spectroscopy (spectrometry) X-ray reflectometry (reflectivity)
14.1 Introduction and General Considerations
14.1 Introduction and General Considerations 14.1.1 Motivation Surfaces are the regions through which materials connect and interact with their surroundings. Transmission of stress, adhesion, friction, abrasion, permeability to gases and liquids, compatibility with biological or harsh, corrosive environments are all properties of polymeric materials which are dominated by the properties of the material within tens of nanometers of the surface. This statement is not unique to polymeric materials. However, surfaces of polymeric materials have several special features, relative to surfaces of other materials, lending particular scientific interest to their study. Large molecular size is the sine qua non of polymers, which at surfaces means that an individual molecule at the surface may also extend relatively deeply into the bulk of the material, thus effectively enlarging the zone of material affected by the surface. Molecular connectivity of macromolecules effectively links the surfaces to the interiors of polymer materials. Surface regions of polymers are usually less sharply delineated than those in other materials. The gradients in structure resulting from the influences of the surface are typically less abrupt in polymers and consequently have a larger characteristic dimension. Furthermore, the three-dimensional coil configurations of macromolecules must arrange themselves in special ways to accommodate the spatial restriction incurred by the surface. This point must be comprehended in order to understand and effectively manipulate the surface properties of polymers to the fullest extent. Polymer materials are usually partially or completely amorphous solids, or viscous or
657
rubbery liquids, so that familiar ideas in ordered solids, of surface crystal structure, reconstruction or defects seldom apply in a simple way to polymers. Near surfaces, polymeric materials are sometimes found to be more ordered than in the bulk (Factor et al., 1991; Menelle et al., 1992), in contrast with other kinds of materials, owing to the influence of the surface constraints on the packing of macromolecules. At the same time, macromolecules are frequently multicomponent molecules, either by design or by accident. This means that an individual molecule may contain more than one (possibly several) distinct chemical functionalities. Polymer molecules that are amphiphilic (meaning multiple chemical affinities within the same molecule) in character, such as block copolymers, can be synthesized (see Chap. 1 of this Volume). Oxidation or other environmental exposure (or intentional chemical modification) can modify portions of individual macromolecules, thus chemically also imparting different surface affinity to those segments. In either case, one finds that certain molecules, or certain portions of molecules, segregate to the surfaces of the material and thereby can assume a dominant role over the average bulk composition in determining how the material interacts with its surroundings. This behavior could be thought of as the counterpart to surface reconstruction of inorganic crystalline materials (Allen and Gobeli, 1962). In technological practice, the tendency toward surface segregation in multicomponent polymer systems provides many real and potential advantageous opportunities for interfacial engineering of polymers, since properly designed macromolecules will self-assemble at the surface where they are desired. Processing can also be used to accomplish what pure thermodynamics only does reluctantly or sluggishly (for example, quenching by cooling or solvent remov-
658
14 Polymer Surfaces and Interfaces with Other Materials
al to trap useful non-equilibrium structures, shearing to induce surface ordering, or surface chemical reactions to tailor interfaces (Bergbreiter et al., 1984; Ward and McCarthy, 1987)). Surfaces are the regions through which polymers make contact with other materials. The large size of polymer molecules, and the related long-range surface gradients, imply that surface polymer molecules often intermingle or entangle substantially with adjacent material in their surroundings. This characteristic is important inter alia in adhesion, in the surfaces of polymer materials exposed to solvents that swell or partially dissolve them, and in the interdiffusion or welding of polymer materials placed in contact with one another.
14.1.2 Scope and Aims The broad area of polymer surfaces and interfaces is too large to cover effectively in a single chapter, so this chapter is intentionally focused, and some arbitrary decisions have been made about the scope of the coverage. Bulk polymer materials are dealt with exclusively, leaving out many interesting aspects arising from interactions with solvent or in solvent-swollen, adsorbed layers of polymers on other solids. Two comprehensive reviews on polymer adsorption have recently appeared (Ploehn and Russel, 1990; Kawaguchi and Takahashi, 1992); surface modification due to grafting chains on to solids has been covered in another review (Halperin et al., 1992). Surface chemistry of polymer materials is discussed in a cursory fashion here; an excellent review is available elsewhere (Ward and McCarthy, 1987). Techniques of experimental characterization and analysis of polymer surfaces are surveyed mainly by describing the kind of information they provide on
polymer surface properties. Thorough and informative reviews of polymer surface analysis methods have recently been published (Russell, 1990,1991;Gardella, 1989; Gardella and Pireaux, 1990; Stamm, 1992). Polymer surfaces and interfaces are frequently discussed in tandem. In the parlance we adopt for this article, "surface" means an external interface of a piece of polymer material, most often in contact with air or another impenetrable solid, whereas "interface" is used to mean internal or buried interfaces which can result from microphase separation in multicomponent polymers, or from the contacting of two different polymers. This chapter deals more with external surfaces than with internal interfaces. Thin films of polymers, where the entire sample experiences restriction in the dimensionality of space, merit special, concentrated attention which they will not fully receive here; however, some additional aspects of thin films, which have two external surfaces, will be brought out in the general discussion of surfaces. The aim of this chapter is to bring out clearly the aspects of polymer surfaces relating most directly to their physical structure and properties. All of the available information on the configurations and surface arrangements of macromolecules in polymeric materials is discussed. Some of the most informative work, albeit on very idealized systems, has been done by computer-simulation experiments, so that this is the point of departure for the chapter. More readily, and therefore more extensively, characterized by laboratory experiments than surface configurations are surface energies and tensions so this is the next topic presented, along with the theoretical approaches that have been used in this area. This is followed by a summary of the arsenal of experimental techniques that are currently being used to study polymer surfaces.
14.2 Computer Simulation
The final section presents information on a range of materials-science issues pertaining to polymer surfaces.
14.2 Computer Simulation 14.2.1 Monte Carlo and MolecularDynamics Results Macromolecules near the surfaces of bulk polymeric materials experience conflicting influences. An undiluted polymer material attempts to maintain its bulk density right up to the surface. An enthalpic penalty must be paid to incorporate void space and reduce the density below the equilibrium bulk value. On the other hand, the constituent macromolecules feel increasing constraints on the number of configurations that they can adopt near the surface, and thus they pay an entropic penalty in trying to maintain bulk density near the surface. Surfaces and interfaces are, of course, zones of finite width over which the transitions in density and polymer configuration are made. The thickness of these zones is governed by the factors just mentioned. For multicomponent polymers, the surface affinities of the various components also play a key role. For a single component, homopolymer material, full knowledge of the state of a polymer surface region would comprise information on the profile of density change through the surface, including its characteristic width, and information on the polymer configuration as a function of position near the surface, including the positions of the centers of mass of the molecules, the distribution of dimensions of the macromolecules, measured both parallel and perpendicular to the surface, and the positions of the chain ends.
659
Over the past several years, a considerable number of computer-simulation experiments have been conducted to probe both the equilibrium and dynamical properties of polymer surfaces. These have the advantage over the current state of laboratory investigations of polymer surface structure in that they produce completely detailed, visualizable and quantifiable pictures of polymers at surfaces. However, in assessing the results of simulations it must always be borne in mind that they are only as good as the rules and representations of interaction from which they are built. Madden opened this line of inquiry, performing extensive lattice Monte Carlo simulations for a film adsorbed against a hard wall and exposed to vacuum (Madden, 1987, 1988). The film thicknesses were greater than several tens of segment diameters, i.e. not so thin that the two surfaces interacted with one another. This work has the feature of comparing directly two kinds of polymer surfaces of interest, i. e. hard, impenetrable solid surfaces and interfaces with vacuum (or air or vapor). In one case, a sharp boundary is imposed, whereas in the other, the density profile can broaden by the incorporation of density defects in the vacuum interfacial region. At most temperatures, the similarities between these two types of surfaces far outweigh the differences. In general, lattice-type simulations may be expected to yield reliable information on large-scale properties where the detailed structure of the lattice is not important. Some representative density profiles from the lattice Monte Carlo simulations of Madden (Madden, 1987) for a melt-vacuum interface are shown in Fig. 14-1. Results are shown as a function of the reduced temperature T* for the simulation of a polymer film having an average degree of polymerization of 100. (Madden also investigated the effects of polydispersity; Figure 14-1
660
14 Polymer Surfaces and Interfaces with Other Materials
1
1
1
1
gment
1
1
1
1
1
.09 !
#**
% ^ gnu
nirnTT
3
I
.70
I
.80
H
.60
•X
D
.50
-
+
.40
GO
1
I
Fract ion
1.0
S$SSS5S$
>
.30 .20
X
.10 .00
+ 1
1
10
20
n \ D I
1^
+
30
40
50
60
70
1^555^555555;
80
90
100
Z/o Figure 14-1. The results of a Monte Carlo simulation of a layer of polymer melt between an impenetrable solid (left) and vacuum (right), shown for several reduced temperatures, 7* (as defined in text): +, 7* = 1.5; #, T* = 1.75; x, T* = 2.0; D, T* = 2.5; V, 7* = 3.0. Z is distance from the solid surface and o is the segment diameter. The density in the middle of the film "reaches that of the bulk" and decreases with increasing temperature. The solid wall enforces an essentially constant total density right up to the surface. At the vacuum surface, the density is also constant and the density profile is then quite sharp, except at the highest temperature (Madden, 1987).
corresponds to a rather low value of polydispersity, i.e. 1.08.) The dimensionless temperature is given by T* = kT/ebb, where k is Boltzmann's constant, T is temperature and £bb is the bead-bead (or segment-segment) interaction parameter on the lattice. In this lattice model, the corresponding interaction energy parameters for segmentvacuum and vacuum-vacuum were set to zero. The results indicate a narrowing of the interfacial region with decreasing temperature and a corresponding increase in the bulk density. Measured in terms of the width of the density gradients, the surface region is narrow, except at high temperature, and covers only several segment diameters. Other results, which are not shown here (Madden, 1987, 1988), demonstrate, inter alia, that there is a peak in the density of chain centers of mass, which is located at a distance comparable to the radius of gyration away from each surface. No such inhomogeneity shows up in the total polymer density (Fig. 14-1).
Lattice Monte Carlo simulations were pursued by ten Brinke et al. (1988), who confirmed this latter feature of Madden's work for a polymer melt (actually a rather sparse one at an average polymer volume fraction of 0.8, which is comparable to Madden's case where T* = 2.0) in contact with a solid wall. Centers of mass are depleted from the immediate vicinity of the hard wall. It was noted in this work that there was a distinct enrichment of chain ends within a few segment diameters away from the wall. For a series of chain lengths (N = 40, 60 and 80 segments), the average degree of enrichment over the bulk density of chain ends was 8 + 2%, apparently independent of chain length, though as we shall discuss directly, this conclusion of a molecular-weight independence appears to be incorrect. Off-lattice Monte Carlo simulations were subsequently pursued in order to examine local structural details (Dickman and Hall, 1988; Yethiraj and Hall, 1990;
14.2 Computer Simulation
Kumar et al., 1988, 1990; Vacatello et al., 1990). We focus on the results of Kumar et al. (1988, 1990) here, as they bring out a range of interesting general properties of polymer surfaces. Kumar and co-workers confirmed that there is enrichment of chain ends near a neutral solid wall. In contrast to the work of ten Brinke et al. (1988), the enrichment found by Kumar et al. was stronger, reaching about 40% for long chains, and was dependent upon the molecular weight. The difference in these findings does not apparently relate to the specific, underlying lattice in the simulations carried out by ten Brinke and his co-workers, since molecular-weight dependent chain-end enrichment at solid surfaces has also been seen in very recent lattice Monte Carlo simulations by Wang and Binder (1992). In all of the discussions of chainend enrichment thus far, we are considering cases where there is no enthalpic factor favoring chain-end adsorption, i. e. there is no extra adsorption affinity of the chain ends, a factor which can be very important if it occurs. Enrichment of chain ends occurs in immediate proximity to the surface, within a distance of about one segment diameter (cr).
661
This produces a depletion in the density of chain ends over the next few segment diameters of distance from the surface. The profile of chain-end density settles down to the homogeneous bulk level beyond about 2Rg (where Rg is the radius of gyration). An example of the density profile of chain ends is shown in Fig. 14-2. An empirical relationship suggested by Kumar et al. (1990) for the molecular-weight dependence of the chain-end density within one segment diameter of the surface [pe(cr = 1)] is pe(<7= 1) = a + b exp(-c/N)
(14-1)
where a, b and c are fitted constants. The density profile of chain centers-of-mass, shown in Fig. 14-3, is peaked at about 3(7 from the wall, less than the bulk Rg for chains of this size, and achieves its bulk value at a distance of about 2 Rg from the wall. The work of Kumar et al. (1990) brings out clearly several additional effects of surfaces on the configurations and shapes of molecules in undiluted polymer materials. The average dimensions of the chains measured in the direction normal to the wall are reduced relative to those measured in the bulk, for those chains where the center of mass is located within Ro of the wall.
1.50
Pi
Figure 14-2. Density of chain ends in a polymer, relative to that in the bulk material, as a function of the distance £ (= Zlti) away from a solid wall, measured as distance divided by the segment diameter, with the chain lengths A, N = 50; O,N= 100; O, N = 200 (adapted from Kumar et al., 1990).
662
14 Polymer Surfaces and Interfaces with Other Materials
Figure 14-3. Density of chain centers of mass in a polymer, relative to that in the bulk material, as a function of the distance £cm of the center of mass (4m ~ Zcm/o) away from a solid wall, measured as distance divided by segment diameter, with chain lengths A, N = 50; O: N - 100 (adapted from Kumar et al., 1990).
Furthermore, as illustrated in Fig. 14-3, the reduction in dimensions normal to the surface scales with Rg. Chain dimensions parallel to the surface are greater on average than those in the bulk. The total radii of gyration appear to be independent of position. Asymmetry appears in the chain configuration for chains within 2Rg of the surface, owing to the fact that the "wall-side", of macromolecules close to the wall, is flattened. These latter two points will be taken up again later in this section and in Section 14.2.2. Chain segments are strongly oriented, but only very near (< 2a) to the wall, where they exhibit a segmental order parameter, SB of -0.2 (SB = [ 3 < c o s 2 0 x > - l]/2, with < cos 2 0 x > being the average square of the cosine of the angle made by all the segments with the x-axis normal to the surface) (Kumar et al., 1990). (As the definition of SB indicates, SB = -0.5 would correspond to perfect orientation in the plane of the surface.) At a > 2, the segmental order parameter is zero. Whole chains are, on average, more weakly oriented than segments. A chain order parameter (defined by Sc = [3 < cos2Oxi > - l]/2, where the average is taken over all the segments of the ith chain)
indicates a weaker parallel tendency at ~ -0.08, but is detectably different from zero out to distances of 2Rg of the center of mass from the wall. These latter features are elaborated more fully in the work of Mansfield and Theodorou (1989). They studied a polymer-vacuum interface, so that the general similarity of their configurational results to those of Kumar and his co-workers for polymersolid interfaces, confirms the essential connection between these two situations. Mansfield and Theodorou extended previous work to investigate the question of whether the flattening that they observed, and which also can be seen in Fig. 14-3, is due to a change in the intrinsic shape of the chains, or whether it is caused by a tendency for chains as whole entities to assume orientations parallel to the surface. To investigate this, they examined two quantities that measure the intrinsic shape of the macromolecules independently from the orientation, namely the spans of the chains and the eigenvalues of the radius-of-gyration tensor. These are quantities (defined precisely by Mansfield and Theodorou) which measure the shape of a macromolecule relative only to its internal coordinates.
14.2 Computer Simulation
They show conclusively that, within the precision of their simulations, the overall shapes of the macromolecules are practically unperturbed, even within one radius of gyration of the surface. Macromolecules in bulk are generally found in a distribution of anisotropic configurations (Sole and Stockmayer, 1971); spherical symmetry is a property only of the average configuration over the entire distribution. The implication of the results of Mansfield and Theodorou (1989) is that the surface region selects from the bulk distribution those molecules with a tendency to align their long axes parallel to the surface. This conclusion is in accord with and illuminates the observation of Kumar et al. (1990) that the overall radius of gyration of the molecules varies little from the bulk values even through the surface region. Carton and Leibler (1990) have shown that the conformations of anisotropic objects adapt in a predictable and general manner to regions of inhomogeneous density, such as polymer surfaces. Another significant piece of work contributing to the understanding of polymer surfaces via computer simulations is the molecular-dynamics study by Bitsanis and Hadziioannou (1990). This is both the most rigorous and the most elaborate means for simulating realistically the properties of polymer-melt surfaces. It is necessary to follow this route if one is interested in dynamic properties (e.g. diffusion near surfaces), as these authors were; structural information, of the type we have discussed here, is also produced by the simulation but it is not necessarily superior to that obtained via Monte Carlo methods. All of the structural features described above were also observed by Bitsanis and Hadziioannou (1990). These workers, along with Wang and Binder (1991, 1992), also studied the effects of preferential surface attraction or
663
repulsion by certain segments. This is important in situations where certain segments are chemically modified, thus producing surface activity. Predictably, preferential attraction of chain ends brings them to the surface (Wang and Binder, 1992); somewhat less obvious is the fact that modifying all of the segments so that they have an attraction for the surface diminishes the enhancement of chain ends at the surface. In retrospect, this is an effect to be anticipated, since a chain of TV segments has only two ends; if 7V»2, as usual, the interior segments will win the competition for the surface. Bitsanis and Hadziioannou (1990) showed that, near the surface, the molecular mobility parallel to the surface increased while that measured perpendicular to the surface decreased; surface attraction reduced the mobility still further. Chain connectivity appears to reduce the surface ordering induced by a solid wall relative to that which would appear in a fluid of disconnected, but otherwise identical, segments (Horn and Israelachvili, 1981). 14.2.2 Implications and Interpretations The overall picture of the facts emerging from computer simulation experiments is rather clear and consistent over a variety of simulation methods. The low compressibility (at all but quite high temperatures) of dense bulk polymer forces the average polymer density (over distances comparable to the segment diameter) near a surface to assume a value very close to the bulk density. The influence of a surface on chain conformations is essentially exclusively exerted on those segments that are within approximately 2crof the surface. These segments have a propensity to align parallel to the surface. This produces some local asymmetry of polymer configurations but does not distort the overall shapes of the macro-
664
14 Polymer Surfaces and Interfaces with Other Materials
molecules very much. The average shapes of macromolecules near the surface are similar to those in bulk but in this case they have their long axes preferentially aligned with the surface. Chain ends are enhanced in density within 2<7 of the surface by as much as 40%, depending somewhat on possible attractive interactions with the surface (de Gennes, 1987). Consequences for adhesion with other materials are to be anticipated (Prager and Tirrell, 1981; Brochard and Pincus, 1992). In the absence of specific interactions with the wall, this can be understood (Helfand and Tagami, 1971) by an argument based on random walks. Simply put, placement of a middle segment of a chain near an impenetrable surface requires that two random walks emanate from the point of that placement without crossing the impenetrable barrier. This is demonstrably (Helfand and Tagami, 1971) more difficult entropically, and therefore of lower probability, than the accommodation necessary for the single, albeit longer, random walk resulting from placing a chain end segment near the surface. Owing to the connectivity of chain molecules, some of the effects exerted on the segments in direct propinquity to the surface are transmitted over distances of the order of the size of the polymer coils in depth away from the surface. Both the density of chain centers of mass and the alignment of anisotropic chain configurations exhibit surface enhancements within 2Rg of the surface. Thus, 2(7is the length scale for surface perturbation of segments, whereas 2Rg is the length scale for surface modification of whole chain properties. Polymer materials exhibit more nonlocal effects of their surfaces than other materials, and these are of some importance in the understanding of polymer surface tensions (Rabin, 1984). However, most features at the segment level of polymer-surface structure are indepen-
Overlappmg polymers in bulk
Creation of a surface by configurational swap maintaining constant density Figure 14-4. Schematic illustration of the idea of configurational swap. In the top picture, two chains are overlapping as in the bulk polymer. To create a surface along the dashed line, while not changing the density of the polymer segments, those portions of the lightly and heavily drawn chains that cross the line are folded, or reflected, back on to their sides of the surface, the effect of which is to leave unaffected the population of rotational states about each bond, but to compress the overall extent of the macromolecules normal to the surface (Silberberg, 1988).
dent of the molecular weight of the entire chain. (Surface chain end enrichment is an exception to this statement; see Eq. 14-1.) Figure 14-4 illustrates schematically a simplified way to think about how and why polymer-surface configurations differ from configurations in the bulk. Essentially all of the above characteristics obtain for both polymer melts and glassy solid surfaces in contact with either hard solid walls or vacuum (or air) (Theodorou, 1988, 1989). The only significant difference between these closely related situations is that in free surfaces the total density profile has somewhat more freedom to accommodate than at solid boundaries (see Fig. 14-1); at many temperatures of interest, the vacuum interface behaves effectively like a solid boundary for the polymer. Crystalline surfaces of polymers have not yet been studied extensively by computational methods. Theoretically, most of the phenomena seen in simulations are understood to a rea-
14.2 Computer Simulation
sonable degree. Self-consistent mean-field (SCMF) theories (in a variety of related forms; see Scheutjens and Fleer, 1980; Freed, 1987; des Cloiseaux and Jannink, 1990) can predict many aspects of the chain configurations observed in simulations, such as the enhancement in density of chain ends near the surface. SCMF theory models the polymer chain as a random walk, distorted by a potential field arising from interactions among neighboring segments (Helfand and Tagami, 1971). Thus the positions of the segments and the effective potential is determined self-consistently. Lattice versions of SCMF theories are good vehicles to predict the observed segment order parameters (Helfand, 1976; Helfand and Weber, 1976). For example, a model due to Helfand (1976), predicts the segment order parameter in the first lattice layer near the surface to be -0.22, then decaying to zero beyond the second layer, in excellent agreement with the results of Kumar et al. (1990). A particularly illuminating and appealing, if somewhat simplistic, theoretical view of polymer configurations at surfaces of undiluted polymers is that offered by Silberberg (1988). He has elaborated a concept known as "configurational swap", which is essentially coincident with the SCMF theory for polymer configurations under the imposition of a reflecting boundary condition at the polymer surface, but is simpler computationally. [It is comparable to the limit of infinitely strong incompatibility and zero compressibility in the model of Helfand and Tagami (1971)]. Imbedded in this analysis is the idea that the surface exerts no influence on the local random orientation of segments, or on the random placement of chain ends, right up to the surface. This, we have seen from simulations, is not true in detail but is valid beyond 2crfrom the surface. "Configura-tional swap" is thus a
665
reliable means to model the global configurational properties of the chain as a whole. The procedure, following Silberberg (1988), is to imagine an isotropic, homogeneous polymer melt in thermodynamic equilibrium through which a surface is then laid in such a way that one infinite plane of segments is traversed by it. If this surface is to mimic an effectively impenetrable solid or vacuum surface, the cutting of the surface by chains must be removed. "Configurational swap", or the reflecting boundary condition, which eliminates crossing of the surface while maintaining bulk density, proceeds by moving along the surface and uncoupling all chains that lie across the surface according to the following procedure. For each chain, or part of a chain, on one side of the surface there is a corresponding mirror-image chain, or part of a chain, on the other. Such chains are uncoupled by removing all parts of the chain from one side of the surface and putting in their place the mirror-image parts from the mirror-image chains. Correspondingly, into the mirror-image parent chains are coupled those parts of the original chains that had been uncoupled. Thus, total density is preserved, as all bonds are replaced by ones that are energetically equivalent; thus, there is no change in the internal energy of the system, although there is an entropy decrease. This idea is depicted in Fig. 14-4. The methodology of the analysis is illustrated in Fig. 14-5. In the absence of a surface, the contribution of the average monomer density 0, in the x-direction from those chains having their center of mass between x0 and x0 + dx0 is given by (Silberberg, 1988) 0 (x, x0) = 0o exp [-p (x - x0 )2]
(14-2)
where 0O is a constant and /?= 3/2 No2. This function arises from the basic Gaussian dis-
666
14 Polymer Surfaces and Interfaces with Other Materials I
0L
Figure 14-5. Segment distribution in the polymer melt. A macromolecule, with its center of mass at JC0, has the segment distribution E'ACD, while a macromolecule with its center of mass at -x0 has the segment distribution D'C'AE. If the plane at x - 0 is a surface, the portions of the distributions E'AO and EAO are, respectively, disallowed and swapped. This gives the distribution OABCD. The center of mass is shifted from x0 to xG, and the mean radius of gyration normal to the surface is reduced from Rgx to Rg x (adapted from Silberberg, 1988).
tribution governing the chain configurations in the bulk polymer. When the polymer is cut by a surface, the configurational statistics of the nearby polymer chains will adjust with the constraint of maintaining bulk density, while not crossing the surface. The resulting density profile for those chains originally at x0 is given by [-/?(*- x0)2]
(x, x0) = exp [-
(14-3)
As illustrated in Fig. 14-5, the net effect is to shift the chain center of mass from x0 to xG = x0 erf (x0 /31/2) + (TT/T 172 exp (-/k 2 ) (14-4) and to reduce the Jt-component of the radius of gyration: Rg,x =Rg,Xo-(*G-Xo)
(14-5)
where Rgx is the x-component of the radius of gyration in the "uncut" polymer. The overall effect is illustrated in Fig. 14-6, where the segmental density contribution is plotted for chains at various distances from the surface. In addition to the local density gradients discussed earlier, it is evident that a surface can exhibit nonlo-
cal or long-range effects on a length scale comparable to the unperturbed size of the polymer coils. The overall effect of Eqs. (14-3) - (14-5) is such that polymer molecules in the near-surface region take on relatively flattened configurations in a direction parallel to the surface, just as seen in the simulations described earlier (Mansfield and Theodorou, 1989). Furthermore, the modified density distribution [see Eq. (14-3)] produces a maximum in the centerof-mass density at a distance corresponding approximately to Rg 0. Figure 14-6 shows clearly that chains which have their centers of mass located at or beyond about 2Rg are essentially unperturbed in their Gaussian distribution of segments found in bulk. These effects, which were seen in many of the simulation results discussed earlier, produce important effects on surface properties, such as non-local contributions to the surface energy, with further ramifications on properties such as adhesion and wetting yet to be fully explored.
14.3 Surface Tension of Polymeric Materials
667
*O/31/2=2.O|
1.0
x o /?1 / 2 = 1 1.0 -
Reduced Distance from Surface
Figure 14-6. Segment distribution at different distances from the surface of a polymer melt. The numbers in boxes refer to the positions of the centers of mass of the unperturbed chains, JC0, scaled by the average size of those chains, with the distance from the surface, x, being scaled in the same way: see text for the definition of /3; , location of center of mass; , radius of gyration, Rg (Silberberg, 1988).
14.3 Surface Tension of Polymeric Materials 14.3.1 Introduction Surface energetics are arguably the most important characteristics of material surfaces. Surface tensions, or for solids, more appropriately, surface energies, express the work that must be done to create surface. In turn, they largely dictate many aspects of how material surfaces interact with one another macroscopically. We will use the terms surface tension and surface energy essentially interchangeably in this chapter, although the latter is more general and the former should strictly be applied only to
liquids. Tension refers to forces arising from the stresses associated with the reversible creation of new surface area, a process which is often experimentally infeasible for solids. Here, however, we will use both surface tension and surface energy to characterize the reversible work required to create unit surface area. In the previous section, attention was focused on the molecular aspects of polymer surfaces. It was observed that although the density drops precipitously from bulk to essentially zero over a distance comparable to several segment diameters, the average conformations of the chains were affected on a much larger scale comparable to the unperturbed dimensions of the coils. The de-
668
14 Polymer Surfaces and Interfaces with Other Materials
crease in configurational entropy associated with these non-local effects, along with the increase in internal energy associated in part with the steep density gradients, as well as new energetic interactions experienced by the segments in the interfacial region, give rise to an excess free energy for the system as a whole. In terms of thermodynamic variables, the excess free energy (per unit area), % is given by the relation y= (E-TS + PVS - X Nt
(14-6)
where E is the internal energy, T absolute temperature, S the entropy, P the pressure, Vs the volume, As the area, and Nt and fa are the number of moles and chemical potential, respectively, of the /th component, all of these quantities referring to the surface region. The first three terms on the right hand side of Eq. (14-6) represent the free energy of the system with the surface. X NiHi represents the free energy of a hypothetically identical mixture of the same components without the surface. Thus, the surface tension, y, represents the excess free energy, per unit area, associated with the presence of the surface. Surface tension can be modeled theoretically by means related to the SCMF methods mentioned in the preceding section. However, before discussing theory, we examine experimental methodology and results for surface tension measurements. 14.3.2 Experimental Techniques for Measuring Surface Tension or Surface Energy The type of experimental technique employed for the measurement of surface energy depends in large part on the state of the polymer sample. Comprehensive reviews of procedures and results have been given by Wu (1982) and by Koberstein (1987). For surfaces of polymers in the liquid state, sur-
face tension is most conveniently measured by one of a variety of methods that applies forces to the surface and measures the resultant distortion of the surface area. Recently invented methods of this type that measure surface distortions by natural capillary waves (Sauer et al., 1987), or by measuring small surface distortions from electromagnetic forcing (Ito et al., 1990), are finding considerable application for polymer solutions, but as yet have not been applied effectively to bulk polymers. For solid polymers, procedures based on the reversible creation of surface by application of force, which work well for liquids, are virtually impossible to implement. For example, cross-linked polymers would resist surface deformation by (frequently irreversible) mechanisms that have nothing to do with surface energy (e.g. viscoelasticity). Therefore, several categories of indirect methods have been employed for solids. Most common among these are the extrapolation methods and the wetting methods. More recently, determination of surface energies of solid polymers by direct measurement of the molecular level contact adhesion has been explored (Merrill et al., 1991). 14.3.2.1 Polymer-Liquid Surface Tension The most versatile, and accurate if done with proper precaution, methods for measuring surface tension of polymer liquids are drop (or bubble) profile methods that are based on the principle that the equilibrium shape of a drop is governed by a balance between the surface-tension forces, acting to minimize surface area, and the effects of gravity, which distort and expand the surface area. Drops are termed either sessile or pendant, depending on whether gravity acts to pull the drop towards or away from a supporting surface, respectively. The
14.3 Surface Tension of Polymeric Materials
force balance for a pendant drop is described by the Bashforth-Adams equations (Bashforth and Adams, 1892; Hartland and Hartley, 1976; Huh and Reed, 1983): dfl _ 2 | gzAp ds ~ R 7 dx -r- = cos o9 ds dz
= sin 6
sin Q x
(14-7 a) (14-7b) (14-7c)
where 0 is the angle measured between the tangent to the profile at the drop apex and the tangent at Cartesian coordinates x and z along the two-dimensional projection of the drop profile (assumed in three dimensions to be axisymmetric), R is the radius of curvature at the apex, g is the acceleration due to gravity (for gravity along the z-direction), Ap is the density difference between the drop and its surroundings, and s is an arclength coordinate, measured from the apex along the projected drop profile. Boundary conditions on the three unknowns, 0, x, and z are that they are equal to zero when s is equal to zero. Surface tension is determined by comparison of a digitized image of the drop profile with the shape of the solution to the set of Eqs. (14-7a, b,c), varying / a s a parameter until good correspondence is achieved (Anastasiadis et al., 1988). Advancements in image-analysis methods are improving the accuracy and convenience of these methods. Other drop (or bubble) configurations can be used, with the physics, if not the details, of the analyses being identical to those described for the pendant-drop method. The choice made among the other drop configurations depends to a large extent on the materials involved (i.e. their physical properties, such as densities); this is a particularly important issue when the methods are used to measure interfacial tension between two polymers and the density differences are small.
