SlideShare a Scribd company logo
Comprehensive comparisons with nonlinear flow
data of a consistently unconstrained Brownian
slip-link model
Jay D. Schieber,a)
Deepa M. Nair, and Thidaporn Kitkrailard
Center of Excellence in Polymer Science and Engineering, Department of
Chemical and Biological Engineering, Illinois Institute of Technology,
Chicago, Illinois 60616-3793
(Received 19 April 2006; final revision received 6 September 2007͒
Synopsis
A consistently unconstrained Brownian slip-link model ͑CUBS͒ with constant chain friction is
used to predict the nonlinear rheological behavior of linear, entangled, polymeric liquids. The
model naturally incorporates primitive-path-length fluctuations, segment connectivity, monomer
density fluctuations, entanglement fluctuations, and constraint release without making any closure
approximations. Constraint release is imposed on the level of the dynamics of the chain, and the
relaxation modulus follows from these rigorously. The model is a mean-field, single-chain slip-link
model, or temporary network model, with a single phenomenological time constant, ␶e, fit by linear
viscoelasticity. The nonlinear flow predictions are made without adjusting any additional
parameters. We find that the addition of constant chain friction noticeably improves the model
predictions in all the flows considered. In contradiction with tube models, the results suggest that
the additional physics of constraint release and convective constraint release are not very important
in predicting the nonlinear shear properties, except at low shear rates ͑close to the LVE regime͒.
© 2007 The Society of Rheology. ͓DOI: 10.1122/1.2790460͔
I. INTRODUCTION
Quantitative agreement with experimental data has become increasingly accurate for
tube models. However, most such models are restricted to either linear viscoelasticity or
to flow, and to a single kind of chain architecture. If chain branching is considered, the
mathematical model is modified substantially from the linear case, or is unknown alto-
gether.
These tube models exist on various levels of description. Doi and Edwards tensor
model ͓Doi and Edwards ͑1978a, 1978b, 1978c͔͒ can be written as six three-dimensional
integrals: one dimension for time history and two over segment orientation; tube models
with stretch have seven coupled ordinary differential equations, including one dimension
for average stretch ͓Ianniruberto and Marrucci ͑2001͔͒; the preaveraged tube model of
͓Graham et al. ͑2003͔͒ uses six coupled partial differential equations. The last model is
sometimes considered to be the most accurate tube model available.
Ideally, one seeks a model on the least-detailed level possible that makes accurate
a͒
Electronic mail: schieber@iit.edu
© 2007 by The Society of Rheology, Inc.
1111J. Rheol. 51͑6͒, 1111-1141 November/December ͑2007͒ 0148-6055/2007/51͑6͒/1111/31/$27.00
predictions for the problem at hand. For example, continuum mechanics equations—
coupled ordinary differential equations—have been a primary goal, historically. If more
complicated equations arise, calculation of nonhomogeneous flows becomes more diffi-
cult. For example, implementation of the coupled partial differential equations of Graham
et al. in finite element codes is presumably computationally prohibitive. Even so, the use
of uncontrolled closure approximations in these models can lead to unexpected errors.
For example, the latter model predicts zero second-normal stress difference in shear flow.
In order to avoid uncontrolled closures, it is necessary to account for fluctuations more
rigorously. Toward this end stochastic models are ideally suited. This is the approach
taken, for example, by Öttinger using a single-segment model ͓Fang et al. ͑2000͒; Öt-
tinger ͑1999͔͒, by Hua and Schieber for a full chain model ͓Hua and Schieber ͑1998a͒;
Hua et al. ͑1998, 1999͒; Neergaard et al. ͑2000͔͒, and by Masubuchi et al. using inter-
acting chains inside a simulation box ͓Masubuchi et al. ͑2003͔͒. Not surprisingly, these
models show agreement with data over a broader range of experiments—for example, a
reasonable second normal-stress difference. However, only the first of these models has
been used in complex geometries ͓Gigras and Khomami ͑2002͔͒.
Nonetheless, all tube models to date exhibit a maximum in the steady shear stress
versus shear rate curve, which is not observed experimentally. Only through tuning pa-
rameters for excessive convective constraint release, or too-early onset of chain stretching
is the maximum avoided in the single-segment tube models. When no such adjustable
parameters are available, the maximum is seen for sufficiently entangled tube models.
More recent work has seen the advance of slip-link models. The first slip-link picture
was drawn by Doi and Edwards. However, it was used only to motivate a certain
network-model expression for the stress tensor in their tube model. The repton model of
Rubinstein ͑1987͒ more closely resembles the slip-link picture than the tube picture, but
here the constraining mesh of chains surrounding the test chain was considered to be
regular. Flow was not considered. Neergaard and Schieber introduced the idea of using
the monomer density between slip-links as a stochastic variable in a stochastic, full chain
model ͓Neergaard and Schieber ͑2000͒; Schieber ͑2003a͒; Schieber et al. ͑2003͔͒, and a
similar multichain simulation was introduced independently by ͓Masubuchi et al.
͑2001͔͒. Greco considered a similar stochastic, single-segment model, in which an infi-
nitely long chain possessed two entanglements, and the ͑infinitely long͒ ends were in a
chemical potential bath ͓Greco ͑2002͔͒. Only step strains have been considered to date
for this latter model.
Likhtman considered a Rouse chain sliding through slip links that are themselves
attached by elastic springs to an affinely deforming background ͓Likhtman ͑2005͔͒. The
picture is a more-detailed depiction of a similar model developed by Panyukov and
Rubinstein for lightly cross-linked elastomers ͓Rubinstein and Panyukov ͑2002͔͒. Likht-
man’s model has been used to study dynamic equilibrium only, to date.
In a previous article, we generalized the slip-link model of Neergaard and Schieber to
include constraint release—the idea that the matrix of constraining chains are relaxing on
the same spectrum of time scales as the probe chain ͓Nair and Schieber ͑2006͔͒. Com-
prehensive comparison was made there with linear viscoelastic experiments and reason-
able agreement was found with data. Here we make comparisons with flow. We will show
that the slip-link model is able to describe accurately all observations ͑known to us͒ in
shear flow, including a monotonic steady shear stress growth with shear rate for linear
chains.
An added benefit to slip-link models is a more natural connection to atomistic level
simulations. Recently, several groups have created and tested algorithms to predict en-
tanglement statistics from equilibrated, long-chain, melt polymers on such a level ͓Ever-
1112 SCHIEBER, NAIR, AND KITKRAILARD
aers et al. ͑2004͔͒. More than one algorithm has been able to predict entanglement
density; however, recent attempts have included calculation of more-detailed statistics,
such as primitive-path length distribution, or entanglement number distribution
͓Foteinopoulou et al. ͑2006͒; Shanbhag and Kroger ͑2007͒; Tzoumanekas and Theodorou
͑2006͒; Zhou and Larson ͑2005͔͒. These comparisons have shown excellent agreement
with our slip-link theories, offering a path for true ab initio calculation of stresses and
dynamics in entangled polymers.
A third advantage to slip-link models is that branched systems or even cross-linked
systems ͓Eskandari ͑2006͔͒ can be modeled without significant modification to the math-
ematics.
II. THE CUBS MODEL
In the consistently unconstrained Brownian slip-link ͑CUBS͒ model ͑Fig. 1͒, entangle-
ments are represented as small rings called “slip links” along the contour length of the
chain. As in the reptation theory, the chain is free to move along its contour length but the
transverse motion is restricted due to the presence of slip links or entanglements. The end
segments which are not constrained by any slip links are free to explore all conformations
in space. The chain is allowed to slide through these entanglements and, hence, the
number of Kuhn steps in an entangled strand fluctuates stochastically. Without constraint
release, the entanglements are assumed to move affinely.
The entanglements are created and destroyed only at the ends of the chain when an
end strand completely slips out of a slip link and the dangling end has zero Kuhn steps.
In reality, a new entanglement is created when a surrounding chain entangles the test
chain at the end. We model this process by an entanglement creation probability that
satisfies detailed balance with entanglement destruction. In the previously proposed
model, entanglements are assumed to be fixed in the medium—i.e., reptation of the
surrounding chains was not considered ͑no constraint release͒. In this model, we consider
the diffusion of the surrounding chains, thereby mimicking the creation and destruction of
entanglements in the middle of the chain ͑constraint release͒. In the previous model it
FIG. 1. Sketch of a short chain of 200 Kuhn steps trapped in three entanglements. Also shown is the notation
used for the vectors in the model. Ri is the position in space of entanglement i, drawn by solid ͑black͒ arrows,
and Qi is the dashed ͑green͒ arrow connecting entanglement i−1 to entanglement i, trapping Ni Kuhn steps.
1113CUBS MODEL FLOW DATA COMPARISON
was also assumed that friction is proportional to the number of entanglements and, hence,
total friction in the chain fluctuated with the creation and destruction of entanglements. In
the present model, friction is assumed to be proportional to the number of monomers or
Kuhn steps in the chain. Hence, total friction in the chain is a constant quantity since the
total number of monomers or Kuhn steps in the chain is a constant.
A. Mathematical description
The evolution equation for the probability density has four contributions, from, re-
spectively, affine motion of the entanglements, sliding of the chain through the entangle-
ments, constraint release, and creation/destruction of the entanglements at the chain ends
‫ץ‬p
‫ץ‬t
= p˙affine + p˙steps + p˙CR + p˙ends. ͑1͒
We give each contribution in turn.
The dynamics of the Kuhn steps in the chain are driven by both Brownian forces and
free energy minimization. Free energy minimization requires the sliding of Kuhn steps
from the strands of higher chemical potential ͑‫ץ‬F/‫ץ‬Ni͒T,͕Qj͖,͕Nj i͖ to those of lower chemi-
cal potential. A friction or mobility for this Kuhn-step-shuffling process is also required,
which introduces the single phenomenological parameter of the theory
p˙steps = ͚i,j=1
Z
‫ץ‬
‫ץ‬Ni
1
kBT͕͑Nk͖͒
Aˆ
ijͫͩ‫ץ‬F
‫ץ‬Nj
ͪT,͕Rj͖,͕Nk j͖
p + kBT
‫ץ‬p
‫ץ‬Nj
ͬ, ͑2͒
where
Aˆ
ij ª
Ά
−
1
␶i−1
, i = j + 1
1
␶i
+
1
␶i−1
, i = j
−
1
␶i
, i = j − 1
0, otherwise.
͑3͒
We have considered two possible implementations for this friction: constant-
entanglement friction and constant-chain friction. For constant-entanglement friction, we
set ␶i=␶Kϳconstant, in which case Aˆ
ij is the Rouse matrix divided by ␶K. Nondimen-
sionalizing the dynamical equations also leads to a more natural fundamental time scale
␶e=Ne
2
␶K.
Since the number of entanglements on a chain fluctuate, the total friction experienced
by the chain will also fluctuate in this case. For various mathematical reasons, tube
models typically assume that the friction of a chain is constant. Hence, we also consider
the possibility that ␶i=␶K͑Ni+Ni−1͒/͑2Ne͒, which we call constant-chain friction. Com-
parison with experiment below suggests that constant-chain friction is a better assump-
tion.
In the absence of constraint release, the entanglements are assumed to move affinely
during flow. Hence, we write
1114 SCHIEBER, NAIR, AND KITKRAILARD
p˙affine = − ͚i=1
Z−1
‫ץ‬
‫ץ‬Ri
· ͓␬ · Ri͔p, ͑4͒
where ␬=ٌ͑v͒†
is the transpose of the velocity gradient. Note that, due to the shuffling of
Kuhn steps through the slip links, the chains themselves do not deform affinely.
The effect of constraint release is incorporated in a self-consistent mean-field way.
Constraint release is mimicked by a diffusive process of the entanglements, which makes
the entanglement motion no longer affine. Hence, we write
p˙CR = ͚i,j=1
Z−1
‫ץ‬
‫ץ‬Ri
·
NeaK
2
12kBT␶j
CRAijͫpͩ‫ץ‬F
‫ץ‬Rj
ͪT,͕Nj͖,͕Ri j͖
+ kBT
‫ץ‬p
‫ץ‬Rj
ͬ, ͑5͒
where ␶j
CR
is the constraint release time scale and Aij is the Rouse matrix ͓Bird et al.
͑1987͔͒. Here we use a dimensionless prefactor, which determines the strength of the
constraint release mechanism. We calculate the value of this prefactor to be 1/12. Details
of this calculation may be found in the supplementary materials of ͓Nair and Schieber
͑2006͔͒. Thus, by considering the diffusion of the surrounding chains, we are mimicking
the birth and destruction of slip links in the middle of the chain. The time constants ␶i
CR
are obtained in a self-consistent way by keeping track of the fraction of surviving en-
tanglements during an equilibrium simulation. Note that this assumption of Rouse-like
motion for the slip links is not necessary for this model, since creation and destruction of
entanglements in the middle of the chain could be implemented rigorously. However,
such an assumption is typically made in tube models, which we will follow here. More
rigorous formulation is left for future work.
During flow, the rate of entanglement destruction can increase—so-called “convective
constraint release” ͓Ianniruberto and Marrucci ͑2000͔͒. Hence, during flow we replace
␶i
CR
with ␶i
CCR
, where
1
␶i
CCR =
1
␶i
CR + ͗wd͘ − ͗wd͘eq, i = 1, ... ,Z − 1, ͑6͒
where ͗wd͘ is the average rate of destruction of entanglements at the dangling ends during
flow and ͗wd͘eq is that at equilibrium.
Ideally, the destruction of entanglements would be given by boundary conditions—
i.e., the entanglement dies when the chain reaches the entanglement. However, such an
implementation creates subtle numerical and thermodynamical problems, because the
chain statistics are no longer determined by the entropy of the chain. On the other hand,
the model is not designed to distinguish length and time scales on the order of when the
final, single Kuhn step passes through the entanglement. To obviate these problems, we
instead give the dangling end an entropy that prevents the last Kuhn step from passing
through the entanglement. Then, we give a probability for entanglement destruction that
increases with decreasing number of Kuhn steps in the last strand
wd =
1
␶e
ͱ3Ne
NE
, ͑7͒
where NE is the number of Kuhn steps in the end strands. Note that Eq. ͑44͒ of ͓Schieber
et al. ͑2003͔͒ contains a typographical error, whereas Eq. ͑20͒ of ͓Nair and Schieber
͑2006͔͒ is correct. This allows us to control the proper statistics of the chain through the
entropy and avoid the problems mentioned. Since the short time scales are most affected,
we do not expect this approximation to have a significant effect. Nonetheless, we are
1115CUBS MODEL FLOW DATA COMPARISON
currently studying a more fundamental implementation of entanglement destruction that
avoids these problems entirely, in order to study the effect of the approximation rigor-
ously.
Related to the above problem, we impose the restriction that the length of an entangle-
ment segment should always be greater than the length of a Kuhn step. The strand length
is kept greater than the length of a Kuhn step ͑aK͒, through a reflecting boundary condi-
tion. We have studied the sensitivity of predictions to this cutoff length and find that
results are indistinguishable within statistical error as long as it remains below the aver-
age equilibrium entanglement spacing ͓Kitkrailard ͑2005͔͒.
The polymer contribution to the stress tensor is given by ͓Nair and Schieber ͑2006͒;
Schieber ͑2003a, 2003b͔͒
␶p
= − n͚Z=3
ϱ
͚j=1
Z−1
͵͵͵Rjͩ‫ץ‬F
‫ץ‬Rj
ͪT,͕Ni͖,͕Ri j͖
p͕͑Ri͖,͕Ni͖,Z;t͒dRZ−1
dNZ
, ͑8͒
where n is the number density of the polymer chains.
B. Brownian dynamics simulation
Solution of the evolution Eq. ͑1͒ for the conformational probability density is, in
general, not possible analytically. Numerical solution of the probability density is also not
feasible given the large number of dimensions possible to describe the conformations.
Therefore, Brownian dynamics is used to calculate the model’s predictions for stress ͑or
other quantities͒. Finding the equivalent stochastic differential equations and other sto-
chastic dynamics from the evolution equation for p are straightforward ͓Gardiner ͑1982͒;
Honerkamp ͑1983͒; Öttinger ͑1996͔͒.
The evolution equation for the number of Kuhn steps in strand i can be written as
dNi =
1
kBT␶i
͑␮i+1 − ␮i͒dt −
1
kBT␶i−1
͑␮i − ␮i−1͒dt + ͱ2
␶i
dWi − ͱ 2
␶i−1
dWi−1, i = 1, ... ,Z,
͑9͒
where ␮iª͑‫ץ‬F/‫ץ‬Ni͒T,͕Rj͖,͕Nj i͖ is the chemical potential of the ith strand, except for the
end strands, where ␮0=␮1, ␮Z+1=␮Z, and ␶0=␶Z=ϱ. An ensemble of chains whose
dynamics are described by this SDE are simulated in discrete time, by using a semi-
implicit stepping algorithm. Details are given in Nair and Schieber ͑2006͒.
The evolution equation for the position of entanglements is given as
dRi = ␬ · Ridt −
NeaK
2
12kBT␶i
CRͩ‫ץ‬F
‫ץ‬Ri
ͪT,͕Ni͖,͕Rj i͖
dt + ͱNeaK
2
6␶i
CR dWi. ͑10͒
The ensemble size used is 5000–10 000 chains and time-step size ͑⌬t͒ is selected as
0.02␶e. Chain conformations are generated from the equilibrium probability density as-
suming zero cutoff length in Q and then segments smaller than aK are reflected. This
procedure gives the proper statistics for Q, but distorts the statistics for the ͕Ni͖ some-
what. In order to correct the Kuhn-step-number statistics, it is necessary to equilibrate a
few time steps. We equilibrate ͗Z͘eq␶e/⌬t time steps, which is sufficient to create an
equilibrium ensemble. Simulations with and without constraint release ͑CR͒ are per-
formed and the effect of convective constraint release ͑CCR͒ is incorporated as shown by
Eq. ͑6͒.
1116 SCHIEBER, NAIR, AND KITKRAILARD
The shear rates used in the simulations are made dimensionless using ␶e. The largest
value of Ne used for calculations is 50, since theoretical predictions are insensitive to Ne
when greater than 50, at fixed ͗Z͘eq.
III. EXPERIMENTAL SYSTEMS
Seven polymer systems are compared with the CUBS model: a polystyrene solution, a
polybutadiene solution, and five polystyrene melts. The different systems are summarized
in Table I.
Test fluid PS1 is a 12 wt % solution in tricresyl phosphate nearly monodisperse ͑poly-
dispersity indexϭ1.2͒ polystyrene with a molecular of weight Mw of 1.92ϫ106
Da at a
polymer density of 0.135 g/cm3
͓Venerus and Kahvand ͑1994a͔͒. In the case of polymer
solutions, the entanglement molecular weight is calculated as
Me
solution
Х
Me
melt
␾1.3 , ͑11͒
where ␾ is the volume fraction of the polymer and Me
melt
Х18 000 g/mol. At any rate, the
entanglement molecular weight is confirmed in LVE comparisons by examination of the
height of the plateau modulus. We use the same fluid for making data comparisons in
reversing double-step strain flow ͓Brown and Burghardt ͑1996͒; Venerus and Kahvand
͑1994a, 1994b͔, inception of steady shear flow, steady shear flow, cessation of steady
shear flow ͓Kahvand ͑1995͔͒, and exponential shear flow ͓Neergaard et al. ͑2000͔͒. Me
for the solution is calculated as given by Eq. ͑11͒ and the value of ͗Z͘eq is estimated
͑͗Z͘eq=Mw/Me͒ to be 10. From the linear viscoelastic comparisons, ␶e for the solution is
obtained as, ␶e
NCR
=0.84 s and ␶e
CR
Х1.7 s at 23 °C.
Comparison between G*
data and theory to determine ␶e for this system is given in
Nair and Schieber ͑2006͒.
The experimental sample used for data comparisons in single-step strain experiments
is the polystyrene melt PS2 of molecular weight 200 kDa ͓Nair ͑2004͔͒. Using Me
TABLE I. Summary of experimental systems compared with theory in this article: Mwϭweight-averaged
molecular weight in kDa, PDIϭMw /Mn, polydispersity index ͑where Mn is the number-averaged molecular
weight͒, Tϭtemperature in degrees Celsius, ͗Z͘eqϭaverage number of entanglements per chain plus one; ␶e
͑NCR͒ϭcharacteristic strand relaxation time assuming no constraint release in seconds; ␶e ͑CR͒ϭcharacteristic
strand relaxation time with constraint release in seconds. WLC indicates use of the worm-like chain free energy;
otherwise, the Gaussian free energy expression was used.
System Mw ͑kDa͒ PDI T ͑°C͒ ͗Z͘eq ␶e ͑NCR͒ ͑s͒ ␶e ͑CR͒ ͑s͒ Source͑s͒
PS1 sol. 1920 1.20 23 10 0.84 1.7 Venerus and Kahvand ͑1994a͒
Venerus and Kahvand ͑1994b͒
Kahvand ͑1995͒
Neergaard et al. ͑2000͒
PS2 melt 200 1.05 160 11 0.4 0.78 Nair ͑2004͒
PS3 melt 500.5 1.02 170 28 0.375 Collis et al. ͑2005͒
PS4 melt 200 1.06 175 11 0.137 Schweizer et al. ͑2004͒
PS5 melt 200 1.05 130 14 56.4 Bach et al. ͑2003͒
56.9 ͑WLC͒
PS6 melt 390 1.06 130 28 23.4 ͑WLC͒ Bach et al. ͑2003͒
PS7 soln 3900 1.05 21 27.5 0.023 ͑WLC͒ Bhattacharjee et al. ͑2003͒
PB soln 813 Ͻ1.05 25 21 0.15 Menezes and Graessley ͑1980͒
1117CUBS MODEL FLOW DATA COMPARISON
=18 100 g/mol ͓Ferry ͑1980͔͒, the ͗Z͘eq for the melt is estimated to be 11 at 160 °C. The
value of ␶e for the melt is obtained as, ␶e
NCR
=0.4 s and ␶e
CR
=0.78 s at 160 °C and Ne is
taken as 22.
Data for a nearly monodisperse polybutadiene in Flexon 391 solution ͑PB/Flexon391,
polydispersity index Ͻ1.05͒ with a concentration of 7 wt % and Mw=813 kDa is taken
from ͓Menezes and Graessley ͑1980͔͒. The experiment was done at 25 °C with a poly-
mer density of 0.972 g/cm3
.
As is typically done, we first estimated the number of entanglements by using the
plateau modulus. However, further refinement was necessary after comparison with LVE
data. Therefore, we performed a detailed fit of the linear viscoelastic prediction with the
experimental data in order to fix the value of ͗Z͘eq at approximately 21 ͑see Fig. 6͒. The
estimated value of Ne for polystyrene solution is 545. However, we take Ne=50 in order
to save computational time. The value of ␶e for the melt is estimated as ␶e
NCR
=0.15 s from
the LVE comparison. Fits for PS1 and PS2 are given elsewhere ͓Nair and Schieber
FIG. 2. A comparison of the storage and loss moduli predictions with and without constraint release and that
for polystyrene melt PS4 ͓Schweizer et al. ͑2004͔͒.
FIG. 3. A comparison of the storage and loss moduli predictions and those measured by ͓Bach et al. ͑2003͔͒ for
PS5.
1118 SCHIEBER, NAIR, AND KITKRAILARD
͑2006͔͒. The remaining fits are shown in Figs. 2–6. Some predictions are found using the
worm-like chain ͑WLC͒ free energy, which is discussed below. We find no difference
between predictions using the Gaussian free energy and the WLC for LVE, which is
expected for small deformation flows.
FIG. 4. A comparison of the storage and loss moduli predictions and those measured by ͓Bach et al. ͑2003͔͒ for
PS6.
FIG. 5. A comparison of the storage and loss moduli predictions and those measured by ͓Bhattacharjee et al.
͑2003͔͒ for PS7 solution.
FIG. 6. Storage and loss moduli predictions of the CUBS model with constraint release, and the experimental
data of ͓Menezes and Graessley ͑1980͔͒ for a PB solution of Mw =813 kDa.
1119CUBS MODEL FLOW DATA COMPARISON
A high-molecular-weight polystyrene melt of ͓Collis et al. ͑2005͔͒ at 170 °C with
Mw=500.5 kDa and PI index of 1.02 is PS3. By using MeϷ18 000, the average number
of entangled strands, ͗Z͘eq, is 28. The G*
data/theory comparisons are shown in Fig. 7
without constraint release. A low-molecular-weight polystyrene melt of ͓Schweizer et al.
͑2004͔͒, MwХ200 kDa is called PS4. The estimated value of ͗Z͘eq is 11. In both systems,
the average number of Kuhn steps, Ne, is 22.
From the predictions obtained from the CUBS model without constraint release, the
characteristic time for the high molecular weight and the low molecular weight system
are 0.0375 and 0.137 s, respectively.
IV. COMPARISON IN FLOW
We now study in detail the nonlinear predictions of CUBS with constant chain friction.
All the flow predictions made here are without adjusting any parameters.
