12.03.2015 Views

View - Martin Kröger - ETH Zürich

View - Martin Kröger - ETH Zürich

View - Martin Kröger - ETH Zürich

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

A thermodynamically admissible reptation model<br />

for fast flows of entangled polymers. II. Model predictions<br />

for shear and extensional flows<br />

Jiannong Fang, <strong>Martin</strong> <strong>Kröger</strong>, and Hans Christian Öttinger a)<br />

Department of Materials, Institute of Polymers, <strong>ETH</strong> <strong>Zürich</strong> and Swiss F.I.T.<br />

Rheocenter, CH-8092 <strong>Zürich</strong>, Switzerland<br />

(Received 30 November 1999; final revision received 11 July 2000)<br />

Synopsis<br />

Numerical predictions of a previously proposed thermodynamically consistent reptation model for<br />

linear entangled polymers are presented for shear and extensional flows. Comparisons with<br />

experimental data and two alternative molecular-based models are given in detail. The model<br />

studied in this paper incorporates the essence of double reptation, convective constraint release, and<br />

chain stretching, and it avoids the independent alignment approximation. Here, no use is made of<br />

the ingredient of anisotropic tube cross sections of the previously proposed model. Simulation<br />

results reveal that the model at a highly simplified level with few structural variables, i.e., four<br />

degrees of freedom, is able to capture qualitatively all features of the available experimental<br />

observations and is highly competitive with recently proposed models in describing nonlinear<br />

rheological properties of linear entangled polymers. © 2000 The Society of Rheology.<br />

S0148-60550000406-5<br />

I. INTRODUCTION<br />

Stress–strain relationships for polymer melts are the main ingredient for the flow<br />

simulation of polymer processing such as injection molding, film blowing, and extrusion.<br />

The reliability and accuracy of these simulations depends crucially on the constitutive<br />

equations which describe the nonlinear viscoelastic properties of the underlying model<br />

polymer. Although the closed-form phenomenological models, such as K-BKZ equations,<br />

have been widely used in research and commercial codes, their degree of success is<br />

limited because of a lack of physical ingredient on the molecular level. For the purpose<br />

of realistic modeling, molecular-based models are uniquely suited. A molecular model<br />

was elaborated by Doi and Edwards 1978a, 1978b, 1978c, 1979 who extended the<br />

reptation idea introduced by de Gennes 1971 to a tube idea in order to describe the<br />

viscoelastic behavior of entangled polymers in the presence of ‘‘obstacles.’’ Within the<br />

tube and reptation pictures, the complex entanglement interaction between polymer<br />

chains has been treated in a rather direct approach, i.e., each chain in the polymer system<br />

is equivalent to a chain restricted to one dimensional motion so called ‘‘reptation’’ in a<br />

confining tube, except for its two ends which can move in any possible direction. In<br />

addition to the reptation mechanism, Doi and Edwards originally assumed instantaneous<br />

and complete chain retraction, affine tube deformation by the flow, and independent<br />

a Author to whom correspondence should be addressed.<br />

© 2000 by The Society of Rheology, Inc.<br />

J. Rheol. 446, November/December 2000 0148-6055/2000/446/1293/25/$20.00<br />

1293


1294 FANG, KRÖGER, AND ÖTTINGER<br />

alignment of tube segments. By doing so, they obtained a closed-form constitutive equation<br />

which only involves the second moment of the orientation vector for a tube segment.<br />

For highly entangled, linear polymers, the original Doi–Edwards DE model has been<br />

extended to incorporate chain contour length fluctuations Doi 1983; Ketzmerick and<br />

Öttinger 1989 and constraint release due to the motion of the surrounding chains so<br />

called ‘‘double reptation’’ Tsenoglou 1987; des Cloizeaux 1988. The combination<br />

of these two effects lead to a refined description of the linear viscoelastic properties<br />

O’Connor and Ball 1992, however, the model is much less successful for the nonlinear<br />

properties. The major experimental observations that the original DE theory fails to<br />

describe in the nonlinear regime are the following:<br />

1 There exist irreversible effects in double-step strain experiments with flow reversal<br />

Osaki and Kurata 1980; Osakia et al. 1981; Venerus and Kahvand 1994a, 1994b;<br />

Brown and Burghardt 1996.<br />

2 Over a wide range of shear rates ˙ above the inverse disentanglement time 1/ d ,<br />

the steady shear stress is nearly constant for very highly entangled ones. The first normal<br />

stress difference N 1 increases more rapidly with shear rate than does the shear stress over<br />

the same range of shear rates. The slope of N 1 vs ˙ increases as the molecular weight<br />

decreases Bercea et al. 1993; Kahvand 1995.<br />

3 The steady-state shear viscosity of different molecular weights merge into a single<br />

curve in the high shear rate, power-law regime Stratton 1966.<br />

4 The shear stress xy shows transient overshoots in the startup of steady shear flow<br />

at low shear rates. The strain p at which the maximum in the overshoot occurs increases<br />

with shear rate at high rates Pearson et al. 1989; Menezes and Graessley 1982;<br />

Kahvand 1995.<br />

5 The first normal stress difference exhibits transient overshoots in the startup of<br />

steady shear flow at moderate shear rates Pearson et al. 1989; Menezes and Graessley<br />

1982; Kahvand 1995.<br />

6 The rate of stress relaxation following cessation of steady shear flow is shear rate<br />

dependent Attane et al. 1985; Menezes and Graessley 1982; Kahvand 1995.<br />

7 The steady-state extinction angle decreases more gradually with shear rate than<br />

predicted by the DE model Mead and Larson 1990; Kahvand 1995.<br />

8 The transient extinction angle shows an undershoot at the startup of steady shear at<br />

high shear rates; it also shows an immediate undershoot when the shear rate is suddenly<br />

decreased after a steady state has been reached, finally it reaches a higher steady-state<br />

value Mead 1996; Oberhauser et al. 1998.<br />

9 Steady-state values of the dimensionless uniaxial extensional viscosity are nonmonotonic<br />

functions of extension rate Munstedt and Laun 1981; Ferguson et al.<br />

1997.<br />

In order to improve the situation, many attempts of modifying the original DE model<br />

have been made during the last years. Several physical effects have been found to be<br />

important for more realistic modeling of nonlinear properties of entangled polymers. A<br />

short summary of the important effects found so far is given in Secs. IA–IE.<br />

A. Avoiding independent alignment „IA…<br />

Recognizing that the large discrepancy between model prediction and experimental<br />

data in double step strain with flow reversal is caused by the IA approximation. Doi<br />

1980a, 1980b, and Doi and Edwards 1986 made a detailed analysis for this situation<br />

and tried to derive a constitutive equation without using the IA approximation. Again, the<br />

instantaneous-chain-retraction assumption was employed in their derivation. It was<br />

shown that the model is able to correct the previous discrepancy whenever the time


THERMODYNAMICALLY ADMISSIBLE REPTATION MODEL<br />

1295<br />

interval between the two applied step strains is much larger than a characteristic time s ,<br />

called the retraction time. Marrucci 1986 and Marrucci and Grizzuti 1986 also showed<br />

that the model without IA predicts the Weissenberg effect correctly. In an attempt to find<br />

a thermodynamically admissible formulation, the reptation model without IA has been<br />

reformulated by Öttinger and Beris 1999 in the general equation for the nonequilibrium<br />

reversible-irreversible coupling GENERIC formalism of nonequilibrium thermodynamics<br />

Grmela and Öttinger 1997; Öttinger and Grmela 1997. For a comprehensive<br />

introduction see Öttinger 1999a. The model has been shown to be thermodynamically<br />

consistent after modifying the production term and introducing an additional term in the<br />

extra stress tensor. A consistent model with a uniform monomer distribution along the<br />

chain was proposed by Öttinger 1999b.<br />

B. Double reptation<br />

Tsenoglou 1987 and des Cloizeaux 1988 derived a successful mixing rule for<br />

polydisperse melts based on the idea of ‘‘double reptation,’’ which represents the relaxation<br />

mechanism for the tube that does not arise from motion of the probe chain, but<br />

rather motion from the surrounding chains. Öttinger 1994 has considered a reptation<br />

model in which this additional constraint release mechanism is mimicked through a noise<br />

term in the time-evolution equation for the orientation of inner chain segments. By doing<br />

so, he recovered the correct mixing rule in the linear regime and extended the idea to the<br />

nonlinear regime. But, the prescribed relaxation spectrum is not fully recovered due to<br />

the lack of contour-length fluctuations in the model. The model predicts an improved<br />

power-law index of 4/3 for the steady-state viscosity in shear flow instead of 3/2 for<br />

the DE model. Accordingly, the shear stress versus shear rate curve still exhibits a<br />

maximum when only double reptation is considered.<br />

C. Chain stretching<br />

Marrucci and Grizzuti 1988 extended the DE model to allow for chain stretching and<br />

predicted steady-state properties. An effect is predicted to result in elongational flows,<br />

giving rise to an expected upturn of the elongational viscosity, but surprisingly, there is<br />

no effect for steady shear flows, in particular, the extended model does not improve the<br />

quality of the model prediction for the power-law index of the steady-state shear viscosity.<br />