669
Another category of drop-profile methods, operating on different physical principles, is that involving spinning- or rotatingdrop methods (Princen et al., 1967). A liquid drop or air (or gas) bubble is suspended in a horizontal cylindrical tube. If one is interested in determining the tension between a polymer material and air, this is done by placing an air bubble in the midst of the polymer fluid. Rotation of the tube at angular frequency (O causes the lower-density material to lie along the axis of rotation and to deform axially to some equilibrium length L under the action of surface tension, buoyancy and centrifugal forces. The surface energy is calculated from (14-8)
AC
where C is determined from 4(Cr 3
3
r
l)
(14-9)
with r being the initial radius of the drop. All drop-profile methods require, and are therefore plagued in their execution by the need for, insurance that the drop configurations being measured actually represent equilibrium conditions. This can necessitate long measurement times with the concomitant need to insure stable, controlled experimental conditions of temperature, humidity, etc., over this period. The other principal way to measure surface tension by physical distortion of the surface is to insert an object, such as a plate or a ring (Wilhelmy, 1863; Dettre and Johnson, 1966) through the surface, and to use this as the means of applying a known force of distortion (usually by lifting the surface against gravity). When a solid plate (in practice, glass or platinum is frequently used) is immersed vertically in a polymer
670
14 Polymer Surfaces and Interfaces with Other Materials
fluid, that is, when it is inserted across the interface between polymer and air (or other controlled environment), if the polymer wets the solid preferentially to air, the polymer fluid will be entrained on the plate, producing a resultant force in the direction of gravity, given by F = pycosO-
pgAd
(14-10)
where p, A and d are the perimeter, crosssectional area and immersion depth, respectively, of the solid plate, and 9 is the contact angle of the wetting polymer fluid on the plate. Surface tension can be determined from Eq. (14-10) by measuring the force on the plate as a function of the immersion depth (Dettre and Johnson, 1966). Insurance of equilibrium is an important consideration here, as well. 14.3.2.2 Polymer-Solid Surface Energy Indirect methods have been applied most often to the determination of polymer-solid surface energies. Extrapolation methods are discussed in detail by Wu (1982); these involve measuring the surface tensions of chemically identical polymers under conditions other than the one of immediate interest, such that the polymer of interest is liquid under those other conditions. The extrapolation involved is the systematic variation of the other conditions back toward the conditions of interest. For bulk polymers, two such approaches have been used and give reasonably consistent results where they have been examined. Variations of both molecular weight and of temperature have been used to create polymer samples in the liquid state; extrapolation to the solid-state condition of interest then proceeds by measuring surface-tensions as temperature decreases or as molecular weight increases. The liquid surface tension methods discussed above are
employed (Wu, 1982). Concentration extrapolation, that is, dissolving a polymer solid in a solvent and measuring the surface energy as a function of solvent content, and extrapolating to bulk, is not advisable since, in binary mixtures where one component can segregate preferentially to the surface, extrapolation is very unreliable. Extrapolation through a glass transition temperature can be carried out reasonably accurately if the slope of the extrapolating line is changed at Tg by the ratio of the thermal-expansion coefficients of the liquid and the glass (Wu, 1974). Some results concerning extrapolation methods will be described briefly in subsequent sections; it is appropriate here, however, to mention some general considerations which affect their utility. Chief among these considerations is that any extrapolation from the liquid state cannot capture the effects ofcrystallinity on the surface energy of the solid (Schonhorn, 1968a, b). Many polymers are at least partially crystalline; furthermore, the state of polymer crystallinity at a surface may be different from that in the bulk, owing both to possible effects of the surface in contact with which the polymer solidified, and to the effects of some of the surface-orientation phenomena (discussed earlier) on crystallization near surfaces. The density of a material is known to be an important factor in determining surface tension [see Eq. (14-15] below] and, for example, the density of crystalline polyethylene is 17% greater than that of amorphous polyethyl-ene at 20°C. These comments on the effects of crystalline order on surface tension also apply, in a modified form, to other microstructured polymer materials. Estimation of the surface tension of an ordered block copolymer material, obtained from measurements made in a disordered liquid state, would be inaccurate, because both sur-
14.3 Surface Tension of Polymeric Materials
face ordering and segregation under the conditions of interest could not be adequately comprehended by an extrapolation from high temperatures. An additional consideration in the use of extrapolation methods is a knowledge of the form of the variation of surface energy with the parameter that is being varied. Whereas there is reasonable theoretical and experimental justification for the idea (see next section) that surface energy varies as the reciprocal of the temperature, the form of the molecularweight dependence of polymer-liquid surface tension is more poorly understood, particularly from a theoretical point of view. This uncertainty makes extrapolation potentially unreliable. Wetting methods are the other major category of indirect routes to the determination of the surface energies of solid polymers. These involve the measurement of the contact angle 0 for sessile drops of various liquids in contact with the surface. Fox and Zisman pioneered this approach and conceived the most widely practiced method of this type, which has come to be known as the measurement of the critical surface tension (Fox and Zisman, 1950). This quantity, yc, is determined by measuring the equilibrium contact angles of a series of liquids of different liquid-vapor surface tensions, 7LV, and extrapolating to the condition of complete spreading, i.e. when 6=0: LV
(14-11)
Young's equation, i. e. ^ v cos 0 = y s v + ySL, relates the liquid-vapor surface tension of the sessile drop to the surface energy of the solid and the solid-liquid interfacial energy via the contact angle 6. Using Young's equation in the limit of zero contact angle gives 7c = 7 - 7 s L - ^ e
(14-12)
where KQ is known as the spreading pressure and is equal to the difference between
671
the solid-vapor surface energy, y sv , and the solid surface energy in vacuum, y. Thus, yc will be a good estimate of /when ySL and 7iQ are small. The former can be minimized by using liquids of polarities similar to that of the solid; the latter becomes larger as the liquid approaches complete wetting of the solid (the implication here is the rather paradoxical conclusion that one should not extrapolate from too close to zero contact angle). This wetting method is a particularly convenient and simple technique to estimate the surface energy of a solid, but is most usefully and appropriately regarded only as an estimate. Direct comparison with the other methods, for example the more rigorous liquid-state methods discussed above, shows that yc is almost always of an order 10% lower than the values of /measured by other methods, which is understandable when based on the considerations embodied in Eq. (14-12). Difficulties in selecting an appropriate series of liquids, and in reliably extrapolating to zero contact angle, beset this method. The best series of liquids are those that are similar in polarity, both to one another and also to the solid in question. Wu (1979) has proposed an alternative, but related, wetting method, known as the "equation of state" method, which mini-mizes the inaccuracies caused by some of these effects. Questions of how the liquid interacts with the solid, i.e. in ways other than through the surface energetics, also need to be recognized when using these techniques. A sessile drop will exert a mechanical tension normal to the solid and possibly induce molecular rearrangement. Furthermore, while immiscibility with the solid is an essential characteristic of the liquid that is to be used, it is sometimes difficult to insure that absolutely no liquid permeates the solid.
672
14 Polymer Surfaces and Interfaces with Other Materials
Recently, progress has been made (see Merrill et al., 1991; Chen et al., 1991; Chaudhury and Whitesides, 1991; Horn and Smith, 1992) in determining the surface (or interfacial) energies of solids by measuring the contact-adhesion forces and shapes of surfaces of identical (or different) solids brought into well-defined, intimate, molecular contact. These methods are based on the Johnson, Kendall, Roberts (JKR) theory (Johnson et al., 1971) of solids in contact with each other, which is an approximate theory in that it assumes that the intermolecular interactions giving rise to the contact adhesion are very shortranged and so act only through those portions of the surfaces in contact (and not through those elements of the surface that are close to one another but not in contact). While this assumption is not defensible in detail, the testable predictions of the JKR theory have been examined thoroughly (Horn and Israelachvili, 1987; Merrill et al., 1991) and appear to be reasonably quantitatively reliable. From the point of view of determination of surface energies of solids, the core results of the JKR theory for contact between two elastic spheres of radii R{ and R2 [where R =RlR2(Ri +^2)] and identical properties, including bulk elastic modulus K, and surface energy, 7 under the action of an external force F, are those for the radius of the contact area between the two solids, a: 3 a
= ±
Table 14-1. Comparison of surface-energy values for PET determined by different methods.
(14-13) and for the (tensile) force required to separate the two solid surfaces: = 3nRy
termination of /by two independent means. In the first method, a is measured as a function of the applied load F, and 7 is determined by fitting the data to Eq. (14-13). The first term is for nonadhesive elastic spheres, whereas the other three terms embody the additional contributions of surface energy to the deformation of the solids. Chaudhury and Whitesides (1991) have examined the validity of Eq. (14-13) for the case of crosslinked spheres of poly(dimethylsiloxane) (PDMS), with a microscope apparatus enabling the accurate measurement of a(F), and found it gave an excellent fit to their data plus a good value of the surface energy (y= 22 mJ/m2, well-known from earlier measurements) (see Wu, 1982). Merrill et al. (1991) have used the surface force apparatus techniques, developed by Israelachvili (1991), to investigate the contact between two identical surfaces of polyethylene terephthalate) (PET). This method involves making thin, smooth sheets of the PET and using the optical interferometry of the method to determine the magnitude of the "jump apart" from adhesive contact, which is directly related to Fs. They, too, found that the JKR theory gave an excellent account of their observations. Measurement of the force to separate [Eq. (14-14)] gave a value of 7 of 61 ± 2 mJ/m2. Table 14-1 compares this with
(14-14)
Measurement of these two quantities (for known values of K and R) enables the de-
7 (mJ/m2) 43 44 45 61 a d
Method
Ref.
Critical surface tension Wu's "equation of state" Melt extrapolation Surface-force measurement
[a] [b] [c] [d]
FoxandZisman(1950); b Wu(1979); c Wu(1982); Merrill et al. (1991).
673
14.3 Surface Tension of Polymeric Materials
values of 7 for PET determined by other methods. The difference between the surfaceforce-determined value and those obtained from the other methods raises some interesting issues. It is possible that this higher value is the most accurate, given the tendencies of wetting methods to underestimate surface energies, as described by Eq. (14-12), and of the inability of extrapolation from the melt to capture influences of the crystallinity in PET on 7. However, there are subtle effects to consider in the application of surface-force methods. The force to separate the surfaces may be elevated by dissipative effects arising from surface rearrangement and/or interpenetration on contact, possibilities it shares with wetting methods. Chen et al. (1991) and Chaudhury and Whitesides (1991) have seen, from examination of Eq. (14-13), that there is an hysteresis effect for some surfaces, i.e. a different contact radius is seen for the same applied force, depending on whether the applied force is in the process of being increased or decreased. They attribute these effects to dynamic rearrangement processes occurring in the zone of contact between the two surfaces.
14.3.3 Experimental Results for the Surface Tension of Polymeric Materials Comprehensive compilations of experimental data for the surface tension of polymer materials are available (see Wu, 1989). Typical experimentally measured surfacetension values for a variety of commercially important polymers are listed in Table 142. Surface tensions for polymers generally range from 20 to 50 mJ/m2; these values are larger than their corresponding small-molecule equivalents. This is in part due to the increased density of the higher molecularweight materials, as surface tension is a sensitive function of the density of the material. Group contribution techniques have been applied for the approximation of surface-tension values, the basis of these being the empirical Macleod relation (Macleod, 1923) connecting surface tension with density r= 7o pn
(14-15)
where 7 is the surface tension, p is the density and 70 and n are constants. Values for the Macleod exponent, n, are generally between 3 and 4, and several of these experimentally measured values are also listed in
Table 14-2. Surface tensions and their temperature coefficients for some common polymers (adapted from Wu, 1989) Polymer Poly(propylene) (atactic) Poly(styrene) (molecular weight, 9300) Poly(tetrafluoroethylene) Poly(methyl acrylate) (molecular weight, 25000) Poly(methyl methacrylate) Poly(ethylene terephthalate) Polycarbonate (bisphenol A) Nylon 66 (molecular weight, 17000) Poly(dimethyl siloxane)
y(mJ/m2)
- d//dr (mJ/(m2K))
Macleod's exponent
29 39 24 41 41 45 43 47 21
0.056 0.065 0.058 0.070 0.076 0.065 0.060 0.065 0.059
3.2 4.0 4.2 3.5
674
14 Polymer Surfaces and Interfaces with Other Materials
Table 14-2. In turn, the constant y0 can be approximated by (14-16) where the quantity P is the combined "parachor" for a repeat unit and M is the repeat unit molecular weight. Extensive lists of the parachors for various atoms and chemical groups are available (Quayle, 1953; Van Krevelen, 1976; Wu, 1982). Surface tension values, calculated with this group-contribution technique, are often in reasonable agreement with experimental values. In general, the surface tension is dependent on temperature, molecular weight and chemical composition. The functional form of these dependencies can be anticipated, in part, by the semi-empirical Macleod equation [Eq. (14-15)]. The chemical composition affects both the constant y0, as described above, and is also a factor in determining the bulk density of the material. However, for a given polymer, the effect of molecular weight and temperature on the surface tension parallel the effects of these variables on the bulk density p. Although detailed predictions for the surface tension of polymers are available (see later), certain semi-empirical relations for the molecular weight and temperature dependence are useful. A semi-empirical expression which adequately describes the experimental molecular weight dependence of surface tension is given by (LeGrandandGaines, 1969, 1973): k M 2/3
(14-17)
where Mn is the number-average molecular weight. Polymers increase somewhat in
density as the average molecular weight increases; this effect has been elucidated by equation-of-state treatments of compressible polymers (Sanchez and Lacombe, 1976, 1977, 1978; Sanchez, 1983 a). By the use of Eq. (14-15), the increase in surface tension with molec-ular weight can be interpreted in terms of the enhanced density of the bulk polymer. Equation (14-17) is most appropriate for lower molecular weight polymers (Wu, 1982), and usually, the magnitude of the constant £e in this equation is such that the surface tension y reaches, within ex-perimental error the limiting value y,, at fairly low molecular weights, of the order of 104, over the same range where the bulk density is also becoming molecular-weight independent. For higher-molecular-weight polymers, the surface tension is rather insensitive to molecular weight. Temperature is also a key variable affecting the surface tension of pure-component polymers. Experimentally it is observed that, to a high degree of approximation, the surface tension decreases linearly with increasing temperature. y=7o-kTT
(14-18)
Some experimentally measured surfacetension values, as a function of temperature, are shown in Figs. 14-7 and 14-8 for a variety of common polymers. Polymeric materials generally have a positive thermal-expansion coefficient. Similar to the effect of molecular weight, the decrease in density with increasing temperature, produces a corresponding decrease in the surface tension of the material [see Eq. (14-15)]. The constant, kT- -dy/dr, varies between 0.05 and 0.07 mJ/(m2 K) for a wide variety of polymers: several values for this temperature coefficient are also listed in Table 14-2. The temperature coefficient for small molecules is generally larger [~ 0.1 mJ/(m2
14.3 Surface Tension of Polymeric Materials
50
200 100 150 Temperature (°C) Figure 14-7. Surface tension as a function of temperature for some polymer melts: PEO, polyethylene oxide); L-PE, linear poly(ethylene); B-PE, branched poly(ethylene); PP, poly(propylene) (Wu, 1992).
675
high-density polymers (Koberstein, 1987). Another contribution is the nonlocal entropy effect. The decrease in configurational entropy of the "flattened" polymer molecules in the near-surface region produces a positive contribution to the temperature coefficient (Silberberg, 1988). However, due to the sensitivity of the surface tension to density, as indicated by Eq. (14-15), the decrease in density with increasing temperature is the dominant effect. The net result is a decrease in the surface tension at elevated temperatures. The details of the origin of this will become clearer in the following section.
14.3.4 Theoretical Aspects of Polymer Surface Tension
100 150 Temperature
200
Figure 14-8. Surface tension as a function of temperature for some polymer melts: PCP, poly(chloroprene); PMMA, poly(methyl methacrylate); PVAc, poly(vinyl acetate); PIB, poly(isobutylene); PnBMA, polyO-butyl methacrylate); PDMS, poly(dimethyl siloxane) (Wu, 1982).
K)] (Padday, 1969). Several factors contribute to the smaller temperature coefficients found for polymers, with perhaps the most important of these being the lower thermalexpansion coefficient associated with
Before describing some theoretical approaches to predicting the surface tensions of polymers in detail, it is useful to obtain some approximate conceptual ideas of how certain of the relevant factors come into play. We will follow closely the line of argument developed by Helfand and Wasserman (1982). For this purpose, we first need a model for the free energy of the material whose surface tension we wish to understand. For this purpose here, and in the subsequent discussion, we will adopt essentially the representation of Sanchez and Lacombe (1978) (though not their detailed lattice-fluid model, which is much more careful and intricate). This approach models a bulk polymer as a mixture of polymer segments and holes in the Flory-Huggins manner (Sanchez, 1983 a). A lattice, of site-size b, is used to compute the free energy of mixing the connected segments and the holes which represent the density, or more exactly, the compressibility due to the free volume of the material. The Flory-Huggins free energy of mixing per unit volume, for a vol-
676
14 Polymer Surfaces and Interfaces with Other Materials
ume fraction of segments 0, with a fraction (l-O) of holes, is (14-19) -
where N is the degree of polymerization of the polymer; AUis the total cohesive contact energy per segment and is (one half of) the increase in internal energy when a segment of size b3 is moved out of the bulk at 0 = 1 to 0= 0 (Silberberg, 1988). AU/kT could be thought of as a Flory-Huggins, ^-parameter for the mixing of the two components, holes and segments. We will use this notation part of the time in this discussion since it enables the discussion of surface tension to be made in a virtually identical way to that of the interfacial tension between two immiscible polymers with interaction parameter x that form an incompressible interface (Helfand, 1975). Consider (following Helfand and Wasserman, 1982) a surface with a sharp density profile, such as the curve with T* = 1.5 shown in Fig. 14-1. Assume that the polymer-melt segment number density falls abruptly from p to 0 over a short distance. Consider further a section of a polymer chain of length Z segments lying in this surface region. A first question to discuss is that concerning the typical length of such a sequence of segments broaching the surface region. A segment entering the surface region will pay an energy cost of order X^T; thus, a sequence of Z segments will suffer an energy penalty of ZxkT. The energy required to pay this penalty to drive the chain segments into the surface region of a melt can come from but one source at equilibrium, namely, thermal fluctuations of
energy kT. The consequence of this is that ZxkT^kT, resulting in Zoc x~l ~ ALT1. If the configuration of the sequence of Z chain segments is roughly that of a random walk [see Eq. (14-2)], then it will span a characteristic spatial dimension of the order bZl/2. This length characterizes the thickness of the surface region, D, which from these considerations must vary as D ^ bx~l/2 ^ AC/""2. If we interpret x a s the energetic cost of introducing holes into, or lowering the density of, the surface region, then we conclude that the surface-region thickness should vary as D °c K~1/2, that is, inversely with the square root of the compressibility of the polymer. For the interface between immiscible polymers, it is the ^-parameter that controls the interfacial thickness (D °c X~1/2)- In order to estimate the surface tension, we observe that the density variation in the surface region of the melt occurs in a volume of DS, where S is the surface area of the material. In this region, there are pDS segments, each with a typical energy of the order of ^fcT, so that the total energy of the region, calculated as the energy of mixing of holes and segments, is
yS=pDSxkT
(14-20a)
or
y=pbxuzkT
(14-20b)
Following the line of argument above, these equations suggest an inverse relationship between surface tension and the square root of the compressibility, or likewise, a proportionality to the square root of the cohesive contact energy between segments. This connection has been demonstrated in an extensive examination of experimental data, and via more detailed theory, by Sanchez (1983b). The serious limitation of this style of argument is its fundamental inconsistency. A
14.3 Surface Tension of Polymeric Materials
sharp interface is assumed, but random-coilsize dimensions of the surface region are found. A more consistent, detailed theory of surfaces and interfaces is necessary and is available (Helfand, 1980; Helfand and Wasserman, 1982; Poser and Sanchez, 1979). There is more than one way to approach the development of this theory but the essential physics is similar. The polymer molecule lays itself out in the surface region in a manner that minimizes the free energy, with the latter containing contributions from the effects of the surface on the configurations of the macromolecules as well as interactions with other molecules in the surface region. Here, we discuss the basic approach, based on the Cahn-Hilliard theory (Cahn and Hilliard, 1958) using the free energy model of Eq. (14-19). The alternative SCMF theory has been worked out only in the limit of infinite molecular weight (Helfand and Wasserman, 1982; Anastasiadis et al., 1988). The detailed theories of surfaces and interfaces (Poser and Sanchez, 1979; Silberberg, 1988) begin by a more detailed calculation of the density profile of segments in the surface region than that given in the simplified argument above. The dependence of 0 on the position JC, normal to the surface in the surface region, is determined by minimizing the Cahn-Hilliard free-energy integral: ^
ck
(14-21)
where A F ^ = AFM(
677
and Sanchez, 1979) have treated K as a parameter to be derived by fitting to experimental data. AFO is the surface free energy per segment. The two terms in Eq. (14-21) represent, first, the free energy of the material, were it homogeneous at the local density, and second, the contribution to the free energy of the surface region due to the fact that the density is varying. This form of the theory has come to be known as "square-gradient theory" and has been a very successful approximation (Davis and Scriven, 1982), despite the fact that one could have imagined that, especially for very sharp interfaces, higher-order gradients of the density profile would have to be considered. Surface tension in this theory is determined by minimizing the free energy of Eq. (14-21) with respect to variation in the profile 0{x) of density of segments through the surface region. This minimization yields AF
- K
(14-22)
dx
so that the surface tension derived from the surface free energy of Eq. (14-21) can be written, with a change of variables, as 7 = 2 J [KAFaM(0)]V
d0
(14-23)
Equations (14-22) and (14-23) provide the means to calculate surface (or interface) density profiles, and surface (or interfacial) tension theoretically. The former can be tested against simulations such as those of Fig. 14-1, or against the growing body of recent experimental results, to be discussed subsequently. The theory is versatile in that different models for the free-energy expression can be employed. For bulk polymer melts at most temperatures of interest, the widths of the surface
678
14 Polymer Surfaces and Interfaces with Other Materials
profiles predicted by square gradient theory are very narrow, of the same order as the segment size b. This is entirely consistent with the conclusions drawn from the simulations discussed earlier. More important are the comparisons with experimental data for surface tension, and its temperature coefficient, which have been explored most extensively by Silberberg (1988) and by Poser and Sanchez (1979). Figure 14-9 shows a comparison of experimental data and theory for several polymers, presented in terms of reduced variables, with K constant at AU/2b. Clearly the agreement is good in several respects; the magnitude of /could be matched nearly perfectly by adjusting K to 0.55 AU/b (Poser and Sanchez, 1979). The negative temperature dependence of y is also captured reasonably accurately. The origin of this is physically illuminating. Several authors (Hong and Noolandi, 1981; Rabin, 1984; Silberberg, 1988) have shown that there is an additional contribution to the surface tension, not included in Eq. (14-23). This results from the nonlocal effects of chain connectivity near the surface; in other words, the effect, captured in simulations but not in square-gradient theory, that
connected segments, by attempting to pack into constrained surface regions, propagate the interactions to distant regions along the chain. Rabin (1984) showed that this could be incorporated into Eq. (14-23) by modifying K to be AU/(2b) + kT/(24®b). This modification adds a positive contribution to the temperature coefficient of the surface tension. Therefore, whereas it may be significant quantitatively, it is not the dominant effect in determining the temperature dependence of polymer surface tensions. This positive effect comes from the fact that, as temperature increases, the constraints on the chain configurations are increasingly important, thus raising the surface free energy with temperature. The dominant effect comes from the fact that, as temperature increases, more holes are introduced into the surface region, so reducing its density. The density profile of the surface broadens, as seen in Fig. 14-1, and this creates increased freedom for the arrangement of segments in the narrow constrained surface region as temperature increases. Sanchez further refined the generalized gradient approach to polymer surfaces and derived a universal relationship between
Figure 14-9. Reduced surface tension y [= yb2/(AU)] as a function of the reduced temperature f [= £77(At/)] shown for several polymers: experimental data compiled by Poser and Sanchez (1979), and compared with theoretical results obtained from a model based on Eq. (14-23), developed by Silberberg (1988). The points are data for the different polymers; the solid line is the theory.
Qt 0.6
0.7
0.8
0.9
Reduced Temperature,!"
1.0
14.4 Techniques to Examine Polymer Surface Properties
surface tension y, bulk density p and isothermal compressibility K (Sanchez, 1983b):
14.4 Techniques to Examine Polymer Surface Properties 14.4.1 General Considerations
1/2
^{K
679
(14-24)
The surface tension is predicted to vary inversely with the square root of compressibility, as we anticipated earlier in this section when based on a qualitative argument (Helfand and Wasserman, 1982). The constant Ao is independent of both temperature and molecular weight, and its value is invariant for certain classes of materials. Figure 14-10 illustrates the universal behavior in experimental data suggested by Eq. (14-24) using surface-tension values for PDMS determined by Bhatia et al. (1985) at different molecular weights and temperatures. For most polymers, as well as other organic liquids, the constant AQ2 varies only slightly between 2 - 3 x 1CT4 (erg cm2/g)1/2. Equation (14-24) is a simple and accurate correlation for estimating the surface tension of polymers.
if)
(p/K-) 1/2 x10- 8 (Nkg/m^)
Figure 14-10. Experimental test of the semi-empirical correlation [Eq. (14-24)] proposed by Sanchez (1983 b) using the data obtained by Bhatia et al. (1985) for PDMS of molecular weights 75000 (O) and 3900
The battery of experimental methods now used to study the characteristics of polymer surfaces, in addition to surface energy, has expanded rapidly in recent years. There is, of course, an intimate relationship between the molecular structure and properties of a surface region and the fundamental thermodynamics discussed earlier. The effect on the surface tension of molecular ordering and surface enrichment in multicomponent materials has been recognized (Wu, 1982, 1989; Bhatia et al., 1985). Surface texture, or roughness, is another important aspect of surface structure which may be probed directly. Consistent with the aims of this chapter, the focus is on those techniques suitable for the study of the external surfaces of polymers, or their interfaces with other materials. For the additional, and interesting, subject of buried interfaces in polymers, the reader is referred to the recent review by Stamm (1992). Objectives in the characterization of polymer surface properties include measurement of roughness, mass density and composition profiles, information on the configuration of the molecules in the surface region, and where appropriate, information on the mechanical, or other material properties. There are two length scales associated with polymer surfaces, which make them particularly challenging experimentally. Certain properties, such as mass density and roughness may vary over the size of a repeat unit; other properties such as chain configuration and compositional variations, may vary over length scales associated with the polymer molecule taken as a whole. In many cases, several different techniques will be required to cover the entire polymer sur-
680
14 Polymer Surfaces and Interfaces with Other Materials
face region. Several excellent reviews of techniques are available for both polymer surfaces and interfaces (see Briggs, 1987; Feast and Munro, 1987; Koberstein, 1987; Russell, 1990, 1991; Stamm, 1992). Table 14-3 summarizes many of the surface-sensitive techniques which are used to study polymer surfaces. The various forms of microscopy are listed first and offer the advantange of a direct image of surface structure and composition. On the other hand, X-ray and neutron reflection yield Fourier- space information on surface structure with very high resolution. Various surface-sensitive spectroscopic techniques can give detailed information on average chemical composition and composition profiles. In some cases, the non-conducting or radiation-sensitive nature of organic polymers requires special precautions. On the other hand, the ability to synthesize polymer molecules bearing isotopic or chemical labels can offer great advantages. Surface sensitivity for many of the techniques is derived from either the limited penetration depth of an incident probe beam, or the stimulated emission of information-carrying particles having very small escape depths. Advances in all these techniques has pushed the resolution into the angstrom range. However, to achieve such resolution, sample quality becomes extremely important, and often some theoretical underpinning is required for complete interpretation of the data. 14.4.2 Imaging Methods Several techniques are available which can produce images of polymer surfaces and enable one to characterize the surface structure at a molecular level. The most conventional, but nevertheless very powerful, imaging technique is scanning electron microscopy (SEM) (Goldstein et al., 1981;
White and Thomas, 1984; Sujata and Jennings, 1991). Nanometer focusing of the incident electron beam offers very high resolution for mapping surface topography. In addition, surface chemical analysis with lateral resolution in the millimeter range may be obtained through energy analysis of Auger electrons (Auger spectroscopy) and X-ray fluorescence (Goldstein et al., 1981). One must take care to avoid specimen damage at higher accelerating voltages, and samples are often coated with small amounts of a heavy element (e.g. platinum) to improve resolution. Environmental scanning electron microscopes that can image at high pressure (20 Torr) offer the exciting possibility of directly studying the effect of changing ambient conditions on polymer-surface structure (Sujata and Jennings, 1991). Transmission electron microscopy (TEM) has also been applied to the study of polymer thin films and surfaces (Sujata and Jennings, 1991; Hasegawa and Hashimoto, 1985a, b; Henkee et al., 1988; Ishizu and Fukuyama, 1989). Both chemical staining and microtoming are required sample preparation for TEM. The technique can offer images of two-dimensional composition profiles with nanometer resolution (Russell, 1991). Atomic resolution under environmental conditions can be achieved with scanning tunneling microscopy (STM) (Rabe, 1989; McMaster et al., 1991; Rabe and Buchholz, 1991). In this method a conducting point probe is rastered across the surface (typically) at a constant tunneling current to give a direct three-dimensional image of the surface topography. Conductivity requirements limit the experimental studies to isolated molecules or thin monolayers on conductive substrates such as graphite or gold. Atomic fofce microscopy (AFM) can image non-conductive polymer surfaces by measuring the force between a fine tip and
14.4 Techniques to Examine Polymer Surface Properties
681
Table 14-3. Techniques for polymer surface analysis. Technique
Acronym Sampling Depth Lateral depth resolution resolution
Scanning electron microscopy
SEM
10 nm
Transmission electron TEM microscopy Scanning tunneling STM microscopy
10 nm
Information content
Comments
5 nm
Direct image of surface topography
3 nm
Two-dimensional profile Molecular imaging and surface topography Molecular imaging and surface topography Composition profiles and roughness Composition profiles
Auger and X-ray mode can give chemical composition Special sample preparation required Conductive substrate required
0.1 nm
0 nm
AFM
0.1 nm
0.5 nm
X-ray reflectometry
XR
1 nm
-
Neutron reflectometry
NR
1 nm
-
Ellipsometry
ELLI VASE
1 nm
-
X-ray photoemission spectroscopy
XPS ESCA
10 nm
0.1 nm
10 Jim
HREELS
1 mm
1 nm
1 jam
Surface composition, vibrational spectra
1 urn
Surface composition
-
Composition profile
Atomic-force microscopy
High-resolution electron-energyloss spectroscopy
Static secondary-ion SSIMS mass spectrometry Dynamic secondary- DSIMS ion mass spectroscopy Forward-recoil FRES spectrometry Rutherford backRBS scattering spectrometry Nuclear-reaction NRA analysis
1 nm
Infra red attenuated total-reflection spectroscopy
IR-ATR
|um
Infra red grazingincidence-reflection spectroscopy
IR-GIR
Film thickness, refractive index profile Surface composition, composition profiles
Interpretation requires model predictions Interpretation requires model predictions Limited contrast Radiation damage and sample charging are problems Radiation damage and charging are problems Complex spectra Etch rate is composition dependent
°°
13 nm
Jim
80 nm
Composition profile
|Ltm
30 nm
Composition profile of marker
Requires heavy elements
jixm
13 nm
Hydrogen and deuterium composition profile Vibrational spectra of polymer surface adjacent to ATR-crystal surface Vibrational spectra of thin films and adsorbed layers
Resolution is depth dependent
682
14 Polymer Surfaces and Interfaces with Other Materials
the surface (Hansma et al., 1988; Meyer et al., 1990; WittmanetaL, 1991; Collinetal., 1992). At present, AFM cannot match the resolution of STM entirely; however, its ability to image dielectric surfaces makes it an invaluable tool in studying polymer surface topography and measurement of surface physical properties. A complementary imaging technique is phase-measurement interference microscopy (PMIM) (Biegen and Smythe, 1988; White et al., 1990; Smith et al., 1991; Stamm, 1992). Here, the interference pattern from a laser beam reflected from the sample surface and a piezoelectrically driven reference surface is monitored with an area detector. The interference pattern can then be related to the thickness and/or topology of the sample film. Although the lateral resolution for PMIM is of the order of microns or greater, height-contour maps with a vertical resolution of 0.6 nm are possible. 14.4.3 Reflection Methods Two related reflection methods which can offer excellent depth resolution in the study of polymer surface structures are X-ray reflectometry (XR) and neutron reflectometry (NR) (Heavens, 1976; Born and Wolf, 1980; Lekner, 1987; Russell, 1990, 1991). Both of these measure the scattered intensity of an incident beam at grazing angle to the polymer surface as a function of the wave vector q = (27t/A)sin0 where A and 0 are the wavelength and incident angle, respectively. For XR, q is scanned by measuring intensity as a function of the angle 6, whereas both X and 0 may be varied in NR, in ways that depend on the neutron sources (Russell, 1990, 1991). Contrast for neutron reflectometry is provided by the scattering-length densities of the constituent nuclei (Felcher et al., 1987).