FIG. 7. A comparison between the storage GЈ and loss GЉ moduli predictions of the model without constraint
release and that for polystyrene melt PS3 ͓Collis et al. ͑2005͔͒, Mw =500.5 kDa, ͗Z͘eq=28 at 170 °C.
FIG. 8. Ratio of nonlinear to linear shear modulus as functions of dimensionless time for different assumptions
in the model and an approximate, analytic result found by assuming a separation of time scales between
retraction and chain reorientation. In this comparison, the chain has an average of 35 entanglements, Ne =22,
and the strain is ␥=6.3.
1120 SCHIEBER, NAIR, AND KITKRAILARD
A. Single-step strain
The earliest tube model derived by Doi and Edwards made a remarkable parameter-
free comparison with the strain-dependent part of single-step strain data. However, future
refinements seemed to worsen this very favorable comparison. Earlier, one of us derived
an analytic expression for the damping function based on the slip-link model under
certain assumptions ͓Schieber ͑2003a͔͒. We now check this analytic result through nu-
merical solution of the model.
A step strain is imposed on a material instantaneously at time t=0. The relaxation of
shear stress can be expressed in terms of the nonlinear relaxation modulus, G͑t,␥͒,
defined as
G͑t,␥͒ = −
␶yx
␥
, ͑12͒
where ␥ is the magnitude of the imposed strain. Beyond a certain time ␶k that depends on
the polymer system, the nonlinear relaxation modulus shows separability into time- and
strain-dependent parts over a wide range of strains
FIG. 9. Shear stress relaxation following single-step strain as functions of time for several strains predicted by
the model with and without CR/CCR and PS2 data ͓Nair ͑2004͔͒.
FIG. 10. Shear stress relaxation following double-step strain of the type-B flow with ␥1 =4 and ␥2 =−2 at
several values of t1 predicted by the model with and without CR/CCR and PS1 data ͓Venerus and Kahvand
͑1994a͔͒.
1121CUBS MODEL FLOW DATA COMPARISON
G͑t,␥͒ = G͑t͒h͑␥͒, t ജ ␶k, ͑13͒
where G͑t͒ is the relaxation modulus of linear viscoelasticity and h͑␥͒ is the damping
function. In the limit of small strains, the nonlinear relaxation modulus becomes G͑t͒.
The damping function is a monotonically decreasing function of strain.
Figure 8 shows a comparison of the ratio of nonlinear to linear shear modulus pre-
dicted by different models for ͗Z͘eq=36 and the analytic result. We see that the model
with CR/CCR and constant chain friction shows better agreement with the analytic h͑␥͒.
In Fig. 9 we show comparisons of shear stress relaxation of PS melt ͑Mw=200 K,
͗Z͘eq=11͒ as functions of time for several strains. Except for the highest strain, the model
predictions with convective constraint release look good. A previous article compared the
damping function predictions with data, and the curves were a bit more strain softening
than either of Doi and Edwards’ predictions. Therefore, our theory appears to continue
the venerable tradition of degrading the Doi–Edwards-model prediction of the damping
function, but improving all other comparisons, as we now show.
B. Reversing double-step strain
Double-step strain experiments provide a more severe test for evaluating the rheologi-
cal constitutive theories for concentrated polymer systems. In these experiments, a strain
of ␥1 is imposed at time t=0, followed by a strain of ␥2 after a time t1, in the same
͑␥2Ͼ0͒ or opposite ͑␥2Ͻ0͒ direction to the first step strain. In this study, we consider
only reversing double-step strain flows where the second strain is applied in a direction
opposite to the first. Our model is used here for predicting two types of reversing flows,
type-B ͑␥2=−␥1/2͒ and type-C ͑␥2=−␥1͒.
It is known that the Doi–Edwards ͑DE͒ model with the independent alignment ͑IA͒
assumption ͓Doi and Edwards ͑1986͔͒ which is a particular case of, and performs iden-
tically to the K-BKZ model ͓Bernstein et al. ͑1963͒; Kaye ͑1962͔͒ in step strains does not
perform well in predicting reversing double-step strain deformation. Double-step strain
experiments have also been used to test the DE model with and without the IA assump-
tion ͓Brown and Burghardt ͑1996͒; Osaki et al. ͑1981͒; Osaki and Kurata ͑1982͒; Venerus
and Kahvand ͑1994a, 1994b͒; Venerus et al. ͑1990͔͒.
Doi derived a rigorous constitutive equation by avoiding the IA assumption ͓Doi
͑1980͔͒ but still assuming instantaneous chain retraction, which was able to explain
reversing double-step strain deformation as long as the time between the two steps is
FIG. 11. First normal stress relaxation for the type-B flow with ␥1 =4 and ␥2 =−2 at t1 =4 s predicted by the
model with and without constraint release, and PS1 data of ͓Brown and Burghardt ͑1996͔͒.
1122 SCHIEBER, NAIR, AND KITKRAILARD
much larger than the retraction time ͑␶R͒. However, when t1ഛ␶R, the model fails ͓Brown
and Burghardt ͑1996͒; Venerus and Kahvand ͑1994a, 1994b͔͒. Also, the equation derived
by Doi is not applicable to any other flows.
Model predictions for type-B flows with ␥1=4 and ␥2=−2 are compared with PS1 in
Fig. 10. We see that except at short times, the model shows good agreement with experi-
mental data and the results are not strongly sensitive to constraint release. The discrep-
ancy shown at short times might be attributed to the coarse-graining introduced in the
model. As mentioned before, the model cannot capture the relaxations taking place below
the strand size. Figure 11 shows the comparison of first normal stress ͑N1͒ relaxation at
t1=4 s. Here the model without CR/CCR gives better predictions compared to the pre-
dictions of K-BKZ and DE theories shown in Fig. 11 of the article by Venerus and
Kahvand ͑1994b͒.
Figure 12 shows comparison of the second normal stress ͑N2͒ relaxation at t1=1 s. As
in the case of N1, we see that the model without CR/CCR gives better predictions for N2,
except at short times.
In Fig. 13, we show the comparisons of the shear stress relaxation for type-C flows
with ␥1=4 and ␥2=−4 at t1=2 and 32 s. Here the data seem to show a similar shape for
FIG. 12. Second normal stress relaxation for type-B reversing double-step strain with ␥1 =4 and ␥2 =−2 at t1
=1 s predicted by the model with and without constraint release, and PS1 data from Brown and Burghardt
͑1996͒.
FIG. 13. Shear stress relaxation for type-C flow with ␥1 =4 and ␥2 =−4 at t1 =2 and 32 s predicted by the model
with and without CR/CCR and PS1 data ͓Venerus and Kahvand ͑1994a͔͒.
1123CUBS MODEL FLOW DATA COMPARISON
different waiting times, and just a vertical displacement. The model without CCR show
the same feature, but CCR does not. Nonetheless, the model does predict the correct
magnitudes for the shear stress. Figure 14 shows the transient normal stress ratio for
type-C flows at t1=1, 4, and 16 s. Here, we see that the normal stress ratio is time
dependent and also t1 dependent. The trend is shown by experimental data in Fig. 6͑a͒ of
Brown and Burghardt ͑1996͒.
C. Inception of steady shear flow
Figures 15 and 16, compare the model predictions to the PS1 solution data for tran-
sient growth of shear viscosity and first normal stress difference, respectively, at several
shear rates. In both cases, at high shear rates the model without CR and CCR captures the
features of the data quantitatively, whereas, at low shear rates the model with CR and
CCR predicts the data well. The linear viscoelastic limit forms an upper envelope for all
the shear rates and the transient overshoot occur at medium and high shear rates, in
accordance with experiments. It is interesting to note that at high shear rates, the model
without CR/CCR captures both the overshoot and undershoot, while CR/CCR makes the
undershoot disappear.
FIG. 14. Second- to first-normal-stress ratio for type-C flow with ␥1 =4 and ␥2 =−4 at t1 =1, 4, and 16 s ͑from
top to bottom͒ predicted by the model with and without constraint release.
FIG. 15. Transient viscosities as functions of time under start up of steady shear at several shear rates predicted
by the model with and without constraint release, and he experimental data of polystyrene solution PS1
͓Kahvand ͑1995͔͒.
1124 SCHIEBER, NAIR, AND KITKRAILARD
These data can also be examined in terms of the extinction angle, Fig. 17. We estimate
the extinction angle as ␹= 1
2 tan−1
͑2␶yx/N1͒, which is valid within our Gaussian approxi-
mation. Here also, the model without CR/CCR shows better agreement at high shear
rates, while the model with CR/CCR gives better predictions at low shear rates. We can
see that the theory is able to capture the undershoot in extinction angle at high shear rates.
High deformation rate appears to suppress the effect of constraint release. At any rate,
constraint release does not have a large effect on predictions, which contradicts the
predictions of tube models ͓Ianniruberto and Marrucci ͑1996, 2000͒; Marrucci ͑1996͔͒.
We also examine polystyrene melts PS3 and PS4. Figure 18 shows a comparison
between the CUBS model and the experimental viscosities of PS4. There is some dis-
crepancy between the LVE predictions and the data at small strains, suggesting some
experimental difficulties in this range. Taking this inconsistency into account the theoret-
ical predictions are reasonable, especially at the higher Deborah numbers.
Figure 19 shows a similar comparison for PS3. These data show an even larger
discrepancy with the LVE curve at small strains, suggesting some experimental error,
possibly an overestimate of sample size. Interestingly, the difference between the data
FIG. 16. Transient growth of first normal stress difference as functions of time under start up of steady shear
flow at several shear rates predicted by the model with and without constraint release, and the experimental data
of polystyrene solution PS1 ͓Kahvand ͑1995͔͒.
FIG. 17. Transient extinction angle as functions of strain under start up of steady shear flow at several shear
rates predicted by the model with and without CR/CCR and experimental data of polystyrene solution PS1
͓Kahvand ͑1995͔͒.
1125CUBS MODEL FLOW DATA COMPARISON
FIG. 18. Transient viscosities as functions of time under start up of steady shear at several shear rates predicted
by the model with and without constraint release, and the experimental data of polystyrene melt PS4 ͓Schweizer
et al. ͑2004͔͒.
FIG. 19. Transient viscosities as functions of time under start up of steady shear at several shear rates predicted
by the model with and without constraint release and the experimental data of polystyrene melt PS3 ͓Collis et
al. ͑2005͔͒.
FIG. 20. Transient viscosities as functions of time under start up of steady shear at several shear rates predicted
by the model with and without constraint release, and the experimental data of the polybutadiene solution
͓Menezes and Graessley ͑1980͔͒.
1126 SCHIEBER, NAIR, AND KITKRAILARD
and theory is nearly identical to the discrepancy between data and the LVE curve, which
would also agree with our speculation about the sample’s being somewhat smaller than
assumed.
Predictions from the CUBS model also show good agreement with the experimental
results for the PB solution data in Figs. 20 and 21.
D. Steady shear flow
The original Doi–Edwards model predicts a maximum in the shear stress with shear
rate at steady state, whereas experiments predict a monotonic increase of shear stress with
shear rate. The DE model predicts a power-law index of 3
2 for the steady-state viscosity
in shear flow ͓Doi and Edwards ͑1979͔͒.
Öttinger considered a modification of the Doi–Edwards and Curtiss–Bird models to
include the effect of constraint release. This modified reptation model predicts an im-
proved power-law index of 4
3 for the steady-state viscosity ͓Öttinger ͑1994͔͒. Ianniruberto
and Marrucci suggested a modification of the Doi–Edwards tube model to include the
mechanism of convective constraint release which is claimed to be important during fast
flows of polymer melts. This model predicted that the steady-state shear stress asymp-
totically approaches a plateau at high shear rates ͓Ianniruberto and Marrucci ͑1996͔͒, and
it has been widely believed that CCR is a necessary mechanism to avoid a local maxi-
mum. More about this point is discussed in the conclusions.
Before making quantitative comparisons with data, we looked at the question of
monotonicity in the steady-state shear-stress curve for a more highly entangled chain:
͗Z͘eq=30 and Ne=22. Figure 22 shows the result in dimensionless form. It should be
noted that the Deborah number is not in terms of the longest relaxation time, but rather
De=␥˙␶e.
Note that things like constraint release and Rouse timescales arise from the model,
rather than being added to the model. Therefore, the monotonic growth of the shear stress
is a result of the assumed physics. In fact, in this calculation constraint release is switched
off. Therefore, we can conclude that CCR is unnecessary to avoid the maximum. Second,
the effect of stretch can clearly be seen for De values greater than approximately 0.2.
From the same curve we can estimate the onset of the longest relaxation time at about
FIG. 21. Transient growth of first normal stress difference as functions of time under start up of steady shear
flow at several shear rates predicted by the model with and without constraint release, data of the polybutadiene
solution ͓Menezes and Graessley ͑1980͔͒.
1127CUBS MODEL FLOW DATA COMPARISON
De=3ϫ10−3
. This gives us a ratio of ϳ67 for the longest relaxation time to the Rouse
time, in accord with the usual estimate of 3͗Z͘eq for tube models, realizing that primitive-
path length fluctuations should lower this estimate.
Figure 23 shows the shear stress predictions and data for PS1. Only a slight depen-
dence is seen on CR. As expected, no maximum is visible. The steady-state viscosity ␩
and the first normal stress difference coefficient ⌿1 for PS1 have power-law indices of
0.79 and 1.4, respectively. The model shows excellent agreement, independent of con-
straint release, with power-law indices of 0.82 and 1.37, respectively, when CR/CCR is
present and 0.82 and 1.38 when CR/CCR are switched off. The small discrepancy could
be due to the small amount of polydispersity present in the polystyrene system used here.
The shear rates used to evaluate the power-law indices range from 0.1 to 40 s−1
.
Similarly good agreement in the power-law indices are found ͑not shown͒ for the
steady state viscosity as a function of shear rate for PS4, with a power-law index of 0.84
for ␩ when CR/CCR is not present in the model. The polybutadiene solution data exhibit
power-law indices of 0.87 and 1.65 respectively, for ␩ and ⌿1 when CR is not present in
the model, in agreement with predictions.
FIG. 22. Dimensionless steady-state shear stress as a function of dimensionless shear rate De=␶e␥˙ predicted by
the theory without constraint release, ͗Z͘eq=30 and Ne =22. The solid line is the linear viscoelastic prediction
obtained from the CUBS model.
FIG. 23. Steady-state values of shear stress as functions of shear rate under steady shear flow predicted by the
model with and without constraint release, and experimental data of polystyrene solution PS1 ͓Kahvand
͑1995͔͒.
1128 SCHIEBER, NAIR, AND KITKRAILARD
E. Cessation of steady shear flow
Figure 24 shows the relaxation of shear stress following cessation of steady shear flow
at two shear rates for PS1. We can see that the relaxation is faster at the higher shear rate
than at the smaller shear rate, which is due to the faster chain retraction. These trends are
consistent with theory. At higher shear rate ͑1.0 s−1
͒, the models with and without CR/
CCR shows good predictions at shorter times since relaxation is mainly due to monomer
density fluctuation and contour-length fluctuation. At larger times, the model with CR/
CCR shows better agreement since relaxation is mainly due to reptation and double
reptation.
F. Exponential shear flow
In exponential shear flow the shear strain increases exponentially with time, given by
␥͑t͒ = exp͑␣t͒ − exp͑− ␣t͒, ͑14͒
where ␣ is a parameter governing the shear acceleration.
Several experimental ͓Demarquette and Dealy ͑1992͒; Doshi and Dealy ͑1987͒; Sam-
urkas et al. ͑1989͒; Sivashinsky et al. ͑1984͒; Venerus ͑2000͒; Zülle et al. ͑1987͔͒ and
theoretical ͓Kwan and Shaqfeh ͑1999͒; Neergaard et al. ͑2000͒; Graham et al. ͑2001͔͒
studies on exponential shear flow have been reported during the past two decades. Neer-
gaard et al. used the full chain reptation model ͓Hua and Schieber ͑1998b͔͒ to interpret
the exponential shear flow data for entangled polystyrene solution. Based on their model
predictions, they reached the conclusion that exponential shear flow stretches the chains
no more than does inception of steady shear ͓Neergaard et al. ͑2000͔͒. In fact, that work
showed that all exponential shear flow data could be described by steady shear flow data
when plotted versus instantaneous shear rate. Hence, we show all comparisons in that
form here.
In Figs. 25 and 26, we show the comparisons of unadulterated shear stress and first
normal stress difference, respectively, as functions of instantaneous shear rates for PS1.
We can see that the model gives excellent predictions of shear stress except at high ␣,
where the model with CR/CCR slightly overpredicts stretch. In the N1 predictions, the
model shows good agreement at low ␣, but again somewhat overpredicts stretch at
high ␣.
FIG. 24. Relaxation of shear stress as functions of time following cessation of steady shear flow at two shear
rates: model predictions with and without constraint release, and experimental data from ͓Kahvand ͑1995͔͒.
1129CUBS MODEL FLOW DATA COMPARISON
FIG. 25. Shear stress as functions of instantaneous shear rate for exponential shearing flow at several values of
␣ predicted by the model with and without constraint release, and PS1 data ͓Neergaard et al. ͑2000͔͒.
FIG. 26. First normal stress difference as functions of instantaneous shear rate for exponential shear flow at
several values of ␣: model predictions with and without constraint release, and PS1 data ͓Neergaard et al.
͑2000͔͒.
FIG. 27. Transient elongational stress as a function of time, predicted by the CUBS model with Gaussian and
wormlike chain free energy and PS5 melt, Bach et al. ͑2003͒ Mw =200 kDa.
1130 SCHIEBER, NAIR, AND KITKRAILARD
G. Elongational flow
Only recently has it been possible to achieve rheometric control for uniaxial extension
of entangled, linear polymers. Concentrated solutions ͑PS7͒ have been examined at Mo-
nash University ͓Bhattacharjee et al. ͑2003͔͒, and melts ͑PS5 and PS6͒ at the Danish
Technical University ͓Bach et al. ͑2003͔͒. Note that systems PS6 and PS7 are particularly
interesting, because they have nearly identical entanglement number, but only one is
solution. We first examine the transient melt data, PS5, in Fig. 27. Here we see the first
clear failure of the model. Whereas the data show a monotonic approach to steady state,
the model exhibits oscillation until very large strains.
Nonetheless, if we compare the steady elongational stresses, we see in Fig. 28, that the
slip-link model is able to capture the steady elongational stresses very well.
It is natural to guess that the infinite-extensibility assumption in the model is the cause
of the discrepancy with the elongational data. Therefore, we also considered a free energy
that prevents any segment from being stretched past its contour length. Namely, we use a
free energy that becomes infinitely large as Q→NaK. Here we use an approximation to
the worm-like chain free energy for an entangled strand
FS͑Q,N,T͒
kBT
=
3Nlp − 2Q
4͑Nlp − Q͒
Q2
Nlp
2 − log J͑N͒, ͑15͒
which we derive from the approximate tension expression of ͓Marko and Siggia ͑1995͔͒.
The function J͑N͒ is a normalization factor dependent upon the number of persistence
lengths in the strand, lp=aK/2 is the persistence length and now N is the number of
persistence lengths instead of the number of Kuhn steps. The function J is defined as
J͑N͒ ª ͫ4␲͑Nlp͒3
͵0
Nlp
Q2
expͭ3Nlp − 2Q
4͑Nlp − Q͒
Q2
Nlp
2 ͮdQͬ−1
. ͑16͒
Since we cannot perform the integration analytically, we use a minimax approximation to
obtain a numerical expression for J. Details about the numerical expression can be found
in the Appendix.
Figures 27 and 28 also show the results for the slip-link model with finite extensibility.
Surprisingly, we see that finite extensibility ͑or, more accurately in this case, “finite
monomer density”͒ plays only a small role in the transient elongational stresses, and
FIG. 28. Steady state elongational stress, predicted by the CUBS model with Gaussian and worm-like chain
free energy and P5 melt, Bach et al. ͑2003͒ Mw =200 kDa.
1131CUBS MODEL FLOW DATA COMPARISON
almost no role in steady elongational stresses. Similar comparisons are made for the more
highly entangled sample, PS6, in Figs. 29 and 30, with similar conclusions.
We now turn to the elongational data for the entangled solution, PS7. Recall that this
sample has the same number of entanglements per chain as the earlier melt, PS5, whose
steady elongational data the theory was able to capture ͑though not the transients͒. Except
for the smallest elongation rate, the slip-link model with finite extensibility is unable to
capture the transient data ͑Fig. 31͒. However, unlike the melt comparison, the predictions
substantially underpredict the steady stress data in Fig. 32, and even show the wrong
shape.
To examine this discrepancy further, we place the solution and melt data on the same
plot by making the strain rate dimensionless by ␶e, and the elongational viscosity dimen-
sionless by the zero-shear-rate viscosity. Recall that both of these dimensional quantities
are determined by comparison at LVE. Figure 33 clearly shows that the melt and the
solution trends are opposite: the solution is strain hardening and the melt is strain soft-
ening at high strain rates.
Theories for concentrated systems to date have assumed an equivalence between en-
tangled solutions and entangled melts, at least implicitly by assuming that entanglements
determine the dynamics. Clearly, such an assumption is not in agreement with our final
FIG. 29. Transient elongational stress as a function of time, predicted by the CUBS model with worm-like
chain model and PS6 melt, Bach et al. ͑2003͒ Mw =390 kDa.
FIG. 30. Steady state elongational stress, predicted by the CUBS model with worm-like chain model and PS
melt, Bach et al. ͑2003͒ Mw =390 kDa.
1132 SCHIEBER, NAIR, AND KITKRAILARD
figure. There are three possibilities: ͑1͒ one of the data sets is ͑or both are͒ incorrect, ͑2͒
additional physics is important in distinguishing solutions from melts, or ͑3͒ an assump-
tion in the model is incorrect.
Since the elongational data have not yet been reproduced in other labs, the first option
cannot be ruled out—easily said for a theoretician.
FIG. 31. Transient elongational stress as a function of time, predicted by the CUBS model with worm-like
chain model and PS7, Bhattacharjee et al. ͑2003͒ Mw =3900 kDa.
FIG. 32. Steady state elongational stress, predicted by the CUBS model with Gaussian chain model and
worm-like chain model and PS7, Bhattacharjee et al. ͑2003͒ Mw =3900 kDa.
FIG. 33. Dimensionless steady state elongational viscosities comparison between polystyrene melt, Bach et al.
͑2003͒ ͑390 kDa at 130 °C͒, and polystyrene solution, Bhattacharjee et al. ͑2003͒ ͑3.9 MDa at 21 °C͒.
1133CUBS MODEL FLOW DATA COMPARISON
We also note that polymer density fluctuations are much larger for the solutions than
they are for the melts. Such fluctuations are not considered in the model, since entangle-
ments are considered fixed in space for LVE, and to move affinely for flows. These
fluctuations might explain the second option. A straightforward way to examine this
assumption is in a multichain simulation of the sort used by ͓Masubuchi et al. ͑2001͔͒.
There are a number of assumptions made in the model: constant chain friction; mim-
icking constraint release through slip-link diffusion, instead of creation and destruction of
entanglements in the middle of the chain; destruction of the entanglements at the end of
the chain through a probability, instead of as a boundary condition; slip links are de-
formed affinely. Discrepancy with data because of the first of these assumptions seems
