17.07.2014 Views

wall slip and boundary effects in polymer shear flows - Journal of ...

wall slip and boundary effects in polymer shear flows - Journal of ...

wall slip and boundary effects in polymer shear flows - Journal of ...

SHOW MORE
SHOW LESS

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

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

WALL SLIP AND BOUNDARY EFFECTS IN<br />

POLYMER SHEAR FLOWS<br />

By<br />

William Brian Black<br />

A dissertation submitted <strong>in</strong> partial fulfillment<br />

<strong>of</strong> the requirements for the degree <strong>of</strong><br />

Doctor <strong>of</strong> Philosophy<br />

(Chemical Eng<strong>in</strong>eer<strong>in</strong>g)<br />

at the<br />

UNIVERSITY OF WISCONSIN – MADISON<br />

2000


i<br />

Abstract<br />

Polymer – surface <strong>in</strong>teractions strongly <strong>in</strong>fluence many important <strong>in</strong>dustrial <strong>and</strong><br />

rheological <strong>flows</strong>. In particular, <strong>polymer</strong> melts <strong>and</strong> solutions <strong>slip</strong> aga<strong>in</strong>st the surface;<br />

this has long been associated with sharksk<strong>in</strong> <strong>and</strong> spurt <strong>in</strong> extrusion, <strong>and</strong><br />

recent experimental observations suggest that <strong>slip</strong> also plays a role <strong>in</strong> the formation<br />

<strong>of</strong> enhanced concentration fluctuations <strong>in</strong> entangled <strong>polymer</strong> solutions. Analyses<br />

<strong>of</strong> melt flow <strong>in</strong> model geometries, <strong>in</strong>corporat<strong>in</strong>g simple <strong>slip</strong> models which <strong>in</strong>clude<br />

the <strong>effects</strong> <strong>of</strong> cha<strong>in</strong> orientation <strong>and</strong> stretch<strong>in</strong>g, <strong>and</strong> employ<strong>in</strong>g common l<strong>in</strong>ear <strong>and</strong><br />

nonl<strong>in</strong>ear viscoelastic models for the <strong>polymer</strong> stress, demonstrate that <strong>slip</strong> can lead<br />

to short-wave hydrodynamic <strong>in</strong>stabilities. The results predict values <strong>of</strong> the critical<br />

recoverable <strong>shear</strong>, the critical <strong>shear</strong> rate, <strong>and</strong> the frequency <strong>of</strong> distortion that are<br />

consistent with experimental observations for sharksk<strong>in</strong> dur<strong>in</strong>g extrusion <strong>of</strong> l<strong>in</strong>ear<br />

polyethylenes. Further analysis, employ<strong>in</strong>g <strong>slip</strong> models generalized to <strong>in</strong>clude the<br />

<strong>effects</strong> <strong>of</strong> normal load (i.e.<br />

pressure) at the surface, confirms <strong>and</strong> unifies previous<br />

theoretical work: <strong>shear</strong> stress dependent <strong>slip</strong> models cannot predict <strong>in</strong>stabilities<br />

consistent with sharksk<strong>in</strong>; <strong>and</strong> pressure dependent <strong>slip</strong> leads to <strong>in</strong>stability due to illposedness<br />

at the <strong>boundary</strong>. Computational studies performed for <strong>polymer</strong> solutions<br />

highlight the importance <strong>of</strong> boundaries on the flow behavior. Computations were


ii<br />

carried out us<strong>in</strong>g a two-fluid model <strong>in</strong> which stress <strong>and</strong> concentration are coupled.<br />

This coupl<strong>in</strong>g gives rise to enhanced fluctuations <strong>in</strong> the bulk, <strong>and</strong> the additional<br />

coupl<strong>in</strong>g with <strong>slip</strong> leads to hydrodynamic <strong>in</strong>stability.<br />

The length scales for <strong>slip</strong><br />

(the extrapolation length) <strong>and</strong> <strong>in</strong>stability (the reciprocal <strong>of</strong> the wavenumber for the<br />

disturbance) are the same <strong>and</strong> are consistent with experiment.<br />

Incorporation <strong>of</strong><br />

thermal fluctuations shows that the fluctuations near the surface are dramatically<br />

enhanced relative to those <strong>in</strong> the bulk. The wavevector for enhanced fluctuations<br />

rotates as the <strong>shear</strong> rate <strong>in</strong>creases, <strong>in</strong> agreement with bulk experiments <strong>and</strong> predictions,<br />

so that the mechanism for near-surface enhancement is essentially the same<br />

as <strong>in</strong> the bulk. As opposed to the flow <strong>in</strong>stability, these enhanced fluctuations are<br />