Deuterons are considerably more efficient scatterers of neutrons than protons, making NR a sensitive technique for evaluating multicomponent systems where one component has been selectively deuterated. On the other hand, X-ray reflectometry is sensitive to differences in electron density, and is amenable to the study of multicomponent systems where there is a difference in electronic structure between the components (Bilderback, 1981). Reflectivities can be measured to as low as 10~6 for neutrons and 10"10 for X-rays (Russell, 1990). Both XR and NR, particularly the former, can give accurate measures of surface roughness (Braslaw et al., 1985; Foster et al., 1990). Although the lateral sampling area for both these techniques is large, the ideal Fresnel reflectivity curve, which decreases as q~4 for an ideal smooth surface, is damped exponentially by the effect of surface roughness (Russell, 1990; Stamm, 1992). The equations relating the measured intensity profiles for both XR and NR to the composition gradients within the sample are very similar, and highly non-linear. As a result, measured intensity or reflectivity profiles cannot be uniquely and directly inverted to obtain composition profiles. Instead, experimental reflectivity profiles are compared with calculated profiles based on an assumed structure. Iterative comparison leads to a best-fit profile, but uniqueness of such a profile cannot be proven. Ellipsometry (ELLI) is a light-reflection technique which has been used to study thin film and surface behavior of polymers. In all ellipsometric measurements, it is the change in polarization state of polarized light after passing through a sample which is studied, either by reflection or transmission (Azzam and Bashara, 1977; Debe, 1987). The use of polarization distinguishes ELLI from XR and NR and confers on it high depth resolution, despite the long-
14.4 Techniques to Examine Polymer Surface Properties
wavelength radiation employed. The polarization state is characterized by the measurement of the amplitude ratio, tan % and the phase difference, A, for the two Cartesian components of the electric field. Thus, at fixed angle and wavelength, two pieces of information characterizing the sample are obtained, and are usually an average thickness and refractive index for an assumed uniform film. The ability to extract much more detailed information has come with the development of variableangle spectroscopic ellipsometry (VASE) (Kimetal., 1989; Arwin and Aspnes, 1986). VASE, at multiple angles, produces additional independent data on the sample and so is particularly useful in examining multilayer surface regions. As with XR and NR, direct inversion of the measured profiles to obtain the spatial variation of optical constants is usually not possible, and model predictions are required to interpret the data. 14.4.4 Spectroscopic Methods Spectroscopic methods of most use for polymer surfaces are either electronic or vibrational in character (Garbassi and Occhiello, 1987). The electron spectroscopic techniques all require high vacuum and so the surface structure to be examined must survive this state. A common-high vacuum surface analytical technique which can usefully measure surface composition and composition profiles at polymer surfaces is X-ray photoemission spectroscopy (XPS), which is also commonly known as electron spectroscopy for chemical analysis (ESCA) (Clark et al., 1973; Clark et al., 1974). In XPS, the polymer surface is irradiated with X-rays from an anode or synchrotron source. The X-rays produce photoelectrons within the sample (the photoelectric effect) which are collected at a take-off angle (6)
683
and the energy is analyzed. Since the kinetic energy of these core-level electrons is directly related to the binding energy, energy analysis of the emitted electrons produces a semi-quantitative spectrum of polymer-surface atomic composition. XPS is not able to detect the presence of extremely light elements, i.e. H and He. XPS is a surface-sensitive tech-nique because the mean free path (for inelastic collision) A, of photoelectrons in a polymer matrix is usually less than 5 nm. Thus, even at normal incidence (6 = 90°), maximum sampling depths are generally not greater than 100 A (Fadley et al., 1974). Furthermore, angle-resolved XPS has depthprofiling capability in that the required escape depth of a photoelectron emitted at depth z for the surface decreases as l/sin#, where 6 is the take-off angle. The measured intensity I*(0) for an electronic core level of a particular nucleus is related to the concentration profile within the sample by (Fadley et al., 1974): (14-25)
7» w (The intensity I*(q) is normalized by the 90° value). The integral prefactor l/sin0, represents the increase in acceptance area with decreasing take-off angle, and should not be included when the entire specimen is being sampled (Fadley et al., 1974). Although faced with an inversion problem similar to XR and NR, when using the above equation, XPS is able to study composition gradients of species which have binding energies which are resolvable in the XPS spectrum. It does, however, require knowledge of the mean free path X. Several values of A have been reported in the literature for a variety of electron kinetic energies and host materials (Clark and Thomas,
684
14 Polymer Surfaces and Interfaces with Other Materials
1977; Bain and Whitesides, 1989; Parsonage et al., 1991). Small-spot sampling techniques have also been used to study lateral variations of surface composition with a resolution of around 10 |im (Kelley, 1991). Some disadvantages of XPS include the inability to probe beyond several tens of angstroms into the sample. Furthermore, angle-resolved XPS measurements require relatively long exposure times, and certain polymers will suffer radiation damage. Charging of the dielectric polymer samples can also cause peak shifts and broadening of spectra. Charging can be partly compensated for by the use of electrostatic flood guns and internal standards. Another high-vacuum technique for surface analysis is high-resolution electronenergy-loss spectroscopy (HREELS) (Wandass and Gardella, 1985; Gardella and Pireaux, 1990; Pireaux et al., 1986, 1988). In addition to XPS-like information on surface composition, HREELS can also probe surface vibrational spectra as well. In this case, the energy loss for a beam of low-energy electrons is analyzed after interacting with the surface. The technique is thus able to probe both vibrational and electronic transitions depending on the electron energy. It is less quantitative than XPS for compositional information, although it does allow for the detection of light elements. Both Raman and IR-active modes are detected in the vibrational spectra. The technique is very surface sensitive, with a sampling depth of about 1 nm (Wandass and Gardella, 1985; Gardella and Pireaux, 1990). However, both charging and radiation damage are problems for polymer samples. Another highly specialized form of vibrational spectroscopy which has found application for polymer surfaces in contact with metals is inelastic electron-tunneling spectroscopy (IETS). In this method, a thin film of polymer is incorporated into a sol-
id/polymer/lead junction and the tunneling current is measured. Inelastic effects in the tunneling through the junction can be related to the vibrational spectrum of the polymer. Colletti et al. (1987) used this method to determine the reactivity and orientation of segments of acrylic polymers in contact with aluminum oxide. There are many other analytical spectroscopic techniques for the study of surfaces and thin films which can be directly applied to polymers under the right circumstances (see, for example, Debe, 1987). Included in these are the various surface and thin-film specific vibrational spectroscopic techniques such as infrared attenuated total reflection spectroscopy (IR-ATR) and infrared grazing-incidence reflection spectroscopy (IR-GIR). IR-ATR measures the vibrational spectrum in a thin surface region of the polymer sample which has been placed in intimate contact with a highrefractive-index material such as silicon, germanium or zinc selenide (Ulman, 1991; Harrick, 1967; Harrick and du Pre, 1966; Haller and Rice, 1970). The technique is not particularly surface sensitive, with a sampling depth comparable to optical wavelengths (> 0.5 jam). It does provide a measure of overall composition in the near-surface (glass-interface) region, and sampling depths can be varied somewhat through the refractive index of the coupling material. IR-GIR measures the vibrational spectrum of a thin film by external reflection(s) from a supporting substrate (Ishitanti et al., 1982). The technique is not in itself surface specific, but it can provide information on the orientation of transition dipoles adjacent to the surface for very thin films and monolayers (Ulman, 1991).
14.4 Techniques to Examine Polymer Surface Properties
14.4.5 Ion-Beam Methods Ion-beam methods require an apparatus for acceleration of the ions, generally operate at high vacuum (as with electron spectroscopic methods), and can be damaging to samples. On the other hand, they have a versatile capacity for probing many aspects of polymer surfaces. An important ion-beam technique for chemical analysis and depth profiling of polymer surfaces is secondary-ion mass spectroscopy (SIMS) (Stamm, 1992; Coulon et al., 1989). In static SIMS (SSIMS), a primary-ion beam (e.g. Ar, Cs, Ga, Xe, etc.) is focused on the surface, where it reacts specifically to produce secondary-ion fragments which are then analyzed by a mass spectrometer. SSIMS spectra for polymers can be very complicated, and chemical identification is accomplished by comparison with spectra of known materials. SSIMS can be very surface sensitive, with a sampling depth of around 1 nm, and in addition, the ability to focus the primary-ion beam allows for two-dimensional studies of chemical composition with a lateral resolution of 1 mm (Kelley, 1991). The basis of dynamic SIMS (DSIMS) is the etching of the sample surface by rastering of a primaryion beam (e.g. 3 keV O2+) across the surface and continuous monitoring of the resulting secondary ions by mass spectroscopy. The SIMS experiment is quite sensitive to the mass difference between the proton and deuteron, and selective deuteration of one component allows for depth profiling of that species (Coulon et al., 1989; Russell et al., 1989). The advantage of DSIMS over angle-resolved XPS is the ability to monitor composition profiles at large depths into the polymer sample (albeit by destructive techniques). However, DSIMS has considerably lower depth resolution than XPS, with a minimum resolution of approximate-
685
ly 125 A. The low resolution is in part due to the necessity of calibrating the etch rate as a function of composition. In addition, problems in obtaining a steady-state etching process demand special preparation of the polymer surfaces (Coulon et al., 1989). Less destructive ion-probe techniques include forward recoil spectroscopy (FRES), Rutherford backscattering (RBS) and nuclear reaction analysis (NRA). These techniques were originally developed for semiconductors but have been transferred to polymer applications with great success. The incident beam for both FRES and RBS consists of helium ions accelerated to a few MeV. (4He is often used to minimize beam damage.) FRES measures the energy spectrum for protons lH and deuterons 2H ejected in the forward direction (Mills et al., 1984; Green and Doyle, 1990). At a fixed acceleration voltage and take-off angle, the energy of the ejected particles is related to its mass and the depth at which it originated within the sample. The ability to differentiate the mass difference of protons from deuterons allows for the study of composition profiles where one component has been selectively deuterated (Mills et al., 1984, 1986; Green et al., 1986; Green and Kramer, 1986a, b; Green and Doyle, 1990; Sokolov et al., 1989; Kausch and Tirrell, 1989). The depth resolution is approximately 80 nm and is limited by both the inelastic scattering processes which relate energy to depth, and the use of a stopper foil to eliminate accessory 4He particles (Stamm, 1992). Recent advances using a time-of-flight technique and removing the stopper foil have considerably improved the resolution (Sokolov et al., 1989). In the case of RBS, it is the 4He ions, backscattered from a heavy-element marker within the sample, which is analyzed to produce composition-profile information (Green and Doyle, 1990; Green et al., 1985). The depth
686
14 Polymer Surfaces and Interfaces with Other Materials
resolution for RBS is around 30 nm and has been used to study polymer interdiffusion by the movement of a gold marker located at a polymer-polymer interface (Green et al., 1985) A new method with improved resolution over FRES and RBS is nuclear reaction analysis (NRA) (Lanford, 1978; Payne et al., 1989; Chaturvedi et al., 1989; Steiner et al., 1990). Like FRES, NRA has the ability to examine composition profiles for partially deuterated polymer samples. Both *H and 2H can be depth profiled, depending on the incident beam: for example, by using a 0.7 MeV 3He beam, protons and 4He resulting from the reaction of the incident beam with deuterium can be energy-analyzed to give a 2H depth profile. The resolution of 3He-NRA varies between 14 and 30 nm, depending on the sampling depth (Stamm, 1992). Alternatively, hydrogen depth profiling can be achieved by using a 15 N beam. In this case, the 15N projectile reacts with protons at a resonance energy of 6.4 MeV to produce a 4.4 MeV y-ray. Hydrogen depth profiling is thus achieved by y-ray detection as a function of the 15N penetration depth, which is controlled through the acceleration voltage. The resolution in this case can vary between 4 and 12 nm, depending on the sampling depth (Stamm, 1992).
14.5 Materials Science Issues Concerning Polymer Surfaces 14.5.1 Overview Four principal areas of materials-science interest concerning the properties of polymer surfaces can be delineated: (1) surface chemistry, including modification and functionalization; (2) surface composition, as in multicomponent materials, where one
species can be enriched at the surface; (3) surface structure and organization and; (4) dynamics of polymer chains near surfaces. As explained in Sec. 14.1, surface chemistry will not be dealt with here in any detail, as it is a large subject and there are other excellent sources of information available (see Ward and McCarthy, 1987). We shall discuss problems that exist in each of the other three areas and in so doing illustrate the methods applicable to their study and explain the current state of knowledge on polymer surfaces. 14.5.2 Surface Composition Multicomponent polymers, including blends, copolymers and additive-containing polymers, are the norm in engineering practice. Indeed, they provide, in many cases, physico-chemical (as opposed to reactivechemical) routes to modify surface properties. Multiple components are often present in materials in order to control bulk properties and in so doing exert adventitious surface effects. The components used have a wide range of mutual miscibilities, from completely miscible to essentially immiscible. One component of a multicomponent mixture will segregate to the surface of the material if it lowers the energy of the system, and in most cases, this means that the component of lower surface energy will be enriched in the surface region. The characteristic dimensions of the depth of the surface-enriched zone will be of the same order as the size of the segregating macromolecules (several hundred angstroms), not just a few segments in depth (or several angstroms), as found for the surface regions of pure polymeric materials. The effects of segregation to surfaces in miscible systems were first observed some time ago during measurement of surface tensions of mixtures and copolymers (Ra-
14.5 Materials Science Issues Concerning Polymer Surfaces
stogi and St. Pierre, 1969). Nonlinear-mixing rules with varying compositions for the surface tensions of blends and block copolymers of poly(ethylene oxide) (PEO) and poly(propylene oxide) (PPO) were measured, whereas random copolymers exhibited linear mixing rules for y. The difference was attributed to the inability of the lower-surface-energy segments (i.e. PPO) to segregate effectively to the surface in random copolymers, although this was possible for blocks and blends. Bhatia et al. (1988) have observed a similar reduction of the surface tension in blends of poly (styrene) (PS) with poly(vinyl methyl ether) (PVME), suggesting segregation of PVME to the surface. Silicone polymers, such as poly(dimethylsiloxane) (PDMS) - are often employed in this way to lower the surface energies of materials. XPS has been used frequently (see Thomas and O'Malley, 1981; Schmitt et al., 1986, 1989; Clark et al., 1989) to detect the presence of surface segregation, relying on its chemical specificity to detect the presence of a characteristic element or chemical functionality (Briggs, 1987). (XPS is employed similarly to study surface contamination and to analyze purposeful surface modification.) XPS was employed very early on to pick up surface enrichment in multicomponent materials containing PEO (Thomas and O'Malley, 1979), and has become a standard tool in assessing the surface composition in many applications, for example in putative biocompatible materials such as polyurethanes (Gardella, 1989). Bhatia et al. (1988) were the first to correlate surfacetension reduction with the direct evidence of segregation provided by XPS. Recent efforts have been aimed at more quantitative measurements of surface enrichment than XPS can provide, and at the determination of the composition profiles in the near-surface region. Forward recoil
687
spectroscopy (FRES), dynamic SIMS and neutron reflectivity have all been used. All of these techniques can detect isotopic substitution and have been applied to blends of hydrogenated and deuterated poly(styrene) (PS and d-PS, respectively). The slightly lower polarizability of the the C-D bonds, relative to the C-H bonds, confers on d-PS a slightly lower surface energy, although, except at high molecular weights (Bates, et al., 1985), PS and d-PS are miscible in all proportions at the temperatures of interest. The first surface-enrichment experiments carried out on this system using FRES (Jones et al., 1989; Sokolov et al., 1989) were able to determine quantitatively the integrated surface excess (expressed as the thickness, z*, of a layer of pure d-PS containing an amount equal to the integrated amount of d-PS in the surface composition profile, using the bulk average composition as baseline). Surface excesses of tens of angstroms are seen and surfacevolume fractions of d-PS can be inferred to exceed their bulk values by a factor of about three for macromolecules with molecular weights of the order of 106, after equilibration for several days at 184°C. Owing to the depth resolution of FRES (~ 500 A), the complete composition profile, decaying as it does from the surface into the bulk over a distance of about the same length, cannot be determined by this method. Jones et al. (1990 a) have applied the higher spatial resolution methods of SIMS and neutron reflectivity. Figure 14-11 shows the reflectivity data for a PS blend containing 15% of the d-PS component, plotted in a format designed to emphasize the quality of the fit near the critical angle (the maximum in the figure). As can be seen in the inset, the fit can be made very good by a slight adjustment of the profile from the one predicted by a mean-field theory (Schmidt and Binder, 1985).
688
14 Polymer Surfaces and Interfaces with Other Materials
100
0.007
0.014
0.021
0.028
0.035
1
q (A" ) Figure 14-11. Reflectivity data obtained for a d-PSPS blend (5% d-PS), plotted as reflectivity times the fourth power of the neutron wavevector, and shown as a function of the wavevector q (Jones et al., 1990 a): the dashed line represents the best fit achievable by using the mean-field theory of Schmidt and Binder (1985); and the solid line is an empirically adjusted fit, constrained to have the same surface and bulk compositions, indicating that some improvement over the mean-field prediction is possible. The real-space forms of the composition profiles are shown in the inset.
The data of Fig. 14-11 pertain to two polymers of relatively high, and approximately equal, molecular weights. Several recent studies have examined the effects of equal molecular weight on this surfaceenrichment problem (see Composto et al., 1989, 1990; Hariharan et al., 1991). From this work, it is clear that lower-molecularweight polymers are enriched in composition near surfaces, all other things being equal. Thus, for two polymers, with different surface affinities and different molecular weights, the two surface-enrichment tendencies could be antagonistic, providing
an opportunity to balance the two effects against each other. Shull et al. (1990) have used FRES to determine the surface and interface excesses of a d-PS-poly(vinyl pyridine) (PVP) block copolymer mixed with PS, both at the PS surface and at the interface with homo-PVP. In this case, the surface-enrichment phenomenon is modified in several ways by the fact that the segregating species is a block copolymer. The d-PS would tend to segregate to the surface of the PS mixture, as is the case for a homopolymer. However, were the d-PS chains to go directly to the surface, they would then expose immiscible PVP to the PS bulk. Thus, what is inferred to happen, which is supported by microscopic evidence, is that micelles of d-PSPVP form, and these are preferentially attracted to the surface. Had the block copolymer been miscible with PS, for example, as in an block copolymer of d-PS and PS, the surface segregation would still have been different from the surfaceenrichment process of a homopolymer blend. Every d-PS block segregating to the surface entrains a tethered block of PS at the surface. The net result is that a brush of PS chains would form at the surface (Halperin et al., 1992); the energetics of this brush must be accounted for when modelling the assembly of the surface layer. This surface-enrichment phenomenon is connected theoretically to the subject of "wetting transitions", on which there has been an enormous body of work for simple fluids and binary alloys. The term "wetting transition" refers to the situation where a binary mixture, which has a miscibility gap, exists in a state of two-phase coexistence, and has a surface region that favors one of the phases. In this case, the surface can be coated with either a microscopic or macroscopic layer of one phase, even if the bulk is in the other phase. The phenomena
14.5 Materials Science Issues Concerning Polymer Surfaces
of transitions from microscopic to macroscopic layer thickness are called wetting transitions and are thought to occur universally when sufficiently close to the critical point. Several variants of mean-field theories are available for polymer mixtures (Nakanishi and Pincus, 1983; Schmidt and Binder, 1985; Carmesin and Noolandi, 1989; Hariharan et al., 1991). The first two are phenomenological, while the latter two are based on self-consistent mean-field (SCMF) theory, which enables a more accurate accounting of polymer configurations. The data of Fig. 14-11 are accounted for reasonably well by the theory of Schmidt and Binder (1985). Recent simulations of wetting transitions in polymer mixtures by Wang and Binder (1991) agree reasonably well with the theory of Carmesin and Noolandi (1989). The data of Fig. 14-11 show a monotonic composition profile, resulting from a longterm annealing of a surface layer to reach equilibrium. Jones et al. (1990b) have recently shown that for a homogeneous mixture, quenched into the two-phase coexistence region below the spinodal, a time-dependent sequence of oscillatory composition profiles develops. In this case, the surface enrichment phenomenon directs the spinodal decomposition phase-separation process to occur preferentially in a direction normal to the surface. The practical implications and possibilities for manipulating surface properties by these means still remain to be explored. 14.5.3 Surface Ordering The state of organization of a polymer surface can be the main determinant of many properties of polymer materials, not only those properties that are directly surface sensitive, such as adhesive or tribological
689
properties, but also properties that depend on transport or stress transmission through the surface. Surface ordering is a very important issue for semicrystalline polymers, and since the properties of these materials often depend strongly on the degree of crystallinity, the local degree of crystallinity at the surface could be expected to be significant for surface properties. In many cases, the degree of surface crystallinity is different from that in the bulk. Surfaces of semi-crystalline polymers often contain a high percentage of the amorphous component (Wu, 1982), and lower surface tension values for amorphous materials promote migration of that component to the surface. The degree of surface crystallinity can be systematically varied by using substrates which promote surface nucleation to varying degrees (Schonhorn, 1968a,b;Gray, 1974). Increasing the degree of surface crystallinity increases the surface energy, understandably in the light of Eq. (14-15), but not necessarily concomitantly with an increase in all surface properties. For example, surface ordering may create a brittle or weak boundary layer in the material. Determining the degree of surface crystallinity or ordering is difficult (AlsNielsen, 1987; Jark et al., 1989). Factor et al. (1991) have studied the surface ordering of the aromatic polyimide, poly(pyromellitic dianhydride oxydianiline) (PMDAODA), a material of significant technological importance as an interlevel dielectric in integrated microelectronics. PMDA-ODA does not crystallize in bulk, though it has been reported (Takahashi et al., 1984) to have local paracrystalline or liquid-crystalline order. X-ray reflectivity was used (Factor et al., 1991) to determine that a thermally cured film from a spin-cast precursor, of ellipsometrically determined thickness of 2600 A, had an rms surface roughness of
690
14 Polymer Surfaces and Interfaces with Other Materials
8 ± 2 A. Using a synchrotron x-ray source, grazing-incidence X-ray diffraction was measured. In this method the photons are incident on the surface of the sample at an angle less than the critical angle for total external reflection. An evanescent wave penetrates the sample and is diffracted by structural elements near the surface of the film. Varying the angle around the critical value can alter the depth of penetration, thus probing order as a function of depth. Near the surface of the PMDA-ODA material, it was found that there were strong reflections corresponding to interchain packing distances, which were not observed in the bulk polyimide. The authors concluded that the surface of this polymer was very highly crystalline, despite the absence of crystallinity in the bulk. Similarly, side-chain liquid-crystalline polymers forming smectic-A layers have been found by X-ray reflectivity to have the layers organized parallel to the surface (Stamm, 1992). STM has shown PEO and other polymers to adopt helical structures in thin layers or solids (Yang et al., 1990). This surface ordering has been explained in part through statistical-mechanical considerations, including segment-surface attractions (Chanetal., 1991). Surface ordering can occur in other microstructured polymers, even in the absence of crystallinity. Henkee et al. (1988) observed by TEM that thin-film samples of microphase-separated block copolymers, of different bulk-phase symmetries (lamellar, cylindrical, spherical), stacked in distinct ways, reflecting their accomodation to the contraints of the thin film. As in previous work by Hasegawa and Hashimoto (1985 a, b), Henkee et al. (1988) found that in the thinnest regions of the film, the formation of the bulk microdomain symmetry was inhibited, owing to the desire of the lower-surface-energy component of the block co-
polymer to occupy the surface preferentially. Russell and co-workers (Green et al., 1989; Coulon et al., 1989, 1990; Anastasiadis et al., 1989, 1990; Russell et al., 1990; Menelle et al., 1992), in a pioneering series of studies using neutron reflection, have studied the thin-film microstructure of poly(styrene)-poly(methyl methacrylate) (PSPMMA) diblock copolymer on silicon substrates. At higher molecular weights, this is a strongly segregating system, due to the relatively large repulsive interaction between the two blocks {% « °- 0 4 a t 25°C). Ordered lamellae were shown to be oriented parallel to the free surface at 25°C, over macroscopic lateral distances, after annealing above or near the glass transition temperature of this material (~ 100°C). Lamellar spacings measured for four different values of (symmetric) block degree of polymerization, N, were found to be comparable to those in bulk samples, all being greater than or equal to values calculated from theories rigorously applicable to the strong segregation limit OfAf» 10) (Helfand, 1975). For conditions in the vicinity of the bulk order-disorder transition (i.e. for lower molecular weights, where %N approaches 10), Menelle et al. (1992) inferred from the reflectivity, not dis-ordering but, rather, that weakly damped oscillatory composition profiles set in, where the bulk block copolymer would have been completely disordered. In this system, PMMA has a strong adsorptive affinity for the silicon substrate, whereas PS preferentially occupies the surface with the air. For very thin films, Menelle et al. (1992), as Henkee et al. (1988), found that the order in the block copolymer was disrupted, in this case probably because of the difficulty in satisfying the opposite surface affinities of the two blocks, while still maintaining order in the film.
14.5 Materials Science Issues Concerning Polymer Surfaces
Work by Foster et al. (1992) on symmetric poly(olefin) diblock copolymers in thin films on silicon enlarges on these observations. The materials in this study were poly (ethylene propylene)-poly(ethyl ethylene) (PEP-PEE) diblock copolymers. Being similar hydrocarbons, they are more weakly segregating than PS-PMMA and do not exhibit strong affinity for either surface; PEE is preferred at both the air and the silicon surface. Figure 14-12 shows the composition profiles across the thin film for the polymer with N = 975, which has an order-disorder transition temperature in the bulk of 125°C. Clearly the thin film maintains a discernible structure well above this temperature, crossing at higher temperatures into a regime of damped oscillatory profiles, as suggested by Menelle et al. (1992). For symmetric diblock copolymers near the order-disorder transition (Leibler, 1980), Fredrickson has developed a theory which predicts oscillatory and exponentially damped composition profiles, with the low-surface-energy sequence having a maximum concentration near the sur-
face (Fredrickson, 1988). Such predicted profiles have been tested experimentally via neutron reflectivity on thin films of PSPMMA diblock copolymers in the disordered state (Anastasiadis et al., 1989). These experiments revealed the free surface of the copolymer film to be enriched with PS. In agreement with the predictions of Fredrickson, the PS composition varies as cos
7=130°C 1 0.5
" = 75°C
0 500
1000
1500
2000
zA)
2500
3000
(2nz —jr-
(14-26)
where £, is the correlation length, L the period, 0e the surface-excess composition, 0 the average composition in the bulk and z the distance from the surface. For nearly symmetric PS-PMMA diblock copolymers, values of L = 150 A and | = 95 A, obtained experimentally, were in good agreement with theoretical predictions for these same quantities (Anastasiadis et al., 1990). A similar enrichment of the PMMA component at the silicon-substrate interface was also observed. In further support of these results, enrichment of the free surface by the PS sequence of a symmetric PS-PMMA
7"=U0°C
0
691
3500
£000
Figure 14-12. Composition profiles across a thin (~ 3600 A) film of PEP-PEE block copolymer shown for several temperatures spanning the bulk order-disorder transition (~ 125°C) (from Foster et al., 1992). Order in the film clearly persists above the bulk disordering temperature. The data obtained at 140°C approximately correspond to the damped oscillatory form suggested by Eq. (14-26) (Fredrickson, 1988).
692
14 Polymer Surfaces and Interfaces with Other Materials
diblock copolymer has been measured using angle-resolved XPS (Green et al., 1989). The surface excess of the PS block increased with increasing molecular weight of the copolymer. Another important observation concerning surface ordering of symmetric (lamellar) block copolymers is that of the appearance of "islands" in thin films (Ausserre et al., 1990; Collin et al., 1992). These islands occur because thin films are typically produced (for example by spin-coating) in arbitrary thicknesses; however, their internal microstructure dictates that they span this thickness in a quantized number of layers (the quantum number depends on whether or not the same or different blocks wet the two surfaces of the thin film). If the total film thickness does not equal the product of the quantum number times the individual lamellar dimension, there is a deficit or surfeit of material that can manifest itself as holes or islands on the surface of the film. The existence of these patches, and detailed studies of their time evolution, have been investigated by both PMIM and AFM methods (Ausserre et al., 1990; Collin et al., 1992). 14.5.4 Thin Films on Solids: Wetting, Spreading, Interactions and Dynamics Dynamics in surface regions of polymeric materials have been little explored, except for the case of interdiffusion between two polymers (Stamm and Majkrzak, 1987; Russell, et al., 1989), a subject that is reviewed thoroughly elsewhere (Kausch and Tirrell, 1989; Russell, 1991; Stamm 1992). Information is beginning to emerge in another area of polymer surface dynamics concerning the processes by which thin films of polymer spread on and wet the surfaces of other solids. These are important
processes for bulk polymer melts used as lubricants, coatings and adhesives. Polymer-melt films confined between mica sheets have been studied recently through force measurements (Horn and Israelachvili, 1988; Montfort and Hadziioannou, 1988; Van Alsten and Granick, 1990). A principal observation in common among these studies is the measurement of a strong repulsive force as the film is compressed down toward the molecular dimensions of the polymer molecules. The consensus of opinion among these investigators is that these long-range repulsive forces are manifestations of nonequilibrium effects, namely that the confined polymer material in contact with the surface is very slow to relax. On the theoretical side, there have been suggestions (Kremer, 1986), and computer simulations in support of them (Bitsanis and Hadziioannou, 1990; Chakraborty et al., 1991; Chakraborty and Adriani, 1992), that polymers in contact with solids can exhibit slow, glass-like relaxations, arising from strong adsorptive interactions between the macromolecules and the solid (Chakraborty et al., 1990; Konstadinidis et al., 1992). Silberzan and Leger (1992) have determined the diffusion coefficient of PDMS in contact with silica, both by observation of the rate of spreading of a molecularly thin film on the surface and also by pattern photobleaching of dye-labelled molecules (Kausch and Tirrell, 1989). They found that a PDMS sample (with a molecular weight of approximately 5 x 104) had a diffusion coefficient on the surface about 30 times smaller than that found for this same molecule in the bulk-PDMS melt. Furthermore, they found that a droplet of this polymer, during spreading on silica, exhibited not a smooth hemispherical cap-like shape but a plateau region [referred to as a "bump" or "step" (Brochard and de Gennes, 1984; Bruinsma, 1990)] around its periphery. This
14.6 Conclusions and Remarks on Future Directions
step has a thickness of about the radius of gyration of the macromolecules (found for several different molecular weights). The step occurs because the monomolecular film appears to have a greatly increased viscosity, owing to the attractive interaction of the polymers with the wall. The state of entanglement in a thin layer of bulk polymer interacting strongly with a wall is poorly understood but is likely to be an important factor in determining the thin-film viscosity. That entanglement may be an important issue is indicated by the observation of Silberzan and Leger (1992) that the radius of the edge of this plateau expands diffusively [r(t) oc t1/2], and, for different molecular weights the apparent diffusion coefficient varies as the inverse square of the molecular weight, as seen for polymer-melt diffusion (Kausch and Tirrell, 1989). Dewetting processes, in which homogeneous thin films of polymers that are trapped by nonequilibrium processing conditions such as spin coating on surfaces spontaneously rupture, form holes and de-wet, have also been observed recently by PMIM investigations (Reiter, 1992; Foster et al., 1992).
14.6 Conclusions and Remarks on Future Directions Polymer surfaces and interfaces are regions where large molecules experience constraints on their configurations and interactions that differ from those that they experience in the bulk. A considerable amount of experimental data and theoretical understanding have been developed over the last decade on the ways in which polymer surfaces affect structure and order in polymer materials. Surfaces are preferentially occupied by those constituents of the polymer materials that can best accomo-
693
date the constraints and interactions. Chain ends tend to be enhanced near surfaces, and the configurations of polymers near surfaces tend toward an orientation of the macromolecule and its segments parallel to the surface, driven largely by the incompressibility of bulk polymers. In multicomponent polymers, the surface region is generally preferentially enriched in the component that lowers the surface energy of the material. Surface energies of polymer fluids have been measured extensively, although assuring equilibrium in viscous materials is always an issue, and can be theoretically modelled reasonably accurately by the square-gradient version of Cahn-Hilliard theory. Existing techniques for measuring surface energies of polymer solids are all indirect and merit further development. Owing to the preference of some components for surfaces, segregation tendencies can produce ordering processes at surfaces, affecting crystallinity and microphase separation, and therefore the physical properties of polymer surfaces. Conversely, the constraints of near-two-dimensional thin films can frustrate the assembly of ordered structures. The dynamic processes at polymer surfaces have been studied extensively only in the context of interdiffusion, wetting and spreading processes. Computer simulations have preceeded experiments in their ability to bring out detailed structural features of polymer surfaces and may be expected to be similarly valuable in exploring polymer surface dynamics. The substantial progress that has been made on polymer surfaces in the last decade derives largely from advances in instrumental methods with new and specific kinds of surface sensitivity. Further new developments in techniques will be necessary for the outstanding problems of polymer-surface science. Particularly
694
14 Polymer Surfaces and Interfaces with Other Materials
lacking are surface- sensitive methods for assessing the physical properties of polymer surfaces. The needs include quantitative and theoretically understandable methods for studying solid surface energy, surface defects, adhesion, hardness, friction and permeability, among others. Beyond that lies the possibility of making connections between the impressive amount of new information on molecular surface structure and surface properties.
14.7 References Allen, F. G., Gobeli,G.W. (1962),Phys.Rev. 127, 150. Als-Nielsen, J. (1987), in: Structure and Dynamics of Surface II. Phenomena, Models and Methods: Schommers, W., vonBlanckenhagen, P. (Eds.). Berlin: Springer, pp. 181-222. Anastasiadis, S. H., Gangarz, I., Koberstein, J. T. (1988), Macromolecules 21, 2980. Anastasiadis, S. H., Russell, T. P., Satija, S. K., Majkrzak, C. F. (1989), Phys. Rev. Lett. 62, 1852. Anastasiadis, S. H., Russell, T. P., Satija, S. K., Majkrzak, C. F. (1990), /. Chem. Phys. 92, 5677. Arwin, H., Aspnes, D. E. (1986), Thin Solid Films 138, 195. Ausserre, D., Chatenay, D., Coulon, G., Collin, B. (1990), J.de Phys. 51,2571. Azzam, R. M. A., Bashara, N. M. (1977), Ellipsometry and Polarized Light. Amsterdam: NorthHolland. Bain, C. D., Whitesides, G. M. (1989), /. Phys. Chem. 93, 1670. Bashforth, S., Adams, J. C. (1892), An Attempt to Test the Theory of Capillary Action. London: Cambridge University Press. Bates, F. S., Wignall, G. D., Koehler, W. C. (1985), Phys. Rev. Lett. 55, 2425. Bergbreiter, D. E., Chen, Z., Hu, H. (1984), Macromolecules 77, 2111. Bhatia, Q. S., Chen, J. K., Koberstein, J. T., Sohn, J. E., Emerson, J. A. (1985), /. Colloid Interface Sci. 106, 353. Bhatia, Q. S., Pan, D. H., Koberstein, J. T. (1988), Macromolecules 21, 2166. Biegen, J. F, Smythe, R. A. (1988), Proc. SPIE-Int. Soc. Opt. Eng. 897, 207. Bilderback, D. H. (1981), Proc. SPIE-Int. Opt. Eng. 315, 90. Bitsanis, I., Hadziioannou, G. (1990), /. Chem. Phys. 92, 3827. Born, M., Wolf, E. (1980), Principles of Optics, Vol. 6. Oxford: Pergamon.