unlikely because constant-entanglement friction has also been tested and found to be
somewhat inferior at predicting shear and elongational stresses. The second and third
assumptions are currently being examined in a more rigorous formulation of the model.
The fourth assumption is more problematic to examine in a mean-field model, such as
assumed here. Again, a simulation similar to that of ͓Masubuchi et al. ͑2003͔͒ might be
more appropriate to examine this assumption, since the motion of the entanglements
could be followed rather than assumed. We do note in passing that we have also exam-
ined alternative assumptions about entanglement motion in light of thermodynamic con-
siderations. However, these calculations found that fluctuating entanglement motion
yielded affine motion, on average.
V. CONCLUSIONS
We introduce a consistently unconstrained Brownian slip-link model with constant
chain friction to study the nonlinear properties of linear entangled polymer melts. The
model uses a single phenomenological parameter ␶e, which is fit by the linear viscoelas-
ticity data and all the flow predictions are made without any adjustable parameters.
We found good to excellent agreement with all shear flows studied: single step, re-
versing double step, inception of steady shear, steady shear, cessation of steady shear, and
exponential shear. Results were not very sensitive to CR—in contrast with tube models.
Studies using various formulations of tube models consistently show that constraint re-
lease substantially reduces the shear thinning seen in steady shear flow. However, the
studies do not make clear whether CCR is sufficient to avoid the maximum altogether.
For example, ͓Marrucci and Ianniruberto ͑2003͔͒ include a prefactor ␤ that should be
order 1, but is set to 2 to avoid the maximum. The original article including CCR
͓Ianniruberto and Marrucci ͑2000͔͒ does not use the factor of 2 and exhibits a plateau in
the shear stress—still implying flow instabilities. The more recent model of ͓Leygue et al.
͑2006͔͒ also predicts a local maximum in the shear stress and the authors point out that
the ␤ parameter of Marrucci et al. could be used as a fix, but decline to do so and write
“We nevertheless believe that a better understanding of convective constraint release,
especially in reversing flows, could lead to improved formulations where such modifica-
tions would not be necessary.” Fang et al. ͑2000͒ decrease the stretching time constant in
their stochastic, single-segment model by a factor greater than 2 to give an onset of
stretch at smaller shear rates to avoid the maximum. Mead et al. ͑1998͒ decrease the ratio
␶d/␶S by approximately 80% to describe their polystyrene solution of 17 entanglements.
They argue that primitive-path length fluctuations justify the change, but the predictions
show a change in slope of stress with shear rate ͑from stretch͒ that are not evident in
either the shear-stress or normal-stress data. In a more-detailed formulation, ͓Milner et al.
͑2001͔͒ adjust their parameter c␯ to a value of 0.1 in order to avoid the maximum. The
physical reasoning given for the parameter’s being less than one is that more than two
1134 SCHIEBER, NAIR, AND KITKRAILARD
chains might make up an entanglement ͓see the Appendix of Nair and Schieber ͑2006͒ for
an alternative explanation͔. However, ten chains would be inconsistent with the number
predicted by the packing criteria of Fetters et al. ͑1999a, 1999b͒ using Rault’s argument
͓Rault ͑1987͔͒, or of the binary interactions found in more recent atomistic simulations
͓Everaers et al. ͑2004͔͒. Our earlier full-chain stochastic model with convective con-
straint release ͓Hua et al. ͑1999͔͒ ͑which mixes tubes and slip links͒ also predicts a
plateau in the shear stress curve. Since the constraint release is found in a self-consistent
way, there are no adjustable parameters like ␤ or c␯. It is possible that tube models predict
instability even with convective constraint release.
On the other hand, in the dual slip-link model ͓Doi and Takimoto ͑2003͔͒, the shear
stress increases monotonically with shear rate in steady shear flow but they do not make
any quantitative comparison with experiments.
Tubologists might wonder why our slip-link model shows something different from
tube models. We mention here three possible explanations. First, we note that tube mod-
els typically assume that the number of entanglements on a chain are fixed at a single
number—there are no fluctuations at equilibrium, and flow does not change it. At equi-
librium, the tube is a random walk with constant step length. It is unclear how this
assumption is modified with flow, given that most tube models are on the level of a single
step in the primitive path.
On the other hand, our slip-link formulation has a distribution of entanglement num-
bers at equilibrium, and a distribution of strand lengths; flow modifies both of these
distributions in well defined ways. An earlier formulation of the model with constant
entanglement friction ͓Schieber et al. ͑2003͔͒ gives results qualitatively similar to this
one, but shows how shear flow modifies the entanglement number distribution. In that
plot we show that the distribution remains Poissonian, but decreases to about half at high
Deborah numbers. The observed disentanglement has two effects that lead in opposite
directions. First, chains with fewer entanglements have longer strands that are largely
aligned with the flow. These would be expected to contribute to more shear thinning.
However, when these chains become dangling ends, they reëntangle rather quickly, lead-
ing to disordered strands, reducing shear thinning.
We mention that the flow stress curves in that paper are in error, though all equations
and dynamic results are correct. This was due to a bug in the portion of the code that
calculated stress from chain conformations. When this bug is fixed, all results we have
studied for constant-entanglement friction vary only slightly from those with constant-
chain friction. However, these results are slightly better for the latter.
It is also possible that the difference might be connected to the fact that the dynamics
in the slip-link model are connected to the stress through different derivatives of the free
energy, whereas in tube models, chain tension itself determines both stress and dynamics.
In other words, in the formulation proposed here the derivative of free energy with
respect to monomer number is the chemical potential, which determines dynamics; the
derivative with respect to slip-link position is tension, which determines stress. As a
result, tension along the chain is not equilibrated for strain rates between the inverse of
the longest relaxation time, and the inverse of the Rouse time—not true for tube models.
For these strain rates, the chemical potential of the chain ends stays near the equilibrium
value of −1/2Ne. Hence, chemical potential balance gives the number of Kuhn steps in
the strand as approximately N/NeХ͑3/2͓͒ͱ1+͑8Q2
/3NeaK
2
͒−1͔. Therefore, the mono-
mer density N/Q is not constant along the chain, but varies in a nonlinear way with Q. In
tube models, equating the tension T=3kBTQ/NaK
2
with the equilibrium Maxwell-demon
1135CUBS MODEL FLOW DATA COMPARISON
tension kBT/ͱNeaK requires that the monomer density N/Q stay constant. The maximum
observed in tube models might be because the tension—and, hence, the monomer
density—is constant along the chain at these strain rates.
A third possibility is that the local maximum is avoided because the strands have a sort
of built-in polydispersity in lengths, which smooths out the stress curve. One might
imagine trying to test these ideas by creating chains with more monodisperse lengths.
However, such a formulation would violate the equilibrium conformation statistics valid
for concentrated chains. It is also difficult to imagine a test for the second hypothesis that
would not violate thermodynamics. The first postulate might be tested by imposing more
entanglements in the middle of the chain in order to keep the entanglement density
constant. We have examined these dynamics, but obtained stress predictions that were
very far from those observed, and different from tube model predictions.
The model was able to describe steady elongational data for polystyrene melts, but
was unable to describe the transient melt elongational data, nor the solution elongation
data. The former discrepancy is likely due to the assumption of how entanglements are
destroyed in the model. Evidence of this is seen in the phase-space plot of Fig. 34, where
the transient elongational viscosity is plotted versus the average number of entangled
strands, and time is parametric. Here we see that disentanglement precedes a drop in the
viscosity, followed by reëntanglement that precedes a rise in the viscosity.
The latter discrepancy between solution and melt data is more puzzling. If the data
hold, it suggests that entangled melts and solutions are not universal and that local
polymer concentration fluctuations can play an important role. A more-detailed model
would be necessary to examine this hypothesis.
ACKNOWLEDGMENT
This material is based upon work supported by the National Science Foundation under
Grant No. NSF-SCI 0506305.
FIG. 34. A phase-space plot of transient dimensionless elongational viscosity versus number of entangled
strands, showing that the number of entanglements is highly correlated with the oscillations seen in the elon-
gational viscosity. For these calculations, ͗Z͘eq=14, Ne =44␶−e␥˙ =0.453, and the ensemble size is 2000.
1136 SCHIEBER, NAIR, AND KITKRAILARD
APPENDIX: WORM-LIKE CHAIN FREE ENERGY
To Eq. ͑16͒, we fit a numerical expression for J,
J͓y͑N͔͒ = expͫA1 + A2y + A3y2
+ A4y3
+ A5y4
1 + A6y + A7y2
+ A8y3
+ A9y4 ͬ, ͑A1͒
where y͑N͒=log N. Values for the constants are given in Table II.
The chemical potential for this free energy is found by taking the differential of the
free energy with respect to N; hence, a derivative of J is necessary. Rather than take a
derivative of our numerical approximation, we first take the derivative analytically
͑A2͒
TABLE III. Constants for the numerical approximation to integral term in
the chemical potential expression ␮0, Eq. ͑A4͒.
Constant Value
M1 −0.5215600145979548
M2 −0.5797035530883076
M3 −0.17660236446248723
M4 −0.10983496030144273
M5 −0.008777278719472151
M6 −0.01894111365316676
M7 0.11070525164370931
M8 0.132323363265643
M9 0.015497437639395373
M10 0.018962828443251357
TABLE II. Constants for the numerical approximation to the normaliza-
tion function J, Eq. ͑A1͒.
Constant Value
A1 0.44080624243588207
A2 2.472397669388139
A3 0.2431635769831712
A4 0.21394499795640387
A5 0.03344400384360319
A6 0.1468954480518177
A7 0.09151779095188171
A8 0.02361837127994027
A9 0.000036340710810897935
1137CUBS MODEL FLOW DATA COMPARISON
where we define g͑x͒ as
g͑x͒ =
3 − 2x
4͑1 − x͒
x2
. ͑A3͒
Then, from Eq. ͑A2͒ we fit a numerical approximation to the ratio of integrals, which we
call ␮0,
␮0 = expͫM1 + M2y + M3y2
+ M4y3
+ M5y4
+ M6y5
1 + M7y + M8y2
+ M9y3
+ M10y4 ͬ, ͑A4͒
where again y=log N. Values for the constants are given in Table III.
In Figs. 35 and 36, respectively, we check our expressions by plotting the relative
residual ͑Jfit−J͒/J between the numerical solutions obtained from MATHEMATICA for the
integrals and the numerical expression of normalization factor, Eq. ͑A1͒, and the integral
term in the chemical potential expression, Eq. ͑A4͒.
FIG. 35. Relative residual between the numerical expression of the normalization factor and the numerical fit
found by using MATHEMATICA.
FIG. 36. Residual between the numerical expression of the integral term of chemical potential expression and
the numerical solution from MATHEMATICA software.
1138 SCHIEBER, NAIR, AND KITKRAILARD
References
Bach, A., K. Almdal, H. K. Rasmussen, and O. Hassager, “Elongational viscosity of narrow molar mass
distribution polystyrene,” Macromolecules 36, 5174–5179 ͑2003͒.
Bernstein, B. E., A. Kearsley, and L. J. Zapas, “A study of stress relaxation with finite strain,” Trans. Soc.
Rheol. 7, 391–410 ͑1963͒.
Bhattacharjee, P., D. Nguyen, G. McKinley, and T. Sridhar, “Extensional stress growth and stress relaxation in
entangled polymer solutions,” J. Rheol. 47, 269–290 ͑2003͒.
Bird, R. B., C. F. Curtiss, R. C. Armstrong, and O. Hassager, Dynamics of Polymeric Liquids Vol. II: Kinetic
Theory, 2nd ed. ͑Wiley-Interscience, New York, 1987͒.
Brown, E. F., and W. R. Burghardt, “First and second normal stress difference relaxation in reversing double-
step strain flows,” J. Rheol. 40, 37–54 ͑1996͒.
Collis, M. W., A. K. Lele, M. R. Mackley, R. S. Graham, A. E. Groves, A. E. Likhtman, T. M. Nicholson, O.
G. Harlen, T. C. B. McLeish, L. R. Hutchings, C. M. Fernyhough, and R. N. Young, “Constriction flows of
monodisperse linear entangled polymers: Multiscale modeling and flow visualization,” J. Rheol. 49, 501–
522 ͑2005͒.
Demarquette, N. R., and J. M. Dealy, “Nonlinear viscoelasticity of concentrated polystyrene solutions: Sliding
plate rheometer studies,” J. Rheol. 36, 1007–1032 ͑1992͒.
Doi, M., “Stress relaxation of polymeric liquids after double-step strain,” J. Polym. Sci., Polym. Phys. Ed. 18,
1891–1901 ͑1980͒.
Doi, M., and S. F. Edwards, “Dynamics of concentrated polymer systems. Part 1. Brownian motions in the
equilibrium state,” J. Chem. Soc., Faraday Trans. 2 74, 1789–1800 ͑1978a͒.
Doi, M., and S. F. Edwards, “Dynamics of concentrated polymer systems. Part 2. Molecular motion under
flow,” J. Chem. Soc., Faraday Trans. 2 74, 1802–1817 ͑1978b͒.
Doi, M., and S. F. Edwards, “Dynamics of concentrated polymer systems. Part 3. The constitutive equation,” J.
Chem. Soc., Faraday Trans. 2 74, 1818–1832 ͑1978c͒.
Doi, M., and S. F. Edwards, “Dynamics of concentrated polymer systems. Part 4. Rheological properties,” J.
Chem. Soc., Faraday Trans. 2 75, 38–54 ͑1979͒.
Doi, M., and S. F. Edwards, The Theory of Polymer Dynamics ͑Clarendon, Oxford, 1986͒.
Doi, M., and J.-I. Takimoto, “Molecular modelling of entanglement,” Philos. Trans. R. Soc. London, Ser. A
361, 641–652 ͑2003͒.
Doshi, S. R., and J. M. Dealy, “Exponential shear: A strong flow,” J. Rheol. 31, 563–582 ͑1987͒.
Eskandari, M., “Optimization of the solid-state shear extrusion process for low-cross-link density natural rubber
with a proposed Gaussian slip-link model,” Ph.D. thesis, Illinois Institute of Technology, Chicago, 2006.
Everaers, R., S. K. Sukumaran, G. S. Grest, C. Svaneborg, A. Sivasubramanian, and K. Kremer, “Rheology and
microscopic topology of entangled polymeric liquids,” Science 303, 823–826 ͑2004͒.
Fang, J., M. Kröger, and H. C. Öttinger, “A thermodynamically admissible reptation model for fast flows of
entangled polymers. II. Model predictions for shear and extensional flows,” J. Rheol. 44, 1293–1317
͑2000͒.
Ferry, J. L., Viscoelastic Properties of Polymers, 3rd ed. ͑Wiley, New York, 1980͒.
Fetters, L. J., D. J. Lohse, S. Milner, and W. W. Graessley, “Packing length influence in linear polymer melts
on the entanglement, critical and reptation molecular weights,” Macromolecules 32, 6847–6851 ͑1999a͒.
Fetters, L. J., D. J. Lohse, and W. W. Graessley, “Chain dimensions and entanglement spacings in dense
macromolecular systems,” J. Polym. Sci., Part B: Polym. Phys. 37, 1023–1033 ͑1999b͒.
Foteinopoulou, K., N. Karayiannis, V. Mavrantzas, and M. Kroger, “Primitive path identification and entangle-
ment statistics in polymer melts: Results from direct topological analysis on atomistic polyethylene mod-
els,” Macromolecules 39, 4207–4216 ͑2006͒.
Gardiner, C. W., Handbook of Stochastic Methods for Physics, Chemistry and the Natural Sciences, 2nd ed.
͑Springer, Berlin, 1982͒.
Gigras, P. G., and B. Khomami, “Adaptive configuration fields: A new multiscale simulation technique for
reptation-based models with a stochastic strain measure and local variations of life span distribution,” J.
Non-Newtonian Fluid Mech. 108, 99–122 ͑2002͒.
1139CUBS MODEL FLOW DATA COMPARISON
Graham, R. S., A. Likhtman, T. C. B. McLeish, and S. T. Milner, “Microscopic theory of linear, entangled
polymer chains under rapid deformation including chain stretch and convective constraint release,” J. Rheol.
47, 1171–1200 ͑2003͒.
Graham, R. S., T. C. B. McLeish, and O. G. Harlen, “Using the pom-pom equations to analyze polymer melts
in exponential shear,” J. Rheol. 45, 275–290 ͑2001͒.
Greco, F., “The stress tensor in entangled polymers,” Phys. Rev. Lett. 88, 108301 ͑2002͒.
Honerkamp, J., Stochastische Dynamische Systeme ͑VCH, Berlin, 1983͒.
Hua, C. C., and J. D. Schieber, “Segment connectivity, chainlength breathing, segmental stretch, and constraint
release in reptation models. I. Theory and single-step strain predictions,” J. Chem. Phys. 109, 10018–10027
͑1998a͒.
Hua, C.-C., and J. D. Schieber, “Viscoelastic flow through fibrous media using the CONNFFESSIT approach,”
J. Rheol. 42, 477–491 ͑1998b͒.
Hua, C. C., J. D. Schieber, and D. C. Venerus, “Segment connectivity, chainlength breathing, segmental stretch,
and constraint release in reptation models. II. Double-step strain predictions,” J. Chem. Phys. 109, 10028–
10032 ͑1998͒.
Hua, C. C., J. D. Schieber, and D. C. Venerus, “Segment connectivity, chain length breathing, segmental stretch,
and constraint release in reptation models. III. Shear flows,” J. Rheol. 43, 701–717 ͑1999͒.
Ianniruberto, G., and G. Marrucci, “On compatibility of the Cox–Merz rule with the model of Doi and Ed-
wards,” J. Non-Newtonian Fluid Mech. 65, 241–246 ͑1996͒.
Ianniruberto, G., and G. Marrucci, “Convective orientational renewal in entangled polymers,” J. Non-
Newtonian Fluid Mech. 95, 363–374 ͑2000͒.
Ianniruberto, G., and G. Marrucci, “A simple constitutive equation for entangled polymers with chain stretch,”
J. Rheol. 45, 1305–1318 ͑2001͒.
Kahvand, H., “Strain coupling effects in polymer rheology,” Ph.D. thesis, Illinois Institute of Technology,
Chicago, 1995.
Kaye, A., “Non-Newtonian flow of incompressible fluids,” Note No. 134, College of Aeronautics, Cranfield,
UK, 1962.
Kitkrailard, T., “Dynamics of semi-flexible entangled polymers,” Master’s thesis, Illinois Institute of Technol-
ogy, Chicago, Illinois, 2005.
Kwan, T. C. B., and E. Shaqfeh, “Brownian dynamics simulations of the stress and molecular configurations of
polymers in exponential and linearly–ramped shear flow,” J. Non-Newtonian Fluid Mech. 82, 139–165
͑1999͒.
Leygue, A., C. Bailly, and R. Keunings, “A tube-based constitutive equation for polydisperse entangled linear
polymers,” J. Non-Newtonian Fluid Mech. 136, 1–16 ͑2006͒.
Likhtman, A., “Single-chain slip-link model of entangled polymers: Simultaneous description of neutron spin-
echo, rheology, and diffusion,” Macromolecules 38, 6128–6139 ͑2005͒.
Marko, J. F., and E. D. Siggia, “Stretching DNA,” Macromolecules 28, 8759–8770 ͑1995͒.
Marrucci, G., “Dynamics of entanglements: A nonlinear model consistent with the Cox–Merz rule,” J. Non-
Newtonian Fluid Mech. 62, 279–289 ͑1996͒.
Marrucci, G., and G. Ianniruberto, “Flow-induced orientation and stretching of entangled polymers,” Philos.
Trans. R. Soc. London, Ser. A 361, 677–688 ͑2003͒.
Masubuchi, Y., G. Ianniruberto, F. Greco, and G. Marrucci, “Entanglement molecular weight and frequency
response of sliplink networks,” J. Chem. Phys. 119, 6925–6930 ͑2003͒.
Masubuchi, Y., J. Ichi Takimoto, K. Koyama, G. Ianniruberto, G. Marrucci, and F. Greco, “Brownian simula-
tions of a network of reptating primitive chains,” J. Chem. Phys. 115, 4387–4394 ͑2001͒.
Mead, D. W., R. G. Larson, and M. Doi, “A molecular theory for fast flows of entangled polymers,”
Macromolecules 31, 7895–7914 ͑1998͒.
Menezes, E. V., and W. W. Graessley, “Nonlinear rheological behavior of polymer systems for several shear-
flow histories,” J. Polym. Sci., Polym. Phys. Ed. 20, 1817–1833 ͑1980͒.
Milner, S. T., T. C. B. McLeish, and A. E. Likhtman, “Microscopic theory of convective constraint release,” J.
Rheol. 45, 539–563 ͑2001͒.
Nair, D., “Brownian dynamics simulation of consistently unconstrainted Brownian sliplinks,” Master’s thesis,
1140 SCHIEBER, NAIR, AND KITKRAILARD
Illinois Institute of Technology, Chicago, 2004.
Nair, D. M., and J. D. Schieber, “Linear viscoelastic predictions of a consistently unconstrained Brownian
slip-link model,” Macromolecules 39, 3386–3397 ͑2006͒.
Neergaard, J., K. Park, D. C. Venerus, and J. D. Schieber, “Exponential shear flow of linear, entangled poly-
meric liquids,” J. Rheol. 44, 1043–1054 ͑2000͒.
Neergaard, J., and J. D. Schieber, “A full chain network model with sliplinks and binary constraint release,”
XIIIth International Congress on Rheology, Cambridge, United Kingdom, 2000, Vol. 2, pp. 120–122.
Osaki, K., S. Kimura, and M. Kurata, “Relaxation of shear and normal stresses in double-step shear deforma-
tions for a polystyrene solution. A test of the Doi–Edwards theory for polymer rheology,” J. Rheol. 25,
549–562 ͑1981͒.
Osaki, K., and M. Kurata, “Stress relaxation of a polymer solution in double-step shear deformations: A
separate evaluation of the stress components due to entangled segments of different orientations,” J. Polym.
Sci., Part A-2 20, 623–632 ͑1982͒.
Öttinger, H. C., “Modified reptation model,” Phys. Rev. E 50, 4891–4895 ͑1994͒.
Öttinger, H. C., Stochastic Processes in Polymeric Fluids ͑Springer, Berlin, 1996͒.
Öttinger, H. C., “A thermodynamically admissible reptation model for fast flows of entangled polymers,” J.
Rheol. 43, 1461–1493 ͑1999͒.
Rault, J., “Space filling an entanglements in polymeric systems,” J. Non-Newtonian Fluid Mech. 23, 229–247
͑1987͒.
Rubinstein, M., “Discretized model of entangled-polymer dynamics,” Phys. Rev. Lett. 59, 1946–1949 ͑1987͒.
Rubinstein, M., and S. Panyukov, “Elasticity of polymer networks,” Macromolecules 35, 6670–6686 ͑2002͒.
Samurkas, T., R. G. Larson, and J. M. Dealy, “Strong extensional and shearing flows of a branched polyeth-
ylene,” J. Rheol. 33, 559–578 ͑1989͒.
Schieber, J. D., “Fluctuations in entanglements of polymer liquids,” J. Chem. Phys. 118, 5162–5166 ͑2003a͒.
Schieber, J. D., “GENERIC compliance of a temporary network model with sliplinks, chain-length fluctuations,
segment-connectivity and constraint release,” J. Non-Equil. Thermodyn. 28, 179–188 ͑2003b͒.
Schieber, J. D., J. Neergaard, and S. Gupta, “A full-chain, temporary network model with sliplinks, chain-length
fluctuations, chain connectivity and chain stretching,” J. Rheol. 47, 213–233 ͑2003͒. Although all dynami-
cal calculations were verified to agree with analytic results at equilibrium, we later discovered that stress
calculations contained a bug. Hence, stress curves in this manuscript should be ignored.
Schweizer, T., J. van Meerveld, and H. C. Öttinger, “Nonlinear shear rheology of polystyrene melt with narrow
molecular weight distribution experiment and theory,” J. Rheol. 48, 1345–1363 ͑2004͒.
Shanbhag, S., and M. Kroger, “Primitive path networks generated by annealing and geometrical methods:
Insights into differences,” Macromolecules 40, 2897–2903 ͑2007͒.
Sivashinsky, N., A. T. Tsai, T. J. Moon, and D. S. Soong, “Some new transient test results from a parallel-plate
rheometer,” J. Rheol. 28, 287–301 ͑1984͒.
Tzoumanekas, C., and D. N. Theodorou, “Topological analysis of linear polymer melts: A statistical approach,”
Macromolecules 39, 4592–4604 ͑2006͒.
Venerus, D. C., “Exponential shear flow of polymer melts,” Rheol. Acta 39, 71–79 ͑2000͒.
Venerus, D. C., and H. Kahvand, “Doi–Edwards theory evaluation in double-step strain flows,” J. Polym. Sci.,
Part B: Polym. Phys. 32, 1531–1542 ͑1994a͒.
Venerus, D. C., and H. Kahvand, “Normal stress relaxation in reversing double-step strain flows,” J. Rheol. 38,
1297–1315 ͑1994b͒.
Venerus, D. C., C. M. Vrentas, and J. S. Vrentas, “Step strain deformations for viscoelastic fluids: Experiment,”
J. Rheol. 34, 657–684 ͑1990͒.
Zhou, Q., and R. Larson, “Primitive path identification and statistics in molecular dynamics simulations of
entangled polymer melts,” Macromolecules 38, 5761–5765 ͑2005͒.
Zülle, B., J. J. Linster, J. Meissner, and H. P. Hürlimann, “Deformation hardening and thinning in both
elongation and shear of a low density polyetyhylene melt,” J. Rheol. 31, 583–598 ͑1987͒.
1141CUBS MODEL FLOW DATA COMPARISON