<strong>in</strong>sensitive to the presence <strong>of</strong> <strong>slip</strong>. Overall, these analyses highlight the importance<br />

<strong>of</strong> bound<strong>in</strong>g surfaces on the macroscopic flow behavior <strong>and</strong> <strong>of</strong> coupl<strong>in</strong>g between<br />

various flow features, such as stress, concentration, <strong>and</strong> <strong>slip</strong>.


iii<br />

Acknowledgements<br />

Iwould like to extend my thanks <strong>and</strong> appreciation to those who have helped make<br />

this possible: my wife, Lisa; my advisor, Pr<strong>of</strong>. Michael Graham; <strong>and</strong> my group<br />

mates: Dr. Gretchen Baier, Arun Kumar, Dr. Venkat Ramanan, Dr. John Kasab,<br />

Philip Stone, <strong>and</strong> Richard Jendrejack, whose help <strong>and</strong> advice have been most beneficial.


iv<br />

Contents<br />

Abstract<br />

i<br />

Acknowledgements<br />

iii<br />

List <strong>of</strong> Tables<br />

vii<br />

List <strong>of</strong> Figures<br />

viii<br />

Summary<br />

xi<br />

1 Extrusion Instabilities 1<br />

1.1 Introduction ............................... 1<br />

1.2 Sharksk<strong>in</strong>................................. 3<br />

1.3 WallSlip................................. 9<br />

1.3.1 ExperimentalEvidence...................... 9<br />

1.3.2 PlaneShearFlowAnalyseswithSlip .............. 13<br />

2 Model<strong>in</strong>g <strong>of</strong> Wall Slip 18<br />

2.1 SlipMechanisms............................. 18<br />

2.2 ConnectionsBetweenNormalStresses<strong>and</strong>Slip............. 21


v<br />

2.3 ASlipModelBasedonNetworkTheory................. 24<br />

2.4 AnisotropicDragSlipModel....................... 27<br />

3 Stability <strong>of</strong> Plane Shear Flow <strong>of</strong> a Polymer Melt with Slip 30<br />

3.1 Relevance<strong>of</strong>ViscometricFlowtoDieExitFlow ............ 30<br />

3.2 Formulation ............................... 32<br />

3.3 AsymptoticSolutionswithGeneralSlipModels............. 36<br />

3.4 ResultswithSpecificSlipModels .................... 44<br />

3.4.1 AnalyticalResultsfortheUCMEquation............ 44<br />

3.4.2 NumericalMethod........................ 47<br />

3.4.3 NumericalResults ........................ 51<br />

3.4.4 Comparisonwithexperiment................... 57<br />

3.5 Mechanism................................ 58<br />

3.6 Summary<strong>of</strong>theMeltAnalysis...................... 63<br />

4 Concentration Fluctuations <strong>in</strong> Semidilute Polymer Solutions 65<br />

5 Concentration Fluctuations <strong>and</strong> Flow Instabilities <strong>in</strong> Sheared Polymer<br />

Solutions 73<br />

5.1 Formulation ............................... 73<br />

5.2 Stability Analysis ............................. 77<br />

5.2.1 NumericalMethod........................ 78<br />

5.2.2 Stability Results ......................... 86<br />

5.3 BrownianFluctuations.......................... 89<br />

5.4 Summary................................. 94


vi<br />

6 Conclud<strong>in</strong>g Remarks 99<br />

Nomenclature 101<br />

A Basic Melt Equations 105<br />

A.1 Nondimensionalization.......................... 105<br />

A.2 Derivation <strong>of</strong> the General Stability Equation .............. 107<br />

A.3 PTTMatrixEigenvalueProblem .................... 110<br />

B 3D Stability Operators 112<br />

Bibliography 114


vii<br />

List <strong>of</strong> Tables<br />

1 Summary<strong>of</strong>experimentalresultsforsharksk<strong>in</strong>............ 4<br />

2 Summary<strong>of</strong>measurements<strong>of</strong><strong>slip</strong>velocities. ............. 10<br />

3 Summary <strong>of</strong> stability analyses <strong>of</strong> flow ................. 15<br />

4 Comparison<strong>of</strong>theresultsfor<strong>slip</strong><strong>and</strong>no-<strong>slip</strong>. ............ 61