Boyce, J. F , Schiirch, R., Rotenberg, Y., Neumann, A. W. (1984), Colloids Surf 9, 307. Braslaw, A.,Deutsch,M.,Pershan,P. S., Weiss, A. H., Als-Nielsen, J., Bohr, J. (1985), Phys. Rev. Lett. 54, 114. Briggs, D. (1987), in: Polymer Surfaces and Interfaces, W. J. Feast and H. S. Munro (Eds.). Chichester, U. K.: Wiley. Brochard, F, de Gennes, P. G. (1984), J. Phys. Lett. 45, L-597. Brochard, F, Pincus, P. (1992), C. R. Acad. Sci. 314, 131. Bruinsma, R. (1990), Macromolecules 23, 276. Cahn, J. W., Hilliard, J. E. (1958), /. Chem. Phys. 28, 258. Carmesin, I., Noolandi, J. (1989), Macromolecules 22, 1689. Carton, J. P., Leibler, L. (1990), /. de Phys. 51, 1683. Chakraborty, A. K., Adriani, P. M. (1992), Macromolecules 25, 2470. Chakraborty, A. K., Davis, H. T., Tirrell, M. (1990), /. Polym. Sci., Polym. Chem. Ed. 28, 3185. Chakraborty, A. K., Shaffer, J. S., Adriani, P.M. (1991), Macromolecules 24, 5226. Chan, H. S., Wattenbarger, M. R., Evans, D. F, Bloomfield, V. A., Dill, K. A. (1991), /. Chem. Phys. 94, 8542. Chaturvedi, U. K., Steiner, U., Zak, O., Krausch, G., Klein, J. (1989), Phys. Rev. Lett. 54, 590. Chaudhury, M. K., Whitesides, G. M. (1991), Langmuir7, 1013. Chen, Y. L., Helm, C. A., Israelachvili, J. N. (1991), /. Phys. Chem. 95, 10736. Clark, D. T., Thomas, H. R. (1977), /. Polym. Sci., Polym. Chem. Ed. 15, 2843. Clark, D. T., Feast, W. J., Kilcast, D., Musgrave, W. K. R. (1973), J. Polym. Sci., Polym. Chem. Ed. 77,389. Clark, D. T., Feast, W. J., Kilcast, D., Musgrave, W. K. R., Modena, M., Ragazzini, N. (1974), J. Polym. Sci., Polym. Chem. Ed. 12, 1049. Clark, M. B., Burkhardt, C. A., Gardell, J. A. (1989), Macromolecules 22, 4495. Colletti, R. F, Gold, H. S., Dybowski, C. (1987), Appl. Spectrosc. 41, 1185. Collin, B., Chatenay, D., Coulon, G., Ausserre, D., Gallot, Y. (1992), Macromolecules 25, 1621. Composto, R. J., Stein, R. S., Kramer, E. J., Jones, R. A. L., Mansour, A. (1989), Physica B 57, 434. Composto, R. J., Stein, R. S., Felcher, G. P., Mansour, A., Karim, A. (1990), Mater. Res. Soc. Symp. Proc. 166, 485. Coulon, G., Russell, T. P., Deline, V. R., Green, P. F (1989), Macromolecules 22, 2581. Coulon, G., Aussere, D., Russell, T. P. (1990), /. de Phys. 51, 111. Davis, H. T., Scriven, L. E. (1982), Adv. Chem. Phys. 49, 357. Debe, M. K. (1987), Prog. Surf. Sci. 24, 1. de Gennes, P. G. (1987), C R. Acad. Sci. 305, 1181.
14.7 References
des Cloiseaux, J., Jannink, G. (1990), Polymers in Solution: Their Modelling and Structure. Oxford: Oxford Science Publishers. Dettre, R. H., Johnson, R. E. (1966), /. Colloid Interface Sci. 21,367. Dickman, R., Hall, C. (1988), J.Chem.Phys. 89,3168. Factor, B. J., Russell, T. P., Toney, M. F. (1991), Phys. Rev. Lett. 66, 1181. Fadley, C. S., Baird, R. J., Siekhaus, W., Novakov, T., Bergstrom, S. A. L. (1974), /. Electron. Spectrosc. Relat. Phenom. 4, 93. Feast, W. J.,Munro,H. S. (Eds.) (1981), Polymer Surfaces and Interfaces. Chichester, U.K: Wiley. Felcher, G. P., Hilleke, R. O., Crawford, R. K., Haumann, J., Kleb, R., Ostrowski, G. (1987), Rev. Sci. Instrum. 58, 609. Foster, M. D., Stamm, M., Reiter, G., Huttenbach, S. (1990), Vacuum 41, 1441. Foster, M. D., Sikka, M , Singh, N., Bates, F. S., Satija, S. K., Majkrzak, C. F. (1992), J. Chem. Phys. 96, 605. Fox, H. W., Zisman, W. A. (1950), J. Coll. Sci. 5,514. Fredrickson, G. H. (1988), Macromolecules 20, 2535. Freed, K. (1987), Renormalization Group Theory of Macromolecules. New York: Wiley . Garbassi, R, Occhiello, E. (1987), Anal. Chim. Acta 197, 1. Gardella, J. A. (1989), Anal. Chem. 61, 589A. Gardella, J. A., Pireaux, J. J. (1990), Anal. Chem. 62, 103. Goldstein, J. I., Newbury, D. E., Ecklin, P., Joy, D. G., Fiori, G., Ipshin, E. (1981), Scanning Electron Microscopy and X-ray Micro analysis, New York: Plenum. Gray, D. G. (1974), /. Polym. Sci., Part B 12, 509. Green, P. E, Doyle, B. L. (1990), in: New Characterization Techniques for Thin Polymer Films: Tong, H. M., Nguyen, L. T. (Eds.). New York: Wiley, p. 139. Green, P. E, Kramer, E. J. (1986a), Macromolecules 19, 1108. Green, P. E, Kramer, E. J. (1986b), J. Mater. Res. 1, 202. Green, P. E, Palmstrom, C. J., Mayer, J. W., Kramer, E. J. (1985), Macromolecules 18, 501. Green, P. E, Mills, P. J., Kramer, E. J. (1986), Polymer 27, 1063. Green, P. E, Christensen, T. M., Russell, T. P., Jerome, R. (1989), Macromolecules 22, 2189. Haller, G. L., Rice, R. W. (1970), /. Phys. Chem. 61, 1460. Halperin, A., Tirrell, M., Lodge, T. P. (1992), Adv. Polym. Sci. 100,31. Hansma, P. K., Elings, V .B., Marti, O., Bracker, C. E. (1988), Science 242, 209. Hariharan, A., Kumar, S., Russell, T. P. (1991), Macromolecules 24, 4909. Harrick, N. J., (1967), Internal Reflection Spectroscopy. New York: Wiley-Interscience. Harrick, N. J., du Pre, F. K. (1966), Appl. Opt. 5,1739.
695
Hartland, S., Hartley, R. W. (1976), Axisymmetric Fluid-Liquid Interfaces. Amsterdam: Elsevier. Hasegawa, H., Hashimoto, T. (1985 a), Macromolecules 18, 323. Hasegawa, H,, Hashimoto, T. (1985 b), Macromolecules 18, 589. Heavens, O. S. (1955), Optical Properties of Thin Solid Films. London: Butterworth. Helfand, E. (1975), Macromolecules 8, 552. Helfand, E. (1976), Macromolecules 9, 307. Helfand, E. (1980), in: Polymer Compatability andlncompatability: Sole, K. (Ed.). MMI Press Symposium Series Midland, MI, pp 143-163. Helfand, E., Tagami, Y. (1971), J. Chem. Phys. 56, 3592. Helfand, E., Wasserman, Z. R. (1982), in: Developments in Block Copolymers I: Goodman, I. (Ed.). Barking, U.K.: Applied Science, pp. 99-125. Helfand, E., Weber, T. A. (1976), Macromolecules 9, 311. Henkee, C. S., Thomas, E. L., Fetters, L. J. (1988), J. Mater. Sci. 23, 1685. Hong, K. M., Noolandi, J. (1981), Macromolecules 14, 1223. Horn, R., Israelachvili, J. N. (1981), /. Chem. Phys. 75, 1400. Horn, R., Israelachvili, J. N. (1987), /. Colloid Interface Sci. 775,480. Horn, R., Israelachvili, J. N. (1988), Macromolecules 27,2836. Horn, R. G., Smith, D. T. (1992), Science 256, 362. Huh, C , Reed, R. L. (1983), /. Colloid Interface Sci. 91, All. Ishitanti, A., Ishida, H., Soeda, E, Nagasawa, Y. (1982), Anal. Chem. 45, 682. Ishizu, K., Fukuyama, T. (1989), Macromolecules 22, 244. Israelachvili, J. N. (1991), Intermolecular and Surface Forces, 2nd ed. New York: Academic. Ito, K., Sauer, B. B., Skarlupka, R. J., Sano, M., Yu, H. (1990), Langmuir 6, 1379. Jark, W., Russell, T. P., Comelli, G, Soohr, J. (1989), Thin Solid Films 170,309. Johnson, K. L., Kendall, K., Roberts, A. D. (1971), Proc. Roy. Soc. London A 324, 301. Jones, R. A., Kramer, E. J., Rafailovich, M. H., Sokolov, J., Schwarz, S. A. (1989), Phys. Rev. Lett. 62, 280. Jones, R. A., Norton, L. J., Kramer, E. J., Composto, R. J., Stein, R. S., Russell, T. P., Mansour, A., Karim, A., Felcher, G. (1990a), Europhys. Lett. 12, 41. Jones, R. A., Norton, L. J., Kramer, E. J., Bates, F. S., Wiltzius, P. (1990b), Phys. Rev. Lett. 66, 1326; erratum (1991) 66, 3087. Kausch, H., Tirrell, M. (1989), Ann. Rev. Mater. Sci. 79,341. Kawaguchi, M., Takahashi, A. (1992), Adv. Colloid Interface Sci. 37,219. Kelley, M. J. (1991), MRS Bull. 16, 46.
696
14 Polymer Surfaces and Interfaces with Other Materials
Kim, M. W., Pfeiffer, D. G., Chen, W., Hsuing, H., Rasing, T., Shen, Y. R. (1989), Macromolecules 22, 2682. Koberstein, J. T. (1987), in: Encyclopedia of Polymer Science and Engineering, Vol. 8. New York: Wiley, p. 237. Konstadinidis, K., Thakkar, B. V., Chakraborty, A. K., Tannenbaum, R., Potts, L. W., Tirrell, M., Evans, J. R (1992), Langmuir 8, 1307. Kremer, K. (1986), /. de Phys. 47, 1269. Kumar, S. K., Vacatello, M., Yoon, D. Y (1988), /. Chem. Phys. 89, 5206. Kumar, S. K., Vacatello, M., Yoon, D. Y. (1990), Macromolecules 23, 2189. Lanford, W. A. (1978), NucL Instrum. Methods 149, 1. LeGrand, D. G., Gaines, G. L., Jr. (1969), /. Colloid Interface Sci. 31, 162. LeGrand, D. G., Gaines, G. L., Jr. (1973), /. Colloid Interface Sci. 42, 181. Leibler, L. (1980), Macromolecules 13, 1602. Lekner, J. (1987), Theory of Reflection. Dordrecht: Martinus Nijhoff. Lotz, B., Wittmann, J. C , Stocker, Y, Magonov, S. N., Cantow, H. J. (1991), Polym. Bull. 26, 209. Macleod, D. B. (1923), Trans. Faraday Soc. 19, 38. Madden, W. G. (1987), /. Chem. Phys. 87, 1405. Madden, W. G. (1988), /. Chem. Phys. 88, 3934. Mansfield, K. F., Theodorou, D. N. (1989), Macromolecules 22, 3143. McMaster, T. J., Carr, H. J., Miles, M. J., Cairns, P., Morris, V. J. (1991), Macromolecules 24, 1428. Menelle, A., Russell, T. P., Anastasiadis, S. H., Satija, S. K., Majkrzak, C. R (1992), Phys. Rev. Lett. 68, 67. Merrill, W. W., Pocius, A. V, Thakkar, B. V, Tirrell, M. (1991), Langmuir 7, 1975. Meyer, E., Heinzelmann, H., Rudin, H., Giintherodt, H. J. (1990), Z.Phys. B 79, 3. Mills, P. J., Green, P. R, Palmstrom, C. J., Mayer, J. W., Kramer, E. J. (1984), Appl. Phys. Lett. 45, 957. Mills, P. J., Green, P. R, Palmstrom, C. J., Mayer, J. W.; Kramer, E. J. (1986), /. Polym. Sci., Polym. Phys. Ed. 24, 1. Montfort, J. P., Hadziioannou, G. (1988), /. Chem. Phys. 55,7187. Nakanashi, H., Pincus, P. (1983), /. Chem. Phys. 79, 997. Padday, J. F. (1969), in: Surface and Colloid Science, Vol. 1: Matijevic, E. (Ed.). New York: Wiley, pp. 39. Parsonage, E., Tirrell, M., Watanabe, H., Nuzzo, R. G. (1991), Macromolecules 24, 1987. Payne, R. S., Clough, A. S., Murphy, P., Mills, P. J. (1989), Nucl. Instrum. Methods B 42, 130. Pireaux, J. J., Thiry, P. A., Caudano, R., Pfluger, P. (1986), /. Chem. Phys. 84, 6452. Pireaux, J. J., Vermeersch, M., Gregoire, C , Thiry, P. A., Candano, R., Clarke, T. C. (1988), /. Chem. Phys. 88, 3353.
Ploehn, H. J., Russel, W. B. (1990), Adv. Chem. Eng. 15, 137. Poser, C. I., Sanchez, I. C. (1979), J. Colloid Interface Sci. 69, 539. Prager,S.,Tirrell,M. (1981),/ Chem. Phys. 75,5194. Princen, H. M., Zia, I. Y. Z., Mason, S. G. (1967), /. Colloid Interface Sci. 23, 99. Quayle, O. R. (1953), Chem. Rev. 53, 439. Rabe, J. P. (1989), Angew. Chem. Int. Ed. Engl. 28, 117, 1127. Rabe, J. P., Buchholz, S. (1991), Phys. Rev. Lett. 66, 2096. Rabin, Y (1984), /. Polym. Sci., Polym. Lett. Ed. 22, 235. Rastogi, A. K., St. Pierre, L. E. (1969), J. Colloid Interface Sci. 31, 169. Reiter, G., (1992), Phys. Rev. Lett. 68, 75. Russell, T. P. (1990), Mater. Sci. Rep. 5, 171. Russell, T. P. (1991), Annu. Rev. Mater. Sci. 21, 249. Russell, T. P., Karim, A., Mansour, A., Felcher, G. P. (1988), Macromolecules 21, 1890. Russell, T. P., Coulon, G., Deline, V. R., Miller, D. C. (1989), Macromolecules 22, 4600. Russell, T. P., Jark, W., Comelli, G., Stohr, J. (1989), Mater. Res. Soc. Symp. Proc. 143, 265. Russell, T. P., Hjelm, R. P., Seeger, P. (1990), Macromolecules 23, 890. Sanchez, I. C. (1983 a), Annu. Rev. Mater. Sci. 13,387. Sanchez, I. C. (1983b), /. Chem. Phys. 79, 405. Sanchez, I. C.,Lacombe,R. H. (1976),/. Phys. Chem. 80, 2352, 2568. Sanchez, I. C , Lacombe, R. H. (1977), J. Polym. Sci., Polym. Lett. Ed. 15,11. Sanchez, I. C , Lacombe, R. H. (1978), Macromolecules 11, 1145. Sauer, B. B., Kawaguchi, M., Yu, H. (1987), Macromolecules 20, 2732. Scheutjens, J. M. H. M., Fleer, G. J. (1980), J. Phys. Chem. 84, 178. Schmidt, I., Binder, K. (1985), /. de Phys. 46, 1631. Schmidt, J. J., Gardella, J. A., Salvati, L. (1989), Macromolecules 22, 4489. Schmitt, R. L., Gardella, J. A., Salvati, L. (1986), Macromolecules 19, 648. Schonhorn, H. (1968 a), Macromolecules 1, 145. Schonhorn, H. (1968b) Adv. Chem. Ser. 87, 140. Shull, K. R., Kramer, E. J., Hadziioannou, G., Tang, W. (1990), Macromolecules 23, 4780. Silberberg, A. (1988),/. Colloid Interface Sci. 125,14. Silberzan, P., Leger, L. (1992), Macromolecules 25, 1267. Smith, C. P., Fritz, D. C , White, H. S., Tirrell, M. (1991), Thin Solid Films 198, 369. Sokolov, J., Rafailovich, M. H., Jones, R. A. L., Kramer, E. J. (1989), Appl. Phys. Lett. 54, 590. Sole, K., Stockmayer, W. H. (1971) /. Chem. Phys. 54, 2756. Stamm, M. (1992), Adv. Polym. Sci. 100, 357. Stamm, M., Majkrzak, C. F. (1987), Polym. Prepr. (Am. Chem. Soc, Div. Polym. Chem.) 28, 18.
14.7 References
Steiner, U., Kraush, G., Schatz, G., Klein, J. (1990), Phys. Rev Lett. 64, 1119. Sujata, K., Jennings, H. M. (1991), MRS Bull. 16, 41. Takahashi, N., Yoon, D. Y., Parrish, W. (1984), Macromolecules 17, 2583. ten Brinke, G., Ausserre, D., Hadziiouannou, G. (1988), J. Chem. Phys. 89, 4374. Theodorou, D. N. (1988), Macromolecules 21, 1391. Theodorou, D. N. (1989), Macromolecules 22, 4578. Thomas, H. R., O'Malley, J. J. (1979), Macromolecules 12, 323. Thomas, H. R., O'Malley, J. J. (1981), Macromolecules 14, 1316. Ulman, A. (1991), An Introduction to Ultrathin Organic Films: From Langmuir-Blodgett to Self-Assembly. San Diego: Academic. Vacatello, M., Yoon, D .Y, Laskowski, B. C. (1990), J. Chem. Phys. 93, 779. Van Alsten, J., Granick, S. (1990), Macromolecules 23, 4856. Van Krevelen, D. W. (1976), Properties of Polymers. Amsterdam: Elsevier. Wandass, J. H., Gardella, J. A. (1985), Surf Sci. 150, 107. Wang, J. S., Binder, K. (1991), /. Chem. Phys. 94, 8537. Wang, J. S., Binder, K. (1992), to be published. Ward, W. J., McCarthy, T. J. (1987), in: Encyclopedia of Polymer Science and Engineering: Supplement I. New York: Wiley, p. 674. White, J. R., Thomas, E. L. (1984), Rubber Chem. Technol. 57, 457. White, H. S., Earl, D. J., Norton, J. D., Kragt, H. J. (1990), Anal. Chem. 62, 1130. Wilhelmy, L. (1863), Ann. Phys. (Leipzig) 119, 111. Wittman, J. C , Stocker, Y, Magonov, S. N., Cantow, H. J. (1991), Polym. Bull. 26, 209. Wu, S. (1974), J. Macromol Sci. C10, 1. Wu, S. (1979), J. Colloid Interface Sci. 71, 605; erratum (1980) 73,590. Wu, S. (1982), Polymer Interface and Adhesion. New York: Dekker. Wu, S. (1989), in: Polymer Handbook, 3rded.: Brandrup, J., Immergut, E. H. (Eds.). New York: Wiley, pp. VI 411-434.
697
Yang, R., Yang, X. R., Evans, D. F., Hendrickson, W. A., Baker, J. (1990), /. Phys. Chem. 94, 6123. Yethiraj, A., Hall, C. K. (1990), Macromolecules 23, 1865.
General Reading Brown, H. R. (1991), The Adhesion Between Polymers, Annual Review of Materials Science, 21, 463-489. Doi, M., Edwards, S. F. (1986), The Theory of Polymer Dynamics. Oxford Science Publications, Oxford. Feast, W. J., Munro, H. S. (Eds.) (1987), Polymer Surfaces and Interfaces. John Wiley, Chichester. Halperin, A., Tirrell, M., Lodge, T. P. (1992), Tethered Chains in Polymer Micro structures, Advances in Polymer Science, 100, 31-71. Israelachvili, J. (1992), Intermolecular and Surface Forces, 2nd ed. Academic Press, London. Kausch, H. H., Tirrell, M. (1989), Polymer Interdiffusion, Annual Review of Materials Science, 19, 341-379. Koberstein, J. T. (1987), Surface and Interfacial Tension of Polymers, in Encyclopedia of Polymer Science and Engineering, Vol. 8. John Wiley, New York, p. 237. Lodge, T. P., Rotstein, N. A., Prager, S. (1990), Dynamics of Entangled Polymer Liquids: Do Linear Chains Reptate?, Advances in Chemical Physics, 79, 1-132. Russell, T. P. (1991), The Characterization of Polymer Interfaces, Annual Review of Materials Science, 21,249-268. Stamm, M. (1992), Polymer Interfaces on a Molecular Scale: Comparison of Techniques and Some Examples, Advances in Polymer Science, 100, 357-400. Ulman, A. (1991), An Introduction to Ultrathin Organic Films. Academic Press, Boston. Wu, S. (1982), Polymer Interface and Adhesion. Marcel Dekker, New York.
15 Crazing and Fracture of Polymers Ikuo Narisawa
Department of Materials Science and Engineering, Yamagata University, Yonezawa, Japan Albert F. Yee
Department of Materials Science and Engineering, University of Michigan, Ann Arbor, MI, U.S.A.
List of Symbols and Abbreviations 15.1 Introduction 15.2 Crazing: Initiation and Growth 15.2.1 What is Crazing? 15.2.2 Morphology and Structure of Crazes 15.2.2.1 Amorphous Glassy Polymers 15.2.2.2 Semicrystalline Polymers 15.2.3 Initiation of Crazes 15.2.3.1 Initiation Model of Crazes 15.2.3.2 Mechanical Criteria of Crazing 15.2.3.3 Crazing and Environment 15.2.3.4 Crazing and Polymer Structure 15.2.3.5 The Effect of Molecular Weight 15.2.4 Crazing: Growth and Theory 15.2.4.1 General Characteristics of Craze Growth 15.2.4.2 The Growth of Crazes at a Crack Tip 15.2.4.3 Microscopic Mechanisms of Crazing 15.2.5 Mechanical Properties of Crazes 15.2.5.1 Stress-Strain Behavior 15.2.5.2 Micromechanics of a Craze 15.2.6 Craze Thickening and Fracture 15.2.6.1 Kramer's Model for Craze Thickening 15.2.6.2 Ductile-Brittle Transitions 15.2.6.3 Craze Fracture Process 15.3 Fracture 15.3.1 Theoretical Strength of Polymers 15.3.2 Fracture Initiation in Uncracked Polymer Specimens (Micromechanism of Fracture) 15.3.2.1 Microvoid Generation due to Tension 15.3.2.2 Chain Scission in Oriented Polymers (Fibers) 15.3.2.3 Relationship Between Chain Scission and Macroscopic Fracture Materials Science and Technology Copyright © WILEY-VCH Verlag GmbH & Co KGaA. Allrightsreserved.
701 705 705 705 706 706 709 711 711 712 717 720 721 722 722 726 727 728 728 729 730 730 731 733 736 736 738 738 738 744
700
15.3.3 15.3.3.1 15.3.3.2 15.3.3.3 15.3.4 15.3.4.1 15.3.4.2 15.3.4.3 15.3.5 15.3.5.1 15.3.5.2
15 Crazing and Fracture of Polymers
Ductile Fracture Fracture of Shear Deformation Bands Plastic Fracture Failure Envelope Nonuniform Fracture of Blunt-Notched Polymers Brittle Fracture of Ductile Materials Stress Distribution Around a Local Plastic Deformation Zone Fracture of a Notched Specimen due to Plastic Constraints Fracture of a Cracked Specimen An Application of Linear Fracture Mechanics The Strain Energy Release Rate gc and Fracture Toughness Kc of Glassy Polymers 15.3.5.3 Plastic Deformation at a Sharp Crack Tip 15.3.5.4 Crazing at a Sharp Crack Tip 15.3.5.5 Various Factors Influencing Fracture Toughness 15.4 Concluding Remarks 15.5 Acknowledgements 15.6 References
745 745 746 747 748 748 750 753 756 756 756 758 758 759 762 762 762
List of Symbols and Abbreviations
List of Symbols and Abbreviations a A , B, C a0 b B c g ,c r C d E F AF GN gc 0craze glc AG (T, t) h /x k k! Kc KY Kcl Kc2 ^craze Ko XIc / /c /e /0 m Mc Me Mn Mo n n' NA Npu p pc
average diameter of micro voids; crack length (temperature-, time-dependent) constants initial crack length average distance between micro voids plate thickness compressibilities of the glassy, rubbery state material-dependent parameter average distance between entanglements tensile modulus force activation energy rubbery plateau modulus critical strain energy release rate critical strain energy release rate for the appearance of crazing gc of mode I fracture activation enthalpy active zone in Kramer's model of craze thickening first stress invariant Boltzmann's constant; constant depending on the polymer species constant critical stress intensity factor mode I stress intensity factor fracture toughness in the plane strain rate fracture toughness in the plane stress state critical stress intensity for the appearance of crazing stress intensity factor Kc of mode I fracture length period of a surface arc in the meniscus instability model; craze length in the environment craze length contour length between entanglements length of the repeat unit constant critical value of the molecular weight entanglement molecular weight number average molecular weight molecular weight of each repeat unit constant constant Avogadro number number of cavities at the ends of microfibrils hydrostatic pressure or mean stress; dilative stress dilative stress for crazing
701
702
15 Crazing and Fracture of Polymers
R rc r0 S s s* t T th Tg U vf Vo x x, y , z X, Z Y
gas constant; chain-scission rate craze length at the crack tip intermolecular distance craze fibril flow stress length of the plastic zone maximum extent of the logarithmic slip-line field time absolute temperature time to fracture glass transition temperature backbone covalent bond energy volume fraction of fibrils molar volume of a solvent crystallinity Cartesian coordinates directions normal to stress direction parallel to stress
a (xg,(xr P P{ P fl0 y F yp ys Ys 3 (5a,(5d,(5h <5 p , (5 S 5t <5pa><5pd <*sa><*sd e ea eb £c e x ,e y ,e z e1 Asy X Amax v
constant; activation volume thermal expansion coefficients of the glassy, rubbery state volume fraction of microvoids initial volume fraction of microvoids microvoiding rate microvoiding rate constant van der Waals surface energy craze surface tension energy of plastic deformation per unit area surface energy obtained surface energy solubility parameter (SP) SP contributions to polarity, dispersion force, hydrogen bonding SP of a polymer, solvent total SP ^ a A of a polymer ^ a A o f a solvent (overall) strain; creep strain strain of amorphous regions strain to fracture critical crazing value of tensile strain tensile strain components first respective principal strain strain distribution around crazes extension ratio maximum extension ratio Poisson's ratio
List of Symbols and Abbreviations
ve vx Q o <jb
o1,o2,o?> T (j) co co0
density of entanglement points density of chemical cross-links density; radius of the notch tensile or external stress stress bias (difference of m a x i m u m a n d m i n i m u m principal stresses); stress to fracture critical tensile stress for craze initiation stress in the radial direction in an U-notched bar axial tensile stress in a notched b a r average axial tensile stress tensile stress components uniaxial yield stress characteristic tensile stress behavior in uniaxial direction after yield maximum of Gy for x < s* critical stress for craze growth respective principal stresses octahedral shear stress crack tip opening displacement notch angle constant
ABS CED DZ ESR HIPS PA PAMSS PBT PC PE PEEK PEI PEOB PES PET PMMA POM PP PPO PS PSAN PSMLA PSMMA PSU
acrylonitrile-butadiene-styrene copolymer cohesive energy density deformation zone electron spin resonance high impact polystyrene polyamide poly(a-methylstyrene styrene) poly(butyl terephthalate) polycarbonate polyethylene poly(ether ether ketone) poly(ether imide) poly[p-(2-hydroxy ethoxy) benzoic acid] poly(ether sulfone) poly(ethylene terephthalate) polymethylmethacrylate poly(oxymethylene) polypropylene poly(phenylene oxide) polystyrene poly(styrene acrylonitrile) polystyrene maleic anhydride) polystyrene methyl methacrylate) polySulfone)
GC GT Gt (it Gx,Gy,Gz GY G'Y 'y(max)
703
704
PTBS PVC PVK PVT PaMS SAN SAXS SBR SEM SLF SP TEM WLF
15 Crazing and Fracture of Polymers
poly(ter£-butylstyrene) poly(vinyl chloride) polyvinylcarbazole polyvinyltoluene poly(a-methylstyrene) styrene acrylonitrile small angle X-ray scattering styrene-butadiene rubber scanning electron microscopy slip-line field solubility parameter transmission electron microscopy Williams-Landel-Ferry equation
15.2 Crazing: Initiation and Growth
15.1 Introduction Polymers are being used in an ever increasing range of applications. Whereas they were initially used as inexpensive substitutes for metal, ceramics and wood, polymers are now most often used because of their intrinsic properties. These include their optical, electric, magnetic, chemical, or specific strength properties, and their processability. However, in these applications, polymers often fail from the mechanical stresses that they are subjected to. It is therefore essential that we are familar with their mechanical behavior, and understand their physical and chemical bases. This might allow us to develop mechanically better polymers on the one hand, and to make allowance for their properties in the design process. The viscoelastic behavior (Chap. 9 of this Volume) and the yielding behavior (Chap. 10) of polymers are treated elsewhere in this volume. The present chapter is concerned with the failure, usually by fracture, of polymers. To understand the failure process, it is essential for us to realize that the mechanical behavior of a given polymer depends strongly on such factors as the molecular weight, its distribution, the degree of crystallinity, the thermal and processing history, the existence of anisotropy, etc. It also depends on such extrinsic factors as the flaw content, and the environment that the polymer is exposed to. Without a thorough knowledge of these materials aspects, it is virtually impossible to understand or predict the failure behavior of a polymer. Some of these aspects are discussed in this chapter. The reader interested in further details should consult the relevant chapters. In all high molecular weight, thermoplastic polymers, fracture is always preceded by crazing. We therefore begin our
705
discussion with this most important topic in Sec. 15.2. We will then proceed to discuss fracture itself in Sec. 15.3. The fundamental concepts of fracture mechanics, which are extremely important for analyzing and quantifying fracture, are treated in Vol. 6 of this Series.
15.2 Crazing: Initiation and Growth The word "craze" originally meant a pattern of fine cracks often found in the glaze on pottery. The phenomenon of craze generation is called crazing. Crazes in polymeric materials, however, are different from those observed on the surface of pottery. It is a characteristic of thermoplastic polymers. It represents an intermediate state between microscopic fracture and yielding, yet it can be considered a precursor to macroscopic fracture. Crazing plays important roles in the fracture of thermoplastic polymers, especially in such cases as brittle fracture, environmental degradation, and fatigue failure. One can say without exaggeration that the fracture of the majority of polymers cannot be explained without mentioning crazing. 15.2.1 What is Crazing?
Crazes exhibiting the pattern shown in Fig. 15-1 are often observed on the surface of glassy polymers such as polystyrene (PS) and polymethylmethacrylate (PMMA). These materials become whitened under additional stress and lose their transparency, the result of the generation of a large number of crazes that are much finer than those in Fig. 15-1. It was believed, due to their appearances, that crazes were microcracks occurring on the surface of a sample. Sauer and co-workers (Sauer et al., 1949; Hsiao and Sauer, 1959) investigated crazes
706
15 Crazing and Fracture of Polymers
Figure 15-1. Crazes occurring on the surface of PMMA.
in PS by optical and electron microscopy and X-ray diffraction techniques and demonstrated that crazes are not simply tiny cracks. However, crazing as a phenomenon was not extensively investigated until about a decade later. Despite extensive investigations, some controversy still exists among scientists about the underlying mechanisms to crazing. The difficulty in uncovering the physical basis of crazing is due to the complexity of the phenomenon, some aspects of which are outlined below: (1) Crazes are observed in high molecular weight glassy polymers except thermosets. They also occur in crystalline polymers such as polypropylene (PP) (Olf and Peterlin, 1974; Friedrich, 1977; Garton et al., 1977, 1978) and polyamide (PA) (Garton et al., 1978). (2) Crazes can be found on the surface and in the interior of a material as well as at crack tips. In glassy polymers, their microstructures are generally independent of where they occur. However, because the mechanical state at the free surface of a material is different from that in the interior, there is ambiguity in the mechanical criteria for craze initiation.
(3) For a given mechanical state, the occurrence of crazing is stochastic in both location and time. It is strongly affected by the distribution of microscopic defects on the sample surface and in its interior; consequently, a statistical treatment is often required in its analysis. (4) Crazes occurring on the surface, in particular, are affected strongly by the environment. Generally speaking, crazing is significantly accelerated by the presence of organic liquids. This phenomenon is called solvent crazing. It is, however, different from stress cracking. The latter refers to the situation where cracking occurs by stress acting alone, usually for a fairly long period of time. Since an induction period is necessary for the generation of a craze, it is affected by the absorption and diffusion of environmental agents. (5) In most cases, crack propagation (subor supercritical) is preceded by crazing ahead of the crack tip. Crazing at the crack tip requires a large amount of energy and therefore it can relax the stress concentration there. It is difficult, however, to relate quantitatively the large amount of energy consumption accompanying crazing to factors such as the polymer species, stress condition, and the environment. 15.2.2 Morphology and Structure of Crazes 15.2.2.1 Amorphous Glassy Polymers
Depending on where crazing occurs, the morphology of crazes can be classified into three types: (1) Surface crazes: This type of crazes often occurs on the surface of a specimen, as seen in Fig. 15-2. They grow in a direction normal to the maximum tensile stress. The length of the crazes can reach about 10 mm, while the thickness
15.2 Crazing: Initiation and Growth
This is illustrated by a schematic sketch and by an interference photomicrograph of a craze (Fig. 15-3). Surface crazes are almost certainly the result of some surface environment. (2) Crazes at a crack tip: Crazes can initiate at a crack tip and then propagated in the direction normal to the maximum tensile stress (Fig. 15-4). The difference between crack tip crazes and surface crazes is that the former often penetrate deep into the sample ahead of the crack, while surface crazes do not. It is also possible, depending on loading conditions, for a craze to grow until it traverses the entire specimen. The thickness of such crazes depends on the crack tip opening displacement and the material and may range from a few to over 10 |iim. (3) Internal crazes: Crazes can occur in the interior of a material. Such crazes are different from surface and crack tip crazes in that their occurrence is due purely to mechanical conditions, i.e., the environment probably plays little or no role in their initiation or growth. Figure 15-5 shows an internal craze in front of a plane strain plastic deformation zone in a notched sample of polycarbonate (PC). The craze plane is normal to the tensile stress direction.
i! '!
'.II f ! i . i ? ' *ii ; I.
I '
;
707
«
Figure 15-2. Surface crazes on a PS film after immersion in ethanol. The stress direction is perpendicular to the craze planes.
(in the stress direction) is extremely small, and is often in the range of 0.1 to several |im. They are also very sharp. The depth perpendicular to the surface is usually in the range of 0.1-0.5 nm.