Later, Pearson et al. 1991 studied the same type of model with chain stretching in<br />

transient situations. They found that both the shear stress and the first normal stress<br />

difference overshoot in the startup of steady shearing flows and the times required to<br />

reach these maxima are independent of the shear rate. Mead and Leal 1995 and Mead<br />

et al. 1995 presented a comprehensive study of the DE model by including chain<br />

stretching and a nonlinear finitely extensible spring law. Numerical predictions were<br />

given for steady two-dimensional flows with a continuously varying degree of extensional<br />

and shear character. The results revealed that significant steady state stretch is<br />

predicted as the flow becomes increasingly extensional in character.<br />

Although chain stretching is important for correcting some of the failings such as an<br />

overshoot in N 1 , it doesn’t solve a long-standing problem in the DE model, namely, the<br />

excessive shear thinning of the viscosity at high shear rates associated with a maximum<br />

in shear stress followed by a region in which shear stress decreases with shear rate<br />

asymptotically as ˙ 0.5 , which leads to constitutive instabilities in shear flow. The<br />

reason for this problem is that, in fast shear flow, the model predicts that the tube<br />

segments become highly oriented in the flow direction and hence present a very slim


1296 FANG, KRÖGER, AND ÖTTINGER<br />

profile to the flow. As a result, the flow ‘‘loses its grip’’ on the molecules, leading to<br />

anomalously low friction and hence low viscosity.<br />

D. Convective constraint release „CCR…<br />

Marrucci 1996 and Ianniruberto and Marrucci 1996 proposed a CCR mechanism<br />

which removes the problem just mentioned above. They proposed a model for which,<br />

under flow conditions, relaxation of chain orientation occurs by two mechanisms. One of<br />

them is ordinary diffusion reptation and double reptation due to thermal motion, which<br />

of course also takes place in the absence of flow. The second mechanism is CCR, i.e., the<br />

topological obstacles on a probe chain are renewed through the relative motion among<br />

chains due to chain retraction. In fast flow situations, this mechanism leaves the chain<br />

much freer to relax than is possible only by the previously described mechanism, and<br />

hence prevents the tube segments from becoming highly oriented in the flow direction.<br />

E. Anisotropic tube cross sections<br />

Ianniruberto and Marrucci 1998 introduced the idea that during deformation an<br />

initial circular tube cross section may become elliptical. They derived the corresponding<br />

expression for the stress tensor, but did not present a time-evolution equation for the tube<br />

cross section in flow. It was shown that the idea of anisotropic tube cross section has an<br />

important influence mainly on the ratio of normal-stress difference in shear flow. In view<br />

of the well know intimate relation between the time evolution of the structural variables<br />

and the stress tensor expression implied by various approaches to nonequilibrium thermodynamics,<br />

Öttinger 2000 developed a thermodynamically admissible reptation<br />

model with anisotropic tube cross section and the constraint release mechanisms associated<br />

with double reptation and CCR. For that model, he proposed relationships between<br />

the ratio of normal-stress differences and the mean-square curvature of the tube cross<br />

section in shear flow.<br />

Very recently, reptation models incorporating all the well-established phenomena except<br />

for anisotropic tube cross sections have been formulated by two groups, based on a<br />

full-chain stochastic approach suitable for computer simulations by Hua and Schieber<br />

1998 and Hua et al. 1998, 1999 and on rather complicated coupled integraldifferential<br />

equations by Mead et al. 1998. It is encouraging that these reptation models<br />

can quite successfully reproduce the experimentally observed rheological behavior in a<br />

large number of flow situations. In a previous work by Öttinger 1999b, which is referred<br />

to as part I of the present work, a new reptation model including anisotropic tube<br />

cross sections, chain stretching, double reptation, and CCR, while avoiding the IA approximation,<br />

has been developed under the guidance of nonequilibrium thermodynamics.<br />

Two versions of the model, referred to as ‘‘uniform’’ and ‘‘tuned,’’ have been proposed.<br />

The purpose of this paper is to give a detailed model evaluation of the recommended<br />

‘‘uniform’’ model in shear and extensional flows by numerical simulations. Because the<br />

experimentally observed features listed in the beginning of Sec. I can be explained<br />

without considering anisotropic tube cross sections their implementation seems to be<br />

nontrivial to us, and they weakly affect the ratio of predicted normal stress differences for<br />

which experimental data is rarely available, the variable Q representing a tube cross<br />

section in the previously proposed model has been omitted here. On this simplified level,<br />

the model has only four degrees of freedom, which will be recalled in Sec. II. Note that<br />

our model takes into account only a single relaxation time reflecting the linear viscoelastic<br />

response of the material. This limitation can be released by considering


THERMODYNAMICALLY ADMISSIBLE REPTATION MODEL<br />

1297<br />

a life span distribution for segments, being deduced from a measured relaxation modulus<br />

as described by Öttinger 2000. In this work, we focus entirely on the basic nonlinear<br />

behavior.<br />

II. MODEL DESCRIPTION<br />

The structural state variables chosen in Part I are the configurational distribution<br />

function f (u,s,r) and the stretching ratio of the chain contour length (r) L/L 0 ,<br />

where u is a unit vector describing the orientation of a tube segment, s is the position<br />

label of a tube segment in the interval 0,1 the values s 0 and s 1 correspond to<br />

the chain ends, r is the space position vector, and L 0 denotes the equilibrium contour<br />

length of the chain. After determining all the state variables the hydrodynamic variables<br />

and the above structural state variables, the model has been formulated in the GENERIC<br />

formalism by constructing its ‘‘building blocks’’ step by step. Here we summarize the<br />

final time-evolution equations for the structural variables.<br />

The equation for the chain contour length stretching and relaxation reads<br />

D<br />

Dt<br />

˙ tot ˙ convect ˙ dissip ,<br />

1<br />

where D/Dt is the material time derivative, and the total stretching rate is split into<br />

convective and dissipative contributions<br />

˙ convect :¯,<br />

2<br />

where is the transpose of the velocity gradient tensor, ¯ is the symmetric secondmoment<br />

orientation tensor defined by<br />

¯ <br />

0<br />

1<br />

ds d 3 ufu,s,ruu.<br />

3<br />

The dissipative contribution reads<br />

˙ dissip 1 s<br />

c<br />

3Z 1,<br />

4<br />

where Z the number of entanglement segments per chain which is given by M/M e M is<br />

the molecular weight of the chain and M e is the average molecular weight between<br />

entanglement points along one chain, s is the characteristic stretching time, and c() is<br />

the effective positive spring coefficient<br />

c 3Z 2<br />

max1<br />

2<br />

max 2 .<br />

5<br />

The spring coefficient is derivable from an entropy expression of the form<br />

3 2 Z ln 2 2<br />

max 1ln 2<br />

max 2<br />

2 ,<br />

max 1<br />

6<br />

such that the spring force F() (1)c() vanishes at 1, and diverges for<br />

→ 0 and → max , respectively, while for small extensions Hookean spring behavior,<br />

and for large extensions FENE type spring behavior is recovered.


