The integrand has a sharp peak at n=m, so the double integral is
evaluated as
N N oo
dn {dm =N | d(n-m). (4.11.11)
[a fan=n J
The final form is written as
122 |
gk, t) =< | du exp[—u —(Ty)7ACU(T)-))-— (4.00.12)
| « x
where
kaT
Tee
kp? (4.11.13)
and
A(u) =2 | x 28H) (1—exp(-x2)). (4,111.14)
0POLARIZABILITY TENSOR OF A GAUSSIAN CHAIN 135
For I, >1, this expression is further simplified to
a(k, #)= atk, 0) { du expl—u — Cu)!*A(0)]
=a(k, Ojexp| - a ro] . (4.11115)
For the Zimm model, the structure factor can be obtained in the same
way if the linearization approximation“ is used. The result for the ©
condition is
a(k, ) =8(k, 0) f du exp[—u — (Tt) *h(u(Tyt)™)] (4.11.16)
o
with
_keT
Te Sn, Re (4.11.17)
h(u) =2 | 2) —exp(-27/V2)). (4.11.18)
oO
For [,t>>1,
g(k, t) = g(k, O)exp(—1.35(T,£)”"). (4.11.19)
Appendix 4.1V_ Polarizability tensor of a Gaussian chain
Let aag(r, N) be the average polarizability of a polymer chain consisting
of N units with its end-to-end vector being fixed at r. By symmetry,
g(r, N) is written as
aap(r, N)= AC, N)atp — 366") (4.1V.1)
where A(r’, N) is a scalar. Equation (4.IV.1) can be expanded with
respect to 77:
Gap(t, N) = (ao(N) + a,(N)P? + ...)(ratp —45ag"). (4.1V.2)
We shall show that ao(N) is proportional to 1/N, and a,(N) = a,(N) =
... =0 provided that (i) the statistics of the chain are Gaussian and that
(ii) the polarizability is additive.
The second condition is written as follows. Suppose that the chain is
divided into two parts each consisting of N, and N, units, then
eran ts Ni + Ns) = [ar drs (r., Ma) + dope, NJIO(N, ) (4.1V.3)136 DYNAMICS OF FLEXIBLE POLYMERS
where ®(r,,™) is the equilibrium distribution of r, and 7, under the
constraint that r, + 7 is fixed at 7. For the Gaussian chain, this is given by
O(n, m) = O( +m — PVH, Ni) (H, No)/(r, N+ Nz) (4-IV.4)
where n 2
3 3
Wr, N)= (=) exp( -s) . (4.1V.5)
If we multiply eqn (4.IV.3) by ep (r, N, + Np) and integrate over r, we
have
Gap (ky Ni + Nz) P(k, Ny + No) = B(k, N2)&ap(k, Ni) Pk, Ni)
+ i(k, Ni)dop(k, N2P(k, N2) (4.1V.6)
where .
W(k, N) = exp(—Nb7k?/6) (4.1V.7)
and & p(k, N) is a differential operator:
= F a0, #
Sap(k, N)= ~(a0- asst. Mae 3g ne az) (4.1V.8)
Using eqn (4.IV.7), eqn (4.IV.6) is written as
2 2
exp(= RN, + Na) Gap (k N+ N)exp( - Fem + ™))
2 2
= exp(= IN, ) ap ky Nvexo( ae eN,)
6 6
Ba). e,
+ exp( k N:) ap Naexp( — ze ) . (41V.9)
For this to be true for arbitrary N, and N,, it must follow that
2 2
exp(= en) dap(k, Nyexp( - ew) =NByp(k) (4.1V.10)
where B,(k) is a certain function of k. Substitution of eqn (4.IV.8) and
a straightforward calculation gives
4 4
= N(kakp - 3baph?)| aol) + (we - 30N)a\(N) tee | = NByg(k).
(4.1V.11)
Comparison of the coefficients of the power of N indicates that this
equality is satisfied only when
aN) =co/N, a(N)=a,(N)=...=0. —(4.1V.12)REFERENCES 137
Choosing the constant cy as Ay/b, we have
A
ean 05 N) = pp (Fala — 48 op?)
Ay
= Pn (ratg — 45ap?”) (4.1V.13)
which is eqn (4.215).
References
1. Rouse, P. E., J. Chem. Phys. 21, 1272 (1953).
2. Verdier, P. H., and Stockmayer, W. H., J. Chem. Phys. 36, 227 (1962); and
ona
FPS pon
ee
Verdier, P. H., J. Chem. Phys. 52, 5512 (1970).
. Orwoll, R. A., and Stockmayer, W. H., Adv. Chem. Phys. 15, 305 (1969).
. Iwata, K., J. Chem. Phys. 54, 12 (1971).
Edwards, S. F., and Goodyear, A. G., J. Phys. A5, 965, 1188 (1972).
See general review, Stockmayer, W. H., Dynamics of chain molecules, in
Molecular Fluids (eds R. Balian and G. Weill). Gordon & Breach, New York
(1976).
. Hilhorst, H. J., and Deutch, J. M., J. Chem. Phys. 63, 5153 (1975).
. Kuhn, W., Kolloid Z. 68, 2 (1934).
. Sole, K., and Stockmayer, W. H., J. Chem. Phys. 54, 2756 (1971); see also
Doi, M., and Nakajima, H., Chem. Phys. 6, 124 (1974).
. Zimm, B. H., J. Chem. Phys. 24, 269 (1956).
. Zimm, B. H., Roe, G. M., Epstein, L. F., J. Chem. Phys. 24, 279 (1962).
Hearst, H. E., J. Chem. Phys. 37, 2547 (1962).
. Yamakawa, H., Modern Theory of Polymer Solutions, Chapter 6. Harper &
Row, New York (1971).
|. Edwards, S. F., and Freed, K. F., J. Chem. Phys. 61, 1189 (1974).
|. Edwards, S. F., and Muthukumar, M., Macromolecules 17, 586 (1984).
. Peterlin, A., J. Chem. Phys. 23, 2464 (1955).
. Kurata, M., Yamakawa, H., and Teramoto, E., J. Chem. Phys. 28, 785
(1958); Chikahisa, Y., and Tanake, T., J. Polym. Sci. C 30, 105 (1970).
. Matsushita, Y., Noda, I., Nagasawa, M., Lodge, T. P., Amis, E. J., and
Han, C. C., Macromolecules 17, 1785 (1984).
. Tanford, C., Physical Chemistry of Macromolecules. Wiley, New York 1961.
. Oono, Y., and Kohmoto, M., J. Chem. Phys. 78, 520 (1983).
}. de Gennes, P. G., Macromolecules 9, 587, 594 (1976).
. de Gennes, P. G., Scaling Concepts in Polymer Physics, Chapter 6. Cornell
University Press, Ithaca (1979).
. Hohenberg, P. C., and Halperin, B. I., Rev. Mod. Phys. 49, 435 (1977).
. Oono, Y., J. Chem. Phys. 79, 4629 (1983).
|. Ceperley, D., Kalos, M. H., and Lebowitz, J. L., Phys. Rev. Lett. 41, 313
(1978); Macromolecules 14, 1472 (1981).
. Berne, B. J., and Pecora, R., Dynamic Light Scattering. Wiley, New York
(1976).138
BIR
41,
&
DYNAMICS OF FLEXIBLE POLYMERS
. Akcasu, Z., and Gurol, H., J. Polym. Sci. Phys. ed. 14, 1 (1976).
. Kirkwood, J. G., J. Polym. Sci. 12, 1 (1954).
. King, T. A., Knox, A., Lee, W. I., and McAdam, J. D. G., Polymer 14, 151
(1973).
. Adam, M., and Delsanti, M., J. Physique 37, 1045 (1976); Macromolecules
10, 1229 (1977).
. Nose, T., and Chu, B., Macromolecules 12, 590 (1979).
. Han, C. C., and Akcasu, A. Z., Macromolecules 14, 1080 (1981).
. Schmidt, M., and Burchard, W., Macromolecules 14, 210 (1981).
. Stockmayer, W. H., and Hammouda, B., Pure & Appl. Chem. 56, 1373
(1984).
. Guttman, C. M., McCrackin, F. L., and Han, C. C., Macromolecules 15,
1205 (1982).
. Yamakawa, H., J. Chem. Phys. 53, 436 (1970).
. Fixman, M., and Mansfield, M. L., Macromolecules 17, 522 (1984).
. Jasnow, D., and Moore, M. A., J. Phys. (Paris) 38, LA67 (1978); Al-Noaimi,
G. F., Martinez-Mekler, G. C., and Wilson, C. A., J. Phys. (Paris) 39, L373
(1978).
. Lee, A., Baldwin, P. R., and Oono, Y., Phys. Rev. A 30, 968 (1984).
. Tsunashima, Y., Nemoto, N., and Kurata, M., Macromolecules 16, 584, 1184
(1983); Nemoto, N., Makita, Y., Tsunashima, Y., and Kurata, M.,
Macromolecules 17, 425 (1984).
des Cloizeaux, J., J. Physique Let. 39, L151 (1978).
Stockmayer, W. H., and Albrecht, A. C., J. Polymer Sci. 32, 215 (1958).
. Kurata, M., and Yamakawa, H., J. Chem. Phys. 29, 311 (1958).
. Weill, G., and des Cloizeaux, J., J. Physique 40, 99 (1979).
. de Gennes, P. G., Physics 3, 37 (1967); Dubois-Violette, E., and de Gennes,
P. G., Physics 3, 181 (1967).
. Akcasu, A. Z., Benmouna, M., and Han, C. C., Polymer 21, 866 (1980).
. Schaefer, D. W., and Han, C. C., in Dynamic Light Scattering (ed. R.
Pecora). Plenum Press, New York (1985).
. Allen, G., Ghosh, R., Higgins, J. S., Cotton, J. P., Farnoux, B., Jannink,
G., and Weill, G., Chem. Phys. Lett. 38, 577 (1976).
. Richter, D., Hayter, J. B., Mezei, F., and Ewen, B., Phys. Rev. Lett. 41,
1484 (1978).
Nicholson, L. K., Higgins, J. S., and Hayter, J. B., Macromolecules 14, 836
(1981).
. Ferry, J. D., Viscoelastic Properties of Polymers. (3rd edn). Wiley, New York
(1980).
. Johnson, R. M., Schrag, J. L., and Ferry, J. D., Polymer J. 742 (1970).
. Tanaka, H., Sakanishi, A., Kaneko, M., and Furuichi, J., J. Polymer Sci.
C15, 317 (1966); Sakanishi, A., J. Chem. Phys. 48, 3850 (1968).
. Osaki, K., Adv. Polym. Sci. 12, 1 (1973).
. Prager, S., J. Chem. Phys. 75, 72 (1971).
. Fixman, M., J. Chem. Phys. 78, 1588 (1983). Wilemski, G., and Tanaka, G.,
Macromolecules 14, 1531 (1981).
. De Wames, R. E., Holland, W. F., and Shen, M. C., J. Chem. Phys. 40,
2782 (1967); see also Zwanzig, R., Kiefer, J., and Weiss, G. H., Proc. Natl.
Acad. Sci. US 60, 381 (1968).57.
58.
59.
. Fixman, M., J. Chem. Phys. 78, 1594 (1983).
61.
62.
REFERENCES 139
Rotne, J., and Prager, S., J. Chem. Phys. 50, 4831 (1969).
Doi, M., unpublished work.
Zimm, B. H., Macromolecules 13, 592 (1980).
Tsvetkoy, V. N., in Newer Methods of Polymer Characterization (ed. B. Ke),
Polymer Rev. 6. Interscience, New York (1964).
Janeschitz-Kriegl, H., Flow birefringence of elastico viscous polymer solu-
tions. Adv. Polym. Sci. 6, 170 (1969).
. Kuhn, W., Kolloid Z 68, 2 (1934); Kuhn, W., and Grin, F., Kolloid Z. 101,
248 (1942); Kuhn, W., J. Polym. Sci. 1, 360 (1946).
. Volkenstein, M. V., Configurational Statistics of Polymeric Chains.
Interscience, New York (1963).
. Flory, P., Statistical Mechanics of Chain Molecules. Interscience, New York
(1969).
. Copic, M., J. Chem. Phys. 26, 1382 (1957).
. Peterlin, A., and Stuart, H. A., Z. Physik 112, 1 (1939); see also Peterlin,
A., in Rheology, (ed. F. R. Eirich) Vol 1, p. 615. Academic Press, New
York (1956).
. Landau, L. D., and Lifshitz, E. M., Electrodynamics of Continuous Media.
Pergamon Press, Oxford (1960).
. Onuki, A., and Kawasaki, K., Physica 111A, 607 (1982); Onuki, A., and
Doi, M., J. Chem. Phys. (1986) to be published.
. Koyama, R., J. Phys. Soc. Jpn 16, 1366 (1961); 19, 1709 (1964). Daudi, S., J.
Physique 38 1301 (1977).
. Janeschitz-Kriegl, H., Macromol. Chem. 40, 140, (1959).
. Frisman, E. V., and Tsvetkov, V. N., J. Polym. Sci. 30, 297 (1958).~
MANY CHAIN SYSTEMS
5.1 Semidilute and concentrated solutions
The physical properties of a polymer solution depend on solvent,
temperature, and concentration. The solvents for polymers are broadly
classified into two categories, good and poor solvents. Good solvents
have strong attractive energy with polymers and dissolve polymers over a
wide range of temperature. In such a solvent, the net interaction between
polymer segments is repulsive (since they tend to contact with solvent
molecules rather than themselves), and the excluded volume parameter v
is positive and large. Poor solvents, on the other hand, are less keen to
accommodate polymers, and precipitate polymers when the temperature
is changed or the polymer concentration is increased. The excluded
volume v in such a solvent can be positive, zero, or negative, depending
on the temperature.
For the purpose of the discussion, polymer solutions in good solvents
can be divided into three regions: dilute, semidilute, and concentrated
(see Fig. 5.1).
A dilute solution is defined as one of sufficiently low concentration that
the polymers are separated from each other; each polymer on average
occupying a spherical region of radius R,. (Fig. 5.1a) In this solution, the
polymer-polymer interaction has only a small effect, and any physical
property is expressed as a power series with respect to the polymer
concentration p (weight of polymer in unit volume). Consider for
example the osmotic pressure II and the viscosity 7. In the dilute limit,
these are written as
l= Gar. (5.12)
n=n,(1 + e[n]) (5.15)
where R is the gas constant, M the molecular weight, T the temperature
and 7,, the solvent viscosity. Equation (5.1@) is van’t Hoff’s law and eqn
(5.1b) is the definition of the intrinsic viscosity [7]. The interaction
among the polymers gives terms of order p”, so that eqns (5.1a) and
(5.16) become
m= Rr(E+ Asp? + . ) (5.22)
n=n(1+ pln] + kxe'[nF +. ..)- (5.2b)SEMIDILUTE AND CONCENTRATED SOLUTIONS 141
e(x) ex) e(x)
é
é
é
x x x
@ ()
Fig. 5.1. Three concentration regimes in good solvent: (a) dilute, (b) semidilute,
and (c) concentrated. c(x) denotes the concentration profile along the dot-dashed
lines.
(o)
The parameters A, and k;, are called the second virial coefficient and the
Huggins coefficient, respectively.
As the concentration increases, the polymer coils come closer and start
to overlap each other. Since the number of polymers in a unit volume is
pNg/M the concentration p* at which the overlap starts is estimated as
p*Na
yr Re= 1. (5.3)
Note that the concentration p* can be quite low. As R, is proportional to
M”, p* depends on the molecular weight as
p*txM'3=M-"5 (for v = 3/5). (5.4)
Thus for large molecular weight, p* becomes quite small; e.g., for
polystyrene of M~10°, p* is about 0.005g/ml (about 0.5% in weight).
Hence we can easily get a solution in which the molecules are strongly
overlapped, but still occupy a small volume fraction. Such a solution is
called semidilute.
A semidilute solution is characterized by the large and strongly
correlated fluctuations in the segment density such as we have in dilute
solutions. Although the fluctuations decrease with increased polymer
concentration, the semidilute solution still retains the same character as
critical phenomena in statistical mechanics, and presents the same kind of
problem as we discussed in Sections 2.5 and 2.6.!
If the concentration becomes sufficiently large, the fluctuations become
small and can be treated by a simple mean field theory. Such a solution is
called concentrated. As we shall show later, the cross-over concentration142 MANY CHAIN SYSTEMS
from semidilute to concentrated is estimated by
pre = OMAN) (5.5)
where M, is the molecular mass of the segment. The high concentration
limit is the melt in which there are no solvent molecules.
We may thus classify the polymer solutions in good solvent into three
Tegimes:
(i) dilute; p , 7p2 Ran ~ Ron-1)° (5.10)
an
is the energy of the chain connectivity and
U,[RanV/keT =4>) v5(Ran — Rom) (5.11)
a,b
am144 MANY CHAIN SYSTEMS
is the excluded volume interaction, which includes both the intramolecu-
lar interaction a = b and the intermolecular interaction a+b. Since eqn
(5.11) uses the pseudo potential, this model needs some alterations at
high concentration where the detailed chemical structure matters, but
here we will proceed using the simplest form.
Although the starting point is clear, theoretical development is not
easy since calculation of any average from eqn (5.9) involves a highly
nonseparable integration which is common to many body problems.
However, under certain conditions we can progress by using collective
coordinates.”
Instead of describing the problem in terms of Ru,, we focus our
attention on the local segment density c(r) defined by
cr) = ¥ (7 — Ran) (5.12)
and consider the distribution function V{c(r)] for c(r). This method is
effective if the physical quantiy under consideration is expressed by c(r)
For the mathematical development, it is convenient to use the Fourier
transform of ¢(r):
= < | dr exp(ik + r)e(r) = z Sesplik - Ror] (5.13)
c= See eme(-ik- = Ges | dkcye*"* (5.14)
where V is the volume of the system.
The component co is equal to the average concentration c, and does
not fluctuate at all, while c, with k #0 denotes the fluctuation of the local
concentration.
A technical remark has to be made here. In the representation of {c,},
not all c, are independent of each other since c, and c_, are related as
Cy~=CE (5.15)
where cj is the complex conjugate of c,.
We shall choose those c, with positive k, values as the independent
components, and use the abbreviated symbol k >0 to denote the set of
the independent components. When the component c, with negative k,
appears in the subsequent calculation, it should be remembered that it is
cy. Thus for example
T= > ccs (5.16)
k>0GAUSSIAN APPROXIMATION 145
is the same as the sum
=> > 5 act (6.17)
K,>0 ky=— kyo
which is expressed by the sum over the entire k space as
> CyC_4 = a> 40-4 — 308= ay cyck — 4c5. (5.18)
k>o k
Now the distribution function of c, is given by
W(4})% { TT 8Ran exp(—(UoLRon] + UilRon)V/kaT)
} en
xT (ce -ZDewlik- R..)) (5.19)
By use of the formula
o(r-r')= > exp(ik -(r—r')), (6.20)
U, [Ran] is rewritten as
Ui(RonlkaT = 5555, 2, explik* Ran — Rim]
k ab
vw 4 . 1 .
=2 (Fz exp(ik RFE, exp(-ik Rin). (6.21)
If this is substituted into eqn (5.19), the underlined terms can be replaced
by c, and c_, because of the delta functions. Equation (5.19) is thus
Tewritten as
W({cx}) & exp[-Fed cxc-a|
x | TT 6Ran exp(—UolRanl/keT) TT 5(c - 3 Sexplik . Ra») 6.20)
° k>0 an
Let Up({c,}) and U,({c,}) be defined by
Up({cx})/keT = -inf T] 6... exp(—UclRen]/ksT)
1 .
x TT 6(cx—pEexrGkRen)) (5.23)
U({cx})/kaT = x Dy CeCe (5.24)
then
W({cx}) © exp[—(Uol {cg}) + Ui({ce}))/ka TI- (5.25)146 MANY CHAIN SYSTEMS
Calculation of Uo({c,}) is difficult and it is only possible to give the
explicit form as a series. A systematic method? is described in Appendix
5.1. (Another method is described in ref. 11). Here we use the Gaussian
approximation (or random phase approximation), which is to say that the
distribution of c, is Gaussian, i.e., the free energy is written as
Up({cx})/kaT = > AokCeC—« + higher order terms inc, (5.26)
ik
where Ao, is a constant to be determined. This approximation is justified
in the concentrated solutions where the density fluctuations are small,
while it is not adequate in the semidilute solutions.
To determine Ay, we use the fact that for noninteracting polymers
(v=0), (cyc_,) can be calculated exactly as a sum of the correlation
functions of independent Gaussian chains,
(cal—4)o= a0lk) (5.27)
where (...)o denotes the average for the state of v = 0, and go(k) is the
scattering function of ideal polymers (see eqn (2.80)),
__2N [ (-mee NBR | |
sdb) = Gapqap [&%P(——g—) +1}
(5.28)
On the other hand, for v =0, the distribution of c, becomes
Pe((c4)) #exp|~E Aacac-+]=exp|-25) Awcact | (5:29)
which gives (see eqn (2.1.31))
(cxe-4)0= lexi do= 5 (5.30)
Comparing eqn (5.30) with eqn (5.27) we obtain Ao, as
Vv
Ax = Zegolky” (5.31)
From eqns (5.24), (5.26), and (5.31) the total energy
U({c4}) = Uo({en}) + Ui({cx}) (5.32)
is obtained as
Vv 1
UleakeT = 53 (Say t)ae-« (5.33)
For large k, the asymptotic form of go(k) gives
+ 4,-KH,, 5.34)
cBo(k) We (5.34)GAUSSIAN APPROXIMATION 147
This expression is also valid for small k because 1/cgo(0)=1/cN is
negligibly small compared to v in the concentrated solution. Thus for the
entire & values we have
242 2
Ue kaT =ZE (St vreu= ELEC +E Mac» (6.5)
where
& =(b7/12cv)"” (5.36)
is called the correlation length.
The distribution function for c, is
Vu _
W({c4}) & exp| -5 Dae +E "exc (5.37)
2 12c¢
We shall now examine some consequences of this expression.
5.2.2 Pair correlation function
From eqn (5.37), (cyc-x) is given by
1 12¢
(cac—z) “TPE (5.38)
whence the scattering function per segment is
_V _ 2
8(k) = c (cac—«) ~ pe +E) e) (5.39)
The Fourier transform of g(k) gives the pair correlation function:*
~2=c{ p(k)
(e(r)e(0)) -e=ef GE saKe
__ famkedk 12 sin(kr)
)@ay BE+E?) kr
= 5 exr(-r/8), (5.40)
Equation (5.40) indicates the physical significance of &:& represents the
correlation length of the concentration fluctuation.
Experimentally, the correlation length can be determined from the
behaviour of g(k) in the small k region:
0)
SD = 1+ 1Elag for k—0. (5.41)148 MANY CHAIN SYSTEMS
The definition of this apparent correlation length can be extended to
dilute regimes, in which case §,p) gives R,/V3 (see eqn (2.72). The
behaviour of &,p, is entirely different in dilute solutions and concentrated
solutions. In dilute solutions, §,)~R, increases with the molecular
weight and the excluded volume, while in concentrated solutions §,5, = &
is independent of the molecular weight and decreases as a function of
concentration and excluded volume (see eqn (5.36)). The reason can be
easily understood from Fig. 5.1: once polymers overlap each other, the
excluded volume interaction tends to make the concentration
homogeneous.
The simple theory given above is valid only at rather high concentra-
tion or at small excluded volume, i.e., near the © condition. At both
these limits there are additional difficulties. At very high concentration
the precise form of the potential matters. Also, near the © conditions the
precise details of the interaction, in the sense of a cluster expansion
passed to the two-body term, can also matter. In both of these limits it is
possible to make an appropriate improvement,®’ and the results have
been found to be in good agreement with experiments.'”"
5.2.3 Osmotic pressure
The free energy A of the system is obtained from the integration of eqn
(5.9). It is convenient to take the system of non-interacting polymers
(uv =0) as a reference state. The difference in the free energy A — Ao
between the two states is
[ Hoa ex(—(WolRan] + Ui (Ranke)
exp(—(A — Ag)/kgP) = R$ SHH
| TAR. exP(-UolRanl/ko7) (5.42)
which can be rewritten as
exp(—(A — Ao)/kpT)
| Haed(ce- 2 exptk- )) {tor
k>0 Van on ao
X exp(~(UolRaq] + UsfRen) Vo)
| a de45 (ce - z E exp(ik- R.,)) T1 ORe, exp(—(UolRan]/koT)
J Hy dee exp(—(WolCee}) + U({ea)))/ko7)
(5.43)
| By, dex exp(—Uol ex) ka T)GAUSSIAN APPROXIMATION 149
In the Gaussian approximation, the integrals on the right-hand side can
be performed rigorously (see Appendix 5.II). Given the free energy, the
osmotic pressure is calculated by
oA c a
I= - $6 = eT (Z) ~5y(A-A0). (5.44
The result is’
V3 (cv)?
I= ket (5 + 4uct— ver). (5.45)
The first two terms are easily understood: the first term represents van’t
Hoff’s law (c/N)ksT (c/N being the number of polymers in unit volume),
and the second term is the excluded volume interaction between the
segments. The last term represents the correction to the second term due
to the chain connectivity: i.e., the effect that the intramolecular excluded
volume interaction does not contribute to the osmotic pressure.
From eqn (5.45) it will be seen that as c—>0, the osmotic compres-
sibility (@11/9c) becomes negative, which is of course incorrect. It follows
that the theory presented above is only valid if
3/2.
er ie, czc* = (5.46)
vc? 2
This gives eqn (5.5). The situation c (2
k>0
sy tH exp(-ik- R,,)
io Cese-a)
+chu>; exp(ik- R.))|
cconp[ D & (cee-1)o*expl—ik-(R, ~Ry)]|
-en[5D > (cxc-4)v? exp[—ik + (R, ~ I} (5.50)
Hence
WLR] exp| 5520 Ry — Rana)? —4D OR, -Rn) | 51)
where
B(r) = vd(7) - 2 (cxc—4)v?e*"*. (5.52)GAUSSIAN APPROXIMATION 151
Using eqns (5.36) and (5.38), we get
-2
Hr) = oo) -7> pepomik -n|
= of Gopnik . vf z=
= »(5@) - one). (5.53)
The potential 0(r) denotes an effective potential between the segments of
the test polymer. The effective potential consists of a strong repulsive
part (v6(r)) of very short interaction range, and a weak attractive part
(—v exp(—r/&)/42&"r) of interaction range &. In the length-scale larger
than &, these two parts cancel precisely: indeed from eqn (5.53) it can be
shown that
Jaro =0. (5.54)
Thus there is no excluded volume interaction among the segments whose
mean separation is larger than &. This effect is called the screening of the
excluded volume interaction.
As a consequence of the screening, the distribution of the conforma-
tion of the test polymer becomes Gaussian. Indeed given the effective
excluded volume potential 0(r), it is easy to calculate the mean size of
the polymer by use of the method described in Section 2.5.'° For
example, the straightforward perturbation calculation gives (see Appen-
dix 5.111)
2 NI/A, c-cls, = bb > BA”. (5.57)
If we know how a certain physical quantity changes under this transfor-
mation, we can discuss the dependence of this quantity on those
parameters.
As an example, let us consider the apparent correlation length &,),
defined by eqn (5.41). We can take the same line of argument as given in
Section 2.6.
(i) From dimensional analysis &,, is written as
Sepp = OF(N, cb). (5.58)
(ii) Since &,,, is invariant under the transformation (5.57) we have
bAYF(NIA, cb°2*-) = F(N, cb). (5.59)
For this to be satisfied &,,, must be written ast
Eapp = N°DF(cb°N?*"?). (5.60)
This is rewritten by using the radius of gyration in infinite dilution
tHere a single symbol F is used to denote various functions for the sake of notational
simplicity.SCALING THEORY—STATICS 153
RY = N° and the segment density c* at the overlap concentration
ct = N/(RO) = N*-"/b3, (5.61)
as
Expp = ROF(c/c*) (dilute and semidilute). (5.62)
This equation is valid in both the dilute and semidilute regimes.
In the semidilute regime, the functional form can be further specified.
(iii) In the overlapped state of polymers, the correlation length &
(=&,pp in this region) should be a function of the segment density c only,
and be independent of N. (To see this, imagine that the polymers in the
highly entangled state shown in Fig. 5.1b are cut into halves. The
behaviour of the concentration fluctuation would not be changed by this
operation.) For eqn (5.60) to be independent of N, it must be written as
£=N°6(cb°N2"!)* (5.63)
with
: Vv
v+(3v—-1)x=0, ie, x= Wor (5.64)
Thus
E= Ro(S) ee VG") «e944 (semidilute). (5.65)
Next we consider the structure factor g(k). Since g(k) is dimensionless,
and changes as g(k)—>g(k)/A under the scaling transformation, it is
written as
g(k) = F(kb, N, cb*) (5.66)
with
F(kbA", NAW, cb°A3"") = A-'F (kb, N, cb). (5.67)
Thus g(k) can be written as
8(k) = cb°N**F(KN’, cb°N**-"), (5.68)
or using eqn (5.60)
8(k) = CE 2p pF (kEapp, c/c*) (dilute and semidilute). (5.69)
In the semidilute regime, g(k) must be independent of N, which gives
&(k)=c&°F(kE) (semidilute). (5.70) _
Similarly it can be shown that the osmotic pressure is given by
ky
= Se” F(ele*) (dilute and semidilute). (5.71)154 MANY CHAIN SYSTEMS
In the semidilute regime, II will again be independent of N, so that eqn
(5.71) predicts
ckgT
N
To compare these results with experiments, it is convenient to express the
concentration by p (the weight of polymer in unit volume). Equation
(5.71) is then rewritten as
T=
(cle*"Gr-) « 0% (semidilute). (5.72)
= PRE Epip*). (5.73)
M
In the dilute region, this must be expanded with respect to p
2-
n= [1+06+0(4) | (5.74)
where a2, a; are some numerical factors. Comparing eqn (5.74) with eqn
(5.1), we have
a 7 —
Aa= ype” Poe MU, (5.75)
In the semidilute region, eqn (5.72) gives
«pm, (5.76)
WBv-1)
n=l (4)
M \p*
These results are confirmed by the experiments of Noda et al.” (see Fig.
5.3).
The radius of gyration of a single chain is also derived by the scaling
argument. By dimensional analysis and the scaling argument, one can
show that the radius of gyration must be written as
R, = ROF(cic*). (5.77)
In the semidilute region, the statistical distribution of the chain becomes
Gaussian and R, must be proportional to N’. This requirement gives
Ry = RO (cic? )8~ 220-39) ce N%e-18, (5.78)
This prediction on the shrinkage of the polymer chains in semidilute
solution was confirmed by neutron scattering.”*
Though the scaling argument gives only a qualitative feature of the
parameter dependence, explicit functional form can be calculated by the
renormalization group method.’°?3* For example Ohta et al.”* gives by €SCALING THEORY—STATICS 155
2.0)
o
log (ITMipRT)
-1.6 0 1.6
log (p/p")
Fig. 5.3. Reduced osmotic pressure of poly(a methylstyrenes) in toluene at 25°C
is plotted against reduced concentration p/p*. Polymers of four molecular
weights ranging from 7.1 x 10* to 1.2 x 10° are shown. The slope of the full line is
1.32, Reproduced from ref. 22.
expansion
M am Loy _ 2m. +3)
wt (3p) +g [9x 2+
* onl; (e (1-X™)In(i + »)] (5.79)
where X is a parameter proportional to c/c* = p/p*, or is more explicitly
expressed by the second virial coefficient A,
X= AnpM. (5.80)
This prediction is in good agreement with the experimental result of
Wiltzius et al.” Explicit calculation for Eapp was done by Nakanishi and
Ohta.156 MANY CHAIN SYSTEMS
5.4 Topological interaction in polymer dynamics
5.4.1 Entanglement effect
Having discussed static properties, we now consider dynamical problems
in many chain systems. Here we have to consider another very important
type of interaction which arises from the very nature of the polymer:
polymers are one-dimensionally connected objects and cannot cross each
other. A good way of looking at this is to imagine that the polymers have
no thickness, and no attractive force, like a mathematical curve in space.
Clearly the excluded volume of such a chain is zero. However, even such
chains can interact strongly due to the topological constraints that chains
cannot cross each other.
Consider two loops initially placed separately in space as in Fig. 5.4a.
If one wants to calculate the partition functions of such systems, one has
to exclude the configurations such as shown in Fig. 5.46 and c because
such configurations are not accessible due to the topological constraints.
The net effect of this is that the effective interaction between the loops is
repulsive; the second virial coefficient A; is positive even if v = 0.78?
The topological interaction is also very important in the problems of
rubbers, in which the configurations of composite chains are severely
restricted by the topological constraints of other chains. This gives an
additional contribution to the elasticity of rubbers.°°!
For linear polymers, the topological constraints do not affect the static
problems at all since all configurations are accessible. However, the
topological interaction seriously affects the dynamical properties since it
imposes constraints on the motion of polymers. Indeed it is a crucial
factor in the dynamics of polymer solutions above the overlap
concentration.
O0 Q) <>
(
Fig. 5.4. Two loops in various topological states.
+ The topological constraints exist in the single chain problem, but the effect is generally
considered not to be serious because the properties in dilute solutions are usually
dominated by the external modes (translation and rotation of the chain as a whole) for
which the topological constraints are not important. This conjecture is supported by the
scaling theory and results of a computer simulation; both indicate that the topological
constraints affect the numerical factor only, although the conclusion is not yet definitive.TOPOLOGICAL INTERACTION IN POLYMER DYNAMICS 157
4 4
PSt-Toluene 40°C PSt-Toluene 40°C
3 8 3 20%
° 18%
°
° 14%
o 7.
ge 2 ° a 2
3 ° 9.7%
é 3 é
& °
2 1
2} s 6%
3 2
o 0 4%
“1 “1
2! 1 1 1 -2 L 1 4 _L
0.0 05 1.0 15 2.0 5.0 55 6.0 65. 7.0
@ log p (g/100m)) o) log M
Fig. 5.5. Steady state viscosity at zero shear rate of polystyrene in toluene
is plotted against (a) concentration and (b) molecular weight. Reproduced from
ref. 66.
This has been well realized in the study of mechanical properties,
which become quite conspicuous above the overlap concentration.°™*!
Firstly, the viscosity of polymer solutions increases steeply (roughly in
Proportion to p*~*) above the overlap concentration, and becomes much
larger than the solvent viscosity. An example is shown in Fig. 5.5a.
Notice that for high molecular weight polymers, even at the concentra-
tion of only 10%, the viscosity is several orders of magnitude larger than
the solvent viscosity, which is about 0.01 poise. Such high viscosity is a
result of molecular entanglement in the state shown in Fig. 5.16. It can
be easily imagined that the viscosity in such a state will depend strongly
on the chain length. Indeed experiments indicate that at constant
Fcentration Pp, the viscosity depends on the molecular weight as (see
ig. 5.5b)
n=M* (5.81)
where the exponent x is about 3.4. This behaviour is rather universal,
independent of temperature, solvent and the molecular species (as long
as the polymer is linear and flexible), which indicates that the phenomena
are governed by the general nature of polymers.158 MANY CHAIN SYSTEMS
Secondly, the polymeric liquids show conspicuous elasticity. For
example, if one stretches a polymeric liquid (say chewing gum) quickly
and holds it, one will feel a restoring force which decreases with time. If
one releases the specimen before the force vanishes completely, it shrinks
like rubber. The characteristic time for such elastic behaviour to be
observed (i.e., the longest relaxation time tax of the restoring force)
depends strongly on the molecular weight of the polymer in the form
Timex® M* with x=3.4 (5.82)
which again indicates that the relaxation of the molecular conformation is
extremely retarded by the molecular entanglement.
There is much experimental evidence which indicates the dominant
role of the topological constraints in the dynamical properties of
polymeric liquids, a comprehensive discussion on which is given in refs 30
and 31.
5.4.2 Rigorous approach
Although experimental results on the viscoelasticity of entangled poly-
mers were established quite early, theoretical explanation was not
successful for a long time. (Some of the early theories are reviewed in
ref. 31). This is due to the difficulty of handling the topological
constraints.
From the theoretical point of view, the topological constraints provide
a unique class of problems. This interaction is quite singular: for
example, in linear polymers its effect is null for static properties but quite
serious for dynamical properties. The interaction has no parameter which
characterizes the strength. Rigorous theoretical treatment of such inter-
action is quite difficult. This may be seen in a simple example.
Consider the problem of calculating the second virial coefficients A, of
the ring polymers shown in Fig. 5.4. The first step in such a calculation is
to find a mathematical expression which distinguishes the non-entangled
state (such as (a)) from the entangled state (such as (b)). How can we do
that? A way of doing it is to use topological invariants which remain the
same as long as the polymers do not cross each other. A classical quantity
found by Gauss is the integral for two loops:**
= Fe fanpam| (FE) x (=) oe (5.83)
where the integral is done over the closed contour of the loops 1 and 2.
The integral is 0 for the configuration (a) in Fig. 5.4, and 1 for (6) and
remains the same if the loops do not cross each other. (This can be
proved by using Ampére’s law in magnetism: the integral is equal to theTOPOLOGICAL INTERACTION IN POLYMER DYNAMICS 159
number of times that the (directed) loop 1 passes through loop 2.)
However, the Gaussian invariant (5.83) is not enough to specify the
topological class of loops because, for example, the configuration (c)
gives the same value of J as the configuration (a). A better specification is
achieved by the use of Alexander polynomials, which amounts to an
analysis of the order of crossing of the projections of the curves on the
surfaces, but the expression becomes too complicated to be used for
analytical calculation.