More Related Content

Viewers also liked

Political Marketing
Political MarketingPolitical Marketing
Political Marketing
Samantha Denbow
 
Strategic planing for competitive advantage
Strategic planing for competitive advantageStrategic planing for competitive advantage
Strategic planing for competitive advantage
Sreang Buntith
 
Political marketing presentation
Political marketing presentationPolitical marketing presentation
Political marketing presentation
Evonne Mwangale Kiptinness
 
INDIAN POLITICAL MARKETING
INDIAN POLITICAL MARKETINGINDIAN POLITICAL MARKETING
INDIAN POLITICAL MARKETING
VedangiBrahmbhatt
 
Political marketing 2014
Political marketing 2014Political marketing 2014
Political marketing 2014
Aashiq Mohamed
 
E-Business & E-Commerce Basics
E-Business & E-Commerce BasicsE-Business & E-Commerce Basics
E-Business & E-Commerce Basics
Abhishek Duttagupta
 
Ebusiness & Information Systems
Ebusiness & Information SystemsEbusiness & Information Systems
Ebusiness & Information Systems
Jose G. Lepervanche, PhD
 
E Business+Models Ppt
E Business+Models PptE Business+Models Ppt
E Business+Models Ppt
md kaiser
 
Political Marketing Presentation
Political Marketing PresentationPolitical Marketing Presentation
Political Marketing Presentation
Sagar Patekar
 
E Business Introduction
E Business IntroductionE Business Introduction
E Business Introduction
Mrirfan
 
E-Supply Chain Management
E-Supply Chain ManagementE-Supply Chain Management
E-Supply Chain Management
Mohamed Aamer
 
E commerce and supply chain management
E commerce and supply chain managementE commerce and supply chain management
E commerce and supply chain management
Saugata Palit
 
Intro to e-commerce and e-business
Intro to e-commerce and e-businessIntro to e-commerce and e-business
Intro to e-commerce and e-business
MoodleAdmin
 
E commerce supply chain management
E commerce supply chain managementE commerce supply chain management
E commerce supply chain management
Zubin Poonawalla
 
E business ,e-commerce, e-marketing (sadiq shariff10@hotmail.com)
E business ,e-commerce, e-marketing (sadiq shariff10@hotmail.com)E business ,e-commerce, e-marketing (sadiq shariff10@hotmail.com)
E business ,e-commerce, e-marketing (sadiq shariff10@hotmail.com)
Sadiq Shariff
 
E business models
E business modelsE business models
E business models
Madhur Choudhary
 
E commerce business models
E commerce business modelsE commerce business models
E commerce business models
Vikram g b
 
E - BUSINESS / E - COMMERCE
E - BUSINESS / E - COMMERCEE - BUSINESS / E - COMMERCE
E - BUSINESS / E - COMMERCE
Paviter Sandhir
 

Viewers also liked (18)

Political Marketing
Political MarketingPolitical Marketing
Political Marketing
 
Strategic planing for competitive advantage
Strategic planing for competitive advantageStrategic planing for competitive advantage
Strategic planing for competitive advantage
 
Political marketing presentation
Political marketing presentationPolitical marketing presentation
Political marketing presentation
 
INDIAN POLITICAL MARKETING
INDIAN POLITICAL MARKETINGINDIAN POLITICAL MARKETING
INDIAN POLITICAL MARKETING
 
Political marketing 2014
Political marketing 2014Political marketing 2014
Political marketing 2014
 
E-Business & E-Commerce Basics
E-Business & E-Commerce BasicsE-Business & E-Commerce Basics
E-Business & E-Commerce Basics
 
Ebusiness & Information Systems
Ebusiness & Information SystemsEbusiness & Information Systems
Ebusiness & Information Systems
 
E Business+Models Ppt
E Business+Models PptE Business+Models Ppt
E Business+Models Ppt
 
Political Marketing Presentation
Political Marketing PresentationPolitical Marketing Presentation
Political Marketing Presentation
 
E Business Introduction
E Business IntroductionE Business Introduction
E Business Introduction
 
E-Supply Chain Management
E-Supply Chain ManagementE-Supply Chain Management
E-Supply Chain Management
 
E commerce and supply chain management
E commerce and supply chain managementE commerce and supply chain management
E commerce and supply chain management
 
Intro to e-commerce and e-business
Intro to e-commerce and e-businessIntro to e-commerce and e-business
Intro to e-commerce and e-business
 
E commerce supply chain management
E commerce supply chain managementE commerce supply chain management
E commerce supply chain management
 
E business ,e-commerce, e-marketing (sadiq shariff10@hotmail.com)
E business ,e-commerce, e-marketing (sadiq shariff10@hotmail.com)E business ,e-commerce, e-marketing (sadiq shariff10@hotmail.com)
E business ,e-commerce, e-marketing (sadiq shariff10@hotmail.com)
 
E business models
E business modelsE business models
E business models
 
E commerce business models
E commerce business modelsE commerce business models
E commerce business models
 
E - BUSINESS / E - COMMERCE
E - BUSINESS / E - COMMERCEE - BUSINESS / E - COMMERCE
E - BUSINESS / E - COMMERCE
 

Similar to Paper_Journal of Rheology

Paper_Macromolecules
Paper_MacromoleculesPaper_Macromolecules
Paper_Macromolecules
Deepa Nair
 
MagnientSagautDeville_PoF_2001
MagnientSagautDeville_PoF_2001MagnientSagautDeville_PoF_2001
MagnientSagautDeville_PoF_2001
Jean-Christophe Magnient
 
Análisis no lineal de estructuras de cables
Análisis no lineal de estructuras de cablesAnálisis no lineal de estructuras de cables
Análisis no lineal de estructuras de cables
Julio Terrones
 
Lattice Polymer Report 20662
Lattice Polymer Report 20662Lattice Polymer Report 20662
Lattice Polymer Report 20662
Tim Jones
 
MBS
MBSMBS
MBS
MB SHAH
 
Continuum Modeling Techniques to Determine Mechanical Properties of Nanocomp...
Continuum Modeling Techniques to Determine Mechanical  Properties of Nanocomp...Continuum Modeling Techniques to Determine Mechanical  Properties of Nanocomp...
Continuum Modeling Techniques to Determine Mechanical Properties of Nanocomp...
IJMER
 