1298 FANG, KRÖGER, AND ÖTTINGER<br />

TABLE I. Values for the maximum possible stretching ratio max via Eq.<br />

6, the entanglement molecular weight M e , the average end-to-end distance<br />

of the entanglement segment at equilibrium d t , and the length of a<br />

Kuhn step l K for three important melts at 140 °C.<br />

Melts M e gmol 1 d t (Å) l K (Å) max<br />

PS 13000 76 16.2 4.7<br />

HDPE 860 33 12.9 2.6<br />

PP atactic 5400 61 10.7 5.7<br />

In Part I, the spring coefficient was deduced from an entropy expression containing a<br />

term squared in the relative stretch (1), together with the constraint of positive<br />

extension, thus motivating two additional parameters, called c 1 ,c 2 . Equation 6 is compatible<br />

with Part I, e.g., for a rather particular choice for c 1 together with c 2 0. The<br />

two approaches for the suggested setting c 1 1, c 2 0 possess measurable differences<br />

in their prediction of rheological quantities at startup of shear flow slightly increased<br />

stress amplitude at overshoot and slightly decreased strain at overshoot—for the<br />

model based on Eq. 6, steady shear flow weak decrease of steady stresses, and extensional<br />

flow critical strain rate postponed by a factor of 1.3–1.5, where the current<br />

setting has overall advantages when compared to experimental data for more details see<br />

Sec. IV.<br />

The parameter max clearly is the maximum possible stretching ratio of the chain<br />

contour length, which is equal to the square root of the number of Kuhn steps per<br />

entanglement N KE ,<br />

max N KE d t<br />

l K<br />

,<br />

7<br />

where d t is the average end-to-end distance of the entanglement segment at equilibrium<br />

and l K is the length of the Kuhn step which is twice the persistence length. If we denote<br />

the molecular weight between entanglements for polymer melts as M e , then M e for<br />

polymer solutions can be estimated by using the relation M e M e / 1.2 , where is the<br />

volume fraction of polymer Ferry 1980. The values of d t and l K for some polymer<br />

melts can be calculated from the relevant experimental data tabulated in the literature<br />

Fetters et al. 1996, where the values of M e are also given. The value of l K is calculated<br />

from the characteristic ratio C by using the relationship<br />

C l K<br />

l 1,<br />

8<br />

where l is the bond length Flory 1988, p.111. Based on these available data, the<br />

values of max for several important melts are given in Table I.<br />

The diffusion equation for the configurational distribution function takes the form<br />

Df<br />

Dt <br />

u • 1 uu<br />

u 2 •–uf <br />

<br />

u •D 1 uu<br />

u 2 • <br />

u f ,<br />

˙ s ṡ tot f dissip<br />

<br />

f 1<br />

2 d<br />

2 f<br />

2 s<br />

9


THERMODYNAMICALLY ADMISSIBLE REPTATION MODEL<br />

1299<br />

where<br />

ṡ tot 1 s 1 2˙ dissip .<br />

10<br />

This drift velocity for s means there is only a rescaling of the position label for the tube<br />

segment when the chain relaxes in the tube. The third term creation/destruction term on<br />

the right side of Eq. 9 compensates for configurations lost or gained at the boundaries.<br />

The terms involving second-order derivatives in Eq. 9 are of irreversible nature and<br />

express the erratic reptational motion along the chain contour second-order derivative<br />

with respect to s with the reptation time d and constraint release second-order derivative<br />

with respect to u with the orientational diffusion coefficient D, respectively. The<br />

form of D is<br />

D 1 6 1<br />

1<br />

˙ dissip<br />

<br />

2<br />

d <br />

H ˙ dissip<br />

<br />

, 11<br />

where H(x) is the Heaviside step function. The 1 term is interpreted as representing<br />

‘‘double reptation,’’ and the 2 term represents the CCR mechanism. The quantities 1/ d<br />

and ˙ dissip / determine the constraint release rate due to the loss of entanglements<br />

caused by reptation motion and chain retraction of side chains, respectively. The parameters<br />

1 and 2 determine the transfer rate from the constraint release rate to the relaxation<br />

rate of chain orientation, here we take 1 2 1/. This choice is motivated by<br />

the work of Mead et al. 1998 in connection with the appearance of their switch function.<br />

The argument is that the constraint release causes not only chain segments reorientation,<br />

but also contour length shortening this effect is not explicitly taken into account<br />

here. The role of the parameters 1 and 2 is to apportion the effects of constraint<br />

release between these two effects. When the chain is unstretched, the constraint release<br />

causes only chain segments reorientation; when the chain is highly stretched, the constraint<br />

release causes mainly chain contour length shortening. Hence, the parameters must<br />

be chosen in such a way that they approach unity when is near unity, and approach zero<br />

when is large.<br />

Conservation of the total probability implies the boundary conditions<br />

f˜ f˜<br />

0,<br />

ss 0<br />

ss 1<br />

12<br />

where f˜(s,r) f (u,s,r)d 3 u 1. At the chain ends, we assume random orientation by<br />

specifying the distribution<br />

fu,s,r 1<br />

u1, s 0,1, 13<br />

4<br />

which is common practice, but has been opened to discussion by <strong>Kröger</strong> and Hess 1993.<br />

The extra stress tensor consists of two contributions, 1 2 , namely, the original<br />

Doi–Edwards contribution<br />

1 r 3Zn p k B T<br />

0<br />

1<br />

uuf u,s,rd 3 uds,<br />

14


1300 FANG, KRÖGER, AND ÖTTINGER<br />

which reflects a proportionality between stress and alignment for fixed Z, and a contribution<br />

associated with the chain stretching,<br />

2 r c1n p k B T<br />

0<br />

1<br />

uuf u,s,rd 3 uds,<br />

15<br />

which may cause deviations from the stress-optic rule when stretching occurs. In these<br />

expressions, n p is the number density of polymers. The plateau modulus G N<br />

0 , to which 2<br />

does not contribute, is given by the Doi–Edwards result<br />

G N<br />

0 <br />

3<br />

5 Zn p k B T. 16<br />

When max approaches infinity, we obtain the simple expression<br />

5G N<br />

0 <br />

2<br />

0<br />

1<br />

uuf u,s,rd 3 uds,<br />

17<br />

which may be regarded as a manifestation of the stress-optic rule, see also Fuller 1995.<br />

The presented model accounts for double reptation, CCR, and chain stretching and avoids<br />

the IA approximation by the drift term and the creation/destruction term. It has been<br />

verified to possess the full structure of GENERIC, in particular, the time–structure invariance<br />

of reversible dynamics. The model has only four structural degrees of freedom,<br />

one from the position label s, two from the unit orientation vector u, and one from the<br />

chain stretching . In this paper, we assume d / s 3Z for our model. In addition to<br />

0<br />

the plateau modulus G N and the reptation time d , the model has four parameters<br />

(Z, 1 , 2 , max ) with the preferable values 1 2 1/ for melts as motivated in<br />

0<br />

Part I and above. In this work we regard G N and d as the key adjustable parameters,<br />

while Z and max are fixed by the chemistry of a particular polymer.<br />

III. SIMULATION ALGORITHM<br />

In this section, we describe the numerical simulation of the model. According to the<br />

theory of stochastic differential equations SDEs, the diffusion equation 9, when ignoring<br />

the creation/loss term discussed separately below, is equivalent to the following<br />

set of Itô SDEs for the stochastic processes u t , and s t :<br />

du t 1 u t u t<br />

u t 2 •–u t 2Du tdt2D 1 u t u t<br />

u t 2 •dW t ,<br />

ds t ṡ tot dt 1 2 d<br />

dW t ,<br />

18<br />

19<br />

where W t and all the three components of the vector W t are independent Wiener processes.<br />

The mentioned equivalence means that the average of an arbitrary function<br />

X(u,s), evaluated as an integral with the solution f of the diffusion equation at the time<br />

t, can be obtained as the expectation of the stochastic process X(u t ,s t )<br />

1<br />

<br />

0 Xu,sfu,s,td 3 uds Xu t ,s t .<br />

20


THERMODYNAMICALLY ADMISSIBLE REPTATION MODEL<br />

1301<br />

Here we assume that we consider the problem in a coordinate system moving with an<br />

arbitrary, given fluid particle, and we hence suppress the position argument.<br />

One should note that Eq. 18 is independent of Eq. 19. However, there exists a<br />

coupling between the processes u t and s t resulting from the boundary condition Eq. 13<br />

and the creation/destruction term in Eq. 9, which need to be discussed for obtaining a<br />

full equivalence between diffusion equations and SDEs. Whenever s t reaches the boundaries<br />

0 or 1, u t must be replaced by a random unit vector. Actually, there is a nonzero<br />

drift through the boundaries, which is exactly compensated by the creation/destruction<br />

term in the diffusion equation. Therefore, and in view of Eq. 12, the configurations<br />

diffusing through the boundaries are reflected back into the range of allowed s values.<br />