Figure 15-3. A schematic crosssection of a craze (a) and a corresponding interference photomicrograph of a craze on the surface of a PS film (b). (a)
Ib)
708
15 Crazing and Fracture of Polymers
| Stress direction
0.1mm Figure 15-4. Crazes occurring at a crack-tip in PMMA.
Internal crazes can occur as isolated crazes or in multitude. The whitening of toughened glassy polymers - high impact polystyrene (HIPS), poly(acrylonitrile-butadiene-styrene copolymer) (ABS), and other styrenics - and sometimes toughened semi-crystalline polymers under stress is often the result of the occurrence of a large number of crazes inside the material. Multiple crazing is therefore an important toughening mechanism. The fact that a craze is not a crack can easily be demonstrated. Even if a craze occurring at a crack tip extends through the
Figure 15-5. A craze formed at the tip of a plane strain plastic zone in front of a notch in PC.
entire cross-section of a specimen, the specimen will still exhibit some load bearing capacity. Crazes are also different from shear localization in that they contain voids. Experimental techniques for the study of the structure of crazes include optical methods, which can estimate the fraction of voids by measuring the refractive index, X-ray and electron beam diffraction methods, which can measure the size of voids in a craze and characterize the orientation of the craze fibrils, and TEM and SEM observation methods, which can directly observe the structure of crazes in thin films and in bulk. Observations on the structure of crazes are too numerous to be described in detail here. The interested reader is referred to the list of General Reading given at the end of Sec. 15.6 if more details are desired. The following deductions can be made regarding the typical craze structure: (1) Crazes consist of elongated voids and fibrils with orientation in the major principal stress direction. This internal structure is therefore like that of a sponge. Surface crazes, internal crazes, and crazes at the crack tip all have this spongelike feature in common. This
15.2 Crazing: Initiation and Growth
means that the craze structure does not depend significantly on the environment. (2) The void content is usually in the range of 50-80%. The diameter of each void is about 20 nm. The voids are interconnected, forming a continuous network interrupted only by craze fibrils. (3) Craze fibrils are highly oriented along the deformation direction when under load. Their diameters are about 6 45 nm.
709
An example of surface crazes formed in PC that has been in contact with ethanol 1 is shown in an SEM micrograph (Fig. 15-6). This micrograph of a section perpendicular to the craze plane was obtained by breaking the specimen in liquid nitrogen. Figure 15-7 is a direct transmission electron micrograph of a craze in a PS thin film which has been immersed in methanol, which shows clearly the characteristic craze features described above. 15.2.2.2 Semicrystalline Polymers In contrast to crazes in amorphous polymers, the morphology and structure of crazes in semicrystalline polymers have received attention only recently. One of the basic problems posed by crazes in semicrystalline polymers is their complicated microstructure: a combination of crystalline and amorphous phases which constitutes many weak spots for craze nucleation. Another problem is that at room
Figure 15-6. SEM micrograph of a section of a craze in PC.
1 Loss of solvent may change the microstructure of a craze. The present examples, however, are adequate representations of both dry and solvent-induced crazes at these scales.
Figure 15-7. Ethanol-induced crazes in a PS thin film.
710
15 Crazing and Fracture of Polymers
temperature many crystalline polymers are above their glass transition and they tend to fail macroscopically in a ductile manner. Consequently, it has long been believed that crazing does not play an important role in the deformation and fracture of crystalline polymers. The appearance of crazes formed above glass temperature tend to be indistinct because of the lack of transparency; they resemble fine cracks because the thickness of individual crazes is usually more than 10 times larger than those observed for amorphous polymers. However, the appearance of the crazes is strongly affected by the crystalline texture which causes the direction and magnitude of local stresses to differ from that of the average far field stress. Figure 15-8 shows typical crazes formed in PP above its glass temperature. The length of individual crazes is shorter than that typically observed in amorphous polymers. These crazes grow preferentially along trans-spherulitic paths which run through the centers of the spherulites. The internal structure of these crazes are shown in SEM micrographs in Fig. 15-9. They are basically similar to those of crazes found in
amorphous polymers, although in this case the craze fibrils are much thinner. They span and interconnect the walls of what would otherwise be a microcrack. When crazes are formed in crystalline polymers composed of much finer spherulites at temperatures above their glass transitions, such as polyethylene (PE) at room temperature, they are short and irregular because the growth of these crazes is confined within a few spherulites and are more affected by local stress directions and the local spherulitic structure (Narisawa and Ishikawa, 1990). The short and irregular crazes are also found in the higher molecular weight crystalline polymers in which the amount of entanglement of amorphous tie chains in interlamellar and interspherulitic regions increases with reducing crystallinity and with increasing molecular weight (Friedrich, 1983). The crazes formed at temperatures below the glass transition temperature in semicrystalline polymers, such as nylon-6 at room temperature, are usually very few in number and are long compared with those formed above the glass transition temperature. They grow normal to the
Figure 15-8. Crazes formed in crystalline PP above its glass transition temperature.
15.2 Crazing: Initiation and Growth
711
Figure 15-9. SEM micrographs showing the internal structure of a craze in PP.
Figure 15-10. The effect
of temperature on crazes formed at the notch tip of a notched specimen of nylon-6 subjected to three point bending. Tg is between 313 K and 323 K. 303 K
313 K
323. K
major principal stress direction, ignoring the spherulitic structure as if it were in an amorphous polymer. Nevertheless, their internal structure is quite similar to those formed above the glass transition temperature. The effect of temperature on the craze pattern in nylon-6 is shown in Fig. 15-10. The glass transition temperature of nylon6 is between 313 K and 323 K. Thus, in general, the morphology and structure of crazes in crystalline polymers are affected by the crystalline structure, but the interspherulitic boundaries or the centers of the spherulites are not necessarily the favored sites for craze nucleation.
333 K
15.2.3 Initiation of Crazes 15.2.3.1 Initiation Model of Crazes
The structure of crazes indicates that its occurrence is accompanied by localized cavitation and fibrillation of molecular chains along the deformation direction. By way of contrasting with shear yielding, which involves no cavitation, Sternstein and Meyers (1974) termed crazing "normal stress yielding". Furthermore, they point out that strong constraints must exist in both of the normal directions (X and Z) in order to generate cavitation. These two types of yielding are illustrated in
712
15 Crazing and Fracture of Polymers
Fig. 15-11. In the case of internal crazes in which the plane strain condition is satisfied, it is easy to understand that crazing is generated due to normal stress yielding. However, on the surface of a material, since there is no such constraint in the Z-direction (which is perpendicular to the surface), one needs to consider a mechanism that includes, firstly, a certain type of deformation occurring on the surface of a material, and secondly, normal stress yielding occurring at a point immediately beneath it. The initial deformation on the surface could be microscopic yielding initiated by surface flaws and defects. The sequence of events leading to surface crazing can be illustrated by an experiment by Wellinghoff and Baer (1975) on the deformation of a polystyrene (PS) thin film by electron microscopy. They made the thin film of PS by casting a dilute solution of it onto a thicker, oriented poly(ethylene terephthalate) (PET) film. This sandwich was then deformed in a jig. The deformation was transferred to the PS film via the stronger PET film. The PS film was then stripped off and observed in the TEM. They found that micronecking oc-
curs first in regions about 30 nm in size, which are randomly dispersed in the material. These regions are in a dilated state although no voids greater than 1.5 nm could be found. Since further deformation of the micronecks was not possible because the elastic regions surrounding the micronecked regions impose constraints in the X- and Z-directions, cavitation and fibrillation of these regions then followed. These observations are depicted schematically in Fig. 15-12. Subsequent cavitation occurs in those regions where plastic deformation has proceeded to the greatest extent. Under such conditions the voids could grow to 10 nm in diameter. A similar observation has been made earlier by Wyzgoski and Yeh (1974). We note here the critical role played by flaws or microdefects in initiating the micronecks. We note also that localized shear plasticity is considered to be a key mechanism in the model on craze initiation proposed by Argon and coworkers which will be discussed in a later section. The observations by Wellinghoff and Baer appear consistent with Argon's model. An alternative mechanism that does not require the occurrence of local shear plasticity has been proposed by Kausch, which will also be discussed in a later section. 15.2.3.2 Mechanical Criteria of Crazing
(a)
(b)
Figure 15-11. A schematic representation of the deformations involved in shear yielding (a) and normal stress yielding (b).
A number of proposals on the criterion for crazing have been reported. They are summarized as follows: (1) Critical stress criterion: This criterion states that crazing occurs when the tensile stress reaches a certain value ac. The critical stress can be measured by optical techniques in a constant tensile stress experiment, i.e., in creep. In reality, the crazing stress depends on the strain rate and temperature, and can
15.2 Crazing: Initiation and Growth PET film Stress concentration region
|E| Shear yielding |
Voids of 10 nm in average
H I Porous micro-neck voids <1.5 nm
Figure 15-12. A schematic representation of shear yielding and crazing in a thin film of PC deformed by stretching it on a PET thick film substrate (Wellinghoffand Baer, 1975).
also be affected by molecular orientation and the environment. The usefulness of this criterion is therefore mainly macroscopic and empirical. (2) Critical strain criterion: This criterion states that crazing occurs when the tensile strain approaches a certain value ec. The critical strain can be obtained either by observing the crazing process in a sample strapped to a specimen holder which has the shape of a quarter ellipse and is immersed in a chemical liquid (or air), or by carrying out a bending test in a chemical liquid. The critical strain criterion is also macroscopic and empirical, but it is preferable to use this criterion to explain why crazing occurs even though the stress is gradually relaxed (Bernier and Kambour, 1968).
713
(3) Criterion based on fracture mechanics parameters: This criterion is useful primarily for crazing at a crack tip, which can be obtained by determining the critical stress intensity for the appearance of crazing Kcmzc or strain energy release rate gcraze. In addition, X craze and gCTaze are sometimes used as parameters determining the growth rate of crazes (Andrews and Bevan, 1972). (4) Dilative stress criterion: This criterion proposes that the mechanical criterion for crazing should include the dilative stress component since it is required if cavitation does occur in crazes. Sternstein and co-workers (Sternstein and Meyers, 1974; Sternstein et al., 1963) proposed that the segmental mobility of the polymer will increase due to dilative stresses and that at a certain point this activated mobility triggers the cavitation and causes molecular segments to orient along the maximum stress direction; as a result, a characteristic structure is formed. The criterion can be expressed as B
(15-1)
where oh is the stress or stress bias required for orienting the fibrils in crazes, It is the first stress invariant and i\ = GX + G2 + G3 > 0, where al9 a2,
+
B
(15-2)
714
15 Crazing and Fracture of Polymers
The regions of crazing and shear yielding obtained under torsion-tension stress conditions for PMMA are shown in Fig. 15-13. The agreement is seen to be excellent. However, a difficulty in the criterion expressed by Eq. (15-2) is that <7b = o1 — o2 is in fact proportionate to the shear stress, and that the strain direction is the maximum shear stress direction. Yet, crazes grow in the direction normal to the maximum principal stress. This obviously cannot explain the normal stress yielding which was originally proposed by Sternstein himself. The crazing criterion proposed by Sternstein was further criticized by Breuer (1979, 1985), who pointed out that it does not yield a significant failure surface in the principal stress space or a significant failure curve in the octahedral stress diagram. Its validity is therefore limited to the plane
(15-1) Crazing and shear yielding
4.0 ^NPure shear \ \
\
Crazing
A >
\ *r 3.0
o
Shear yielding
V N
= 2.0
\ y^—Von Mises shear yielding locus
1.0 -
No crazing no shear yielding \ \ \ \ \ \ \ \
0
-3.0
-2.0
-1.0
a2 d0" 9 N/m2)
Figure 15-13. Effect of biaxial stress state on the mode of deformation of PMMA (Sternstein and Meyers, 1974). (o) No crazing or shear yielding; (•) crazing only; (A) shear yielding only; (A) crazing and shear yielding. The temperature is 60 °C.
stress case. Therefore, Breuer considers this criterion to be useful for describing surface crazing only. Noting some of these difficulties, Oxborough and Bowden (1973, 1974) performed experiments on PS in the tension-compression quadrant (a3 = 0). They found that the data could be fitted to an equation of the form —V(7j
=
(15-3)
where v is Poisson's ratio, and A and B are temperature- and time-dependent constants. Since a1 — va2 = Eel9 Oxborough and Bowden proposed that the craze criterion can be written as Y'
(15-4)
Although Eq. (15-4) has the same form under plane stress as that proposed by Sternstein (Fig. 15-13), the physical implication here is more explicit, i.e., it is a critical strain criterion. Furthermore, the strain direction is normal to the craze plane which is consistent with experimental observations. Another advantage of this criterion is that it is three dimensional. However, it should be noted that Breuer (1979) also objects to this criterion on the basis that a critical strain criterion cannot be circumscribed by a unique three-dimensional failure boundary. Using the same assumption that cavitation from dilative stress is the primary cause of crazing, Gent (1970, 1973) proposed a quite different criterion. In this criterion, the dilative stress decreases the glass transition temperature of materials to the level of the testing temperature. He further proposed that it is the transition, i.e., from the glassy state to the rubbery state, that makes cavitation possible. The dila-
715
15.2 Crazing: Initiation and Growth
tive stress for crazing pc is given by
- No crazing
_20
o
Pc = P(Tg-T)
with p =
cr
^ ^ (15-5) cg
2
(15-6)
Given the experimental conditions described, Zhurkov probably detected scattering from crazes that had already been formed, rather than from craze nuclei. Recent results from Kramer and co-workers using synchrotron radiation found no evidence of precursor voiding.
•o
o-
18 \
where ag, and ar are the thermal expansion coefficients of the glassy and rubbery states, respectively, and cg, and cr are the compressibilities of the two states. In ordinary polymers, fi is about 5 MPa • °C~l, thus for materials with glass transition temperature of 100-150 °C, pc is in the range 400-500 MPa. According to Gent's proposed criterion, crazing should be suppressed if an external hydrostatic pressure of 400-600 MPa is applied to a material. Such an experiment was performed by Ishikawa and Narisawa (1983). Their experimental data, given in Fig. 15-14, indicate that crazing can be eliminated at a much lower pressure. This observation is obviously inconsistent with the calculation based on Eq. (15-5). Argon and co-workers (Argon, 1974; Argon et al., 1977; Argon and Hannoosh, 1977) proposed that voids with diameters of about 10 nm may first occur in the material because of chain breakage due to the mechanical stress. Their assumption is based on SAXS results obtained by Zhurkov 2 (Zhurkov et al., 1969 b). Furthermore, he regarded voiding as a stressdependent kinetic process, and he assumed the voiding rate dependence on temperature to obey Arrhenius' equation. That is, the microvoiding rate /? can be expressed by: kT
Quenched sample
/
crazing / — S l o w cooled sample
o
0
s
10 20 30 40 50 60 External hydrostatic pressure (MPa)
70
Figure 15-14. Effect of external hydrostatic pressure on internal crazing in PC at 23 °C. Crazing is completely suppressed when sufficiently high pressure is applied (Ishikawa and Narisawa, 1983).
Elastic area
Surrounding area of microvoids
Figure 15-15. Argon's model for craze initiation assumes plastic dilatation of microvoids (Argon et al., 1977).
where fi0 is a constant, AG (T, p) is the activation enthalpy, T is the temperature, and k is Boltzmann's constant. AG depends on the dilative stress p and the octahedral shear stress t. Consequently, the number of microvoids increases linearly with time, and will become, after time t,
[-^TT4]
(15-7)
Accordingly, once the fraction of voids in a material reaches a certain value, the microvoids (shown in Fig. 15-15 by a circle with a radius r) will plastically expand themselves as they unload the elastic stress in
716
15 Crazing and Fracture of Polymers
the surrounding area. By assuming a material to be strain-hardening, the characteristic behavior after yield can be given by <7v = <7Y + A ( X2 — y
(15-8)
where oY stands for the uniaxial yield stress, A is the extension ratio, and A is a constant. On the basis of the Mises criterion, the stress in the radial direction can be obtained by solving the equation of equilibrium, which gives the stress p required for these microvoids to plastically expand and transform into the nuclei of crazes: (15-9)
As defined in Fig. 15-15, if the average diameter of the microvoids is a, and the average distance between the microvoids is b, then the volume fraction P of the microvoids is P = (a/b)3. ft{ is the initial volume fraction of the microvoids at the instant of plastic expansion. In the case of no strain hardening, Eq. (15-9) can be simplified as follows: 2(7 Y
(15-10)
In reality, the actual dilative stress required for the nucleation of crazes, when a material is in a state near yield, is much lower than the value estimated from Eq. (15-10) (Argon and Hannoosh, 1977). Once they arrive at a certain size, the nuclei that have homogeneously expanded in the initial stage will deform preferentially in the direction normal to the tensile stress axis, because of the greater stress concentration in this direction. Fibril orientation will also become easier. The overall shape of the craze will become oblate and grow in the normal direction. The Argon criterion is quite complicated because many param-
eters need to be determined when it is to be used for prediction. However, it is the first theory that gives an explanation for the time-dependence of crazing. In addition, the concept that the voids can plastically expand as the surrounding stress is elastically unloaded is based on the observation of void expansion in ductile metallic materials. The difference between the two cases lies in the addition of the mechanism, in the case of polymers, wherein the voids do not evolve into a porous patch but instead become crazes by collective deformation. This mechanism is possible due to the strong strain hardening capability of high molecular weight polymers. Kausch proposed an initiation mechanism that is similar to that of Argon but does not require the use of plasticity concepts (Kausch, 1976). This model takes into account molecular structure, chain rigidity, conformational changes, and intermolecular interactions. The initiation process (Fig. 15-16) itself is proposed to consist of three sequential steps, viz.: (1) Decohesion of molecular coils in regions of low entanglement (B1 in Fig. 15-16), which forms unstable microvoids, i.e., these microvoids could collapse due to elastic strain. (2) Anelastic deformation of adjacent molecular coils which are more entangled (B2). (3) Transmission of strain to adjacent coils which stabilizes the microvoids. With sustained or increased load, the microvoids coalesce and leave fibrils which connect the opposite craze surfaces. Narisawa and co-workers (Ishikawa et al., 1977; Narisawa et al., 1980 a) considered more simple reasons for cavitation. They proposed that cavitation before crazing must be preceded by some type of shear yielding. For example, if yielding occurs in the localized area in front of a notch in a
15.2 Crazing: Initiation and Growth *
•
•
^
-
—
"
—
*
—
.
-
-
"
• '
717
-
•
v
>
I X
(
[
5I
^ \
j
/
Y
/
I
-
\
—-r
r -4—/
•
'
-
"
i*
J
>
X
\
>
-
/
•
3« *
\
•k
10 nm IPMMA M = 60000 g/mol.
-
.
nm. a = 4.A nml
-" ,
-
-
Figure 15-16. Kausch's model of craze initiation and fibril formation assumes that in an elastically strained region A, there exist regions B, which have insufficient interpenetration of molecular coils 1 and 2. B : are then possible nucleation sites. B2 are regions where there is sufficient interpenetration of molecular coils 3. The stresses in Bt are transferred to B 2 . This process continues to propagate to surrounding areas. (Kausch, 1976.)
material, the elastic area that surrounds the yielded area will act in such a way that the plastic shear yield stress will be raised. If this occurs in a sufficiently thick specimen (the plane strain state), a large dilative force in front of the yielded area will act to initiate microvoids. (For a detailed calculation of the dilative force, refer to Hill's slipline field theory described in the section on ductile fracture.) Most of the criteria proposed so far pertain to crazing on the surface. The effects of dilative stresses are implicit in these criteria. However, since no dilative stress can physically exist on the outermost surface of a material, experimental observations would seem to be contradictory to most of the proposed criteria. But, if one considers the flaws and defects on the surface as small notches, then a dilative stress can still be obtained. This is possible because local-
ized shear yielding can occur first in front of flaws and defects on the surface of a specimen, and if the front radius is small enough so that a triaxial stress field is generated inside the specimen, then this stress field induces a stress concentration and eventually leads to cavitation (voiding). The time-dependence of crazing can also be explained on this basis if we consider the fact that shear yielding that must precede voiding is a kinetic process. 15.2.3.3 Crazing and Environment
Crazing can be accelerated significantly under the influence of organic liquids. The effects of environmental liquids can be classified inter two categories in relation to the crazing mechanism: (1) the environmental liquids accelerate both the shear yielding process, which precedes crazing,
718
15 Crazing and Fracture of Polymers
and the fibril orientation process in crazes; (2) the environmental liquids decrease the surface energy during cavitation. The first category describes the role of environmental liquids as plasticizers due to absorption. Generally speaking, the glass transition temperature Tg of polymers will decrease because of the swelling due to the absorption of chemical liquids; consequently, it may be possible for molecular segments to move at lower stresses than if there was no swelling. Figure 15-17 shows the decrease of Tg of PS after being swollen to equilibrium in various chemical liquids versus the critical strain of crazing gc (Kambour et al., 1973). The ec for PS, which has been plasticized by chlorobenzene in advance, is also plotted in the same figure. The results clearly indicate that lowering the Tg will have identical effects on crazing, no matter what method is used to swell the sample. It has also been shown for materials other than PS that accelerated crazing is a combined effect of the decrease in Tg due to swelling and the interaction between the material and chemical liquids. The crazing behavior at crack tips exhibits exactly the same behavior.
100 90 80 70 60 50 40 30 20 10 0-10-20-30
Figure 15-17. Decrease in Tg of PS due to solvent absorption, and its effect on the critical strain for crazing (Kambour et al., 1973). (o) Swollen samples in various chemical solvents; (•) extended in air mixed with PS-dichlorobenzene.
It takes a long time to swell a material until equilibrium is reached; moreover, it is difficult to simultaneously measure the Tg. For those chemical liquids that have slow absorption rates, the time to reach equilibrium becomes especially long. Therefore, in order to evaluate accelerated crazing by a chemical liquid, it is convenient to quantify the interaction between the material and the chemical liquid by the solubility parameter (SP). The solubility parameter is often denoted by 5, and 3 = ^JCED where CED is the cohesive energy density of the material or the solvent. Figure 15-18 schematically depicts the relationship between the critical crazing strain sc and the SP value. Let the SP values of polymer and solvent be 3p and 3S, respectively. When 3p becomes nearly equal to 3S, the polymer will very likely become dissolved in the liquid or, if not dissolved, will be cracked intensely. With the increase in the difference between 3p and (5S, crazing begins to occur, and sc will also increase and the chemical-acceleration effect becomes less significant. The relationship depicted in Fig. 15-18 is especially useful for the situation in which a nonpolar material and a nonpolar solution are involved. In a situation where polarity or hydrogen bonding is involved, however, a different relationship might be exhibited. In Fig. 15-19, for instance, the crazing ability of the solvents are expressed in terms of the SP values and the hydrogen bonding parameter with respect to a polar polymer, PMMA, in a two-dimensional plot (Vincent and Raha, 1972). If the SP value contains contributions from the dispersion force, polarity, as well as hydrogen bonding, denoted as 3d, 3a and 3h respectively, then it is useful to consider the effective total solubility parameter 3t (Henry, 1974): <5? = <5d + <5a2 + <5h
(15-H)
15.2 Crazing: Initiation and Growth
719
We also need to consider the enthalpic contribution from each of these components, which is proportional to:
Crack or dissolution Craze
V0(Sp-Ss)2
Solubility parameter $s (\/J/mJ
Figure 15-18. Schematic relationship between the critical crazing strain of a polymer and the solubility parameter SP of solvents.
U
(15-12)
where Vo is the molar volume of a solvent. In the case of polymers with polarity, it has also been proposed (Jacques and Wyzgoski, 1979) that it is reasonable to consider the contributions from dispersion 3d and polarity (5a separately and express them as Void,* ~
22 20 3 18
w w
'V
v
o en
•-§
o _o
1.5
u 12
gMO
10 11 12 SP value
13 U
15 16
Figure 15-19. The combined effect of hydrogen bonding parameter and SP on crazing behavior of PMMA (Vincent and Raha, 1972). (•) Dissolution, ( x) crack, (v) small crazing ec, (A) large crazing sc.
Figure 15-20. A crazing strain map for a polar polymer (PC-) polar solvent system (Jacques and Wyzgoski, 1979).
720
15 Crazing and Fracture of Polymers
crazes on the surface of a material at sufficiently high pressure (Matsushige et al., 1975). It is therefore possible that results from some previously reported mechanical experiments conducted under superimposed hydrostatic pressure may contain artefacts. The structure of crazes generated also depends on the environment itself. A combination of a material and an environmental agent that can lower the Tg significantly will exhibit larger volume fraction of voids (Yaffe and Kramer, 1981).
melt state and the melt as if it were a rubber network, the rubbery plateau modulus GN can be related to the entanglement molecular weight Me by
15.2.3.4 Crazing and Polymer Structure
where NA is Avogadro's number. The contour length le between entanglements is
The effect of molecular structure on crazing behavior is of utmost interest, and has been explained in detail by Kramer and his coworkers in terms of molecular entanglements. Although the direct relationship between the chemical structure and intermolecular interactions which result in networklike behavior, i.e., entanglements, is not explained by these workers, the microscopic model produced by them is both intuitive and consistent with experimental observations. Kramer's model is most clearly explained by two reviews written by himself (Kramer, 1983; Kramer and Berger, 1990). Due to space limitations, only a brief summary of the part of the model that deals specifically with molecular entanglements is presented in this and a later section. The reader interested in more details and other aspects of the model should consult the reviews and original references cited therein. A key feature of Kramer's model is that, in order for a craze to be stable, there must be molecular entanglements. Thus, the deformation of a stable craze can be related to the deformation of the molecular segment between entanglements. For polymers with a sufficiently high molecular weight so that interactions occur in the
QRT
(15-13)
where Q is the density. In such a network, the density of chains between entanglement points, ve, is ve = g^
le = l0—
(5-14)
(15-15)
where /0 is the length of the repeat unit and M o is the molecular weight of each repeat unit. If the polymer chain can be treated as a Gaussian coil, then the average distance between entanglements, d, is =
C(Me)112
(15-16)
where C is a constant. It is clear from these definitions that the maximum extent that the chain between entanglements can be stretched, Amax, is X
™ = -d
(15-17)
The actual extension of craze fibrils X can be calculated from densitometry measurements on transmission micrographs of crazes formed in thin polymer films. The results of such experiments on a number of polymers and copolymers are shown in Fig. 15-21 plotted against the theoretical Xmax. These results show that the general trend of X agrees very well with the Xmax calculated. The deviation at high X is suggested by Kramer to be due to chain scission, a process to be explicitly considered in a later section. As the entanglement den-
15.2 Crazing: Initiation and Growth
721
10
4 -
3 ~
2 ~ PSAN2 1 -
sity increases (i.e., as Amax becomes smaller), the polymers tend to deform by shear yielding (at temperatures well below Tg) because the crazing stress at the craze tip increases beyond the shear yielding strength of the material. The craze to shear yield transition can then be understood in terms of the relative ease of shear yielding vs. crazing as the entanglement density is changed. This has been demonstrated in a series of miscible blends of polystyrene and poly(phenylene oxide) (PPO) (Donald and Kramer, 1982) where the transition to shear yielding behavior when the fraction of PPO is increased is explained in terms of a gradual increase in the average entanglement density. 15.2.3.5 The Effect of Molecular Weight
Since entanglements are so important in the deformation of crazes, it is not surpris-
Figure 15-21. Experimental extension ratio of crazes in various homo- and copolymers plotted against the theoretical maximum extension ratio (Kramer, 1983).
ing that the molecular weight strongly influences the number and configuration of crazes. At low molecular weights, both the width and length of crazes are large but their number density is small. The number of fibrils inside such a craze is also very low (Brady and Yeh, 1973). In contrast, at high molecular weights, a number of thin crazes are generated (Sato, 1965; Fellers and Kee, 1974; Rudd, 1973). It was generally thought that the more chain entanglements there are, i.e., the higher the molecular weight is, the more difficult it is for fibrils to form within a craze. The crazing stress might therefore be expected to increase with increasing molecular weight. Nevertheless, an examination of the surface crazes of PS indicates that there is no molecular weight dependence as long as Mn > 2M e (Fig. 15-22). When Mn < 2M e , no crazing was observed and only load-in-
722
|
15 Crazing and Fracture of Polymers
30
S 20 Zn 10
-M.
0
5 6 7 8 9 10 1 1 1 2 J 3 U Number average molecular weight Afnx10~*
15
Figure 15-22. Molecular weight dependence of crazing and fracture stress of PS (Fellers and Kee, 1974). (•) Failure stress, (•) crazing stress.
18 9-"
from crazing to shear yielding (Henkee and Kramer, 1984; Glad, 1986; Berger and Kramer, 1988). The fact that crazing stress is independent of molecular weight if M n > 2 M e has been interpreted to mean that crazing is not accompanied by any large scale segmental movement. This is consistent with Kramer's model presented in the last section. An important corollary of this observation is that craze growth must take place primarily by drawing fresh chains from the uncrazed surrounding into the craze once Amax is approached. A discussion on craze growth is the subject of the next section. 15.2.4 Crazing: Growth and Theory 15.2.4.1 General Characteristics of Craze Growth
Length (mm)
Figure 15-23. The growth of both length and thickness of crazes at a crack tip in PMMA (Kramer et al., 1978).
duced fracture occurred. Where crazing occurs, failure starts from these crazes, and the failure stress increases as the molecular weight increases. This implies that the stress required for extending the fibrils to their critical strains increases with molecular weight. The effect of molecular weight and entanglement density can be extended to the case of chemically cross-linked polymers. If a linear polymer is gradually cross-linked, the deformation mode can gradually shift
Both the length and the thickness of crazes increase with loading time, as illustrated by results on PMMA (Kramer et al., 1978) in Fig. 15-23. Since the thickening process eventually leads to fracture, this process is discussed in conjunction with craze fracture in a later section. In the present section we focus on some general characteristics of craze growth, especially in the longitudinal direction. Unless otherwise indicated, the term growth rate refers in this section specifically to longitudinal growth. Surface crazes do not continue growing indefinitely; instead, their dimensions will approach a certain equilibrium as time elapses. It has also been found that crazes evolve with a constant length to thickness ratio (Verheulpen-Heymans, 1979). The growth behavior of crazes is related to the stress state around them, especially near the tips of growing crazes. Therefore, it can be expected that the stress state of the surface craze is much different from that of the craze at the crack tip. Figure 15-24 shows
15.2 Crazing: Initiation and Growth 50°
40°
30°
20°
10° 0=0
/?=!0
(a) 50° 60°
70°
/,0°
/
30°
20° >^
10° 6-
<> Ay x h >
--——
2/
^--—
A
\
\
80°
/
1I
/
/
723
stress direction coincides with the maximum strain direction, it is not possible to distinguish which one is related to the growth direction of the crazes. To clarify this, one can use an anisotropic specimen in which the molecular chains are oriented in a certain direction, and carry out tensile tests at different directions to determine whether the principal stress or principal strain direction is related to the growth of crazes. Figure 15-25 gives the results of one such experiment (Beardmore and Rabinowitz, 1975). It indicates that crazes do not necessarily grow in the direction normal to the principal stress. The crazes grown in the direction normal to the maximum strain either cease to grow if their tips encounter a shear deformation band (Fig. 1526), or continue to grow, but in a different direction, i.e., along the shear deformation band. We now consider briefly the growth kinetics of crazes in the thickness direction.
(b)
Figure 15-24. Principal stress directions and growth directions of crazes in PC with (a) a hole and (b) a rigid cylinder in the specimen. R is the radius of the hole or cylinder (Miltz et al., 1978).
how the crazes in a PC specimen, which contains a hole or a rigid cylinder, evolve. The maximum principal stress direction is indicated in the figure in order to relate the growth of crazes with the stress direction. It is apparent that crazes grow in the direction normal to the maximum principal stress (Miltz et al., 1978). The same results have been obtained for PMMA (Wang et al., 1971) and PS (Sternstein et al., 1963). From these results, it seems sufficient to simply conclude that crazes grow in the direction normal to the maximum principal stress. However, in the case of isotropic materials, since the maximum principal
Isotropic
65 0
90
Figure 15-25. The effect of orientation on the craze growth direction with respect to the stress axis (Beardmore and Rabinowitz, 1975). a is the angle between the extension direction and the craze plane, 6 is the angle between the orientation direction and the extension direction.
724
15 Crazing and Fracture of Polymers
Figure 15-26. A craze in PC which has ceased to grow due to the formation of a pair of shear bands ahead of it (Narisawa, 1982).
The time dependence of the increase in thickness of surface crazes in PC under a constant load is shown in Fig. 15-27 (Verheulpen-Heymans, 1979). Two possible mechanisms for the growth of crazes in the thickness direction have been proposed: (1) creep of craze fibrils, and (2) gradual transformation of the uncrazed material in the interfacial area into fibrils. Although the kinetics suggests that the creep of craze fibrils is dominant (Verheulpen-Heymans
and Bauwens, 1976 b), electron microscopy observations of craze structure favor the second mechanism for both crazes on the surface and at the crack tip (Chan et al., 1981; Donald et al., 1981; Lauterwasser and Kramer, 1979; Trent et al., 1981). In other words, the fibrils in a craze are stretched at first, stabilized by the entanglements; this process, however, will become more difficult because of the strainhardening effect induced by the advanced
MPa
A
+ & + A
o DA DA
o log t (h)
Figure 15-27. Time dependence of thickness d of surface crazes in PC formed under different conditions (VerheulpenHeymans, 1979). ( + ) No treatment; (A) annealed, (o) thickness after unloading.