Despite the difficulty, progress has been made in the static problems
(e.g., the calculation of A, of ring polymers” and the rubber
elasticity“), by using the Gauss integral as a principal index for the
topological class. Rigorous results can be obtained for the entanglement
between a polymer on the plane and a line standing on it,**“* and the
two-dimensional problem can generally be solved using Riemann
surfaces. **
In the case of linear polymers, the entanglement effect appears only in
the dynamical properties, and it appears at first sight one does not know
where to start. Edwards” suggested that the entanglement effect can be
described by the Smoluchowski equation,
ow Qo
Soo ape Hom (ho
bm
+
ow aU
ORom ORom ¥), 6.84)
because the mobility tensor H,,., is zero for the pair of segments which
are going to cross each other. To see how this works, consider the
one-dimensional motion of two particles, 1 and 2 (see Fig. 5.6). If we
x x
Oo Oo
@ x
%
=X.
My
(b)
Fig. 5.6. (a) One-dimensional Brownian motion of two particles which cannot
cross each other, (b) phase space of the system.160 MANY CHAIN SYSTEMS
assume that the diffusion constant D, =k,TH, is a function of lx —x,|
only, the Smoluchowski equation is written as
ow a ow a
“Ot Ox, DOT, xz DOs
a wi a ow
+ 3y, D2) 5, + 5y, Pa ZT (5.85)
Using
1 1
Xa +x2), %= 7 i — 2) (5.86)
eqn ee is rewritten as
& = (00) + D&) + OO - Dewy) se (5.87)
at a
The underlined diffusion constant, which corresponds to the relative
motion of the particles, vanishes when X,=0, i.e., when x,;=x. This
guarantees that the flux of the representative point across the line x, = x2
is zero, i.e., the particles cannot cross each other. Notice that this
conclusion holds for any functional form of D (provided it is finite at
X,=0). An attractive feature of this approach is that it guarantees that
the equilibrium properties with the topological constraints are the same
as the ideal chains. However, rigorous solution of eqn (5.84) is not easy,
and approximate treatment such as the preaveraging approximation does
not give the correct molecular weight dependence for 7 and Tmax. At this
moment, this idea still remains to be explored.
5.4.3 The tube model
Though rigorous theory for the topological interaction is extremely
difficult, it has been shown that a highly entangled state can be treated by
an effective model, the tube model. This model assumes that due to the
topological constraints, the motion of the chain is essentially confined in a
tube-like region made of the surrounding polymers (see Fig. 1.6). This
model, originally proposed for the problem of rubbers,“ offers a basis for
the dynamics of chains in a network,*” which has been quite successful in
explaining many dynamical properties of entangled polymers.** We shall
give a detailed account of it in the subsequent chapters.
Though the tube model is successful, our present understanding of the
dynamics in entangled systems is still incomplete. Agreement between
theory and experiments is not yet complete as we shall discuss later.
More seriously, the tube model does not describe all aspects of the
dynamics: it describes properties which depend on a single chainDYNAMICS OF CONCENTRATION FLUCTUATIONS 161
dynamics, but cannot so far handle problems which involves the
collective motion of many chains. Progress can be made on aspects of this
problem, particularly in concentrated and semidilute solutions, as against
melts, and this we shall study in the rest of this chapter.
5.5 Dynamics of concentration fluctuations
5.5.1 Kinetic equation
In Section 5.2, it was shown that the Gaussian approximation for the
collective coordinates c, is quite useful for describing the static properties
of polymers in concentrated solutions. It is a natural temptation to
generalize this approach to dynamical problems. The central assumption
to be made in this approach is that the set of coordinates {c,} are good
variables for characterizing the state of the system, and that therefore a
closed equation can be constructed for their time evolution. This
approach is quite analogous to the critical dynamics for binary solutions
of low molecular weight,” where the dynamics of the system is described
by the phenomenological Langevin equation
Butte),
Bcy: (5.88)
a
a = 2 — Ly
where U({c,}) is the free energy, Lj, are phenomenological kinetic
coefficients and 7, are Gaussian random variables satisfying
(nm) =0, (re (t)re(t')) = 2kpTLyy-d(t - t'). (5.89)
In the polymer problem, the validity of this equation is not obvious
since the description of the polymeric system by {c,} disregards the chain
connectivity and, therefore, neglects the entanglement effect. However,
as far as the dynamics in the short time-scale is concerned, this will not be
a serious problem} since, as we shall discuss later, the topological
constraints are not important in the short time-scale dynamics, Indeed it
will be shown that the initial slope in the dynamical structure factor is
correctly described in this approach. In the long time-scale, on the other
hand, the validity of eqn (5.88) is not clear, and it may well be that the
theory has to be modified in future. Fortunately, many experiments
Telated to concentration fluctuations are concerned with the short
time-scale motion, so that it is worthwhile to pursue the idea in detail.
t Here short time means that it is shorter than the reptation time t, which will be discussed
in Chapter 6. For the phenomena which involve long time-scale, or large length-scale
motion of polymers (as happens in the problem of phase separation™” and @ regimes’), the
Present theory is not valid.162 MANY CHAIN SYSTEMS
Now for the polymer problems, U({c,}) has been calculated in Section
5.2.1. To determine L,,-, we use the Langevin equation for R,, (see eqn
(4.42).
op - 3U[Ren]
ay Ren = = Hantm (- ORom + hn)
=_ _ 2U [Rand
= Haro * Sp Tn (5.90)
where Hontm = H(Ran — Rom) and
Ton = 2 Hono * Son (5.91)
whose time correlation function is
(Fan (t)) = 0, (Tan (t)Fom(t")) = 2k e THanomd(t — t'). (5.92)
Equations (5.13) and (5.90) give
a 1y., IRan .
arch) = 7 Dik exp ik - Ran)
U[Ran]
Rom
aly ih (—Hoaom + in) exPCiK * Ran). (5.93)
V an,bm
Comparing eqn (5.93) with eqn (5.88), we have
ly. .
nt) = pm ik * Tan (t)exp(ik * Ran). (5.94)
From eqn (5.92) and (5.94), it follows that
Cralyrel@')) = Fa Be (lll) BE
X exp(ik * Roy + ik’ « Rym)) (5.95)
= 55D 2a TO — 1 RK: Hanon
an,bm
x exp(ik + Ran + ik’ + Rom):
Using
Haan = | aes a ttenptig = (Rax—Rim)) (5:96)DYNAMICS OF CONCENTRATION FLUCTUATIONS 163
we have
—26 k-k’—(k-@)(k'-
trade) = a | EE EOE
“5 oxi +q)+Ra, t+ i(k’ —q)+ Rom]
og kek’ -(k- OK)
= —25(t— ‘ykeT | Gay nea? a
(5.97)
From eqns (5.89) and (5.97) we have
Lye = —[ Ot KDR) (5.98)
(any ng? Cueeeme
Equations (5.88) and (5.98) give a nonlinear equation for c,. To proceed
further, we use the preaveraging approximation, i.e., replace Ly, by its
average in equilibrium
_dg_k-k—(k-Ak'-4
Lye (Lie eq = 2 4) (Cergle'—g)eq:
(2m)° 14
G.99)
Since (Cy+gr'~g)eq is given by (c/V)64_«(k +), (Lax')eq is written as
(Lit eq = On-0 Le (5.100)
where
_dq B-(k- ay
Ly S GY ng? g(k+q). (5.101)
Therefore the Langevin equation becomes
9 yey, Wud)
my o(t)=—-Ly 3c +7(t). (5.102)
The savivaen form of the Smoluchowski equation is
= W((c})= Lae Stal kat woth)! 5.103)
—k
In the Gaussian oo an is given by
U({cx}) = Xkp Tap ok) CRE-k (5.104)
a(k)164 MANY CHAIN SYSTEMS
Thus the Langevin equation (5.102) becomes
2 Cy = Tee, + 4 (t) (5.105)
where
Tyo kyT EE (5.106)
why "
or using eqn (5.101)
2 2.
Ty=keT og glk+ qk Gy (6.107)
ay gk) ng?
Equation (5.105) is the basic equation for the dynamics of concentration
fluctuation. We now study the consequence of this equation.
5.5.2 Dynamic light scattering
Equation (5.105) is a linear Langevin equation which has been studied in
Section 3.5. The time correlation function (c;,(t)c_,(0)) is thus easily
calculated as
(ce(t)e—4(0)) = (cxc-x)exP(—Tit). (5.108)
Hence the dynamical structure factor is
wth =F (ele) =e erp(—Te). 6.109)
In this approximation, the decay of g(k, ¢) is single exponential.
Note that I, given by eqn (5.107) is precisely the same as the exact
initial decay rate given in Chapter 4 (eqn (4.110)). Thus eqn (5.109) is
exact for t— 0. The explicit form of I, is evaluated from eqns (5.39) and
(5.107) (see eqn (4.111)*) as
__keT f 1418? (eee, kal
The 4n°n, | da pe 1+ get (z 1g | 1]
AeT Re
RF 5.110
- Gant (ks) (5.110)
where
F(x)=> 1+eY + (x? - Dtan7“(x)). (5.111)
The function is plotted in in Fig. 5.7.DYNAMICS OF CONCENTRATION FLUCTUATIONS 165
FidDecopk®
0.1 1 10
ke
Fig. 5.7. F(x) =T;/Dcopk” is plotted against x = ké.
The expression is simplified in the two limiting cases
(i) For x >> 1, F(x) =32x/8, whence,
_ ket 3
Te=T69, © KE 1. (5.112)
This is precisely the same as eqn (4.113). This must be so since in the
region of k&>>1, the dynamics is dominated by the single chain
behaviour. The condition of k&>>1 is not easily attainable by light
scattering, but the result has been confirmed by neutron scattering.”
(ii) For x «1: F(0) =1, so that
kT
Te Gane
The k dependence of I, is again the same as in dilute solutions (eqn
(4.99)). In fact this k dependence holds quite generally in the small k
Tegion, for consider the Taylor series expansion of I’, with respect to k.
In the isotropic state, the most general form is
Ty=anta,k’+a(ePPt....
Since T, must vanish at k=0, a)=0, whence the above equation is
written as
Ro kEKI. (5.113)
Ty =Dappk? (k>0). (5.114)
This defines an apparent diffusion constant D,,, for the entire concentra-
tion region.166 MANY CHAIN SYSTEMS
In dilute solutions, it was shown that D,,, agrees with the self-diffusion
constant defined by
Do = lim (Ro(0 ~ Ro(0)?) (6.115)
where Rg is the position of a test polymer. On the other hand, in
concentrated solutions, eqn (5.113) indicates that D,,, is given by’?
kT
627n,€
which is called the cooperative diffusion constant. It should be noticed
that D.oop has no relation to Dg. Indeed Dg should decrease with
concentration (because of the molecular entanglement), while Daoop
increases with the concentration. The increase Of D.oop With concentra-
tion results from the fact that the restoring force for the concentration
fluctuations is larger at higher concentration.
The concentration dependence of Dzoop is obtained from eqns (5.36)
and (5.116):
Deoop = (5.116)
Doop *c'” (Gaussian approximation). (5.1174)
On the other hand, in the semidilute regime, use of eqn (5.65) gives
Decop® 0%. (5.117)
From experiments, the exponent in D.oop*c* has been found to be
between 0.5 and 0.75°**” and this is discussed in detail by Schaefer and
Han.
It must be mentioned that although the behaviour of the dynamic light
scattering in the short time-scale is well understood, theory is still lacking
for the behaviour in the whole time-scale. Experimentally it has been
observed" that in some systems the structure factor does not decrease
in a single exponential manner and has a long tail. The long time-scale
behaviour is considered to be related to the topological interaction, but
quantitative theory is not yet given.
5.5.3 Form birefringence
As mentioned in Chapter 4, the birefringence of polymer solutions has
two origins: the intrinsic birefringence which arises from the orientation
of polymer segments and the form birefringence which comes from the
anisotropy in the correlation of the segment density. In dilute solutions,
the relative contribution of each of these terms depends both on polymer
molecules and solvent molecules. In concentrated solutions, the form
birefringence decreases with increasing concentration,” while the intrin-DYNAMICS OF CONCENTRATION FLUCTUATIONS 167
sic birefringence increases steeply like the viscosity. (Note that the
intrinsic birefringence is proportional to the stress (see eqn (4.238).) This
leads to an important relation called the stress optical law® which states
that the stress and birefringence are proportional to each other. Detailed
discussion of this relationship is given in Chapter 7. Here, as an
application of the Langevin equation (5.105), we shall calculate the form
birefringence and show that it actually decreases with the concentration.
According to eqn (4.237), the form birefringence is given by
on\? 1
n=—25 (Z) aap | teCEaky—Wenlek) 6.118)
where g,(k) is the (static) structure factor at time t,
gh) =~ (exlte-4(0)- 6.119)
To calculate g,(k), we first consider the Langevin equation under the
macroscopic velocity field
Ualt, t) = Kap(trp- (5.120)
The macroscopic velocity causes a flow of c(r, ¢),
2D) 8 -r0(4, 0). (5.121)
Ot | acite or
The Fourier transform of which gives the drift for c,(r) as
Sex) =k-K() Sali. (5.122)
drift
Adding this to eqn (5.105), we have the Langevin equation in the
presence of the velocity gradient,
é c(t) = —Tycx(t) + y(t) +k w(t) + Sal. (5.123)
The equation for g,(k) is obtained from eqn (5.123) as
a
50) = 2 (P00) + (00 *S9)]
=~ gh) + © (ree-a(O)) + (eaCOr-0(0)1
t+k-K(Q)- Ae (&). (5.124)168 MANY CHAIN SYSTEMS
The underlined terms can be calculated in the same way as in eqn
(4.141), and gives 2k,TL,. Thus, using eqn (5.106) we have
3 gk) = = 20 u(8:(k) — Seq(k)) + & + w(t) alk) (5.125)
where g.q(k) is the equilibrium structure factor. Since eqn (5.125) is a
first-order differential equation, it can be solved by a standard method,}
but here we consider the simple case of «(t) small and independent of
time.
The steady-state solution of eqn (5.125) is written as
B(K) = Beq(k) + 81(k) (5.126)
with
3 Kap ke
Bull) = 5p KH Se Bal) =e KE gah). (5-127
Substituting this into ™ (5.118), we have
9 = ~2£ Eire
p= 25 (ZY | abby ~ 1509) EM ES gah). 6.128)
The integral over the direction & can be calculated by using
x | dkk ks =450p (5.129)
& | hk hy hy ky = 48(Sap5yv + Sau. 6v + Sav pu). (5-130)
The result is written as
19, = AM Kap + Koa) (5.131)
with
any? 1
aQ= -£(2) in ) sant sas [awe 7 3 FBealk). (5.132)
t Let Fig(t, t’) be the solution of the equation
a
Bao lt 0) = Fault, )Kpu(t)
with the boundary condition F,9(¢, t) = 5,9, then the solution of eqn (5.125) is
: .
Ver { ar
a(h)=2 0 kyr f at exp| 2] aT pe a]briey
I tSCALING THEORY—DYNAMICS 169
In semidilute solutions, I, and g.q(k) are given by the following scaling
form (see 5.3 and 5.6):
T= ne Per, UKE), Beq(k) = c8°F,(KE). (5.133)
Equations (5.132) and (5.133) give
3
40 = nea aa (2 ay" (5.134)
If On/dc is regarded as independent of concentration, A{? depends on the
concentration as
AQ x 783 a eo 4 (5.135)
which decreases with increasing concentration.
On the other hand, for concentrated solutions, substitution of eqns
(5.39) and (5.110) gives
1_(8 £ (2) = poth (By
= npr) § 3c) kgTun \de (6.136)
which also decreases with concentration since 9n/dc decreases with
concentration.
Therefore in both the semidilute and concentrated solutions, the
contribution from the form birefringence decreases with the concentra-
tion. Experimental results are summarized in refs 62 and 63.
5.6 Scaling theory—dynamics
The dynamical scaling law discussed in Section 4.3 can also be used for
semidilute solutions.'~* The hypothesis is that any physical quantities
characterizing the dynamics satisfy the scaling transformation
A> HA (5.137)
when the basic parameters are changed as
N>N/A, b> ba’, cel A. (5.138)
It is worthwhile to ask whether dynamical scaling is valid in a system in
which the topological constraints are important since it is well known that
topology is extremely dimension dependent; for example, vorticity is
conserved in two dimensions, not in three dimensions; knots exist in
three dimensions but not in four dimensions, etc. Such effects do not
appear in the renormalization group calculations which are quite con-
tinuous in variation of the dimensionality of the system. On the other170 MANY CHAIN SYSTEMS
hand, one can argue that since the topological constraints introduce no
new length-scale, it would not upset the scaling arguments; rather the
scaling argument would give a common restriction on the theoretical
results for systems both with and without topological constraints (the
latter of course being a hypothetical system in which chains can pass
through each other like phantoms.) Here we shall take this viewpoint.
Various experimental results are quite consistent with the scaling
assumption in good solvents, though the validity is not yet established in
© solvents.
As an example let us consider the apparent diffusion constant D,p,
measured by dynamic light scattering (see eqn (5.114)). First we use a
dimensional analysis. The relevant parameters in the problem are b, c,
N, kgT, and n,, 80 Dapp is written as
Dapp =
, N). (5.139)
This must be invariant under the transformation (5.138), whence
; F(cb, N) ag hcorn, N/A) (5.140)
which leads to
= kT 3ay—14+3
Dapp = aN F(cb°N7**9"). (5.141)
Using the diffusion coefficient in the dilute limit D9’ ~k,T/N’b and
c* = N/(N’b)’, eqn (5.141) is written as
Dapp = DYF(cie*) (dilute and semidilute). (5.142)
This equation is valid both in dilute and semidilute solutions. Unlike with
dilute solutions, the scaling argument does not predict molecular weight
dependence explicitly. (This must be so since one can derive the same
formula for the self-diffusion constant Dg. Further information is needed
to distinguish the functional form of D,,, from that of Dg.) If one
assumes that D,p, is independent of the molecular weight, then one has
=H ( B3)"r-D ~The (5.143)
D, ‘app
which agrees with eqn (5.117b). On the other hand, Dg depends on the
molecular weight. In the next section we shall show that D « N~? by the
tube model. For this to be written as DF (c/c*), Dg must be written as
kgT
De= nN b(cleye OD *
ee N72¢7™ (5.144)SCALING THEORY—DYNAMICS 171
which decreases with concentration. The self-diffusion coefficient in dilute
and semidilute solutions has been measured by forced Rayleigh
scattering.” The result is shown in Fig. 5.8, which includes the data of
Deoop and displays the different concentration dependence of Dg and
coop:
Likewise the viscosity 7 of the solution is shown to be
n =,F(cle*). (5.145)
Again, the molecular weight dependence is not determined by the scaling
argument alone, but eqn (5.145) predicts that if 7/7, is plotted against
«x @
10%
Dg, Dogo (em? s°')
ay
SLOPE -1.75 \
y,
+My = 78300 \
©” = 245000
x = 598000 ¥
y= 754000
x
I
10° F © from ORS.
10"
10° 10? 107 10°
Weight fraction of polymer
Fig. 5.8. The apparent diffusion constant D,,, obtained by the dynamic light
scattering® and the self-diffusion constant D, obtained by the forced Rayleigh
scattering of the chains for four different molecular weights. Reproduced from
ref. 64.172 MANY CHAIN SYSTEMS
c/c* = p/p*, then the viscosity curves of various molecular weight can be
superimposed. Such superposition had been known prior to the scaling
theory.* If one uses p*[n] = constant, eqn (5.145) can be rewritten as
n =1,F(oe[n)). (5.146)
This form also has been established experimentally for a long time.®
5.7 Effective medium theory
5.7.1 Failure of the scalar field description
The collective coordinate formalism given in Section 5.5.1 is not capable
of describing the whole aspect of polymer dynamics. For example, the
quantities which are related to the orientation of the bond vector
OR,,/dn, such as stress, birefringence, and electric displacement, are not
expressed by {c,(t)}, and therefore cannot be dealt with in this
framework. This is in contrast to the critical dynamics of binary
solutions® in which the dynamics is entirely specified by an equation of
motion for the concentration fluctuation. To discuss the orientation
dynamics of polymers, other collective coordinates are needed. A
possible collective coordinate is®
=1y cos(e™ Vexpti
Cap(t) = 750 a expCik Run) (5.147)
which can describe orientation of bonds. For example, the tensor
Rana od)
2 ( on On 6.148)
is expressed as
er) (etn)
x ( N/ \dky kp | (5.149)
However, the usefulness of such coordinates has not yet been fully
worked out.
An alternative idea proposed by Edwards and Freed” is to consider
the motion of a single chain in an effective medium which includes the
effect of the other chains. The property of the effective medium is
determined self-consistently from the single chain dynamics. Though this
method fails to describe the entanglement effect appropriately, it
indicates an important aspect of the hydrodynamic interaction in the
concentrated system, which is the screening of the hydrodynamic
interaction.EFFECTIVE MEDIUM THEORY 173
5.7.2. Hydrodynamic screening
The concept of hydrodynamic screening is seen if one considers the
velocity field created by a point force in a fluid (see Fig. 5.9). If the fluid
is a pure solvent of viscosity y,, the velocity field u(r) is long range,
decreasing like 1/n,r as predicted by the Oseen formula. On the other
hand if there are polymers in the fluid, the situation is different. On a
small length-scale, the velocity profile will still be like 1/n,7 since it is not
disturbed by the polymers. However, on a large length-scale, the velocity
profile will obey the macroscopic hydrodynamics, so that the velocity
must decrease like 1/nr, where n is the macroscopic viscosity of the
solution. If 7 >> n,, the velocity field falls quickly and the effect caused by
the point force becomes very weak beyond a certain characteristic length
&,, called the hydrodynamic screening length (see Fig. 5.9). In such a
situation, the hydrodynamic interaction becomes negligible between
i c 7
r
@) ()
al
(a) pure solvent
(ce)
Fig. 5.9. (a), (b) Velocity profiles caused by a point force in a pure solvent and a
polymer solution. (c) The velocity v,(r) on the axis shown in (a), (6) is plotted
against 7.174 MANY CHAIN SYSTEMS
segments whose distance apart is larger than &,. This is called the
hydrodynamic screening. As a consequence of the hydrodynamic screen-
ing the dynamical behaviour of a single chain becomes Rouse-like as
against the Zimm-like behaviour in dilute solution.
Though the idea of hydrodynamic screening is easy to understand, the
actual formulation needs somewhat elaborate methods. In the following,
we shall explain the method using the crudest approximation. The reader
who is interested in a complete form of the formulation is advised to see
the original papers.'°7”7
5.7.3 Effective medium theory
First let us consider how the velocity profile caused by a point force is
affected when a small number of polymers are present in solution. For
simplicity we consider the steady state in the velocity field, though this
assumption is not essential.”"”” Let F.. be the point force acting at the
origin. For pure solvent, the velocity perturbance is given by
vo(r) = H(r) * Foxe (5.150)
where H(r) is the Oseen tensor given by eqn (3.106). By the Fourier
transform,
Uy = | drup(r)exp(ik +r). (5.151)
eqn (5.150) is written as
1-kk
Von Te Fox. (5.152)
Our problem is how this field is affected by the polymers. In principle the
answer is given by solving the Smoluchowski equation
ow aU
a
2 aR. [> Ham . (ssa t oR.
) - w(Ra) =0.
(5.153)
The hydrodynamic interaction is included in Myjom = H(Ran — Rom). If ¥
is obtained for a given u(r), the average velocity field is calculated as
B= w+ DD (H(r ~ Ram) * Fam) (5.154)
where
a
Fam =~ 5p— (koT In B+ U) (5.155)
and the average {...) is taken for the distribution function WV.EFFECTIVE MEDIUM THEORY 175
Now it is difficult to solve the diffusion equation (5.153) which includes
polymer-polymer interactions. So we first focus our attention on a
particular polymer, which we shall call the test polymer, assuming that
the velocity field created by the other polymers is known. Let 0,(r) be
the velocity field created by the external force and the polymers
excluding the test polymer, i.e.,
Bale) = vole) + DD (HU — Ron) Fon). (5.156)
The effect of the test polymer is solved by the Smoluchowski equation for
the single chain:
a aw au a _
2 oR. [Hom . («sR ™ an) | ~ 2 5p,” PaRan)]=0.
(5.157)
The velocity perturbation created by the test polymer is then calculated
by
50,(r) = > (H(r — Ram) * Fam): (5.158)
Since 6%,(r) depends on @,(r) linearly, it must be written using a spatial
response function =(r),
60,(r) = - 5 | dr’S(r—r')a,(r’). (5.159)
Equation (5.156) is then written as
B_(r) = vo(r) + D, 50,(r) = w(r) -i5 J dr'Z(r — r')v,(r'). (5.160)
b&b 6
Now we assume that the perturbation 5%,(r) is not large, and replace
0,(r) and @,(r’) by #(r) and v(r’) respectively:
a(n) =v) -55 | dr'2(r— r')5(r’)
= nr) -£ | dr’E(r— r')0(r’) (5.161)
where c/N is the number of polymers in unit volume.
Equation (5.161) is easily solved by the Fourier transform as
Vow.
v(k)= c
1+55(%)
= Ak): Fox176 MANY CHAIN SYSTEMS
where
(5.162)
and
(Kk) = | drz(re*"t (5.163)
The function A(k) represents the effect of the point force in the presence
of the polymers. Comparing eqn (5.162) with eqn (5.152) we may
interpret
nea(k)=n [1+ £20] (6.164)
as an effective viscosity of the polymer solution on the length-scale of
1/|k|. The macroscopic viscosity is given in the limit of k> 0 as
n=n[1+<=)|
So far we have been considering the dilute solution. This allowed us to
use the Oseen tensor H(r) in solving the single chain problem, eqn
(5.157). In the concentrated solution, however, the hydrodynamic
interactions among the segments are affected by the surrounding poly-
mers. Thus we have to replace Hyram in eqn (5.157) by the effective
interaction Ajram Which is derived from A(k) given by eqn (5.162).
(5.165)
k=0
dk . -
Ham > Rana = A (Ran ~ Ran) = | ns DLR * (Ran — Ran).
(5.166)
This gives a closed equation for the effective hydrodynamic interaction.
5.7.4 Example
As an example, we consider a solution in @ condition. For simplicity of
notation we omit the suffix a to denote the quantity of the test polymer.
Since the general solution of eqn (5.157) is involved, we use a crude
approximation: we assume that in the velocity field 0(r), the test polymer
behaves as a rigid chain, and moves with constant velocity V (the
rotational motion being neglected for simplicity.). The equations deter-
mining V are
V=> Aum? Fn + 0(Rm) (5.167)EFFECTIVE MEDIUM THEORY 177
id
an DF, =0. (5.168)
™
Equation (5.167) indicates that the fluid velocity at R,, must be V and
eqn (5.168) represents the condition that the total force acting on the
polymer is zero.
To solve the equations we use the preaveraging approximation
Fm ~ (Bln a= Bl = J as (exBCik (RyRy) og R)- (5.169)
From eqns (5.162) and (5.169) it follows that
2
Pum =| 3 qe (ex CIK + (Ra — Rm) eq. (5-170)
(ny anh 5 (k) “
Equations (5.167) and (5.168) then give
Fn = — > Snnd (Rr) (5.171)
where
San =F Dan = (EEMaE ms) (EE). a7)
Thus the perturbed flow is given by eqns (5.158) and (5.171) as
88 (7) = —B (Hr Rn) - SnD (Rn) eq
=~ fae [arr Hie P)8(0" — Rp) 8(r' — Ry) exon * (0)
= AE for fae aH (OO Rade RD oak)
(5.173)
Comparing eqn (5.173) with eqn om we get
2) =VE far HH U0 ~Rn)B(Rs))oabam 5.174)
or in Fourier transform,
3(6)= | dro Be) = 57a D (emit Rn RDaabame (5.175)
on (5.170), (5.172), and (5.175) give a nonlinear equation for
(k).
To proceed further, we use the diagonal approximation for the normal178 MANY CHAIN SYSTEMS
coordinate representation of vr
where oe = “wk, 0 Joos?) (5.176)
ins ” Py 00s ES cos( PF Vn (5.177)
In this approximation, eqn (5.172) gives
Brn = wo oos( 777 )eos(™) (5.178)
(Note that the term of p = 0 is omitted from the summation.) Thus in the
® condition, eqn (5.175) becomes
~ 11 Per) (=)
2 Lp2p pam
k OPC 37k? |n ml arg, oo w oS .
(5.179)
The sum over n and m are calculated as in eqn (4.52)
> cos(P5"Joos( PE Jexp(—ab?&? |n-m))
nm
~ fan *cos( p (Fos (= (n +) exp(—36°%? |m'l)
dm
NT b?k716
ay J , vos expt 367K? |’) => + pal CGF
(5.180)
Then
ayy = 1 1
50) - SW omnes ORR, 6.181)
Similarly, h, is calculated from eqns (5.170) and (5.177),
- 1 2
4=3D | oF
Noum J Qz) 3nde(1 + pe)
x exp(—4b7k? |n — ml)oos( >= eos ‘4 pm)
2 2(b7k2/6)
) (pxINy? + (B16)?
=! ow Om anae(1+£20)EFFECTIVE MEDIUM THEORY 179
2
b? ¢ dk 1 1 (6.182)
Equations (5.181) and (5.182) determine E(k). Since this is a nonlinear
integral equation the general solution is quite hard to obtain,” but in
special cases, the characteristic results can be discussed. Without going
into the detail, we quote the main results.
(i) In dilute solution, eqns (5.181) and (5.182) can be solved by
expanding E(k) with respect to c:
E(k) = Eq(k) +c, (k). (5.183)
The first term gives the intrinsic viscosity [q] and the second term gives
the Huggins coefficient (see eqn (5.2)). The result is
ANa(VN OY (Sas
(=< View (2? ) 18)
and
3,0 -2
bua (Se) =0.757... (5.185)
Equation (5.184) is larger than eqn (4.149) by a factor V2. Equation
(5.185) is slightly larger than the experimental value which is between 0.4
and 0.6.” Thus the numerical factors differ by about 50%, but this can be
improved by removing the rigid chain approximation used in the
calculation.
(ii) Given the effective hydrodynamic interaction, one can calculate
the relaxation time 1, for the normal mode as
— ep (1 + ac(2)” +. ) (5.186)
where t{ is the longest relaxation time in the dilute limit (c— 0) and p is
about 0.5 in © condition. Equation (5.186) explicitly shows that the p
dependence of 1, changes from the Zimm-like (1, «p~*”) to Rouse-like
(1, «p~) as the concentration increases. This result has been compared
with experiments on dynamic flow birefringence.” _
(iii) At higher concentration, over a wide region of k, H(k) is
expressed as”
l-kk
AW) = Es Ee)
(5.187)180 MANY CHAIN SYSTEMS
or in the r spacet
A(r) = exp(—r/En)/nr (5.188)
which explicitly shows the screening of the hydrodynamic interaction.
The length &,, is called the hydrodynamic screening length, and given
by”?
&y = 2/acb?. (5.189)
In this regime the steady-state viscosity is obtained’? as
7b°
n= (1 + SN) (5.190)
This again indicates that the molecular weight dependence of the
viscosity becomes Rouse-like (7 « M) for high concentration.
Appendix 5.1 Transformation to collective coordinates
The free energy U, is formally calculated from
exp(—Un({ea) ko) = | TT 8Ran TT 8[ e473 exptik- Re)
a k>o
x exp(—Us[Ran]/kaT). (5.1.1)
From the definition of the complex integral given in eqn (2.1.28), it is
shown that
= doe) lid (ey —
Tae-4n~ J (I B)ew(id(-Ave+) 612)
where @_, means @,*. Using this, eqn (5.1.1) can be rewritten as
exp(—Us({oa})/kaT)* f TT doef TT OR on
k>0 a
x enp(id @-a6e- 5 Dx explik - Re.)
k+0 k+0
x exp(—Up[Ran]/ksT). (5.1.3)
To evaluate the integral on R,,, let
or) = “5 » o-.8". (6.14)
k#t0
+ Note that eqn (5.188) is not correct for 7 >> &, since the asymptotic behaviour of H(r)
must be 1/yr, but the important region of integration in (5.181) and (5.182) is not this
region, so (5.187) is a good self-consistent solution.TRANSFORMATION TO COLLECTIVE COORDINATES 181
Then eqn (2.46) gives
N 2
[on exp| E 60Ron)~ 553 [an (Bet) | = far dr'Gee #', Ns)
" 0
where G(r, r’, N;[@]) satisfies
2
[2 -Fe-omlowr.wten=ae-r900y. 6.15)
Thus
exp(-Unl(ea) ke) = [TT deeexp|id 6-24]
k>0 +0
Np
x [ | ar | ar'G(r, ', N; top| (5.1.6)
where N,=cV/N is the total number of polymers in the system.
Equation (5.1.5) can be solved in a power series of # as
G(r’, Ns) = Golr= 2's) + [an
0
N n;
x | dr, Gor — 43 N —2,)9(4)Go(r, — r'31) + | dn, | dn,
0 0
x [an f ars Gale— nsN — m1) O(R)Gilr~ 05m — m2) H(IGOLOa— sm)
fees 6.17)
Using the relation
[arcice -r;N)=1
we have
N n
far arco, rN: [oD = v+N{ dro(e) + [anf dng ar,
0 0
x | drab (n)o()G(r, 95m —M). 6.18)
Using eqn (5.1.4) and the definition of go(k) (see eqn (2.70)), this is
written as
farfaroe.r. nstop=v[i-3 Zbeb-s8k)] 6.19)182 MANY CHAIN SYSTEMS
Thus
exp(—Ual(ex))lka)= | TT aoe exp] id, #-1ce
+N, tog(1 - x = $480k). (5.1.10)
Assuming that the fluctuation in @, is small, we expand the logarithm and
evaluate the integral over @,:
ths = | Tam ex0[ iS, eee - ey $1$-180(0))
~ | Thao exo] d (ities +iguct ~£ ox6-s80(h))|
|- Da (5.1.11)
which gives eqn (5.31).
Appendix 5.11 Osmotic pressure in concentrated solution
Using the formula in Appendix 2.1, we have
{z Tl dc, exp(- Vv = Eat)
exp(—(A — Ao)/kaT) = exo] —F ve? | Hdexeno(-V 2, = G5)
= exp[-F ve] TT @tayent)
(&)
=exp| —7 ue? + 5 Hog & (5.11.1)
He [-7 2 aot
voaiv kag?
(A~AbllkeT = 700 +5505 | dk og(<+F-), (5.11.2)
The osmotic pressure is thus given by
=£ _ (A= Ao)
WksT = 7, av
T,cV constant
Va 24 &? Vv 3g?
+ ; say | a[toe(* Pa =) +R? ov
c uv
“N 2°
< 5¢ eH 3G a | seo [loe(1 + gp) - FFI: (5.11.3)PERTURBATION CALCULATION OF (R?) 183
The underlined integral is performed by the integration by parts:
[onr[re(t+ 7 2) nel fe ws
x [roe( + 3) - sel -%. (14)
l+x
Hence
T=ksT [E+5e - Tap (5.11.5)
Appendix 5.111 Perturbation calculation of (R7)
Let (...)o be the average for the ideal chain and U, be defined by
He aa 5(R,— Rn) (5.11.1)
then the perturbation method described in Appendix 2.III gives
(R?) = (Rw —Ro))of1 + (D,)o] — (Ry —Roy’G,)o. (5.11.2)
The first term gives
N m
(Ry ~ Re) + (O;)o] =O 1+ | dm ano, ~Rn))0)
8 (5.11.3)
and the second term becomes
(By -Ro)?0:)o= j aim | dn, ~ Ry (V ~m)b2-+ (Ry — Ry?
° nb?})o.
(5.11.4)
Hence
(R°) = Nb?+ [ dm dn(0(R, — Rn )[ Gm — 28? — (Rn — Ra'Do
N m
= Nb? + | dm [ dn { dri(r)((m —n)b?—P)Ge(r, m—n) (5.1115)
where
Golr, 2) = (3/2nnb2)22 exp(- (5.11.6)
an)184 MANY CHAIN SYSTEMS
Since the integrand of eqn (5.III.5) depends only m —n, this is rewritten
as
N
(R®) = Nb? + [ dn(N—n) | dri(r)(nb?— F)Gy(r, n). (5.111.7)
0
By use of the Fourier transform
Gor, n)(nb? — P= lan yo
and eqn (5.53), eqn (5.III.7) is written as
dk b*h? , OR \ a,
aL exp(- Z nje (5.1118)
(ee
(5.11.9)
Since the underlined part decreases quickly with n, the integral over 7 is
evaluated for large N as
an
(e)=Nnot+u a]
N °°
fanqv-n)...=Nfan.... (5.11.10)
0 0
Hence, we finally get
2 2 dk 267k? / 6 \>
(Rt) = Nb [ela > (es) aE
ool
= Nor[1+ (5.11.11)
References
1. de Gennes, P. G., Scaling Concepts in Polymer Physics, Chapter 3. Cornell
Univ. Press, Ithaca, New York (1979).
2. Flory, P., J. Chem. Phys. 10, 51 (1942); see also Principles of Polymer
Chemistry, Chapter 12. Cornell Univ. Press, Ithaca, New York (1953).
. Huggins, M. L., J. Phys. Chem. 46, 151 (1942); J. Am. Chem. Soc. 64, 1712
(1942).
. A review on polymer collapse is given by Williams, C., Brochard, F., and
Frisch, H. L., Ann. Rev. Phys. Chem. 32, 433 (1981).
. Daoud, M., and Jannink G., J. Phys. (Paris) 37, 973 (1976).
fen” D. W., Joanny J. F., and Pincus, P., Macromolecules 13, 1280
1980).
. Moore, M. A., J. Phys. (Paris) 38, 265 (1977).
XN AW F w10.
11.
13.
BS RRSB RB
ss
38,
41.
42.
43.
REFERENCES 185
. Schafer, L., Macromolecules 47, 1357 (1984).
. Edwards, S. F., Proc. Phys. Soc. London 88, 265 (1966).