viii<br />

List <strong>of</strong> Figures<br />

1 Extrusion <strong>in</strong>stabilities. ......................... 2<br />

2 FlowcurveforLLDPE.......................... 5<br />

3 Critical stresses for LLDPE for dies constructed <strong>of</strong> different metals. 6<br />

4 Effect<strong>of</strong>DFLcoat<strong>in</strong>gontheflowcurve<strong>of</strong>HDPE........... 11<br />

5 Effect<strong>of</strong>Dynamarcoat<strong>in</strong>gontheflowcurve<strong>of</strong>HDPE. ....... 12<br />

6 Schematic<strong>of</strong>thetwopr<strong>in</strong>cipalmechanismsfor<strong>slip</strong>. ......... 19<br />

7 K<strong>in</strong>etic<strong>slip</strong>model. ........................... 24<br />

8 Typicaldieexit.............................. 31<br />

9 Stress <strong>boundary</strong> layers <strong>in</strong> corner <strong>flows</strong>. ................ 31<br />

10 Basic parallel <strong>shear</strong> flow geometry with <strong>slip</strong> at the solid surfaces. . . 33<br />

11 Snapshot <strong>of</strong> the destabiliz<strong>in</strong>g disturbance at the onset <strong>of</strong> <strong>in</strong>stability 41<br />

12 Stability diagram for pressure- <strong>and</strong> normal stress- dependent <strong>slip</strong>. . 42<br />

13 Critical Weissenberg number as a function <strong>of</strong> s for several values <strong>of</strong><br />

We s forthenetwork<strong>slip</strong>model. .................... 45<br />

14 Critical We n as a function <strong>of</strong> ɛs for the UCM equation <strong>and</strong> the network<br />

<strong>slip</strong>model................................. 45


ix<br />

15 Growth rate versus wavenumber for the UCM equation <strong>and</strong> the N<br />

<strong>slip</strong>model................................. 46<br />

16 Typical eigenvalue spectra obta<strong>in</strong>ed analytically <strong>and</strong> numerically. . 47<br />

17 Semi-analytical neutral curves for the plane Couette flow <strong>of</strong> the UCM<br />

fluidwiththeanisotropicdrag<strong>slip</strong>model. .............. 48<br />

18 Comparison <strong>of</strong> the growth rate curves predicted numerically <strong>and</strong> analyticallyfortheUCMequation<strong>and</strong>thenetwork<strong>slip</strong>model.....<br />

52<br />

19 Numerical <strong>and</strong> analytical neutral curves for the UCM equation <strong>and</strong><br />

thenetwork<strong>slip</strong>model ......................... 53<br />

20 Neutral curves for the PTT constitutive equation with the network<br />

modelasthe<strong>slip</strong>relation. ....................... 53<br />

21 Numerical neutral curves for the PTT constitutive equation with the<br />

anisotropicdrag<strong>slip</strong>model. ...................... 54<br />

22 Master curve <strong>of</strong> k x b versus We t forthenetworkmodel......... 55<br />

23 Phase shift between the <strong>slip</strong> velocity <strong>and</strong> the <strong>shear</strong> <strong>and</strong> normal stress<br />

componentsatthecriticalpo<strong>in</strong>t. ................... 62<br />

24 Enhanced concentration fluctuations <strong>in</strong> a semidilute polystyrene solution.<br />

.................................. 66<br />

25 Schematic<strong>of</strong>typicallightscatter<strong>in</strong>gexperiments. .......... 67<br />

26 Scatter<strong>in</strong>g<strong>in</strong>tensityasafunction<strong>of</strong><strong>shear</strong>rate............. 67<br />

27 Scatter<strong>in</strong>gpatternasafunction<strong>of</strong>the<strong>shear</strong>rate. .......... 68<br />

28 Physical picture <strong>of</strong> the HF hydrodynamic mechanism for enhanced<br />

concentrationfluctuations........................ 70


x<br />

29 Plane Couette geometry show<strong>in</strong>g <strong>slip</strong> between the solution <strong>and</strong> solid<br />

surface. ................................. 74<br />

30 Eigenvalue spectrum for 2D disturbances with the stress diffusion<br />

termdropped............................... 84<br />

31 Comparison<strong>of</strong>thest<strong>and</strong>ard<strong>and</strong>SIDtechniques............ 85<br />

32 Atypicaleigenvalueobta<strong>in</strong>edus<strong>in</strong>gtheSIDtechnique. ....... 87<br />

33 Neutral curves for k z =0<strong>and</strong>S =10 −2 for various values <strong>of</strong> b. ... 88<br />

34 Neutral curves for three dimensional disturbances at the given values<br />

<strong>of</strong> k x <strong>and</strong> b. ............................... 89<br />