2014 Reinhardt
2014 Reinhardt2014 Reinhardt
2014 Reinhardt
james reinhardt
 
A New Method Based On the Comparison of the Unique Chain Code to Detect Isom...
A New Method Based On the Comparison of the Unique Chain  Code to Detect Isom...A New Method Based On the Comparison of the Unique Chain  Code to Detect Isom...
A New Method Based On the Comparison of the Unique Chain Code to Detect Isom...
IJMER
 
Effect of Geometry on Variation of Heat Flux and Drag for Launch Vehicle -- Z...
Effect of Geometry on Variation of Heat Flux and Drag for Launch Vehicle -- Z...Effect of Geometry on Variation of Heat Flux and Drag for Launch Vehicle -- Z...
Effect of Geometry on Variation of Heat Flux and Drag for Launch Vehicle -- Z...
Abhishek Jain
 
Identification of isomorphism and detection of distinct mechanism of kinemati...
Identification of isomorphism and detection of distinct mechanism of kinemati...Identification of isomorphism and detection of distinct mechanism of kinemati...
Identification of isomorphism and detection of distinct mechanism of kinemati...
eSAT Publishing House
 
Identification of isomorphism and detection of distinct mechanism of kinemati...
Identification of isomorphism and detection of distinct mechanism of kinemati...Identification of isomorphism and detection of distinct mechanism of kinemati...
Identification of isomorphism and detection of distinct mechanism of kinemati...
eSAT Journals
 
Auricolari, svelato il mistero dei nodi nei cavi
Auricolari, svelato il mistero dei nodi nei caviAuricolari, svelato il mistero dei nodi nei cavi
Auricolari, svelato il mistero dei nodi nei cavi
Dario Caliendo
 
6. Rezaiguia et al.pdf
6. Rezaiguia et al.pdf6. Rezaiguia et al.pdf
6. Rezaiguia et al.pdf
OmNarayanSingh4
 
LaSiGMA_Poster-Ethan , Divya
LaSiGMA_Poster-Ethan , DivyaLaSiGMA_Poster-Ethan , Divya
LaSiGMA_Poster-Ethan , Divya
Ethan Sullivan
 
VasasElekMarkus_RSAR_JSPI_final
VasasElekMarkus_RSAR_JSPI_finalVasasElekMarkus_RSAR_JSPI_final
VasasElekMarkus_RSAR_JSPI_final
Vasas Krisztina
 
PH600 Project report
PH600 Project reportPH600 Project report
PH600 Project report
Luke Moore
 
Mechanics of chain-link wire nets
Mechanics of chain-link wire netsMechanics of chain-link wire nets
Mechanics of chain-link wire nets
Juan Escallón
 
Asme2009 82287 - Porous Media - Forced Convection Flow
Asme2009 82287 - Porous Media - Forced Convection FlowAsme2009 82287 - Porous Media - Forced Convection Flow
Asme2009 82287 - Porous Media - Forced Convection Flow
HIIO
 
Experimental Analysis of Superlooping in Twisted Polymer Line and its Applica...
Experimental Analysis of Superlooping in Twisted Polymer Line and its Applica...Experimental Analysis of Superlooping in Twisted Polymer Line and its Applica...
Experimental Analysis of Superlooping in Twisted Polymer Line and its Applica...
Allison Stiller
 
Experimental Analysis of Superlooping in Twisted Polymer Line and its Applica...
Experimental Analysis of Superlooping in Twisted Polymer Line and its Applica...Experimental Analysis of Superlooping in Twisted Polymer Line and its Applica...
Experimental Analysis of Superlooping in Twisted Polymer Line and its Applica...
Marlen Mahendraratnam
 

Similar to Paper_Journal of Rheology (20)

Paper_Macromolecules
Paper_MacromoleculesPaper_Macromolecules
Paper_Macromolecules
 
MagnientSagautDeville_PoF_2001
MagnientSagautDeville_PoF_2001MagnientSagautDeville_PoF_2001
MagnientSagautDeville_PoF_2001
 
Análisis no lineal de estructuras de cables
Análisis no lineal de estructuras de cablesAnálisis no lineal de estructuras de cables
Análisis no lineal de estructuras de cables
 
Lattice Polymer Report 20662
Lattice Polymer Report 20662Lattice Polymer Report 20662
Lattice Polymer Report 20662
 
MBS
MBSMBS
MBS
 
Continuum Modeling Techniques to Determine Mechanical Properties of Nanocomp...
Continuum Modeling Techniques to Determine Mechanical  Properties of Nanocomp...Continuum Modeling Techniques to Determine Mechanical  Properties of Nanocomp...
Continuum Modeling Techniques to Determine Mechanical Properties of Nanocomp...
 
2014 Reinhardt
2014 Reinhardt2014 Reinhardt
2014 Reinhardt
 
A New Method Based On the Comparison of the Unique Chain Code to Detect Isom...
A New Method Based On the Comparison of the Unique Chain  Code to Detect Isom...A New Method Based On the Comparison of the Unique Chain  Code to Detect Isom...
A New Method Based On the Comparison of the Unique Chain Code to Detect Isom...
 
Effect of Geometry on Variation of Heat Flux and Drag for Launch Vehicle -- Z...
Effect of Geometry on Variation of Heat Flux and Drag for Launch Vehicle -- Z...Effect of Geometry on Variation of Heat Flux and Drag for Launch Vehicle -- Z...
Effect of Geometry on Variation of Heat Flux and Drag for Launch Vehicle -- Z...
 
Identification of isomorphism and detection of distinct mechanism of kinemati...
Identification of isomorphism and detection of distinct mechanism of kinemati...Identification of isomorphism and detection of distinct mechanism of kinemati...
Identification of isomorphism and detection of distinct mechanism of kinemati...
 
Identification of isomorphism and detection of distinct mechanism of kinemati...
Identification of isomorphism and detection of distinct mechanism of kinemati...Identification of isomorphism and detection of distinct mechanism of kinemati...
Identification of isomorphism and detection of distinct mechanism of kinemati...
 
Auricolari, svelato il mistero dei nodi nei cavi
Auricolari, svelato il mistero dei nodi nei caviAuricolari, svelato il mistero dei nodi nei cavi
Auricolari, svelato il mistero dei nodi nei cavi
 
6. Rezaiguia et al.pdf
6. Rezaiguia et al.pdf6. Rezaiguia et al.pdf
6. Rezaiguia et al.pdf
 
LaSiGMA_Poster-Ethan , Divya
LaSiGMA_Poster-Ethan , DivyaLaSiGMA_Poster-Ethan , Divya
LaSiGMA_Poster-Ethan , Divya
 
VasasElekMarkus_RSAR_JSPI_final
VasasElekMarkus_RSAR_JSPI_finalVasasElekMarkus_RSAR_JSPI_final
VasasElekMarkus_RSAR_JSPI_final
 
PH600 Project report
PH600 Project reportPH600 Project report
PH600 Project report
 
Mechanics of chain-link wire nets
Mechanics of chain-link wire netsMechanics of chain-link wire nets
Mechanics of chain-link wire nets
 
Asme2009 82287 - Porous Media - Forced Convection Flow
Asme2009 82287 - Porous Media - Forced Convection FlowAsme2009 82287 - Porous Media - Forced Convection Flow
Asme2009 82287 - Porous Media - Forced Convection Flow
 
Experimental Analysis of Superlooping in Twisted Polymer Line and its Applica...
Experimental Analysis of Superlooping in Twisted Polymer Line and its Applica...Experimental Analysis of Superlooping in Twisted Polymer Line and its Applica...
Experimental Analysis of Superlooping in Twisted Polymer Line and its Applica...
 
Experimental Analysis of Superlooping in Twisted Polymer Line and its Applica...
Experimental Analysis of Superlooping in Twisted Polymer Line and its Applica...Experimental Analysis of Superlooping in Twisted Polymer Line and its Applica...
Experimental Analysis of Superlooping in Twisted Polymer Line and its Applica...
 