Edwards, S. F., and Muthukumar, M., Macromolecules 17, 586 (1984).
Leibier, L., Macromolecules 13, 1602 (1980).
. Okano, K., Wada, E., Kurita, K., Hiramatsu, H., and Fukuro, H., J.
Physique Lett. 40, L171 (1979).
Okano, K., Wada, E., Kurita, K., and Fukuro, H., J. Appl. Cryst. 11, 507
(1978).
. Muthukumar, M., and Edwards, S. F., J. Chem. Phys. 76, 2720 (1982).
. Des Cloizeaux J., and Jannink G., Les Polyméres en Solution: Modelisation
et leur structures, Chapter 13, §2.5.5. Edition de Physique, Paris (1986 or 7).
. Edwards, S. F., J. Phys. A8, 1670 (1975); Edwards, S. F., and Jeffers, E. F.,
J. C. S. Faraday Trans. 11 75, 1020 (1979).
. Flory, P., J. Chem. Phys. 17, 303 (1949); see also Principles of Polymer
Chemistry, Chapter 12. Cornell Univ. Press, Ithaca, New York (1953).
. Richards, R. W., Maconnachie, A., and Allen, G., Polymer 19, 266 (1978).
. Oono, Y., Adv. Chem. Phys. 61, 301 (1985).
Freed, K. F., Accounts Chem. Res., to be published.
. Daoud, M., Cotton, J. P., Farnoux, B., Jannink, G., Sarma, G., Benoit, H.,
Duplessix, R., Picot, C., and de Gennes, P. G., Macromolecules 8, 804
(1975).
Noda, I., Kato, N., Kitano, T., and Ngasawa, M., Macromolecules 14, 668
(1981).
des Cloizeaux J., and Noda, I., Macromolecules 15, 1505 (1982).
Schafer, L., Macromolecules 15, 652 (1982).
Ohta, T., and Oono, Y., Phys. Lett. 89A, 460 (1982).
. Wiltzius, P., Haller, H. R., Cannell, D. S., and Schaefer, D. W., Phys. Rev.
Lett, 51, 1183 (1983).
Nakanishi, A., and Ohta., T., J. Phys. A18, 127 (1985).
- Vologodskii, A., Lukashin, A., Frank-Kamenetskii, M. F., and Anshelevich,
V., Sou. Phys. JETP 39, 1059 (1975); 40, 932 (1975).
Roovers, J., and Toporowski, P. M., Macromolecules 16, 843 (1983).
(1980) J. D., Viscoelastic Properties of Polymers (3rd edn). Wiley, New York
1980).
. Graessley, W. W., Adv. Polym. Sci. 16, 1 (1974).
Baumgartner, A., Polymer 22, 1308 (1981).
. Edwards, S. F., J. Phys. Al, 15 (1968).
}. Alexander, J. W., Trans. Am. Math. Soc. 30, 275 (1928).
. Ball, R., and Mehta, M. L., J. Physique. 42, 1193 (1981).
. Brereton, M. G., and Shah, S., J. Phys. A14, L51 (1981).
. des Cloizeaux, J., J. Phys. (Paris) 42, LA33 (1981).
Iwata, K., and Kimura, T., J. Chem. Phys. 74, 2039 (1981); Iwata, K., J.
Chem. Phys. 78, 2778 (1983); Macromolecules 18, 115 (1985).
. Tanaka, F., Prog. Theor. Phys. 68, 148, 164 (1982).
. Edwards, S. F., Proc. Phys. Soc. 92, 9 (1967).
Iwata, K., J. Chem. Phys. 76, 6363, 6375 (1982).
Edwards, S. F., Proc. Phys. Soc. 91, 513 (1967).
Prager, S., and Frisch, H. L., J. Chem. Phys. 46, 1475 (1967).
. Saito, N., and Chen, Y., J. Chem. Phys. 59, 3701 (1973).186
45.
47.
49.
SASH BAS s
a)
70.
Th.
MANY CHAIN SYSTEMS
Ito, K., and McKean, H. P., Diffusion Processes and their Simple Paths.
Springer, Berlin (1965).
. Edwards, S. F., Proc. R. Soc. Lond. A385, 267 (1982).
de Gennes, P. G., J. Chem. Phys. 55, 572 (1971).
- Doi, M., and Edwards, S. F., J. C. $. Faraday Trans. II 74, pp. 1789, 1802,
and 1818 (1978).
Kawasaki, K., Annals Phys. 61, 1 (1970); Kawasaki, K., and Gunton, J. D.,
Critical dynamics, in Progress in Liquid Physics (ed. C. Croxton). Wiley,
New York (1978).
. Pincus, P., J. Chem. Phys. 75, 1996 (1981).
. Brochard, F., and de Gennes, P. G., Macromolecules 10, 1157 (1977);
Brochard, F., J. Physique 44, 39 (1983).
. Richter, D., Hayter, J. B., Mezei, F., and Ewen, B., Phys. Rev. Lett. 41,
1484 (1978).
. de Gennes, P. G., Macromolecules 9, 587, 594 (1976).
. King, T. A., Knox, A., and McAdam, J. D. G., Polymer 14, 293 (1973).
. Ford, N. C., Karasz, F. E., and Owen, J. E. M., Discuss. Faraday Soc. 49,
228 (1970).
. Adam, M., and Delsanti, M., Macromolecules 10, 1229 (1977).
. Nose, T., and Chu, B., Macromolecules 12, pp. 590, 599, and 1122 (1979);
13, 122 (1980); Nemoto, N., Makita, Y., Tsunashima, Y., and Kurata, M.,
Macromolecules 17, 2629 (1984).
. Schaefer, D. W., and Han, C. C., in Dynamic Light Scattering (ed. R.
Pecora). Plenum Press, New York (1985).
Amis, E. J., and Han, C. C., Polymer 23, 1403 (1982); see also Amis, E. J.,
Janmey, P. A., Ferry, J. D., and Yu, H., Macromolecules 16, 441 (1983).
Eisele, M., and Burchard, W., Macromolecules, 17, 1636 (1984).
Brown, W., Macromolecules 17, 66 (1984).
Tsvetkov, V. N., in Newer Methods of Polymer Characterization (ed. B. Ke).
Interscience, New York (1964).
. Janeschitz-Kriegl, H., Adv. Polym. Sci. 6, 170 (1969).
Léger, L., Hervet, H., and Rondelez, F., Macromolecules 14, 1732 (1981).
. Debye, P., J. Chem. Phys. 14, 636 (1946); see also ref. 75.
. Onogi, S., Kimura, $., Kato, T., Masuda, T., and Miyanaga, N., J. Polym.
Sci. C 381 (1966).
Onogi, S., Masuda, T., Miyanaga, N., and Kimura, S., J. Polym. Sci. A2 5,
899 (1967).
. See for example, Dreval, V. E., Malkin, A. Y., and Botvinnik, G. O., J.
Polym. Sci. Phys. ed. 11, 1055 (1973).
. Edwards, S. F., Faraday Symp. Chem. Soc. 18, 145 (1983).
Edwards, S. F., and Freed, K. F., J. Chem. Phys. 61, 1189 (1974), Freed, K.
F., and Edwards, S. F., J. Chem. Phys. 61, 3626 (1974); ibid 62, 4032 (1975).
Freed, K. F., Polymer dynamics and the hydrodynamics of polymer
solutions, in Progress in Liquid Physics (ed C. Croxton). Wiley, New York
(1978).
. Muthukumar, M., and Freed, K. F., Macromolecules 10, 899 (1977); 11, 843
(1978).
. Perico, A., and Freed, K. F., J. Chem. Phys. 81, 1466, 1475 (1984).REFERENCES 187
74. A discussion for the solution is given by Yoshimura, T., J. Math. Phys. 24,
2056 (1983).
75. Frisch, H. L., and Simha, R., The viscosity of colloidal suspensions and
macromolecular solutions, in Rheology (ed. F. R. Eirich) Vol. 1, Chapter 14.
Academic Press, New York (1956); see also Bohdanecky, M., and Kovar, J.
Viscosity of Polymer Solutions, Chapter 3. Elsevier, Amsterdam (1982).
76. Martel, C. J. T., Lodge, T. P., Dibbs, M. G., Stokich, T. M., Sammler, R.
L., Carriere C. J., and Schrag, J. L., Faraday Symp. Chem. Soc. 18, 173
(1983).6
DYNAMICS OF A POLYMER
IN A FIXED NETWORK
6.1 Tube model
6.1.1 Tube model in crosslinked systems
As mentioned in Section 5.4.3, the highly entangled state of polymers can
be effectively described by the tube model. The idea of the tube model
originated in studying the problem of rubber elasticity."? A rubber is a
huge molecular network which is formed when a polymeric liquid is
crosslinked by chemical bonds.? An important problem in the theory of
rubber elasticity is to calculate the entropy, which is essentially the
number of allowed conformations of the chains constituting the rubber.
The topological constraints play an important role in such a problem.
Consider a lightly crosslinked rubber which consists of long strands of
polymers between crosslinks. A strand in such a rubber is schematically
shown in Fig. 6.1. In Fig. 6.16 the strand is placed on a plane and the
cross-sections of other strands are shown by dots. Due to the topological
constraints, the strand cannot cross the dots, so that the number of
conformations allowed for the strand is much less than that in free space.
How can we estimate it?
Suppose for a moment that the other chains are frozen, then the dots
can be regarded as fixed obstacles. One can see that the allowed
conformation of the strand is almost confined in a tube-like region shown
by the dotted lines: the conformations which go outside the tube are
likely to violate the topological constraints. The axis of the tube can be
defined as the shortest path connecting the two ends of the strand with
the same topology as the strand itself relative to the obstacles. Such a
path represents a group of conformations which are accessible to each
other without violating the topological constraints imposed by the other
chains, and is called the primitive path.” If the topological constraints are
replaced by the tube, the number w of the allowed conformations can be
calculated easily by the method described in Chapter 2 (see Section 6.4).
In real rubbers, the situation is more complicated since the other
strands are mobile. However, even in such a case, a self-consistent
picture will be that the range in which each part of the strand can move
around will remain finite. The range is perhaps larger than the mean
separation between the frozen strands discussed above. What diameter
one should assign to the tube is a question which has not been answered
with absolute certainty. However, as long as the strand is long enough,
the diameter is determined by local conditions, and will be independentTUBE MODEL 189
©
Fig. 6.1. (a) A strand in a rubber. A and B denote the crosslinks. (b) Schematic
picture of (a): the strand under consideration is placed on a plane and the other
strands intersecting the plane are shown by dots. (c) The tube model.
of the length of the strand. Detailed discussions'** on rubber elasticity
based on this idea and its appraisals”® are given in the literature. An
important point here is the proposition that the tube concept will be a
self-consistent picture in the system of topologically interacting system.
6.1.2 Tube model in uncrosslinked systems
The idea of the tube is intuitively appealing, and one can imagine that the
same picture will be useful for uncrosslinked system such as polymer
melts. However, one has to face a new problem that, in melts, the tube
itself changes with time because all conformations in a melt are
accessible. A key concept to solve this problem was introduced by de
Gennes’ who discussed the Brownian motion of an unattached chain190 DYNAMICS OF A POLYMER IN A FIXED NETWORK
primitive chain
{b)
Fig. 6.2. (a) A chain in a fixed network of obstacles (denoted by dots). (b) The
defects and the primitive chain (dashed line).
moving through a fixed network (see Fig. 6.2). His idea was that in the
situation shown in Fig. 6.2, the motion of the chain is almost confined in
a tube-like region denoted by dotted lines in Fig. 6.2b. Since the chain is
rather longer than the tube, the slack will constitute a series of ‘defects’
which can flow up and down the tube (Fig. 6.2b). De Gennes visualized
this as a gas of non-interacting defects running along like the arch in a
caterpillar (see Fig. 6.3a). As a result of such motion, the tube itself
changes with time (Fig. 6.3b): for example if the chain moves right, the
part B)B can choose a random direction, and create a new part of the
tube which will be a constraint for the rest of the chain, while the part of
the previous tube AgA becomes empty and disappears. This type of
motion was called reptation by de Gennes after the Latin reptare, to
creep.
Our current understanding of the dynamics of the highly entangled
state is based on the concept of reptation. This picture is rigorously
correct for the system that has been presented, i.e., a single chain in aREPTATION 191
SN
—_N\___
NL
f@
()
Fig. 6.3. (a) Motion of a single defect (b) Motion of the tube.
fixed network. Whether the picture does indeed hold for concentrated
polymer solutions or melts still remains a matter for debate, but many
experimental results suggest that reptation is the dominant mechanism
for the dynamics of a chain in the highly entangled state. Leaving the
detailed discussion of this problem to the next chapter, we shall first
consider the simple situation of a chain moving in a fixed network.
6.2 Reptation
6.2.1 Primitive chain
Let us consider a polymer moving in a fixed network of obstacles. For the
convenience of later discussion, we shall specify the problem in slightly192 DYNAMICS OF A POLYMER IN A FIXED NETWORK
different terms from those used by de Gennes. We assume that the
intrinsic properties of the polymer are represented by the Rouse model
consisting of N segments with bond length b and friction constant €. The
obstacles are assumed to be thin lines, so they have no effect on static
properties, but have a serious effect on dynamical properties by imposing
topological constraints.
The characteristic feature of the dynamics can be visualized by the tube
model. For a given conformation of the polymer, we can draw the
primitive path, i.e., the shortest path connecting the two ends of the
chain with the same topology as the chain itself relative to the obstacles
(see the dashed line in Fig. 6.2b). In the short time-scale the motion of
the polymer is regarded as wriggling around the primitive path. On a
longer time-scale, the conformation of the primitive path changes as the
polymer moves, creating and destroying the ends of the primitive path.
Even though such a picture is clear, the mathematical treatment of the
problem is still complicated since the time evolution of the primitive path
is governed by the wriggling motion of the polymer, and the wriggling
motion itself is limited by the primitive path. However, if we are
interested in the large-scale motion of long chains, we may disregard the
small-scale fluctuations, and discuss only the time evolution of the
primitive path. Since the primitive path at any moment represents the
conformation of the chain with the small-scale fluctuations omitted, we
shall use the term ‘primitive chain’ to denote the dynamical equivalent of
the primitive path. At this level of description, the details of the wriggling
motion are irrelevant, and we can start with a simpler model.
To denote a point on the primitive chain, we use the contour length s
measured from the chain end and call this the primitive chain segment s.
If R(s, £) is its position at time ¢, the vector
u(s, t)= ZR, t) (6.1)
is the unit vector tangent to the primitive chain.
The dynamics of the primitive chain is characterized by the following
assumptions.
(i) The primitive chain has constant contour length L.
(ii) The primitive chain can move back and forth only along itself with
a certain diffusion constant D..
(iii) The correlation of the tangent vectors u(s, t) and u(s’,t) de-
creases quickly with |s — s’|.
The first assumption corresponds to neglecting the fluctuations of the
contour length. The second states that the motion of the primitive chain
is reptation. The third guarantees that the conformation of the primitiveREPTATION 193
chain becomes Gaussian} on a large length-scale. This assumption
introduces a new parameter into the problem. Since the mean square
distance between two points on the Gaussian chain is proportional to
|s —s'|, it is written as
((R(s, 1) — R(s', DY) =als—s'| for |s—s’|>>a. (6.2)
The length a is called the step length of the primitive chain.
The primitive chain is thus characterized by three parameters L, D,
and a, which must be expressed by the Rouse model parameters N, b, €
and the parameters characterizing the network. The parameter D, can be
identified as the diffusion coefficient of the Rouse model
ke
NE
because the motion of the primitive chain corresponds to the overall
translation of the Rouse chain along the tube. The length L is expressed
by a since the mean square end-to-end vector of the primitive chain,
which is La according to eqn (6.2), must be the same as that of the Rouse
chain Nb?. Thus
D,= (6.3)
Nb?
=. (6.4)
We are left with a single parameter a, which depends on the statistical
nature of the network. Though precise calculation of this parameter is
difficult, it is obvious that @ is of the order of the mesh size of the
network and much less than L. This knowledge is enough for the purpose
of the present discussion.
6.2.2 Simple application
We now study the dynamics of the primitive chain and show that certain
time correlation functions can be calculated by a straightforward
method.’ For example, consider the time correlation function of the
end-to-end vector P(t) = R(L, t)— R(0, t). Figure 6.4 explains the prin-
ciple of calculating this correlation function. At t = 0, the chain is trapped
in a certain tube. As time passes, the primitive chain reptates and at a
certain later time (Fig. 6.4d), the part of the chain CD remains in the
original tube while the parts AC and DB are in a new tube. To calculate
t It must be remembered that despite the Gaussian behaviour with a large length-scale, the
Primitive chain cannot be modelled bya continuous Gaussian chain since the contour length
of the continuous Gaussian chain is infinite and has no physical significance, while the
contour length of the primitive chain has a definite physical significance and appears in
various dynamical results.194 DYNAMICS OF A POLYMER IN A FIXED NETWORK
@)
a) OY
© <
Fig. 6.4. Four successive situations of a reptating chain. (a) The initial conforma-
tion of the primitive chain and the tube which we call the original tube. (b) and
(c) As the chain moves right or left, some parts of the chain leave the original
tube, The parts of the original tube which have become empty of the chain
disappear (dotted line). (d) The conformation at a later time t. The tube segment
vanishes when it is reached by either of the chain ends: €.g., the tube segment P
and Q vanish at the instance (6) when &(t) =sp and at (c) when &(t)=sg—L,
Tespectively.
(P(t) P(0)), we express P(t) and P(0) as
P(0) =A,C + CD + DB, (6.5)
P(t)=AC+CD +DB (6.6)
Since the vectors AC and DB are uncorrelated with P(0), (P(t) - P(0))
will have the form
(PQ) PO) = (CD*) = a(o(t)) (6.7)
where o(t) is the contour length of CD, i.e., the part in the original tube.
To calculate (o(t)) we focus attention on a certain segment s of the
original tube. This tube segment disappears when it is reached by either
end of the primitive chain. Let p(s, ¢) be the probability that this tubeREPTATION 195
segment remains at time ¢. The average (o(t)) is calculated as
L
(o(9) = fasy6s, 0. (68)
0
Let ¥(E,t;s) be the probability that the primitive chain moves the
distance & while its ends have not reached the segment s of the original
tube. The probability satisfies the one-dimensional diffusion equation
ow ay
a7 Diag (6.9)
with the initial condition
W(E, 0; 5) = 6(&). (6.10)
When §=s, the tube segment s is reached by the end of the primitive
chain and W(&, t;s) vanishes (see Fig. 6.4). Similarly when § =5 — L, the
tube segment is reached by the other end and W(E, t;s) vanishes.} Thus
WE, ts)=0 at E=s and E=s-L. (6.11)
The solution of eqn (6.9) with these boundary conditions is
WE, t5)= > 2 sin( PE 7) sin(2*™S— 5) as L PHS 9) exp(—pitit,) (6.12)
where
tg = L?/D,n. (6.13)
For the tube segment s to remain, & can be anywhere between s — L and
5, so that
5
v,0= | awe, 45)
sok
=> S sin(P *)exp(—p't!t4). (6.14)
podd PIE
Thus from eqns (6.7), (6.8), and (6.14)
(P(t): P(O)) = Lay(t) = Nb?y(0) (6.15)
where
1f 8 ,
vOrT | dsy(s, t)= ae pe OPP tI). (6.16)
+ Strictly speaking this argument is valid in the limit of a—>0. If a is finite, the boundary
condition is not written in a simple form, but the correction is of the order of a/L.196 DYNAMICS OF A POLYMER IN A FIXED NETWORK
The longest relaxation time of (P(#) + P(0)) is given by ty. This is called
the reptation or disengagement time, since it is the time needed for the
primitive chain to disengage from the tube it was confined to at t= 0.
Equation (6.15) can be compared with eqn (4.35) for the Rouse chain
without constraints:
(P(t) - P(0)) =e? Soa exp(—p"t/T,r) (6.17)
where Tz is the Rouse relaxation time,
TR= ce (6.18)
On the other hand, eqn (6.13) is rewritten by eqns (6.3) and (6.4) as
1 EN*o" 6.19)
w=sa lS:
4° kpTa’
Note that t, is proportional to N® and becomes much larger than tp for
large N. This demonstrates the crucial effect of topological constraints on
the conformational change of polymers.
Let us define the number of steps in a primitive chain by
Z=7=7. (6.20)
Then the ratio between t, and tp is written as
Tal Tr = 3Z. (6.21)
Equation (6.19) has been confirmed by computer simulation.”
The function y(s, t) will appear frequently in the subsequent discus-
sions. This function has been defined as the probability that the original
tube segment s remains at time ¢. As will be shown in Section 6.3.3,
(s, t) also represents the probability that the primitive chain segment s
is in the original tube at time ¢. (Note the distinction between the tube
segment and the primitive chain segment; the former is fixed in space,
while the latter moves with the primitive chain.)
The behaviour of p(s, t) is shown in Fig. 6.5. The tube segments in the
middle (s = L/2) have long lifetimes of order tz, while the tube segments
near the chain ends have very short lifetimes: the end segment is almost
instantaneously replaced. This fact will be used in the subsequent
discussions. The function w(t) represents the average fraction of the
original tube that remains at time 4. This function is also equal to the
average fraction of the primitive chain contour that remains in the
original tube.REPTATION DYNAMICS 197
0.01
w(st)
ae
0 tL
s
Fig. 6.5. The probability y(s, t) that the tube segment s is remaining at time t.
This is also equal to the probability that the primitive chain segment s remains in
the original tube at time ¢.
6.3 Reptation dynamics
6.3.1 Stochastic equation for reptation dynamics
Although the above probabilistic description is quite useful in under-
standing the essence of reptation dynamics, it becomes progressively
more difficult to proceed with the calculation for other types of time
correlation function. For example, it is not easy to calculate the mean
square displacement of a primitive chain segment ((R(s, t) — R(s, 0))?)
by this method. In this section we shall describe a convenient method’!
for calculating general time correlation functions.
First we derive a simple mathematical equation for reptation dynamics.
Let A&(t) be the distance that the primitive chain moves in a time
interval between ¢ and ¢ + At, then
R(s,t+ At) =R(s + A&(2), 0), (6.22)
which states simply that if the primitive chain moves the distance A&(#)
along itself, the segment s comes to the point where the segment s + AE
was at time ¢ (see Fig. 6.6). The variable A&(¢) takes random values, the198 DYNAMICS OF A POLYMER IN A FIXED NETWORK
[Ree
“qa
Ag(t)
Rieteaty—/
Fig. 6.6. When the primitive chain moves a distance A&(#) along itself, the
segment s comes to the point where the segment s + A was at time t.
distribution of which is Gaussian characterized by the moments
(A&())=0, — (AE(t)?) = 2D. At, (6.23)
ie.,
2
W(AE) = (42D. At)"exp ( - i) . (6.24)
Equation (6.22) is not correct for all s: if s + A&(t) is not between 0
and L, R(s, t+ At) should be on the newly created part of the tube (see
Fig. 6.6). Writing down this condition in a formal mathematical equation
is cumbersome. However, such an expression is not needed in practice
since the condition can usually be accounted for by the fact that the
distribution of the tangent vectors at the chain ends u(0, ¢) and u(L, t)
are independent of the previous conformation of the primitive chain since
their correlation time is infinitesimal.}
6.3.2 Segmental motion
To see how eqn (6.22) works, we calculate the mean square displacement
of a primitive chain segment s((R(s, t) — R(s, 0))*).
In this case it is more convenient to calculate the following time
correlation function,
$(s, 8°58) = (ROS, 1) - RG’, OY). (6.25)
The time evolution equation for (s, 5’; ¢) is written as
o(s, s'; t+ At) = ([R(s + AE(1), 2) -— R(s', OP)
= (p(s + A(t), 55 1)). (6.26)
The bracket in the last expression represents the average over A&(t). To
tit must be remembered, however, that the dynamics of the chain end is extremely
important in reptation dynamics since eqn (6.22) only describes the one-dimensional
motion, and the three-dimensional property of the primitive chain is entirely governed by
the dynamics of the chain ends.REPTATION DYNAMICS 199
calculate this average, we expand the right-hand side of eqn (6.26) and
use eqn (6.23):
hy = a AR ' )
(o(s + A&,5'32)) =((1 +ags+S SF Sa)ol,s a)
= (1408) 2 + FOZ )965, 59
= (140.4 3) 9(6,5°s d. (6.27)
From eqns (6.26) and (6.27) we have
3 #
3p POS 85) = De za O(S, 8°31). (6.28)
To solve this, we need the initial condition and the boundary condition:
(i) The initial condition is given by eqn (6.2),
$(s, 53 Dleo= [3 —s’] a. (6.29)
(ii) To obtain the boundary condition, we write 3/ds at s = L as
Zs] =2uL,0- (RE) -RO', 0)
as enk
=2(u(L, t) - (R(L, t) — R(s", t)))
+ 2(u(L, t) + (R(s", t)— R(s', 0))) (6.30)
where s” is an arbitrary value betwen 0 and L. Since the correlation time
of u(L, t) is infinitesimal, the underlined part gives
(u(L, t) - (R(s", t) — R(s', 0))) = (u(Z, 8) - ((RG", 1) — RG’, 0))) = 0.
(6.31)
The first term in eqn (6.30) gives
2(u(L, (RU, )- Rs", D)) = 2 (RG, - RE")
s=
a Ww
= 3,408 —s") on a. (6.32)
Hence
8 una =
3s o(s,s';t)=a at s=L. (6.33)200 DYNAMICS OF A POLYMER IN A FIXED NETWORK
Similarly a
35 p(s, s';Q=-a at s=0. (6.34)
The solution of eqn (6.28) with the boundary conditions (6.33) and
(6.34), and the initial condition (6.29) is
$(s, 53 t) =|s—s'|a +25 Dit Se 4a
p=1P
(pms
x cos( 2H PT cos PA) — exp(—ip*/1.) (6.35)
Hence
$66.83 yazne+ S seo EE 4) —exp(—w2l2)}. (636)
The expression becomes simple in two extreme cases:
(i) For t«t,, $(s,5:t) is dominated by the terms with large p.
Replacing cos*(pxs/L) by the average 1/2, and converting the sum into
the integral, we have
1. 4La
(8,830) [dp S24 exp(-w'lea)
0
= 2a(D,t/n)", (6.37)
This result is easily derived, for if t 7,, the first term in eqn (6.36) dominates, so that
G(s, 53t) = 2D,t/Z. (6.39)
Hence the diffusion constant Dg of the centre of mass is given as
=n POO) _ keTa?
De lim a D,/3Z = aN7EbE" (6.40)
Thus the self-diffusion constant is proportional to N~*, which has also
been confirmed by computer experiment.”REPTATION DYNAMICS 201
6.3.3 Correlation function of the tangent vectors
Next let us consider the time correlation function of the tangent vector
u(s, t):
G(s, s'; t) = (u(s, t) + u(s’, 0)) (6.41)
which can be related to $(s, 5‘; £) by eqns (6.1) and (6.25):
cone A a
1 #
= 2a ((R(s, t) ~ R(s", 0))?)
1
= 355 3s" p(s, s';t). (6.42)
Substituting eqn (6.35), one gets
G(s, st) =ad(s —s')- 2 Pisin PE sin PE) ~ exp(—tp’/tz)]
— 3 24., (pas (pms 2
= L = sin (22 Pe sin( P= “)exp(- (p"!t4). (6.43)
An important conclusion is derived from eqn (6.43) by a geometrical
interpretation of (u(s, t)- u(s’, 0)). First note that if u(s) and u(s') are
the tangent vectors of a Gaussian chain of step length a, the integral
4
[=> | ds(u(s)-u(s’)) — (61>52) (6.44)
is equal to 1 when s’ is between s, and s2. Now consider
L
V6", =3 | ds(a(s, 1) mls’, 0). (6.48)
0
This is equal to unity at time =0. As time passes, the original tube
segment s’ becomes empty of the primitive chain. If this happens u(s', 0)
becomes totally uncorrelated to u(s, t) and the contribution from such
case to p(s’, t) is zero. Therefore ~(s’, t) represents the probability that
the original tube segment s’ is remaining at time ¢. Indeed from eqns
(6.43) and (6.45), it follows that
vO", = SS sin( PE esp(—p'ur,) (6.46)
which agrees with eqn 6.14),202 DYNAMICS OF A POLYMER IN A FIXED NETWORK
Now, one can show by a similar argument that the integral
L
~ 1
HG, )=7 [ ds"(u(s,£)-(s', 0)) (6.47)
0
Tepresents the probability that the primitive chain segment s is in the
original tube. From eqns (6.43), (6.45), and (6.47) it follows that
HG, )= Ys.) (6.48)
i.e., the probability that the original tube segment s remains is equal to
the probability that the primitive chain segment s is in the original tube
that remains.
6.3.4 Dynamic structure factor
As the final example, we shall consider the dynamic structure factor of a
single chain:
L L
ab =X fas dor(explik-(R(.)-RG',091). 649)
To calculate this quantity, we again consider $(s, s‘; t) defined by
$(s, 53 t) = (explik - (R(s, 2) — R(s’, 0))]). (6.50)
By the same trick as that used in Section 6.3.2, we can show that
£96, 850=D F960. (6.51)
Since the distribution of R(s, 0) — R(s’, 0) is Gaussian with the variance
a|s—s'|, the initial condition for $(s,s';t) is obtained as (see eqn
(2.79)
2
os, $s Dlno=exp(— a \s -s'\) (6.52)
The boundary condition for $(s,s';t) is again obtained by the same
method as in eqn (6.30):
206 8'30|_ sik (w(L, fexplik - (RL, 2)~ RG’, 0))])
=ik- (u(L, Dexplik- (RL) - RG", 8)]
x explik - (R(s", t) — R(s", 0))1) (6.53)
where s” is again an arbitrary value between 0 and L. If s” is taken to beREPTATION DYNAMICS 203
close to L, the average of the underlined part can be taken separately
from the rest since its correlation time is very short. Hence
ZO6.5's0] ike (w(L, Dexplik RL, RE", NY)
x (explik -(R(s", RG", O)))
= 5; (empl R(6,)—RG" OM] 6" 5'30)
= — 2a exp[ - 3k’a(L—- s")]9(6", 532). (6.54)
In the limit of s"— L, this gives
a o(s, 8's o| = —Wag(L, 5’). (6.55)
S s=L
Similarly
246s 85 o| =2ka9(0, 5‘; t). (6.56)
S s=0
The solution of eqn (6.51) under these conditions is
0950-3 [eral t (°-3)]
cool (0 ~5) oP) + sy
(cine
cl e(—E) lee E) 22] cs
n= x La= £ Nb 4(KR,)* (6.58)
where
and @, and §, are the positive solutions of the equations
a, tan a, =p, B, cot Bp = —p. (6.59)
From eqn (6.57) we get”
2uN 2 ( “Duse)
=. ETF - . (6.
g(k, th= 3S at ab ep pia ‘a, exp| ie (6.60)
The expressions are simplified for the two limiting cases.
@) u<1, ie., KRg<«K1. Here eqn (6.60) is dominated by the term204 DYNAMICS OF A POLYMER IN A FIXED NETWORK
including a, = Vu <1, thus
a(k; t) = N exp(—4D,tu/L?) = N exp( - Peat R)= N exp(—Dok’t)
(6.61)
which is natural since, for KR, <1, the dynamical structure factor is
governed by the diffusion of the centre of mass.
(ii) w>>1, kR,>>1. In this case, the @, are approximated as
(p — 1/2), and eqn (6.60) reduces to
96N
ak, th= Pala aL 2p J exp(- tp?!)
= mi vl). (6.62)
This result is also derived by a simple argument. In the limit AR, >> 1, the
average of
L
7] &exolik- RG", ~ RG, 0)
is equal to 12/k?b? if R(s', t) remains in the original tube, and zero if it
has already disengaged from it. Thus using the probability (s, t) that it
remains, we have
L
ds 12 12
8(k, t)= | Lp VS = pape VO- (6.63)
0
Notice that the decay rate of g(k, t) is 1/Tz and depends strongly on the
molecular weight. This behaviour is entirely different from the result of
the Rouse dynamics, according to which g(k, t) becomes independent of
the chain length for kR, >>1 (see eqn (4.III.12)).
6.3.5 General time correlation functions
Time correlation functions of more complicated form can be calculated in
a similar way. For example, the time correlation function
C(s, s', 5"; t) = (RG, t) ~ R(s", 0) - (RGs', 1) — R(s", 0))) (6.64)
can be calculated from
a 2
Zc= p(2+ +a) (6.65)
under appropriate boundary conditions. In general the time correlationCONTOUR LENGTH FLUCTUATION 205
function C(s,,52,...55m,t) which depends on the m-points on the
primitive chain is obtained from
a m™ 9\?
Sc= v{ 55) Cc (6.66)
with appropriate boundary conditions.
6.4 Contour length fluctuation
6.4.1 Statistical distribution of the contour length
In the previous sections we regarded the primitive chain as an inexten-
sible string of contour length L. In reality, the contour length of the
primitive chain fluctuates with time, and the fluctuation sometimes plays
an important role in various dynamical processes.
First we consider the statistical distribution of the contour length. Since
the primitive chain represents a set of conformations of the Rouse chain,
the probability that a certain conformation of the primitive chain is
realized is proportional to w, the number of the conformations of the
Rouse chain which are represented by that primitive chain. The simplest
hypothesis is to take the polymer as a random walk confined within a
tube. Then @ is calculated by the method described in Section 2.3.2 (see
Appendix 6.1). The result is
2 2
3 Nb ) 6.61)
w(L) = @ exp( — ine? oo
where Wo is the number of the configuration in the free space, ao, the
diameter of the tube, and a, a certain numerical factor which depends
on the shape of the cross-section of the tube. The logarithm of w gives
the entropy of the primitive chain
2. ast)
2Nb? °° ad
At first sight eqn (6.68) may seem to imply a paradoxical result: since
the entropy increases with decreasing L, the chain will contract to the
state of L=0 if its ends are not fixed. This argument is wrong. The
collapse happens if the chain is confined in an infinitely long tube of given
conformation, but does not happen in the case of a network, where the
Rouse chain explores many tubes. To calculate the statistical distribution
of L in the network, one has to take into account the multiplicity of the
State specified by L. Let Q(L) be the number of primitive paths which
have length L, then the probability that a primitive chain has a contour
S(L) = 5 ko (6.68)206 DYNAMICS OF A POLYMER IN A FIXED NETWORK
length L is
W(L) & w(L)Q(L). (6.69)
Q(L) can be estimated by the number of random walks consisting of L/ao
steps:
Q(L) = exp(a *) (6.70)
where «, is a certain numerical constant which depends on the structure
of the network. From eqns (6.67), (6.69), and (6.70)
W(L) x exp( - Bieta) « «exp| — a3 (L- DF (6.71)
which has a maximum at
- _ a, Nb?
L= 3 ap : (6.72)
From eqns (6.4) and (6.72)
3
a= a, Q. (6.73)
This shows explicitly that the step length a of the primitive chain is of the
same order as the tube diameter ao. The average of the fluctuation is
calculated from eqn (6.71) as
AL= (AL?) = [ j dL Y(L)(L— oy" =(Nb?/3)'? for L>>VNb.
(6.74)
This statistical property of the primitive chain has been studied by
computer simulation.’® Under certain conditions, the statistical distribu-
tion of the primitive chain can be calculated analytically.'7""* Both studies
show the relations
(L)*N and (AL?)«N (6.75)
in agreement with eqns (6.72) and (6.74).
6.4.2 Dynamics of the contour length fluctuation
Having studied the static distribution of the contour length, we now
examine the dynamics of the contour length fluctuation. As before, we
use the Rouse model for the dynamics (see Fig. 6.7). Let s, be the
curvilinear coordinate of the n-th Rouse segment measured from aCONTOUR LENGTH FLUCTUATION 207
iN
Sy
Fig. 6.7. A Rouse model describing the contour length fluctuation. The point O
denotes the origin of the curvilinear coordinate s,.
certain fixed point on the tube. The contour length of the primitive chain
is defined by
L(t) = sy(t) — So(t)- (6.76)
The dynamics of s, are described by the Langevin equation for the Rouse
model,
a 3kpT #
C59 GE Byatt) that) (6.7)
where
(fit) =O and (f,()fm(t')) = 26keT5(n — m)d(t— 1"). (6.78)
Since eqn (6.77) is the same linear equation as appeared in the Rouse
dynamics, analysis of this equation is straightforward. An important
difference, however, must be mentioned. In the present case, the
equilibrium average of the contour length is L:
(sv ~50) = £, (6.79)
while in the Rouse model the corresponding quantity (Ry — Ro) is zero.
To get eqn (6.79) we have to modify the boundary condition which
corresponds to eqn (4.11). This is obtained as follows.
If we take the average of both sides of eqn (6.77) for the equilibrium
state, we have
BS (sq) = AES (a0) + 600). (6.80)
Both 3({s,,)/dt and (f,(t)) vanish at equilibrium, so that
& (Sn) = 0. (6.81)208 DYNAMICS OF A POLYMER IN A FIXED NETWORK
For this to be consistent with eqn (6.79), the boundary condition must be
a L
ay at n=OandN. (6.82)
The boundary condition (6.82) can be visualized as a hypothetical
tensile force acting on the chain ends (see Fig. 6.82):
_3ksT , _ 3keT
“a Nb? a
The origin of this force is again the multiplicity of the tube. One can
intuitively understand it by considering the dynamical process at the
chain end. Suppose in a time At the chain end moves one step in a
random direction (see Fig. 6.85). If it moves to the (z — 1) positions Ai,
A2,...,Az-1, the contour length increases, whilst if it moves to Ao, the
contour length decreases. Thus there is an imbalance in the change,
which tends to increase the contour length and causes the force (6.83).