35 Unstable eigenfunction for the concentration. The parameters are<br />

k x =0.4, b = 10, We = 10, S =0.01, N =96.............. 90<br />

36 Thenoisecorrelationfunction...................... 92<br />

37 Correlation function at equilibrium. .................. 92<br />

38 Series <strong>of</strong> concentration correlation functions for <strong>in</strong>creas<strong>in</strong>g We. ... 96<br />

39 Series <strong>of</strong> snapshots <strong>of</strong> typical concentration pr<strong>of</strong>iles as We is <strong>in</strong>creased. 97<br />

40 Eigenvalue spectra for the <strong>slip</strong> <strong>and</strong> no-<strong>slip</strong> cases <strong>in</strong> the bounded flow<br />

doma<strong>in</strong>. ................................. 98


xi<br />

Summary<br />

Polymer extrusion processes are severely limited by flow <strong>in</strong>stabilities <strong>and</strong> product<br />

distortions. The first flow <strong>in</strong>stability observed upon <strong>in</strong>creas<strong>in</strong>g the flow rate from<br />

zero dur<strong>in</strong>g the extrusion <strong>of</strong> high molecular weight, l<strong>in</strong>ear <strong>polymer</strong>s, such as high<br />

density polyethylene (HDPE) <strong>and</strong> l<strong>in</strong>ear low density polyethylene (LLDPE) is sharksk<strong>in</strong>,<br />

which is manifested as a short wavelength, periodic distortion <strong>of</strong> the surface <strong>of</strong><br />

the product. This happens at very low Reynolds numbers, due to the large viscosity<br />

<strong>of</strong> <strong>polymer</strong> melts, <strong>and</strong> at flow rates where the <strong>shear</strong> stress is a monotonic function<br />

<strong>of</strong> the <strong>shear</strong> rate. The onset conditions are affected by the die material, <strong>in</strong>dicat<strong>in</strong>g<br />

that the <strong>in</strong>stability is <strong>in</strong>terfacial <strong>in</strong> orig<strong>in</strong>. In addition, the <strong>in</strong>stability is triggered <strong>in</strong><br />

the die exit region, where the stresses are largest due to the <strong>boundary</strong> s<strong>in</strong>gularity<br />

<strong>and</strong> velocity pr<strong>of</strong>ile rearrangement. As sharksk<strong>in</strong> is the first <strong>in</strong>stability seen <strong>in</strong> the<br />

process<strong>in</strong>g <strong>of</strong> these materials, underst<strong>and</strong><strong>in</strong>g this phenomenon is key to improv<strong>in</strong>g<br />

the productivity <strong>of</strong> these processes.<br />

There have been many theories put forth to expla<strong>in</strong> the onset <strong>of</strong> sharksk<strong>in</strong> distortion,<br />

but, clearly, <strong>wall</strong> <strong>slip</strong> <strong>in</strong>fluences the critical conditions for <strong>in</strong>stability to occur.<br />

Two possible mechanisms for <strong>slip</strong> exist, desorption <strong>of</strong> adsorbed molecules from the<br />

surface <strong>and</strong> disentanglement <strong>of</strong> anchored cha<strong>in</strong>s from the bulk material. In the past,


xii<br />

<strong>slip</strong> has been modeled macroscopically us<strong>in</strong>g <strong>slip</strong> relations where the <strong>slip</strong> velocity is<br />

a function <strong>of</strong> the <strong>shear</strong> stress at the <strong>wall</strong>. However, analyses employ<strong>in</strong>g these types<br />

<strong>of</strong> models do not reveal any hydrodynamic <strong>in</strong>stabilities which are consistent with<br />

these distortions, rais<strong>in</strong>g the question <strong>of</strong> whether <strong>slip</strong> has been modeled correctly.<br />

Physically, the tension <strong>in</strong> the cha<strong>in</strong>s at the surface, or equivalently, the cha<strong>in</strong> extension<br />

at the <strong>wall</strong>, should <strong>in</strong>fluence both possible mechanisms for <strong>slip</strong> <strong>and</strong> should<br />

be <strong>in</strong>cluded <strong>in</strong> the <strong>slip</strong> model. This implies that the <strong>slip</strong> velocity is a function <strong>of</strong><br />

the normal stresses, a fact <strong>in</strong>cluded <strong>in</strong> few <strong>slip</strong> models to date. As the die exit is<br />

a region <strong>of</strong> extensional flow <strong>and</strong> large normal stresses, due to velocity pr<strong>of</strong>ile rearrangement,<br />

these types <strong>of</strong> models may be particularly relevant for predict<strong>in</strong>g <strong>and</strong><br />

model<strong>in</strong>g sharksk<strong>in</strong> behavior.<br />

Normal stress-dependent <strong>slip</strong> models arise generally from simple mesoscopic <strong>and</strong><br />

microscopic treatments <strong>of</strong> <strong>polymer</strong>-surface <strong>in</strong>teraction. From a mesoscopic po<strong>in</strong>t<br />