In order to calculate the statistical error of the results, we perform simulations for a<br />

number of independent blocks (N block ). In each simulation block, a number of trajectories<br />

(N sample ) of the stochastic processes u t and s t are propagated as follows. First, the<br />

drift in Eq. 19 is treated by a deterministic method. Given an s value at time t, the<br />

intermediate value after this treatment is<br />

s sṡ tot tt.<br />

21<br />

If there is a net flux of configurations out of the interval 0, 1, the lost configurations are<br />

randomly replaced by existing ensemble members; if there is a net flux of configurations<br />

into the interval 0, 1, the s values of the gained configurations are set in equal distance<br />

from the two outest existing points to the corresponding chain ends and the u vectors are<br />

randomly oriented according to the boundary condition. In order to keep the ensemble<br />

size constant in the latter case, the same number of configurations are randomly selected<br />

from the existing ensemble members and discarded. This first step takes care of both the<br />

drift in Eq. 19 and the creation/destruction term in Eq. 9. In the next step, we construct<br />

the new configurations at time tt<br />

s new s 1 2 d<br />

W,<br />

22<br />

u u–ut2DW,<br />

u new u<br />

u .<br />

23<br />

24<br />

If s new leaves the interval 0, 1, it will be reflected back into it. Upon any reflection, a<br />

new random unit vector u new is chosen according to the boundary condition. The accuracy<br />

of this straightforward Euler discretization of the stochastic part in Eq. 19 is of<br />

order t which is due to unobserved reflections Öttinger 1989. In order to obtain a<br />

higher-order (t) algorithm, we use the improved scheme proposed by Öttinger 1989,<br />

in which the effect of the unobserved reflections is taken into account through the conditional<br />

probability for their occurrence. Finally, the chain stretching at time tt is<br />

obtained as<br />

new ˙ convect ˙ dissip t.<br />

25<br />

For the results in this paper, we choose N sample 100 000, N block 10, and the time<br />

step size such that the strain is a maximum of 0.02 per time step and the maximum of the<br />

time step is 0.01 for small deformation rates. In each block, transient values are obtained<br />

by ensemble average, and steady values are obtained by average over time after reaching<br />

steady state in addition to ensemble average. The relative statistical errors of the simu-


1302 FANG, KRÖGER, AND ÖTTINGER<br />

lation data are within 2%, except for those with very small absolute values smaller<br />

than 10 3 at the very beginning of the startup of steady flows. There is no detectable<br />

difference in the results by increasing ensemble size or decreasing time step size.<br />

IV. RESULTS<br />

The experimental data used for comparison is from Kahvand 1995 which were also<br />

used by Hua et al. 1999. The test fluid is a solution in tricresyl phosphate of nearly<br />

monodisperse polystyrene with a molecular weight M w of 1.910 6 polydispersity index<br />

of 1.2 at a polymer density of 0.135 g/cm 3 . This fluid has a zero-shear-rate viscosity 0<br />

of 6.810 3 Pa s and a zero-shear-rate first normal stress coefficient 1,0 of 2.0<br />

10 5 Pa s 2 . The longest relaxation time and the average number of entanglement of the<br />

fluid are estimated to be 15 s and 7, respectively. Correspondingly, for our model we will<br />

0<br />

use the set of parameter values Z 7, 21, d 15 s, and G N 1160 Pa beside<br />

1 2 1/, as stated above to compare the simulation results with the experimental<br />

data. The value of the finite extensibility parameter max corresponds to 441 Kuhn steps<br />

per entanglement segment, or 6 monomers per Kuhn step. This value of monomers per<br />

Kuhn step is consistent with the value calculated from the data on polystyrene PS melts<br />

in Table I. The results given here except for extensional flows are not sensitive to finite<br />

extensibility. This set of parameter values is used for most of the cases, otherwise to be<br />

mentioned. In addition to the comparisons with the experimental data, comparisons with<br />

the two other models are also presented. These two models are the full chain simulation<br />

FCS model by Hua and Schieber 1998, Hua et al. 1999, and the simplified S form<br />

of the theory by Mead et al. 1998. Here, we shall denote them as ‘‘FCS’’ model and<br />

‘‘MLDS’’ model, respectively. For the FCS model, the simulations were done by Neergaard<br />

and co-workers, with the parameter b 3N K /(N1) 150, where N K denotes<br />

the number of Kuhn steps per chain and N is the number of beads for the FCS model<br />

compare Table II. The chosen value is more reasonable than the value b 25 used by<br />

Hua et al. 1999, while the other parameters remain the same. The former value of b<br />

corresponds to 150 Kuhn steps per entanglement, or 17 monomers per Kuhn step. This<br />

value of monomers per Kuhn step is larger than the value we estimated above. However,<br />

as for our model, the results given here for the FCS model are not affected significantly<br />

by choosing an even larger value of b. For the MLDS model in which double reptation is<br />

not taken into account, one may regard d and s both as adjustable parameters. The<br />

0<br />

values d 9s, d / s 10, and G N 800 Pa are obtained by a best overall fit to the<br />

available experimental data, where the exponential switch function is used. For d / s<br />

3Z 21 the fit would deteriorate considerably. Prior to further discussion of the<br />

different models, we refer to Table II for a characterization of the three models: FCS,<br />

MLDS, and the one under study.<br />

A. Start-up of steady shearing<br />

Figure 1 shows the growth of the dimensionless shear viscosity for several dimensionless<br />

shear rates predicted by our model. The linear viscoelastic limit forms an upper<br />

envelope for all the shear rates, the transient overshoots occur at medium and high shear<br />

rates, and the undershoots occur only at high shear rates. These undershoots are also<br />

predicted by the FCS model and reported by the experimental results. In Fig. 2 we<br />

compare our model predictions to the experimental data for the transient growth of the<br />

shear viscosity for several shear rates. Actually the model is able to capture the features<br />

of the data quantitatively. The predictions of different models for the same property are<br />

plotted in Fig. 3 where only the curves for the highest shear rate are shown because all


THERMODYNAMICALLY ADMISSIBLE REPTATION MODEL<br />

1303<br />

TABLE II. Characterization of the three models discussed in this work. Concerning the model predictions, the<br />

main discrepancies between FCS and our model are: larger stress overshoot predicted by FCS in the startup of<br />

steady shear flow Fig. 3 and larger second undershoot of the extinction angle predicted by FCS Fig. 12 due<br />

to segment connectivity only incorporated into FCS. The main discrepancies between predictions of MLDS and<br />

our model are: undershoots in transient stress and extinction angle only predicted by our model due to differences<br />

in the incorporation of CCR. For a detailed comparison see text.<br />

Parameters<br />

Model FCS MLDS This work<br />

Adjustable<br />

parameters<br />

The reptation time d<br />

and the number of<br />

beads per chain N fit<br />

from linear viscoelasticity<br />

alone<br />

The plateau modulus<br />

G N<br />

0 , the reptation<br />

time d , and the<br />

Rouse time s fit<br />

from nonlinear data<br />

The plateau modulus<br />

G N<br />

0 and the reptation<br />

time d fit from<br />

nonlinear data<br />

Physical<br />

effects<br />

incorporated<br />

and<br />

corresponding<br />

mathematical<br />

means<br />

Parameters<br />

fixed by the<br />

chemistry of<br />

the polymer<br />

or physical<br />

arguments<br />

Reptation<br />

The number of<br />

entanglements<br />

per chain Z,<br />

the Kuhn step length<br />

a K , and the number<br />

of Kuhn steps per<br />

chain N K<br />

None The number of<br />

entanglements<br />

and the chain<br />

Incorporated into a<br />

set of Langevin equations<br />

for the chain motion<br />

The second term in<br />

the equation for the<br />

tube survival probability,<br />

Eq. 10 of the<br />

reference for MLDS<br />

per chain Z,<br />

the constraint release<br />

parameters 1 , 2 ,<br />

stretchability<br />

parameter max<br />

The stochastic term<br />

in Eq. 19<br />

Avoiding IA<br />

approximation<br />

As for ‘‘reptation’’<br />

plus an equation for<br />

the tube motion<br />

Not considered<br />

The drift term in Eq.<br />

19 and the creation/<br />

destruction term in<br />

Eq. 9<br />

Chain<br />

stretching<br />

As for ‘‘avoiding<br />

IA approximation’’<br />

The equation for<br />

stretch, Eq. 12 of the<br />

reference<br />

The equation for<br />

stretch, Eq. 1<br />

Double<br />

reptation<br />

By a random,<br />

instantaneous constraint<br />

release algorithm<br />

Not considered<br />

The 1 -term in the<br />

orientational diffusion<br />

coefficient, Eq. 11<br />

Convective<br />

constraint<br />

release<br />

CCR<br />

As for ‘‘double<br />

reptation’’<br />

The term with the<br />

‘‘switch’’ function in<br />

Eq. 10 and the last<br />

term in Eq. 12 of the<br />

reference<br />

The 2 -term in the<br />

orientational diffusion<br />

coefficient, Eq. 11<br />

Chain-length<br />

breathing<br />

and segment<br />

connectivity<br />

Incorporated into a<br />

set of Langevin equations<br />

for the chain<br />

motion<br />

Not considered<br />

Not considered<br />

the models give similar results for the two lower shear rates. It turns out that the FCS<br />

model overpredicts the overshoot, our model predicts the magnitude precisely, and the<br />