725
15.2 Crazing: Initiation and Growth
orientation. As a result, the material at the two ends of the fibrils must be drawn out to relax the stress until it eventually approaches a new equilibrium. Figure 15-28 shows the extension ratio of the craze fibrils at a crack tip in PS and the variation of stress there. In this figure, the increase in the thickness of the craze away from the crack tip is not accompanied by a concomitant increase in the extension ratio. This is consistent with the entanglement model proposed by Kramer and his coworkers (Kramer, 1983). The growth of surface crazes in the longitudinal direction has been reported to be proportional to time (case 1), or proportional to the logarithm of time (case 2). For the surface crazes of PS at room temperature in case 1, the observation results indicate that, except in the very initial period, the growth rate of the crazes can be expressed by (Sauer and Hsiao, 1953) m
(15-18)
where m is a constant, and a0 is the critical stress for craze growth and has a value close to that of the crazing stress. The growth behavior of crazes has also been investigated in a microscopic view for crazing in PMMA and PS at room temperature or lower (Argon and Salama, 1977). In case 2, the notion that craze length is proportional to the logarithm of time can be expressed as = fclogt or d/c
k
df
t
(15-19)
(15-20)
Equation (15-20) shows that the growth rate decreases with time. This kind of behavior has been observed in PMMA (Regel, 1956) and PC which has been plas-
Craze front m
200
150
100 Craze rear (crack tip)
50
0
1
2
3 4 Extension ratio
5
6
Figure 15-28. The extension ratio of craze fibrils at a crack tip in PS and their associated fibril stresses (Lauterwasser and Kramer, 1977).
ticized to various extents (Sato, 1966; Verheulpen-Heymans and Bauwens, 1976 a). k is a constant that depends on the polymer species, the extent of plasticization, and the stress as well as the temperature. According to Kambour (1973), the fact that the growth rate of crazes decreases with time is due to creep of the overall specimen during the growth of crazes. That is, assuming that the growth rate of crazes is inversely proportional to the increase of the creep strain, we have k
(15-21) +e where k is a constant which depends on stress only, and s is dependent on both the stress and time. If we denote k' and e by k = fc'V, and e = Aea°t", Eq. (15-21) becomes df
d/.
dt
1+Ae«atn
(15-22)
where A, a, and n are constants obtained from creep experiments. Since the approximation log(l + Ad1" f)~logt can be made over a wide range of temperatures,
726
15 Crazing and Fracture of Polymers
Eq. (15-22) can be reduced to Eq. (15-21). Indeed, the reason for the growth rate of crazes being inversely proportional to the loading time is that most experiments carried out so far have been either on materials which exhibit significant creep deformation or in conditions under which creep deformation easily occurs. The growth of crazes at a crack tip can be treated by using the stress intensity factor as a parameter since the stress at the crack tip is the driving force. Specifically, for the growth of crazes in environmental liquids, a critical Kc exists below which the growth of crazes will cease. From many experimental results, the length of the crazes in this case (ignoring the very initial period) can be expressed as follows (Miltz et al., 1978; Marshall et al, 1970; Narisawa and Kondo, 1973; Kitagawa, 1972; Kitagawa and Motomura, 1974,1976; Andrews and Levy, 1974; El-Hakeem et al., 1975): lc = A tn>
60
90 120 Time (min)
180
150
(a)
(15-23)
where n' is a constant that is independent of the stress intensity factor, n' ~ 0.5 in most cases for reasons given in the following sections and is related to the mechanism for the growth of crazes.
30
60
90 120 Time (min)
150
180
(b)
15.2.4.2 The Growth of Crazes at a Crack Tip
Consider the growth of crazes which occurs at a crack tip in a liquidlike environmental agent. In this case, growth of the crazes as a whole is governed by the stress intensity factor Ko = a sjnao due to an initial crack with length aQ. The growth curves for the crazes in PMMA and PC in kerosene are presented in (a) and (b) of Fig. 15-29, respectively. When KO>KC, growth continues until the specimen fails. Marshall and co-workers (Marshall et al., 1970) proposed a simple mechanism for the
Figure 15-29. Craze length vs. time for (a) PMMA (b) PC exposed to kerosene and subjected to various levels of stress intensity factor (Narisawa and Kondo, 1973).
growth of crazes due to an environmental fluid in terms of fracture mechanics:
= AknJt
(15-24)
where A is a constant which depends on the viscosity of the fluid and the stress required for craze initiation. The driving force is considered by them to be atmospheric pressure.
15.2 Crazing: Initiation and Growth
Equation (15-24) explains why n' = 0.5 in Eq. (15-23), that is, why the flux rate of the environmental agent determines the growth of crazes. Another mechanism, which resembles that described above, has been proposed. It states that the flow of the environmental agent into the crazes is driven by the much higher capillary pressure arising from the size of the craze voids (Kramer and Bubeck, 1978; Kambour and Yee, 1981). The pressure gradient due to capillary pressure is about 100 times larger than one atmosphere. Consequently, the flow of the environmental agent into the tip of crazes through the voids could be much faster.
727
(a)
(b)
15.2.4.3 Microscopic Mechanisms of Crazing
The ability of fluids to flow into the tips of crazes must be because the voids are interconnected. Two-dimensional electron microscopy observations sometimes give the impression that a craze grows by nucleating single voids at its tip. The reason for the voiding was proposed to be the lowering of the glass transition temperature of the region around the craze tip to a level below the testing temperature due to the dilative stress (Gent, 1970,1973). However, the model proposed by Argon et al. (Argon and Salama, 1976, 1977; see also Vol. 6, Chap. 10 of this Series) shown schematically in Fig. 15-30 provides a more reasonable explanation for the formation of the characteristic structure of crazes and their growth. The basis of the Argon model is the phenomenon known as meniscus instability (Saffman and Taylor, 1958; Fields and Ashby, 1976). It is observed when a regular arced surface is created between air and liquid, or when two rigid plates with a liquid in between are pulled apart, or when a
(c)
(d)
Figure 15-30. A model of craze growth based on meniscus instability (Argon and Salama, 1977).
liquid climbs up a capillary tube. In Argon's model, instead of the liquid, the material at the craze tip is considered to be plastically deformed. Because the extent of the instability is related to the surface energy of the interfaces, when an environmental liquid is
728
15 Crazing and Fracture of Polymers
allowed to flow into the craze in place of air, the surface energy of the interface is lowered and this effectively accelerates the meniscus instability. The mechanism proposed by Argon can therefore explain the environmental effect quite well. The length period of these arcs, /, is determined by the viscosity and the surface energy in the case of liquids, and by the viscoelastic characteristics of the unfibrillated areas and the surface energy of the interfaces in the case of crazes. / at the craze tip is in fact the gap between the fibrils, which, according to Argon's calculation, is about 140 nm with fibrils of 50 nm diameter for PMMA and PS. Electron microscopy observation of the craze tip in PS with specimens of varied thicknesses by Kramer and co-workers (Chan etal., 1981; Donald and Kramer, 1981) found no independent voids at the craze tip, i.e., all the voids are interconnected. Moreover, the latter authors actually observed the fingerlike structure of voids at the craze tip as predicted by the meniscus instability model. Such results led Kramer to conclude that Argon's microscopic model is probably the most reasonable mechanism for the growth of crazes (Kramer, 1983). 15.2.5 Mechanical Properties of Crazes 15.2.5.1 Stress-Strain Behavior The reason why a craze, even though it is about 50% in void content, can still carry load is explained by the fact that the fibrils, which are oriented in the loading direction, can bear stress. Figure 15-31 shows the stress-strain relationship of crazes in PC measured after the specimen has been removed from ethanol (Kambour and Kopp, 1969). The result indicates that the longitudinal modulus of craze fibrils is almost the same as that of uncrazed fibrils.
0
k
12 16 20 2k 28 32 36 40 U 48 52 Strain (%)
Figure 15-31. Stress-strain behavior of crazes in PC formed after exposure to ethanol, then removed. The stress-strain behavior of the uncrazed material is also shown for comparison. (Kambour and Kopp, 1969.)
Moreover, the stress-strain behavior undergoes a yieldlike process after which the strain shows a drastic jump. As can be seen from the loading and unloading curves, the fibrils in crazes have much higher ductility than the uncrazed materials. The fact that the hysteretic loss is larger in crazes than in uncrazed materials should also be noted. A similar stress-strain curve has also been obtained for the crazes in PS (Hoarse and Hull, 1972). In the measurement of the stress-strain relationship of crazes in PC, we note, however, that the removal of ethanol from crazes will cause them to shrink, and plasticizing effects will definitely be involved in the crazes generated under the influence of environmental liquids. Therefore, the results obtained from this experiment do not necessarily reflect the real mechanical behavior of the craze as it goes through the processes of growth and fracture, especially in the absence of an environmental liquid.
15.2 Crazing: Initiation and Growth Monochromatic light
15.2.5.2 Micromechanics of a Craze
One of the methods for measuring the displacement of crazes at the crack tip is the interference pattern method, which is shown in Fig. 15-32. The distribution of displacement of the crazes at the crack tip obtained from the interference pattern measurement is given in Fig. 15-33 for PMMA (Brown and Ward, 1973; Morgan and Ward, 1977; Frazer and Ward, 1978; Weidman and Doll, 1978; Israel et al, 1979; Andrews and Levy, 1974; Young and Beaumont, 1976; Shirrer and Goett, 1980). Similar data have been obtained for PS (Kreuz et al., 1976), in HIPS and ABS (Hoffman and Richmond, 1976; Newmann and Williams, 1980), and PC (Pitman and Ward, 1979; Kambour and Miller, 1976). The patterns in these materials are quite consistent with the prediction from Dugdale's model. An extremely sensitive method - laser holographic interferometry - has been used to directly measure the strain distribution, Ae^, around crazes (Peterson et al., 1974; Krenz et al, 1976). Figure 15-34 illustrates the distribution of strain in a craze at a crack tip in PS and the corresponding stress distribution calculated from the relation oy = E sy. This figure shows crazing in (a) methanol, and (b) rc-heptane. One can see that the shapes of the two distributions are quite different. In the case of rc-heptane, the stress decreases sharply to a value comparable to that in the vicinity of the crack tip. This means that n-heptane has a significant plasticizing effect on the craze fibrils in PS. The stress distribution in this case is also quite different from that predicted by Dugdale's model. A contrasting situation exists for methanol wherein the stress distribution obtained deviates only slightly from that predicted by Dugdale's model. For this reason, the stress at the craze tip
729
Reflected light
tn)
Figure 15-32. (a) A schematic of the interference technique used to measure the thickness of crazes, w is the thickness of the craze at the position of the beam drawn in the figure, (b) An interference fringe pattern of a craze at a crack tip in PMMA. (Doyle, 1975.)
1.00
0.80
0.60
0.40
0.20
0
0
0.20
0.40
0.60
0.80
1.00
Figure 15-33. The normalized distribution in a PMMA craze determined from interference fringe data such as those shown in Fig. 15-32. The abscissa is the position X along the craze normalized with respect to the length of the craze lc. The ordinate is the thickness of the craze W (x) at position X normalized with respect to the width at the crack tip Ao. (Morgan and Ward, 1977.)
730
15 Crazing and Fracture of Polymers
culated stress compares well with that from a finite elements method (Bevan, 1982), although the latter cannot yet give as detailed a profile as from the technique used by Kramer and co-workers. These results show that there is a significant increase in stress at the tip of the craze, depending on the extension of the interconnected fibrils within the craze.
(a) 10 x 10"3
-Before crazing
Q
~ m-3
*•> After crazing
10
2 x 10-
H«-Cra
0 /A
i "0 02 ( U 0 6 08 Distance from crack tip (mm)
(b)
After crazing—1 / •
15.2.6 Craze Thickening and Fracture
\
\
\
-Before crazing /
7 x 10"'
\
Crazes eventually transform into cracks, given sufficient time and stress. To understand this transformation process, we must focus our attention on how the craze thickens and eventually breaks down. Some phenomenological aspects of longitudinal craze growth have been presented in a previous section. We discuss here the phenomenology of craze thickening specifically in the context of their behavior just prior to their transformation into cracks.
10" 0
06 1.2 1.8 IX Distance from crack tip (mm)
Figure 15-34. The distribution of strain in a craze at a crack tip in PS and the corresponding stress distribution. Crazing occurred in (a) methanol and (b) nheptane. (Krenz et al., 1976.)
has to be divided into two steps to calculate the distribution of the displacement; as a result, a modified Dugdale's model has been proposed (Graham et al., 1976). The stress distribution along a single craze can be calculated using either the Fourier transform procedure (Lauterwasser and Kramer, 1979) or the distributed dislocation method provided that the strain distribution along the craze can be determined. Kramer and co-workers were successful in this endeavor using TEM (Lauterwasser and Kramer, 1979). The cal-
15.2.6.1 Kramer's Model for Craze Thickening
A recent model proposed by Kramer and co-workers on craze growth is presented here because of its direct relationship to craze failure and its ability to explain ductile-brittle transitions. A schematic of Kramer's model is shown in Fig. 15-35. In this model, the craze thickens by drawing molecular matter from a certain layer h at the craze interface into the fibrils. The value of the width h can be determined from TEM of the thin polymer film decorated with gold particles, the average distance between which is changed by deformation. It was observed that the active zone h is quite narrow, on the order of the fibril diameters, and is dependent on the local strain rate and the temperature. This experiment further reveals that the stress transforms the material in the thin
15.2 Crazing: Initiation and Growth
731
polymer glass;
Figure 15-35. A schematic drawing of the craze-bulk polymer interface in Kramer's model. The hydrostatic tensions (cro)s and (cro)m form a pressure gradient. (Kramer and Berger, 1990.)
zone h so that it "flows" into the fibrils. While the craze as a whole thickens, the volume fraction of fibrils, vf, remains a constant, hence the extension ratio X also remains constant. Starting from this model, a detailed calculation involving assumptions of power-law fluid behavior, pressure gradients between the top of the fibril and the top of the void surface, and the surface tension there yields a craze fibril flow stress, i.e., the crazing stress of the polymer, S, which is approximately proportional to the square root of the craze surface tension per unit width of the active zone, F/h, viz., 05-25) and = y+ ^
v = v e = vx
(15-26)
where y is the van der Waals surface energy and U is the backbone covalent bond energy, and vx is the density of chemical cross-links. C" is a material dependent parameter that is sensitive to temperature and strain rate. This model allows us to understand the role played by the entanglement density on the crazing stress. Increasing v raises the crazing stress. If craze growth involves chain scission, e.g., when a strand crosses an interface, then the second term becomes important. On the other hand, if the kinetics of disentanglement is very rapid and there are no chemical crosslinks, then the surface tension is entirely
due to the van der Waals force involved in disentanglement, viz., F=y
(15-27)
Kramer calls this "van der Waals crazing". Note that since the term F/h is the energy density near the craze surface, it follows that its square root is the solubility parameter. Equation (15-27) then relates the crazing stress to the solubility parameter of the polymer. A solvent which reduces the van der Waals term would reduce the stress required for crazing to occur. This is, of course, consistent with experimental observations of solvent crazing. An important consequence of differentiating between scission dominated crazing and van der Waals crazing is that it allows us to understand the brittle-to-ductile transitions when chemical cross-link densities are increased and when the temperature is increased. Surprisingly, these concepts also allow us to understand ductileto-brittle transitions with increasing temperature observed in certain polymers. 15.2.6.2 Ductile-Brittle Transitions
Brittle to ductile transitions with crosslink density in the case of normally brittle thermoplastics are easy to explain in view of the model presented above. Some results collected by Kramer's group are presented in Fig. 15-36 to illustrate such transitions. We note that care should be taken in extrapolating such results to thermosets which begin as monomers and eventually
15 Crazing and Fracture of Polymers
732
1 PTBS PSMLA 3 PVT 4 PAMSS 5 PSAN1 6 PMMA 7 PSAN1 8 PSAN2 9 PPO 10 PC CVJ
1.5
_ \ E
1.0
[3 T * Crazes
0.5 Crazes-
; Crazes DZ's
0
1
1
10 I/U10 2 5
20 chains/m3)
form networks with very high cross-link densities without going through the thermoplastic stage. At very high cross-link densities, such thermosets are usually quite brittle. Kramer's results should then be taken to mean that, for low entanglement density polymers, strengthening the network with chemical cross-links will shift the deformation mode to shear yielding. However, if the cross-link density is sufficiently high, then shear yielding once again becomes difficult, and brittle fracture without crazing occurs. In this case, chain scission clearly dominates the fracture mechanism. The effect of temperature on ductilebrittle transitions is somewhat more complex. One normally expects a polymer to shear yield more easily at temperatures near the Tg. Thus, a polymer that is brittle at temperatures well below Tg would undergo a brittle-to-ductile transition as the Tg is approached. In such a polymer, the low temperature craze growth should involve a certain amount of chain scission,
#-DZ'S
Figure 15-36. The true strain ratio £R in crazes and deformation zone (DZs) in various polymers and copolymers plotted against the network strand density v = ve 4- vx. The open triangles and hexagons represent blends of PS and PPO. The filled triangles and circles represent crosslinked PS. (Henkee and Kramer, 1984.)
30
while at high temperatures, disentanglement should dominate. This situation is illustrated by the experiments of Plummer and Donald (1989). In their experiment, crazing stresses for previously undeformed and previously crazed but "healed" (at 130°C for lOmin) PS having molecular weights of 260000 and 1 150000 were determined. The ratio of these two stresses is plotted against the temperature in Fig. 15-37. To explain their results, Plummer and Donald argue that at low temperatures, the initial crazing causes a significant amount of chain scission in the craze, and the subsequent healing is insufficient to reestablish the network. Crazing of the healed glass therefore involves only the van der Waals term. Hence, the initial crazing stress is higher. At high temperatures, crazing is primarily by disentanglement, therefore the deformation history makes little difference to the crazing stress. For the higher molecular weight PS, the scissiondominated crazing continues to higher temperatures. At these high temperatures,
15.2 Crazing: Initiation and Growth
shear yielding becomes relatively easy, therefore a transition to this mode of deformation occurs. In the case of the low molecular weight PS, a transition to disentanglement crazing requiring lower stresses occurs at low temperatures; therefore, a transition to shear yielding does not occur. Some ductile glassy polymers have been reported to exhibit a reverse transition with temperature, i.e., undergo crazing as their Tg is approached (Wellinghoff and Baer, 1978). On the basis of the entanglement model described above, Donald and Kramer (1982) suggested that this phenomenon is due to a transition to disentanglement-dominated crazing at high temperatures. At lower temperatures, crazing would require stresses sufficiently high to cause extensive chain scission, therefore shear yielding would be preferred. This explanation was given strong support by a series of experiments by Plummer and Donald (1989). Their results on poly(ether sulfone) (PES) illustrate these ideas beautifully (Fig. 15-38). In Fig. 15-38 a, the tensile strain and the corresponding deformation mechanisms are plotted against tempera1.6 1.5
_, n.
U
8
O
r 1.3 1.2 -
\
1.1 1.0 20
i
30
40
50
60
70
80
90
100
rra Figure 15-37. A plot of the ratio r of the initial crazing stress to the crazing stress for van der Waals crazing as a function of temperature for PS having Mn of 260000 (•) and 1115 000 (o) (Plummer and Donald, 1989).
733
ture for two PES resins with molecular weights of 47000 and 69000. In Fig. 15-38b, the corresponding craze fibril extension ratios are plotted. At low temperatures, the thin films deform by forming deformation zones (DZs), which is an indication of ductile behavior in the bulk state. The DZs form instead of crazes because the crazing stress is too high - a result of high entanglement density. At higher temperatures, disentanglement becomes easier, and eventually the crazing stress becomes lower than the yielding stress, and brittleness results. The temperature at which the transition to disentanglement crazing occurs increases with the molecular weight because chain-scission dominated crazing persists to higher temperatures in this situation. Similar results have been obtained by this group on PC. 15.2.6.3 Craze Fracture Process
Usually, in the tension-induced fracture of polymers, especially those that craze easily, a number of crazes already exist on the surface of the specimen before catastrophic failure occurs. Fracture then begins with one of these crazes. Two fracture initiation mechanisms are possible in such a situation: One is that fibril breakage inside a craze occurs and a void forms; the void thus formed then expands by joining up with other voids, leading to fracture; another is that the craze itself acts as the cause for creating the stress concentration around it, just like surface flaws and cracks. This is considered to be an initiation mechanism for relatively brittle fractures that occur at low temperatures. It has generally been believed that fibril breakage occurs in the midrib of the craze. However, recent TEM observations by Kramer and co-workers (Yang et al., 1986 a; Yang etal, 1986 b; Kramer and Berger,
734
15 Crazing and Fracture of Polymers
0.05
20
40
60
80
100
120
140
160
180
200
Figure 15-38. (a) Tensile strain at the onset of plastic deformation is PES as a function of temperature for samples with an Mn of 47 000 (o) and 69000 (•). (b) The craze extension ratios corresponding to the specimens in (a). (Kramer and Berger, 1990.) 20
(b)
40
60
80
100 120 T(°C)
140
160
1990) show that without exception, all fibril breakdowns initiate at the craze-bulk polymer interface, i.e., at the active zone /z, whether or not it is initiated by a flaw (Fig. 15-39). In these TEM micrographs, the midribs are clearly below the lower end of the pear-shaped voids and hence are not involved in the initial breakdown process. Such pear-shaped voids have also been observed in bulk specimens (Behan et al., 1975; Hull, 1970). Similar breakdown morphologies have been observed in other vinyl polymers such as PMMA, poly(a-methyl styrene) (PaMS), and poly(styrene acrylonitrile) PS AN (Kramer and Berger, 1990).
180
200
With the exception of these recent observations, the craze breakdown process has been described by Murray and Hull (1969, 1970 a, b, c). Their observations are modified here to reflect the results of Kramer and co-workers, and summarized as follows: (1) As the craze grows in the thickness direction, fibril breakage initiate at the craze-bulk interface, and expand to form larger voids. The initiation usually occurs around heterogeneities in the craze such as foreign matter and impurities. The initiation stress can be described by Weibull statistics.
15.2 Crazing: Initiation and Growth
(b) Figure 15-39. TEM micrographs of craze fibril breakdown in the absence of dust inclusion (a) and associated with a dust inclusion (b) (Kramer and Berger, 1990).
735
(2) Hach of these voids initially grows independently, but they eventually impinge on each other. Finally, a crack is formed in a direction normal to the stress. (3) Once the cracks reach a certain size, they pass through the midrib or the interfacial regions between crazed and uncrazed materials along the craze plane and propagate by breaking more and more fibrils. (4) Several cracks, which have undergone the above three steps, become connected to each other, and this leads to macroscopic fracture. Ductile fracture of craze fibrils can also occur when the time between crazing and fracture initiation is long enough to allow the craze fibrils to be plasticized by environmental agents, which could facilitate the disentanglement process. Likewise, high temperatures could facilitate the disentanglement process. Although most of the observations described above have been made on carefully crazed specimens and then fractured in quasistatic loading or at low strain rates, in more practical loading conditions crazing
736
15 Crazing and Fracture of Polymers
always precedes cracking as long as the cross-link density is not too high. This can be identified by the interference pattern found at the propagating crack tip or the interference color on the fracture surface (Murray and Hull, 1970 a; Graham et al., 1976; Kambour, 1965). The green and pink interference color is a characteristic indication that the fracture surface is covered by a thin broken craze layer which has a lower refractive index (Higuchi, 1958) than that of the uncrazed material (Kambour, 1964 a, b). Even in dynamic fracture, crazing is always found (Takahashi et al., 1978). For different polymers the thickness of the craze layer on the fracture surface varies measurably with the crack propagating velocity and temperature, and therefore the fracture surface often exhibits complicated features. The evidences presented above demonstrate that, apparently without exception, for thermoplastic glassy polymers over a wide range of temperatures, crazes are always created at the crack tip and the crack then propagates through the body of the craze.
15.3 Fracture Polymeric materials can be classified according to their fracture type and characteristics into four categories: amorphous polymers, crystalline polymers, highly oriented polymers such as fibers, and elastomeric materials such as rubber. In terms of the macroscopic distribution of internal stress and strain, polymer solids in the glassy state are more homogeneous compared with those possessing crystallinity. It is often convenient to regard crystalline polymers as a two-phase system wherein two components of different mechanical properties are appropriately combined. In this case, fracture will always oc-
cur in the weakest phase. The fracture of fibers, however, can be considered essentially elastic because of the high degree of anisotropy and the small capacity for plastic deformation. In the fracture of elastomers, the strain to failure is considerably larger, and a particular relationship exists between the strain to failure and the strength at that point. This distinguishes the fracture of elastomers from that of any other material. In this chapter, we shall first review the microscopic mechanisms of polymer fracture. In later sections, we shall focus on the fracture of materials with a crack or a flaw in terms of fracture mechanics. 15.3.1 Theoretical Strength of Polymers
There are a number of observations that help to elucidate the microscopic mechanisms that cause polymers to fracture. (A) Whether a polymer exhibits brittle or ductile behavior depends on how the molecular chains respond to the external force, either in concert or individually. The molecular response of a polymer to stress is determined by its chemical (primary) structure (molecular species, molecular weight, molecular weight distribution, whether crystalline or amorphous, cross-linking, entanglement density, etc.), secondary structure (crystallinity, morphology and the dimensions of crystals, orientation, etc.), the state of the material (with respect to the glass transition temperature and other secondary or higher transition temperatures), and conditions of the external force (the type of loading, loading rate, etc.). (B) When a material in which the molecular chains are randomly oriented is subjected to a tensile stress, the chains tend to orient along the stress direc-
15.3 Fracture
tion as they are drawn. In this process, chain slippage and separation occur preferentially to chain scission since the covalent bonds which constitute the backbone chain are much stronger than those due to intermolecular interactions such as van der Waal's force. (C) Crystals and chemical cross-linking prevent the molecular chains from slippage. At temperatures below the glass transition temperature, physical entanglements among molecular chains have a similar, but smaller, effect than that of cross-links, even though they are not permanent bonds that can prevent chain slippage. For this reason, higher stress is needed to induce sufficient chain slippage in the glassy state than in the rubbery or melt states. (D) When the fraction of crystals and cross-links in a material is large, it will be difficult for chain slippage to occur and the molecular chains will be subjected to tension elastically from the onset of deformation. Some of the chains may begin to break. A similar situation occurs when drawn materials are subjected to an external force in the direction of drawing. Such materials usually exhibit macroscopically brittle fracture. (E) In those materials where chain slippage is a main cause of fracture and chain scission is a minor effect, either crazing, where the deformation is usually localized, or plastic (shear) deformation, where the deformation often involves a substantial portion of the sample, will occur. Macroscopically, these materials are ductile. Even in the case of plastic deformation, only those molecular chains which have been oriented in advance will bear the external force. Therefore, a nonuniform stress
737
distribution on a molecular scale has already been generated in the material at this point, what we shall call the first microscopic stress concentration. Similarly, the molecular chains in crystalline and amorphous phases also differ in the amounts of stress they bear. (F) When the stress is further increased, some of the chains that are under tension will begin to break. This chain scission is generally regarded as a stress-dependent kinetic process in which the rate R can be expressed by AF — a a = co0 exp —
kT
(15-28)
where AF is the activation energy, a is the activation volume which includes the stress concentration factor, a is the external stress, k is Boltzmann's constant, T is the absolute temperature, and co0 is a constant. (G) Upon further loading, more and more "newly" stretched chains break as chain scission progresses. When the number of broken chains reaches a certain density, a microvoid can be formed; consequently, more slippage occurs because of freer molecular motion around the void. This process will further accelerate chain scission or will lead to the merging of the voids, which eventually generates a crack. We shall call stress concentration due to microvoids the secondary microscopic stress concentration. (H) To initiate macroscopic fracture, a single large crack is sufficient. The crack will propagate by breaking either molecular chains or intermolecular interactions at its tip. The microvoids that have already formed in (G) can render crack propagation easier. (I) If there exist some foreign particles or impurities that lead to macroscopic
738
15 Crazing and Fracture of Polymers
stress concentrations in the material, processes (E) through (G) will proceed around them locally. If there is any crazing, the deformation will be concentrated in the crazes, in which case processes (E) through (G) will occur inside the crazes. 15.3.2 Fracture Initiation in Uncracked Polymer Specimens (Micromechanism of Fracture) 15.3.2.1 Micro void Generation due to Tension
In polymers, even if there are no cracktype defects such as voids and flaws generated during processing or service, inhomogeneities in the structure will still cause microscopic stress concentration and could lead to the formation of Griffith-type defects. This will give rise to a considerable difference between the theoretical and the practical strengths. For specific examples of structural inhomogeneities, one can cite the disordered lamellar crystals which constitute the microcrystals or spherulites, the amorphous region between lamellae and between lamellar bundles, and the interface between spherulites, etc. Other examples are found in glassy polymers such as PS and PMMA, etc., which often exhibit brittle fracture. The strength increase in these materials will eventually disappear as the size of the intentionally induced crack is decreased; this can be regarded as evidence for the existence of an inherent flaw in the material (Berry, 1961 a, b, c). The size of this inherent flaw calculated from Griffith's theory depends on the species of amorphous polymer, but it is usually in the range of 0.1-1.0 mm, which is unreasonably large. Because inherent flaws of this size have never been observed, stress concentration due to structural inhomogeneities (much
smaller in scale than the inherent flaws) must be present, which could lead to and accelerate the initiation and growth of larger defects such as crazes, etc., around them. Defects grown in this way would act as a quasicrack, and could eventually lead to macroscopic brittle fracture. This hypothesis explains why, in the brittle fracture of glassy polymers, it is difficult to achieve the theoretical strength since crazing is generally involved. In polymers that do not exhibit brittle fracture at the initial stages of deformation, molecular chains in locations where microscopic stress concentrations exist due to structural inhomogeneities probably rearrange themselves to relax the stress. Once this process ceases and plastic deformation has proceeded to a sufficient extent, microvoids generated from chain scission due to the structural flaws will grow and eventually lead to the catastrophic failure of the material as described previously. Microscopic voids generated during the initial stages of failure were first discovered by Zhurkov and co-workers (Zhurkov and Kuksenko, 1965; Zhurkov et al., 1969 a). They investigated void generation during deformations such as simple tension and creep for various oriented polymers, and estimated the number and size of the microvoids as a function of load or time. The results in Fig. 15-40 show that the number of microvoids in nylon-6 increases sharply as the load or time is increased. 15.3.2.2 Chain Scission in Oriented Polymers (Fibers)
If no chain slippage occurs, microvoids are generated by chain scission. The generation of these microvoids plays an important role in fracture initiation, particularly for fibers in which the molecular chains have been well-oriented with little room
15.3 Fracture
739
Figure 15-41. Fiber structure model in terms of crystal lamellae and amorphous regions (Peterlin, 1972). A, B are tie-molecules. 0.5 1.0 Creep t (min) 0
20 40 Constant strain rate a (kg/mm2)
0
5 10 Constant load rate £ (%)
3
Figure 15-40. The number density of micro voids generated in nylon-6 under various loading conditions.
for plastic deformation. Several fine structure models for fibers have been proposed so far. We shall consider in the following the model proposed by Peterlin (Peterlin, 1969, 1971a, b, 1972, 1974 b). The fiber structural models proposed by Peterlin for PE and PA-6 are shown in Figs. 15-41 and 15-42. In Fig. 15-41 a fibrous structure is depicted as an alternating combination of crystal lamellae composed of folded chains with orientation in the fiber axis and tie-chain regions. The tiechain regions incorporate chain ends and loops that are not included in the crystals and taut chains that connect the crystals. In other words, the amorphous area is repeated alternatively along the fiber axis to form a number of microfibrils with widths of 10-20 nm and lengths of 5-20 |im. These microfibrils will further aggregate to
Figure 15-42. Microfibrils and the associated fiber structure model (Peterlin, 1974 b).
form fibrils as depicted in Fig. 15-42 with widths of 0.1-1.0 |im. The stress in the fiber direction will be transmitted through tie-chains, which are thought to comprise 10-30% of the number of chains in the lamellae. Tie-chains, therefore, are subjected to quite a large force. The tie-chains between microfibrils are much fewer, and in PE amount to about 10% by volume
740
15 Crazing and Fracture of Polymers
(Meinel and Peterlin, 1968; Meinel et al, 1968). In PA-6, the fraction of tie-chains is also about 10%. However, this number generally depends on the polymer species as well as the type of processing that the material has undergone (Peterlin, 1974 a). The stress is distributed through microfibrils to the serially connected sandwich structures consisting of the lamellaamorphous regions. The amount of deformation in the lamellae is negligible compared with that in the amorphous regions. Let x be the crystallinity. When an overall strain s is applied, the strain of amorphous regions sa can be expressed by a
(15-29)
1-x
For example, if we suppose that the crystallinity is 50%, a strain twice as large as that of the overall strain will be applied to amorphous regions, according to Eq. (15-29). Since one cannot expect that all the tiechains have the same length, it is easy to imagine that shorter chains will be tightly stretched while longer ones are almost unstrained. In other words, even if the strain in the amorphous regions is constant, a distribution can still form within the region due to the distribution of lengths of initial tie-chains. In fact, Zhurkov (Zhurkov et al, 1969 a) has demonstrated that a stress ten times as large as the average stress is applied to some tie-chains in PE and PP. Therefore, if the strain applied to the amor-
a
b
• t
'/I •
B
phous regions increases, as shown in Fig. 15-43, some of the tie-chains will begin to break (Peterlin, 1972). The chain scission will generate free radicals which can be detected by ESR. Since Zhurkov's first reports (Zhurkov and Kuksenko, 1965; Zhurkov et al., 1964; Zhurkov and Tomashevskii, 1966), the generation of free radicals under tension has been extensively investigated by Peterlin and co-workers (Campbell and Peterlin, 1968; Verma and Peterlin, 1970), Becht and co-workers (Becht and Fischer, 1969; 1970), Kausch and co-workers (Kausch and Becht, 1970; Kausch, 1970) and Takayanagi and coworkers (Nagamura et al., 1973; Nagamura and Takayanagi, 1974). The following is a summary of the results: (1) Free radicals are not generated immediately after the application of tension. They appear upon reaching a stress level that is about 40-60% of the fracture stress, and increase sharply thereafter (DeVries et al, 1971) (Fig. 15-44). (2) In creep, free radicals increase sharply immediately after loading. At low stresses the rate of free radical generation is almost constant, but increase again up to near fracture at high stresses (DeVries et al, 1971) (Fig. 15-45). (3) In cyclic tension, free radicals can be accumulated (DeVries et al, 1971) (Fig. 15-46). (4) In step-wise tension, free radicals can vary with the step, and as a result, a
Figure 15-43. Schematic representation of interlamellar chain scission which proceeds according to the length distribution of tie chains (Peterlin, 1972). a, b, and c are undeformed, slightly deformed, and greatly deformed amorphous regions, respectively. A, B, and C are unstressed, stressed, and ruptured tie chains, respectively.