<strong>of</strong> view, the <strong>in</strong>teractions between the <strong>wall</strong> <strong>and</strong> the <strong>polymer</strong> can be modeled as<br />

network junction po<strong>in</strong>ts, giv<strong>in</strong>g a <strong>slip</strong> velocity dependent on the number <strong>of</strong> network<br />

str<strong>and</strong>s connected to the surface. Successful network theories for bulk constitutive<br />

behavior assume that the lifetime <strong>of</strong> str<strong>and</strong>s depends upon the extension <strong>of</strong> the<br />

str<strong>and</strong>s, typically measured by the trace <strong>of</strong> the extra stress tensor.<br />

Application<br />

<strong>of</strong> this assumption to the surface-tethered str<strong>and</strong>s yields a simple, normal stressdependent<br />

<strong>slip</strong> model.<br />

Microscopic theories for <strong>polymer</strong> cha<strong>in</strong>s <strong>in</strong>teract<strong>in</strong>g with<br />

solid surfaces also lead to normal stress dependent <strong>slip</strong> models.<br />

Coarse-gra<strong>in</strong>ed<br />

bead-spr<strong>in</strong>g theories beg<strong>in</strong> with a force balance on the beads which expresses the<br />

equality between the drag force on the beads due to flow <strong>and</strong> the restor<strong>in</strong>g force<br />

due to the spr<strong>in</strong>gs.<br />

Generally, the drag is assumed isotropic, but <strong>in</strong> reality, the


xiii<br />

drag is highly anisotropic because the cha<strong>in</strong>s are distorted <strong>and</strong> oriented by flow.<br />

Inclusion <strong>of</strong> anisotropic drag leads directly to a normal stress-dependent <strong>slip</strong> model.<br />

These models, while simple <strong>in</strong> formulation, underscore the generality <strong>of</strong> normal<br />

stress- dependent <strong>slip</strong> models <strong>and</strong> the simplicity <strong>of</strong> <strong>in</strong>corporat<strong>in</strong>g the <strong>effects</strong> <strong>of</strong> cha<strong>in</strong><br />

orientation <strong>and</strong> stretch<strong>in</strong>g <strong>in</strong>to current methodologies for model<strong>in</strong>g <strong>slip</strong>.<br />

The <strong>slip</strong> models mentioned above can be generalized to <strong>in</strong>clude pressure <strong>effects</strong><br />

<strong>and</strong> analyzed formally <strong>in</strong> planar <strong>shear</strong> <strong>flows</strong> when the bulk behavior is described by<br />

the upper convected Maxwell (UCM) constitutive equation. The flow is unstable<br />

to short wavelength (i.e.<br />

large wavenumber) disturbances at large Weissenberg<br />

numbers. The perturbations are localized near the surfaces <strong>and</strong> are convected along<br />

the channel at the steady state <strong>slip</strong> velocity.<br />

Growth rates are bounded for all<br />

wavenumbers, so that the model is well-posed, if the <strong>slip</strong> model does not <strong>in</strong>clude<br />

pressure <strong>effects</strong>.<br />

The model becomes ill-posed if pressure <strong>effects</strong> are considered,<br />

lead<strong>in</strong>g to unbounded growth rates for large wavenumbers, consistent with previous<br />

analytical work by other authors. The flow is unconditionally stable if the normal<br />

stress <strong>and</strong> pressure dependencies are removed so that the <strong>slip</strong> velocity only depends<br />

on the <strong>shear</strong> stress.<br />

Exam<strong>in</strong>ation <strong>of</strong> the network <strong>and</strong> anisotropic drag models mentioned above <strong>in</strong><br />

plane Couette flow for general wavenumbers reveals several other <strong>in</strong>terest<strong>in</strong>g facets<br />

<strong>of</strong> the <strong>in</strong>stability.<br />

Stability results for both l<strong>in</strong>ear (UCM) <strong>and</strong> nonl<strong>in</strong>ear (Phan-<br />