MLDS model slightly overpredicts the overshoot and does not predict an undershoot.<br />

Comparisons between our model predictions and experimental data are also made for the


1304 FANG, KRÖGER, AND ÖTTINGER<br />

FIG. 1. Dimensionless shear viscosity as function of dimensionless time after startup of steady shearing at<br />

many dimensionless shear rates predicted by our model.<br />

first-normal-stress difference coefficient 1 in Fig. 4. The overshoot in 1 occurs at a<br />

higher shear rate than the shear viscosity does. Here, our model underpredicts the magnitude<br />

of the overshoot at the highest shear rate. Not shown are the predictions by FCS<br />

and MLDS models: they are very comparable with our model for the two lower shear<br />

rates. For the highest rate, again the FCS model overpredicts the overshoot and the<br />

MLDS model gives a better but underpredictive fit. In Fig. 5 we plot the magnitude of the<br />

strain p at which the maximum in the overshoot of the stress occurs as functions of<br />

shear rate predicted by the models and the experiment. It appears that the values of p<br />

stay nearly constant at low shear rates and increase with the increasing shear rate at high<br />

rates. Here, our model predictions are in good agreement with the experimental data,<br />

while the MLDS and FCS models overpredict. Not shown are the model predictions and<br />

FIG. 2. Transient growth of normalized viscosity as function of time under startup of steady shear flow at<br />

several shear rates predicted by our model lines and experiment symbols.


THERMODYNAMICALLY ADMISSIBLE REPTATION MODEL<br />

1305<br />

FIG. 3. Transient growth of normalized viscosity as functions of time under startup of steady shear flow at<br />

˙ 10 s 1 predicted by the models and experiment.<br />

the experimental data of p for the first normal stress difference, which have a qualitative<br />

shape similar to that seen in Fig. 5 but differ by approximately a factor of 2.<br />

B. Cessation of steady shearing<br />

In Fig. 6 we plot the relaxation of the shear stress, normalized by its initial value, after<br />

cessation of steady shear flow for two shear rates. All the model predictions not shown<br />

are the predictions by MLDS which almost overlap with ours compare rather well with<br />

the experimental data. The stress relaxation rate increases with the increasing previous<br />

shear rate. Later in the relaxation process, at times much larger than s , the relaxation<br />

rate is governed only by the reptation process and constraint release due to double<br />

FIG. 4. Transient growth of normalized first-normal-stress difference coefficient as functions of time under<br />

startup of steady shear flow at several shear rates predicted by our model lines and experiment symbols.


1306 FANG, KRÖGER, AND ÖTTINGER<br />

FIG. 5. Magnitude of the strain p at which the maximum in the stress overshoot occurs as functions of shear<br />

rate for shear stress under startup of steady shear flow predicted by the models and experiment.<br />

reptation represented by a smaller value of d in the MLDS model. This behavior is<br />

also in accordance with the other experimental results Attane et al. 1985; Menezes and<br />

Graessley 1982.<br />

C. Steady shear flow<br />

In Fig. 7 the steady-state values of the shear stress xy and the first normal stress<br />

difference N 1 are plotted as functions of shear rate. Our model and the FCS model<br />

predict a slight decrease of the shear stress over a range of shear rates that extends from<br />

roughly ˙ 0.3 to ˙ 2s 1 , where the MLDS model predicts a slight increase which<br />

is consistent with the experimental data. This new observation of the MLDS model may<br />

be attributed to the fact that we use a relatively small ratio d / s 10 here. Over the<br />

same range of shear rates, all the model predictions for the first normal stress difference<br />

FIG. 6. Relaxation of the shear stress reduced by its value at steady state as a function of time after cessation<br />

of steady shear flow at two shear rates predicted by experiment and the models.


THERMODYNAMICALLY ADMISSIBLE REPTATION MODEL<br />

1307<br />

FIG. 7. Steady-state values of the shear stress and first normal stress difference as functions of shear rate<br />

predicted by the models and experiment.<br />

N 1 increase gradually with shear rate, which are consistent with the experimental data.<br />

For the shear stress at rates after the range, our model predicts an increase, the FCS<br />

model shows a lesser increase, and the MLDS model shows a decrease. This decreasing<br />

behavior of MLDS is caused by the exponential switch function used and the additional<br />

term accounting for the CCR effect in the equation for stretch.<br />

We examine the behavior of the steady-state shear stress and its dependence on choice<br />

of our model parameters in more detail in order to understand how to avoid the prediction<br />

of an instability region as mentioned earlier. In Fig. 8, we see that the inclusion of chain<br />

FIG. 8. Steady-state values of the dimensionless shear stress as functions of dimensionless shear rate predicted<br />

by the DE model, the model without any constraint release effects 1 0, 2 0, the model with double<br />

reptation effect 1 1/, 2 0, the model with double reptation and convective constraint release effects<br />

( 1 2 1/), and the model with double reptation and enhanced convective constraint release effects<br />

( 1 1/, 2 2/. The other parameters of the model are set as Z 20, and max 10.


1308 FANG, KRÖGER, AND ÖTTINGER<br />

FIG. 9. Steady-state shear viscosity vs dimensionless shear rate which here is proportional to a physical shear<br />

rate rather than expressed in terms of a relaxation time which depends on molecular weight, for Z 20 and<br />

Z 40 with the other parameters 1 1/, 2 1/, and max 10.<br />

stretching alone yields the onset of an upturn of the shear stress at high shear rates, but<br />

the prediction of the power-law index is worse. When double reptation is considered, the<br />

situation is improved. After CCR is added and adjusted to proper intensity, the maximum<br />

in the steady-state shear stress disappears and the model predicts an almost constant<br />

stress over a wide range of shear rates, which is also observed in experiments of a highly<br />

entangled polymer solution Bercea et al. 1993. We thus confirm the previous conclusion<br />

that the CCR effect is critical for predicting the right shape of the shear stress curve.<br />

In Fig. 9 we plot the steady-state shear viscosity for Z 20 and Z 40 in order to<br />

resolve the effect of molecular weight. As a basis for comparison, we assume that the<br />

reptation time scales like d M 3 . It turns out that the steady-state viscosity of two<br />

different molecular weights coincides in the power-law region, in agreement with the<br />

behavior observed in experiments Stratton 1966 and, e.g., molecular dynamics simulations<br />

<strong>Kröger</strong> et al. 1993. Similar results are found by the FCS and MLDS models.<br />

The effect of molecular weight on N 1 is reported in Fig. 10, the model prediction is in<br />

agreement with the second part of observation 2 in Sec. I.<br />

D. Steady and transient extinction angle<br />

The extinction angle is given by (1/2)arctan(2 xy /N 1 ). In Fig. 11 we show the<br />

steady-state extinction angle as a function of shear rate. Obviously, in our model, CCR<br />

and double reptation prevent the chain segments from aligning with flow dramatically at<br />

high steady shear rates, and thus solve the overestimated steady shear orientation problem<br />

of the original DE model. The predictions of our model compare very favorably with the<br />

experimental results, while the predictions of the FCS and MLDS models go back to the<br />

DE model predictions at very high shear rates in steady flow. Not shown here are the<br />

predictions of our model with the parameters 1 2 exp((1)), which also coincide<br />

with the predictions of the DE model at high rates. It can be concluded from the<br />

above observations that the exponential switch function tunes down the constraint release<br />

effect on orientation too strongly, and hence is not appropriate. The reorientation algorithm<br />

in the FCS model for constraint release is successful at low and intermediate rates,


THERMODYNAMICALLY ADMISSIBLE REPTATION MODEL<br />

1309<br />

FIG. 10. Steady-state first-normal-stress difference vs dimensionless shear rate for Z 20 and Z 40 with<br />

the other parameter values as for Fig. 9.<br />

but may profit from an adjustment if a plateau for the steady time-averaged extinction<br />

angle at high rates caused by unsteady rotation of chains would be experimentally confirmed.<br />

In Fig. 12 we show the transient extinction angle predicted in the startup of shear flow<br />

followed by a step down in shear rate. In both predictions by our model and the FCS<br />

model, the extinction angle shows an undershoot at the startup of the first shear rate and<br />

an immediate undershoot at the inception of the lower shear rate before reaching a<br />

steady-state value. These predictions are in agreement with the experimental data Mead<br />

1996; Oberhauser et al. 1998. However, the second undershoot predicted by our<br />

model is too small compared with both the experiment and the prediction of the FCS<br />

FIG. 11. Steady-state extinction angle as a function of shear rate predicted by the models and experiment.