Paper_Journal of Rheology

  • 1. Comprehensive comparisons with nonlinear flow data of a consistently unconstrained Brownian slip-link model Jay D. Schieber,a) Deepa M. Nair, and Thidaporn Kitkrailard Center of Excellence in Polymer Science and Engineering, Department of Chemical and Biological Engineering, Illinois Institute of Technology, Chicago, Illinois 60616-3793 (Received 19 April 2006; final revision received 6 September 2007͒ Synopsis A consistently unconstrained Brownian slip-link model ͑CUBS͒ with constant chain friction is used to predict the nonlinear rheological behavior of linear, entangled, polymeric liquids. The model naturally incorporates primitive-path-length fluctuations, segment connectivity, monomer density fluctuations, entanglement fluctuations, and constraint release without making any closure approximations. Constraint release is imposed on the level of the dynamics of the chain, and the relaxation modulus follows from these rigorously. The model is a mean-field, single-chain slip-link model, or temporary network model, with a single phenomenological time constant, ␶e, fit by linear viscoelasticity. The nonlinear flow predictions are made without adjusting any additional parameters. We find that the addition of constant chain friction noticeably improves the model predictions in all the flows considered. In contradiction with tube models, the results suggest that the additional physics of constraint release and convective constraint release are not very important in predicting the nonlinear shear properties, except at low shear rates ͑close to the LVE regime͒. © 2007 The Society of Rheology. ͓DOI: 10.1122/1.2790460͔ I. INTRODUCTION Quantitative agreement with experimental data has become increasingly accurate for tube models. However, most such models are restricted to either linear viscoelasticity or to flow, and to a single kind of chain architecture. If chain branching is considered, the mathematical model is modified substantially from the linear case, or is unknown alto- gether. These tube models exist on various levels of description. Doi and Edwards tensor model ͓Doi and Edwards ͑1978a, 1978b, 1978c͔͒ can be written as six three-dimensional integrals: one dimension for time history and two over segment orientation; tube models with stretch have seven coupled ordinary differential equations, including one dimension for average stretch ͓Ianniruberto and Marrucci ͑2001͔͒; the preaveraged tube model of ͓Graham et al. ͑2003͔͒ uses six coupled partial differential equations. The last model is sometimes considered to be the most accurate tube model available. Ideally, one seeks a model on the least-detailed level possible that makes accurate a͒ Electronic mail: schieber@iit.edu © 2007 by The Society of Rheology, Inc. 1111J. Rheol. 51͑6͒, 1111-1141 November/December ͑2007͒ 0148-6055/2007/51͑6͒/1111/31/$27.00
  • 2. predictions for the problem at hand. For example, continuum mechanics equations— coupled ordinary differential equations—have been a primary goal, historically. If more complicated equations arise, calculation of nonhomogeneous flows becomes more diffi- cult. For example, implementation of the coupled partial differential equations of Graham et al. in finite element codes is presumably computationally prohibitive. Even so, the use of uncontrolled closure approximations in these models can lead to unexpected errors. For example, the latter model predicts zero second-normal stress difference in shear flow. In order to avoid uncontrolled closures, it is necessary to account for fluctuations more rigorously. Toward this end stochastic models are ideally suited. This is the approach taken, for example, by Öttinger using a single-segment model ͓Fang et al. ͑2000͒; Öt- tinger ͑1999͔͒, by Hua and Schieber for a full chain model ͓Hua and Schieber ͑1998a͒; Hua et al. ͑1998, 1999͒; Neergaard et al. ͑2000͔͒, and by Masubuchi et al. using inter- acting chains inside a simulation box ͓Masubuchi et al. ͑2003͔͒. Not surprisingly, these models show agreement with data over a broader range of experiments—for example, a reasonable second normal-stress difference. However, only the first of these models has been used in complex geometries ͓Gigras and Khomami ͑2002͔͒. Nonetheless, all tube models to date exhibit a maximum in the steady shear stress versus shear rate curve, which is not observed experimentally. Only through tuning pa- rameters for excessive convective constraint release, or too-early onset of chain stretching is the maximum avoided in the single-segment tube models. When no such adjustable parameters are available, the maximum is seen for sufficiently entangled tube models. More recent work has seen the advance of slip-link models. The first slip-link picture was drawn by Doi and Edwards. However, it was used only to motivate a certain network-model expression for the stress tensor in their tube model. The repton model of Rubinstein ͑1987͒ more closely resembles the slip-link picture than the tube picture, but here the constraining mesh of chains surrounding the test chain was considered to be regular. Flow was not considered. Neergaard and Schieber introduced the idea of using the monomer density between slip-links as a stochastic variable in a stochastic, full chain model ͓Neergaard and Schieber ͑2000͒; Schieber ͑2003a͒; Schieber et al. ͑2003͔͒, and a similar multichain simulation was introduced independently by ͓Masubuchi et al. ͑2001͔͒. Greco considered a similar stochastic, single-segment model, in which an infi- nitely long chain possessed two entanglements, and the ͑infinitely long͒ ends were in a chemical potential bath ͓Greco ͑2002͔͒. Only step strains have been considered to date for this latter model. Likhtman considered a Rouse chain sliding through slip links that are themselves attached by elastic springs to an affinely deforming background ͓Likhtman ͑2005͔͒. The picture is a more-detailed depiction of a similar model developed by Panyukov and Rubinstein for lightly cross-linked elastomers ͓Rubinstein and Panyukov ͑2002͔͒. Likht- man’s model has been used to study dynamic equilibrium only, to date. In a previous article, we generalized the slip-link model of Neergaard and Schieber to include constraint release—the idea that the matrix of constraining chains are relaxing on the same spectrum of time scales as the probe chain ͓Nair and Schieber ͑2006͔͒. Com- prehensive comparison was made there with linear viscoelastic experiments and reason- able agreement was found with data. Here we make comparisons with flow. We will show that the slip-link model is able to describe accurately all observations ͑known to us͒ in shear flow, including a monotonic steady shear stress growth with shear rate for linear chains. An added benefit to slip-link models is a more natural connection to atomistic level simulations. Recently, several groups have created and tested algorithms to predict en- tanglement statistics from equilibrated, long-chain, melt polymers on such a level ͓Ever- 1112 SCHIEBER, NAIR, AND KITKRAILARD
  • 3. aers et al. ͑2004͔͒. More than one algorithm has been able to predict entanglement density; however, recent attempts have included calculation of more-detailed statistics, such as primitive-path length distribution, or entanglement number distribution ͓Foteinopoulou et al. ͑2006͒; Shanbhag and Kroger ͑2007͒; Tzoumanekas and Theodorou ͑2006͒; Zhou and Larson ͑2005͔͒. These comparisons have shown excellent agreement with our slip-link theories, offering a path for true ab initio calculation of stresses and dynamics in entangled polymers. A third advantage to slip-link models is that branched systems or even cross-linked systems ͓Eskandari ͑2006͔͒ can be modeled without significant modification to the math- ematics. II. THE CUBS MODEL In the consistently unconstrained Brownian slip-link ͑CUBS͒ model ͑Fig. 1͒, entangle- ments are represented as small rings called “slip links” along the contour length of the chain. As in the reptation theory, the chain is free to move along its contour length but the transverse motion is restricted due to the presence of slip links or entanglements. The end segments which are not constrained by any slip links are free to explore all conformations in space. The chain is allowed to slide through these entanglements and, hence, the number of Kuhn steps in an entangled strand fluctuates stochastically. Without constraint release, the entanglements are assumed to move affinely. The entanglements are created and destroyed only at the ends of the chain when an end strand completely slips out of a slip link and the dangling end has zero Kuhn steps. In reality, a new entanglement is created when a surrounding chain entangles the test chain at the end. We model this process by an entanglement creation probability that satisfies detailed balance with entanglement destruction. In the previously proposed model, entanglements are assumed to be fixed in the medium—i.e., reptation of the surrounding chains was not considered ͑no constraint release͒. In this model, we consider the diffusion of the surrounding chains, thereby mimicking the creation and destruction of entanglements in the middle of the chain ͑constraint release͒. In the previous model it FIG. 1. Sketch of a short chain of 200 Kuhn steps trapped in three entanglements. Also shown is the notation used for the vectors in the model. Ri is the position in space of entanglement i, drawn by solid ͑black͒ arrows, and Qi is the dashed ͑green͒ arrow connecting entanglement i−1 to entanglement i, trapping Ni Kuhn steps. 1113CUBS MODEL FLOW DATA COMPARISON
  • 4. was also assumed that friction is proportional to the number of entanglements and, hence, total friction in the chain fluctuated with the creation and destruction of entanglements. In the present model, friction is assumed to be proportional to the number of monomers or Kuhn steps in the chain. Hence, total friction in the chain is a constant quantity since the total number of monomers or Kuhn steps in the chain is a constant. A. Mathematical description The evolution equation for the probability density has four contributions, from, re- spectively, affine motion of the entanglements, sliding of the chain through the entangle- ments, constraint release, and creation/destruction of the entanglements at the chain ends ‫ץ‬p ‫ץ‬t = p˙affine + p˙steps + p˙CR + p˙ends. ͑1͒ We give each contribution in turn. The dynamics of the Kuhn steps in the chain are driven by both Brownian forces and free energy minimization. Free energy minimization requires the sliding of Kuhn steps from the strands of higher chemical potential ͑‫ץ‬F/‫ץ‬Ni͒T,͕Qj͖,͕Nj i͖ to those of lower chemi- cal potential. A friction or mobility for this Kuhn-step-shuffling process is also required, which introduces the single phenomenological parameter of the theory p˙steps = ͚i,j=1 Z ‫ץ‬ ‫ץ‬Ni 1 kBT͕͑Nk͖͒ Aˆ ijͫͩ‫ץ‬F ‫ץ‬Nj ͪT,͕Rj͖,͕Nk j͖ p + kBT ‫ץ‬p ‫ץ‬Nj ͬ, ͑2͒ where Aˆ ij ª Ά − 1 ␶i−1 , i = j + 1 1 ␶i + 1 ␶i−1 , i = j − 1 ␶i , i = j − 1 0, otherwise. ͑3͒ We have considered two possible implementations for this friction: constant- entanglement friction and constant-chain friction. For constant-entanglement friction, we set ␶i=␶Kϳconstant, in which case Aˆ ij is the Rouse matrix divided by ␶K. Nondimen- sionalizing the dynamical equations also leads to a more natural fundamental time scale ␶e=Ne 2 ␶K. Since the number of entanglements on a chain fluctuate, the total friction experienced by the chain will also fluctuate in this case. For various mathematical reasons, tube models typically assume that the friction of a chain is constant. Hence, we also consider the possibility that ␶i=␶K͑Ni+Ni−1͒/͑2Ne͒, which we call constant-chain friction. Com- parison with experiment below suggests that constant-chain friction is a better assump- tion. In the absence of constraint release, the entanglements are assumed to move affinely during flow. Hence, we write 1114 SCHIEBER, NAIR, AND KITKRAILARD
  • 5. p˙affine = − ͚i=1 Z−1 ‫ץ‬ ‫ץ‬Ri · ͓␬ · Ri͔p, ͑4͒ where ␬=ٌ͑v͒† is the transpose of the velocity gradient. Note that, due to the shuffling of Kuhn steps through the slip links, the chains themselves do not deform affinely. The effect of constraint release is incorporated in a self-consistent mean-field way. Constraint release is mimicked by a diffusive process of the entanglements, which makes the entanglement motion no longer affine. Hence, we write p˙CR = ͚i,j=1 Z−1 ‫ץ‬ ‫ץ‬Ri · NeaK 2 12kBT␶j CRAijͫpͩ‫ץ‬F ‫ץ‬Rj ͪT,͕Nj͖,͕Ri j͖ + kBT ‫ץ‬p ‫ץ‬Rj ͬ, ͑5͒ where ␶j CR is the constraint release time scale and Aij is the Rouse matrix ͓Bird et al. ͑1987͔͒. Here we use a dimensionless prefactor, which determines the strength of the constraint release mechanism. We calculate the value of this prefactor to be 1/12. Details of this calculation may be found in the supplementary materials of ͓Nair and Schieber ͑2006͔͒. Thus, by considering the diffusion of the surrounding chains, we are mimicking the birth and destruction of slip links in the middle of the chain. The time constants ␶i CR are obtained in a self-consistent way by keeping track of the fraction of surviving en- tanglements during an equilibrium simulation. Note that this assumption of Rouse-like motion for the slip links is not necessary for this model, since creation and destruction of entanglements in the middle of the chain could be implemented rigorously. However, such an assumption is typically made in tube models, which we will follow here. More rigorous formulation is left for future work. During flow, the rate of entanglement destruction can increase—so-called “convective constraint release” ͓Ianniruberto and Marrucci ͑2000͔͒. Hence, during flow we replace ␶i CR with ␶i CCR , where 1 ␶i CCR = 1 ␶i CR + ͗wd͘ − ͗wd͘eq, i = 1, ... ,Z − 1, ͑6͒ where ͗wd͘ is the average rate of destruction of entanglements at the dangling ends during flow and ͗wd͘eq is that at equilibrium. Ideally, the destruction of entanglements would be given by boundary conditions— i.e., the entanglement dies when the chain reaches the entanglement. However, such an implementation creates subtle numerical and thermodynamical problems, because the chain statistics are no longer determined by the entropy of the chain. On the other hand, the model is not designed to distinguish length and time scales on the order of when the final, single Kuhn step passes through the entanglement. To obviate these problems, we instead give the dangling end an entropy that prevents the last Kuhn step from passing through the entanglement. Then, we give a probability for entanglement destruction that increases with decreasing number of Kuhn steps in the last strand wd = 1 ␶e ͱ3Ne NE , ͑7͒ where NE is the number of Kuhn steps in the end strands. Note that Eq. ͑44͒ of ͓Schieber et al. ͑2003͔͒ contains a typographical error, whereas Eq. ͑20͒ of ͓Nair and Schieber ͑2006͔͒ is correct. This allows us to control the proper statistics of the chain through the entropy and avoid the problems mentioned. Since the short time scales are most affected, we do not expect this approximation to have a significant effect. Nonetheless, we are 1115CUBS MODEL FLOW DATA COMPARISON
  • 6. currently studying a more fundamental implementation of entanglement destruction that avoids these problems entirely, in order to study the effect of the approximation rigor- ously. Related to the above problem, we impose the restriction that the length of an entangle- ment segment should always be greater than the length of a Kuhn step. The strand length is kept greater than the length of a Kuhn step ͑aK͒, through a reflecting boundary condi- tion. We have studied the sensitivity of predictions to this cutoff length and find that results are indistinguishable within statistical error as long as it remains below the aver- age equilibrium entanglement spacing ͓Kitkrailard ͑2005͔͒. The polymer contribution to the stress tensor is given by ͓Nair and Schieber ͑2006͒; Schieber ͑2003a, 2003b͔͒ ␶p = − n͚Z=3 ϱ ͚j=1 Z−1 ͵͵͵Rjͩ‫ץ‬F ‫ץ‬Rj ͪT,͕Ni͖,͕Ri j͖ p͕͑Ri͖,͕Ni͖,Z;t͒dRZ−1 dNZ , ͑8͒ where n is the number density of the polymer chains. B. Brownian dynamics simulation Solution of the evolution Eq. ͑1͒ for the conformational probability density is, in general, not possible analytically. Numerical solution of the probability density is also not feasible given the large number of dimensions possible to describe the conformations. Therefore, Brownian dynamics is used to calculate the model’s predictions for stress ͑or other quantities͒. Finding the equivalent stochastic differential equations and other sto- chastic dynamics from the evolution equation for p are straightforward ͓Gardiner ͑1982͒; Honerkamp ͑1983͒; Öttinger ͑1996͔͒. The evolution equation for the number of Kuhn steps in strand i can be written as dNi = 1 kBT␶i ͑␮i+1 − ␮i͒dt − 1 kBT␶i−1 ͑␮i − ␮i−1͒dt + ͱ2 ␶i dWi − ͱ 2 ␶i−1 dWi−1, i = 1, ... ,Z, ͑9͒ where ␮iª͑‫ץ‬F/‫ץ‬Ni͒T,͕Rj͖,͕Nj i͖ is the chemical potential of the ith strand, except for the end strands, where ␮0=␮1, ␮Z+1=␮Z, and ␶0=␶Z=ϱ. An ensemble of chains whose dynamics are described by this SDE are simulated in discrete time, by using a semi- implicit stepping algorithm. Details are given in Nair and Schieber ͑2006͒. The evolution equation for the position of entanglements is given as dRi = ␬ · Ridt − NeaK 2 12kBT␶i CRͩ‫ץ‬F ‫ץ‬Ri ͪT,͕Ni͖,͕Rj i͖ dt + ͱNeaK 2 6␶i CR dWi. ͑10͒ The ensemble size used is 5000–10 000 chains and time-step size ͑⌬t͒ is selected as 0.02␶e. Chain conformations are generated from the equilibrium probability density as- suming zero cutoff length in Q and then segments smaller than aK are reflected. This procedure gives the proper statistics for Q, but distorts the statistics for the ͕Ni͖ some- what. In order to correct the Kuhn-step-number statistics, it is necessary to equilibrate a few time steps. We equilibrate ͗Z͘eq␶e/⌬t time steps, which is sufficient to create an equilibrium ensemble. Simulations with and without constraint release ͑CR͒ are per- formed and the effect of convective constraint release ͑CCR͒ is incorporated as shown by Eq. ͑6͒. 1116 SCHIEBER, NAIR, AND KITKRAILARD
  • 7. The shear rates used in the simulations are made dimensionless using ␶e. The largest value of Ne used for calculations is 50, since theoretical predictions are insensitive to Ne when greater than 50, at fixed ͗Z͘eq. III. EXPERIMENTAL SYSTEMS Seven polymer systems are compared with the CUBS model: a polystyrene solution, a polybutadiene solution, and five polystyrene melts. The different systems are summarized in Table I. Test fluid PS1 is a 12 wt % solution in tricresyl phosphate nearly monodisperse ͑poly- dispersity indexϭ1.2͒ polystyrene with a molecular of weight Mw of 1.92ϫ106 Da at a polymer density of 0.135 g/cm3 ͓Venerus and Kahvand ͑1994a͔͒. In the case of polymer solutions, the entanglement molecular weight is calculated as Me solution Х Me melt ␾1.3 , ͑11͒ where ␾ is the volume fraction of the polymer and Me melt Х18 000 g/mol. At any rate, the entanglement molecular weight is confirmed in LVE comparisons by examination of the height of the plateau modulus. We use the same fluid for making data comparisons in reversing double-step strain flow ͓Brown and Burghardt ͑1996͒; Venerus and Kahvand ͑1994a, 1994b͔, inception of steady shear flow, steady shear flow, cessation of steady shear flow ͓Kahvand ͑1995͔͒, and exponential shear flow ͓Neergaard et al. ͑2000͔͒. Me for the solution is calculated as given by Eq. ͑11͒ and the value of ͗Z͘eq is estimated ͑͗Z͘eq=Mw/Me͒ to be 10. From the linear viscoelastic comparisons, ␶e for the solution is obtained as, ␶e NCR =0.84 s and ␶e CR Х1.7 s at 23 °C. Comparison between G* data and theory to determine ␶e for this system is given in Nair and Schieber ͑2006͒. The experimental sample used for data comparisons in single-step strain experiments is the polystyrene melt PS2 of molecular weight 200 kDa ͓Nair ͑2004͔͒. Using Me TABLE I. Summary of experimental systems compared with theory in this article: Mwϭweight-averaged molecular weight in kDa, PDIϭMw /Mn, polydispersity index ͑where Mn is the number-averaged molecular weight͒, Tϭtemperature in degrees Celsius, ͗Z͘eqϭaverage number of entanglements per chain plus one; ␶e ͑NCR͒ϭcharacteristic strand relaxation time assuming no constraint release in seconds; ␶e ͑CR͒ϭcharacteristic strand relaxation time with constraint release in seconds. WLC indicates use of the worm-like chain free energy; otherwise, the Gaussian free energy expression was used. System Mw ͑kDa͒ PDI T ͑°C͒ ͗Z͘eq ␶e ͑NCR͒ ͑s͒ ␶e ͑CR͒ ͑s͒ Source͑s͒ PS1 sol. 1920 1.20 23 10 0.84 1.7 Venerus and Kahvand ͑1994a͒ Venerus and Kahvand ͑1994b͒ Kahvand ͑1995͒ Neergaard et al. ͑2000͒ PS2 melt 200 1.05 160 11 0.4 0.78 Nair ͑2004͒ PS3 melt 500.5 1.02 170 28 0.375 Collis et al. ͑2005͒ PS4 melt 200 1.06 175 11 0.137 Schweizer et al. ͑2004͒ PS5 melt 200 1.05 130 14 56.4 Bach et al. ͑2003͒ 56.9 ͑WLC͒ PS6 melt 390 1.06 130 28 23.4 ͑WLC͒ Bach et al. ͑2003͒ PS7 soln 3900 1.05 21 27.5 0.023 ͑WLC͒ Bhattacharjee et al. ͑2003͒ PB soln 813 Ͻ1.05 25 21 0.15 Menezes and Graessley ͑1980͒ 1117CUBS MODEL FLOW DATA COMPARISON
  • 8. =18 100 g/mol ͓Ferry ͑1980͔͒, the ͗Z͘eq for the melt is estimated to be 11 at 160 °C. The value of ␶e for the melt is obtained as, ␶e NCR =0.4 s and ␶e CR =0.78 s at 160 °C and Ne is taken as 22. Data for a nearly monodisperse polybutadiene in Flexon 391 solution ͑PB/Flexon391, polydispersity index Ͻ1.05͒ with a concentration of 7 wt % and Mw=813 kDa is taken from ͓Menezes and Graessley ͑1980͔͒. The experiment was done at 25 °C with a poly- mer density of 0.972 g/cm3 . As is typically done, we first estimated the number of entanglements by using the plateau modulus. However, further refinement was necessary after comparison with LVE data. Therefore, we performed a detailed fit of the linear viscoelastic prediction with the experimental data in order to fix the value of ͗Z͘eq at approximately 21 ͑see Fig. 6͒. The estimated value of Ne for polystyrene solution is 545. However, we take Ne=50 in order to save computational time. The value of ␶e for the melt is estimated as ␶e NCR =0.15 s from the LVE comparison. Fits for PS1 and PS2 are given elsewhere ͓Nair and Schieber FIG. 2. A comparison of the storage and loss moduli predictions with and without constraint release and that for polystyrene melt PS4 ͓Schweizer et al. ͑2004͔͒. FIG. 3. A comparison of the storage and loss moduli predictions and those measured by ͓Bach et al. ͑2003͔͒ for PS5. 1118 SCHIEBER, NAIR, AND KITKRAILARD
  • 9. ͑2006͔͒. The remaining fits are shown in Figs. 2–6. Some predictions are found using the worm-like chain ͑WLC͒ free energy, which is discussed below. We find no difference between predictions using the Gaussian free energy and the WLC for LVE, which is expected for small deformation flows. FIG. 4. A comparison of the storage and loss moduli predictions and those measured by ͓Bach et al. ͑2003͔͒ for PS6. FIG. 5. A comparison of the storage and loss moduli predictions and those measured by ͓Bhattacharjee et al. ͑2003͔͒ for PS7 solution. FIG. 6. Storage and loss moduli predictions of the CUBS model with constraint release, and the experimental data of ͓Menezes and Graessley ͑1980͔͒ for a PB solution of Mw =813 kDa. 1119CUBS MODEL FLOW DATA COMPARISON
  • 10. A high-molecular-weight polystyrene melt of ͓Collis et al. ͑2005͔͒ at 170 °C with Mw=500.5 kDa and PI index of 1.02 is PS3. By using MeϷ18 000, the average number of entangled strands, ͗Z͘eq, is 28. The G* data/theory comparisons are shown in Fig. 7 without constraint release. A low-molecular-weight polystyrene melt of ͓Schweizer et al. ͑2004͔͒, MwХ200 kDa is called PS4. The estimated value of ͗Z͘eq is 11. In both systems, the average number of Kuhn steps, Ne, is 22. From the predictions obtained from the CUBS model without constraint release, the characteristic time for the high molecular weight and the low molecular weight system are 0.0375 and 0.137 s, respectively. IV. COMPARISON IN FLOW We now study in detail the nonlinear predictions of CUBS with constant chain friction. All the flow predictions made here are without adjusting any parameters. FIG. 7. A comparison between the storage GЈ and loss GЉ moduli predictions of the model without constraint release and that for polystyrene melt PS3 ͓Collis et al. ͑2005͔͒, Mw =500.5 kDa, ͗Z͘eq=28 at 170 °C. FIG. 8. Ratio of nonlinear to linear shear modulus as functions of dimensionless time for different assumptions in the model and an approximate, analytic result found by assuming a separation of time scales between retraction and chain reorientation. In this comparison, the chain has an average of 35 entanglements, Ne =22, and the strain is ␥=6.3. 1120 SCHIEBER, NAIR, AND KITKRAILARD