15.3 Fracture
741
Stress i
Radicals Time
Figure 15-46. Free radical accumulation due to cyclic loading (DeVries et al., 1971).
Time
Figure 15-44. The time dependence of stress and strain and the number of free radicals generated (DeVries et al., 1971).
-Radicals -Stress
High stress Strain
Radicals
Time
Strain
Figure 15-45. Radical generation due to creep (DeVries et al., 1971).
Figure 15-47. Frequency distribution of free radical generation due to stepwise tension (DeVries et al., 1971).
frequency distribution of free radical generation corresponding to strain can be obtained (DeVries et al, 1971) (Fig. 15-47). By combining the fiber structure with the occurrence of chain scission, the microscopic mechanism for fracture initiation can be modeled in two ways. Model I: DeVries et al. attempted to quantitatively relate the characteristics of the free radical generation described above to the stress-strain curve of PA-6 by changing the temperature and rate of deformation under various loading con-
ditions (constant stress rate, constant strain rate, creep, and cyclic deformation) (DeVries et al, 1971; Lloyd et al, 1972; DeVries and Farris, 1970; DeVries, 1971). By assuming that amorphous regions in the microfibrillar structure deform uniformly, and that the results illustrated in Fig. 15-48 reflect the distribution of lengths of tie-chains in the interlamellar regions, they proposed that the stress generated when a certain strain is applied is equal to the sum total of the stresses to which each stretched tie-chain is subjected. They considered the breakage of an individual tie
742
15 Crazing and Fracture of Polymers
Figure 15-48. The strain (a/E), the number of free radicals (R/R^) and the rate of generation of free radicals (dR/dX) as functions of amorphous region extension ratio X when a normal distribution of tiechains (W/l) is assumed.
chain to occur as a result of a thermally activated stochastic process before it reaches its critical strain. The length distribution of tie-chains obtained by ESR is further approximated by a normal distribution (with a standard deviation of 0.01 0.02); when an arbitrary strain is applied to the amorphous region, a distribution of strain corresponding to that of the lengths of the tie-chains is assumed. In practical calculations, the continuous distribution of strain is obtained by splitting it up into discontinuous blocks and taking the summation. The calculated results are in good agreement with experimental results with the exception of the initial period of tension. Similarly, Kausch et al. (1972) also assumed that stress is uniformly transmitted in amorphous regions in a microfibril.
However, they found that the detected radicals due to chain scission could account for only about 0.3% of all the tie-chains. Of the remainder, only 1.1% were stretched but unbroken, and the other 98.6% of the tie-chains were only slightly stretched. They concluded that a major part of the stress is borne by only a small fraction of the tie-chains. Takayanagi and co-workers (Nagamura etal., 1975) also carried out a calculation similar to that done by DeVries but for poly[p-(2-hydroxy ethoxy) benzoic acid] (PEOB) fibers. The difference between the two calculations is that the effect of a crystal-crystal phase transition (which occurs during tension), the dangling chain ends, the broken chains, or any unstrained chains were all taken into consideration in the case of PEOB. Since tie-chains were assumed to break at their critical strains, the effect of time was ignored. Moreover, all the above effects were added independently. The calculated stress-strain curve was found to be quite consistent with the experimental results, although the curve itself was somewhat jagged. Model II: The early models proposed by Peterlin and co-workers (Peterlin, 1971 a,b; Verma and Peterlin, 1972) were the same as those postulated by DeVries and Kausch. Later, however, Peterlin proposed a new model. Using Model I, he first obtained the stress-strain curve, the frequency distribution and the accumulation of the free radicals (Fig. 15-48) by assuming that the length distribution of tie-chains follows the normal distribution, and that each tiechain breaks at its critical strain. He then examined and critically analyzed this model in light of these observations: (1) In an ideal model, the stress will gradually decrease to zero after reaching the maximum. In DeVries' model, because the microvoids are generated immedi-
15.3 Fracture
ately after the stress had reached its maximum, the right side of the void generation frequency distribution could not be observed, and therefore the point of maximum stress coincides with the center of the distribution. In actuality, however, the point of maximum radical generation shifts significantly to the right side of the point of maximum stress. (2) In the experiments determining the stress-strain curve, failure occurred at a stress level beyond the maximum stress because a bundle of fibers was used. If a single fiber had been used, the maximum stress would have been completely equal to the fracture stress. In the case of fiber bundles, fluctuations in the mechanical properties and in the lengths of single fibers should be taken into consideration. (3) The number of radicals generated varies significantly with the amount of heating and the mechanical history of the fiber. In Model I, it was shown that the strength grows with the width of the distribution of radical generation frequency when the strain is increased. This, however, ignored the effect of mechanical history. (4) According to Model I, if a once-fractured sample is retested, its strength must be lower than that of the original sample since most of the tie-chains have already been broken. No such experimental results, however, have been observed. (5) According to calculations based on Model I as well as observations, the number of radicals generated until fracture occurs is about 1% of the tiechains. Peterlin considered the above problems to be caused by the inappropriate assumption of uniform tension in the amorphous
743
regions within microfibrils, and proposed an alternative which we shall call Model II. In this model, which addresses this difficulty, the ends of microfibrils are thought to give rise to two types of defects, i.e., cavities and dislocations, and amorphous regions around these cavities are assumed to be under greater strain than those regions away from the fibril ends. This model is depicted in Fig. 15-49. The frequency distribution of radical generation detected by ESR actually reflects the nonuniformity of the stress rather than the length distribution of tie-chains. Indeed, according to Model II, the tensile strength of once-fractured samples or the number of radicals occurring until fracture may be explained since the tie-chains in the amorphous regions other than the ends of microfibrils may remain unbroken. However, it is not clear how far-reaching the effect of cavities at the ends of microfibrils might be along the microfibrils and it is therefore difficult to carry out a quantitative analysis. Because Model II cannot provide a quantitative calculation as can Model I, it is still unclear which model is more reasonable. Nevertheless, the idea in Model II that cavities can be formed at the ends of microfibrils thereby introducing a nonuniform stress field in the material, may be considered as a new direction in understanding the mechanical properties of fibers. Since Model II cannot explain the initial Young's modulus of fibers, a revised model has recently been proposed (Peterlin, 1975) in which tie-chains with lengths shorter than the average lamellar thickness are assumed to make the surface of the lamellae uneven. According to this model, the number of extended tie-chains in the initial stage will be large and this leads to a higher Young's modulus.
744
15 Crazing and Fracture of Polymers Cross section DD'
Cross section CC
B1
Cross section AA1
-4f-4f-4C'
C4-4f-t
(a) Cavity
•ODD D1
an ana
Figure 15-49. Defects caused by the ends of microfibrils: (a) cavities and (b) dislocations (Peterlin, 1971b).
D'
(b) Dislocational defect
15.3.2.3 Relationship Between Chain Scission and Macroscopic Fracture
Macroscopic fracture does not necessarily occur even if all the tie-chains in the cross-section have been broken. It is reasonable to imagine that some sort of Griffith-type microcracks, which can also lead to macroscopic fracture, can be generated during the tension. This process can be described by the following sequence of events: (1) Stress concentrations arise in the structural heterogeneities in fibers. (2) Chain scission occurs due to the stress concentration, leading to the formation of microvoids. (3) Microvoids grow and evolve into cracks by aggregation. Zhurkov and co-workers (Zhurkov et al., 1969 a) investigated how microvoids grow and become a crack through coales-
cence, and obtained a parameter for the aggregation state of microvoids at which catastrophic fracture can occur. This type of microvoids is thought to be generated throughout the amorphous regions. This is deduced from the fact that the surface area of a microvoid estimated from the number of tie-chain scission, 10 6 -10 7 m 2 /m 3 , is roughly consistent with that calculated from the size and the number of microvoids, 10 5 -10 6 m 2 /m 3 . On the other hand, Model II argues that the number of cavities at the ends of microfibrils, Npu = 1021/m3, which is in agreement with the number of microvoids calculated above. If this is true, then the number of microvoids is determined in the beginning, i.e., it depends only on the structure. However, experiments demonstrate that the number of microvoids increases with extension, which is inconsistent with this postulation of Model II.
15.3 Fracture
Whatever the case may be, catastrophic fracture of a sample will occur once the transformed microvoids grow into a large crack. This fracture may proceed along the path of least resistance as shown in Fig. 1550. In the case of PA-6, which has a large interfibrillar adhesion because of the relatively strong hydrogen bonding, this path can be the ends and amorphous regions of the microfibrils. For PE, where the microfibrils are held together by van der Waal's force only, the path can be the interfibrillar areas. This theory would perhaps explain why the number of radicals generated during fracture is fewer for PE than for PA-6. 15.3.3 Ductile Fracture
In the previous section, we focused on the micromechanism of brittle fracture initiated from chain scission mainly for crystalline polymers such as fibers in which sufficiently uniform orientation has already been established by processing. If the starting material for deformation is unori-
n n
n
(a)
(b)
Figure 15-50. A schematic representation of the fracture path in PE (a) and PA-6 (b) (Peterlin, 1975). A crack propagates between the ends of microfibrils (A), parallel between two microfibrils (inter-microfibril) (C), and through a microfibril (intra-microfibril) (B).
745
ented, then fracture occurs after it has undergone macroscopic plastic deformation. Although this fracture is ductile, it is different from plastic flow in which molecular chains slide past each other. In this section, we shall concentrate on ductile fracture again in the sense of chain slippage, but in the context of a slightly different phenomenon. 15.3.3.1 Fracture of Shear Deformation Bands
When the ductile deformation is constrained, shear deformation bands (slipbased) occur as a nonuniform plastic deformation. Two different configurations of this type of shear deformation bands can be observed. One is called coarse shear bands in which the width of the deformation band is large, and the boundaries between the sliding and the surrounding areas are distinct. The other configuration is called the diffuse shear band in which a shear band is composed of a number of fine shear bands. In the former case, shear fracture along the deformation band often occurs; in the latter case, a large kink will appear in the specimen rather than fracture (Chau and Li, 1976). Although shear fracture occurs when the deformation band completely traverses the specimen, in the interior of the deformation and, deformations such as strands (fibril-like deformation bands) and void sheets may have already formed during the growth of the shear deformation bands (Chau and Li, 1980) (Fig. 15-51), and catastrophic fracture will proceed along the boundary between the void sheets and the undeformed areas (Chau and Li, 1981). The phenomena described above occurs in compression. If we reverse the compression, i.e., apply a tensile load before compressive shear fracture occurs, fracture will
746
15 Crazing and Fracture of Polymers
Figure 15-51. A shear deformation band in PS which was formed in compression (Chau and Li, 1980).
still occur along the band, but the strands inside the band are now stretched and exhibit a remarkable fracture morphology (Fig. 15-52) where the strands are formed in steps (Chau and Li, 1981). When deformation is not constrained, plastic deformation due to tensile loading will begin with necking which is due to postyield softening. Thus, necking unloads the surrounding unyielded material and the elastic strain energy there is released. In materials with a significant orientation-induced strain hardening effect, a neck will propagate regardless of the tempera-
ture elevation; however, in materials where the strain hardening effect is not strong, temperature elevation will lower the strength in the neck area. In the neck itself heat is generated by the plastic deformation. At sufficiently high rates of deformation (if yielding is still possible) fracture may result from localized heating. This type of fracture has been reported for poly(vinyl chloride) PVC and PC, and is usually called thermal fracture (Cross and Haward, 1973). The molecular basis for the reason why various plastics have different tendencies for strain hardening is still not known. Even at moderate rates of loading, neck formation may cause fracture in the neck itself, probably because the neck induces a higher triaxial tensile stress due to the plastic constraint. 15.3.3.2 Plastic Fracture
At the beginning of tensile deformation, crazes sometimes occur on the surface of the specimen, but do not cause fracture. A
(b) Figure 15-52. Fracture surface (a) and side view (b) of a shear band in PC which was initially formed in compression, then fractured in tension (Chau and Li, 1980).
15.3 Fracture
possible mechanism for this is that the craze is blunted by a pair of shear bands. In this case the crazes may persist in the necking area. Once the neck has propagated it is difficult for the crazes to grow either along the chain orientation or in the normal direction. They may then remain in the neck area without growth. When the stress in the specimen is further increased, cavities may occur in some of the crazes (Fig. 15-53). In accordance with their appearance, they are called diamond cavities. With increasing tensile stress, a diamond cavity will grow first along the tensile direction until it stabilizes and then in the transverse direction until the specimen fails. Examples for this type of fracture after plastic deformation are found in PVC (Cornes and Haward, 1974; Walker et al., 1979 a), PC (Cornes et al., 1977; Walker et al, 1981), PSU (Cornes et al, 1977), PET (Walker et al, 1979 b), and at higher temperatures, in PMMA and PS (Cornes et al, 1977; Smith and Haward, 1977). The generation of diamond cavities may not always be due to crazes in plastic deformation. The reason for the formation of this particular type of cavity is generally thought to be due to the pure shear deformation in the orientation direction at the tip of the cavity (Walker et al, 1979 a). However, since it occurs in anisotropic materials after plastic deformation, it is difficult to estimate the states of stress or strain around it. 15.3.3.3 Failure Envelope
The fracture of elastomeric materials such as vulcanized rubber, etc, in uniaxial tension may exhibit an extremely large strain to failure, although it does not involve permanent plastic deformation, i.e., ductility. In the tensile deformation of this type of materials, an important experimen-
747
Figure 15-53. A SEM micrograph of a diamond cavity in PMMA (Cornes et al., 1977).
tal result has been found, i.e., changing the deformation rate at a constant temperature has the equivalent effect on the failure strain as changing the temperature at a constant deformation rate. A master curve, shown in Fig. 15-54, which is called the failure envelope, can be constructed for any elastomer using the scheme described (Smith, 1958, 1969). The phenomenological basis of the failure envelope is the time-temperature superposition principle well-known for describing the viscoelastic behavior of polymers. In this particular application of the principle, the stress and time to fracture, i.e., <jb and th of elastomeric materials under uniaxial tension at various temperatures, can be collected into a single master curve by using a shift factor based on the WLFequation. Similarly, the strain to fracture £b and the time to fracture can be collectively described by using the same shift factor. Thus, the relationship ah — th — eh, i.e., ah — £b, can be obtained through the master curve. At constant rates of extension,
748
15 Crazing and Fracture of Polymers
Symbol
Temp. °C
X
140 100 70 40 25 10 -5 -20 -35 -45
•
1.5
Q
O
1.0
©
a 9
0.5
15.3.4 Nonuniform Fracture of Blunt-Notched Polymers
99
a y a jf a
/
/ a
0
a a|
-
8^r
o
0.5 i
-0.2
0
0.2
i
0.4 log CVb-1)
1
1
0.6
0.8
Figure 15-54. The failure envelope of SBR rubber (Smith, 1969).
th = sb/s; therefore, altering the strain rate is equivalent to altering the temperature. The failure envelope is valid if the structural change prior to fracture is kept in the equilibrium or nearly equilibrium state. Moving the curve towards the upper righthand direction (see Fig. 15-54) corresponds to ah — sh behavior at higher strain rates or lower temperatures. At the top of the curve, the experimental data begin to deviate from the envelope. This is due to deviations from equilibrium (Smith, 1965). For example, when an elastomer is deformed at low strain rates or high temperatures, a sudden increase in the strain rate during extension (Knauss, 1965) or the occurrence of a fracture due to cyclic deformation (Valanis and Yilmazer, 1978) results in deviations from the envelope due to departure of the deformation mode from the equilibrium state.
Structural components often contain notches of various geometries. Brittle failure sometimes occurs as a result of the presence of such notches, especially under impact conditions. Notched impact strength testing has therefore become an important practical tool for evaluating the fracture resistance of structural components containing blunt notches or abrupt changes in geometry. However, the apparent ductile or brittle behavior is highly dependent on the geometry of the notches, the thickness of the specimen, the rate of deformation, and the temperature. The origin of macroscopic behavior governing the fracture of notched polymers and the causes of geometric effects are discussed in this section. 15.3.4.1 Brittle Fracture of Ductile Materials
Fracture can occur in an unnotched (smooth) specimen in tension after uniform plastic deformation. Such phenomena have been discussed in a previous section. The same material with a notch in it may fracture in a brittle fashion. This is generally called notch (-induced) brittleness and occurs in many different types of polymers. Figure 15-55 shows the stress-strain relationship of a PE bar with a U-shaped circumferential notch for various notch depths. When the notch depth exceeds a certain value, a ductile-brittle fracture transition occurs (Narisawa et al., 1979). The phenomenon of notch brittleness in ductile has been explained by Orowan (Orowan, 1949). When a U-notched bar is pulled in tension, an extension occurs in the pulling direction and at the same time a contraction occurs in the transverse direction as the pulling stress is increased (Fig. 15-56). In elastic deformation, the
15.3 Fracture
p- 1.0 mm
749
elastic compared with that at the base of the notch since the cross-sectional areas here are larger, hence the stresses are lower, moreover, plastic deformation has not yet occurred at these locations. Consequently, the material in the notch will experience a tensile stress in the radial direction, and simultaneously a tensile stress of nearly the same magnitude will arise in the circumferential direction. As a result, in spite of the fact that the stress originally applied was uniaxial tension, a triaxial stress is actually set up in the cross-section of the notch. Since all the normal stresses are tensile, a negative hydrostatic pressure (dilative stress) is in effect exerted. When the axial stress is increased, and if the material is not so brittle as to fracture almost immediately, then plastic deformation will proceed inward along the notch cross-section, and the stresses in the radial and circumferential directions will also increase accordingly. When the maximum stress or dilative stress exceeds the critical value for
Strain (%)
Figure 15-55. The stress-strain behavior of a PE specimen with circumferential notches of various depths d: (A) 27 mm, (B) 21mm, (C) 15 mm, (D) 0.9 mm, (E) 0.7 mm, (F) 0.2 mm, (G) 0 mm. (Narisawa etal., 1979.)
contraction is due to the Poisson's ratio effect. If the tensile stress continues to be increased, plastic deformation will occur first at the base of the notch as a result of stress concentration. If the plastic deformation resulting from the longitudinal stress is shear yielding instead of crazing, an even larger contraction will occur in the transverse direction since the volume in the area of plastic deformation must remain approximately constant. On the other hand, in the upper and lower ends of the notch, the lateral contraction is very small and
Figure 15-56. A schematic illustrating how a radial stress
750
15 Crazing and Fracture of Polymers
cohesive failure of the material, the specimen will fail at the notch. In this case, the plastic deformation is limited only to part of the cross-section of the notch; therefore the fracture is still macroscopically brittle. When shear yielding occurs at the tip of the notch, the axial stress becomes as high as (7t == aY + ox (GT is the stress in the radial direction) where
a;
p
°z
a notch exists in a specimen and is not in a shape that can maintain sufficient constraint in the radial and circumferential directions, then plastic yielding in the notch area will proceed until the entire cross-section has yielded. Fracture in this case can be regarded as ductile in a real sense. 15.3.4.2 Stress Distribution Around a Local Plastic Deformation Zone
We now consider a plate with a blunt notch, as shown in Fig. 15-57. If it is in the elastic range, it will contract in the z-direction due to the Poisson's effect. If its thickness is sufficient to maintain a plane strain state during the deformation, i.e., sz = 0, then a tensile stress oz which is to resist the contraction will appear in the thickness direction of the specimen. Also, a contraction in the width (x-direction) will occur and for the same reason, a tensile stress ax is generated in the notch, because sx inside the notch is constrained by the smaller £x outside the notch. oz has a maximum in the center of the plate and is expressed by G = v z (°x + Gy) (v i s Poisson's ratio). On the surface of the specimen, the stress oz is, of course, zero. The stress ax increases from zero at the base of the notch, reaches
1
Figure 15-57. Elastic stress distribution in the interior of a thick plate with a notch whose radius is Q.
15.3 Fracture
a maximum, and will decrease again farther away from the notch. Thus, inside the region near the base of the notch, ay> az> ax and the maximum shear stress T max = iGy — °x)/2 if the plane strain condition (ez = 0) is satisfied. If shear yielding occurs at the base of the notch, the slip planes will be normal to the x y-plane and propagate at some angle to the notch direction, as illustrated in Fig. 15-58b. This mode of deformation is said to be plane strain. On the other hand, if the specimen is thin, constraint will not exist in the z-direction, thus GZ will be zero throughout the thickness of the specimen. In this case, shear yielding will occur through the thickness, once again because Tmax = (
751
Figure 15-58. Plane stress (a) and plane strain (b) deformation (Hahn and Rosenfeld, 1965).
tion by Nimmer (1991) has shown that the analytical solution provided by the SLFtheory is surprisingly accurate for two ductile glassy polymers, PC and poly(ether imide) PEL According to the SLF-theory, the slip lines, i.e., lines of maximum shear stress, at the notch tip can be described as a logarithmic spiral, and ay inside the plastic zone (x < s) is given by Gy = (jY 1 + l n M + - j
(15-30)
Figure 15-59. Plastic deformation zone at a blunt notch tip in a thick PC specimen.
752
15 Crazing and Fracture of Polymers
Figure 15-60. Plastic deformation and stress distribution according to the slipline field theory. The notch tip radius Q and the thickness are such that the specimen is in plane strain, s is the extent of the plastic zone.
where Q is the radius of the notch, and oY is the uniaxial yield stress. If we further assume, for the sake of simplicity, that the material obeys Tresca's yield criterion, then, from the relationship oy — aY = ax, we have
i\
(15-31)
The maximum extent of the logarithmic slip-line field s* is determined by the geometry of the notch tip, and is given by S* = Q
ex
PU - ^
-
(15-32)
where co is the notch angle. As the load is further increased, i.e., s > s*, the stress in the range s* < x < a is given by K
CO
(15-33)
The value of ay is the maximum value of the stress in the tensile direction when a plastic zone exists at the notch tip, and depends on the opening angle of the notch CD only. When x < s*, the maximum of ay is at the elastic-plastic boundary x = s, i.e., = (7,
L
1 + In ( 1 + -
V Q.
(15-34)
Since there is little volume change during the plastic deformation, the relationship (15-35) holds. We therefore have az = ^ |~1 + 2 In (1 + -Yl (15-36) 2 L \ QJA The hydrostatic stress p due to the triaxial tensile stress at the notch tip is P=
(15-37)
and is consistent with the value of Eq. (15-36). In materials with cohesive strength substantially higher than shear yield strength, e.g., metals, the plastic deformation at the notch tip will simply grow with the applied load. In bending, a plastic hinge will develop. However, isotropic thermoplastics generally possess cohesive strengths (i.e., crazing strengths) not substantially different than the shear yield strength. Consequently, the formation of a plastic hinge under plane strain conditions is seldom, if ever, observed. Instead, fracture due to cohesive failure truncates the plastic deformation process. This phenomenon is described in the next section.
15.3 Fracture
15.3.4.3 Fracture of a Notched Specimen due to Plastic Constraints
In deformation under plane strain, if plastic deformation at the base of a notch proceeds, a large dilative stress expressed by Eq. (15-36) will occur at the tip of the notch. In Fig. 15-5 we showed a micrograph of the plastic zone at the root of a notch in a thick PC specimen after the deformation has proceeded to a sufficiently large extent. One can see that cracks are generated in the direction normal to the tensile axis (Ishikawa et al, 1977; Narisawa et al., 1980 c). The corresponding fracture surface is presented in Fig. 15-61. There are three significant features on this surface: (a) a spot that is the nucleus of fracture at the tip of the plastic zone where cracking started, (b) a smooth, flat zone that surrounds the nucleus formed by the fracture of a craze that had grown relatively slowly, and (c) a zone which clearly indicates that a rapid propagation of cracks has proceeded from the smooth area, (c) Is in part caused by unstable fracture as a result of the use of a notched 3-point-bend specimen.
Figure 15-61. Fracture surface of a thick PC specimen with a blunt notch.
753
The type of fracture surface in PC described above has been observed in thick plates of U-notched PVC and PMMA as well. In thick plates of PE and PP, a number of cracklike features can be observed starting at the root of the notch (Fig. 15-62) rather than in the interior as is the case of most glassy polymers. This behavior is observed in many semicrystalline polymers (Narisawa and Ishikawa, 1990). Figure 15-63 shows the fracture surface in PE
0.1 mm Figure 15-62. Multiple crazing at the notch root of a thick PE specimen subjected to bending (Narisawa etal., 1979).
Figure 15-63. Fracture surface of a PE specimen broken as in Fig. 15-62 (Narisawa et al., 1979).
754
15 Crazing and Fracture of Polymers
(Narisawa et al., 1980 c) obtained by bending a thick, notched specimen. Although no fracture nucleus can be clearly identified, it is apparent that fracture began ahead of the base of the notch rather than at its base. Again, in this case, plastic constraint became the cause of fracture, which is clearly demonstrated in the stress-strain curves of notched PE specimens in Fig. 15-55. When the location of a fracture nucleus can be identified on the fracture surface, its distance from the bottom of the notch is equal to s* in Eq. (15-33). From this, one can obtain the critical dilative stress during fracture from Eq. (15-36). Figure 15-64 shows an example in which the critical stress of PC so obtained is plotted against the strain rate. Although the yield stress generally increases with increase in strain rate, the critical dilative stress exhibits almost no change (Narisawa et al., 1980 a). In thermoplastics, the fracture nucleus has been shown to begin as a craze (Narisawa etal, 1980c; Ishikawa et al., 1981). Table 15-1 gives the critical dilative stresses of various polymers determined in this manner. We have shown that the use of a critical dilative stress criterion is quite reasonable for craze initiation. The dilative stress at the tip of the plastic zone in plane strain first generates crazes. The crazes will at first grow thicker in the normal direction, and upon further thickening the crazes evolve into cracks. It has been argued that crazing is the result of a strong tensile stress concentration like that due to the intersection of the slip lines in metals (Chau and Li, 1981; Kramer, 1975; Camwell and Hull, 1973). However, the slip line field at the notch tip does not always involve distinct shear bands. For example, in quenched PMMA, the plastic zone exhibits only diffuse slip lines (Fig. 15-65). This clearly
Table 15-1. Critical dilative stress due to plastic constraint. Polymer (20 °C)
Critical dilative stress (MPa) 37 40 56 69 74 67 87 96
PS (slow cooled) PS (quenched) PSAN (slow cooled) PSAN (quenched) PMMA (quenched) PVC (quenched) PC (slow cooled) PC (quenched)
100
Q_
50
OVA
-I*
-3
-2
Strain rate (s"1)
Figure 15-64. Critical dilative stress for crazing for PC and its dependence on the strain rate (Narisawa etal., 1980 a).
Figure 15-65. Polarized (A) and normal (B) light micrographs of a thin section of quenched PMMA from a thick notched specimen deformed to initiate a craze.
755
15.3 Fracture
demonstrates that even when there are no well-defined slip lines in plastic deformation, crazing can still occur. The existence of slip lines is therefore not essential for crazing. Instead, the dilative stress due to plastic constraint is the real cause for crazing (Orowan, 1949; Ishikawa et al., 1977; Narisawa et al., 1980 a, b; Ishikawa et al., 1981; Grade and Weiss, 1972; Mills, 1976). Since a dilative stress is necessary for craze initiation, it stands to reason that if this dilative stress is counteracted by an externally imposed hydrostatic pressure, then crazing should be suppressed. In Fig. 15-66, the bending moment at which crazing occurs in a thick notched PC specimen under an external hydrostatic pressure is shown as a function of displacement. The bending moment required to cause crazing increases with the hydrostatic pressure until yielding occurs through the entire cross-section of the specimen, i.e., a plastic hinge is formed (Ishikawa and Narisawa, 1983). The testing temperature also changes the location of the craze nucleus. Figure 15-67 shows such an example for PMMA. By extrapolating to the point where the nucleus is at the surface of the notch, the ductile-to-brittle transition temperature can be found. Generally speaking, small flaws with a large tip radius in a specimen can be safely ignored in considering the strength of polymer products. This is particularly true in those polymers which exhibit ductile behavior. However, as we have indicated, even a small plastic deformation will generate plastic constraints, thus crazing or cracking. It is therefore important to realize that the dilative stress at the tip of the plastic zone will increase due to these constraints and will lead to macroscopic brittle fracture of the specimen.
20 ^ S l o w cooled
29.4 MPa/
16 --
1 ^^-Quenched
9.8 MPa/ 12 /
A
t 5.9 MPa 3.9 MPa
V
7.8 MPa 39.2 MPa
MPa
MPa
f
i 1.0
i
i
4.0
2.0 3.0 Displacement (mm)
Figure 15-66. Effect of hydrostatic pressure (numbers in the figure) on bending moment at which crazing occurs in slow-cooled and quenched PC (Ishikawa and Narisawa, 1983).
o [Bending
•
•"I
X
speedi
1 mm/min
1.0
o
-
x
o o
•
•
x
10 mm/min o
X
•
•
X X
•
0.5 -o/
o
100 mm/min X
/
X /
i
n
300
320 Temperature (K)
i
340
Figure 15-67. Effect of temperature and testing rate on the location of the craze nucleus relative to the notch radius, R/Q for PMMA (Ishikawa et al., 1981).
756
15 Crazing and Fracture of Polymers
15.3.5 Fracture of a Cracked Specimen 15.3.5.1 An Application of Linear Fracture Mechanics
According to Griffith's criterion (Griffith, 1921), fracture occurs when the elastic strain energy released per unit increment area of a crack with length a exceeds the increase in surface energy ys, i.e., 2KG2
a>4ys
and can be rewritten as (15-39)
and for a three-dimensional crack, it can be expressed by G >
-v2)na
(15-41)
(15-38)
Thus, the stress for fracture is
/2ysE
tionship between the radius of the crack tip and the interatomic distance. This can be seen if we let the radius of the crack tip be Q, the stress at the crack tip then becomes
(15-40)
where v is Poisson's ratio. Equations (1539) and (15-40) give a necessary condition for fracture of a perfectly elastic body in terms of energy conditions. Another necessary condition is the stress. We can estimate the level of stress by considering what happens when the crack tip radius approaches the interatomic distance. Equations (15-39) and (15-40) imply that the maximum stress almost reaches the level of the interatomic bond strength and is thus a sufficient condition for fracture in terms of the mechanics of continuous media. This consideration suggests to us that one more condition must be satisfied other than Griffith's condition, i.e., the local stress at the crack tip must exceed the interatomic bonding force (Yokobori, 1975). The failure strength should be obtained by satisfying both conditions, and determined by the one which gives a larger critical stress. Which condition should be applied in a particular situation depends on the rela-
o>
(15-42)
Comparing Eq. (15-39) with Eq. (15.42), one can clearly see that when 7tg(8r0) < 1, i.e., Q < 3ro/n, the stress for Griffith's criterion Eq. (15-39) becomes larger, and is therefore the critical stress, since the local stress condition will be automatically satisfied; while if Q > 8ro/7i, the local stress in Eq. (15-41) will become larger, and thus will be required instead as the fracture criterion. In practice it would be difficult to find Q < 3ro/7i. Thus, the stress for fracture is always greater than the value determined by Eq. (15-39). Furthermore, the energy required to grow the crack exceeds the surface energy by a considerable amount due to the plastic energy that must accompany the crack growth. These considerations are described in the following sections. 15.3.5.2 The Strain Energy Release Rate gc and Fracture Toughness Kc of Glassy Polymers
In early studies, polymers which tended to exhibit brittle fracture, such as PMMA (Irwin et al., 1952; Benbow and Roesler, 1957; Benbow, 1961; Svensson, 1961; Berry, 1961a, 1962, 1963; Broutman and McGarry, 1963; Berry, 1964), and PS
15.3 Fracture
(Irwin etal, 1952; Benbow, 1961; Berry, 1961 b, 1963; Murray and Hull, 1971) were investigated to see if Griffith's criterion was applicable or not. This was accomplished by preparing specimens with cracks of various lengths and by determining if Eq. (15-39), i.e., (Th~l/y/a9 holds. Recently, with the development of testing methods for fracture toughness in metals, more attention has been paid to the preparation of cracks, the method of loading, and the shape of the specimen (Marshall et al., 1973), and more exact measurements of the critical strain energy release rate gc or the critical stress intensity factor Kc has become possible since then. Table 15-2 summarizes the values of Klc for a number of polymers. The additional subscript I is used to indicate Mode I (opening mode) fracture. In general, the value given is for plane strain fracture. Because the values of glc or Klc are affected by: (1) the thickness of the specimen,
757
(2) strain rate, (3) temperature, (4) molecular weight, (5) cross-linking, (6) whether oriented or not, and (7) for crystalline polymers, the morphology, it is difficult to directly compare the values in Table 15-2 on the basis of molecular structure alone. Nevertheless, some generalizations could be made. For example, thermosetting materials which have a cross-linked structure usually exhibit lower glc and Klc than thermoplastic materials. Among thermoplastic materials, vinyl polymers such as PMMA and PS show lower glc and Klc; by contrast, PC and PSU, and toughened materials such as ABS, exhibit high glc and X Ic . The theoretical surface energy ys for polymer solids is less than 0.1 J/m2. Even for brittle thermosetting materials the measured ys is higher by at least two orders of magnitude. The reason is that the fracture of glassy polymers does not occur in an elastic fashion; instead, there are microdeformation mechanisms which can act as
Table 15-2. X Ic of various polymers. Polymer
Specimena
Temperature b (°C)
Crosshead speed0 (mm/min)
References
SENB SENB SEN4B CT SENB SENB SENT SENB SENB SENB SENB CT SENB
23 23 RT RT 20 20 RT -60 -60 20 -40 RT RT
1.0 5.0 0.076 0.03 0.5 0.5 SS 0.1 0.5 0.1 0.1 0.03 0.254
Misaki et al. (1989) Chan and Williams (1981) Gupta etal. (1986) Serrano et al. (1982) Hashemi and Williams (1985) Hashemi and Williams (1985) Hobbs and Bopp (1980) Hashemi and Williams (1985) Fernando and Williams (1980) Hashemi and Williams (1985) Hashemi and Williams (1985) Serrano et al. (1982) Wu and Schultz (1989)
(MPa m1/2) Epoxy HDPE PS PMMA PVC PBT PP POM PA66 PPO PEEK a
1.12 1.18 1.29 1.00 1.97 2.81 2.73 3.37 4.1 4.05 3.84 4.85 7.0
SENB; single edge notch bending (3-point bending), CT: compact tension (tension), SENT: single edge notch tension (tension), SEN4B: single edge notch bending (4-point bending); b RT: room temperature; c SS: slow speed.