1310 FANG, KRÖGER, AND ÖTTINGER<br />

FIG. 12. Time-dependent extinction angle as a function of dimensionless time after steady shearing and<br />

attainment of steady state followed by a step down in dimensionless shear rate predicted by our model and the<br />

FCS model.<br />

model. The MLDS model with the switch function either exp((1)) or 1/ does not<br />

possess these two observed undershoots. The undershoots predicted by our model is due<br />

to chain retraction and its effect on chain orientation through the noise term in Eq. 18.<br />

E. Extensional flows<br />

The steady-state uniaxial extensional viscosity ( 1 ( zz xx )/(3˙ )) as a function<br />

of extensional rate is considered next, see Fig. 13, where z is the direction of extension.<br />

One observes four regimes of extension rate: i a very low rate region, ˙ 0.1/ d ,<br />

where the extensional viscosity is a constant equal to one third of the Trouton value; ii<br />

FIG. 13. Steady-state values of the dimensionless uniaxial extensional viscosity as function of dimensionless<br />

extension rate for 1 2 1/, Z 20, and max 10. The figure inside is the amplification of the region<br />

of extension rate range from 0.01 to 2.


THERMODYNAMICALLY ADMISSIBLE REPTATION MODEL<br />

1311<br />

FIG. 14. Dimensionless uniaxial extensional viscosity vs dimensionless time after startup of steady uniaxial<br />

elongation at many dimensionless elongational rates with the parameter values as for Fig. 13.<br />

a low rate region, 0.1/ d ˙ 1/ d , where the viscosity shows a weak extension<br />

hardening induced by orientation the maximum factor above the first constant region is<br />

around 1.05 here; iii an intermediate region, 1/ d ˙ 1/ s , in which the extensional<br />

viscosity is extension thinning, scaling as ˙ 1 ; and iv a high rate region, ˙<br />

1/ s , where the extensional viscosity increases with increasing strain rate and<br />

reaches another constant value finally. The same conclusion for regimes i, iii, and iv<br />

has been drawn from the MLDS model by Mead et al. 1998. Regime ii is caused by<br />

double reptation, because at low strain rates the other effects have a minor influence.<br />

Accordingly, regime ii is not predicted by the MLDS model and DE theory. Because of<br />

the lack of experimental data on monodisperse, linear entangled polymers, these unusual<br />

predictions have not yet been fully established. Two of the few works showing a pronounced<br />

regime ii were performed by Munstedt and Laun 1981 and Ferguson et al.<br />

1997. The first of these papers stated that two of the four HDPE investigated in their<br />

study do not show a maximum of the steady-state elongational viscosity; for the others it<br />

is less pronounced than in the case of LDPE.<br />

In Fig. 14 we plot the transient elongational viscosity as a function of time in the<br />

startup of uniaxial extensional flow. After the extensional rate exceeds 1/ s , the viscosity<br />

shows strain-hardening behavior and finally saturates because of the finite chain stretchability.<br />

To study the strain-hardening property of our model in detail, a strain-hardening<br />

parameter versus Hencky strain at the extensional rate of 100 is plotted in Fig. 15. The<br />

strain-hardening parameter is defined as the ratio of the strain rate-dependent nonlinear<br />

elongational viscosity to the strain rate-independent linear elongational viscosity at the<br />

same time Koyama and Ishizuka 1983. The figure reveals that the first departure from<br />

linearity occurs at a Hencky strain value around 0.5, fully in accord with the experimentally<br />

observed value by Takahashi et al. 1999. Figure 16 shows the influence of the<br />

maximum stretch max on the transient uniaxial extensional viscosity. An increase of<br />

max does not affect the occurrence of strain hardening, but postpones the saturation and<br />

increases the saturation value. Not shown are the predictions of the steady-state and<br />

transient viscosity in equibiaxial extension, which have qualitative shapes similar to those


1312 FANG, KRÖGER, AND ÖTTINGER<br />

FIG. 15. The strain-hardening parameter as a function of the Hencky strain at the dimensionless elongational<br />

rate ˙ d 100 with the parameter values as for Fig. 13.<br />

seen in the uniaxial extension, except that i the second regime of steady-state viscosity<br />

is absent and ii it is ‘‘softer’’ and less strain hardening than uniaxial extension Nishioka<br />

et al. 2000. Simulations of planar extension of our model are also performed. The<br />

steady-state planar viscosity p1 ( zz xx )/(4˙ ) direction of extension z and<br />

p2 ( yy xx )/(2˙ ) neutral direction y are plotted in Fig. 17. The viscosity p2<br />

decays much faster than the p1 does in the intermediate region, 1/ d ˙ 1/ s .<br />

Again, there is no second regime for both of them. In Fig. 18 we show the transient<br />

FIG. 16. Dimensionless uniaxial extensional viscosity vs dimensionless time after startup of steady uniaxial<br />

elongation at the dimensionless elongational rate ˙ d 200 for 1 2 1/, Z 20, and three different<br />

values of max .


THERMODYNAMICALLY ADMISSIBLE REPTATION MODEL<br />

1313<br />

FIG. 17. Steady-state values of the dimensionless planar extensional viscosity p1 along the direction of<br />

extension z and p2 along the neutral direction y as functions of dimensionless extensional rate with the<br />

parameter values as for Fig. 13.<br />

<br />

viscosity p2 not shown is the<br />

p1 , which is similar to the transient viscosity in<br />

uniaxial extension. The model predicts overshoots at medium and high extensional rates.<br />

V. CONCLUSION<br />

A new thermodynamically admissible reptation model that includes chain stretching,<br />

double reptation, and convective constraint release, and that avoids IA approximation is<br />

numerically investigated for transient and steady properties in shear and extensional<br />

flows. Quantitative comparisons are made with experimental data of entangled PS solution<br />

in shear flows. We find that the model is able to capture qualitatively, or quantita-<br />

FIG. 18. Dimensionless planar extensional viscosity p2 along the neutral direction y versus dimensionless<br />

time after startup of steady planar elongation at many dimensionless elongational rates with the parameter<br />

values as for Fig. 13.