  • 11. A. Single-step strain The earliest tube model derived by Doi and Edwards made a remarkable parameter- free comparison with the strain-dependent part of single-step strain data. However, future refinements seemed to worsen this very favorable comparison. Earlier, one of us derived an analytic expression for the damping function based on the slip-link model under certain assumptions ͓Schieber ͑2003a͔͒. We now check this analytic result through nu- merical solution of the model. A step strain is imposed on a material instantaneously at time t=0. The relaxation of shear stress can be expressed in terms of the nonlinear relaxation modulus, G͑t,␥͒, defined as G͑t,␥͒ = − ␶yx ␥ , ͑12͒ where ␥ is the magnitude of the imposed strain. Beyond a certain time ␶k that depends on the polymer system, the nonlinear relaxation modulus shows separability into time- and strain-dependent parts over a wide range of strains FIG. 9. Shear stress relaxation following single-step strain as functions of time for several strains predicted by the model with and without CR/CCR and PS2 data ͓Nair ͑2004͔͒. FIG. 10. Shear stress relaxation following double-step strain of the type-B flow with ␥1 =4 and ␥2 =−2 at several values of t1 predicted by the model with and without CR/CCR and PS1 data ͓Venerus and Kahvand ͑1994a͔͒. 1121CUBS MODEL FLOW DATA COMPARISON
  • 12. G͑t,␥͒ = G͑t͒h͑␥͒, t ജ ␶k, ͑13͒ where G͑t͒ is the relaxation modulus of linear viscoelasticity and h͑␥͒ is the damping function. In the limit of small strains, the nonlinear relaxation modulus becomes G͑t͒. The damping function is a monotonically decreasing function of strain. Figure 8 shows a comparison of the ratio of nonlinear to linear shear modulus pre- dicted by different models for ͗Z͘eq=36 and the analytic result. We see that the model with CR/CCR and constant chain friction shows better agreement with the analytic h͑␥͒. In Fig. 9 we show comparisons of shear stress relaxation of PS melt ͑Mw=200 K, ͗Z͘eq=11͒ as functions of time for several strains. Except for the highest strain, the model predictions with convective constraint release look good. A previous article compared the damping function predictions with data, and the curves were a bit more strain softening than either of Doi and Edwards’ predictions. Therefore, our theory appears to continue the venerable tradition of degrading the Doi–Edwards-model prediction of the damping function, but improving all other comparisons, as we now show. B. Reversing double-step strain Double-step strain experiments provide a more severe test for evaluating the rheologi- cal constitutive theories for concentrated polymer systems. In these experiments, a strain of ␥1 is imposed at time t=0, followed by a strain of ␥2 after a time t1, in the same ͑␥2Ͼ0͒ or opposite ͑␥2Ͻ0͒ direction to the first step strain. In this study, we consider only reversing double-step strain flows where the second strain is applied in a direction opposite to the first. Our model is used here for predicting two types of reversing flows, type-B ͑␥2=−␥1/2͒ and type-C ͑␥2=−␥1͒. It is known that the Doi–Edwards ͑DE͒ model with the independent alignment ͑IA͒ assumption ͓Doi and Edwards ͑1986͔͒ which is a particular case of, and performs iden- tically to the K-BKZ model ͓Bernstein et al. ͑1963͒; Kaye ͑1962͔͒ in step strains does not perform well in predicting reversing double-step strain deformation. Double-step strain experiments have also been used to test the DE model with and without the IA assump- tion ͓Brown and Burghardt ͑1996͒; Osaki et al. ͑1981͒; Osaki and Kurata ͑1982͒; Venerus and Kahvand ͑1994a, 1994b͒; Venerus et al. ͑1990͔͒. Doi derived a rigorous constitutive equation by avoiding the IA assumption ͓Doi ͑1980͔͒ but still assuming instantaneous chain retraction, which was able to explain reversing double-step strain deformation as long as the time between the two steps is FIG. 11. First normal stress relaxation for the type-B flow with ␥1 =4 and ␥2 =−2 at t1 =4 s predicted by the model with and without constraint release, and PS1 data of ͓Brown and Burghardt ͑1996͔͒. 1122 SCHIEBER, NAIR, AND KITKRAILARD
  • 13. much larger than the retraction time ͑␶R͒. However, when t1ഛ␶R, the model fails ͓Brown and Burghardt ͑1996͒; Venerus and Kahvand ͑1994a, 1994b͔͒. Also, the equation derived by Doi is not applicable to any other flows. Model predictions for type-B flows with ␥1=4 and ␥2=−2 are compared with PS1 in Fig. 10. We see that except at short times, the model shows good agreement with experi- mental data and the results are not strongly sensitive to constraint release. The discrep- ancy shown at short times might be attributed to the coarse-graining introduced in the model. As mentioned before, the model cannot capture the relaxations taking place below the strand size. Figure 11 shows the comparison of first normal stress ͑N1͒ relaxation at t1=4 s. Here the model without CR/CCR gives better predictions compared to the pre- dictions of K-BKZ and DE theories shown in Fig. 11 of the article by Venerus and Kahvand ͑1994b͒. Figure 12 shows comparison of the second normal stress ͑N2͒ relaxation at t1=1 s. As in the case of N1, we see that the model without CR/CCR gives better predictions for N2, except at short times. In Fig. 13, we show the comparisons of the shear stress relaxation for type-C flows with ␥1=4 and ␥2=−4 at t1=2 and 32 s. Here the data seem to show a similar shape for FIG. 12. Second normal stress relaxation for type-B reversing double-step strain with ␥1 =4 and ␥2 =−2 at t1 =1 s predicted by the model with and without constraint release, and PS1 data from Brown and Burghardt ͑1996͒. FIG. 13. Shear stress relaxation for type-C flow with ␥1 =4 and ␥2 =−4 at t1 =2 and 32 s predicted by the model with and without CR/CCR and PS1 data ͓Venerus and Kahvand ͑1994a͔͒. 1123CUBS MODEL FLOW DATA COMPARISON
  • 14. different waiting times, and just a vertical displacement. The model without CCR show the same feature, but CCR does not. Nonetheless, the model does predict the correct magnitudes for the shear stress. Figure 14 shows the transient normal stress ratio for type-C flows at t1=1, 4, and 16 s. Here, we see that the normal stress ratio is time dependent and also t1 dependent. The trend is shown by experimental data in Fig. 6͑a͒ of Brown and Burghardt ͑1996͒. C. Inception of steady shear flow Figures 15 and 16, compare the model predictions to the PS1 solution data for tran- sient growth of shear viscosity and first normal stress difference, respectively, at several shear rates. In both cases, at high shear rates the model without CR and CCR captures the features of the data quantitatively, whereas, at low shear rates the model with CR and CCR predicts the data well. The linear viscoelastic limit forms an upper envelope for all the shear rates and the transient overshoot occur at medium and high shear rates, in accordance with experiments. It is interesting to note that at high shear rates, the model without CR/CCR captures both the overshoot and undershoot, while CR/CCR makes the undershoot disappear. FIG. 14. Second- to first-normal-stress ratio for type-C flow with ␥1 =4 and ␥2 =−4 at t1 =1, 4, and 16 s ͑from top to bottom͒ predicted by the model with and without constraint release. FIG. 15. Transient viscosities as functions of time under start up of steady shear at several shear rates predicted by the model with and without constraint release, and he experimental data of polystyrene solution PS1 ͓Kahvand ͑1995͔͒. 1124 SCHIEBER, NAIR, AND KITKRAILARD
  • 15. These data can also be examined in terms of the extinction angle, Fig. 17. We estimate the extinction angle as ␹= 1 2 tan−1 ͑2␶yx/N1͒, which is valid within our Gaussian approxi- mation. Here also, the model without CR/CCR shows better agreement at high shear rates, while the model with CR/CCR gives better predictions at low shear rates. We can see that the theory is able to capture the undershoot in extinction angle at high shear rates. High deformation rate appears to suppress the effect of constraint release. At any rate, constraint release does not have a large effect on predictions, which contradicts the predictions of tube models ͓Ianniruberto and Marrucci ͑1996, 2000͒; Marrucci ͑1996͔͒. We also examine polystyrene melts PS3 and PS4. Figure 18 shows a comparison between the CUBS model and the experimental viscosities of PS4. There is some dis- crepancy between the LVE predictions and the data at small strains, suggesting some experimental difficulties in this range. Taking this inconsistency into account the theoret- ical predictions are reasonable, especially at the higher Deborah numbers. Figure 19 shows a similar comparison for PS3. These data show an even larger discrepancy with the LVE curve at small strains, suggesting some experimental error, possibly an overestimate of sample size. Interestingly, the difference between the data FIG. 16. Transient growth of first normal stress difference as functions of time under start up of steady shear flow at several shear rates predicted by the model with and without constraint release, and the experimental data of polystyrene solution PS1 ͓Kahvand ͑1995͔͒. FIG. 17. Transient extinction angle as functions of strain under start up of steady shear flow at several shear rates predicted by the model with and without CR/CCR and experimental data of polystyrene solution PS1 ͓Kahvand ͑1995͔͒. 1125CUBS MODEL FLOW DATA COMPARISON
  • 16. FIG. 18. Transient viscosities as functions of time under start up of steady shear at several shear rates predicted by the model with and without constraint release, and the experimental data of polystyrene melt PS4 ͓Schweizer et al. ͑2004͔͒. FIG. 19. Transient viscosities as functions of time under start up of steady shear at several shear rates predicted by the model with and without constraint release and the experimental data of polystyrene melt PS3 ͓Collis et al. ͑2005͔͒. FIG. 20. Transient viscosities as functions of time under start up of steady shear at several shear rates predicted by the model with and without constraint release, and the experimental data of the polybutadiene solution ͓Menezes and Graessley ͑1980͔͒. 1126 SCHIEBER, NAIR, AND KITKRAILARD
  • 17. and theory is nearly identical to the discrepancy between data and the LVE curve, which would also agree with our speculation about the sample’s being somewhat smaller than assumed. Predictions from the CUBS model also show good agreement with the experimental results for the PB solution data in Figs. 20 and 21. D. Steady shear flow The original Doi–Edwards model predicts a maximum in the shear stress with shear rate at steady state, whereas experiments predict a monotonic increase of shear stress with shear rate. The DE model predicts a power-law index of 3 2 for the steady-state viscosity in shear flow ͓Doi and Edwards ͑1979͔͒. Öttinger considered a modification of the Doi–Edwards and Curtiss–Bird models to include the effect of constraint release. This modified reptation model predicts an im- proved power-law index of 4 3 for the steady-state viscosity ͓Öttinger ͑1994͔͒. Ianniruberto and Marrucci suggested a modification of the Doi–Edwards tube model to include the mechanism of convective constraint release which is claimed to be important during fast flows of polymer melts. This model predicted that the steady-state shear stress asymp- totically approaches a plateau at high shear rates ͓Ianniruberto and Marrucci ͑1996͔͒, and it has been widely believed that CCR is a necessary mechanism to avoid a local maxi- mum. More about this point is discussed in the conclusions. Before making quantitative comparisons with data, we looked at the question of monotonicity in the steady-state shear-stress curve for a more highly entangled chain: ͗Z͘eq=30 and Ne=22. Figure 22 shows the result in dimensionless form. It should be noted that the Deborah number is not in terms of the longest relaxation time, but rather De=␥˙␶e. Note that things like constraint release and Rouse timescales arise from the model, rather than being added to the model. Therefore, the monotonic growth of the shear stress is a result of the assumed physics. In fact, in this calculation constraint release is switched off. Therefore, we can conclude that CCR is unnecessary to avoid the maximum. Second, the effect of stretch can clearly be seen for De values greater than approximately 0.2. From the same curve we can estimate the onset of the longest relaxation time at about FIG. 21. Transient growth of first normal stress difference as functions of time under start up of steady shear flow at several shear rates predicted by the model with and without constraint release, data of the polybutadiene solution ͓Menezes and Graessley ͑1980͔͒. 1127CUBS MODEL FLOW DATA COMPARISON
  • 18. De=3ϫ10−3 . This gives us a ratio of ϳ67 for the longest relaxation time to the Rouse time, in accord with the usual estimate of 3͗Z͘eq for tube models, realizing that primitive- path length fluctuations should lower this estimate. Figure 23 shows the shear stress predictions and data for PS1. Only a slight depen- dence is seen on CR. As expected, no maximum is visible. The steady-state viscosity ␩ and the first normal stress difference coefficient ⌿1 for PS1 have power-law indices of 0.79 and 1.4, respectively. The model shows excellent agreement, independent of con- straint release, with power-law indices of 0.82 and 1.37, respectively, when CR/CCR is present and 0.82 and 1.38 when CR/CCR are switched off. The small discrepancy could be due to the small amount of polydispersity present in the polystyrene system used here. The shear rates used to evaluate the power-law indices range from 0.1 to 40 s−1 . Similarly good agreement in the power-law indices are found ͑not shown͒ for the steady state viscosity as a function of shear rate for PS4, with a power-law index of 0.84 for ␩ when CR/CCR is not present in the model. The polybutadiene solution data exhibit power-law indices of 0.87 and 1.65 respectively, for ␩ and ⌿1 when CR is not present in the model, in agreement with predictions. FIG. 22. Dimensionless steady-state shear stress as a function of dimensionless shear rate De=␶e␥˙ predicted by the theory without constraint release, ͗Z͘eq=30 and Ne =22. The solid line is the linear viscoelastic prediction obtained from the CUBS model. FIG. 23. Steady-state values of shear stress as functions of shear rate under steady shear flow predicted by the model with and without constraint release, and experimental data of polystyrene solution PS1 ͓Kahvand ͑1995͔͒. 1128 SCHIEBER, NAIR, AND KITKRAILARD
  • 19. E. Cessation of steady shear flow Figure 24 shows the relaxation of shear stress following cessation of steady shear flow at two shear rates for PS1. We can see that the relaxation is faster at the higher shear rate than at the smaller shear rate, which is due to the faster chain retraction. These trends are consistent with theory. At higher shear rate ͑1.0 s−1 ͒, the models with and without CR/ CCR shows good predictions at shorter times since relaxation is mainly due to monomer density fluctuation and contour-length fluctuation. At larger times, the model with CR/ CCR shows better agreement since relaxation is mainly due to reptation and double reptation. F. Exponential shear flow In exponential shear flow the shear strain increases exponentially with time, given by ␥͑t͒ = exp͑␣t͒ − exp͑− ␣t͒, ͑14͒ where ␣ is a parameter governing the shear acceleration. Several experimental ͓Demarquette and Dealy ͑1992͒; Doshi and Dealy ͑1987͒; Sam- urkas et al. ͑1989͒; Sivashinsky et al. ͑1984͒; Venerus ͑2000͒; Zülle et al. ͑1987͔͒ and theoretical ͓Kwan and Shaqfeh ͑1999͒; Neergaard et al. ͑2000͒; Graham et al. ͑2001͔͒ studies on exponential shear flow have been reported during the past two decades. Neer- gaard et al. used the full chain reptation model ͓Hua and Schieber ͑1998b͔͒ to interpret the exponential shear flow data for entangled polystyrene solution. Based on their model predictions, they reached the conclusion that exponential shear flow stretches the chains no more than does inception of steady shear ͓Neergaard et al. ͑2000͔͒. In fact, that work showed that all exponential shear flow data could be described by steady shear flow data when plotted versus instantaneous shear rate. Hence, we show all comparisons in that form here. In Figs. 25 and 26, we show the comparisons of unadulterated shear stress and first normal stress difference, respectively, as functions of instantaneous shear rates for PS1. We can see that the model gives excellent predictions of shear stress except at high ␣, where the model with CR/CCR slightly overpredicts stretch. In the N1 predictions, the model shows good agreement at low ␣, but again somewhat overpredicts stretch at high ␣. FIG. 24. Relaxation of shear stress as functions of time following cessation of steady shear flow at two shear rates: model predictions with and without constraint release, and experimental data from ͓Kahvand ͑1995͔͒. 1129CUBS MODEL FLOW DATA COMPARISON
  • 20. FIG. 25. Shear stress as functions of instantaneous shear rate for exponential shearing flow at several values of ␣ predicted by the model with and without constraint release, and PS1 data ͓Neergaard et al. ͑2000͔͒. FIG. 26. First normal stress difference as functions of instantaneous shear rate for exponential shear flow at several values of ␣: model predictions with and without constraint release, and PS1 data ͓Neergaard et al. ͑2000͔͒. FIG. 27. Transient elongational stress as a function of time, predicted by the CUBS model with Gaussian and wormlike chain free energy and PS5 melt, Bach et al. ͑2003͒ Mw =200 kDa. 1130 SCHIEBER, NAIR, AND KITKRAILARD
  • 21. G. Elongational flow Only recently has it been possible to achieve rheometric control for uniaxial extension of entangled, linear polymers. Concentrated solutions ͑PS7͒ have been examined at Mo- nash University ͓Bhattacharjee et al. ͑2003͔͒, and melts ͑PS5 and PS6͒ at the Danish Technical University ͓Bach et al. ͑2003͔͒. Note that systems PS6 and PS7 are particularly interesting, because they have nearly identical entanglement number, but only one is solution. We first examine the transient melt data, PS5, in Fig. 27. Here we see the first clear failure of the model. Whereas the data show a monotonic approach to steady state, the model exhibits oscillation until very large strains. Nonetheless, if we compare the steady elongational stresses, we see in Fig. 28, that the slip-link model is able to capture the steady elongational stresses very well. It is natural to guess that the infinite-extensibility assumption in the model is the cause of the discrepancy with the elongational data. Therefore, we also considered a free energy that prevents any segment from being stretched past its contour length. Namely, we use a free energy that becomes infinitely large as Q→NaK. Here we use an approximation to the worm-like chain free energy for an entangled strand FS͑Q,N,T͒ kBT = 3Nlp − 2Q 4͑Nlp − Q͒ Q2 Nlp 2 − log J͑N͒, ͑15͒ which we derive from the approximate tension expression of ͓Marko and Siggia ͑1995͔͒. The function J͑N͒ is a normalization factor dependent upon the number of persistence lengths in the strand, lp=aK/2 is the persistence length and now N is the number of persistence lengths instead of the number of Kuhn steps. The function J is defined as J͑N͒ ª ͫ4␲͑Nlp͒3 ͵0 Nlp Q2 expͭ3Nlp − 2Q 4͑Nlp − Q͒ Q2 Nlp 2 ͮdQͬ−1 . ͑16͒ Since we cannot perform the integration analytically, we use a minimax approximation to obtain a numerical expression for J. Details about the numerical expression can be found in the Appendix. Figures 27 and 28 also show the results for the slip-link model with finite extensibility. Surprisingly, we see that finite extensibility ͑or, more accurately in this case, “finite monomer density”͒ plays only a small role in the transient elongational stresses, and FIG. 28. Steady state elongational stress, predicted by the CUBS model with Gaussian and worm-like chain free energy and P5 melt, Bach et al. ͑2003͒ Mw =200 kDa. 1131CUBS MODEL FLOW DATA COMPARISON
  • 22. almost no role in steady elongational stresses. Similar comparisons are made for the more highly entangled sample, PS6, in Figs. 29 and 30, with similar conclusions. We now turn to the elongational data for the entangled solution, PS7. Recall that this sample has the same number of entanglements per chain as the earlier melt, PS5, whose steady elongational data the theory was able to capture ͑though not the transients͒. Except for the smallest elongation rate, the slip-link model with finite extensibility is unable to capture the transient data ͑Fig. 31͒. However, unlike the melt comparison, the predictions substantially underpredict the steady stress data in Fig. 32, and even show the wrong shape. To examine this discrepancy further, we place the solution and melt data on the same plot by making the strain rate dimensionless by ␶e, and the elongational viscosity dimen- sionless by the zero-shear-rate viscosity. Recall that both of these dimensional quantities are determined by comparison at LVE. Figure 33 clearly shows that the melt and the solution trends are opposite: the solution is strain hardening and the melt is strain soft- ening at high strain rates. Theories for concentrated systems to date have assumed an equivalence between en- tangled solutions and entangled melts, at least implicitly by assuming that entanglements determine the dynamics. Clearly, such an assumption is not in agreement with our final FIG. 29. Transient elongational stress as a function of time, predicted by the CUBS model with worm-like chain model and PS6 melt, Bach et al. ͑2003͒ Mw =390 kDa. FIG. 30. Steady state elongational stress, predicted by the CUBS model with worm-like chain model and PS melt, Bach et al. ͑2003͒ Mw =390 kDa. 1132 SCHIEBER, NAIR, AND KITKRAILARD
  • 23. figure. There are three possibilities: ͑1͒ one of the data sets is ͑or both are͒ incorrect, ͑2͒ additional physics is important in distinguishing solutions from melts, or ͑3͒ an assump- tion in the model is incorrect. Since the elongational data have not yet been reproduced in other labs, the first option cannot be ruled out—easily said for a theoretician. FIG. 31. Transient elongational stress as a function of time, predicted by the CUBS model with worm-like chain model and PS7, Bhattacharjee et al. ͑2003͒ Mw =3900 kDa. FIG. 32. Steady state elongational stress, predicted by the CUBS model with Gaussian chain model and worm-like chain model and PS7, Bhattacharjee et al. ͑2003͒ Mw =3900 kDa. FIG. 33. Dimensionless steady state elongational viscosities comparison between polystyrene melt, Bach et al. ͑2003͒ ͑390 kDa at 130 °C͒, and polystyrene solution, Bhattacharjee et al. ͑2003͒ ͑3.9 MDa at 21 °C͒. 1133CUBS MODEL FLOW DATA COMPARISON