Thien – Tanner) viscoelastic constitutive equations collapse to one master curve<br />

relat<strong>in</strong>g the critical recoverable <strong>shear</strong>, def<strong>in</strong>ed as the critical <strong>shear</strong> stress divided<br />

by the <strong>shear</strong> modulus, to the disturbance wavenumber.<br />

The critical recoverable<br />

<strong>shear</strong> is O(10), which is the same order <strong>of</strong> magnitude measured experimentally.


xiv<br />

The frequency <strong>of</strong> distortion scales with the bulk <strong>polymer</strong> relaxation time, a f<strong>in</strong>d<strong>in</strong>g<br />

corroborated by several experiments <strong>and</strong> once argued to conclusively prove that the<br />

mechanism for <strong>slip</strong> was disentanglement. These results suggest, however, that this<br />

scal<strong>in</strong>g is common to both mechanisms for <strong>slip</strong> <strong>and</strong> the processes <strong>of</strong> reentanglement<br />

<strong>and</strong> readsorption are both dom<strong>in</strong>ated by <strong>polymer</strong> relaxation. Scal<strong>in</strong>g comparisons<br />

with experiment demonstrate that the predicted temperature <strong>and</strong> molecular weight<br />

dependence are <strong>in</strong> agreement with experimental results for LLDPE <strong>and</strong> HDPE.<br />

These results are encourag<strong>in</strong>g, <strong>and</strong> suggest that normal stresses play an important<br />

role <strong>in</strong> sharksk<strong>in</strong> formation.<br />

Flow <strong>in</strong>stabilities due to surface <strong>in</strong>teractions may also be present <strong>in</strong> the flow <strong>of</strong><br />

entangled <strong>polymer</strong> solutions as enhanced concentration fluctuations. Experiments<br />

have only recently demonstrated the <strong>in</strong>terfacial nature <strong>of</strong> this phenomenon – fluctuations<br />

were visually observed to <strong>in</strong>itiate near the surfaces <strong>and</strong> chemically modify<strong>in</strong>g<br />

the surfaces to <strong>in</strong>crease <strong>slip</strong> delayed the onset <strong>of</strong> enhancement. Prior to this experiment,<br />

all explanations for <strong>shear</strong> enhanced concentration fluctuations focused<br />

on the coupl<strong>in</strong>g between <strong>polymer</strong> stress <strong>and</strong> concentration <strong>in</strong> the bulk <strong>of</strong> the solution.<br />

Analyses <strong>of</strong> <strong>polymer</strong> solution models which employ these ideas show that the<br />

diffusion <strong>of</strong> r<strong>and</strong>om fluctuations is retarded <strong>in</strong> certa<strong>in</strong> directions giv<strong>in</strong>g rise to enhanced<br />

fluctuations <strong>in</strong> those directions. These theories are successful <strong>in</strong> describ<strong>in</strong>g<br />

low <strong>shear</strong> rate behavior <strong>in</strong> scatter<strong>in</strong>g experiments which exam<strong>in</strong>e the bulk region,<br />

but are unable to describe the critical <strong>shear</strong> rates observed <strong>in</strong> experiments which<br />

sample the near-surface regions.<br />

Stability analysis us<strong>in</strong>g a Navier <strong>slip</strong> <strong>boundary</strong> condition has revealed that <strong>slip</strong><br />

can couple to the stress <strong>and</strong> concentration, giv<strong>in</strong>g rise to flow <strong>in</strong>stability.<br />

The


xv<br />

<strong>in</strong>stability is localized near the bound<strong>in</strong>g surfaces <strong>and</strong> has length scales consistent<br />

with the experimentally observed values. The characteristic length scale is √ D tr λ,<br />

where D tr is the translational diffusivity <strong>of</strong> the <strong>polymer</strong> cha<strong>in</strong> <strong>and</strong> λ is the relaxation<br />

time, <strong>and</strong> represents a crossover size from short wavelength fluctuations whose decay<br />

is dom<strong>in</strong>ated by diffusion to longer wavelength fluctuations whose decay is <strong>in</strong>fluenced<br />

by <strong>polymer</strong> cha<strong>in</strong> relaxation.<br />

For low values <strong>of</strong> the wavenumber, <strong>in</strong>creas<strong>in</strong>g <strong>slip</strong><br />

results <strong>in</strong> flow stabilization, <strong>in</strong> agreement with experiment. The orientation <strong>of</strong> the<br />

destabiliz<strong>in</strong>g perturbation is similar to the enhanced r<strong>and</strong>om fluctuations that arise<br />