1314 FANG, KRÖGER, AND ÖTTINGER<br />

tively in most cases, all the nonlinear properties observed for shear flows summarized in<br />

Sec. I from item 2 to item 9 item 1 has already been checked in Part I. For extensional<br />

flows, the model exhibits unusual predictions. In order to confirm these, further measurements<br />

of the stress for ideal samples linear, monodisperse, entangled polymers in various<br />

extensional flows are required. There are difficulties in performing this kind of<br />

experiment: a monodisperse polymer is usually sold in powder form, which is often more<br />

difficult to shape into bubble free samples than granules; the polymer breaks at much<br />

lower Hencky strains and is less homogeneous than an industrial melt Schweizer<br />

1999.<br />

Comparisons with two recently formulated models, FCS Hua and Schieber 1998<br />

and MLDS Mead et al. 1998, respectively, in shear flows are also presented. It turns<br />

out that all three models are competitive with each other; in many cases, they show<br />

similar behaviors. Very detailed comparisons are given in the previous section, e.g., for<br />

the MLDS and our model it is shown that a weak switch function proportional to the<br />

inverse of relative stretch 1/ predicts more realistic results than obtained by applying<br />

a stronger one.<br />

In addition to the plateau modulus G N<br />

0 and the reptation time d , our model has two<br />

basic parameters, namely Z, the number of entanglements per chain which is proportional<br />

to the molecular weight and can be expressed in terms of the Rouse and reptation time,<br />

and the maximum stretch max , being equal to the square root of the number of Kuhn<br />

steps per entanglement segment. The remaining two model parameters 1 , 2 have been<br />

introduced but actually assigned the values 1 2 1/.<br />

A ‘‘stable’’ steady-state shear stress curve is obtained by setting 2 2/. Predictions<br />

of the steady shear stress at very high shear rates can be also improved by considering<br />

the stretching time s to be a decaying function in , to account for the effect of<br />

contour length shortening by constraint release in situations where highly stretched conformations<br />

occur.<br />

Notably, our model is a thermodynamically consistent single-segment theory which<br />

has only four degrees of freedom. In the sense of considering the time and memory<br />

requirements for the computation, it is very suitable to utilize the model to simulate<br />

complex flows by using the CONNFFESSIT idea Laso and Öttinger 1993. To our<br />

knowledge, variance reduction techniques Hulsen et al. 1997; Öttinger et al. 1997;<br />

van Heel et al. 1999; Gallez et al. 1999; and Bonvin and Picasso 1999 may not be<br />

directly applied to the model. The dynamics for the position label—and in particular its<br />

jump events upon touching the chain ends—depend on the local value for the macroscopic<br />

velocity gradient such that a synchronization of trajectories by cancellation of<br />

fluctuations is prevented at first glance.<br />

However, it should be a straightforward exercise to extend the presented model to<br />

account for polydispersity effects by a mean-field approach, such that the properties of<br />

surrounding chains enter into the dynamics of the test chain. The resulting model will<br />

necessarily locate beyond simple superposition models.<br />

ACKNOWLEDGMENTS<br />

The authors would like to thank Professor D. C. Venerus for providing the experimental<br />

data, Professor R. G. Larson for providing the simulation program of the MLDS<br />

model, J. Neergaard for providing calculations of the FCS model, and Professor J. D.<br />

Schieber for extremely helpful comments. J. Fang is grateful for the financial support<br />

from the European Community Program BRITE-EuRam.


THERMODYNAMICALLY ADMISSIBLE REPTATION MODEL<br />

1315<br />

References<br />

Attané, P., M. Pierrand, and G. Turrel, ‘‘Steady and transient shear flows of polystyrene solutions I: Concentration<br />

and molecular weight dependence of non-dimensional viscometric functions,’’ J. Non-Newtonian<br />

Fluid Mech. 18, 295–317 1985.<br />

Bercea, M., C. Peiti, B. Dimionescu, and P. Navard, ‘‘Shear rheology of semidilute polymethylmethacrylate<br />

solutions,’’ Macromolecules 26, 7095–7096 1993.<br />

Bonvin, J. and M. Picasso, ‘‘Variance reduction methods for CONNFFESSIT-like simulations,’’ J. Non-<br />

Newtonian Fluid Mech. 84, 191–215 1999.<br />

Brown, E. F. and W. R. Burghardt, ‘‘First and second normal stress difference relaxation in reversing doublestep<br />

strain flows,’’ J. Rheol. 40, 37–54 1996.<br />

de Gennes, P. G., ‘‘Reptation of a polymer chain in the presence of fixed obstacles,’’ J. Chem. Phys. 55,<br />

572–579 1971.<br />

des Cloizeaux, J., ‘‘Double reptation vs simple reptation in polymer melts,’’ Europhys. Lett. 5, 437–442 1988.<br />

Doi, M., ‘‘Stress relaxation of polymeric liquids after double step strain,’’ J. Polym. Sci., Polym. Phys. Ed. 18,<br />

1891–1905 1980a.<br />

Doi, M., ‘‘A constitutive equation derived from the model of Doi and Edwards for concentrated polymer<br />

solutions and polymer melts,’’ J. Polym. Sci., Polym. Phys. Ed. 18, 2055–2067 1980b.<br />

Doi, M., ‘‘Explanation for the 3.4-power law for viscosity of polymeric liquids on the basis of the tube model,’’<br />

J. Polym. Sci., Polym. Phys. Ed. 21, 667–684 1983.<br />

Doi, M. and S. F. Edwards, ‘‘Dynamics of concentrated polymer systems. Part 1. Brownian motion in the<br />

equilibrium state,’’ J. Chem. Soc., Faraday Trans. 2 74, 1789–1801 1978a.<br />

Doi, M. and S. F. Edwards, ‘‘Dynamics of concentrated polymer systems. Part 2. Molecular motion under<br />

flow,’’ J. Chem. Soc., Faraday Trans. 2 74, 1802–1817 1978b.<br />

Doi, M. and S. F. Edwards, ‘‘Dynamics of concentrated polymer systems. Part 3. The constitutive equation,’’<br />

J. Chem. Soc., Faraday Trans. 2 74, 1818–1832 1978c.<br />

Doi, M. and S. F. Edwards, ‘‘Dynamics of concentrated polymer systems. Part 4. Rheological properties,’’ J.<br />

Chem. Soc., Faraday Trans. 2 75, 38–54 1979.<br />

Doi, M. and S. F. Edward, The Theory of Polymer Dynamics Clarendon, Oxford, 1986.<br />

Ferguson, J., N. E. Hudson, and M. A. Odriozola, ‘‘Interpretation of transient extensional viscosity data,’’ J.<br />

Non-Newtonian Fluid Mech. 68, 241–257 1997.<br />

Ferry, J. D., Viscoelastic Properties of Polymers Wiley, New York, 1980.<br />

Fetters, L. J., D. J. Lohse, and R. H. Colby, ‘‘Chain dimensions and entanglement spacings,’’ in Physical<br />

Properties of Polymers Handbook, edited by J. E. Mark AIP, New York, 1996.<br />

Flory, P. J., Statistical Mechanics of Chain Molecules Hanser, Munich, 1988.<br />

Fuller, G. G., Optical Rheometry of Complex Fluids Oxford University Press, Oxford, U.K., 1995<br />

Gallez, X., P. Halin, G. Lielens, R. Keunings, and V. Legat, ‘‘The adaptive Lagrangian particle method for<br />

macroscopic and micro-macro computations of time-dependent viscoelastic flows,’’ Comput. Methods<br />

Appl. Mech. Eng. 68, 345–364 1999.<br />

Grmela, M. and H. C. Öttinger, ‘‘Dynamics and thermodynamics of complex fluids. I. Development of a<br />

general formalism,’’ Phys. Rev. E 56, 6620–6632 1997.<br />

Hua, C. C. and J. D. Schieber, ‘‘Segment connectivity, chain-length breathing, segmental stretch, and constraint<br />

release in reptation models. I. Theory and single-step strain predictions,’’ J. Chem. Phys. 109, 10018–10027<br />

1998.<br />

Hua, C. C., J. D. Schieber, and D. C. Venerus, ‘‘Segment connectivity, chain-length breathing, segmental<br />

stretch, and constraint release in reptation models. II. Double-step strain predictions,’’ J. Chem. Phys. 109,<br />

10028–10032 1998.<br />

Hua, C. C., J. D. Schieber, and D. C. Venerus, ‘‘Segment connectivity, chain-length breathing, segmental<br />

stretch, and constraint release in reptation models. III. Shear flows,’’ J. Rheol. 43, 701–717 1999.<br />

Hulsen, M. A., A. P. G. van Heel, and B. H. A. A. van den Brule, ‘‘Simulation of viscoelastic flows using<br />

Brownian configuration fields,’’ J. Non-Newtonian Fluid Mech. 70, 79–101 1997.<br />

Ianniruberto, G., and G. Marrucci, ‘‘On compatibility of the Cox-Merz rule with the model of Doi and Edwards,’’<br />

J. Non-Newtonian Fluid Mech. 65, 241–246 1996.<br />

Ianniruberto, G. and G. Marrucci, ‘‘Stress tensor and stress-optical law in entangled polymers,’’ J. Non-<br />

Newtonian Fluid Mech. 79, 225–234 1998.<br />

Kahvand, H., ‘‘Strain Coupling Effects in Polymer Rheology,’’ Ph.D. thesis, Illinois Institute of Technology,<br />

1995.<br />

Ketzmerick, R. and H. C. Öttinger, ‘‘Simulation of a Non-Markovian process modelling contour length fluctuation<br />

in the Doi-Edwards model,’’ Continuum Mech. Thermodyn. 1, 113–124 1989.<br />

Koyama, K. and O. Ishizuka, ‘‘Nonlinearity in uniaxial elongational viscosity at a constant strain rate,’’ Polym.<br />

Proc. Eng. 1, 55–70 1983.<br />

<strong>Kröger</strong>, M. and S. Hess, ‘‘Viscoelasticity of polymeric melts and concentrated solutions. The effect of flowinduced<br />

alignment of chain ends,’’ Physica A 195, 336–353 1993.