For eqn (6.82) to be satisfied, the normal mode is now defined as
(6.83)
N
1
hay J dns, (6.84)
17 7 LD
= pan\( nh =
Y= anoos( Fi") 4) for p=1,2,..., (6.85)
es e ° °
@ (b)
Fig. 6.8. (2) Equilibrium tension acting on the primitive chain. (6) Physical
origin of the force F,,. The case of z = 4 is shown.CONTOUR LENGTH FLUCTUATION 29
or
= : pan) nb
m= %+2 D Yoo oos( 2) + (6.86)
The coordinate Yq represents the position of the ‘curvilinear center of
mass’,
N
sol) =% [ anss()= (0, a)
0
while the other coordinates Y, (p >0) represents fluctuations along the
tube.
The equation of motion for Y, is precisely the same as eqn (4.19):
9Y,
62 ri =-k,Y, +h (6.88)
with
So=NE, = 2NE for p=1,2,...5 6.89)
67k,
p= ge Ps (6.90)
and
($A) =0, (M(t!) = 26pkT6p,5(t — t'). (6.91)
Thus the time correlation functions for Y, are obtained as
((%) — %O)?) = 2at t (6.92)
Nb?
«¥5(0%(0)) = Zaz exP(—1P Vt), (6.93)
where Tz is the Rouse relaxation time given by eqn (6.18).
Equation (6.92) justifies the previous assumption that the diffusion
coefficient D, is equal to kgT/N€, the diffusion constant of the Rouse
chain.
From eqn (6.86) the contour length of the primitive chain is written as
LQ) = sy(t) ~ So(t) (6.94)
=L-4 a Y,(t). (6.95)
The time correlation function of L(t) is calculated from eqns (6.93) and
(6.95). The result is
8Nb? 1
=
= exp(—tp”/ tp). (6.96)
(HOLO) = B+ Se So210 DYNAMICS OF A POLYMER IN A FIXED NETWORK
In particular,
8Nb? 1 _ Nb?
==> 6.97
aa pgp? 3 0)
which agrees with eqn (6.74). The characteristic time of the contour
length fluctuation is tg, the Rouse relaxation time. Since
teil Ab (2).
ta 2 L NL vz’
the effect of the contour length fluctuation can be neglected for Z >> 1. It
is at this limit that the dynamics described in Section 6.3 is valid.
(AL?) = (12) ==
(6.98)
6.4.3 Effect of the contour length fluctuation on reptation
Though the contour length fluctuation becomes negligible for very large
Z, its effect is not negligible for usual values of Z, which are typically less
than 100. An important effect is found in the disengagement time t,.'°
Consider the two situations shown in Fig. 6.9, one represents the motion
of the chain with fluctuations, and the other without. Obviously the life
time of the tube becomes shorter for the chain with fluctuation than for
the chain without, i.e., the contour length fluctuation reduces the
disengagement time. This effect is estimated as follows.
If the contour length fluctuation is neglected, the disengagement time
is estimated as the time necessary for the chain to move the distance L,
7") = [/D, (6.99)
Oe, CIEE, momma aeE
LLL SOLE,
time
Fig. 6.9, The Brownian motion of a primitive chain with (a) fixed contour length,
and (b) fluctuating contour length. The oblique lines denote the region that has
not been reached by either end of the primitive chain. Obviously the length of
this region o(t) decreases faster in (b) than in (a). Reproduced from ref. 15.CONTOUR LENGTH FLUCTUATION 211
(The superscript (NF) stands for ‘no fluctuation’.) On the other hand if
there are fluctuations, the chain can disengage from the tube when it
moves the distance L — AL because the chain ends are fluctuating rapidly
over the distance AL. Hence the disengagement time is estimated as
oP = (L- AL)/D,. (6.100)
(The superscript (F) stands for ‘fluctuation’.) From eqns (6.98) and
(6.100), it follows that
TP) = (1 - xy (6.101)
where X is a certain numerical constant.
Precise calculation of + requires the first passage problem in
multidimensional phase space. A variational calculation’® for the Rouse
model shows that X is larger that 1.47. Hence the effect of the contour
length fluctuation is significant even if Z is as large as 100.
The effect of the contour length fluctuation has been studied for
slightly different models'”"* and it has been reported that the effect is less
significant than in the case of the Rouse model. The discrepancy is
perhaps due to the difference in the dynamics of the fluctuation,
especially the short-time dynamics, which is quite important in the first
passage time problem.’?”
6.4.4 Other small-scale fluctuations and their effects on the segmental
motion
Due to the small-scale fluctuations around the primitive path, the actual
dynamics of the Rouse segment in a network is much more complicated
than that described in Section 6.3.2. As an example, let us consider the
mean square displacement of a Rouse segment.
alt) = ((Ralt) — Rn(0)))- (6.102)
The precise calculation of ¢,,(t) is difficult, but its characteristic features
can be inferred easily.
(i) For a very short time the segment does not feel the constraints of
the network, so that ¢,(t) is the same as that calculated for the Rouse
model in free space. Hence a is given by eqn (4.III.6) as
4Nb? (pan
alt) = 6D.1+ 22 S 5 c0s*( PZ) 1 — exp(—tp*/re)). (6.103)
p=1P
Since t<< tr, eqn (6.103) is approximated in the same way as in eqn212 DYNAMICS OF A POLYMER IN A FIXED NETWORK
(6.37),
on | dp H(1~ exp(-9°/=)
= ONB*CI tg)!” = (kpTb7t/£)'. (6.104)
This formula is correct when the average displacement is much less than
the tube diameter. Let 1, be the time at which the segmental displace-
ment becomes comparable to a:
a? = y(t) = (kp Tb7/E)'Vr, (6.105)
ie.,
t= aC lkpTb?. (6.106)
The time 1, denotes the onset of the effect of tube constraints: for t< t.,
the chain behaves as a Rouse chain in free space, while for t>, the
chain feels the constraints imposed by the tube,
(ii) For t>+, the motion of the Rouse segment perpendicular to the
primitive path is restricted, but the motion along the primitive path is
free. The mean square displacement along the tube is calculated from
eqns (6.86)—(6.93) as
Cat) ss(O)?) = ((Hle)— YAO?) +4 3 cos "CHC — HOP)
kpT sn? a
a2 ae pees oos?(P= a "a — exp(-tp?/tp))
=f" St, 6.107
keTt/NE tt. (6.108)
From this, $,(¢) is estimated as in Section 6.3.2. If the segment moves
5,(¢t)—s,(0) along the tube, the mean square displacement in the
three-dimensional space is a |s,(¢) — s,(0)|. Hence
n(t) = a s(t) — 5n(0)|) =a ((n(2) — sn)". (6.109)
From eqns (6.107)—(6.109)
a(kpTb7t/t)"* 4,StStp,
nll) (eaenney” tp StS Ty. (6.110)
Note that ¢,(¢) is proportional to ¢” in the time region 1, St tp. This
specific diffusion behaviour, first predicted by de Gennes,’ is a conse-
quence of the two effects, the Rouse-like diffusion equation (6.107), and
the tube constraints equation (6.109). On the other hand, the behaviourCONTOUR LENGTH FLUCTUATION 213
log<(R(t)—Fa(0) }?>
log t
Fig. 6.10. Mean square displacement of a chain segment is plotted logarithmically
against time.
for tt, agrees with that predicted by the primitive chain dynamics
equation (6.37).
(iii) For t= 1,4, the dynamics is governed by the reptation process. As
discussed in the previous section
2. kpyTa?
¢,(t) = ze = NAEp! tet (6.111)
Equations (6.104), (6.110), and (6.111) are summarized as follows”
Nb*(t/tp)? tst,
Nb(t/Z?tz)"* 1 St St,
PO ~4 np2e/24)! n*”
On the other hand there is much experimental evidence which
indicates that reptation is the dominant motion governing the dynamics in
the highly entangled state.
Clear evidence comes from the study of diffusion. Klein ef al.®°
measured the diffusion constant of a deutrated polyethylene in a
polyethylene matrix, and found
Dg *M~? (7.1)
in agreement with the result of reptation dynamics (eqn 6.40). The
self-diffusion constant of a labelled polymer in solutions and melts has
been measured by other techniques (forced Rayleigh scattering’® and
field gradient Nuclear Magnetic Resonance,""”) and eqn (7.1) is fully
confirmed in these systems. (A comprehensive review is given by
Tirrell.>)
More evidence comes from the study of viscoelasticity, which has been
done extensively in the past and established the characteristic aspects
common to all flexible polymers.”> The reptation model has succeeded in
explaining many of these features and also predicting some of the
behaviour in nonlinear viscoelasticity. In this chapter we shall describe
the reptation theory'*> for viscoelasticity in detail, and discuss the
validity of the reptation model in solutions and melts.
The viscoelastic properties of polymers are quite important in polymer
technology. A great deal of experimental work has been done, and at the
Same time phenomenological theories have been developed to a highly
sophisticated level. These are summarized in various monographs.”!*1®
An overview of this field can be obtained in the excellent textbook by
Bird et al.’ Here we shall limit ourselves to the molecular aspects of the
Problem, i.e., how the viscoelastic properties are related to the molecular
dynamics and how they depend on molecular parameters such as
molecular weight, concentration, and molecular structure.220 MOLECULAR THEORY OF POLYMERIC LIQUIDS
7.2 Microscopic expression for the stress tensor
7.2.1 Stress in polymeric liquids
To understand the mechanical properties of materials, it is important
to consider first the microscopic origin of stress. In the usual gas or liquid
of small molecules, the stress comes from the momentum transfer due to
the intermolecular collision. In polymeric liquids, the stress is mainly due
to the intramolecular force, and is directly related to the orientation of
the bond vectors of the polymer. This idea, originated from the theory of
rubber elasticity,” is fundamental in the physics of polymeric materials.+
To demonstrate the point, let us first consider a dilute polymer
solution. As shown in Section 4.5.2, the stress tensor is written as
Cap (t) = (Kap (t) + Kpa(t))ns + OF}(t) + POap (7.2)
where 7, is the viscosity of the solvent, x, is the velocity gradient tensor
and o® represents the contribution from the polymer
= 2 aE (Hn Ree)
oC) N 2 b? an an | (7-3)
This is directly related to the orientation of the vector OR,,/dn, and it is
this term that gives the viscoelasticity of dilute solutions.
In the dilute solution, the viscoelasticity has only a small effect; the
major contribution to the stress is the purely viscous stress given by the
first term in eqn (7.2). The situation changes with increasing polymer
concentration. The contribution of o®} increases steeply (because the
polymers become more easily oriented due to the entanglement), while
the contribution of the first term remains essentially unchanged. When
of} dominates the first term, the total stress is simply related to the
orientation of the bond vectors:
cm EE OE), a
Note that eqn (7.4) is valid even if the excluded volume effect is
accounted for since, as shown in eqn (4.135), the pseudo-potential
described by the delta function does not change the expression for the
stress tensor (apart from the isotropic stress, i.e., the pressure).
+ Historically, the molecular theory for the mechanical properties of polymeric materials
was first constructed for rubbers, and it was realized that the stress is related to the
orientation of the polymer segments. That the stress in polymer melts has the same
molecular origin as in rubbers was noted by Green and Tobolsky,”' who regarded the
polymer melt as a kind of network with temporary junctions. This idea has been the base of
successful phenomenological theories'®- and it is indeed used in the present theory.MICROSCOPIC EXPRESSION FOR THE STRESS TENSOR 221
The above argument is essentially for semidilute solutions, and may
not apply to concentrated solutions and melts in which the intermolecular
interaction cannot be described by the pseudo-potential. However, since
the intermolecular interaction does not have any specific polymeric effect
(for example the associated relaxation times will be as short as in a liquid
monomer), one can argue that their effect will be only to give a constant
viscous stress. Indeed eqn (7.4) is derived from the fact that even in the
highly entangled state the short-time dynamics of polymers is described
by the Rouse model. For, as was discussed in Sections 3.7.4 and 3.8.4,
the expression of the stress tensor is entirely determined by the
short-time dynamics: for example, the elastic stress is given by the change
in the free energy for an instantaneous deformation. Therefore if the
short-time dynamics is given by the Rouse model, the expression for the
elastic stress must be the same as for the Rouse model, which is eqn
(7.4). Thus eqn (7.4) holds quite generally in polymeric liquids in which
the viscous stress is negligibly small.+
Since the short-time dynamics of polymers is unaffected by the global
molecular structure, such as branching or crosslinks, the same argument
will also hold for branched polymers and gels. In these systems, the stress
is written as
_ aT (Rael) OR p(t)
onl) = all aoe b on on ) (7-5)
in unit volume
where R, is the position of the segments which are chosen small
compared to the distance between the branching points (or crosslink
points).
7.2.2 Stress optical law
The above argument is crucially supported by the stress optical law,”**
which is obtained by comparing eqn (7.4) with the formula for the
intrinsic birefringence (eqn (4.238))
n, = Coze(t) (7.6)
where
2n(n? + 2)?
c 27kpTn 7.7)
That the viscous stress of polymeric liquids is negligibly small compared to the elastic
Stress is well established experimentally. Indeed this fact has been taken as a basic postulate
in Coleman’s phenomenological theory” (see also Chapter 4, ref. 17). It must be noted that
this is the result of the elastic stress contribution becoming so large relative to the viscous
Stress: the absolute magnitude of the viscous stress will not differ considerably between
simple liquids and polymeric liquids.222 MOLECULAR THEORY OF POLYMERIC LIQUIDS
In concentrated solutions and melts, the intrinsic birefringence n{) can be
regarded as the total birefringence since the form birefringence is
negligibly small as was shown in Chapter 5. Thus eqn (7.6) is written as
Nap = COng + isotropic tensor. (7.8)
This is the stress optical law, which has been found experimentally to
hold for rubbers and the more general problem of polymeric liquids.”
Two important aspects of eqn (7.8) must be mentioned.
(i) The simple linear relation between the stress and the birefringence
holds even when the relation between the stress and the velocity gradient
becomes quite complex (generally a nonlinear functional). Thus eqn (7.8)
has much deeper physical significance than the similar relation for
amorphous solids, which is derived by the symmetry argument and is
limited to the Hookian range of elasticity.
Gi) The proportional constant C is a function of the local condition
such as the temperature, solvent, and polymer concentration but is quite
independent of the features of molecular structure on a large length-scale
such as the molecular weight, molecular weight distribution, branching,
and degree of crosslinking.t
These results are fully confirmed by experiment. The only exptanation
for such a general relation is that both the stress and the birefringence
have the same physical origin, i.e., the orientational ordering in the
polymer segments.
In deriving eqn (7.8) we assumed that (a) the relation between the
orientation of the bond vectors and that of the end-to-end vector of the
Rouse segments is linear (eqn (4.1V.13)) and that (b) the form
birefringence is neglected. The stress optical law breaks down when
either of these conditions is not met.”4 For example, the first condition is
Not satisfied when the stress is very large, as often happens in experi-
ments near the glass transition temperature. The second condition is not
satisfied when the sample includes a spatial inhomogeneity as in the case
of block copolymers or polymer blends which include microphase
separation. Except for these situations, the stress optical Jaw holds quite
generally in polymeric liquids.
7.3 Linear viscoelasticity
7.3.1 Background of phenomenological theory
The viscoelastic properties of a material are entirely characterized by the
constitutive equation which relates the stress tensor 0,¢(¢) to the velocity
+ Though the excluded volume interaction (and other Van der Waals’ type of interaction
among the segments) do not violate the validity of eqns (7.4) and 7.8)" they do affect the
stress optical coefficient C. For example C is sensitive to the nematic-like interaction which
tends to orient the neighbouring segments in the same direction.”*?LINEAR VISCOELASTICITY 223
gradient tensor x,g(t). In usual fluids the relationship is simple,
Oap(t) = (Kap(t) + Kpa(t))n- (7.9)
In polymeric liquids the constitutive equation becomes quite complicated:
the stress depends on the previous values of the velocity gradient tensor
and the dependence is generally not linear; hence the relation is written
only by a nonlinear functional relation:
Oap(t) = Fg[Kap] for Kag(t') with t' <¢. (7.10)
However, in the special case that the perturbation by the velocity
gradient is small, the relation between the stress and the velocity gradient
becomes linear and is written as””?
Oup(t) = | dt'G(t — t’)(Kap(t’) + Kga(t'))- (7.11)
This constitutive equation includes only one material function G(t),
called the shear relaxation modulus.
Though the applicability of eqn (7.11) is limited, linear viscoelasticity
has been studied in great detail as it represents a property of the material
in a well-defined limit.
For shear flow
v,=K(t)y, vy=0, vu, =0, (7.12)
equation (7.11) reduces to the form given in Chapter 4 (eqn (4.116)),
O,y(t) = { dt’G(t — t')k(t'). (7.13)
Let y(t) be the shear strain measured from the state at t= 0:
y= fareey (-@0, (7.17)
for which, eqn (7.16) gives
O,(t) = YoG(*). (7.18)
This provides a direct determination of G(t).
(b) Oscillatory shear (Fig. 7.1b). This has already been discussed in
Chapter 4. The shear strain is give by
Y(t) = Yo cos (wt) = yo Re(e). (7.19)
The response defines the storage modulus G'(w), loss modulus G’(w),
and the complex modulus G*(w),
Ory(t) = yo(G" cos( wt) — G"(w)sin(wt))
= yo Re(G*(w)e'”) (7.20)
where
G*(@) = G'(w) + iG"(w). (7.21)
It is easy to show that eqn (7.16) gives
G*() =io | dte“"G(t) (7.22)
0LINEAR VISCOELASTICITY 225
and i =
G'(o)=0f dt sin(wt)G(0) Grw)=0f dt cos(wt)G(). (7.23)
(c) Creep (Fig. 7.1c). When a constant shear stress 0 is applied to the
system at equilibrium, the system starts to flow. The shear strain y(t) is
obtained from the integral equation:
oo= j aro — 2) 2, (7.24)
dt
This can be solved by the Fourier—Laplace transform:
-id+0
do el
n= 00 |e act@ (7.25)
where 6 is an arbitrary positive constant. The behaviour of y(t) for r—> ©
is governed by the contribution from the pole at w = 0. Since for small w,
G*(@) is given by
G*(@)= iw axc1 —iot+...)G(t)=iwgotw’git... (7.26)
oO
with
B= [dG and g.= [aGey, (7.2)
oO 0
the contribution from the pole at w = 0 is given by
81
= 0+ +8). 1.28
(0) = 00) ao ge (7.28)
For large t, eqn (7.28) gives o)=gody/dt, whence gy can be identified
with the steady state viscosity mo. The constant term g,/g2 is called the
steady-state compliance and written as J{. Thus
o= j 1G (0)= fin, FO,
JO= | dG (t / [ j aio] =i Jim a oor (7.30)
(7.29)
7.3.2 Calculation by Rouse model
We now study the linear viscoelasticity from the molecular viewpoint.
First we consider Rouse dynamics. This corresponds to the case of short226 MOLECULAR THEORY OF POLYMERIC LIQUIDS
polymers in melts. The basic equations and their development are
precisely the same as in Sections 4.5.3 and 4.5.4. The only distinction is
that the viscous stress is now negligibly small, so that G(s) in eqn
(4.158), which corresponds to the polymer contribution to the relaxation
modulus in Chapter 4, is now regarded as the relaxation modulus of the
system. Thus
Gi)= cnet > exp(—2tp?/ tp) (7.31)
pol
where
__EN%?
ET (7.32)
The viscosity and the steady-state compliance are calculated from G(¢t),
_x = (1), cf
ww) ta 5g NO? (7.33)
no= [ AGU) = yKoT He 3,2
Oo
and
JO= 5 facor=— =
Noy
#(Se)(Ze7) = wer (7:34)
ckpT \m1
These are written in terms of the molecular weight M, the weight of
polymers in unit volume p (=cM/NN,), and the gas constant R (=N,kg)
as
Te & M2, (7.35)
x” (pR
N0= (=) TR* pM, (7.36)
IO= oar (137)
These results are confirmed for polymer melts with low molecular
weight.2>
7.3.3 Calculation by reptation model
Next we consider a polymer melt of high molecular weight in which
entanglement is very important. To calculate G(f), it is convenient to
consider the stress relaxation after a step strain. Suppose at t= 0 a shear
strain y is applied to the system in equilibrium. The strain causes the
deformation of the molecular conformation, and creates the stress, which
Telaxes with time as the conformation of polymers goes back toLINEAR VISCOELASTICITY 227
equilibrium. Though description of this process in the general case is
slightly involved (see Section 7.5), the relaxation modulus in the linear
regime can be obtained by a simple argument.
(A) For small ¢ (¢<1,), the dynamics is described by the Rouse model
and the relaxation modulus is given by eqn (7.31). Since t, «tz, eqn
(7.31) is approximated as
2 Tr\'?
Gt) ha | do cat 20p*Ite) = say ko ot(=) (tt,
is only due to the disengagement. This can be evaluated as follows (see
Fig. 7.2).
At t=+t,, the whole polymer is confined in a deformed tube. As time
passes, the polymer reptates, and at time ¢ the parts of the polymer near
the ends have disengaged from the deformed tube, while the part in the
middle is still confined in the tube. Since only the segments in the
deformed tube are oriented and contribute to the stress, the stress is
Proportional to the fraction w(t) of the polymers still confined in the
LE
LL
“Lily
Fig. 7.2. Explanation of the stress relaxation after small step strain. (a) Before
deformation, the conformation of the tube is in equilibrium. (6) Immediately
after the deformation, the whole tube is deformed. The deformed part is
indicated by the oblique lines. For small strain, the contour length of the tube is
unchanged. (c) At a later time #, the chain is partly confined in a deformed tube.
The average of the contour length o(s) of this part is equal to Ly(t).228 MOLECULAR THEORY OF POLYMERIC LIQUIDS
deformed tube, i.e.,
G(t)=GPy) zt) (7.39)
where G® is a certain constant and *(t) has already been calculated in
Section 6.2.2 (see eqn (6.16)),
vor S, ap eor(-Pt!t) (7.40)
with
‘N36? (b
= (°) . (7.41)
To obtain G®, we utilize the fact that at t= t,, the Rouse-like behaviour
(eqn (7.38)) smoothly crosses over to the reptation behaviour (eqn
(7.39), ie., ”
GY ~ G(s.) = 5 ket (=) (7.42)
Te
or using eqn (6.106)
b?
CP =F ka®. (7.43)
Equation (7.38) is then written as
G(t)= ae(%)" for t! (see Fig. 7.4).
Maximum relaxation time, viscosity, and steady-state compliance. The
longest relaxation time Tmax Of G(#) is ta:
“N35? /b\? Nb?
ax = a= Seat (a) “3° te (7.46)
The viscosity and the steady-state compliance are calculated from eqns
(7.29) and (7.30). Since the contribution to the integral from the region
t°5 An
example of the viscosity in melts is given in Fig. 7.6. The reason for the
discrepancy will be discussed later.
7.3.5 Tube diameter in melts
Given the general agreement in the shape of the relaxation modulus, it is
possible to determine the step length a of the primitive chain’ Though
various ways are conceivable, a direct way is to use the plateau modulusLINEAR VISCOELASTICITY 231
log J. (om?/dyne)
log My,
Fig. 7.5. Steady-state compliance of polystyrene at 160°C. Dotted line represents
the result of the Rouse model.
O Plazek-O’Rourke @ Mills—Nevin © O’Reilly-Prest
® Murakami et al. © Tobolsky et al. @ Nemoto
® Akovali ® Onogi et al. © Mieras-Rijn
Reproduced from ref. 35.
G®, which is related to a by eqn (7.43). Since
cb? = £0? = 9M no", (7.50)
eqn (7.43) is written as
2
GP= eRt we (7.51)
Experimentally G® is often expressed by a characteristic molecular
weight M,, called the molecular weight between entanglements, which is
defined by
RT
M,= eo . (7.52)) |»
—T T_T
Poly(di-methyl
siloxane)
Constant + log 19
T
Poly(iso-butylene)
|. Poly(ethylene)
Poly(butadiene)
Poly(tetra-methyl
p-silphenylene
[ _ siloxane)
Poly(methyl
| methacrylate)
x
Poly(ethylene
glycol)
Poly(vinyl acetate)
Poly(styrene)
log 10'7x, '
Fig. 7.6. Steady-state viscosity of various polymers in melt, where X, is a
parameter proportional to M,. The curves are shifted vertically so as not to
overlap each other. Reproduced from ref. 33.LINEAR VISCOELASTICITY 233
It follows from eqns (7.51) and (7.52) that
M, wm.
ax (Gene?) =u, (7.53)
where Ry, is the root mean square of the end-to-end distance of the
polymer with the molecular weight M,.
The precise numerical coefficient in eqn (7.53) is not given by the
simple argument given in Section 7.3.3, To determine the coefficient, a
further assumption is needed about the deformation of the tube under
strain. A specific model described in the next section gives
a= ait No =0.8R%, (7.54)
which we shall now use.
The value of M, in polymer melts is tabulated in the literature,” from
which a is found to be 82 A for polystyrene and 34 A for polyethylene.**
This distance is much larger than the correlation length measured by
neutron scattering or X-ray scattering. At present, it is not fully
understood what factors determine a, but various semiempirical relations
are available for M,.°>°78 Here we shall proceed regarding @ as an
adjustable parameter.
Graessley®* showed that the above estimation of a is consistent with the
results of diffusion experiments. According to the reptation theory, the
self diffusion constant Dg is given by (see eqn (6.40))
thet
°°" 3Nb? NE-
To eliminate , it is convenient to consider the zero shear rate viscosity
calculated by the Rouse dynamics
n§(M) = one? = EN NER, (7.56)
(7.55)
where R%, is the mean square end-to-end distance of the polymer of
molecular weight M. Experimentally, {°(M) can be obtained by
extracting the viscosity for low molecular weight according to the
following equation (see Fig. 7.7)
nM) = no Mas) (7.57)
where M,,¢ is an arbitrary molecular weight in the Rouse regime.234 MOLECULAR THEORY OF POLYMERIC LIQUIDS
og 0
log M
Fig. 7.7. Viscosity curve. Solid line: experimental formula eqns (7.71) and (7.72).
Dashed line: theoretical curve, eqn (7.73).
From eqns (7.54)-(7.56), it follows that
1 a(R us) M,_pRT_
6735 M n®(M)"
According to Graessley, eqn (7.58) gives values in reasonable agree-
ment with experimental results*”? (e.g. for the case of polyethylene the
error being only about 30%). Considering the different nature of the
experiments, the agreement is remarkable.
That the length a is rather large seems to be consistent with the fact
that both neutron scattering® and computer simulation’! do not find it
easy to detect reptation: the characteristic time-scale in these experiments
is often shorter than t,. Recently, however, it has been reported that
indications of reptation are observed in some situations.“
(7.58)
7.3.6 Semidilute and concentrated solutions
So far we have been considering polymer melts. It is expected that the
same picture will hold in semidilute and concentrated solutions. Though
this is generally believed to be the case, a few remarks must be made.
(i) In semidilute solutions, the excluded volume effect and the
hydrodynamic interaction become important for dynamics on a length-
scale shorter than the correlation length & (or the hydrodynamic
screening length &,). The problem of how this affects the reptationLINEAR VISCOELASTICITY 235
picture is delicate. However, in the case of a good solvent, certain
conclusions can be drawn from scaling arguments.“*+
Consider for example the maximum relaxation time Tnx, the steady
state viscosity mo, and the plateau modulus G{. If one imposes the
condition that they are invariant under the scaling transformation given
by eqn (5.138), one can show that
_te N25? .
Fax = Ailcle*), (7.59)
No = nsh(e/e*), (7.60)
and
Ge =< keTfi(c/c*), (7.61)
where c* is the overlap concentration
1-3v
ctx B (7.62)
The concentration dependence can be determined if one imposes a
further condition that eqns (7.59)-(7.61) must be consistent with the
reptation prediction:
Tmax ®N?, No N2, GY N°, (7.63)
then
N32"? / c \O-2GV—D
ton" (5) = Mp™, (7.64)
c\¥Gr-0)
No= 1s >) = M*p%4, (7.65)
and Bi d
GQ =LkyT(£) ep (7.66)
NO? \ct eo .
(The last expressions indicate the result when v is equal to 3/5).
Experimentally, eqns (7.64)-(7.66) seem to hold at least approximately in
the semidilute regime,***’ though the exponents in the molecular weight
dependence of tna, and 19 are slightly larger than 3, and the exponent in
the concentration dependence of G{? is somewhat higher than 9/4
depending on the quality of the solvent.
+ It must be mentioned that whether dynamical scaling holds for systems dominated by the
topological interaction is still a matter of debate, though eqns (7.64)-(7.66) seem to be in
agreement with experimental results.236 MOLECULAR THEORY OF POLYMERIC LIQUIDS
(ii) At higher concentration, the effect of the excluded volume and the
hydrodynamic interaction becomes less important. If we disregard these
effects, the dynamics is described by the same model as in melts except
that a and € now depend on concentration. Experimentally it has been
found*“*° that
G® « p? (7.67)
which implies
axp-'?, (7.68)
The concentration dependence Of Tmax and Mo is delicate since the friction
constant £ depends on the concentration in a nontrivial way.
7.4 Other relaxation modes
7.4.1 Discrepancy between the theory and experiments
Although the theory given in the previous section is largely in agreement
with experiments on linear viscoelasticity, there remain certain
discrepancies.
(i) The theoretical molecular weight dependence TloX MP and Tmax
M? is weaker than the experimental one; the measured exponent is
higher than 3, ranging from 3 to 3.7.>*
(ii) The theoretical relaxation modulus G(t) is too close to a single
exponential compared to the experimental modulus.” For example, the
experimental value of JG, which measures the deviation from the
single exponential behaviour of G(t), is between 2 and 3**" as against 6/5
for the theoretical result of eqn (7.48).
Part of these discrepancies can be attributed to the molecular weight
distribution, which seriously affects the value of J&.> On the other hand
detailed comparison with experimental results indicates that not all the
discrepancy can be resolved by the molecular weight distribution.”
The discrepancy in the exponent of the viscosity has been a matter of
debate which is not yet settled. Various modifications of the reptation
picture have been proposed. For example, Wendel and Noolandi®
argued that if the polymers are trapped by some tight knots with an
extremely long lifetime, the diffusion along the tube becomes non-
Fickian and this gives a higher exponent in 7) M*. However, such tight
knots, if they exist, would create a rubbery plateau with extremely long
+A similar problem exists in the melt, where it is often observed that the viscosity does not
follow the Rouse behaviour at small molecular weights much less than M,. This is attributed
to the fact that for short polymers the segmental friction constant € depends on the
molecular weight.*?OTHER RELAXATION MODES 237
relaxation time, which no viscoelastic data seem to support. Curtiss and
Bird™ suggested that the segmental friction constant may depend on
the molecular weight, but this hypothesis seems to contradict the result of
the diffusion experiment. Another possibility suggested by Ball® takes
into account the long-range correlation in the motion of vacancies needed
for the reptation motion to take place. This has been studied by
computer simulation,** but the result is not yet conclusive.
At present, a more consistent explanation seems to be that given by
Graessley,° who pointed out that the observed viscosity and the
relaxation time are smaller than the calculated ones. Using eqns (7.32),
(7.33), (7.46), and (7.47), one can show that
3 2
nol) =" no) ) = nay 2) (7.68)
and
3
Tan aoe MA(gr) GEM). (7.70)
On the other hand, a widely accepted empirical formula is
no(M.) aM for MM., (7.72)
where M, is a certain molecular weight which is two or three times larger
than M,. Using M,=2M., and no(M.) = no(Mc)(M./M-), eqn (7.69) is
written as
nst(M) = 15 M(H)» 7.73)
nf**(M) is about 15 times larger than n{*°)(M) at M,, but the discrepancy
decreases with increasing M and diminishes at M =(15)'°*M, ~ 800M,
(see Fig. 7.7). Graessley thus conjectured that although the
Pure reptation behaviour will be observed for very large molecular
Weight, there is a large cross-over region in the viscosity from the
Rouse-like behaviour to the pure reptation behaviour, which gives an
apparent exponent larger than 3.
Unfortunately since no data are available for molecular weights higher
than 800M,, which is 3 x 10’ for a polystyrene melt, the crucial test of
Graessley’s conjecture has not been given. However, it is obvious that238 MOLECULAR THEORY OF POLYMERIC LIQUIDS
any relaxation process which occurs concurrently with reptation de-
creases the viscosity and hence reduces the discrepancy between the
theory and experiment.
7.4.2 Contour length fluctuation and tube reorganization
Two relaxation processes have been suggested to alter the pure reptation
behaviour discussed in Section 7.3.3.
Contour length fluctuation. As was shown in the previous chapter, the
contour length fluctuation reduces the disengagement time ty
significantly. From eqns (6.20), (6.101), and (7.54), the disengagement
time +? of a chain with fluctuation is given by
‘M, 12,2
P= o7(1 -x(F) ) (7.74)
M
where X’ is estimated as 1.47+* V(4/5) =~ 1.3.°’ For M/M, =50 the ratio
between 1 and r{" is about 0.67 which displays the considerable effect
of the contour length fluctuation in the region where the polymers are
usually regarded as ‘fully entangled’. A crude calculation’”* indicates
that the discrepancy in the viscosity and the steady state compliance is
significantly improved if the contour length fluctuation is taken into
account.
Tube reorganization. So far, it has been assumed that the tube is fixed in
the material and its conformational change occurs only at the ends. It is
conceivable that the conformational change of the tube can occur in the
middle.“ For example:
(i) Constraint release: The topological constraints for a polymer can be
teleased (or created) by the reptation of the surrounding polymers as
shown in Fig. 7.8. This will cause the conformational change of the tube
in the middle. A model describing this process is to regard the
conformational change as a local jump of the primitive chain. Since the
jump rate is of the order 1/t,, this process has a negligible effect on the
longest relaxation time. However, the process gives an additional
relaxation to G(t) in the plateau region, and improves the value of
IOGY.
(ii) Tube deformation: If a polymer is in a strained conformation, it
will tend to relax the strain by creating the deformation of the
surrounding polymers. This effect may be handled by considering the
deformation of a strained chain placed in a viscoelastic medium. So far
NO quantitative estimation of this effect has been done.
Though these processes are conceivable, estimation of their effect isSTRESS RELAXATION AFTER LARGE STEP STRAIN 239
©)
Fig. 7.8. Release and creation of the topological constraints. (a) The topological
constraints imposed on the chain A by C is released and recreated by the motion
of C. (6) In the two-dimensional representation, this process can be represented
by the disappearance and reappearance of the obstacle C. The process causes the
deformation of the tube in the middle.
still at the level of conjecture, and there are other theoretical
treatments. Experimentally, in linear polymers with narrow molecular
weight distribution, it seems that the major difficulty of the theory can be
Tesolved by including the contour length fluctuation. On the other hand
the tube reorganization is believed to be important for polymers with
broader molecular weight distribution,” or long branches, which will be
discussed later.
7.5 Stress relaxation after large step strain
7.5.1 Experimental setup
Having seen the characteristic features of the linear viscoelasticity, we
shall now study the nonlinear viscoelasticity. Before studying the general
situation, we shall first consider a simple case, the stress relaxation after
Stepwise deformation.“ Suppose that at time ¢=0, a polymeric liquid
is suddenly deformed homogeneously. The deformation creates a stress
which gradually relaxes with time. Our problem is to find how this
telaxation takes place.
For a homogeneous deformation, we may assume without loss of240 MOLECULAR THEORY OF POLYMERIC LIQUIDS
fa) (b)
Fig. 7.9. (a) Shear and (6) elongation.
generality that a point r in the material is displaced to
ror=E-r. (7.75)
The tensor E is called the deformation gradient.t Two particular cases
are often studied experimentally.
(i) Shear (Fig. 7.92), for which the material is deformed to
Tea het My Ty Hy TE Te (7.76)
This deformation is characterized by a single parameter y, the shear
strain. Since the deformation has a reflection symmetry with respect to
the xy plane, the stress components 0,,(=0,,) and o,,(=0,,) vanish
identically. However, the shear stress o,, = 0,, and the diagonal stresses
Ox, yy and o,, generally do not vanish. Since the isotropic part of the
stress has no significance, only two of the diagonal components have
meaning. The stresses
N, = Ox — Oyy (7.77)
and
Ny = Oy — 02, (7.78)
are called the first and the second normal stress differences. Thus the
tesponse of the shear deformation is characterized by three stress
components 0,,, N, and N;, each of which are nonlinear functions of y
and t.
(ii) Uniaxial elongation (Fig. 7.9b). Here the sample is stretched in the
+ A general deformation is described by the function r’ = r’(r) which connects the position
vectors r and r’ before and after the deformation. In such a case Ez, is given by 4rg/ dr.STRESS RELAXATION AFTER LARGE STEP STRAIN 241
z direction by factor A. Since the volume is unchanged, this causes a
contraction in the x and y directions by a factor 1/VA. Hence, the
deformation is described by
iat rt '
mate WA Tyte ee ar. (7.79)
In this deformation, the off-diagonal components of the stress vanish and
two of the diagonal components o,, and 0,, are equal to each other. Thus
the response is characterized by a component
Or = Oz, — One (7.80)
which is called the tensile stress.