<strong>in</strong> the bulk as mentioned above, <strong>and</strong> therefore, the mechanism <strong>of</strong> <strong>in</strong>stability is clearly<br />

related to the bulk mechanism <strong>of</strong> enhancement.<br />

Surfaces can enhance fluctuations even when the basic flow is stable. Time <strong>in</strong>tegration<br />

<strong>of</strong> the l<strong>in</strong>earized equations with r<strong>and</strong>om forc<strong>in</strong>g revealed the formation <strong>of</strong><br />

<strong>boundary</strong> layers <strong>in</strong> the <strong>polymer</strong> concentration pr<strong>of</strong>ile, even though fluctuations were<br />

imposed uniformly throughout the doma<strong>in</strong>. These <strong>boundary</strong> layers form on length<br />

scales similar to those <strong>in</strong> experiments, <strong>and</strong> are most easily quantified <strong>in</strong> terms <strong>of</strong> the<br />

spatial correlation function for the concentration, which is related to the structure<br />

factor <strong>and</strong>, therefore, conta<strong>in</strong>s scatter<strong>in</strong>g <strong>in</strong>formation. Exam<strong>in</strong>ation <strong>of</strong> the correlation<br />

functions suggests that r<strong>and</strong>om fluctuations are selectively, <strong>and</strong> dramatically,<br />

enhanced near the surface, even at equilibrium. The scatter<strong>in</strong>g <strong>in</strong>tensity near the<br />

surface <strong>in</strong>creases as the <strong>shear</strong> rate <strong>in</strong>creases. The enhanced fluctuations have a preferred<br />

orientation at low <strong>shear</strong> rates <strong>and</strong> appear to rotate clockwise as the <strong>shear</strong> rate<br />

<strong>in</strong>creases. This prediction is similar to those made for bulk behavior <strong>and</strong> observed<br />

experimentally <strong>in</strong> the bulk region. The preferred orientations are confirmed by plott<strong>in</strong>g<br />

representative concentration pr<strong>of</strong>iles. Interest<strong>in</strong>gly, the enhancement <strong>of</strong> r<strong>and</strong>om


xvi<br />

fluctuations is <strong>in</strong>sensitive to the presence <strong>of</strong> <strong>slip</strong> – neither the magnitude nor the<br />

orientation appears to be affected – so that any complete explanation <strong>of</strong> enhanced<br />

concentration fluctuations must <strong>in</strong>clude <strong>boundary</strong> <strong>effects</strong>. Overall, these analyses<br />

po<strong>in</strong>t out the importance <strong>of</strong> <strong>polymer</strong> – surface <strong>in</strong>teractions <strong>in</strong> underst<strong>and</strong><strong>in</strong>g the<br />

macroscopic flow behavior <strong>of</strong> <strong>polymer</strong> melts <strong>and</strong> solutions.


1<br />

Chapter 1<br />

Extrusion Instabilities<br />

1.1 Introduction<br />

Flow <strong>in</strong>stabilities <strong>and</strong> product distortions limit the productivity <strong>of</strong> many <strong>polymer</strong><br />

process<strong>in</strong>g applications, such as extrusion, film blow<strong>in</strong>g, <strong>and</strong> fiber sp<strong>in</strong>n<strong>in</strong>g. The<br />

product becomes deformed upon pass<strong>in</strong>g a critical flow rate, thereby hamper<strong>in</strong>g its<br />

usefulness <strong>and</strong>/or marketability. Hence, both <strong>in</strong>dustry <strong>and</strong> academia have expended<br />

a great deal <strong>of</strong> effort attempt<strong>in</strong>g to underst<strong>and</strong> the root causes <strong>of</strong> these <strong>in</strong>stabilities.<br />

Still, no general consensus has been reached on explanations for many <strong>of</strong> them.<br />

The general sequence <strong>of</strong> <strong>in</strong>stabilities for high molecular weight, l<strong>in</strong>ear <strong>polymer</strong>s<br />

<strong>in</strong> extrusion is outl<strong>in</strong>ed below. For these <strong>polymer</strong>s, sharksk<strong>in</strong> is the first distortion<br />

typically seen <strong>and</strong> a picture <strong>of</strong> this phenomenon for a high density polyethylene<br />

(HDPE) is shown <strong>in</strong> Fig. 1(a). It is a short wavelength, wavy distortion conf<strong>in</strong>ed<br />

to the outside edge <strong>of</strong> the extrudate.<br />

The depth <strong>of</strong> severe sharksk<strong>in</strong> is at most<br />

about 10% <strong>of</strong> the thickness <strong>of</strong> the product. Increas<strong>in</strong>g the flow rate further results