1316 FANG, KRÖGER, AND ÖTTINGER<br />

<strong>Kröger</strong>, M., W. Loose, and S. Hess, ‘‘Rheology and structural changes of polymer melts via nonequilibrium<br />

molecular dynamics,’’ J. Rheol. 37, 1057–1079 1993.<br />

Laso, M. and H. C. Öttinger, ‘‘Calculation of viscoelastic flow using molecular models: the CONNFFESSIT<br />

approach,’’ J. Non-Newtonian Fluid Mech. 47, 1–20 1993.<br />

Marrucci, G., ‘‘The Doi-Edwards model without independent alignment,’’ J. Non-Newtonian Fluid Mech. 21,<br />

329–336 1986.<br />

Marrucci, G., ‘‘Dynamics of entanglements: A nonlinear model consistent with the Cox-Merz rule,’’ J. Non-<br />

Newtonian Fluid Mech. 62, 279–289 1996.<br />

Marrucci, G. and N. Grizzuti, ‘‘The Doi-Edwards model in slow flows. Predictions on the weissenberg effect,’’<br />

J. Non-Newtonian Fluid Mech. 21, 319–328 1986.<br />

Marrucci, G. and N. Grizzuti, ‘‘Fast flow of concentrated polymers: Predictions of the tube model on chain<br />

stretching,’’ Gazz. Chim. Ital. 118, 179–185 1988.<br />

Mead, D. W., ‘‘Orientation angle transients in step shear rate experiments on lightly entangled systems of linear<br />

flexible polymers,’’ In Proceeding of the 12th International Congress on Rheology, Aug. 18–23, 1996,<br />

edited by Aït-Kadi, A., J. M. Dealy, D. F. James, and M. C. Williams Canadian Rheology Group, Quebec<br />

City, Canada, 1996.<br />

Mead, D. W. and R. G. Larson, ‘‘Rheooptical study of isotropic solutions of stiff polymers,’’ Macromolecules<br />

23, 2524–2533 1990.<br />

Mead, D. W. and L. G. Leal, ‘‘The reptation model with segmental stretch I. Basic equations and general<br />

properties,’’ Rheol. Acta 34, 339–359 1995.<br />

Mead, D. W., D. Yavich, and L. G. Leal, ‘‘The reptation model with segmental stretch II. Steady flow<br />

properties,’’ Rheol. Acta 34, 360–383 1995.<br />

Mead, D. W., R. G. Larson, and M. Doi, ‘‘A molecular theory for fast flows of entangled polymers,’’ Macromolecules<br />

31, 7895–7914 1998.<br />

Menezes, E. V. and W. W. Graessley, ‘‘Nonlinear rheological behavior of polymer systems for several shear<br />

flow histories,’’ J. Polym. Sci., Polym. Phys. Ed. 20, 1817–1833 1982.<br />

Munstedt, H. and H. M. Laun, ‘‘Elongational properties and molecular structure of polyethylene melts,’’ Rheol.<br />

Acta 20, 211–221 1981.<br />

Nishioka, A., T. Takahashi, Y. Masubuchi, J.-I. Takimoto, and K. Koyama, ‘‘Description of uniaxial, biaxial,<br />

and planar elongational viscosities of polystyrene melt by the K-BKZ model,’’ J. Non-Newtonian Fluid<br />

Mech. 89, 287–301 2000.<br />

Oberhauser, J. P., L. G. Leal, and D. W. Mead, ‘‘The response of entangled polymer solutions to step changes<br />

of shear rate: signatures of segmental stretch?,’’ J. Polym. Sci., Part B: Polym. Phys. 36, 265–280 1998.<br />

O’Connor, N. P. T. and R. C. Ball, ‘‘Confirmation of the Doi-Edwards model,’’ Macromolecules 25, 5677–<br />

5682 1992.<br />

Osaki, K. and M. Kurata, ‘‘Experimental appraisal of the Doi-Edwards theory for polymer rheology based on<br />

the data for polystyrene solutions,’’ Macromolecules 13, 671–676 1980.<br />

Osaki, K., S. Kimura, and M. Kurata, ‘‘Relaxation of shear and normal stresses in double-step shear deformations<br />

for a polystyrene solution. A test of the Doi-Edwards theory for polymer rheology,’’ J. Rheol. 25,<br />

549–562 1981.<br />

Öttinger, H. C., ‘‘Computer simulation of reptation theories. I. Doi-Edwards and Curtiss-Bird models,’’ J.<br />

Chem. Phys. 91, 6455–6462 1989.<br />

Öttinger, H. C., ‘‘Modified reptation model,’’ Phys. Rev. E 50, 4891–4895 1994.<br />

Öttinger, H. C., ‘‘Nonequilibrium thermodynamics—A tool for applied rheologists,’’ Appl. Rheol. 9, 17–26<br />

1999a.<br />

Öttinger, H. C., ‘‘A thermodynamically admissible reptation model for fast flows of entangled polymers,’’ J.<br />

Rheol. 43, 1461–1493 1999b.<br />

Öttinger, H. C., ‘‘Thermodynamically admissible reptation models with anisotropic tube cross sections and<br />

convective constraint release,’’ J. Non-Newtonian Fluid Mech. 89, 165–185 2000.<br />

Öttinger, H. C. and A. N. Beris, ‘‘Thermodynamically consistent reptation model without independent alignment,’’<br />

J. Chem. Phys. 110, 6593–6596 1999.<br />

Öttinger, H. C. and M. Grmela, ‘‘Dynamics and thermodynamics of complex fluids. II. Illustrations of a general<br />

formalism,’’ Phys. Rev. E 56, 6633–6655 1997.<br />

Öttinger, H. C., B. H. A. A. van den Brule, and M. A. Hulsen, ‘‘Brownian configuration fields and variance<br />

reduced CONNFFESSIT,’’ J. Non-Newtonian Fluid Mech. 70, 255–261 1997.<br />

Pearson, D. S., A. D. Kiss, and L. J. Fetters, ‘‘Flow-induced birefringence of concentrated polyisoprene solutions,’’<br />

J. Rheol. 33, 517–535 1989.<br />

Pearson, D. S., E. A. Herbolzheimer, G. Marrucci, and N. Grizzuti, ‘‘Transient behavior of entangled polymers<br />

at high shear rates,’’ J. Polym. Sci., Polym. Phys. Ed. 29, 1589–1597 1991.<br />

Schweizer, T. personal communication, 1999.<br />

Stratton, R. A., ‘‘The dependence of non-Newtonian viscosity on molecular weight for ‘monodisperse’ polystyrene,’’<br />

J. Colloid Interface Sci. 22, 517–530 1966.


THERMODYNAMICALLY ADMISSIBLE REPTATION MODEL<br />

1317<br />

Takahashi, T., J. Takimoto, and K. Koyama, ‘‘Elongational viscosity for miscible and immiscible polymer<br />

blends. I. PMMA and AS with similar elongational viscosity,’’ J. Appl. Polym. Sci. 73, 757–766 1999.<br />

Tsenoglou, C., ‘‘Viscoelasticity of binary homopolymer blends,’’ ACS Polym. Preprints 28, 185–186 1987.<br />

van Heel, A. P. G., M. A. Hulsen, and B. H. A. A. van den Brule, ‘‘Simulation of the Doi–Edwards model in<br />

complex flow,’’ J. Rheol. 43, 1239–1260 1999.<br />

Venerus, D. C. and H. Kahvand, ‘‘Doi–Edwards theory evaluation in double-step strain flows,’’ J. Polym. Sci.,<br />

Part B: Polym. Phys. 32, 1531–1542 1994a.<br />

Venerus, D. C. and H. Kahvand, ‘‘Normal stress relaxation in reversing double-step strain flows,’’ J. Rheol. 38,<br />

1297–1315 1994b.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!