7.5.2 Calculation by Rouse model
First we calculate the stress relaxation using the Rouse model. The stress
tensor (eqn 7.4) is expressed by the normal coordinates X, of the Rouse
model (see Section 4.5.2, eqn (4.137))
ce
Oat = Fy D Ky Xpelt)Xoa(0)) (7.81)
p=1
where
6x7kyT
k= ge? (P=1,2,--.). (7.82)
According to Rouse dynamics, the position of the segments are
changed in the same way as the macroscopic point;} thus if R,(—0) and
R,(+0) are the positions of the segment before and after the
deformation,
R,(+0) = E+ R,(—0) (7.83)
X,(+0) = E+ X,(-0). (7.84)
Since the system is in equilibrium for «<0
(Xpa(—0)Xpa(-0)) = "27 5,,, (7.85)
kp
t This i is often called the affine deformation assumption, but it is actually derived from the
equation for the Rouse model (eqn (4.139)). For an instantancous deformation,
the velocity gradient Kap (t) becomes so large that it dominates the other terms on the right-
hand side of eqn (4. 139) ‘The equation for X, then becomes the same as that for the macroscopic
Point (see eqn (7.151)), and therefore the change of X, becomes affine.w2 MOLECULAR THEORY OF POLYMERIC LIQUIDS
From eqns (7.84) and (7.85), it follows that
(Xpa(-+0)Xp(+0)) = Ea Epe(Xou(—0)Xo(—0)) = Ey Er Spe “at
‘P
= Byg(E) “2 (7.86)
where °
Bap (E) = Eau E py (7.87)
which is called the Finger strain. For t>0, (X,a(t)Xpp(t)) satisfies eqn
(4.141) with Kap =0:
a 2p kp
> (Xpalt)Xpp(t)) = —=— ( (Xpul%o) — Sap}. (7.88)
ot tr k,
Equation (7.88) is solved with the initial condition (7.86) by
kpT
(Xpe(t)Xpp(t)) = 7 [Bap(E)exp(—2pt/tp)
‘P
+ Sap(1—exp(—2p7t/t,))]. (7.89)
Substituting eqn (7.89) into eqn (7.81) and dropping the isotropic term
we have
Gap (t) = kp TBap(E) D exp(—2p"t/ tx) (7.90)
N ie
= Baa{E)G(t) (7.91)
where G(t) is the linear relaxation modulus given by eqn (7.31).
For a shear deformation, E is given by
1 y 0
E=|0 1 0 (7.92)
001
so that
1+y? y 0
B= Y 10
0 o1
and
Ory = YG(t), (7.93)
M=/G), (7.94)
Np =0. (7.95)
Note that the shear stress is a linear function of y even if y is large.STRESS RELAXATION AFTER LARGE STEP STRAIN 243,
For the uniaxial elongation, the tensile stress is given by
or= (2? -7)60. (7.96)
The Rouse-like behaviour is expected to be seen in the initial stage of the
relaxation (t< t,).
7.5.3 Calculation by reptation model
Now we consider the behaviour for t > t, using the reptation model. To
calculate the stress, we have to know
(i) How the conformation R(s, t) of the primitive chain is changed by
the macroscopic deformation, and
(ii) How the stress is calculated for a given conformation R(s, t).
We shall discuss these problems separately.
Expression for the stress tensor. The microscopic expression for the stress
tensor can be obtained by taking the average of eqn (7.4) for a given
conformation of the primitive chain.“ Alternatively, it can be derived by
an elementary argument explained in Fig. 7.10. In both cases the result is
F(t)
Fig. 7.10. Consider a plane of area A, normal to the z axis. The stress
component o,, is given by the force (per area) S,/A acting through the plane.
Consider a part of the primitive chain between s and s + As. If this part is within
the distance u,(s,t)As from the plane, it penetrates the plane and gives a
contribution of F,(s, t) = F(t)u.(s, t) to S,, where F(t) is the tensile force acting
along the primitive chain. Since the number of the primitive chain in unit volume
is (c/N), the number of such part is (c/N)Au,(s, t)As, whence
s.=5 (ZAu(s, )ask(s, )= Lal [arouc, outs. ).
Thus the force per area S,/A gives eqn (7.97).244 MOLECULAR THEORY OF POLYMERIC LIQUIDS
written as
L
Oup(t) =<( | dsF(0)ua(s, t)up(s, 0) (7.97)
0
where u(s, t) = AR(s, t)/s is the unit vector tangent to the primitive
chain and F(t) is the tensile force acting along the primitive chain. In the
equilibrium state F(t) is given by (see eqn (6.83))
3kaT ;
aE L, (7.98)
while in the non-equilibrium state in which the contour length is L(t),
F(t) is given byt
F(t)=
F(t)=
ae Th). (7.99)
From eqns (7.97) and (7.99)
)
apt) = SE (fare 1p (6s #)—38ap)). (7-10038
This formula shows that the stress is determined by two quantities, the
contour length L(t) and the orientation of u(s, t).
+ Here it is assumed that the tensile force is constant along the chain. This assumption is not
correct immediately after the deformation because initially the Rouse segments are
stretched or compressed along the primitive path depending on their direction (see Fig.
7.11). However, such local imbalance in the segment density is adjusted in the time t,, and
for t>t,, the tensile force F(s, t) can be regarded as independent of s.
+ Curtiss and Bird™ derived a slightly different stress formula, which includes an adjustable
parameter called the link tension coefficient «. However, this formula is not consistent with
the stress optical law unless ¢=0. In the case of ¢=0, the formula becomes essentially
equivalent to eqn (7.120) which is a special case of eqn (7.100).
Equation (7.100) is also derived from the principle of virtual work. The free energy per
unit volume is
= £ 3keT(LO")
a
Under a virtual deformation 6€,,, L(t) changes by
L
oL= J ds5eqgua(S, t)ug(s, t)
°
whence
La)
dat = PERF (LOL) = = beag £3kat (Le) | dsu,(s, t)up(s, »)
Oo
which gives eqn (7.100).STRESS RELAXATION AFTER LARGE STEP STRAIN 25
Deformation of the primitive path. Our next question is how the
conformation R(s, t) of the primitive chain is deformed by the macro-
scopic strain. The simplest assumption is that the deformation is affine,
ie., that the primitive chain (or the central axis of the tube) is deformed
in the same way as the macroscopic deformation. Thus the point
R(s, —0) on the primitive chain is displaced as
R(s, -0) > E R(s, -0). (7.101)
Let us now study how this changes the contour length L(t) and the
orientation of u(s, t). This is illustrated in Fig. 7.11.
(i) The change of the contour length: The transformation (7.101)
changes the length As of a line segment on the primitive chain to
As = As |E + u(s, —0)|. (7.102)
The length AS can be larger or smaller than the original length As
depending on u(s, —0). Since the distribution of u(s, —0) is isotropic, the
average ratio between AS and As is given by
a(E) = (|E-ul)o (7.103)
a(e)L
mt
0
Fig. 7.11. Deformation of the primitive chain by macroscopic strain. Here for the
Purpose of explanation, the primitive chain is represented by randomly connected
ime segments. According to the affine deformation assumption, a line segment
Tepresented by r is transformed to E + r. Thus the length As = |r| and the direction
u=r/|r| are changed as As—>|E + u| As and u— E- u/|E- ul.246 MOLECULAR THEORY OF POLYMERIC LIQUIDS
where {...)o denotes the average of u over the isotropic state,
du
seslo=] ares 104)
(.do= | S (7.104)
For any deformation which conserves the volume (det |E|= 1), it can be
shown that’?
a(E)>1. (7.105)
Thus the contour length of the primitive chain immediately after the
deformation is given by
(L(+0)) = (EL (7.106)
which is always larger than L.
(ii) The change of the orientation: To denote the orientation of the
primitive path, we define the orientational tensor
Sap (S, t) = (ua(s, t)ug(s, t) — 356) (7.107)
which vanishes in the isotropic state, but does not vanish in the oriented
state. Since the distribution of u(s, —0) is independent of s, S,g(s, +0)
will also be independent of s and can be written as
Sup(s, +0) = Qup(E). (7.108)
To calculate Q,(£), let us consider the probability distribution function
f(u, s, t) for the tangent vector u(s, t). (f(u, s, t) is the probability that
the tangent vector at s and t is in the direction u.) Obviously
1
f(u, s, -0) "ke (7.109)
By the deformation, the unit vector u changes as
._ E-u
w= TE (7.110)
The probability that an arbitrarily chosen point of the deformed primitive
chain is in the direction of u’ = E - u/|E - ul is proportional to [E - u|, the
length of such a part. Thus
flu’, s, +0)= cf du |E + ul 3(w Ea s,-0) (7.111)
where C is the normalization constant which is determined from the
condition
1=C | du'f(u’, s, t) (7.112)STRESS RELAXATION AFTER LARGE STEP STRAIN 247
to be
Ct=(4x) | du [E+ ul = (|E-ul)o. (7.113)
Then
flu, s +0 = [du |e -w'| 6(u- =~") (7.114)
” 4n(|E-ul)o |E-'|
so that
Qne(E)= { dutagflin, 5, +0)~ 3806
= (EEO) Eula Wops (715)
Stress relaxation. Now it is easy to calculate the stress relaxation.
According to the model described in Section 6.4, the relaxation of L(t)
occurs on the time-scale of tz, while that of orientation occurs on the
time-scale tz. Thus the stress relaxation for t > t, occurs in two steps (see
Fig. 7.12).
(i) Contour length relaxation: In the time-scale of tz, S.g(s, t) can be
regarded as equal to the initial value Q,,(E), so that eqn (7.100) is
(@) (b) (o) @
Fig. 7.12. Explanation of the stress relaxation after large step strain. (2) Before
deformation the conformation of the primitive chain is in equilibrium (¢ = —0).
(6) Immediately after deformation, the primitive chain is in the affinely deformed
conformation (t= +0). (c) After time tz, the primitive chain contracts along the
tube and recovers the equilibrium contour length (t= tz). (d) After the time t,,
the primitive chain leaves the deformed tube by reptation (f= 1,). The oblique
lines indicates the deformed part of the tube. Reproduced from ref. 107.28 MOLECULAR THEORY OF POLYMERIC LIQUIDS
written as
Oop = EBT (1(0))*Qap(E) for e tp, L(t) is at the equilibrium value L so
that eqn (7.100) is written as
£
€ 3kT -
apt) = Rts L| ds (uals, tps) —4800)
0
£
1
=65 J dsSz(s, 1). (7.120)
Now S,g(s, t) is equal to Q,.6(E) if the primitive chain segment s in the
deformed tube, and is zero if it has left the tube. Since the probability
that the primitive chain segment s is in the deformed tube is 1(s, t) (see
Section 6.2.2)
Saa(s, t) = Qup(E)Y(s, t). (7.121)
Hence
op) = G.Qan(E)z | d5¥65, 0)
=GQu(E\Y(t) > tr. (7.122)
Here y(t) is given by eqn (7.40). Combining eqns (7.118) and (7.122) we
finally have
Oap(t) = G.Qap(E)(1 + (a(E) — lexp(-t/tp))° YO) (> 4,). (7.123)STRESS RELAXATION AFTER LARGE STEP STRAIN 249
In the case of a shear deformation, we may write a(E) and Q.,(E) as
a(y) and Q.9(y), respectively. For small y, Q,,(y) and a(y) are easily
calculated.
a(y) = (1+ 2yuguy + y7u5)'?)o
= (14 yuu, — dy?uzuy + 4y7u5)o=1+ Ky? + O(y*) — (7.124)
and
a (__ et ry
Qo) = a(y) (a + 2yu,uy + aay),
= (uy — uus)o + O(Y°) = By + O(Y’). (7.125)
Thus eqn (7.123) becomes
Oxy (t) = YG. WY) + O(y’). (7.126)
Hence, the relaxation modulus in the linear viscoelasticity is given by
GO =8G.y(). (7.127)
This determines the plateau modulus to be
2
GP =48G, =f Skat (7.128)
which gives eqn (7.54).
We shall now compare the results for large strain with experimental
results.
7.5.4 Comparison with experimental results
Shear, Extensive experiments on the stress relaxation for shear deforma-
tion have been done by Osaki et al. For convenience of comparison,
we shall represent the relaxation of the shear stress by the nonlinear
relaxation modulus defined by
1
Gi y) = 7 alt y)- (7.129)
In the limit of y—>0, this reduces to the relaxation modulus of linear
viscoelasticity.
Equation (7.123) gives
Gt 1) = G, 22D + (aC) ~ 1pexp(—ra WO
=A(y)GO(1 + (ey) — Yexp(—t/ta)? (7.130)Git. WG,
10? 107" 10° 10
the
Fig. 7.13. Theoretical curve of the nonlinear relaxation modulus G(r, y). The
case of T,/Tp = 100 is shown. Reproduced from ref. 64.
10!
G(ty) (Pa)
10°
107
10° 10! 10° 10°
t(sec)
Fig. 7.14. Nonlinear relaxation modulus G(t, y) for polystyrene solution of
chlorinated biphenyl at 30°C. The molecular weight of the polymer is 8.42 x 10°
and the concentration is 0.06 g/cm. Magnitudes of shear y are <0.57, 1.25, 2.06,
3.04, 4.0, 5.3, and 6.1, from top to bottom. Reproduced from ref. 69.STRESS RELAXATION AFTER LARGE STEP STRAIN 251
where G(t) is the relaxation modulus of linear viscoelasticity and
(7) = Q,,(y)/ Gy). (7.131)
Equation (7.130) is plotted in Fig. 7.13. For small y, G(t, y) decays
roughly in a single exponential manner with relaxation time t,. For large
y, Gt y) shows another relaxation characterized by tz corresponding to
the relaxation of the contour length. Such behaviour has actually been
observed experimentally® as shown in Fig. 7.14. Detailed comparison
teveals good agreement between the theory and experiment.
(i) Osaki et al® found that at large t, curves in Fig. 7.14 for various y
can be superimposed by a vertical shift (see Fig. 7.15). This implies that,
for large t, G(t, y) can be written as a product of two functions, one
depending on time and the other on strain. This agrees with eqn (7.130),
which is written for t > tz as
GO y) =hA()GO. (7.132)
The function A(y), called the damping function, is found to be
independent of the molecular weight and concentration over a wide
range.7° Figure 7.16 shows the comparison between the theoretical
damping function and the experimental one. The agreement is very good
considering that h(y) includes no adjustable parameters.
10°
107
Gn'Pa)
10!
10° 10" 10° 10°
t (sec)
Fig. 7.15. Reduced relaxation modulus G(s, y)/h(y) derived from Fig. 7.14.
Each curve for y > 1.25 in Fig. 7.14 is shifted vertically by an amount —log h(y)
so that it superposes on the top curve in the long time region. t? indicates the
longest relaxation time, and t, the characteristic time below which the superposi-
tion is not possible. Reproduced from ref. 69.252 MOLECULAR THEORY OF POLYMERIC LIQUIDS
AY)
10"
107!
10" 10 10"
y
Fig. 7.16. h() determined from the procedure explained in Fig. 7.15. Filled
circles represent polystyrene of molecular weight 8.42 x 10° and the unfilled
circles of 4.48 x 10°. Directions of pips indicate concentrations which range from
0.02 gcm’ to 0.08gcm*. The solid curve represents the theoretical value (eqn
(7.131)), and the dashed curve the result of the independent alignment
approximation (eqn 7.187). Reproduced from ref. 69.
(ii) Experimentally, the first relaxation can be characterized by the
time 1, below which the factorization of G(t, y) is not possible (see Fig.
7.15). Osaki et al. found that the ratio between t, and the Rouse
relaxation time tg is about 4.5 and essentially independent of the
molecular weight and concentration.
(iii) The relaxation of the other stress components N,(t, 7) and
Nt, y), measured by birefringence, have precisely the same time
dependence as g,,(f, y), and their ratio depends only on y. This agrees
with eqn (7.123), according to which
MG y= Galt) rc Ooty) (7.133)
(7)
Q»(y) — Q.2(Y)
NXt, Y) = O.(y) O,,(t, Y). (7.134)
Since it can be proved™* that
Qux(¥) — Qy(Y) = YOxy(Y), (7.135)STRESS RELAXATION AFTER LARGE STEP STRAIN 253
107
—Nelt,yV/Nilty)
10%
10° 10° 10°
y
Fig. 7.17. Quantity —N,/N, is plotted against magnitude of shear y. Sample:
polystyrene solution in chlorinated biphenyl (M,, = 6.7 x 10°, p =0.40gcm™°).
The number of entanglements Z corresponds to about 14. The solid line is the
theoretical value, —(Q,,(y) — Q..(7))/(Qux(7) — Q,,(y)). The dashed line is the
Tesult of the independent alignment approximation. Reproduced from ref. 67.
eqn (7.133) is written as
MG Y) = 78n(t Y)- (7.136)
This relation, first found by Lodge and Meissner”! using a phenomenolo-
gical argument, has been well confirmed.’ The ratio between the
second normal stress difference N,(t,y) and the first normal stress
difference N,(t, y) is shown in Fig. 7.17. The experimental values are
again in reasonable agreement with the theory.
Uniaxial elongation. For uniaxial elongation, the tensile stress is given by
Or (t, 4) =[1 + (@(A) — Dexp(—t/te) PF(A)G (0). (7.137)
Here @(A) denotes a(E) of uniaxial elongation and
f@)=4(@..A) - Qn). (7.138)
Explicit formulae for w(A) and f(A) can be calculated analytically,”
a(A) =44(1 + A(A)) (7.139)
and
_ 15(43+1/2) 1 4-1
fO=TGi-y Ts4@) (1-F74@) 140)254 MOLECULAR THEORY OF POLYMERIC LIQUIDS
where
sinh™[(a? — 1)"7]
eae— HP?
The stress relaxation for uniaxial elongation has been studied by Ferry
et al.,” and the result has been well fitted by eqn (7.137) if the effect of
the molecular distribution is taken into account.”
A(A) = (7.141)
7.5.5 Discussion
As we have seen the theory has predicted many aspects of the nonlinear
stress relaxation. However, there are some experimental results which
are not in accordance with the theory and need some discussion.
Anomalous stress relaxation in shear flow. Osaki et al.® found that the
nonlinear relaxation modulus G(f, y) of polystyrene solutions does not
agree with the theory for very-high-molecular-weight samples for which
Mp > 10° g/cm?. (7.142)
For these samples, the stress near t, decreases much more steeply than
predicted by the theory, and shows complex M, p dependence. A similar
anomaly is also reported by Vrentas ef al.” This result is puzzling since
the theory should become most valid in the high molecular weight limit.
A possible explanation, however, was given by Marrucci and Grizzuti,”*
who pointed out that the theoretical damping function h(y) has a region
where the differential rigidity modulus d(yh(y))/dy is negative. In this
tegion, the elastic energy can be lowered by microscopic phase separa-
tion, each phase having different local shear strains. Indeed the observed
anomalous behaviour can be reproduced with suitable assumptions.”>”°
Experimental evidence of such microphase separation is, however,
lacking and further study is expected.
Tube reorganization. The theory described in Section 7.5 includes two
essential assumptions; (i) the conformation of the tube remains un-
changed and (ii) the contour length of the tube returns to the equilibrium
value L even if the environment is not in equilibrium. The validity of
those assumptions is not established and it is worthwhile to study the
consequences of the theory based on other assumptions.
Marrucci et al.”7’ assumed that the volume of the tube remains
constant by deformation, and derived a result which has the same time
dependence as eqn (7.122) but different strain dependence. Though an
experiment on PMMA” seems to fit with the modified formula, caution
is needed in accepting the modification since critical experiments needNONLINEAR VISCOELASTICITY 255
data over a large time-scale and for a monodisperse sample, but the
quoted experiment does not meet these conditions.
Viovy et al.” argued that as the contour length of the surrounding
primitive chain contracts, there will be an extra relaxation by the release
of the topological constraints. They proposed a theory which gives a
slightly different relaxation behaviour for ¢> M,. In
practice, the condition (7.169) is not always satisfied, and the elongation
of the contour length can be important, but this will not be considered
here.
For the sake of simplicity, we shall use a slightly different notation in
this and the following two sections: the equilibrium contour length will be
denoted by L (because L and L need not be distinguished for the
inextensible model), and the segments of the primitive chain are labelled
from —L/2 to L/2. (Thus the segment 0 corresponds to the middle of the
chain.)
7.7.1 Deformation of the primitive chain
First we express the transformation rule of the inextensible primitive
chain in mathematical terms. Let R(s) and R(s) be the conformations of
the primitive chain before and after the deformation. The transformation
tule is explained in Fig. 7.21, i.e.,
(a) The segment in the middle changes its position affinely, i.e.,
RO) =E-R(0). (7.170)
(b) The segment § lies on the curve E - R(s), so that
R(5) = E-R(s) (7.171)
Fig. 7.21. The deformation of an inextensible primitive chain by a macroscopic
strain. The new conformation A’O’B’ is on the curve A"O”B”, which is the affine
transformation of AOB. The new position of a segment, say C, is determined
from the condition that the contour length O'C' is equal to OC, where O’ is the
affine transformation of O and coincides with O”. Reproduced from ref. 108.262 MOLECULAR THEORY OF POLYMERIC LIQUIDS
where § is the contour length along the curve E- R(s’) from s’=0 to
s'=S, ie.,
s= fas’ |E-u(s’)|. (7.17)
From eqn (7.171), the transformation rule for the direction u(s) is
obtained as
a) =A) =-2e- RG) =e.
a(S) = gph O =57F R(s) aps © R(s) (7.173)
os
=F -u(s). (7.174)
Equation (7.172) gives
1 -Se - u(s)}. (7.175)
From eqns (7.174) and (7.175) it follows that
E-u(s)
|E-u(s)|"
This is equivalent to the transformation given in Fig. 7.11.
a6) = (7.176)
7.7.2 Independent alignment approximation
According to eqn (7.172) s and § are not equal to each other. This leads
to a constitutive equation of a rather complicated form (see Section 7.9).
If we disregard the difference between s and § and assume the following
transformation rule
E-u(s)
Ewe)
(7.177)
the constitutive equation is obtained in a simple form. This prompts the
study of the approximation (7.177), which we call the independent
alignment approximation (IA approximation).'*!5 Though the physical
justification of this approximation is not clear, its error usually turns out
to be small except for a few cases which will be discussed later. Therefore
we shall first proceed using this approximation.
According to eqn (7.100) the stress for the inextensible model is givenCONSTITUTIVE EQUATION FOR REPTATION MODEL 263
by
c 3kpTL’ 1
NW NB? if ds (ua(s, t)up(s, t) — 4506)
Capt) =
L2
=G.z | as(uals, Qupls, 1) —45ap)
-L2
L2
1
Gr , J dsSap(5, t)- (7.178)
ay
To calculate S,,(s,t) we need to obtain the probability distribution
function f(u,s,t) that the tangent vector at the segment s is in the
jirection u at time t. The time evolution equation for f(u, s, t) is obtained
n the same way as in Section 6.3.
Suppose that in the time interval between ¢ and ++ At, the chain
egment s moves to the position at which the chain segment s + A& was
ocated at time ¢, then u(s, t + Af) is given by
E(t + At, t)- u(s + AG, t)
ms, 6+ At) = JE(e+ At, t)- u(s + AE, 1)"
(7.179)
since the distribution function of u(s+A&,t) and A& are given by
"u, s + AE, t) and W(A8&), respectively (see eqn (6.24)), the distribution
function of u(s, t + At) is given by
_ re(,, Et At teu"
f(u,s, 1+ At)= faAew(as) | du 5(w Bera al a)
x f(u', s + AE, t). (7.180)
To assess the accuracy of the IA approximation, let us consider the
stepwise deformation E imposed at t=0. In this case the orientational
distribution before the deformation is
fu, s, -0) = = (7.181)
Since (A) becomes 6(A&) for an infinitesimally small time-interval At,
eqn (7.180) gives
fu, s, +0) = ia a(u - Eni) (7.182)264 MOLECULAR THEORY OF POLYMERIC LIQUIDS
Hence S,¢(s, +0) is given by
Sag(s, +0) = | du(ugity —48ap)f(us 8, +0)
du’ E-u'
= [eu Sy (tetia~480p)8(u 7g 7)
du (E+ u)a(E + u)p
4a |E- ul?
(E-u).(E + u),
«era,
= OME). (7.183)
On the other hand the correct value of S.g(s, +0) is Q.g(E) given by eqn
(7.115). Thus in the case of stepwise deformation, the LA approximation
amounts to a decoupling approximation
—4ap
(E-u)a(E-u)\ 1 _ (E+ u)a(E-u)p
( E-ul ) de-anc=4 E-u? ). (7-184)
It can be seen that the error of this approximation will not be large for
any form of E. Indeed for the case of shear deformation, the IA
approximation gives
OM)=3y «KD, (7.185)
OS) - OY) = BY « 1). (7.186)
The damping function is thus given by
AA” y) = OSM(y) /' (2). (7.187)
This is shown by the dashed line in Fig. 7.16. It is seen that the error of
the IA \ approximation is small over a wide range of y. A useful formula
for Q%9(E) for general E is given in ref. 87.
7.7.3 Constitutive equation
To obtain f(u, s,¢) in the general case, we rewrite eqn (7.180) in a
differential form. For small At, E(t + At, t) is written as 1+ «(t)At.
Hence
E(t+At,th-u ut K(t)-udt
|E(@ + Az, t)- val |u + «(0) - wAt|
=u + (K(t)+u — (uu: K(t))u)At=u+ AM(u, t) (7.188)
T(u, t) = K(@) + u — (uu: K(t))u. (7.189)
whereCONSTITUTIVE EQUATION FOR REPTATION MODEL 265
Thus if we neglect the terms of order (Af)*, eqn (7.180) is rewritten as
flu s, + At) = | dAEW(AE) du’d(u—u' — AMT(w’, )f(u', s+ AE, 1)
= | angw(agy(1 - are Ty, )ftw, s+A6&,0)
2
= | dagw(agy(1 + age +4(AE)? S)(1 - are Tu, 0) fw st)
=(1 + AID, 25 z )(a - are E(u, 0) ) fl 50). (7.190)
Comparing the terms of order At, we have
a # a
3 5,0 =D, gpfu 5, t)- aa" T(u, Of(u,s,t). (7.191)
The boundary condition is that the tangent vector at the chain end is
isotropic,
1
_ = . A
f(u,s, Dare at s=+L/2. (7.192)
Equation (7.191) can be rigorously solved to give’®
t
= [{ar(2 wo, 1-19) { (py - FOO
flu s, 1) J at ( <.w,1-1) | a 5(u oP 1) (7.193)
where p(s, t) is given by eqn (6.14). This solution could also have been
arrived at by physical argument. Suppose that a tube segment is created
in the direction u' at either of the chain ends between time ¢’ and ¢’ + dr’.
If this still survives at time ¢ and is now occupied by the primitive chain
segment s, u(s, t) must be E(t, t’)-u’/|E(t, t')-u’| (this is the result of
the IA approximation). Since the probability that this happens is
pe t—t’')/at'] dt’ and the distribution of u’ is 1/4a, we get eqn
(7.193).
From eqn (7.193), the orientation of the primitive chain segment is
given by
Sag(s, t) = J a(S w(s, t- 1) f auf
E(,t’)-u'
X (Ugg — 4809)(u ieeryel I)
~ fa (ye Viotot OPEC, r"). (7.194)266 MOLECULAR THEORY OF POLYMERIC LIQUIDS
Substituting this into eqn (7.178), we finally have
Cap(t) = Ge | ar(2 wt -1)) OPEC, r) (7.195)
where p(t) is given by eqn (7.40). Equation (7.195) is the constitutive
equation which comes out of the IA approximation. We shall now
compare this equation with experimental results.
7.7.4 Comparison with experiments
General feature. Equation (7.195) can be written in a more convenient
form. Consider that a stepwise strain E is applied at time 0, then E(¢, t')
is given by
— ift>Oand?'<0,
t, t’)= .
Fe) ti otherwise. (7.196)
Let Pap(t, E) be the stress caused by this deformation at a positive time ft.
Equations (7.195) and (7.196) give
beplt 5-6. a(S ve-) ogre)
= Gy (NOvp(E). (7.197)
This is very similar to eqn (7.122) except that Q,(E) is now replaced by
Q%(E). In fact eqn (7.197) can be derived much more simply by the
reasoning given in Section 7.5.3, if it is noted that S,,(s, +0) is given by
Q%#)(E) in the independent alignment approximation.
Using p(t, E), the stress for an arbitrary flow history is given by
t
Oxp(t) = | a(S bap (t—t', 6) . (7.198)
2, ot E=£(,"')
Equation (7.198) agrees with the empirical equation proposed by
Bernstein, Kearseley, and Zapas (BKZ),* who found that the stress
response for various flow histories can be predicted by eqn (7.198) using
the stress relaxation function @,g(t, E) determined experimentally.
Subsequent experiments done by many authors®-** revealed that the
BKZ equation is one of the most successful empirical constitutive
equations.
As was shown in Section 7.5, the empirical stress relaxation function
gap(t, E) is in good agreement with the reptation theory for linear
polymers of narrow molecular weight distribution. This, together with the
success of the BKZ equation, indicates that the constitutive equationCONSTITUTIVE EQUATION FOR REPTATION MODEL 267
derived by the reptation theory works well for general flow histories.
Indeed eqn (7.198) reproduces many characteristic features of the
nonlinear viscoelasticity. Leaving detailed comparison to refs 51 and
94-97, we shall study the main features briefly.
Nonlinear viscoelastic behaviour. To see the characteristic features of the
constitutive equation (7.195), we approximate w(t) by
vo~{
1 for tty. (7.199)
Then eqn (7.195) gives
Fup ~ G-OT~(E(t, t—ta)) (7.200)
which simply says that the stress is given by the elastic deformation
caused between ¢ — ty and ¢.
(i) Steady shear flow. In the steady shear flow, eqn (7.200) gives the
shear stress
G,y(K) = G.QGM(Kt4) ~ Gxt gh (Kr4). (7.201)
Thus the viscosity becomes
n(x) = 10h (Kt,). (7.202)
Equation (7.202) indicates the shear thinning occurs at kK ~1/t,, which
becomes very small for large molecules. This explains why the nonlinear
response is important in polymeric liquids.
The first normal stress coefficient is also estimated as
W(k) = G_t7h (Kt) (7.203)
which again decreases with the shear rate as shown in Fig. 7.18. The
second normal stress coefficient
W(x) = —G.tah(Kta),
where
WS") =~ (OM) OY), (7.204)
is negative and again decreases with x, in agreement with experiments.
The ratio W,(0)/'¥,(0) is precisely evaluated as™*+
20) _ O97) - OY)
Hi) OL) - OF(y) 1-0
The experimental value is between —0.1 and —0.3.%°-"
tIf the IA approximation is not used, the precise value of W,(0)/¥,(0) becomes
-1/7=-0.14.
=-3=-~0.3. (7.205)268 MOLECULAR THEORY OF POLYMERIC LIQUIDS
(ii) Stress growth in shear flow. When a shear flow is started with
constant shear rate x at time ¢=0, eqn (7.200) gives
1
GK) Gt t— Tah tt). (7.206)
Since
Kt for tt,, (7.207)
therefore
GthM(kt) for tty. ( )
Since yh(y) has a maximum at y=2, n*(¢;«) shows a maximum at
t=2/Kk provided Kt, 2. The height of the maximum decreases as 1/x.
These features are in agreement with the experimental results shown in
Fig. 7.19.
(iii) Steady elongational flow. The steady elongational viscosity is given
as
ne(e) =~ (OM(Er.) - OW(Et) (7.208)
which first increases slightly with é and then decreases with é. This was in
contradiction with earlier data® for low density polyethylene, which
indicated a sharp rise of nz(é), but recent data for monodisperse linear
polymers®*™ are consistent with the theory.
7.7.5 Discussion
Though the theoretical constitutive equation (7.195) explains many
features of nonlinear viscoelasticity, there are some discrepancies which
are worth discussing.
Steady shear flow. The predicted steady state viscosity n(x) depends on
the shear rate x too strongly. In fact eqn (7.195) predicts that at high
shear rate of Kt, >>1™*
n(x) =~ n(O)(«r,) >? (7.210)
i.e., the shear stress o,,(K) = (k)k decreases with increasing shear rate,
which means that the shear flow is not stable at high shear rate. At first
sight this conclusion may seem to contradict the many experiments whichCONSTITUTIVE EQUATION FOR REPTATION MODEL 269
show stable shear flow up to very high shear rate. However, the
theoretical prediction is not entirely ruled out for various reasons:
(i) Equation (7.195) indicates that the form of (x) is sensitive to the
relaxation spectra of the linear relaxation modulus G(t): the broader the
relaxation spectra is, the smaller the exponent x in n(x) «x~* becomes.
If the sample has broader molecular weight distribution, the relaxation
spectra of G(t) becomes broad and the anomalous behaviour of n(x)
disappears. Also, even for the monodisperse sample, the various relaxa-
tion processes discussed in Section 7.4.2 broaden the relaxation spectra
and weaken the shear rate dependence of n(x).
(ii) Equation (7.210) is derived under the condition
WtyKK K 1/tr. (7.211)
If Ktg becomes of the order of unity, the contour length L(x) increases
with the shear rate, and the stress starts to increase according to
Oy & L(K)* (7.212)
(see eqn (7.100)). Therefore if ty/tg=M/M, is not sufficiently large,
which is the case in many experiments with monodisperse samples, the
minimum of g,,(x) will not be observed and the flow will be stable.
(iii) On the other hand if the system is monodisperse and if M/M, is
large enough, the theory predicts the shear stress shown in Fig. 7.22.
Such behaviour has indeed been proposed by Vinogradov and by Ball
and McLeish,” to interpret the finding that the flow rate of polymers
through pipes changes abruptly as the shear rate is raised if the molecular
weight of the polymer is high and has a narrow distribution. Though the
log oy, (x)
Ke = Vi
log x
Fig. 7.22. Shear stress predicted by the theory for monodisperse systems with
M/M, large.270 MOLECULAR THEORY OF POLYMERIC LIQUIDS
detailed analysis has not been done, it should be worthwhile to study the
phenomenon under well-controlled conditions.
Stress overshoot. According to eqn (7.195), the stress maximum at the
start of the shear flow appears in the shear stress, but not in the first
normal stress difference N,(t, K) = Oz,(t; K) — Gyy(t; K),* whilst experi-
mentally the maximum is often observed in N,(t;K). This is possibly
due to the elongation of the contour length. Indeed the overshoot in the
normal stress appears at a higher shear rate than in the shear stress."
That the elongation of the contour length is important under usual flow
conditions is indicated by the stress relaxation after the steady shear flow.
It has been observed™ that when the shear rate becomes larger than
1/tg, the relaxation curves begin to show a short-time component which
corresponds to the relaxation in the contour length.
7.8 Stress relaxation after double step strain
Though the BKZ-type constitutive equation has been quite successful in
many phenomena, it has been reported that under certain flow history,
the equation gives unsatisfactory predictions. One such experiment is the
stress relaxation after application of double step strain.'*" The flow
history of this experiment is illustrated in Fig. 7.23.
Two step shears y, and y, are applied with time interval ¢,, one at time
—t, and the other at time 0. The BKZ equation (7.198) predicts
Oap(t) = Paplt, Y2) + Paplt +t, Y2t 1) ~ Paplt +t, Y2)- (7.213)
wt)
Fig. 7.23. Double step strain experiments.STRESS RELAXATION AFTER DOUBLE STEP STRAIN 271
According to Osaki et al.,%° eqn (7.213) predicts the stress with
reasonable accuracy when 7,72 > 0, i.e, when the sense of the two shears
are the same, while when y,y,<0, a large discrepancy is found. It has
been shown’ that this discrepancy is caused by the IA approximation
and the rigorous analysis of the model gives good agreement with
experiments.
We consider the inextensible chain model. Figure 7.24 explains the
change of polymer conformation under the double step strain. Figure
7.24a shows the undeformed state just before the first deformation.
Figure 7.24 represents the state immediately after the deformation: the
primitive chain is deformed by the shear y;. Figure 7.24c indicates the
state just before the second deformation; the inner part AB still remains
in the deformed tube, while the outer parts are in the undeformed tube.
Now when the second deformation is applied, the inner part AB is
deformed by the shear y, + y2 from the equilibrium state, while the outer
part is deformed by the shear 2. It is important to note that the second
shear stretches the contour length of the outer part by the factor a(y2),
but that of the inner part by the factor
B= (yi + Y2)/a(y), (7.214)
since the inner part is already stretched by the factor a(y,). If the
coordinates of A and B are s, and sz, respectively, the coordinates of A’
and B’ are fs, and Bs2. Hence the probability that a primitive chain
segment s is between A’B’ is equal to the probability that it is between
time
@) -t-0
o -.. LY
[jo Zn BR Bn
Fig. 7.24. Microscopic process in the relaxation of double step strain. The
deformation of the primitive chain at various times are shown. The deformation
of the primitive chain relative to the equilibrium state is shown by oblique lines.272 MOLECULAR THEORY OF POLYMERIC LIQUIDS
AB at time t= —0, and is given by (s/f, t,). Hence the average of
Sap(s, +0) is given byt
p(s/B, t1)Qap (ys + ¥2) + [1 — (s/B, 4,)]Qap(y2)
Sup(s, +0) = |s/B| < L/2,
als, +0) {onus \sie|>Lr2. ©7219)
For t>0, S,p(s, £) satisfies
a e
9 500 (5) 1) = Dea Seals, 6) (7.216)
and the boundary condition
Sap(s, t)=O at s= +e. (7.217)
Hence S,9(s, ¢) is given by
12
Sap(s, t) = | ds’G(s, 5’, t)Sap(s', +0) (7.218)
-L2
where
25... (pa *))si pal, )) 2
jak pr & px & —
G(s, s', t) Le sin( L (s+ 2 sin( L (s + 3 jexp(—p7t/ta).
(7.219)
From eqns (7.178), (7.218), and (7.219), the stress at time ¢(t>0) is
given by
Ln
“Le
Oup(t) = Get i ds’ | dsG(s, 5’, t)Sap(s", +0)
~L2
LR
~L2
=G.z [| As’ V6s', DS", +0). (7.220)
-La
Substituting eqn (7.215) we have
Oap(t) = GeQap(Y2) P(t) + GelQaa(¥1 + ¥2) — Qaa(y2)]Y'(t, 4, B)
(7.221)
+Here it is assumed that if |s/8| is larger than L/2, the primitive chain segment s is
deformed by y2. Strictly speaking this is only approximately correct. Actually there is a
small correction term to eqn (7.215),!”” which is neglected here because it is numericallySTRESS RELAXATION AFTER DOUBLE STEP STRAIN 273
where
ne
i [ ave. )y(s/B,t1) for B<1,
VG B=4 Ain (7.222)
1 [ ave. )ws/B,)) for B>1.