  • 24. We also note that polymer density fluctuations are much larger for the solutions than they are for the melts. Such fluctuations are not considered in the model, since entangle- ments are considered fixed in space for LVE, and to move affinely for flows. These fluctuations might explain the second option. A straightforward way to examine this assumption is in a multichain simulation of the sort used by ͓Masubuchi et al. ͑2001͔͒. There are a number of assumptions made in the model: constant chain friction; mim- icking constraint release through slip-link diffusion, instead of creation and destruction of entanglements in the middle of the chain; destruction of the entanglements at the end of the chain through a probability, instead of as a boundary condition; slip links are de- formed affinely. Discrepancy with data because of the first of these assumptions seems unlikely because constant-entanglement friction has also been tested and found to be somewhat inferior at predicting shear and elongational stresses. The second and third assumptions are currently being examined in a more rigorous formulation of the model. The fourth assumption is more problematic to examine in a mean-field model, such as assumed here. Again, a simulation similar to that of ͓Masubuchi et al. ͑2003͔͒ might be more appropriate to examine this assumption, since the motion of the entanglements could be followed rather than assumed. We do note in passing that we have also exam- ined alternative assumptions about entanglement motion in light of thermodynamic con- siderations. However, these calculations found that fluctuating entanglement motion yielded affine motion, on average. V. CONCLUSIONS We introduce a consistently unconstrained Brownian slip-link model with constant chain friction to study the nonlinear properties of linear entangled polymer melts. The model uses a single phenomenological parameter ␶e, which is fit by the linear viscoelas- ticity data and all the flow predictions are made without any adjustable parameters. We found good to excellent agreement with all shear flows studied: single step, re- versing double step, inception of steady shear, steady shear, cessation of steady shear, and exponential shear. Results were not very sensitive to CR—in contrast with tube models. Studies using various formulations of tube models consistently show that constraint re- lease substantially reduces the shear thinning seen in steady shear flow. However, the studies do not make clear whether CCR is sufficient to avoid the maximum altogether. For example, ͓Marrucci and Ianniruberto ͑2003͔͒ include a prefactor ␤ that should be order 1, but is set to 2 to avoid the maximum. The original article including CCR ͓Ianniruberto and Marrucci ͑2000͔͒ does not use the factor of 2 and exhibits a plateau in the shear stress—still implying flow instabilities. The more recent model of ͓Leygue et al. ͑2006͔͒ also predicts a local maximum in the shear stress and the authors point out that the ␤ parameter of Marrucci et al. could be used as a fix, but decline to do so and write “We nevertheless believe that a better understanding of convective constraint release, especially in reversing flows, could lead to improved formulations where such modifica- tions would not be necessary.” Fang et al. ͑2000͒ decrease the stretching time constant in their stochastic, single-segment model by a factor greater than 2 to give an onset of stretch at smaller shear rates to avoid the maximum. Mead et al. ͑1998͒ decrease the ratio ␶d/␶S by approximately 80% to describe their polystyrene solution of 17 entanglements. They argue that primitive-path length fluctuations justify the change, but the predictions show a change in slope of stress with shear rate ͑from stretch͒ that are not evident in either the shear-stress or normal-stress data. In a more-detailed formulation, ͓Milner et al. ͑2001͔͒ adjust their parameter c␯ to a value of 0.1 in order to avoid the maximum. The physical reasoning given for the parameter’s being less than one is that more than two 1134 SCHIEBER, NAIR, AND KITKRAILARD
  • 25. chains might make up an entanglement ͓see the Appendix of Nair and Schieber ͑2006͒ for an alternative explanation͔. However, ten chains would be inconsistent with the number predicted by the packing criteria of Fetters et al. ͑1999a, 1999b͒ using Rault’s argument ͓Rault ͑1987͔͒, or of the binary interactions found in more recent atomistic simulations ͓Everaers et al. ͑2004͔͒. Our earlier full-chain stochastic model with convective con- straint release ͓Hua et al. ͑1999͔͒ ͑which mixes tubes and slip links͒ also predicts a plateau in the shear stress curve. Since the constraint release is found in a self-consistent way, there are no adjustable parameters like ␤ or c␯. It is possible that tube models predict instability even with convective constraint release. On the other hand, in the dual slip-link model ͓Doi and Takimoto ͑2003͔͒, the shear stress increases monotonically with shear rate in steady shear flow but they do not make any quantitative comparison with experiments. Tubologists might wonder why our slip-link model shows something different from tube models. We mention here three possible explanations. First, we note that tube mod- els typically assume that the number of entanglements on a chain are fixed at a single number—there are no fluctuations at equilibrium, and flow does not change it. At equi- librium, the tube is a random walk with constant step length. It is unclear how this assumption is modified with flow, given that most tube models are on the level of a single step in the primitive path. On the other hand, our slip-link formulation has a distribution of entanglement num- bers at equilibrium, and a distribution of strand lengths; flow modifies both of these distributions in well defined ways. An earlier formulation of the model with constant entanglement friction ͓Schieber et al. ͑2003͔͒ gives results qualitatively similar to this one, but shows how shear flow modifies the entanglement number distribution. In that plot we show that the distribution remains Poissonian, but decreases to about half at high Deborah numbers. The observed disentanglement has two effects that lead in opposite directions. First, chains with fewer entanglements have longer strands that are largely aligned with the flow. These would be expected to contribute to more shear thinning. However, when these chains become dangling ends, they reëntangle rather quickly, lead- ing to disordered strands, reducing shear thinning. We mention that the flow stress curves in that paper are in error, though all equations and dynamic results are correct. This was due to a bug in the portion of the code that calculated stress from chain conformations. When this bug is fixed, all results we have studied for constant-entanglement friction vary only slightly from those with constant- chain friction. However, these results are slightly better for the latter. It is also possible that the difference might be connected to the fact that the dynamics in the slip-link model are connected to the stress through different derivatives of the free energy, whereas in tube models, chain tension itself determines both stress and dynamics. In other words, in the formulation proposed here the derivative of free energy with respect to monomer number is the chemical potential, which determines dynamics; the derivative with respect to slip-link position is tension, which determines stress. As a result, tension along the chain is not equilibrated for strain rates between the inverse of the longest relaxation time, and the inverse of the Rouse time—not true for tube models. For these strain rates, the chemical potential of the chain ends stays near the equilibrium value of −1/2Ne. Hence, chemical potential balance gives the number of Kuhn steps in the strand as approximately N/NeХ͑3/2͓͒ͱ1+͑8Q2 /3NeaK 2 ͒−1͔. Therefore, the mono- mer density N/Q is not constant along the chain, but varies in a nonlinear way with Q. In tube models, equating the tension T=3kBTQ/NaK 2 with the equilibrium Maxwell-demon 1135CUBS MODEL FLOW DATA COMPARISON
  • 26. tension kBT/ͱNeaK requires that the monomer density N/Q stay constant. The maximum observed in tube models might be because the tension—and, hence, the monomer density—is constant along the chain at these strain rates. A third possibility is that the local maximum is avoided because the strands have a sort of built-in polydispersity in lengths, which smooths out the stress curve. One might imagine trying to test these ideas by creating chains with more monodisperse lengths. However, such a formulation would violate the equilibrium conformation statistics valid for concentrated chains. It is also difficult to imagine a test for the second hypothesis that would not violate thermodynamics. The first postulate might be tested by imposing more entanglements in the middle of the chain in order to keep the entanglement density constant. We have examined these dynamics, but obtained stress predictions that were very far from those observed, and different from tube model predictions. The model was able to describe steady elongational data for polystyrene melts, but was unable to describe the transient melt elongational data, nor the solution elongation data. The former discrepancy is likely due to the assumption of how entanglements are destroyed in the model. Evidence of this is seen in the phase-space plot of Fig. 34, where the transient elongational viscosity is plotted versus the average number of entangled strands, and time is parametric. Here we see that disentanglement precedes a drop in the viscosity, followed by reëntanglement that precedes a rise in the viscosity. The latter discrepancy between solution and melt data is more puzzling. If the data hold, it suggests that entangled melts and solutions are not universal and that local polymer concentration fluctuations can play an important role. A more-detailed model would be necessary to examine this hypothesis. ACKNOWLEDGMENT This material is based upon work supported by the National Science Foundation under Grant No. NSF-SCI 0506305. FIG. 34. A phase-space plot of transient dimensionless elongational viscosity versus number of entangled strands, showing that the number of entanglements is highly correlated with the oscillations seen in the elon- gational viscosity. For these calculations, ͗Z͘eq=14, Ne =44␶−e␥˙ =0.453, and the ensemble size is 2000. 1136 SCHIEBER, NAIR, AND KITKRAILARD
  • 27. APPENDIX: WORM-LIKE CHAIN FREE ENERGY To Eq. ͑16͒, we fit a numerical expression for J, J͓y͑N͔͒ = expͫA1 + A2y + A3y2 + A4y3 + A5y4 1 + A6y + A7y2 + A8y3 + A9y4 ͬ, ͑A1͒ where y͑N͒=log N. Values for the constants are given in Table II. The chemical potential for this free energy is found by taking the differential of the free energy with respect to N; hence, a derivative of J is necessary. Rather than take a derivative of our numerical approximation, we first take the derivative analytically ͑A2͒ TABLE III. Constants for the numerical approximation to integral term in the chemical potential expression ␮0, Eq. ͑A4͒. Constant Value M1 −0.5215600145979548 M2 −0.5797035530883076 M3 −0.17660236446248723 M4 −0.10983496030144273 M5 −0.008777278719472151 M6 −0.01894111365316676 M7 0.11070525164370931 M8 0.132323363265643 M9 0.015497437639395373 M10 0.018962828443251357 TABLE II. Constants for the numerical approximation to the normaliza- tion function J, Eq. ͑A1͒. Constant Value A1 0.44080624243588207 A2 2.472397669388139 A3 0.2431635769831712 A4 0.21394499795640387 A5 0.03344400384360319 A6 0.1468954480518177 A7 0.09151779095188171 A8 0.02361837127994027 A9 0.000036340710810897935 1137CUBS MODEL FLOW DATA COMPARISON
  • 28. where we define g͑x͒ as g͑x͒ = 3 − 2x 4͑1 − x͒ x2 . ͑A3͒ Then, from Eq. ͑A2͒ we fit a numerical approximation to the ratio of integrals, which we call ␮0, ␮0 = expͫM1 + M2y + M3y2 + M4y3 + M5y4 + M6y5 1 + M7y + M8y2 + M9y3 + M10y4 ͬ, ͑A4͒ where again y=log N. Values for the constants are given in Table III. In Figs. 35 and 36, respectively, we check our expressions by plotting the relative residual ͑Jfit−J͒/J between the numerical solutions obtained from MATHEMATICA for the integrals and the numerical expression of normalization factor, Eq. ͑A1͒, and the integral term in the chemical potential expression, Eq. ͑A4͒. FIG. 35. Relative residual between the numerical expression of the normalization factor and the numerical fit found by using MATHEMATICA. FIG. 36. Residual between the numerical expression of the integral term of chemical potential expression and the numerical solution from MATHEMATICA software. 1138 SCHIEBER, NAIR, AND KITKRAILARD
  • 29. References Bach, A., K. Almdal, H. K. Rasmussen, and O. Hassager, “Elongational viscosity of narrow molar mass distribution polystyrene,” Macromolecules 36, 5174–5179 ͑2003͒. Bernstein, B. E., A. Kearsley, and L. J. Zapas, “A study of stress relaxation with finite strain,” Trans. Soc. Rheol. 7, 391–410 ͑1963͒. Bhattacharjee, P., D. Nguyen, G. McKinley, and T. Sridhar, “Extensional stress growth and stress relaxation in entangled polymer solutions,” J. Rheol. 47, 269–290 ͑2003͒. Bird, R. B., C. F. Curtiss, R. C. Armstrong, and O. Hassager, Dynamics of Polymeric Liquids Vol. II: Kinetic Theory, 2nd ed. ͑Wiley-Interscience, New York, 1987͒. Brown, E. F., and W. R. Burghardt, “First and second normal stress difference relaxation in reversing double- step strain flows,” J. Rheol. 40, 37–54 ͑1996͒. Collis, M. W., A. K. Lele, M. R. Mackley, R. S. Graham, A. E. Groves, A. E. Likhtman, T. M. Nicholson, O. G. Harlen, T. C. B. McLeish, L. R. Hutchings, C. M. Fernyhough, and R. N. Young, “Constriction flows of monodisperse linear entangled polymers: Multiscale modeling and flow visualization,” J. Rheol. 49, 501– 522 ͑2005͒. Demarquette, N. R., and J. M. Dealy, “Nonlinear viscoelasticity of concentrated polystyrene solutions: Sliding plate rheometer studies,” J. Rheol. 36, 1007–1032 ͑1992͒. Doi, M., “Stress relaxation of polymeric liquids after double-step strain,” J. Polym. Sci., Polym. Phys. Ed. 18, 1891–1901 ͑1980͒. Doi, M., and S. F. Edwards, “Dynamics of concentrated polymer systems. Part 1. Brownian motions in the equilibrium state,” J. Chem. Soc., Faraday Trans. 2 74, 1789–1800 ͑1978a͒. Doi, M., and S. F. Edwards, “Dynamics of concentrated polymer systems. Part 2. Molecular motion under flow,” J. Chem. Soc., Faraday Trans. 2 74, 1802–1817 ͑1978b͒. Doi, M., and S. F. Edwards, “Dynamics of concentrated polymer systems. Part 3. The constitutive equation,” J. Chem. Soc., Faraday Trans. 2 74, 1818–1832 ͑1978c͒. Doi, M., and S. F. Edwards, “Dynamics of concentrated polymer systems. Part 4. Rheological properties,” J. Chem. Soc., Faraday Trans. 2 75, 38–54 ͑1979͒. Doi, M., and S. F. Edwards, The Theory of Polymer Dynamics ͑Clarendon, Oxford, 1986͒. Doi, M., and J.-I. Takimoto, “Molecular modelling of entanglement,” Philos. Trans. R. Soc. London, Ser. A 361, 641–652 ͑2003͒. Doshi, S. R., and J. M. Dealy, “Exponential shear: A strong flow,” J. Rheol. 31, 563–582 ͑1987͒. Eskandari, M., “Optimization of the solid-state shear extrusion process for low-cross-link density natural rubber with a proposed Gaussian slip-link model,” Ph.D. thesis, Illinois Institute of Technology, Chicago, 2006. Everaers, R., S. K. Sukumaran, G. S. Grest, C. Svaneborg, A. Sivasubramanian, and K. Kremer, “Rheology and microscopic topology of entangled polymeric liquids,” Science 303, 823–826 ͑2004͒. Fang, J., M. Kröger, and H. C. Öttinger, “A thermodynamically admissible reptation model for fast flows of entangled polymers. II. Model predictions for shear and extensional flows,” J. Rheol. 44, 1293–1317 ͑2000͒. Ferry, J. L., Viscoelastic Properties of Polymers, 3rd ed. ͑Wiley, New York, 1980͒. Fetters, L. J., D. J. Lohse, S. Milner, and W. W. Graessley, “Packing length influence in linear polymer melts on the entanglement, critical and reptation molecular weights,” Macromolecules 32, 6847–6851 ͑1999a͒. Fetters, L. J., D. J. Lohse, and W. W. Graessley, “Chain dimensions and entanglement spacings in dense macromolecular systems,” J. Polym. Sci., Part B: Polym. Phys. 37, 1023–1033 ͑1999b͒. Foteinopoulou, K., N. Karayiannis, V. Mavrantzas, and M. Kroger, “Primitive path identification and entangle- ment statistics in polymer melts: Results from direct topological analysis on atomistic polyethylene mod- els,” Macromolecules 39, 4207–4216 ͑2006͒. Gardiner, C. W., Handbook of Stochastic Methods for Physics, Chemistry and the Natural Sciences, 2nd ed. ͑Springer, Berlin, 1982͒. Gigras, P. G., and B. Khomami, “Adaptive configuration fields: A new multiscale simulation technique for reptation-based models with a stochastic strain measure and local variations of life span distribution,” J. Non-Newtonian Fluid Mech. 108, 99–122 ͑2002͒. 1139CUBS MODEL FLOW DATA COMPARISON
  • 30. Graham, R. S., A. Likhtman, T. C. B. McLeish, and S. T. Milner, “Microscopic theory of linear, entangled polymer chains under rapid deformation including chain stretch and convective constraint release,” J. Rheol. 47, 1171–1200 ͑2003͒. Graham, R. S., T. C. B. McLeish, and O. G. Harlen, “Using the pom-pom equations to analyze polymer melts in exponential shear,” J. Rheol. 45, 275–290 ͑2001͒. Greco, F., “The stress tensor in entangled polymers,” Phys. Rev. Lett. 88, 108301 ͑2002͒. Honerkamp, J., Stochastische Dynamische Systeme ͑VCH, Berlin, 1983͒. Hua, C. C., and J. D. Schieber, “Segment connectivity, chainlength breathing, segmental stretch, and constraint release in reptation models. I. Theory and single-step strain predictions,” J. Chem. Phys. 109, 10018–10027 ͑1998a͒. Hua, C.-C., and J. D. Schieber, “Viscoelastic flow through fibrous media using the CONNFFESSIT approach,” J. Rheol. 42, 477–491 ͑1998b͒. Hua, C. C., J. D. Schieber, and D. C. Venerus, “Segment connectivity, chainlength breathing, segmental stretch, and constraint release in reptation models. II. Double-step strain predictions,” J. Chem. Phys. 109, 10028– 10032 ͑1998͒. Hua, C. C., J. D. Schieber, and D. C. Venerus, “Segment connectivity, chain length breathing, segmental stretch, and constraint release in reptation models. III. Shear flows,” J. Rheol. 43, 701–717 ͑1999͒. Ianniruberto, G., and G. Marrucci, “On compatibility of the Cox–Merz rule with the model of Doi and Ed- wards,” J. Non-Newtonian Fluid Mech. 65, 241–246 ͑1996͒. Ianniruberto, G., and G. Marrucci, “Convective orientational renewal in entangled polymers,” J. Non- Newtonian Fluid Mech. 95, 363–374 ͑2000͒. Ianniruberto, G., and G. Marrucci, “A simple constitutive equation for entangled polymers with chain stretch,” J. Rheol. 45, 1305–1318 ͑2001͒. Kahvand, H., “Strain coupling effects in polymer rheology,” Ph.D. thesis, Illinois Institute of Technology, Chicago, 1995. Kaye, A., “Non-Newtonian flow of incompressible fluids,” Note No. 134, College of Aeronautics, Cranfield, UK, 1962. Kitkrailard, T., “Dynamics of semi-flexible entangled polymers,” Master’s thesis, Illinois Institute of Technol- ogy, Chicago, Illinois, 2005. Kwan, T. C. B., and E. Shaqfeh, “Brownian dynamics simulations of the stress and molecular configurations of polymers in exponential and linearly–ramped shear flow,” J. Non-Newtonian Fluid Mech. 82, 139–165 ͑1999͒. Leygue, A., C. Bailly, and R. Keunings, “A tube-based constitutive equation for polydisperse entangled linear polymers,” J. Non-Newtonian Fluid Mech. 136, 1–16 ͑2006͒. Likhtman, A., “Single-chain slip-link model of entangled polymers: Simultaneous description of neutron spin- echo, rheology, and diffusion,” Macromolecules 38, 6128–6139 ͑2005͒. Marko, J. F., and E. D. Siggia, “Stretching DNA,” Macromolecules 28, 8759–8770 ͑1995͒. Marrucci, G., “Dynamics of entanglements: A nonlinear model consistent with the Cox–Merz rule,” J. Non- Newtonian Fluid Mech. 62, 279–289 ͑1996͒. Marrucci, G., and G. Ianniruberto, “Flow-induced orientation and stretching of entangled polymers,” Philos. Trans. R. Soc. London, Ser. A 361, 677–688 ͑2003͒. Masubuchi, Y., G. Ianniruberto, F. Greco, and G. Marrucci, “Entanglement molecular weight and frequency response of sliplink networks,” J. Chem. Phys. 119, 6925–6930 ͑2003͒. Masubuchi, Y., J. Ichi Takimoto, K. Koyama, G. Ianniruberto, G. Marrucci, and F. Greco, “Brownian simula- tions of a network of reptating primitive chains,” J. Chem. Phys. 115, 4387–4394 ͑2001͒. Mead, D. W., R. G. Larson, and M. Doi, “A molecular theory for fast flows of entangled polymers,” Macromolecules 31, 7895–7914 ͑1998͒. Menezes, E. V., and W. W. Graessley, “Nonlinear rheological behavior of polymer systems for several shear- flow histories,” J. Polym. Sci., Polym. Phys. Ed. 20, 1817–1833 ͑1980͒. Milner, S. T., T. C. B. McLeish, and A. E. Likhtman, “Microscopic theory of convective constraint release,” J. Rheol. 45, 539–563 ͑2001͒. Nair, D., “Brownian dynamics simulation of consistently unconstrainted Brownian sliplinks,” Master’s thesis, 1140 SCHIEBER, NAIR, AND KITKRAILARD
  • 31. Illinois Institute of Technology, Chicago, 2004. Nair, D. M., and J. D. Schieber, “Linear viscoelastic predictions of a consistently unconstrained Brownian slip-link model,” Macromolecules 39, 3386–3397 ͑2006͒. Neergaard, J., K. Park, D. C. Venerus, and J. D. Schieber, “Exponential shear flow of linear, entangled poly- meric liquids,” J. Rheol. 44, 1043–1054 ͑2000͒. Neergaard, J., and J. D. Schieber, “A full chain network model with sliplinks and binary constraint release,” XIIIth International Congress on Rheology, Cambridge, United Kingdom, 2000, Vol. 2, pp. 120–122. Osaki, K., S. Kimura, and M. Kurata, “Relaxation of shear and normal stresses in double-step shear deforma- tions for a polystyrene solution. A test of the Doi–Edwards theory for polymer rheology,” J. Rheol. 25, 549–562 ͑1981͒. Osaki, K., and M. Kurata, “Stress relaxation of a polymer solution in double-step shear deformations: A separate evaluation of the stress components due to entangled segments of different orientations,” J. Polym. Sci., Part A-2 20, 623–632 ͑1982͒. Öttinger, H. C., “Modified reptation model,” Phys. Rev. E 50, 4891–4895 ͑1994͒. Öttinger, H. C., Stochastic Processes in Polymeric Fluids ͑Springer, Berlin, 1996͒. Öttinger, H. C., “A thermodynamically admissible reptation model for fast flows of entangled polymers,” J. Rheol. 43, 1461–1493 ͑1999͒. Rault, J., “Space filling an entanglements in polymeric systems,” J. Non-Newtonian Fluid Mech. 23, 229–247 ͑1987͒. Rubinstein, M., “Discretized model of entangled-polymer dynamics,” Phys. Rev. Lett. 59, 1946–1949 ͑1987͒. Rubinstein, M., and S. Panyukov, “Elasticity of polymer networks,” Macromolecules 35, 6670–6686 ͑2002͒. Samurkas, T., R. G. Larson, and J. M. Dealy, “Strong extensional and shearing flows of a branched polyeth- ylene,” J. Rheol. 33, 559–578 ͑1989͒. Schieber, J. D., “Fluctuations in entanglements of polymer liquids,” J. Chem. Phys. 118, 5162–5166 ͑2003a͒. Schieber, J. D., “GENERIC compliance of a temporary network model with sliplinks, chain-length fluctuations, segment-connectivity and constraint release,” J. Non-Equil. Thermodyn. 28, 179–188 ͑2003b͒. Schieber, J. D., J. Neergaard, and S. Gupta, “A full-chain, temporary network model with sliplinks, chain-length fluctuations, chain connectivity and chain stretching,” J. Rheol. 47, 213–233 ͑2003͒. Although all dynami- cal calculations were verified to agree with analytic results at equilibrium, we later discovered that stress calculations contained a bug. Hence, stress curves in this manuscript should be ignored. Schweizer, T., J. van Meerveld, and H. C. Öttinger, “Nonlinear shear rheology of polystyrene melt with narrow molecular weight distribution experiment and theory,” J. Rheol. 48, 1345–1363 ͑2004͒. Shanbhag, S., and M. Kroger, “Primitive path networks generated by annealing and geometrical methods: Insights into differences,” Macromolecules 40, 2897–2903 ͑2007͒. Sivashinsky, N., A. T. Tsai, T. J. Moon, and D. S. Soong, “Some new transient test results from a parallel-plate rheometer,” J. Rheol. 28, 287–301 ͑1984͒. Tzoumanekas, C., and D. N. Theodorou, “Topological analysis of linear polymer melts: A statistical approach,” Macromolecules 39, 4592–4604 ͑2006͒. Venerus, D. C., “Exponential shear flow of polymer melts,” Rheol. Acta 39, 71–79 ͑2000͒. Venerus, D. C., and H. Kahvand, “Doi–Edwards theory evaluation in double-step strain flows,” J. Polym. Sci., Part B: Polym. Phys. 32, 1531–1542 ͑1994a͒. Venerus, D. C., and H. Kahvand, “Normal stress relaxation in reversing double-step strain flows,” J. Rheol. 38, 1297–1315 ͑1994b͒. Venerus, D. C., C. M. Vrentas, and J. S. Vrentas, “Step strain deformations for viscoelastic fluids: Experiment,” J. Rheol. 34, 657–684 ͑1990͒. Zhou, Q., and R. Larson, “Primitive path identification and statistics in molecular dynamics simulations of entangled polymer melts,” Macromolecules 38, 5761–5765 ͑2005͒. Zülle, B., J. J. Linster, J. Meissner, and H. P. Hürlimann, “Deformation hardening and thinning in both elongation and shear of a low density polyetyhylene melt,” J. Rheol. 31, 583–598 ͑1987͒. 1141CUBS MODEL FLOW DATA COMPARISON