758
15 Crazing and Fracture of Polymers
energy sinks before the propagation of cracks. Even when the fracture appears brittle by macroscopic observation, the contributions from these mechanisms are still significant. Therefore, the experimentally obtained surface energy y's = ys + yp (7P > 7s) (Orowan, 1949; Irwin, 1948), where yp is the energy of plastic deformation per unit area in a thin layer ahead of a crack tip. 15.3.5.3 Plastic Deformation at a Sharp Crack Tip
The occurrence of plastic deformation at and ahead of a sharp crack tip has been identified as the main source for a large strain energy release rate gc. When a thin PC specimen with a crack (hence it is in the plane stress state) is extended (Fig. 15-68), a plastic deformation zone that increases with the increasing load will appear before the crack propagates (Narisawa et al., 1976, 1978 a, b). By cutting a thin section from this well-developed plastic zone in the direction normal to the growth direction and observing it under a polarized microscope, one can obtain the type of pattern shown in Fig. 15-69 (Ishikawa et al., 1976), where the symmetric slip of the deformation zone in the maximum shear stress direction as well as their intersection zones
/Cx=3.56
4.93
are plainly visible. This is a typical pattern of plastic deformation in the plane stress state as seen in Fig. 15-58. Such a characteristic shape and structure of plastic deformation at a crack tip has also been observed in other materials such as PVC (Narisawa et al., 1977). 15.3.5.4 Crazing at a Sharp Crack Tip
Another cause for the very large gc and Kc values compared with theoretical predictions is crazing that could occur at the crack tip. When crazing already exists at a crack tip, the glc is found to be 40 ~ 50% lower than if no crazing occurred (Narisawa and Kondo, 1972; Narisawa, 1972 a, b). When the specimen thickness is large (plane strain state) crazing usually occurs at the crack tip before crack propagation. For the most part glc is consumed in the formation and deformation of the crazes if account is taken of the energy required for the deformation of fibrils within the crazes until failure (Kambour, 1966 b). The shape of the crazed zone that occurs at the crack tip in a thick plate (plane strain) is generally consistent with that calculated by the Dugdale model. In practice, the lengths of crazes rc are often such that rc <^ a, where rc = n Kf/(& al). Thus, the crack opening displacement
6.20
0.2 mm
Figure 15-68. Plastic zone growth under plane stress conditions in a thin specimen of PC with a sharp crack.
15.3 Fracture
759
tained from the shape of the crazes is less than half of the value actually measured (Mills and Walker, 1976; Brown and Chin, 1980; Brown and Stevens, 1978). 0.20
15.3.5.5 Various Factors Influencing Fracture Toughness
The following is a summarized list of the various factors that influence gc or Kc. (1) Effects of specimen geometry and crack initiation. If one compares the plastic 0.79 deformation zone and the crazes that occur at a crack tip in the plane stress state with 1.12 those in the plane strain state, one would observe that the former requires more energy than the latter simply because the vol1.40 ume of material involved in the plastic deformation is larger. The glc measured for Figure 15-69. Polarized light micrographs of crossPMMA and PS exhibit almost no thicksections of the plane stress deformation zone ahead of ness effect (Berry, 1964); in the case of ducthe crack in a thin PC specimen. The numbers inditile polymers such as PC (Brown and cate distances from the crack tip. Ward, 1973; Parvin, 1981; Frazer and Ward, 1978; Parvin and Williams, 1975a,b; proximated by <j> = K?(l — v2)/(£ ac) and Pitman and Ward, 1979; Kambour and the calculation can be treated linearly. This Miller, 1976; Kambour et al., 1978), PPO has been demonstrated in a number of ex(Kambour and Smith, 1979), PSU (Ting, periments on PC (Frazer and Ward, 1978; 1981), and PE (Bandyopadhyay and Pitman and Ward, 1979; Kambour et al., Brown, 1982), the fracture toughness was 1978), PMMA (Murray and Hull, 1971; found to be dependent on the thickness of Brown and Ward, 1973; Weidmann and the specimen. This is because the specimen Doll, 1978; Doll and Weidmann, 1979), surface deforms in plane stress, while the PVC (Weidmann and Doll, 1978), and ABS interior deforms in plane strain if sufficient (Doll and Weidmann, 1979). The relationconstraint exists. If the specimen is very ship between KY and (j) thus becomes thick, then the bulk of the specimen deforms in the plane strain state, and a more brittle type of behavior dominates the de(15-43) 1 8 r c ( l - v 22\2 ) formation. However, if the specimen is so thin that the plane strain region never Thus the fracture toughness can be obforms because of the lack of sufficient contained if we know the shape of the crazes straint, then the specimen could deform in immediately before fracture. For PVC, it is a very ductile manner. At intermediate known that the opening displacement is thicknesses, a mixed mode deformation smaller than that calculated from the Dugcould occur, especially near the notch tip dale model and the plastic zone tends to be where some elastic constraint could exist. long and sharp. The fracture toughness ob0.29
760
15 Crazing and Fracture of Polymers
The fracture surface of a PC specimen that fractured in a mixed mode is shown in Fig. 15-70. Remnants of shear deformation that started from the two side surfaces (shear lips) can be seen. The crack initiated by crazing in the mid-thickness of the specimen near the crack tip where the elastic constraint prevented the shear lips from forming. Consequently, the apparent fracture toughness, Kc, of such specimens decreases as the thickness of the specimen increases, and approaches a constant plateau value (Fig. 15-71). This plateau value should be cosnsidered as the true plane strain fracture toughness of the material.
Plane stn zone
Figure 15-70. Mixed mode fracture of a PC specimen.
It has been shown that if the thickness of the plate is B, and if half of the thickness is greater than twice the dimension of the plastic deformation zone, then crack propagation will occur in the plane strain state in the middle of the plate (Irwin et al., 1958); that is, (15-44) 7i
aY
The dimension of the plastic zone is thus given by the Dugdale model as s = 7TKC2/(80"Y) f° r the surface and s = Kl/(2n GY) for the interior where a state of plane strain exists. Kc2 is the fracture toughness in the plane stress state. Assuming that the plane stress plastic zone (shear lips) and the plane strain deformation zone (crazes or cracks) form independently of each other, the fracture toughness of a specimen of arbitrary thickness may be estimated by assuming that it is a simple linear combination of the two contributions. With such an assumption, experimentally obtained fracture toughness in the case of a mixed plane stress and plane strain deformation can be expressed by (Parvin, 1981; Parvin and Williams, 1975 a, b) n Gy B
Ductile fracture Plane stress
Thickness
Figure 15-71. The thickness dependence of apparent fracture toughness of ductile polymers.
(15-45)
where Kcl is the fracture toughness in the plane strain. In the case of PC, for example, Kc2 = 2.5 Kcl (Parvin and Williams, 1975 a). If the plane stress deformation zone is able to grow to a significant extent then the plane strain deformations may be prevented from taking place. When interaction between the two modes occurs, the additivity assumption underlying Eq. (1545) may not be valid. Clearly, then, it is important for the person performing the test to note at what thickness the fracture toughness is evaluated.
15.3 Fracture
In the experimental evaluation of the toughness of polymers, for the same reason that specimen thickness affects the apparent fracture toughness, the manner with which the crack is introduced into the specimen could also significantly affect the measured apparent toughness. For example, pressing a sharp razor blade into the specimen could create a compressive yield zone and produce artificially high apparent toughness even if the crack does appear to be very sharp. The use of fatigue loading could also create a damage zone at the crack tip, once again increasing the apparent toughness. By the same token, blunt notches will usually produce higher apparent toughness than sharp notches. Therefore, in fracture toughness testing, it is essential that a sharp crack free of the above-mentioned defects be used. The failure of many investigators to exercise caution in this respect has caused a certain amount of confusion regarding the true fracture toughness of polymers. There is no set rule on how to produce a sharp crack, except that the crack tip should be examined to see if plastic deformation has been introduced by the crack initiation technique, and that the specimen dimensions are such that the measured toughness is independent of the geometry. In the case of extremely ductile polymers such as PE and highly toughened plastics, nonlinear fracture mechanics techniques must be used. A discussion of such techniques is beyond the scope of this review. (2) The effect of strain rate. We note that, although fracture toughness values are sometimes reported as a function of testing rate, the effective strain rate at a crack tip is usually much higher than the nominal strain rate. In general, as Fig. 15-72 indicates, glc decreases as the strain rate or testing rate increases (Narisawa and Kondo, 1972; Narisawa, 1972; Ting, 1981;
761
Yamani and Young, 1977, 1979; Phillips etal., 1978; Cherry and Thomson, 1981a; Arad etal., 1973). The strain rate effect arises because the toughness depends strongly on whether there is sufficient time for the crazes or other forms of plasticity at the crack tip to develop fully; the larger the amount of plasticity, the higher the toughness will be. (3) The effect of molecular weight. gc is relatively independent of the molecular weight above a critical value Mc = 2M e , where M e is the entanglement molecular weight, but diminishes below this value. For PMMA it has been found that a relationship between the number average molecular weight Mn and gc holds: gc = A — (B/Mn) (Berry, 1964) where A and B are constants. The reason why gc is low when Mn is less than Mc was thought to be the increased probability of finding molecular chain ends inside the crazes; thus, it is more difficult to form fibrils bridging the entire thickness of the crazes (Kramer, 1978; Haward and Daniels, 1978). Kramer's work, which is described in Sec. 15.2, shows that crazes are stabilized by entanglements (Kramer and Berger, 1990). For those polymers with low entanglement densities, having a high Mn causes the polymer to preferentially fail by crazing which involves chain scission, thus increasing the crazing strength. For those polymers that have high entanglement densities, having a high enough M n prevents crazing from becoming the easier deformation process, so shear yielding occurs instead. Therefore, regardless of the mode of deformation preferred by a given polymer, satisfying the relationship Mn > Mc always results in a higher toughness.
762
15 Crazing and Fracture of Polymers
15.4 Concluding Remarks We have attempted to outline for the reader in a limited amount of space the extremely complicated phenomena of crazing and fracture of polymers. Although we have attempted to give a generally accepted overview of these phenomena, some controversies regarding both the phenomenology and the underlying mechanisms undoubtedly still exist, and many specific experimental observations pertaining to different materials, morphologies, environmental conditions, stress or strain histories have been omitted. Certain aspects of the microscopic mechanisms of crazing have been carefully explained via experiments and theories by Kramer and his outstanding collaborators at Cornell. But much of the remainder of this vast subject still awaits clarification and the development of models. The reader who is interested in exploring this subject further is encouraged to consult the general reference listed at the end of this Chapter.
15.5 Acknowledgements The authors are indebted to Dr. Liang Bao Liu for rendering into English a first draft of this review from the Japanese original (Narisawa, 1982). They are also grateful to Ms. Sonia Janich and Ms. Susan Wright for their help in editing the manuscript.
15.6 References Andrews, E. H., Bevan, L. (1972), Polymer 13, 337. Andrews, E. H., Levy, G. M. (1974), Polymer 15, 599. Arad, S., Radon, J. C , Culver, L. E. (1973), / Appl. Polym. Sci. 17, 1467. Argon, A. S. (1974), The Solid State of Polymers: Geil, P. H., Baer, E., Wada, Y. (Eds.). New York: Marcel Dekker, p. 573.
Argon, A. S., Hannoosh, J. G. (1977), Phil. Mag. 36, 1195. Argon, A. S., Salama, M. M. (1976), Mater. Sci. Eng. 23, 219. Argon, A. S., Salama, M. M. (1977), Phil. Mag. 36, 1217. Argon, A. S., Hannoosh, J. G., Salama, M. M. (1977), Fracture 1977, ICF4, Waterloo, Canada, Vol. 1. New
York: Pergamon, p. 445. Bandyopadhyay, S., Brown, H. R. (1981), /. Polym. Sci. Phys. 19, 749. Beardmore, P., Rabinowitz, S. (1975), J. Mater. Sci. 10, 1763. Becht, I, Fischer, H. (1969), Roll. Z. Z. Polym. 229, 167. Becht, I, Fischer, H. (1970), Roll. Z. Z. Polym. 240, 776. Behan, P., Bevis, M., Hull, D. (1975), Proc. Roy. Soc. A 343, 525. Benbow, J. J. (1961), Proc. Roy. Soc. 78, 970. Benbow, J. X, Roesler, F. C. (1957), Proc. Roy. Soc. 70 B, 201. Berger, L. L., Kramer, E. J. (1988), J. Mater. Sci. 23, 3536. Bernier, G. A., Kambour, R. P. (1968), Macromol. 1, 393. Berry, J. P. (1961a), J. Polym. Sci. 50, 107. Berry, J. P. (1961b), /. Polym. Sci. 50, 313. Berry, J. P. (1961c), SPE J. Trans. 1, 3. Berry, J. P. (1962), J. Appl. Phys. 33, 1741. Berry, J. P. (1963), /. Appl. Phys. 34, 62. Berry, J. P. (1964), /. Polym. Sci. A2, 4069. Bevan, L. (1982), J. Appl. Polym. Sci. 27, 4263. Brady, T. E., Yeh, G. S. Y. (1973), /. Mater. Sci. 8, 1083. Breuer, H. (1979), Angew. Makromol. Chem. 78, 45. Breuer, H. (1985), in: Polymer Blends and Mixtures: Walsh, D. I, Higgins, J. S., Maconnachie, A. (eds.). Dordrecht: Martinus Nijhoff, p. 375. Broutman, L. J., McGarry, F. C. (1963), J. Appl. Polym. Sci. 9, 589. Brown, H. R., Chin, T. H. (1980), J. Mater. Sci. 15, 611. Brown, H. R., Stevens, G. (1978), J. Mater. Sci. 13, 2373. Brown, H. R., Ward, I. M. (1973), Polymer 14, 469. Campbell, D., Peterlin, A. (1968), /. Polym. Sci. B6, 481. Camwell, L., Hull, D. (1973), Phil. Mag. 27, 1135. Chan, M. K. V., Williams, J. G. (1981), Polym. Eng. Sci. 21, 1019. Chan, T, Donald, A. M., Kramer, E. J. (1981), J. Mater. Sci. 16, 676. Chau, C. C , Li, J. C. M. (1976), J. Mater. Sci. 11,434. Chau, C. C , Li, J. C. M. (1980), /. Mater. Sci. 15, 1898. Chau, C. C , Li, J. C. M. (1981), J. Mater. Sci. 16, 1858. Cherry, B. W, Thomson, K. W. (1981a), J. Mater. Sci. 16, 1913.
15.6 References
Cherry, B. W., Thomson, K. W. (1981b), /. Mater. Sci. 16, 1925. Cornes, P. L., Haward, R. N. (1974), Polymer 15, 149. Cornes, P. L., Smith, K., Haward, R. N. (1977), /. Polym. Sci. Phys. 15, 955. Cross, A., Haward, R. N. (1973), /. Polym. Sci. Phys. 11, 2423. DeVries, K. L. (1971), / Polym. Sci. C32, 325. DeVries, K. L., Farris, S. J. (1970), Intern. J. Fract. Mech. 6,411. DeVries, K. L., Lloyd, B. A., Williams, M. L. (1971), /. Appl. Phys. 42, 4644. Doll, W, Weidmann, G. W. (1979), Prog. Colloid. Polym. Sci. 66, 291. Donald, A. M., Kramer, E. J. (1981), Phil. Mag.A43, 857. Donald, A. M., Kramer, E. J. (1982), J. Mater. Sci. 17, 1871. Doyle, M. J. (1975), J. Polym. Sci. Phys. 13, 2429. El-Hakeem, H., Marshall, G. P., Zichy, E. L., Culver, L. E. (1975), J. Appl. Polym. Sci. 19, 3039. Fellers, J. F, Kee, B. F. (1974), J. Appl. Polym. Sci. 18, 2355. Fernando, P. L., Williams, J. G. (1980), Polym. Eng. Sci. 20, 215. Fields, R. X, Ashby, M. F. (1976), Phil. Mag. 33, 33. Frazer, R. A. W, Ward, I. M. (1978), Polymer 19, 220. Friedrich, K. (1977), Fracture 1977, ICF4, Waterloo, Canada, Vol. 3. New York: Pergamon, p. 1119. Friedrich, K. (1983), Adv. Polym. Sci. 52/53, 225. Garton, A., Carlsson, D. I, Sturgeon, P. Z., Wiles, D. M. (1977), J. Polym. Sci. Phys. 15, 2013. Garton, A., Carlsson, D. I, Stepaniak, R. F , Wiles, D. M. (1978), Polym. Eng. Sci. 18, 923. Gent, A. N. (1970), J. Mater. Sci. 5, 925. Gent, A. N. (1973), J. MacromoL Sci. - Phys. B8, 597. Glad, M. D. (1986), Ph.D. Thesis, Cornell University. Ithaca, NY. Grade, A. M., Weiss, V. (1972), Metall. Trans. 3, 2810. Graham, I. D., Williams, J. G., Zichy, E. L. (1976), Polymer 17, 439. Griffith, A. A. (1921), Phil. Trans. Roy. Soc. A221, 163. Gupta, S. K., Quesnel, D. J., Lambropoulos, J. (1986), /. Appl. Polym. Sci. 32, 5585. Hahn, G. T, Rosenfeld, A. R. (1965), Acta Metall. 13, 293. Hashemi, S., Williams, J. G. (1986), /. Mater. Sci. 20, 922. Haward, R. N., Daniels, H. E. (1978), /. Polym. Sci. Phys. 16, 1169. Henkee, C. S., Kramer, E. J. (1984), J. Polymer Sci., Polym. Phys. Ed., 22, 721. Henry, L. F. (1974), Polym. Eng. Sci. 14, 167. Higuchi, M. (1958), Repts. Res. Inst. Appl. Mech., Kyushu Univ. 6, 173. Hill, R. (1950), The Mathematical Theory of Plasticity. Oxford: Oxford Univ. Press.
763
Hoarse, J., Hull, D. (1972), Phil. Mag. 26, 443. Hobbs, S. Y, Bopp, R. C. (1980), Polymer 21, 559. Hoffman, R. D., Richmond, O. (1976), J. Appl. Phys. 47, 4289. Hsiao, C. C , Sauer, J. A. (1959), J. Appl. Phys. 21, 1071. Hull, D. (1970), J. Mater. Sci. 5, 357. Irwin, G. R. (1948), Fracturing of Metals. Cleveland: ASM, p. 152. Irwin, G. R., Kies, J. A., Weld, J. (1952), Res. Suppl. 31, 95. Irwin, G. R., Kies, J. A., Smith, H. L. (1958), Proc. ASTM 58, 640. Ishikawa, M., Narisawa, I. (1983), /. Mater. Sci. 18, 1947. Ishikawa, M., Narisawa, I., Ogawa, H. (1976), Polym. J. 8, 391. Ishikawa, M., Narisawa, I., Ogawa, H. (1977), /. Polym. Sci. Phys. 15, 1791. Ishikawa, M., Ogawa, H., Narisawa, I. (1981), J. MacromoL Sci. Phys. B19, 421. Israel, S. J., Thomas, E. L., Gerberich, W. W (1979), J. Mater. Sci. 14, 2128. Jacques, C. H. M., Wyzgoski, M. G. (1979), J. Appl. Polym. Sci. 23, 1153. Kambour, R. P. (1964a), /. Polym. Sci. A2, 4159. Kambour, R. P. (1964b), Polymer 5, 143. Kambour, R. P. (1965), /. Polym. Sci. A3, 1713. Kambour, R. P. (1966a), /. Polym. Sci. A-2, 4, 17. Kambour, R. P. (1966 b), J. Polym. Sci. A-2, 4, 349. Kambour, R. P. (1973), J. Polym. Sci. D., Macromolecular Reviews, Vol. 7, 1. Kambour, R. P., Kopp, R. W. (1969), J. Polym. Sci. A-2, 7, 183. Kambour, R. P., Miller, S. (1976), J. Mater. Sci. 11, 823. Kambour, R. P., Smith, S. A. (1979), 4th Int. Conf Deform. Yield, Fract. of Polymers, Churchill College, Cambridge. Kambour, R. P., Yee, A. F. (1981), Polym. Eng. Sci. 21, 218. Kambour, R. P., Gruner, C. L., Romagosa, E. E. (1973), J. Polym. Sci. Phys. 11, 1879. Kambour, R. P., Holik, A. S., Miller, S. (1978), /. Polym. Sci. Phys. 16, 91. Kausch, H. H. (1970), Revs. MacromoL Chem. C4, 243. Kausch, H. H. (1976), Kunststoffe 66, 538. Kausch, H. H., Becht, J. (1970), Rheol. Acta. 9, 137. Kausch, H. H., Becht, J. (1972), Roll. Z. Z. Polym. 250, 1048. Kitagawa, M. (1972), Zairyo {Materials, in Japanese) 21, 1012. Kitagawa, M., Montomura, K. (1974), J. Polym. Sci. Phys. 12, 1979. Kitagawa, M., Motomura, K. (1976), Nihoonkiron (in Japanese) 42, 676. Knauss, W. G. (1965), Proc. 1st Int. Conf Fract., Sendai, Vol. 2, p. 1139. Kramer, E. J. (1975), /. Polym. Sci. Phys. 13, 509. Kramer, E. X (1978), J. Mater. Sci. 14, 1381.
764
15 Crazing and Fracture of Polymers
Kramer, E. J. (1983), Adv. Polym. Sci. 52/53, 1. Kramer, E. I, Berger, L. L. (1990), Adv. Polym. Sci. 91/92, 1. Kramer, E. X, Bubeck, R. A. (1978), /. Polym. Sci. Phys. 16, 1195. Kramer, E. I, Krenz, H. G., Ast, D. G. (1978), /. Mater. Sci. 13, 1093. Krenz, H. G., Ast, D. G., Kramer, E. J. (1976), /. Mater. Sci. 11, 2198. Lauterwasser, B. D., Kramer, E. J. (1979), Phil. Mag. A 39, 469. Lloyd, B. A., DeVries, K. L., Williams, M. L. (1972), /. Polym. Sci. A-2, 10, 1415. Marshall, G. P., Culver, L. E., Williams, J. G. (1970), Proc. Roy. Soc. Lond. A 319, 165. Marshall, G. P., Culver, L. E., Williams, J. G. (1973), Intern. J. Fract. 9, 295. Matsushige, K., Baer, E., Radcliffe, S. V. (1975), /. Macromol. Sci. Phys. Bll, 566. Meinel, G., Peterlin, A. (1968), /. Polym. Sci., A-2, 6, 587. Meinel, G., Peterlin, A., Sakaoku, K. (1968), Analytical Calorimetry: Porter, R. S., Johnson, J. F. (Eds.). New York: Plenum Press, p. 15. Mills, N. J. (1976), /. Mater. Sci. 11,363. Mills, N. X, Walker, N. (1976), Polymer 17, 335. Miltz, X, DiBenedetto, A. T, Petrie, S. (1978), J. Mater. Sci. 13, 2037. Misaki, T., Hirohata, T., Yoshii, M., Hamasaki, T. (1989), / Appl. Polym. Sci. 37, 2617. Morgan, G. P., Ward, I. M. (1977), Polymer 18, 87. Murray, X, Hull, D. (1969), Polymer 10, 431. Murray, X, Hull, D. (1970a), /. Polym. Sci. A-2, 8, 159. Murray, X, Hull, D. (1970b), J. Polym. Sci., A-2, 8, 583. Murray, X, Hull, D. (1970c), /. Polym. Sci. A-2, 8, 1521. Murray, X, Hull, D. (1971), J. Mater. Sci. 6, 1277. Nagamura, T., Takayanagi, M. (1974), /. Polym. Sci. Phys. 12, 2019. Nagamura, T., Fukitani, K., Takayanagi, M. (1975), J. Polym. Sci. Phys. 13, 1515. Nagamura, T, Kusumoto, N., Takayanagi, M. (1973), J. Polym. Sci. Phys. 11, 2357. Narisawa, I. (1972 a), Zairyokyodogakkai-shi {Proc. Jpn. Soc. Strength Mat., in Japanese) 7, p. 8. Narisawa, I. (1972b), J. Polym. Sci. A-2, 10, 1789. Narisawa, I. (1982), Strength and Fracture of Polymers (in Japanese). Tokyo: Ohm Publishing Co., p. 162. Narisawa, I., Ishikawa, M. (1990), Adv. Polym. Sci. 91/92, 353. Narisawa, I., Kondo, T. (1972), Zairyo (Materials, in Japanese) 21, 315. Narisawa, I., Kondo, T. (1973), /. Polym. Sci. Phys. 11, 223. Narisawa, I., Ishikawa, M., Ogawa, H. (1976), Polym. J. 8, 181.
Narisawa, L, Ishikawa, M., Ogawa, H. (1977), J. Polym. Sci. 15, 2227. Narisawa, I., Ishikawa, M., Ogawa, H. (1978 a), Zairyokyodagakkai-shi (Proc. Jpn. Soc. Strength Mat., in Japanese) 12, p. 156. Narisawa, I., Ishikawa, M., Ogawa, H. (1978 b), Zairyokyodogakkai-shi (Proc. Jpn. Soc. Strength Mat., in Japanese) 13, p. 104. Narisawa, I., Ishikawa, M., Ogawa, H., Sato, T. (1979), Materials 28, 543. Narisawa, I., Ishikawa, M., Ogawa, H. (1980a), J. Mater. Sci. 15, 2059. Narisawa, I., Ishikawa, M., Ogawa, H. (1980b), Phil. Mag. A41, 331. Narisawa, I., Ishikawa, M., Murayama, B., Ogawa, H. (1980 c), Karon 36, 331. Newmann, L. V., Williams, X G. (1980), /. Mater. Sci. 15, 773. Nimmer, R. P. (1991), General Electric, private communication. Olf, H. G., Peterlin, A. (1974), /. Polym. Sci. Phys. 12, 2209. Orowan, E. (1949), Repts. Prog. Phys. 12, 185. Oxborough, R. X, Bowden, P. B. (1973), Phil. Mag. 28, 547. Oxborough, R. X, Bowden, P. B. (1974), Phil. Mag. 30, 111. Parvin, M. (1981), J. Mater. Sci. 16, 1796. Parvin, M., Williams, X G. (1975 a), Intern. J. Fract. 11, 963. Parvin, M., Williams, X G. (1975b), /. Mater. Sci. 10, 1883. Peterlin, A. (1969), /. Polym. Sci. A-2, 7, 1151. Peterlin, A. (1971a), J. Polym. Sci. C-32, 297. Peterlin, A. (1971b), Intern. J. Fract. Mech. 7, 496. Peterlin, A. (1972), J. Macromol. Sci. Phys. B-6, 583. Peterlin, A. (1974a), Pure and Appl. Chem. 39, 239. Peterlin, A. (1974 b), The Solid State of Polymers: Geil, P. H., Baer, E., Wada, Y. (Eds.). New York: Marcel Dekker, p. 83. Peterlin, A. (1975), Int. J. Fract. 11, 761. Peterson, T. L., Ast, D. G., Kramer, E. X (1974), /. Appl. Phys. 45, 4420. Phillips, D. C , Scott, X M., Jones, M. (1978), /. Mater. Sci.
13,311.
Pitman, G. L., Ward, I. M. (1979), Polymer 20, 895. Plummer, C. X G., Donald, A. M. (1989), /. Polym. Sci., Polym. Phys. Ed. 27, 327. Regel, V. R. (1956), /. Tech. Phys. USSR 26, 359. Rudd, X F. (1973), Polym. Lett. 1, 1. Saffman, P. G., Taylor, G. I. (1958), Proc. Roy. Soc. Lond. A 245, 312. Sato, K. (1965), Kobunshi Kagaku (Polym. J, in Japanese) 22, 148. Sato, Y. (1966), Kobunshi Kagaku (Polym. J, in Japanese) 23, 69. Sauer, H. A., Marin, X, Hsiao, X X (1949), J. Appl. Phys. 20, 507. Sauer, X A., Hsiao, C. C. (1953), Trans. ASME75, 895.
15.6 References
Serrano, A. M., Welsch, G. E., Gibala, R. (1982), Polym. Eng. Sci. 22, 946. Shirrer, R., Goett, C. (1980), Intern. J. Fract. 16, R133. Smith, K., Haward, R. N. (1977), Polymer 18, 745. Smith, T. L. (1958), J. Polym. Sci. 32, 99. Smith, T. L. (1965), Polym. Eng. Sci. 5, 270. Smith, T. L. (1969), Rheology, Theory and Application, Vol. 5: Eirich, F. R. (Ed.). New York: Academic Press, p. 127. Sternstein, S. S., Meyers, F. A. (1974), The Solid State of Polymers: Geil, P. H., Baer, E., Wada, Y. (Eds.). New York: Marcel Dekker, p. 539. Sternstein, S. S., Onchin, L., Silverman, A. (1963), Appl. Polym. Symp. 7, 175. Svensson, N. L. (1961), Proc. Roy. Soc. 77, 876. Takahashi, K., Kimura, M., Hyodo, S. (1978), J. Mater. Sci. 13, 2718. Ting, R. Y. (1981), /. Mater. Sci. 16, 3059. Truss, R. W, Chadwick, G. A. (1977), J. Mater. Sci. 12, 1383. Valanis, K. C , Yilmazer, U. (1978), Eng. Fract. Mech. 10, 659. Verheulpen-Heymans, N. (1979), Polymer 20, 356. Verheulpen-Heymans, N., Bauwens, J. C. (1976a), /. Mater. Sci. 11, 1. Verheulpen-Heymans, N., Bauwens, J. C. (1976 b), J. Mater. Sci. 11, 1. Verma, G. S. P., Peterlin, A. (1970), Roll. Z. Z. Polym. 236, 111. Verma, G. S. P., Peterlin, A. (1972), /. Macromol. Sci.-Phys. B6, 583. Vincent, P. L, Raha, S. (1972), Polymer 13, 283. Walker, N., Haward, R. N., Hay, I N . (1979a), J. Mater. Sci. 14, 1085. Walker, N., Hay, J. N., Haward, R. N. (1979 b), Polymer 20, 1056. Walker, N., Haward, R. N., Hay, J. N. (1981), /. Mater. Sci. 16, 817. Wang, T. T., Matsuo, M., Kwei, T. K. (1971), J. Appl. Phys. 42,4188. Weidmann, G. W, Doll, W (1976), Colloid Polym. Sci. 254, 205. Weidmann, G. W, Doll, W. (1978), Intern. J. Fract. 14, R189. Wellinghoff, S. T, Baer, E. (1975). J. Macromol. Sci. Phys. Bll, 367. Wellinghoff, S. T, Baer, E. (1978), /. Appl. Polym. Sci. 22, 2025.
765
Wu, G. M., Shultz, J. M. (1989), Polym. Eng. Set 29, 405. Wyzgoski, M. G., Yeh, G. S. Y (1974), J. Macromol. Sci. Phys. B10, 641. Yaffe, M. B., Kramer, E. J. (1981), /. Mater. Sci. 16, 2130. Yamani, S., Young, R. J. (1977), Polymer 18, 1075. Yamani, S., Young, R. J. (1979), / Mater. Sci. 14, 1609. Yang, A.-C. M., Kramer, E. I, Kuo, C. C , Phoenix, S. L. (1986a), Macromolecules 19, 2010. Yang, A.-C. M., Kramer, E. 1, Kuo, C. C , Phoenix, S. L. (1986b), Macromolecules 19, 2020. Yokobori, T. (1975), Strength of Materials (in Japanese), 2nd ed. Tokyo: Iwanami, p. 121. Young, R. J., Beaumont, W. R. (1976), Polymer 17, 111. Zhurkov, S. N., Kuksenko, V. S. (1965), Intern. J. Fract. Mech. 1, 629. Zhurkov, S. N., Tomashevskii, E. E. (1966), Proc. of the Conf on the Physical Basis of Yield and Fracture, Oxford: Stickland, A. C. (Ed.). London: Institute of Physics, p. 200. Zhurkov, S. N., Tomashevskii, E. E., Zakrevskii, V. A. (1964), Fiz. Tver. Tera. 6, 1912. Zhurkov, S. N., Kuksenko, V. S., Slutsker, A. I. (1969b), Fracture: Pratt, P. L. (Ed.). London: Chapman and Hall, p. 531. Zhurkov, S. N., Vettegren, V. I., Korsukov, E., Novak, 1.1. (1969a), Fracture: Pratt, P. L. (Ed.). London: Chapman and Hall, p. 545.
General Reading Kambour, R. P. (1973), A Review of Crazing and Fracture in Thermoplastics, J. Polym. Sci.: Macromol. Rev. 7, 1. Kausch, H. H. (Ed.) (1983), Crazing in Polymers, Vol. 52/53 of Adv. Polymer Sci. Berlin: SpringerVerlag. Kausch, H. H. (Ed.) (1990), Crazing in Polymers, Vol. 2, Vol. 91/92 of Adv. Polymer Sci. Berlin: Springer-Verlag. Kausch, H. H. (1978), Polymer Fracture. Berlin: Springer-Verlag. Kinloch, A., Young, R. (1984), Fracture of Polymers. London: Applied Science.