L
-in
Using eqn (6.14) and doing the integral, we get
2 _yp+q-aya Sinl(/2)(q — PB)]
Pea yn pq’ - p’B’)
—(p7i 2, al fe 1,
ven Bad “MMO Payed for P<
32 > (—)ote-2y Sin /2)(0 — 4/8)
prod qe" - 7/6?)
x exp[—(p7t+ q7t)/ta] for B>1.
To simplify the equation we consider the case of large t and t,. If t> ty
and t, > ty, only the first term in the sum of eqn (7.223) is important and
'(t, t1, B) is approximated by
SE exol-t +t)/ta] for B<1
v'G ty B= (7.224)
a exp[-(¢+4)/ta] for B>1
or it may be written as
¥'(b tr, B)= AB) VE +4) (7.225)
where
ee for B<1,
A(B) = (7.226)
4 12,
ieee for B>1.
Although eqn (7.225) is obtained under the condition t> t, and t, > ty, it
turns out that eqn (7.225) is actually a good approximation for the entire
regime of ¢ and t,.'7 If eqn (7.225) is used, eqn (7.221) is written as
Fup (t) = Paplts Y2) + ACB) Pop(t + tr» Ya + ¥2)— Paplt +t, ¥2)). (7.227)
Equation (7.227) has been thoroughly checked by Osaki et al. An
example is given in Fig. 7.25. This indicates that the theory described in274 MOLECULAR THEORY OF POLYMERIC LIQUIDS
107|
10°
10' 10°
t (sec)
Fig. 7.25. Shear stresses for double step shear deformation. —y, = y. = 11.6, and
t, is indicated in the figure. Sample polystyrene solution in diethyl phthalate
M=3.10x 10° and p =0.221gcm *. The heavy lines represent stress for single
step deformation. The light solid line represents eqn (7.227) and the light broken
lines the result of the BKZ equation (eqn (7.213)). Reproduced from ref. 68.
Section 7.7 correctly reflects the reality of polymer dynamics in an
entangled state.
7.9 Rigorous constitutive equation for reptation model
Having seen that the IA approximation causes a serious error in certain
situations, we now derive a constitutive equation without using the IA
approximation. "°°
In a small time-interval At, E(t + At, ¢) is given by
E(t + At, t)=1+ «(t)At. (7.228)
Thus the transformation rule described by eqn (7.171) is written as
R(, t+ Ad) =R(s, t) + K(t)- R(s, DAt=R(s, t) + x(t) RG, At
(7.229)CONSTITUTIVE EQUATION FOR REPTATION MODEL 275
The second equality holds since s — § is of order At. Similarly eqn (7.172)
becomes s
5= [av |a(s’, t) + x(t) + u(s', t)At|
oO
| ds'(1 + «(t):u(s', t)u(s’, t)At) + O(AP)
=stAt | ds’x(t):u(s', u(s', 2). (7.230)
0
Therefore, to the order of At, s is expressed by § as
&
s=5— At} ds'x(t):u(s’, t)u(s', t)
0
=5— At&(, t) (7.231)
where
&s,)= | ds'x(t):u(s’, t)u(s’, t). (7.232)
0
From eqns (7.229) and (7.230), the change in the tangent vector u(s, )
becomes
u(5, t+ At) ~2RG, t+ At)= 2 (R(s, t) + K(f) RG, t)At)
os 3 3
= peas RO t+K(o- ag. tat
=(1- ars & 0))ms, t)+x(t)-u(% NAL — (7.233)
Using eqn (7.232),
u(S, t + At) =u(s, t) —[K(t): a(S, Ou, tu(s, 0) — x(t) - a, O)At
=u(s, t) —[a(t):u(5, Dus, Du, t)— KQ)- u(G, At + O(A)
=u(s, t) +1 (u(5, 1), At (7.234)
where I'(u, t) is given by eqn (7.189)
In eqn (7.234) the effect of Brownian motion was not taken into
account. If this is included, the final equation becomes
u(S, t+ At)=u(s + AE, t)+T(u(s, 1), OAt
=u(5 — £5, At + AE, )+T(u(s, 1), At, (7.235)276 MOLECULAR THEORY OF POLYMERIC LIQUIDS
or replacing 5 by s
u(s, t+ At)=u(s — E(s, t)At+ AE, t)+T(u(s, ), At. (7.236)
This is the time-evolution equation for the tangent vector u(s, t) of the
inextensible primitive chain. Thus the equation for f(u, s, t) is
f(u, s, t+ At) = (d[u —u(s, t+ A2)])
= (6[u —u(s — E(s, t)At + A&(t), t) —T(u(s, t), At)
= (d[u — u(s — E(s, t)At + AE(2), £)])
~ are - (u, Of, 0). (7.237)
The first term is written
X= (d[u—u(s — &(s, t)At + AED, D])
=([1 +ag 2488 2 at? lou -u(s, ))
e a
= [: + pea |fu, s,t)—-At (2 d[u —u(s, ol). (7.238)
The underlined term is rewritten as
Y= (é2 d[u—u(s, a) -2 (E6{u—a(s, ))) - (ou als, 2)] 38),
(7.239)
Since the correlation between u(s, t) and u(s', t) decreases quickly with
an increase in |s ~ s‘|, the first average in eqn (7.239) becomes
(&d[u — u(s, t)]) = [asrxey: (u(s', thu(s', t)d[u — u(s, t)])
m= fase: (u(s', t)u(s’, )) (d[u —u(s, 2)))
0
= [asrac dats’, t)u(s’, t))f(u, s, t). (7.240)
0
From eqn (7.232) it follows that
3 =K(t):u(s, Du(s, t). (7.241)CONSTITUTIVE EQUATION FOR REPTATION MODEL 277
From eqns (7.239)-(7.241), one has
y=2 [favrace: cue t)u(s’, t))f(u, s, |
asl} , , ”
— (x(t): u(s, t)u(s, t)d6[e — u(s, t)])
= a():(uls, Quls, D)f(e, 5,4 (EG, 0) Zhlw, 5,
—K(t): uuf(u, s, t). (7.242)
Hence the time evolution equation for f(u, s, t) is obtained ast
a
2 (0. 5- (86.9) 2\pte,5,)-Z (ew, ftw, 5.)
+ x(t): (ua — (u(s, thu(s, t)))f(u, s, t). (7.243)
The average in eqn (7.243) can be expressed by f(u, s, t) as
(6, )) = [ ds" | dux(t):wuf(a, s', 0) (7.244)
and °
(ug(s, t)up(s, t)) = | duu, ugf(u, s, t). (7.245)
Hence eqn (7.243) is a nonlinear integro-differential equation for
Fu, s, t).
Equation (7.243) can be rewritten into more tractable form. By a
similar technique described in ref. 108, eqn (7.243) can be transformed
into a closed equation for S.g(s, t):
Sup(S, 1) = | a'(2 KG, 11°) Qap(EC 1) (7.246)
where K(s, t, t’) is the solution of the differential equation
a e 3 ,
(5- DeSa+ (Es,9) <)KG. tr)=0 (7.247)
with the initial condition
K(s,t,’)=1 at t=0’ (7.248)
+In ref. 108 the terms in the last parenthesis are erroneously omitted. This gives a
constitutive equation which includes Q°)(E) instead of Q(E) in eqn (7.246). Since the
difference in QE) and Q(E) is small, the error caused by this is not serious.278 MOLECULAR THEORY OF POLYMERIC LIQUIDS
and the boundary condition
K(s,t,t')=0 at s=—-L/2 and s=L/2. (7.249)
Finally (&(s, 2) is given by
(86.9) = | ds" )Su066'. 1). (7.250)
0
Equations (7.246)-(7.250) determine S,,(s, t). Given S,g(s, t), the stress
can be calculated by eqn (7.178).
In the special case of step strain, one can solve the set of equations
(7.246)-(7.250) rigorously and obtain the results given in eqn (7.122). In
the general case, the solution of the equation needs numerical calcula-
tion. It turns out that the difference between eqns (7.194) and (7.246) is
not large for the usual flow history discussed in Section 7.6. For such
flows, the simple constitutive equation will be useful.
The effect of the IA approximation has also been examined for large
amplitude oscillatory shear deformation.’ In this case the result of the
constitutive equation without using the IA approximation is shown to be
in better agreement with experimental results.""°
It has been shown by Marrucci"!™ and Marrucci and Grizzuti that
analysis at the level of accuracy of this section is required to derive
Weissenberg effect correctly.
7.10 Further applications
Here we shall briefly discuss some pending problems which have not been
discussed in the previous sections.
7.10.1 Branched polymers
As discussed in Section 6.4.5, reptation is severely suppressed if the
polymer has long branches. Indeed it has been observed that the
dynamical properties of branched polymers are quite distinct from those
of linear polymers. So far studies have been done for branched polymers
of the simplest type, the star-shaped polymer in which f chains are
connected to a centre. The observed phenomena are:
(i) The diffusion constant Dg of a star polymer in a high molecular
weight matrix is much smaller than that of a linear polymer of the same
molecular weight,'" and the molecular weight dependence of Dg is much
stronger than that of linear polymers. This is consistent with the
prediction of the reptation theory’? (eqn (6.118)).
+The same constitutive equation has recently been derived by G. Marrucci (J. Non-
Newtonian Fluid Mech, to appear) by a different method.FURTHER APPLICATIONS 279
LE Wipes imal
(a) (o) (2)
Fig. 7.26. Relaxation of a star polymer. Figures show the states (a) before the
deformation, (b) immediately after the deformation, (c) at a later time ¢. The
deformed part of the tube which contributes to the stress is denoted by oblique
lines.
plateau region and the steady-state compliance J® is Rouse like (i.e.,
proportional to M) even if the molecular weight becomes quite high.’"
The anomalous behaviour in the linear viscoelasticity has been ex-
plained by the tube model."2""56 Figure 7.26 shows schematically how
the stress relaxation takes place in star polymers. In the crude theory,'’*
it is assumed that the centre of the star is fixed during the viscoelastic
relaxation time and that the relaxation takes place only by the contour
length fluctuation, i.e., by the process that the polymer retracts its arm
down the tube and evacuates from the deformed tube as shown in Fig.
7.26.
Let y(s,f) be the probability that the tube segment s which is
separated from the centre by the contour length s still remains at time ¢,
then the relaxation modulus is written as
GW) = GY x [ove. ) (7.251)
7%
where Z, is the equilibrium length of the tube for the arm of the star
polymer, and the constant G{? can be identified, in a first approximation,
with the plateau modulus for linear polymers:
pRT
GM = (7.252)
A simple approximation for #(s, ¢) is
w(s, t) = exp(—t/r(s)) (7.253)
where t(s) is the average time at which the chain end first reaches the
tube segment s, i.e., the contour length L,(¢) first becomes equal to s. As280 MOLECULAR THEORY OF POLYMERIC LIQUIDS
was discussed in Section 6.4.5, the motion of L,(t) can be regarded as a
Brownian motion of a particle in a harmonic potential
3kpT
2N,b”
where N, is the number of Rouse segments in the arm. Hence the time
1(s) is estimated as
U(L,) = (L,-L,) (7.254)
(£. -s)
t(s) = D ° exp((U(L, =s) — U(L,)\VkpT). (7.255)
The maximum relaxation time is given by
rc 3. 5N2 b4 b\?
Tmax = 15 = 0) =" exe 5552 i) kya exp[3n.(2) | (7.256)
Equation (7.255) is then written as
b 2
1(s) = Tmax ~ 8)? exp(3N.(2) (2-28) (7.257)
where
E=s/L,. (7.258)
Consider the case
2
a=in,(2) >1, (7.259)
then the viscosity is evaluated as
~ Le 1
no= | JG) = GPz- | dst(s) = G9 tax | dE (1 - §)* exp(-2a8 + a8%)
= GO tex d& exp(—2aé) = x G2 an. (7.260)
0
Similarly, the steady-state compliance is obtained as
1 a
(0) — =—
= | a(n = 5. (7.261)
Since the molecular weight of an arm is M/f, it follows from eqns (7.54)
and (7.259)
15M
oa (7.262)FURTHER APPLICATIONS 281
Thus eqns (7.260) and (7.261) are written as
No® (#) ox (Sa) (7.263)
and
O) =e SER (7.264)
The results of eqns (7.263) and (7.264) are in qualitative agreement with
experimental results: the viscosity increases steeply because of the
exponential factor, and the steady state compliance is proportional to M.
However, the quantitative agreement is not satisfactory. The observed
viscosity is smaller than the calculated one, and the best fit with
experiments is obtained only when the numerical coefficient in the
exponential of eqn (7.263) is replaced by a smaller number (about 1/2)
instead of 15/8.16 This suggests that relaxation mechanisms other than
the contour length fluctuations are important for star polymers. Indeed it
has been pointed out” that in the case of star polymers the constraint
release, and perhaps other tube reorganization processes, are as impor-
tant as the contour length fluctuation.
That the tube reorganization is important for star polymers is indicated
by another experiment. Kan et al.’’ found that the relaxation time of a
star polymer dispersed in a crosslinked system is by orders of magnitude
larger than that in the melt, while for linear polymers the former is larger
only by a factor of 2 or 3.
The constraint release or other mechanisms of the tube reorganization
are supposed to be important in other branched polymers such as
H-shaped polymers’ or ring polymers." Theoretical prediction for the
theological properties of these polymers is interesting and challenging.
7.10.2 Molecular weight distribution
Various experimental data suggest that the tube reorganization is im-
portant in linear polymers with molecular weight distribution.
(i) The diffusion constant of a polymer (of molecular weight M) in a
matrix (of molecular weight P) has been found to be essentially
independent of P if P is larger than a certain value P, which is between M
and M,.°!° This indicates that the tube reorganization is weak in
monodisperse systems (M = P). On the other hand, if P becomes smaller
than P., the diffusion constant increases with decreasing P.'?°!71
(ii) The linear viscoelasticity of a mixture of two polymers of the same282 MOLECULAR THEORY OF POLYMERIC LIQUIDS
Fig. 7.27. Discrepancy between experimental results and eqn (7.265).
species, but different molecular weights M, and Mg (M,> Ms) is not
explained by the model which includes only reptation. According to the
fixed tube model, the relaxation modulus of the mixture is the weight
average of that of the pure melts of individual components:
GO (0) = Wa Ga(t) + Wa Galt) (7.265)
where
Pa PB
w=——, =. 7.266)
“pats “ Pat PB ( )
The discrepancy between the experimental results” and eqn (7.265) is
schematically explained in Fig. 7.27: G™™(t) shows two characteristic
telaxations, each corresponding to the disengagement of polymer A and
B. Though this feature is in agreement with eqn (7.265), the relaxation
time of the larger component tg, is shorter than that in the pure A
component 1}, and the plateau modulus for the larger polymer is lower
than expected from eqn (7.265). Kurata’ suggested that the experimen-
tal data can be fitted by
GOD) = (1 - w)Ga(t) + WAGa(t/ ma). (7.267)
These results clearly indicate that the tube constraint for a polymer
becomes weaker if it is made of shorter polymers. The weakening of the
tube can be expressed either by an increase in the step length," or by
an increase in the constraint release process,°* or both.» 7° However,
the interpretation seems to be still at a tentative level.
7.10.3 Future problems
The reptation model has been applied to various problems other than the
problems of viscoelasticity and diffusion that have been discussed. TheseREFERENCES 283
include dielectric relaxation,” spinodal decomposition," polymer—
polymer welding,“°*!_ diffusion controlled _ reaction,“ and
crazing."%*"55 A concise review of various applications is given by de
Gennes and Léger.'*
On the whole the reptation model works well qualitatively, and for
several problems it gives quantitatively successful predictions. However,
many problems remain unsolved.
Perhaps the most important problem is the tube reorganization. We
have seen that the tube reorganization is important in branched polymers
and in linear polymers with polydispersity. It will also be important in a
nonuniform system such as polymer mixtures. So far the reptation theory
is based on the assumption that there is a tube which is characterized by a
single parameter a, the step length of the tube. Though the outcome of
this simple assumption is quite fruitful, one could ask: to what extent is
this picture correct?
A complete answer to this question will be given when the tube is
derived from more basic equations such as eqn (5.84) by a kind of mean
field approximation. This will require a new development of statistical
mechanics since the tube is a dynamical concept rather than static.
(Notice that the mean force acting on the polymer vanishes if it is
averaged over a time longer than ty, so that the average of the
surrounding field must be taken over a finite time.) Perhaps the tube is
better understood as representing the effect of dynamical correlation of
the environment rather than the usual mean field.
A slightly different, but closely related, problem is rubber elasticity.
Here the dynamical problem does not arise since the topological
constraints are permanent. However, the correlation plays an essential
role in the problem. Indeed it is the correlation in the topological
structure between the undeformed state and the deformed state that gives
rise to the rubber elasticity. In the modern theory of rubber elasticity,
this correlation is neatly handled by the replica method.'°7** Generali-
zation of this method to dynamical problems might be quite useful.
On the other hand, apart from that purely theoretical approach, it will
be quite promising to develop a theory by closely studying experimental
results. Collaboration between experiment and theory will be essential
for further progress.
References
1. (19), J. D., Landel, R. F. and Williams, M. L., J. Appl. Phys. 26, 359
2. Ferry, J. D., Viscoelastic Properties of Polymers (3rd edn). Wiley, New
York (1980).
3. Graessley, W. W., Adv. Polym. Sci. 16, 1 (1974).aus
Seo
11.
12.
13.
14.
15.
16.
17.
18.
19.
MOLECULAR THEORY OF POLYMERIC LIQUIDS
. Baur, M. E., and Stockmayer, W. H., J. Chem. Phys. 43, 4319 (1965).
. Higgins, J. S., Nicholson, L. K., and Hayter, J. B., Polymer 22, 163 (1981).
. Richter, D., Hayter, J. B., Mezei, F., and Ewen, B., Phys. Rev. Lett. 41,
1484 (1978).
. Baumgartner, A., Kremer, K., and Binder, K., Faraday Symp. Chem. Soc.
18, 37 (1983),
. Klein, J., Nature (London), 271, 143 (1978); Phil. Mag. 443, 771 (1981).
. Klein, J., and Briscoe, B. J., Proc. R. Soc. London A365, 53 (1979).
. Hervet, H., Léger, L., Rondelez, F., Phys. Rev. Lett. 42, 1681 (1979);
Léger, L., Hervet, H., and Rondelez, F., Macromolecules 14, 1732 (1981).
Tanner, J. E., Macromolecules 4, 748 (1971); Tanner, J. E., Liu, K. J., and
Anderson, J. E., Macromolecules 4, 586 (1971).
Bachus, R., and Kimmich, R., Polymer 24, 964 (1983).
Tirrell, M., Rubber Chem. Tech. 57, 523 (1984).
Doi, M., and Edwards, S. F., J. C. S. Faraday Trans. 274, 1802 (1978).
Doi, M., and Edwards, S. F., J. C. S. Faraday Trans. 274, 1818 (1978).
Walters, K., Rheometry. Chapman & Hall, London. New York (1975).
Astarita, G., and Marrucci, G., Principles of Non-Newtonian Fluid
Mechanics, McGraw-Hill, London (1974).
Lodge, A. S., Elastic Liquids. Academic Press, London (1964).
Bird, R. B., Armstrong, R. C., Hassager, O., and Curtiss, C. F., Dynamics
of Polymeric Liquids, Vols. 1, 2. Wiley, New York (1977).
. See for example Treloar, L. R. G., Rep. Prog. Phys. 36, 755 (1973); and
Treloar, L. R. G., The Physics of Rubber Elasticity, (3rd edn). Clarendon
Press, Oxford (1975).
. Green, M. S., and Tobolsky, A. V., J. Chem. Phys. 14, 80 (1946).
. Yamamoto, M., J. Phys. Soc. Jpn 11, 413 (1956), 12, 1148 (1957); 13, 1200
(1958). Lodge, A. S., Rheol. Acta 7, 379 (1968).
- Coleman, B. D., Arch. Ratl. Mech. 17, 1 (1964), 17, 230 (1964). Truesdell,
C., and Noll, W., The nonlinear field theories of mechanics, in
Encyclopedia of Physics III/3. Springer (1965).
|. Janeschitz-Kriegl, H., Polymer Melt Rheology and Flow Birefringence.
Springer, New York (1983).
. Janeschitz-Kriegl, H., Adv. Polym. Sci. 6, 170 (1969).
. Historical development of the stress optical law is described in refs 20 and
24. A recent study is given by Wales J. L. S., The Application of Flow
Birefringence to Rheological Studies of Polymer Melts. Delft Univ. Press,
Rotterdam (1976).
. DiMarzio, E. A., J. Chem. Phys. 3, 1563 (1962).
. Fukuda, M., Wilkes, G. L., and Stein, R. S., J. Polym. Sci. A2, 9, 1417
(1971).
. Jarry, J. P., and Monnerie, L., Macromolecules 12, 316 (1979).
. Tobolsky, A. V., Properties and Structure of Polymers. Wiley, New York
(1960).
. Onogi, S., Masuda, T., and Kitagawa, K., Macromolecules 3, 109 (1970).
. Doi, M., Chem. Phys. Lett. 26, 269 (1974).
. Berry, G. C., and Fox, T. G., Adv. Polym. Sci. 5, 261 (1968).
. Casale, A., Porter, R. S., and Johnson, J. F., J. Macromol. Sci. Rev.
Macromol. Chem. CS, 387 (1971).35.
36.
37.
38.
39.
40.
51,
52.
33.
RESVSSSSIRAL
&
REFERENCES 285
Odani, H., Nemoto, N., and Kurata, M., Bull. Inst. Chem. Res. Kyoto
Univ. 50, 117 (1972).
Graessley, W. W., J. Polym. Sci. 18, 27 (1980).
van Krevelen, D. W., Properties of Polymers, p. 338. Elsevier, Amsterdam
(1976).
Graessley, W. W., and Edwards, S. F., Polymer 22, 1329 (1981).
Bartels, C. R., Crist, B., and Graessley, W. W., Macromolecules 11, 2702
(1984).
Kremer, K., Macromolecules 16, 1632 (1983).
. A recent review on the computer simulation for polymer dynamics is given
by Baumgirtner, A., Ann. Rev. Phys. Chem. 35, 419 (1984).
. Deutsch, J. M., Phys. Rev. Lett. 49, 926 (1982).
. Higgins, J. §., and Roots, J. E., J. C. S. Faraday Trans II 81, 757 (1985).
. de Gennes, P. G., Macromolecules 9, 587, 594 (1976).
- Adam, M., and Delsanti, M., J. Phys. (Paris) 44, 1185 (1983).
. Raju, V. R., Menezes, E. V., Marin, G., Graessley, W. W., and Fetters, L.
J., Macromolecules 14, 1668 (1981).
. Onogi, S., Masuda, T., Miyanaga, N., and Kimura, Y., J. Polym. Sci. Part
A2 5, 899 (1967); Onogi, S., Kimura, S., Kato, T., Masuda, T., and
Miyanaga, N., J. Polym. Sci. C15, 381 (1966).
. Masuda, T., Toda, N., Aoto, Y., and Onogi, S., Polymer J. 3, 315 (1972).
. Nemoto, N., Ogawa, T., Odani, H., and Kurata, M., Macromolecules 5,
641 (1972).
). See for example, Osaki, K., Fukuda, M., and Kurata, M., J. Polym. Sci.
Phys. ed. 13, 775 (1975). Recent data for melt are given by Lin, Y. H.,
Macromolecules 17, 2846 (1984); J. Rheol. 28, 1 (1984).
Graessley, W. W., Faraday Symp. Chem. Soc. 18, 7 (1983).
Bernard, D. A., and Noolandi, J., Macromolecules 15, 1553 (1982); 16, 548
(1983).
Wendel, H., and Noolandi, J., Macromolecules 15, 1318 (1982); Wendel,
H., Colloid Polym. Sci. 259, 908 (1981).
Curtiss, C. F., and Bird, R. B., J. Chem. Phys. 74, 2016, 2026 (1981).
. Ball R., private communication.
Deutsch, J. M., Phys. Rev. Lett. 54, 56 (1985).
Doi, M., J. Polymer Sci. 21, 667 (1983); J. Polym. Sci. Lett. 19, 265 (1981).
Lin, Y. H., Macromolecules 19, 159, 168 (1986).
Klein, J., Macromolecules 11, 852 (1978).
Daoud, M., and de Gennes, P. G., J. Polym. Sci. Phys ed. 17, 1971 (1979).
Graessley, W. W., Adv. Polym. Sci. 41, 67 (1982).
Marrucci, G., J. Polym. Sci. Phys. 23, 159 (1985).
Viovy, J. L., J. Physique Lett. 46, 847 (1985).
. Doi, M., J. Polym. Sci. 18, 1005 (1980).
. Einaga, Y., Osaki, K., Kurata, M., Kimura, S., and Tamura, M., Polymer.
J. 2, 580 (1971); Fukuda, M., Osaki, K., and Kurata, M., J. Polym. Sci.
Phys. 13, 1563 (1975).
Osaki, K., Bessho, N., Kojimoto, T., and Kurata, M., J. Rheol. 23, 617
1979).
. Osaki, K., Kimura, S., and Kurata, M., J. Polym. Sci. Phys. ed. 19, 517
(1981).Be
$8
MOLECULAR THEORY OF POLYMERIC LIQUIDS
Osaki, K., and Kurata, M., Macromolecules 13, 671 (1980).
Osaki, K., Nishizawa, K., and Kurata, M., Macromolecules 15, 1068 (1982).
. Takahashi, M., Nakamura, H., Masuda, T., and Onogi, S., Polymer
Preprints Japan 30, 1970 (1981).
. Lodge, A. S., and Meissner, J., Rheol. Acta, 11, 351 (1972); Lodge, A. S.,
ibid. 14, 664 (1975).
. Marrucci, G., and de Cindio, B., Rheol. Acta, 19, 68 (1980).
. Taylor, C. R., Greco, R., Kramer, O., and Ferry, J. D., Trans. Soc. Rheol.
20, 141 (1976); Noordermeer, J. M., and Ferry, J. D., J. Polym. Sci. 14,
509 (1976).
. Vrentas, C. M., and Graessley, W. W., J. Rheol. 26, 359 (1982); Pearson,
D. S., IUPAC Proceedings, 28th Macromolecular Symposium, July, p. 866
(1982).
. Marrucci, G., and Grizzuti, N., J. Rheol. 27, 433 (1983).
. McLeish, T. C. B., and Ball, R. C., J. Polym. Sci., to be published.
. Marrucci, G., and Hermans, J. J., Macromolecules 13, 380 (1980).
. Viovy, J. L., Monnerie, L., and Tassin, J. F., J. Polym. Sci. Phys. ed. 21,
2427 (1983); Viovy, J. L., J. Polym. Sci. Phys. ed. 23, 2423 (1985).
. Boué, F., Adv. Polym. Sci. to be published; Boué, F., Nierlich, M., and
B82 8 S$ BSR R SSRAS
98.
Osaki, K., Faraday Symp. Chem. Soc. 18, 83 (1983): Boué, F., Nierlich,
M., Jannink, G., and Ball, R. C., J. Phys. (Paris) 43, 137 (1982); J. Phys.
Lett. 43, L585, L593 (1982).
Bastide, J., Herz, J., and Boué, F., J. Physique 46, 1967 (1985).
Sekiya, M., and Doi, M., J. Phys. Soc. Jpn 51, 3672 (1982).
Noolandi, J., and Hong, K. M., J. Physique Lett, 45, L149 (1984).
Boué, F., Osaki, K., and Ball, R. C., J. Polym. Sci. 23, 833 (1985).
Takahashi, M., Masuda, T., Bessho, N., and Osaki, K., J. Rheol. 24, 517
(1980).
Takahashi, M., Masuda, T., Oono, H., and Onogi, S., Polymer Preprints
Japan 33, 871 (1984).
Meissner, J., Rheol. Acta, 10, 230 (1971).
Currie, P. K., J. Non-Newtonian Fluid Mech. 11, 53 (1982).
on B., Kearsley, E. A., and Zapas, L. J., Trans. Soc. Rheol. 7, 391
1963).
Osaki, K., Ohta, S., Fukuda, M., and Kurata, M., J. Polym. Sci. A14, 1701
(1976).
Chang, W. V., Bloch, R., and Tschoegl, N. W., Rheol. Acta 15, 367 (1976);
J. Polym. Sci. A18, 923 (1977).
. Wagner, M. H., Rheol. Acta 15, 136 (1976); 16, 43 (1977).
Phillips, M. C., J. Non-Newtonian Fluid, Mech. 2, 109, 123, 139 (1977).
. Osaki, K., Proceeding of the 7th International Congress on Rheology (eds C.
Klason and J. Kubat). Chalmers University of Technology, Gothenberg,
p. 104 (1976).
. Doi, M., and Edwards, S. F., J. C. S. Faraday Trans. 275, 38 (1979).
. Bird, R. B., Saab, H. H., and Curtiss, C. F., J. Phys. Chem. 86, 1102
(1982); J. Chem. Phys. 7, 4747, 4758 (1982).
Osaki, K., and Doi, M., Polym. Eng. Rev. 4, 35 (1984).
. Marrucci, G., Adv. Transport Processes § (eds. A. S. Mujumdar and R. A.
Mashelkar). New York (1985).
Camachandran, S., Gao, H. W., and Christiansen, E. B., J. Rheol. 25, 213
981).101.
102.
103.
104.
105.
106,
107.
108.
109.
110.
REFERENCES 287
. Tanner, R. I., Trans. Soc. Rheol. 17, 365 (1973).
100.
Laun, H. M., and Munstedt, H., Rheol. Acta 18, 427 (1979); Munstedt, H.,
J. Rheol. 23, 421 (1979).
Vinogradov, G. V., Rheol. Acta 12, 273 (1973); see also Lin, Y. H., J.
Rheol. 29, 65 (1985).
Menezes, E. V., and Graessley, W. W., J. Polym. Sci. Phys. 20, 1817
(1982).
Takahashi, M., Masuda, T., and Onogi, S., Polymer Preprints Japan 29,
1807 (1980).
Osaki, K., and Kurata, M., J. Polym. Sci. Phys. 18, 2421 (1980).
Zapas, L. J., Deformation and Fracture of High Polymers (eds H. H.
Kausch, J. A. Hassell, and R. I. Jaffe). Plenum Press, New York (1974);
see also McKenna, G. B., and Zapas, L. J., J. Rheol. 23, 151 (1979); 24,
367 (1980).
Osaki, K., Kimura, S., and Kurata, M., J. Rheol. 25, 549 (1981).
Doi, M., J. Polym. Sci. 18, 1891 (1980).
Doi, M., J. Polym. Sci. 18, 2055 (1980).
Helfand, E., and Pearson, D. S., J. Polym. Sci. 20, 1249 (1982).
Pearson, D. S., and Rochefort, W. E., J. Polym. Sci. Phys. ed. 20, 83
(1982).
110a. Marrucci, G., J. Non-Newtonian Fluid Mech. 21, 329-36 (1986).
110b. Marrucci, G. and Grizzuti, N., J. Non-Newtonian Fluid Mech, 21, 319-28
111.
112.
113.
114.
115.
116.
117.
118.
119.
. Smith, B. A., Samulski, E. T., Yu, L. P., and Winnik, M. A., Phys. Rev.
121.
125.
126.
(1986).
Klein, J., Fletcher, D., and Fetters, L. J., Faraday Symp. Chem. Soc. 18,
159 (1983).
de Gennes, P. G., J. Phys. (Paris) 36, 1199 (1975).
Kraus, G., and Gruver, J. T., J. Polym. Sci. A3, 105 (1965); J. Appl.
Polym. Sci. 9, 739 (1965).
Graessley, W. W., Masuda, T., Roovers, J. E. L., and Hadjichristidis, N.,
Macromolecules 9, 127 (1976); Graessley, W. W., and Roovers, J.,
Macromolecules 12, 959 (1979); Raju, V. R., Menezes, E. V., Marin, G.,
Graessley, W. W., and Fetters, L. J., Macromolecules 14, 1668 (1981).
Doi, M., and Kuzuu, N., J. Polym. Sci. Lett. 18, 775 (1980).
Pearson, D. S., and Helfand, E., Macromolecules 17, 888 (1984).
Kan, H. C., Ferry, J. D., and Fetters, L. J., Macromolecules 13, 1571
(1980).
Roovers, J., Macromolecules 17, 1196 (1984).
Roovers, J., J. Polym. Sci. 23, 1117 (1985).
Lett, 52, 45 (1984).
Green, P. F., Mills, P. J., Palmstrgm, C. J., Mayer, J. W., and Kramer, E.
J., Phys. Rev. Lett. 53, 2145 (1984).
. Masuda, T., Takahashi, M., and Onogi, S., Appl. Polym. Symp. 20, 49
(1973); Bogue, D. C., Masuda, T., Einaga, Y., and Onogi, S., Polym. J. 1,
563 (1970).
ba Kurata, M., Macromolecules 17, 895 (1984).
Montfort, J. P., Marin, G., and Monge, P., Macromolecules 17, 1551
(1984).
Watanabe, H., and Kotaka, T., Macromolecules 17, 2316 (1984).
Masuda, T., Yoshimatsu, S., Takahashi, M., and Onogi, S., Polymer
Preprints Japan 33, 2699 (1984).288
127,
128.
129.
130.
131,
132.
133.
134.
135,
136.
137.
138.
MOLECULAR THEORY OF POLYMERIC LIQUIDS
Adachi, K., and Kotaka, T., Macromolecules 17, 120 (1984); 18, 466 (1985).
de Gennes, P. G., J. Chem. Phys. 72, 4756 (1980).
Pincus, P., J. Chem. Phys. 75, 1996 (1981).
de Gennes, P. G., C. R. Acad. Sci. Paris B291, 219 (1980).
Prager, S., and Tirrell, M., J. Chem. Phys. 75, 5194 (1981); Adolf, D.,
Tirrell, M., and Prager, S., J. Polym. Sci. 23, 413 (1985).
Tulig, T. J., and Tirrell, M., Macromolecules 14, 1501 (1981).
de Gennes, P. G., J. Chem. Phys. 16, 3316, 3322 (1982).
Kramer, E. J., Adv. Polym. Sci. 52/53, 1 (1983).
Evans, K. E., and Donald, A. M., Polymer 26, 101 (1985).
de Gennes, P. G., and Léger, L., Ann. Rev. Phys. Chem. 33, 49 (1982).
Deam, R. T., and Edwards S. F., Phil. Trans. R. Soc. A280, 317 (1976).
Ball, R. C., Doi, M., Edwards S. F., and Warner, M., Polymer 22, 1010
(1981).8
DILUTE SOLUTIONS OF RIGID
RODLIKE POLYMERS
8.1 Rodlike polymers
Though many polymers are flexible and take a random coil structure,
there is a large class of polymers which are not flexible and assume a
rodlike structure. For example, some polypeptides or polynucleotides
form a helix structure which can be regarded effectively as a rigid rod. If
the chemical bonds in the backbone chain consist of double bonds or
phenylene rings, the internal rotation of the polymer is severely restricted
and the polymer takes an elongated form. These latter type of rodlike
polymers are quite important in polymer technology because of their
capability of creating very strong fibres, and an increasing amount of
research is being done as a result.
The physical properties of the rodlike polymers differ from those of
flexible polymers in many respects.
Firstly, an obvious distinction is that rodlike polymers are much larger
than flexible polymers with the same molecular weight. If the polymer is
a straight rod, its radius of gyration R, is proportional to the contour
length of the polymer, or the molecular weight M, as compared to the
relation R,« M” (v~0.6) for flexible polymers. The elongated form of
the polymer is reflected in various dilute solution properties such as the
larger intrinsic viscosity, larger relaxation time, or smaller diffusion
constant as compared to those of flexible polymers.’
Secondly, due to the large molecular anisotropy, rodlike polymers are
much more easily oriented by an external field and show large birefrin-
gence. This enables us to use electric or magnetic birefringence as a
practical tool to study the rotational motion of these polymers.”
Thirdly, the distinction between rodlike polymers and flexible polymers
becomes more pronounced as concentration increases. Due to their
larger size, the interaction of the rodlike polymers becomes important at
a much lower concentration than with flexible polymers, and, as we shall
show later, the effect of the entanglement is much more remarkable.
Fourthly, but not least, when the concentration becomes sufficiently
high, rodlike polymers spontaneously orient towards some direction, and
form a liquid crystalline phase.° It is this capability of forming a highly
ordered phase that produces strong fibres.
In this and the following two chapters, we shall discuss the physical
Properties of such polymers. Although real polymers have finite rigidity
and can bend to some extent, we shall mainly consider the extreme290 DILUTE SOLUTIONS OF RIGID RODLIKE POLYMERS
situation, i.e. tigid rodlike polymers. The effect of the flexibility will be
discussed only briefly.
Theoretical treatment of rodlike polymers is much easier than for
flexible polymers since rodlike polymers can have only two kinds of
motion, i.e., translation and rotation. Once the basic equation is set up,
mathematical analysis is easy. However, important physics is included in
an essential way in the problems of rodlike polymers. In particular, the
importance of the orientational degrees of freedom and the peculiar
nature of the topological constraints will be seen clearly in this system.
In this chapter we shall discuss the properties of dilute solutions. The
properties at higher concentrations will be discussed in later chapters.
8.2 Rotational diffusion
8.2.1 Rotational Brownian motion
Rodlike polymers do two kinds of Brownian motion, translation and
rotation. The translational Brownian motion is the random motion of the
position vector R of the centre of mass, and the rotational Brownian
motion is the random motion of the unit vector u which is parallel to the
polymer.
To visualize the rotational Brownian motion we imagine the trajectory
of u(t), which is on the surface of the sphere |u| =1 (see Fig. 8.1). For
short times, the random motion of u(t) can be regarded as Brownian
motion on a two-dimensional flat surface, and the mean square displace-
Fig. 8.1. Rotational diffusion.ROTATIONAL DIFFUSION 291
ment of u(¢) in time ¢ is written as
((u(@) — u))?) =4D,t (for D,t «1). (8.1)
The coefficient D, is called the rotational diffusion constant. Note that the
dimension of D, is (time)', and is not the same as that of the
translational diffusion constant, which is (length)?/(time).
Equation (8.1) is correct only for D,t <1. To discuss the general case,
we have to study the Smoluchowski equation for the rotational Brownian
motion. This equation can be derived straightforwardly according to the
Kirkwood theory* described in Section 3.8. Such a derivation is given in
Appendix 8.1. Here we derive it by an elementary method to clarify the
underlying physics.
8.2.2. Hydrodynamics of rotational motion
As was discussed in Chapter 3, the first step in deriving the Smoluchow-
ski equation is to obtain the phenomenological relation between the force
and flux by using the hydrodynamics of the problem.
Consider a rod placed in a quiescent viscous fluid. If an external field
exerts a torque N on the rod, the rod will rotate with certain angular
velocity w. For thin rod, we may neglect the rotation around u, and
assume that both w and N are perpendicular to u. If N is small, @ is
linear in N, and by symmetry, parallel to N.
1
o E N. (8.2)
The coefficient €, is called the rotational friction constant.
A simple estimation of & is done for the ‘shish-kebab model’
illustrated in Fig. 8.2: the rod is regarded as made up of N = L/b ‘beads’,
which are numbered from —N/2 to N/2, When the rod rotates with
angular velocity w, the bead n which is separated from the centre by the
distance nb moves with velocity V, =(@ X nbu).
If the hydrodynamic interaction is neglected, the frictional force acting
on the segment n is —€)V,, where €=3,7,6 is the translational friction
constant of the bead. Thus the total torque due to the hydrodynamic
friction is given by
NI2
Nevction=— 2, _nbu X £oV,
n=—-NI2
NI
=- > nbuX (foo X nbu)
n=-N2
NZ 2 3 aL?
=- 2620 = — 22 (NV yan Ze .
bo 3 n'b?@ = —Gan,b)6?=(X) o=-n wo (83)
n=-Ni2292 DILUTE SOLUTIONS OF RIGID RODLIKE POLYMERS
ue
Fig. 8.2. Rodlike polymer and shish-kebab model, which consists of N = L/b
beads of diameter b placed along a straight line.
which must balance with the external torque N. Hence
ansL?
t,=2be
4
If the hydrodynamic interaction among the beads is taken into account,
€, is shown to be (see Appendix 8.1):
(8.4)
an,L?
3in(L/2b)
More precise hydrodynamic calculation for the cylinder gives a correction
y to the denominator:+
c= (8.5)
p=
* 3(in(L/b) — y)”
+ For a prolate elipsoid, an exact calculation can be done and the result is”
16x -( yl 2p-1 (ete vn) T
= 0,°(1-4)|—*— -1
beng me 3p(p?= 1p P=)
where 2a is the length of the long axis and p is the aspect ratio. For p >2, the above
equation is approximated as
(8.6)
te 1627,a*
* 3{2tn(2p) — 1]
which agrees with eqn (8.6)ROTATIONAL DIFFUSION 293
The calculation based on the point force approximation® gives y =0.8,
while more recent calculation” indicates that y weakly depends on L/b.
The torque N is now expressed by the potential U(u) of the external
field. Consider a small rotation Sw which changes u to u+ dw Xu. The
work needed for this change is —N- dy, which must be equal to the
change in U, i.e.,
-N- by = Ulu + by Xu) — Ulu) = (69 Xu) “2 U=5y- (ux)
Hence 87
N=-RU 8.
wheret 68)
R=ux 2 : (8.9)
ou
The operator ®, called the rotational operator, plays the role of the
gradient operator 9/AR in translational diffusion.
An important property of ® is the formula of integration by parts, i.e.,
for the integral over the entire surface of the sphere of |u| = 1,
| duA(u)RB(u) = — | du[RA(u)]B(u). (8.10)
(In quantum mechanics, —i@ corresponds to the angular momentum
operator so that eqn (8.10) is equivalent to the Hermitian property of this
operator.)
Now if the fluid surrounding the rod is flowing with a certain velocity
gradient, there will be an additional angular velocity w» of the rod, which
again can be calculated by hydrodynamics.} For a slender rod, Wp is
obtained by a simple geometrical reasoning explained in Fig. 8.3:
Oo =UuX Kou. (8.11)
tIn eqn (8.9) 3/u, means the partial derivative in which u,, u,, u, are regarded as
indent variables. Since w is a unit vector, there are many ways to express U. For
example, consider the following three quantities
= = uy =(1— 22-42)
Rate Boca B= G-a- un
All represent the same quantity for the unit vector w. It is easily checked that, though
OF,/du,, 9F,/du, and 3F,/3u, are not equal to each other, RA, RE, and RF, are all
equal. Thus the derivative RF has no ambiguity.
+For a spheroid of aspect ratio p = a/b, wy is given by**
=ux p Ke -age-s)
won mT Marre “#)-
In the limit of p>, this reduces to eqn (8.11).294 DILUTE SOLUTIONS OF RIGID RODLIKE POLYMERS
Fig. 8.3. Geometrical meaning of eqn (8.11). If the rod follows the macroscopic
velocity gradient, its direction changes as = K - u — (uu:K)u. Hence the angular
velocity @o is given by @ =u X a =u X(K-u).
Therefore the angular velocity of a rod immersed in a fluid with
velocity gradient « and subject to an external potential U(u) is given as
o= —-lausuxnen. (8.12)
é
This consequence of the hydrodynamics corresponds to eqn (3.118) for
translational motion.
8.2.3 Smoluchowski equation for rotational motion
Now it is easy to give an account of the Brownian motion. If Y(u; t) is
the probability distribution function of u, the Brownian motion is
included by adding the ‘Brownian potential’ k,T In W to U. The angular
velocity @ is now given by
w= ~ EM kel In W +) + ux Kw (8.13)
For given , u changes with the velocity w Xu, and the equation for the
conservation of the probability becomes
ow 9 a
= -Z-(oxuy= -(w x=) ‘aU =-R-(W). (6.14)
From eqns (8.13) and (8.14), we have the Smoluchowski equation forTRANSLATIONAL DIFFUSION 295
rotational diffusion
SEER: [kpTRY + RU] — R(X nu)
= DR [RY + RU] - R-(wXHe-u) (8.15)
where D, is defined by
kpT _ 3kgT(in(L/b) — y)
& a,b?
We shall later show that D, agrees with the rotational diffusion constant
defined by eqn (8.1).
Note the formal similarity between the rotational diffusion equation
and the usual translational diffusion equation: if the gradient operator
9/AR in the translational diffusion equation is replaced by the operator
&, the rotational diffusion equation is obtained.
The rotational Brownian motion can also be described by the Langevin
equation, but it is rarely used in the problem of rodlike polymers because
it is less convenient for calculation than the Smoluchowski equation.
D,=
. (8.16)
8.3 Translational diffusion
8.3.1 Hydrodynamics of translational motion
It is straightforward to include the translational motion into the Smol-
uchowski equation. Again the hydrodynamics is considered first. Suppose
the rod is moving with the velocity V in a quiescent fluid (see Fig. 8.4). If
the rod moves along u, the rod will feel a hydrodynamic drag, which is
parallel to V and is written as €V. On the other hand if V is
perpendicular to u, the drag is again parallel to V and is written as €, V.
AY
@F=CV (oe) F=C.V OFHCV,+0V,
Fig. 8.4. Anisotropy in the translation friction constant. (a) V||u, (b) V Lu, and
(c) general direction V= V+ Vi.296 DILUTE SOLUTIONS OF RIGID RODLIKE POLYMERS
In general the coefficients £, and €, are not equal to each other. They
are called the parallel and perpendicular components of the translational
friction constant, respectively.
Given €, and ¢,, the hydrodynamic drag for a translational motion in
a general direction is obtained as follows. Since the Stokes equation (eqn
(3.102)) is a linear equation, the hydrodynamic drag must be linear in V.
Thus if Vj, and V, are the parallel and the perpendicular components of
V, the drag is written as
FaQVjtoih. 6.17)
Substituting
Vi=(V-uu and V,=V-Y, (8.18)
we have
F=Cyuu-V+6.(l—uu)-V. (8.19)
The calculation of £, and £, based on the Kirkwood theory is given in
Appendix 8.I. The result is
2an,L
Si inci) (8.20)
6. = 26). (8.21)
Equation (8.19) is solved for V as
v= le uu +e- wu)| “F. (8.22)
If there is a macroscopic flow v(r)=« +r, there is an additional velocity
xR for the rod at point R, and eqn (8.22) becomes
Vm aut Om) eR, (8.23)
This is the result of hydrodynamic calculation.}
8.3.2 Smoluchowski equation including both translational and rotational
diffusion
We can now write down the Smoluchowski equation which includes both
the rotational and translational motions. Let W(R, u; t) be the probabil-
+Note that in the case of a rod, V is independent of the torque N acting on it. In the
general case, this is not true. (Consider for example a screw: the torque turns the screw and.
causes a translational motion.) The formula for the general case is given in refs 6 and 11,
and an example of its application to the Brownian dynamics is given in ref. 12TRANSLATIONAL DIFFUSION 297
ity distribution function for the rod in the configuration (R, u). The
velocity V is given by
v=-[pmet ze (-m)|- ZeTinw +0) +K- +R. (8.24)
The angular velocity is again given by eqn (8.13). Substituting this in the
continuity equation
awa
a — 5p VE) BR (a) (8.25)
we get
aw aw WOU
ae aR’ Pame + DG uw) (Rt ee an RY)
+DQR- (av + er?) R-(uXn«-u), (8.26)
where
_keT _keT n(L/b)
‘oT I“) 8.27
= ty lanl (8.27)
and
_kpT_keT in(L/b) (628)
1b, 4am
The constants Dy and D, characterize the diffusion parallel and
perpendicular to the rod axis: if the rod is along the z axis, then the
displacement of R in a small time interval At is given as
((Rx(At) — R(0))?) = ((Ry(At) — Ry(0))?) = 2D At
{(R,(At) — R,(0))*) = 2D, Az.
Since Dy >D_, the rod can move more easily in the direction parallel to
the axis than that perpendicular. Due to this anisotropy in the diffusion
constant, translational and rotational motions of a rod are generally
coupled with each other. For example, a concentration gradient of the
rodlike polymer can induce an anisotropy in the orientational distribution.
However, the reverse is not true: in a homogeneous system (in which the
Positional distribution is uniform), the translation—rotation coupling has
No effect: if the system is homogeneous, it will remain homogeneous even
if the orientational distribution is not isotropic. Thus in a homogeneous
system, one can discuss the rotational diffusion using eqn (8.15) instead
of the full Smoluchowski equation (8.26).
(8.29)298 DILUTE SOLUTIONS OF RIGID RODLIKE POLYMERS
8.4 Brownian motion in the equilibrium state
Having obtained the Smoluchowski equation, we now study the charac-
teristic features of the Brownian motion of a free polymer (U =0 and
«=0).
84.1 Vector correlation function (u(t) -u(0))
To see the rotational motion, let us consider the time correlation function
(u(t)-u(0)). According to the general prescription given in Chapter 3,
this is calculated by
(u(t) w(0)) = du du'a - 'G(u, u's 1).,(u"), (8.30)
where W,, is the equilibrium distribution function
1
Weq(u) = re (8.31)
and G(u, u';t) is the conditional probability that the polymer is in the
direction of u at time ¢, given that it was in the direction u’ at time t= 0.
This probability is the Green function of the diffusion equation
2G, u'; t)=D,R’G(u, u’; t) (8.32)
with the initial condition
G(u, u';¢=0) = 5(u—u'). (8.33)
Though the explicit form of G(u, u’; t) is available (see for example ref.
13 Chapter 7), the time correlation function can be calculated directly as
before. The time derivative of (u(t) - u(0)) becomes
3 (u(e)-0(0)) = feu duu -w'| 2 Gu, ws 0) ea(u’)
=D, [du du'u + w'[R°G(w, u's eu’). (8:34)
Using eqn (8.10) for the right-hand side, we get
2 (u(t) « u(0)) = D, i du du'[Ru- u']G(u, w'; (uw). (8.35)
By a straightforward calculationt
Relig = — CapyUly. (8.36)
t Here e,gy is Levi Civita’s symbol, i.e. e.g, =e, * (eg X€,), where e, is the unit vector in
the direction of the a axis. oer (eo erBROWNIAN MOTION IN THE EQUILIBRIUM STATE 299
Applying this twice, we have
Ritts = = — C apy Rally = Capyl ayully = —2ilp- , (8.37)
Hence eqn (8.35) is written as
Suto -u(0)) = ~2b, {du du'u-u'G(u, u'; t)Y,,(u')
= —2D, (u(t) - u(0)). (8.38)
Since (u(t) + u(0)) is equal to 1 at time t= 0, eqn (8.38) gives
(u(t) - u(0)) = exp(—2D,t). (8.39)
The rotational correlation time 1, is thus given by
1, =1/2D, (8.40)
From eqn (8.39), it follows that
((u(t) — u(0))?) = 2— 2(u(?) - u(0))
= 2(1 — exp(—2D,t)). (8.41)
For tD, <1, eqn (8.41) reduces to eqn (8.1), which gives a clear physical
meaning of D,.
In the same way, one can show?
3([u(e) - (0)P —3)) = exp(—6D,2) (8.42)
or in general
(P,(u(f) + u(0))) = exp(—D,n(n + 1)t) (8.43)
where P,(x) is the Legendre polynomial of n-th order
1 da n
Pa) = eq gyn DY (8.44)
8.4.2 Translational diffusion
Consider the mean square displacement of the centre of mass:
9) = ((R() - RO)’) (8.45)
This is calculated by essentially the same method as before. Let
G(R, u, R', a t) be the Green function for the configuration (R, 4),
c-[p, wR? + * (Dyuu + DI ua) ale (8.46)
with the initial condition
G(R, u, R', u’;t=0)=6(R—R')6(u -w'). (8.47)300 DILUTE SOLUTIONS OF RIGID RODLIKE POLYMERS
Then ¢(¢) is calculated from
e()= far du dR’ du’(R — R'PG(R, u, R', u’;t)Y.,(R’, u’). (8.48)
We again evaluate the time derivative of #(t) using eqn (8.46). The
resulting equation is written, after integration by parts, as
a - o
< o()= [aR du ar du’GY,,
a a
x [Dw + 5 -(Djuu + Dy (I~ wu) 5 |(R = RP
= far du dR' du’GW,,(2D, + 4D) = 2(D, + 2D,). (8.49)
Hence
$(t) = 2D, + 2D, )t. (8.50)
Thus $(f) increases linearly with ¢ and the diffusion constant Dg defined
by
Dg =lim x ((RQ) - RO))) (8.51)
is given by
Dg = Dit2Ps _ (Lib), (8.52)
3 32n,L
This formula can also be directly derived from the Kirkwood formula for
the diffusion constant eqn (4.102).
It must be noted that although $(#) increases linearly with time, the
diffusion of R is not Fickian because of the translation—rotation coupling.
Indeed as will be shown below, G(R, u, R', u';t) is not Gaussian in
R-—R’'. The Fickian diffusion is recovered if the relevant length-scale is
much larger than L.
8.4.3 Dynamic light scattering
As before, the Brownian motion of the polymer can be studied by
dynamic light scattering.’* If we take the shish-kebab model shown in
Fig. 8.2, the dynamical structure factor is given by
a D= pS explik R= ROM). 65)
(Note the normalization of eqn (8.53) is chosen so that g(0, 0) = 1, whichBROWNIAN MOTION IN THE EQUILIBRIUM STATE 301
is different from that chosen for flexible polymers.) Since
R,=R+nbu (8.54)
the sum over n is
NR NIZ
> exp(ik- R,) = exp(ik +R) | dn exp(ik + unb)
n==NI2 -Ne
- «py Sin(k uNB/2) _ py si(k + wL/2)
2exp(ik - R) rr Nexp(ik +R) aL (8.55)
Thus g(k, f) is expressed by R(t) and u(t) as
_ loan tpi) sin[K - u(s)] sin[K - u(0)]
ak, #) = (expGk [R(® - RCO) SEE TA
where
) (8.56)
K=kL/2. (8.57)
Let G,(u,u’;t) be the Fourier transform of the Green function
G(R, u, R’, u';t), which depends only on R — R’,
G,(u, ust) = | dRe*®G(R, u, 0, u’;1) (8.58)
then
sin(K + w) sin(K « u’)
= du’ " y
a(h 1) = [edu fau’Gy(u, u's) SESE Hy wy (8.59)
From eqns (8.46) and (8.58), G,(u, u'; t) satisfies
(2+ F)G., ust) =0 (8.60)
with
F=—D,R?+ Dy(k-u)?+ Di[- (ku) (8.61)
The solution of eqn (8.60) is involved, so here we shall briefly describe its
characteristic aspects. The limiting cases are:
(i) |K| K1, ie., [AI LL.
In this case, it is intuitively obvious that g(k, t) is described by the
Fickian diffusion with the diffusion constant Dg, so that
gk, t) = exp(—Dgk’). (8.62)
A formal justification of this is made by considering the eigenfunction
expansion of G,. Let , and A, be the eigenfunctions and the
corresponding eigenvalues of I.
Pp = op (8.63)302. +=DILUTE SOLUTIONS OF RIGID RODLIKE POLYMERS
Then
Ge= z exp(—A,1) (4) ¥p(u')- (8.64)
Substituting this into eqn (8.59) and using K «1, we have
Bh, =D enol —Agt( day, cw) A):
-23 ents Jawy,(u) (8.65)
If k=0, the eigenfunctions are given by spherical harmonics Y,,,(u) with
the eigenvalues A{2) = D,J(/ +1) and only the term of the lowest eigen-
function Yoo(u) = 1/V(4s) remains in the sum of eqn (8.65). If k is small,
the relaxation is still dominated by the eigenfunction of the lowest
eigenvalue
8(k,t) = exp(—Aot). (8.66)
Ag is obtained by perturbation theory:
Jo= | duYool Yon / | duY2,= PUA ee Do (8.67)
which justifies eqn (8.62).
For the above perturbation calculation to be justified, Aj must be much
smaller than the next smallest eigenvalue AS, = 2D,, i.e.,
Dok? «2D, (8.68)
which is rewritten using eqns (8.16) and (8.52) as
|A|L<«1. (8.69)
Equation (8.69) indicates that if the relevant length-scale is much larger
than L, the diffusion can be regarded as Fickian with the diffusion
constant Dg.
Gi) [K|>>1, ie., |k| L>>1.
In this case, g(k, t) is not expressed by a single exponential. However,
the initial decay rate is calculated easily:
TP =-Ling(&, D)loa
“75
~ fost _sin(K8) sink) sin( KE) sin) / I ag (ae (8.70)
sink »f pee Dy ¥.,(u)ORIENTATION BY AN ELECTRIC FIELD 303
where
£=K-u/|K| and K=|k|. (8.71)
By using the relation
OF | uXKOoF
72)
RF(E)= (RE) BE TK] 29€ (8.72)
we have, after some senate
[ae =[-p SE 20-92 3et DPE + Dude e|\e®).
(8.73)
The integral over & is carried out analytically by using the fact that K is
large: for example,
2B) = fan(S24) = fon) =
jae KE K “Kia 7) K
Straightforward calculation gives finally’
2
TP=Dke+ 5 DR. (8.74)
More detailed studies are given in the literature.'* 16
Since rodlike polymers have a large optical anisotropy, they have a
significant depolarized light scattering, which is particularly suitable for
studying rotational diffusion. In the small-angle regime |k| L<1, the
dynamic structure factor is written as’?
Saep(k, t) « exp(—Dgk"t — 6D,t), (8.75)
the decay of which is mainly determined by D,,.
8.5 Orientation by an electric field
85.1 The effect of an electric field
An electric or magnetic field can orient the polymer, and measurement of
this process gives information on the rotational motion of polymers in
solution.’*° Using the Smoluchowski equation, we shall consider the
orientation caused by an electric field.
Elementary electrostatics says that if an object with dipole moment p is
placed in an electric field E(t), it feels a torque
N=pXE. (8.76)
The dipole moment p consists of two parts, the permanent dipole p, and304 DILUTE SOLUTIONS OF RIGID RODLIKE POLYMERS
the induced dipole p;. For thin, rodlike polymers the permanent dipole is
always parallel to u, the direction of the polymer, and is written as
Pp = Hu (8.77)
where y is the magnitude of the permanent dipole moment. On the other
hand, the induced dipole moment is written using the parallel and
perpendicular polarizability a and a, as
Pi= a(u + E)u+ w(E —(u-E)u)
= [utes 14+ (q- a,)(uu -%)| -E=a-E. (8.78)
Hence the torque acting on the polymer is written as
N=(p, +p) XE
=puxE+Aa(E-uuxeE (8.79)
where
Aa =a - a. (8.80)
The potential which gives such a torque is
= —pu- E—4Aa(E -u)*. (8.81)
The orientation of the polymers can be studied by measurement of the
dipole moment
(p) = (ws +a +E) (8.82)
or the birefringence. The refractive index tensor A, is written as
Aap = NS ap + A( Ua — 48ap) (8.83)
where A is a constant which includes contributions from both the form
birefringence and the intrinsic birefringence.
8.5.2 Dielectric relaxation
We now calculate the average dipole moment under a given electric field
E(t). The diffusion equation to be solved is
ow w
= =px-[av+ au). (8.84)
To obtain (u), we multiply both sides of eqn (8.84) by wu and integrate
over w.
5 (u) =D, [auu( 2 + 8S. ev) (8.85)ORIENTATION BY AN ELECTRIC FIELD 305
Integration by parts leads to
2 uya 2) -—_ (gu - )
S (uy =D,(( 1) ~ papi Bu BU). (8.86)
As before, Ru gives —2u, and (Ru-RU) is calculated from eqn
(8.81). Hence
2 yy =D,(- ~ 5p (u-B- 8) - 2% ((E-uy'u-E-w)B))
= (u)=0,( 2(u) — pg (EE) — Fo ((E “uu (Eu).
(8.87)
We shall consider the linear response, in which case we can neglect the
third term and evaluate the average in the second term for the isotropic
distribution function of u:
((u-B)u—E)y= | 2 [(u- Bu E]=— 8. (8.88)
Hence
apy 2D,
5 = 2D.u) +30 FE. (8.89)
The solution of this equation is
{ 2,
(u)= f ae’ exp(—20.0--9) Tepe) (8.90)
From eqns (8.82) and (8.90), the dipole moment is given by
i
ip) -2e | dt’ exp(-2D,(t = #*)E(") +7 t 1 gp, (8.91)
B
For an oscillating electric field where
E(t) = Re{E exp(iot)], (8.92)
the dipole moment is calculated as
(p) = Re[a*(w)E exp(iat)] (8.93)
where
2
* __# 1 ay t 2a,
()=FeTitior* 3 654)
with
an,L?
6kyT(In(L/b) — 7)"
@*(w) is called the complex polarizability. Note that the induced dipole
t=1/2D,= (8.95)306 DILUTE SOLUTIONS OF RIGID RODLIKE POLYMERS
Moment shows no frequency dependence. Dielectric relaxation experi-
ments have been carried out for various polymers."
8.5.3 Electric birefringence
Calculation of the electric birefringence is slightly more complicated
because it is a second-order effect in the electric field. Let us consider
the case that a time-dependent, weak electric field E(t) is applied in the z
direction. Straightforward perturbation calculation gives**”*
:
(ub 02) =3D,25 | des expl-6D.(¢~ A IEC?
B
+402(45) | deexpl-60,¢— se)
x ie exp[-2D,(ti — )JE(4). (8.96)
The difference in the refractive index An =,, —f,, is proportional to
(u2- uz). The response of An for various histories is immediately
calculated from eqn (8.96).
(i) Steady state:
If the electric field is constant, eqn (8.96) gives
2_ 2) aa { A, (HL ‘| 2
(u2 — u2) alet+ (4) Ee, (8.97)
Thus
An = KoE”
with
=Aal 425 (4 ‘|
Ko= ase (Zz) 6.98)
which is called the Kerr constant. It consists of a permanent and an
induced dipole term. If the polymer has a dipole moment, their ratio
2
r=
AakgT
is much larger than unity since both 4 and Aq increase in proportion to
the molecular weight.
(8.99)
(ii) Decay: a constant field E is switched off at t=0, ie., E(t)=
E@(—1), and
An(t) = KoE? exp(—6D,t). (8.100)LINEAR VISCOELASTICITY 307
(iii) Rise: a constant field is applied at t = 0, i.e., vO = EQ(t), and
An(t)= KoE*(1- exp(—2D,t) + exp(—6D, .t)).
(8.101)
(iv) Reversal: the direction of a constant field is changed at t = 0, i.e.,
E=-—E@(-—t)+ EQ@(t), and
_3R_
2(R +1) aes 5
An(t)= Koe|1 + a (exp(-6D,1) — exp(~20,1))} . (8.102)
(v) Oscillation: E(t) = E cos(wt). The response for this field is written
as
An(t) = E?[Ka() + Re Kt,(o)exp(2iot)] (8.103)
where
K R
Ka“ oR + 1) G +(atpt 1) (6.104)
and
a a Se
Kul) =F R40 la + 2iwt/3\(1+ ior) * a+ iim (8.105)
where t is given by eqn (8.95).
8.6 Linear viscoelasticity
86.1 Expression for the stress tensor
We now consider the viscoelasticity of a solution of rodlike polymers. As
was discussed in Chapter 3, the stress tensor consists of two terms, the
elastic stress a) and the viscous stress 0.
The elastic stress is related to the change in the free energy of (per
volume) for a virtual deformation dé,, as
6A = CDSE qp- (8.106)
Since the free energy is given by
=v | du(keTY In ¥ + WU) (8.107)
(where v is the number of polymers in unit volume), the change in & is
written as
b= du(kgTS¥ In V + kgTSW + 5WU). (8.108)308 DILUTE SOLUTIONS OF RIGID RODLIKE POLYMERS
The change 5 is calculated using the Smoluchowski equation (8.15).
For the instantaneous deformation, the velocity gradient Kg = d&q/6t
(6t being the duration time of the deformation) dominates the time
evolution of WY, so that
= =-R-(uxXK -uW) during the deformation. _(8.109)
Hence
bY =-R-(uXK-uP)dt=—-R-(ux de -uV). (8.110)
Substituting eqn (8.110) and integrating by parts, we have
b= v felu(ksT(u x Be + u) > OE + (u x be -ul)- RU)
=v faw(-kaT «(ux be +n) +(uX de +u)+ RU). (8.111)
The underlined term can be calculated using eqn (8.36):
Rs (uX 5E + u) = —35Ey6(Uglig — 45g). (8.112)
Hence
65 = vbExg | du'Y (3k eT (Untip — 45 ap) — ( X RU) qtg)
= Vbeap(3kgT (Ualls — 4605) — (UX RU)qutg)). (8.113)
Thus
OGD = 3vkgT (Ugg -45ap) — V((u X RU)aup). (8.114)
If U=0, the stress is given by
o® =3vkp TS xp (8.115)
Sop = (Ualtp - 45ep), (8.116)
which is called the orientational tensor. Note that this stress comes
entirely from the Brownian potential. That the Brownian motion of a rod
can produce a stress may be understood in the following way. Suppose a
rod is placed at the origin along the z axis (see Fig. 8.5). If the rod
rotates by Brownian motion, it will create flow of the surrounding fluid as
in (a) or (6) of Fig. 8.5, depending on the direction of the rotation.
Whichever direction the rod rotates, the fluid around the z axis comes
towards the rod while the fluid in the x — y plane goes away from the rod
as shown in (c). This fluid motion is equivalent to the appearance of the
stress 0,, — 0,,-
The viscous stress is related to the hydrodynamic energy dissipation W
by (see Section 3.8.4)
with
W = kapoSe (8.117)LINEAR VISCOELASTICITY 309
om °°
yf
Mw _ a ws
A a a
@ 0) ©)
Fig. 8.5. Explanation of the stress expression (8.115). The rotation of the rod
causes fluid flow as in (a) or (b). The direction of the rotation is random, but on
average, the fluid moves as shown by thick arrows in (c).
A crude estimation of W is done easily again by neglecting the
hydrodynamic interaction in the shish-kebab model. Under the velocity
gradient «, the rod rotates with the angular velocity @»)=u X (m+).
Hence the velocity of the n-th bead relative to the fluid is
Var = nb (Wo X u — K+ u) = nb([u X (K+ u)] Xu — K+ u) = —nbu(K: un).
(8.118)
The frictional force acting on the segment is F, = €)V,,. Hence the work
done by the frictional force in unit time and unit volume is
WH vd (Ey Vir) =v 3 bn? (na?) = veg (0 ay?)
n n=-N/2
(8.119)
where
NiI2 3
bar= Dont? = TL (8.120)
n=-NR2 4
Hence
OR = VE secKuv (Uy ty taltg (8.121)
Note that in this case, €,., is equal to the rotational friction constant 6,
(see eqn (8.4)). If the hydrodynamic interaction is taken into account, Ey,
is given by £,/2 (see Appendix 8.1). We shall consider only this case,310 DILUTE SOLUTIONS OF RIGID RODLIKE POLYMERS
whence
OS) =F Ekne (atl tty) (8.122)
Thus the stress due to the polymer is
of = off + off
= 3vkpTSyp — v((6X RU)aup) + 5 E,Kyy (Ue Uiplty tly) (8.123)
and the total macroscopic stress in the solution is
up = 09} + 0, (Kap + Kpa)- (8.124)
86.2 Calculation for weak velocity gradient
Let us now calculate the stress tensor under a weak velocity gradient. If
we consider the case U =0, then eqn (8.15) becomes
ow
a DRWV-R+ (ux n-u). (8.125)
To calculate S,, we again multiply u,ug — 54/3 by eqn (8.125) and
integrate over u. The equation is then rewritten by integration by parts as
a
F Sug = [du PDB (usp — $59) + Rlitatig — HBug) «(UX H- W))
(8.126)
The terms in the bracket is calculated directly giving
3
3500 = —6D,(Uqttg — 35 ap) + Kay (Upp) + Key (Upla) — 2Kyy (Ua lipUyptly )
= —6D,Sop +4(Kap + Kpa) + Kau Spu + Kpp Say — 2Kuy (Ugg ly Uy)
(8.127)
To calculate the first order in x, the coefficients of kas can be replaced by
their equilibrium values. Since at equilibrium
Sap =O and (ugtg,t) = i(Sap Suv + San Spy + SavOpn) (8.128)
eqn (8.127) is approximated by
2 Sap = —6D,Sap + 4(Kap + Kpa) — %Kuv(Sap Suv + Sap Spy + Sav5pu)
= -6D,Sap +}(Kap + Kea)» (8.129)LINEAR VISCOELASTICITY 311
(Here the incompressible condition x... =0 has been used.) This is solved
by
Sap(t) = i | dr’ exp[—6D,(t — t')(Kap(t') + Kpa(t')). (8.130)
From eqns (8.123), (8.128), and (8.130), the stress is obtained, to the first
order in Kgg, aS
OC) = SvkaT | ae’ expl-60.(¢— 1 )Ifkap(t?) + Kpa()]
+ 55 S(Kaa(0) + Kpa()- (8.131)
For the shear flow
x ={ro for w=x and B=y,
lo otherwise.
Equations (8.124) and (8.131) give
(8.132)
Ony(t) = mex(t) + ivkyT | ai’ expl-6D-(¢= FeO) + C x(0).
(8.133)
Let us consider two special cases.
(i) Steady shear flow: x(t) is constant, for which eqn (8.133) gives
a(n, + veer A) -
Oxy = (x. + 10D, +39 )K= (n. + &VE,)K. (8.134)
Thus the viscosity of the solution is
2an,L>
= dvl= —_—
nen, +ive, "+ S5Gn(Liby= 1)" (8.135)
The intrinsic viscosity is defined by
1
= lim —(n- n,. 8.136
[n] tim on” ns) (8.136)
where p is the weight of polymer in unit volume, which is expressed in
terms of v and the molecular weight M of the polymer as
M
pax. (8.137)312 DILUTE SOLUTIONS OF RIGID RODLIKE POLYMERS
Thus
[n]= 28 Na
1) 45(In(L/b) — y) M*
(ii) Oscillatory flow: x(t) = Ko Re(e'™). The stress for this flow defines
the complex viscosity n*(w) by
Ory (t) = Ko Re(n*(w)e'). (8.139)
(Note: The complex viscosity is related to the complex modulus by
n*(w) = iwG*(w).) From eqn (8.133), it is easy to show that
(8.138)
ety =n 4 VEBT (__1 1)
10) = 16+ Top, (Fon (8.140)
where
a1. mt?
7 6D, 18(In(L/b) — y)keT” 8.141)
The intrinsic complex viscosity is defined by
1
*(w)] = lim *(@) — ns 8.142)
[n*(@)] tim) 7s) (8.142)
which is given by
3001 1
“(oy =tal(> 1
"l=" Frsioe ta): 143)
Note that [n*()] has a finite value as @—» ©, which arises from the rigid
constraints of the rods. The relaxation time for (n*(@)] is one third of the
relaxation time in the dielectric relaxation. Since L is proportional to M,
[n] and t, depend on the molecular weight,
[n] C x82, =bvkeT( a) . (8.171)
Thus the intrinsic viscosity decreases in proportion to x-**. On the other
hand, rigorous asymptotic analysis”? and the numerical calculation? show
that the intrinsic viscosity decreases as x".
8.8 Effect of flexibility
So far we have been considering the dynamics of rigid rods. Real
polymers have greater or lesser flexibility. This can be modelled by a rod
which has an elastic energy for bending. For the polymer which has
constant contour length, the simplest possible model will be the follow-
ing. Let R(s) be the position of a point on the chain at the contour length
s. The vector
oR
u(s)= =a (8.172)
is a unit vector tangent to the chain. The straight rod corresponds to
u(s)=constant, or du/ds=0. Thus the bending energy must be a
quadratic of Ou/ds. Since u + du/ds =0, the only quadratic form is
=1E fa(2ey (8.173)
where E is a constant. The conformational distribution of the polymer isEFFECT OF FLEXIBILITY 317
thus given by the Boltzmann distribution for this s energy:
warson( St) eoml -Zfa(@)] avo
where
keT
==. 5.17:
I= (8.175)
This is called the Kratky-Porod model, and the length (2A)~'
referred to as the persistence length.
Equation (8.173) indicates the analogy between the change of u(s) of
the Kratky-Porod model and the time evolution of u(t) in rotational
Brownian motion: both processes are Gaussian with the constraint
u’ = 1,> From eqn (8.174) it can be shown that for small s
((u(s) — #(0))) = 4a (8.176)
which indicates that A plays the role of D, in the rotational Brownian
motion. Thus the correlation functions of u(s) are immediately obtained
from the result of eqn (8.43). In particular
(u(s) + u(0)) = exp(—2As) (8.177)
from which the mean square end-to-end distance is calculated as
R= (RL) - RO)? = [sf d5"(u(s)-w("))
1) 0
L s
Loi
=] as [a exp(-2i(s - 5!) =F sal —exp(-2AL)]- 8.178)
The two limiting cases are:
(i) LA>>1 (the random flight limit),
R?=L/A (8.179)
(ii) LA «1 (the rigid rod limit),
=L’, (8.180)
To develop a complete theory for the dynamical properties is difficult,
but various approximate treatments have been proposed. Crudely speak-
ing, the flexibility affects the dynamical properties in two respects.
Firstly, the flexibility changes the transport coefficients. As the size of
the chain decreases with the increase in the flexibility, the diffusion
constant Dg, which is roughly proportional to R;', increases, and the318 DILUTE SOLUTIONS OF RIGID RODLIKE POLYMERS
intrinsic viscosity [n] («R3) decreases. The results of various approaches
are summarized in refs 1 and 36.
Secondly, the flexibility gives a relaxation in the high-frequency region.
For example, according to the rigid rod model, the complex intrinsic
viscosity [)*(@)] approaches the finite value
[a2] = lim [n*(@)] =4[7]. (8.181)
For the flexible polymer, [7*(w)] decreases as the frequency becomes
comparable to that of the bending mode.”* The experimental situation is
summarized in ref. 29 and theoretical analysis is in ref. 37.
Appendix 8.1 Derivation of the Smoluchowski equation by the
Kirkwood theory
In this appendix we derive the kinetic equation based on the Kirkwood
theory described in Section 3.8, and using the model shown in Fig. 8.2.
The position vector of the n-th segment (—N/2 Him * Fon (8.1.3)
From the definition of H,,, (eqn (3.106)) and eqn (8.1.1), it follows that
Hy, = (uu t+ Diam nm (8.1.4)
with
liam = 1/820, |n — ml b. (8.1.5)
In eqn (8.1.4), we neglected the term h,,,. The validity of this approxima-
tion is discussed later.
Now the total force acting on the centre of mass is given by 2, F-
Equating this with that given by the thermodynamic potential we get
> — 2 (kgT in¥+ U). (8.1.6)
a OR
Similarly the balance in torque is written as
> nbu X F, = -—R(kgT In W + U). (8.1.7)SMOLUCHOWSKI EQUATION 319
Our aim is to express V and @ in terms of the quantity on the right-hand
side of eqns (8.1.6) and (8.1.7). For that purpose we solve eqn (8.1.3) for
Fi
B= D(H am Ym = Rr) (8.1.8)
where (H™") um is the inverse of Hymn, i.e.,
2 Ham * (H7")mic = Snel (8.1.9)
From eqn (8.1.4), (H~)am is written as
(Ham = CD am( Ht -3) (8.1.10)
where (h7)am is the inverse of Aam,
2 Fam (h Dnt = Sut (8.1.11)
We substitute eqn (8.1.8) into eqns (8.1.6) and (8.1.7), and use eqn
(8.1.2) to obtain
ZAM (V+ mb Xu) —K-(R + mbu)]= -SlkeT In W+U)
(8.1.12)
> nbu X (H™")am+ [V+ mbw X u— K+ (R+mbu)| = —RksT In W + V).
“m (8.1.13)
Using eqn (8.1.10), the right-hand side is rewritten as
2 Stan 1 +(V-K-R)= -ZTInw+ U) (81.14)
DO agrmb?u x (1-2). (@X u— K+ u) = —R(kyT In ¥ + U).
(8.1.15)
(Here we used the property D,m1(h~")am=0, which follows from
hom = h—n—m-) Defining
=D ams (8.1.16)
¢.=0? > (A74),,,nm, (8.1.17)320 DILUTE SOLUTIONS OF RIGID RODLIKE POLYMERS
we rewrite eqns (8.1.14) and (8.1.15) as
tI -=). (V-K-R)= - Flo in +U) (8.1.18)
C.uXx(oxXu—K+u)=6(w—uXK-u)=—-RksTIn¥+U), (8.1.19)
which is solved for V and o giving
V=k- ee a 2 (kgT In ¥ + U), (8.1.20)
¢
o=uX K-u—1 RkpT In + U). (8.1.21)
Comparing eqn (8.1.20) with eqn (8.24), we have
C120, by =6,/2. (8.1.22)
Substituting this into the conservation equation (8.25), we finally get
= DR: [av += av] - R- (ux k-uP)
kpT
ow A oO
Ds +(i+uu)- [Set eetorl~ ore “RY (8.1.23)
where
D,=ksT/6, and D,=kgT/€,. (8.1.24)
Equation (8.1.23) agrees with eqn (8.26).
Next we obtain the expression for the stress tensor. According to eqn
(3.134),
of =—v >) (FraeRns) + (8.1.25)
Substituting eqns (8.1.1), (8.1.2), and (8.1.10) into eqn (8.1.8), we obtain
the force F, as
F,=DOm(I-Z)- [Vtmbaxu—K-(R+mbu)]. (8.1.26)
Here we consider the case when the total force acting on the polymer is
zero,
DF,=0, (8.1.27)
which gives V=K-R, and
= Le "San( =) « [mboxXu—K-mbul. (8.1.28)SMOLUCHOWSKI EQUATION 321
From egns (8.1.25) and (8.1.28), it follows that
08 = —v >) (kD amnmb> (00 X ut) qt — (K+ U)qUg + 4U,ugu-K-u).
(8.1.29)
Finally, substituting eqns (8.1.17) and (8.1.21) we get
09} = V{(CouvtyRulkaT In ¥ + U))ug) +> 5 EA otgt tly Kyv)- (8.1.30)
The first term on the right-hand side of eqn (8.1.30) is rewritten using
integration by parts (eqn (8.10) to give
(Up aut, R, In VW) = ete Fy
= Cane [du —W)Ry pty
= 3(uglts —35ap). (8.1.31)
Hence
Of = 3vkgT (alg — 45ap) — v{(u X RU) attp)
+3 ” iyy (Ualtpltly)- (8.1.32)
The first two terms represent the elastic stress, and the last term is the
viscous stress.
Finally we calculate the friction constants ¢, and , using eqns (8.1.16)
and (8.1.17). Since Kym decreases quickly with |n—m|, we may ap-
proximate it by
Fram = AS nm (8.1.33)
with
“e In(N/2)
B=2 | don =" (8.1.34)
(Aum = Sam he (8.1.35)
Therefore from eqns (8.1.16), (8.1.17), and (8.1.35)
N _ 4an.Nb _ Ann L
b= a “in(N/2) ~In(L/2b) 6.1.36)
and
m? an (Nby _ an, L?
= 2° { dm “31n(N/2) 3 In(L/2b)° 6.1.37)322
DILUTE SOLUTIONS OF RIGID RODLIKE POLYMERS
References
wn
SN awe
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
24,
27,
. Yamakawa, H., Modern Theory of Polymer Solutions. Harper & Row, New
York (1971).
. Tsvetkov, V. N., and Andreeva, L. N., Adv. Polym. Sci. 39, 95 (1981).
. See for example, Blumstein, A., (ed.) Liquid Crystalline Order in Polymers.
Academic Press, New York (1978); and Ciferri, A., Krigbaum, W. R., and
Meyer, R. B., (eds.) Polymer Liquid Crystals. Academic Press, New York
(1982).
. Kirkwood, J. G., and Auer, P. L., J. Chem. Phys. 19, 281 (1951).
. Jeffery, G. B., Proc. R. Soc. London A102, 161 (1922).
Happel, J., and Brenner, H., Low Reynolds Number Hydrodynamics, Chap.
5. Prentice-Hall, Englewood Cliffs, New Jersey (1965).
Perrin, F., J. Phys. Radium. 5, 497 (1934); 7, 1 (1936).
, Burgers, J. M., Ver. Kon. Ned. Akad. Wet. (1). 16, 113 (1938). See also
Second Report on Viscosity and Plasticity, pp. 113. North-Holland, Amster-
dam (1938).
. Broersma, S., J. Chem. Phys. 32, 1626 (1960).
Yoshizaki, T., and Yamakawa, H., J. Chem. Phys. 72, 57 (1980).
Bremner, H., Int. J. Multiphase Flow 1, 195 (1974).
Rallison, J. M., J. Fluid Mech. 84, 237 (1978).
Beme, B. J., and Pecora, R., Dynamic Light Scattering. Wiley, New York,
(1976).
Maeda, H., and Saito, N., J. Phys. Soc. Jpn 27, 984 (1969); Polymer J. 4, 309
(1972).
Schaefer, D. W., Benedek, G. B., Schofield, P., and Bradford, E., J. Chem.
Phys. 55, 3884 (1971).
Wilcoxon, J., and Shurr, J. M., Biopolymers 22, 849 (1983).
Maeda, T., and Fujime, S., Macromolecules 17, 1157 (1984); Kubota, K.,
Urabe, H., Tominaga, Y., and Fujime, S., Macromolecules 17, 2096 (1984).
Rallison, J. M., and Leal, L. G., J. Chem. Phys. 74, 4819 (1981).
A classical textbook on dielectric relaxation is Debye, P., Polar Molecules.
Dover, New York (1945). A recent review on dielectric properties including
rodlike polymers is given by Mandel, M., and Odijk, T., Ann. Rev. Phys.
Chem. 35, 75 (1984).
Fredericq, E., and Houssier, C., Electric Dichroism and Electric Birefrin-
gence. Oxford Univ. Press, London (1973). A recent review is given by
Watanabe, H., and Morita, A., Adv. Chem. Phys. 56, 255 (1984).
Wada, A., J. Chem. Phys. 30, 329 (1959); 31, 495 (1959).
Sakamoto, M., Kanda, H., Hayakawa, R., and Wada, Y., Biopolymers 15,
879 (1976); 18, 2769 (1979).
. Bur, A. J., and Roberts, D. E., J. Chem. Phys. 51, 406 (1969).
Peterlin, A., and Stuart, H. A., Hand und Jahrbuch der Chemischen Physik,
Vol 81. Akademische Verlagses, Leipzig (1943).
. Benoit, H., Ann. Phys. 6, 561 (1951); Tinoco, I. Jr., and Yamaoka, K., J.
Phys. Chem. 63, 423 (1959).
. Doi, M., J. Polym. Sci. Phys. ed. 19, 229 (1981).
Nemoto, N., Schrag, J. L., Ferry, J. D., and Fulton, R. W., Biopolymers 14,
409 (1975).31.
32.
33.
. Kratky, O., and Porod, G., Rec. Trav. Chim. 68, 1106 (1949).
. Saito, N., Takahashi, K., and Yunoki, Y., J. Phys. Soc. Jpn 22, 219 (1967).
. Yamakawa, H., Ann Rev. Phys. Chem. 25, 179 (1974); Pure Appl. Chem. 46,
37.
REFERENCES 323
. Ookubo, N., Komatsubara, M., Nakajima, H., and Wada, Y., Biopolymers
15, 929 (1976).
. Ferry, J. D., Viscoelastic Properties of Polymers (3rd edn), Chap. 9. Wiley,
New York (1980).
. See for example Bird, R. B., et al. Dynamics of Polymeric Liquids Vol 2,
Chap. 11. Wiley, New York (1977).
Hinch, E. J., and Leal, L. G., J. Fluid Mech. 52, 683 (1972).
Hinch, E. J., and Leal, L. G., J. Fluid Mech. 76, 187 (1976).
Stewart, W. E., and Sgrensen, J. P., Trans. Soc. Rheol. 16, 1 (1972).
135 (1976).
Yamakawa, H., Ann. Rev. Phys. Chem. 35, 23 (1984).9
SEMIDILUTE SOLUTIONS OF RIGID
RODLIKE POLYMERS
9.1 Semidilute and concentrated solutions of rodlike polymers
In Chapter 8, we discussed the dynamics of a single rodlike polymer. Let
us now consider the interaction between the polymers at finite
concentration.
Solutions of slender rodlike polymers of length L and diameter b may
be classified into four concentration regimes (Fig. 9.1). Let p be the
weight of polymers in unit volume of the solution, then the number of
polymers per volume is given by
v =a (9.1)
where M is the molecular weight.
(i) Dilute solution (Fig. 9.12). A dilute solution is defined as one
having a sufficiently low concentration that the average distance between
-13 5
the polymers v~*° is much larger than L, i.e.,
vsy=UL? (9.2)
In such a solution each polymer can rotate freely without interference
by other polymers. The effect of the interaction can be expressed
by a power series expansion with respect to v as in the case of flexible
polymers.
(ii) Semidilute solution (Fig. 9.1b). If v>>¥v, the rotation of each
polymer is severely restricted by other polymers, so that the dynamics of
the polymers will be entirely different from that in dilute solution.
However, the static properties will not be affected seriously until the
concentration reaches another characteristic concentration v2. This is
easily seen if one considers that the polymers are mathematical lines with
no thickness. The equilibrium distribution of such polymers is entirely
independent of each other at all concentrations. Thus the effect of the
interaction becomes important in static properties only for polymers with
finite diameter. Indeed the excluded volume of rigid rods is shown to be
of the order of bL? (see Fig. 9.2)), so that the static properties are
unaffected if vbL? is small. We call the concentration regime
VS VK V2= 1/bL? (9.3)SOLUTIONS OF RODLIKE POLYMERS 325
(@)
Fig. 9.1, Four concentration regimes of rodlike polymers: (a) dilute solution, (b)
semidilute solution, (c) isotropic concentrated solution, and (d) liquid crystalline
solution.
semidilute. (Note that this definition of semidilute is different from that
used for flexible polymers.) In this concentration regime, the effect of
intermolecular interaction can be neglected in the static properties, while
the dynamical properties are completely changed by the constraint that
polymers cannot cross each other. We call such an interaction the
entanglement interaction, although rods do not entangle with each other
literally.
(iii) Concentrated (isotropic) solution (Fig. 9.1c). The static correla-
tion of the polymers becomes important at a higher concentration v > v2.
Here, as will be shown later, polymers tend to orient in the same
Fig. 9.2. Excluded volume between the rods in the direction u and uw’. For a
given position of the rod in the direction u, the centre of mass of the other rod is
not allowed in the parallelepiped region shown in the figure. The volume of this
region is 2bL” |sin ©| where @ is the angle between u and u’.326 SEMIDILUTE SOLUTIONS
direction as their neighbours. If the concentration becomes larger than a
certain critical value v*, which is of the order of 1/bL”, the polymers
align on a macroscopic scale in equilibrium, and the solution becomes an
anisotropic liquid. In the concentration regime
veSvsv" (9.4)
the solution is still isotropic, but the excluded volume interactions among
the polymers are important in both the static and dynamic properties.
Such a solution is called a concentrated isotropic solution.
(iv) Liquid crystalline solution (Fig. 9.1d). The anisotropic solution
above v* is called a liquid crystalline solution. This solution shows a
range of interesting properties quite distinct from those of isotropic
solutions, and will be discussed in the next chapter.
In this chapter we shall discuss the semidilute solution. This is in fact
an ideal system for studying the entanglement effect since the dynamics
of rodlike polymers is much simpler than that of the flexible polymers
and the excluded volume effect is negligibly small.
9.2 Entanglement effect in rodlike polymers
9.2.1 Tube model
In the semidilute region, the dominant interaction is caused by the
topological constraint that the polymers cannot cross each other. This
effect can be treated’” by the same model as that for flexible polymers.
In the situation shown in Fig. 9.1b, the motion along a polymer is almost
aN, a
aN wt B B
OX T a yy
oe os a e
NA Le
Y A
@ ©)
Fig. 9.3. (2) Tube model for the rodlike polymer. (b) The mechanism of rotation
in semidilute solution. The polymer can change its direction when it disengages
from a tube, e.g., by moving from AB to A’B’ and then to A”B”. Reproduced
from ref. 10.ENTANGLEMENT EFFECT IN RODLIKE POLYMERS 327
free, while the motion perpendicular to the polymer is severely limited by
surrounding polymers. Such a characteristic feature of the Brownian
motion can be again represented by a tube which surrounds the polymer
(Fig. 9.3). The tube radius a corresponds to the average distance that the
polymer can move perpendicularly to its own axis without being hindered
by other polymers.
Let us now study how the tube constraint affects the translational and
the rotational Brownian motion.
9.2.2 Translational diffusion
The translational motion is easily analysed. The motion of the polymer
parallel to the tube axis, which is nearly parallel to u, is not hindered by
the tube, so that the parallel component Dj will be nearly equal to the
diffusion constant in dilute solution Djo.t On the other hand, the
perpendicular motion is limited to within the distance a, so that the
perpendicular component D, can be regarded as zero provided the
small-scale motion of order @ is neglected. Thus
Di=Dyo, _D, = 0. (9.5)
9.2.3 Rotational diffusion
The rotational diffusion can be discussed using the following model (Fig.
9.3b). As long as the polymer stays in a certain tube, its direction is
essentially fixed in the direction of the tube axis. If the polymer moves
the distance L/2 along itself, it disengages from the old tube and goes
into a new tube which is generally tilted from the old tube with an angle
of about e~a/L. If the polymer leaves this tube, it again changes
direction by order ¢. Thus overall rotation of the polymer is attained by
repetition of this process. If ty is the average time necessary for the
polymer to leave a certain tube, the polymer repeats the process ¢/tz
times in a time interval ¢. Since the direction u(t) changes about € per
step, the mean square displacement of u(t) is estimated as
({u)-wOP) == (0.6)
Comparing this with eqn (8.1), the rotational diffusion constant is
estimated as
d,=£=(2) Ire @.7)
+ Strictly speaking, Dy is different from Do due to the hydrodynamic interaction among
polymers. However, this effect gives only a logarithmic correction.?*328 SEMIDILUTE SOLUTIONS
T c c
oct fe,
KH
@ ) ()
Fig. 9.4. Constraint release process. The constraint on the test polymer T (thick
line) imposed by a surrounding polymer A can be released if the polymer A
moves away as shown in (6), or a new constraint can be created by an incoming
polymer D as shown in (c).
The time t, is estimated as the time necessary for the polymer to move
the distance L: t= LD, ~ L/D yp. (9.8)
Since L?/Dyo = D7o' (see eqns (8.16) and (8.27)), eqn (9.7) is written as
D ~D.(2) (9.9)
i.e., the rotational diffusion constant in the semidilute region becomes
smaller than that in dilute solution by a factor (a/L)*.
In the above estimation, it is implicitly assumed that the tube is fixed
during the time t,. Actually this assumption is not correct because during
the time t,, the surrounding polymers constituting the tube will also
move a distance of the order of L, and the constraints imposed by them
will be released. At the same time, new constraints are created by
incoming rods (see Fig. 9.4). Thus the tube itself changes with correlation
time tz. However, this effect only changes the numerical coefficients and
does not change the functional form in the result of eqn (9.9) because the
characteristic time and the step length associated with this process are
again tg and e.' (Note that this situation is entirely different from that of
the flexible polymers, for which the release of tube constraints has
negligible effect if the polymer is sufficiently long.)
9.2.4 Estimation of the tube radius and the rotational diffusion constant
To complete the analysis, we express the tube radius a in terms of v and
L. A crude estimation of a is done as follows. We consider a tube of
radius r around the test polymer and calculate the average number N(r)ENTANGLEMENT EFFECT IN RODLIKE POLYMERS 329
)
Fig. 9.5. (a) Tube enveloping the test polymer (thick line). (b) The polymers in
the direction u’ intersect the area AS if their centres of mass are in the region
shown here. The volume of this region is ASL |u’ -s|, where s is the unit vector
normal to the region AS.
of the surrounding polymers which intersect this tube. If r=a, the
number N(r) will be of order unity. Hence a is estimated by
N(a)=1. (9.10)
To calculate N(r), we consider a smail region of area AS on the
surface, and count the number AN(r) of surrounding polymers which
penetrate this area (see Fig. 9.5). Let s be a unit vector normal to this
region. As shown in Fig. 9.5, the surrounding polymers in the direction
u' intersect the region AS if their centres of mass are in the region of
volume LAS |u’ - s|. Hence
AN = | W,(w/WLAS |’ «s| du’ (9.11)
where W,(u’) is the orientational distribution function of the surrounding
polymers.
| First we consider that the distribution of the surrounding polymers is
isotropic. In this case AN is independent of s, and given by
x
AN= vias [ do} |cos 6| sin 8 ="Zas. (9.12)
o
Summation of AN over the entire surface element gives vLS/2, (S=
2zxrL being the total surface area). This is twice the number of polymers330 SEMIDILUTE SOLUTIONS
penetrating the tube since most polymers intersect the surface of the tube
twice. Hence N(r) is obtained as
2
M(t) =3VLS = “=e . (9.13)
From eqns (9.10) and (9.13), it follows that
a=1/vL?, (9.14)
Equation (9.14) indicates that in the semidilute region 1/L?« v «1/bL’,
the tube radius a is much smaller than L, but much larger than b.
Substituting eqn (9.14) into eqn (9.9) we finally get
D, = D(VL?). (9.15)
Thus D, is smaller than Dj» by a factor (vL*)~?.
In the above estimation we neglected the numerical factor entirely. To
write eqn (9.15) as an equality, we have to put in a numerical factor B
D, = BD,(vL’)?. (9.16)
The precise value of the numerical factor B is not known. Various data
suggest that B is rather large (of the order of 10°) as will be shown later in
comparisons with experiments.
If the distribution of the surrounding polymers is not isotropic, the
tube radius becomes a function of u, and so does the rotational diffusion
constant, which will be written as D,(u). This is calculated in Appendix
9.1,
B,(u) = p|4 | du! [us X 1'| ww] (9.17)
where D, is the rotational diffusion constant in the isotropic environment,
eqn (9.16).
9.3 Brownian motion in equilibrium
9.3.1 Time correlation functions
Having obtained the translational and rotational diffusion constants, we
can write down a kinetic equation for the probability Y(R, wu; t) that a
given test polymer is in the configuration (R, u) at time t.+
ay
= =D u-—) V+ R-D,- RV. (9.18)
a Di(t 3p)
OR:
+D, must be placed between the two rotational operator %’s because (i) the equilibrium
distribution of W must be isotropic and (ii) the integral of 9W/dt over u and R must vanish.BROWNIAN MOTION IN EQUILIBRIUM 331
Equation (9.18) describes the Brownian motion of a test polymer in a
given environment whose distribution is specified by W,(u). In the
isotropic solution, D, becomes a constant given by eqn (9.16). The
conditional probability G(R, u, R’, u'; t) that the test polymer which was
in the configuration (R’, w’) at time ¢ = 0 is in the configuration (R, u) at
time ¢ satisfies
8G 9
a7 i aR
Since eqn (9.19) has the same form as eqn (8.46), the time correlation
functions for the test polymer are immediately obtained from the result
of Section 8.4.
yo + DRG. (9.19)
(i) Translational motion.
The mean square displacement of the centre of mass is given by
((R() — R(D)?) = 2Dyt ~ 2Dyot. (9.20)
Thus the diffusion constant Dg is
Da = lim = {(R(0) ~ R(O)??) = Dye (9.21)
Hence the ratio between Dg and Dgo, the diffusion constant in dilute
solution, is given by
De __ Py 1
Doo Djo+2Dio 2° (6.22)
Note that in the rodlike polymers, the entanglement does not change the
molecular weight dependence of Dg. This is in contrast to the result for
flexible polymers, in which the molecular weight dependence changes
from Dgo* M~” to Dg «M7.
(ii) Rotational motion
The correlation function of u(t) is calculated as
(u(t) - u(0)) = exp(—2D,t) (9.23)
Thus the rotational relaxation time t, = 1/2D, is larger than that in dilute
solution by
1, lt = (vLP/B. * (9.24)
This shows a strong entanglement effect.
It must be mentioned that in eqns (9.21) and (9.23) we have neglected
the motion inside the tube. As in the case of flexible polymers, the tube332 SEMIDILUTE SOLUTIONS
constraint is ineffective in a very short time-scale, so that
{RO — R))’) = 6Deot (9.25)
and
(u(t) - u(0)) = exp(—2Dyot). (9.26)
These equations hold if the time-scale is less than
a ¢
le 9.27
Da De’ 27)
In Section 9.6, we shall show how to describe both short-time (¢ < t.) and
long-time (t¢ > t,) behaviour.
9.3.2 Dynamic light scattering
Now we shall consider the dynamic light scattering in semidilute solution.
In general the light scattering from polymer solutions of finite concentra-
tions includes both the intramolecular and the intermolecular inter-
ferences. In the semidilute regime, however, even though the average
distance between the polymers is smaller than the polymer size, the
intermolecular interference is negligible because there is no correlation
between the configurations of different polymers. Therefore the dynamic
structure factor is again given by eqn (8.59)
ah, 1)= fan {aw (S8E2)) (RA) Gu, ws Naga!) (9.28)
K-u
where
K=kL/2 (9.29)
and
3 G,(u, u’; t) = (D,R? — Dy(k > u))G,(u, u'; 1). (9.30)
The distinction between the dilute and semidilute regimes is in the
magnitude of the ratio between the translational term and the rotational
term in the equation for the Green function G,(u, u’; t). Let r be defined
by:
r= Dgk?/D,. (9.31)
In dilute solutions, r is given by
19 = Dgok"/Dyo = (KL? (9.32)
which is usually of the order of unity. In the semidilute regime, on the
other hand, r can be quite large because
1 = Dgk?/D, = (kLP(vL?? >> (RLY for vL?>> 1. (9.33)ORIENTATION BY EXTERNAL FIELDS 333
To see the characteristic feature of the semidilute regime, let us
consider the extreme case of r = © (i.e., D, = 0). Equation (9.30) is then
solved by
G,(u, u’; t) = 6(u — u’)exp(—Dj(k + u)’t). (9.34)
Hence
atk, = [oH (OY exp(—Dyck- 0)
K|
-| a(t 1) exp(—D 78). (9.35)
Thus the dynamical structure factor has a very broad distribution of
decay rates ranging from 0 to D,k’.
Experimental study of dynamic light scattering has been carried out by
several groups.** Though some qualitative features of the above predic-
tions are indeed observed, clear interpretation of experimental results
has been hindered by various factors inherent in real polymers, such as
polydispersity, partial flexibility, and association. These effects will be
discussed later in connection with rotational motion. An important factor
which will not affect the rotational motion, but will be important in the
dynamic light scattering is the effect of weak, long-range repulsive force.°
As in the case of flexible polymers, such interaction tends to keep the
segment density homogeneous, and increases the decay rate of g(k, t)
with the concentration. This is indeed observed in several systems.>*
9.4 Orientation by external fields
9.4.1 Linear regime
So far we have been considering the motion of a test polymer in an
isotropic environment. We now consider a slightly different problem:
how does the orientational distribution function of polymers change
under external fields such as a potential field U,(u) or a velocity gradient
x. Let W(u; t) be the probability that an arbitrarily chosen polymer is in
the direction w. Since each polymer feels the external field as in eqn
(8.15), the time evolution of Y(u; t) can be described by*"?
ow
A= R-D, [avs auyy] - R-[ux(K-wV)]. (9.36)
An important point here is that in this problem, the environmental
distribution function , is the same as Y(u; ¢) itself, and D, is now given334 SEMIDILUTE SOLUTIONS
by
4 2
b,=D|* | du'V(u'; t) |u xu'|| . (9.37)
Equations (9.36) and (9.37) give a nonlinear equation for W. The
nonlinearity indicates a mean field character of the present theory: it
comes from identifying the distribution of the surrounding polymers with
that of the test polymer. Equation (9.36) is thus different from the usual
Smoluchowski equation, which is always linear in WV.
In the linear response regime, however, the nonlinearity of the kinetic
equation is not important because there D, can be replaced by D, since
the change in D, appears only in the higher order perturbation. Therefore
the linear response function is given by the same form as that in dilute
solution except that D, is much smaller than D,o. For example, consider a
todlike polymer which has permanent dipole moment and isotropic
polarizability (a = a, =«a.). The complex polarizability and the dyna-
mic Kerr constant (per polymer) are given in the same form as eqns
(8.94) and (8.104),
wool
(= STi + ior (9.38)
aKo_ 1
Ke) =F 7 Fon? (9.39)
with
+=1/2D,. (9.40)
The rotational diffusion constant can be obtained from these expressions.
9.4.2 Nonlinear regime—tube dilation
The nonlinearity in the kinetic equation becomes important if the
external perturbation is large. In this case precise mathematical analysis
becomes quite difficult. A convenient approximation is to replace D, by
the average D,
2
b,=B,=0]* few du"B(u; E(u’; 1) \uxw'l] . @41)
Since D, is independent of u, the kinetic equation can be written as
pe. [aw + au] —R-[ux(w-w)) (9.42)
ot kgT
This can be handled much more easily than eqn (9.36) (see refs 2 and
10). Though the approximation (9.41) is crude, it takes into account theORIENTATION BY EXTERNAL FIELDS 335
(a) (b)
Fig. 9.6. Tube dilation. The tube radius increases when the surrounding polymers
are oriented in the same direction as the test polymer.
following effect. As the polymers orient in the same direction, the
average diameter of the tube becomes larger, so that the average
rotational diffusion constant increases (see Fig. 9.6). This effect is called
the tube dilation.
The tube dilation may be seen in, for example, the relaxation of
birefringence from the highly oriented state; the initial relaxation rate is
larger than the final one. A theoretical analysis of this effect has been
done in ref. 2
9.4.3 Experimental study of the rotational diffusion constant
The rotational diffusion constant has been measured by relaxation of the
Kerr effect,'? and by dynamic light scattering.” The experimental
results are in accordance with the theoretical predictions
D,|Djo= B(VL*)? « pM, (9.43)
or, since Dj «In(M)/M°,
D,« p~?M~" In(M). (9.44)
However, the absolute magnitude of D, has turned out to be quite large:
experimental values of 6 range from 10° to 10*. Such large values of B
have also been found by a computer simulation for thin polymers (with
no thickness).'*15 Figure 9.7 shows the concentration-dependence of D,
obtained by dynamic electric birefringence and computer simulation. The
solid line indicates eqn (9.43) with B = 1.3 x 10°, which was obtained by
Hayakawa et al."? by detailed study of the tube statistics. Though the
preciseness of this value can be questioned, it is clear that the numerical336 SEMIDILUTE SOLUTIONS
1 © 00.8 ge oe
Bh
LN
é a
o \
01 y
X
\
0.01 pe woh
04 1 10 100 1000
vl?
Fig. 9.7. Reduced rotational diffusion constant D,/D, is plotted against reduced
concentration vL*. Filled circles: the experimental results’? of dynamic electric
birefringence of poly(y-benzyl-L-glutamate, molecular weight ranging from 7.3 x
10* to 1.5 x 10°). Open circles: the result of computer simulation.’* The solid line
is the theoretical result.’* (Courtesy of Prof. Hayakawa, Tokyo University.)
factor is large and that the semidilute regime starts at rather large values
of vL.
As has already been mentioned, precise comparison with experiments
is hindered by various effects such as polydispersity and bending of
polymers. Quantitative theory for these effects is difficult, but the
qualitative aspects have been discussed in refs 16-18
9.5 Viscoelasticity
9.5.1 Expression for the stress tensor
To discuss the viscoelastic properties in semidilute solutions, we first
consider the expression for the stress tensor. This is obtained from the
principle of virtual work as in Section 8.6.1.
In the semidilute region, the expression for the free energy is the same
as that in the dilute solution since the excluded volume effect is
negligible:
A= J du'U(kpT InW + U,). (9.45)
The change in W by instantaneous deformation dé, is again given by
OW =-R- (ux de-u). (9.46)VISCOELASTICITY 337
Hence the elastic stress is given by precisely the same form as eqn (8.114)
0G = 3vkgT (uglip — 40a) — V{(u X RU, )atip )- (9.47)
The viscous stress is obtained from the hydrodynamic energy dissipa-
tion under a given velocity gradient. If we assume that the hydrodynamic
interaction is completely screened, this is calculated in the same manner
as in eqn (8.119):
OG) = VEcKuy (Up ly Ua lp ) (9.48)
with
an, L3
be (9.49)
On the other hand if we neglect the hydrodynamic screening entirely, €..,
is given by the same formula as eqn (8.122)
__an,L?
~ 6In(L/b)*
The actual form will be between the two. Indeed the effective medium
theory‘ indicates that
Cote (9.50)
fog numbers? (9.51)
se In(1/vbL?)* .
It is important to note that €,, is not affected as seriously as D,. This is
because the viscous stress reflects very fast motions, for which the tube
constraint is not effective.+
For simplicity we proceed using eqn (9.50). The stress tensor is given in
precisely the same manner as for dilute solutions (see eqn (8.123)).
Ong = 3VK pT (Ugttg — 36g) — V{(u X RU, )altp )
F VE acKuy (Up ly alls) + Is(Kap + Kea)» (9.52)
9.5.2 Linear viscoelasticity
To calculate the stress, we have to solve
aw
FRB, RV-R- UX Kw). (9.53)
7 It has been suggeted””° that better agreement with some experiments can be obtained if
baz is replaced by kpT ID, = n,L(vL*)*/In(L/b). However, this agreement is perhaps
superficial, caused by non ideal effects of real polymers such as molecular weight
distribution, or flexibility. Theoretically there is no reason to believe that f,,, is given by
k,TID,-338 SEMIDILUTE SOLUTIONS
In the linear viscoelastic region, DB, can be replaced by D,, and the
solution is obtained in the same way as in Section 8.6.
In steady-state shear flow with shear rate x, the terms in eqn (9.52) are
given by (the case of U, =0 is being considered)
vkpT _ a (vL°)
EB) = =
o®) = 3vkpT (ugly) 10D, K 30B in(L/b) nsK; (9.54)
3
Vv) — 242) e eck a _VE
OG) = Vbsur( wey) K = HVE uK ~ op inhi ™ (9.55)
and
of = nx.
Hence their ratio is
oF): oS): $9 = B-'(vL?)?: (vL?):1. (9.56)
In the semidilute regime vL*>>1, the contribution of the elastic stress is
much larger than the viscous stress and the solvent stress. Thus in the
ideal semidilute region of 1/L?« v<« 1/bL?, the stress can be written as
Cap = 3Vk pT (Ugtts — 45ap) = 3vk5TSap- (9.57)
Since the characteristic time of S,g is very large (being of the order of
1/D,), the viscoelastic behaviour becomes quite pronounced in the
semidilute region.
From egns (9.54) and (9.57), the steady-state viscosity is given by
_vkeT_ (vL39°
= F0D, 306" in(L/b)
which depends on the molecular weight and concentration as
n= p°M*/In(N). (9.59)
The strong molecular weight dependence of 7 is in accordance with
experiments,”"?> though the precise exponent is difficult to extract
because of the nonideal effect discussed previously. The numerical factor
B can also be obtained from the viscosity and has again turned out to be
rather large ranging from 10° to 10*. (Comparison between the theoreti-
cal prediction and the experimental results are given in refs 1, 19, 24, and
26) In some cases the concentration dependence of 7 is stronger than
predicted by eqn (9.59) at higher concentration.~%75 One possible
explanation for this is the rod jamming effect.””
The complex viscosity defined by eqn (8.139) is calculated as
3pRT_ ¢
5M 1+iot
(9.58)
n*(@) = (9.60)VISCOELASTICITY 339
16?
ms gt gy
S O6h
04
o.2t-
bev eit ii ei
0 05 1.0 15
Fig. 9.8. Real part of the complex viscosity n'(w) of poly(y-benzyl-L-glutamate)
in m-cresole is plotted against concentration p at various frequencies (from top to
bottom, 0, 2.2, 6.6, 20, 58, 202, and 525 KHz). Here n, is the viscosity of the pure
solvent. Reproduced from ref. 28.
with
t= 1/6D,. (9.61)
Figure 9.8 shows the real part of the complex viscosity at various
frequencies.% As the concentration increases, the viscosity at low
frequency increases sharply, while the viscosity at high frequency
increases only in proportion to the concentration. This is in accordance
with eqns (9.54) and (9.55).
9.5.3 Nonlinear viscoelasticity
Since D, becomes small in the semidilute region, the nonlinear visco-
elasticity becomes quite important. To handle the nonlinear visco-
elasticity, we need the full solution of eqn (9.53). This has been done by340 SEMIDILUTE SOLUTIONS
solving eqn (9.53) numerically’°”® with the approximation (9.41). It turns
out that the qualitative features of the nonlinear viscoelasticity are quite
similar to those of flexible polymers, for example:
(i) Shear thinning: In the steady shear flow, the viscosity (x), the first
normal stress coefficient Y,(x), and the absolute value of the second
normal stress coefficients W(x) (which is negative) all decrease with the
shear rate. The characteristic shear rate for the shear thinning is about
D,.
(ii) Stress overshoot: When a constant shear flow is started, the shear
stress shows overshoot if the shear rate is sufficiently high.
(iii) Elongational viscosity: The elongational viscosity first increases
slightly with the elongational rate and then decreases. These features
agree at least qualitatively with the observed behaviour.74-?43°%
Unlike the case of flexible polymers the constitutive equation cannot
be written in a simple closed form unless we use the decoupling
approximation. Since the equation for Sg is given by the same equation
as that in dilute solution, eqns (8.149) and (9.57) give a closed equation
for Oz:
a = 1
> Cap = —6D,0ng + Kay Ogu + Key Fay + 3 G.(Kap + Kpa)
at
— 20 ,0Krv( P+ 3808) (9.62)
where
G, = 3vkpT. (9.63)
The qualitative features described above can be checked by this
approximate constitutive equation.
9.6 Short time-scale motion
9.6.1 Chopstick model
So far we have neglected the small-scale fluctuation that the polymer
makes inside the tube. To include such motion, we consider the following
model™ (see Fig. 9.9). We separate the direction of the polymer and the
direction of the tube, and consider two vectors wu representing the
direction of the polymer and a the direction of the tube. The diffusion
constant of n is D,, while the diffusion constant of w is Dj» since the
polymer can move freely inside the tube. The condition that wu is
fluctuating inside the tube is represented by the coupling potentialSHORT TIME-SCALE MOTION 341
Fig. 9.9. The chopstick model.
between u and a.
U(u —n) = Su —ny. (9.64)
Equation (9.64) guarantees that at a ticum, the deviation between u
and n is of order e:
(aay
Jomese(—“s2")
The kinetic equation is now given for the two-vector distribution function
®(u, n; t) as
(@-ay)= =2e, (9.65)
= DI: (20+; RyU-(u— ”)
+ Do, * (2. o+e = Ra Ue(u — n) + U .(u))) (9.66)
where ®, and &, are written as
a a
R, =ux oa" R, =nX oa (9.67)
and for simplicity the tube dilation is neglected.
The model described by eqn (9.66) is easy to visualize: the two vectors
u and m move together like a pair of chopsticks. The external field,
represented by the potential U,(u), affects the rapidly moving vector u,
which drags the slowly moving vector a through the coupling potential
U.(u — n).342 SEMIDILUTE SOLUTIONS
An important property of the kinetic equation (9.66) is that at
equilibrium, the distribution of u is not affected by the vector n. In fact,
the equilibrium solution of eqn (9.66) is
®,,(u, n) « exp[—(U.(u) + U.(u — n))/kpT). (9.68)
Hence the equilibrium distribution of u is
Wea(u) = [dn q(u, 9) exp[-U-(u)/koT] (0.68)
which is the same as the isolated polymer without the tube constraint.
This must be so since the entanglement interaction does not affect the
static properties.
9.6.2 Local equilibrium approximation
Although the kinetic equation (9.66) is conceptually simple, it is not
easily handled mathematically. However, if we focus our attention on
slow motion, a simple treatment is possible.
If the external field changes slowly compared to t,~€7/D,o, the
distribution of the vector u can be assumed to be in a local equilibrium
for given n. Then ®(u, n; t) can be written as
O(u, n; 1) = Bn; t)peq(u, 2) (9.70)
where ., represents the local equilibrium distribution of u for given n,
ice.,
_Usu—n) + Ulu)
elu ye ar on)
“~ fou [Hea + Ho) .
xP ket
arf _ Ulu —n) + U,(u) - O(n)
= exp| ee | OD)
Here
O.(n) = —kyT in fu exo| - Rigen (9.73)
B
and Weq(u, m) is normalized such that
fama n)=1. (9.74)
To determine W(n;¢) we substitute eqn (9.70) into eqn (9.66). The