2<br />

(a)<br />

(b)<br />

(c)<br />

Figure 1: General sequence <strong>of</strong> extrusion <strong>in</strong>stabilities for l<strong>in</strong>ear <strong>polymer</strong>s.<br />

From Polymer Process<strong>in</strong>g, Pr<strong>in</strong>ciples <strong>and</strong> Model<strong>in</strong>g by Agassant, Avenas, Sergent, <strong>and</strong> Carreau [2]<br />

<strong>in</strong> more severe distortion, either stick-<strong>slip</strong> or bamboo melt fracture, depend<strong>in</strong>g on<br />

the experimental setup. Bamboo melt fracture, Fig. 1(b), has alternat<strong>in</strong>g regions <strong>of</strong><br />

smooth, non-distorted extrudate followed by regions with distortions that resemble<br />

sharksk<strong>in</strong>, <strong>and</strong> arises if the volumetric flow rate, or <strong>shear</strong> rate, is the flow variable<br />

controlled. If, <strong>in</strong>stead, the pressure drop, <strong>and</strong> hence, the <strong>shear</strong> stress is controlled,<br />

there is a dramatic <strong>in</strong>crease <strong>of</strong> the flow rate at a critical <strong>shear</strong> stress. At higher<br />

<strong>shear</strong> rates the extrudate becomes grossly distorted, as shown <strong>in</strong> Fig. 1(c). This is a<br />

large wavelength, large amplitude distortion, <strong>and</strong> the entire extrudate cross section


3<br />

is deformed. However, the surface is smooth <strong>and</strong> sharksk<strong>in</strong> free. Further <strong>in</strong>creases<br />

<strong>in</strong> <strong>shear</strong> rate result <strong>in</strong> more severe gross fracture.<br />

This sequence <strong>of</strong> <strong>in</strong>stabilities is not unique. Highly branched systems, such as<br />

polystyrene (PS), polypropylene (PP), <strong>and</strong> low density polyethylene (LDPE) do<br />

not exhibit the sharksk<strong>in</strong> <strong>in</strong>stability at all. Instead a spiral or corkscrew <strong>in</strong>stability<br />

sets <strong>in</strong> after a critical flow rate. There is a second critical flow rate after which the<br />

extrudate becomes grossly distorted.<br />

Numerous attempts have been made, both theoretically <strong>and</strong> experimentally, to<br />

underst<strong>and</strong> these <strong>in</strong>stabilities, generally with mixed results. This body <strong>of</strong> work is<br />

reviewed below with emphasis on sharksk<strong>in</strong>. As this is typically the first <strong>in</strong>stability<br />

seen <strong>in</strong> extrusion operations, an underst<strong>and</strong><strong>in</strong>g <strong>of</strong> this phenomenon is crucial for<br />

improv<strong>in</strong>g <strong>polymer</strong> extrusion processes.<br />

1.2 Sharksk<strong>in</strong><br />

A partial list<strong>in</strong>g <strong>of</strong> experimental work on sharksk<strong>in</strong> is shown <strong>in</strong> Table 1. By far the<br />

most extensive body <strong>of</strong> work exists for l<strong>in</strong>ear low-density polyethylene (LLDPE)<br />

res<strong>in</strong>s, but other res<strong>in</strong>s, such as high density polyethylene (HDPE), polybutadiene<br />

(PB), polydimethylsiloxane (PDMS), <strong>and</strong> poly(methylmethacrylate) (PMMA)<br />

exhibit sharksk<strong>in</strong> distortions as well.<br />

Sharksk<strong>in</strong> is first observed as a loss <strong>of</strong><br />

gloss <strong>of</strong> the extrudate surface above a critical <strong>shear</strong> stress or <strong>shear</strong> rate. the distortions<br />

become more pronounced as the flow rate is <strong>in</strong>creased, until the surface is<br />

clearly distorted <strong>and</strong> wavy. The loss <strong>of</strong> gloss can usually be associated with a slope<br />

change on a plot <strong>of</strong> the apparent <strong>shear</strong> stress, τ a , versus the apparent <strong>shear</strong> rate,


4<br />

Author(s) Polymer Die Proposed Mechanism<br />

Tordella [94] PMMA Capillary<br />

Sieglaff [89] PVC Capillary Cohesive failure<br />

Ajji et al. [4] LLDPE Capillary Sporadicloss <strong>of</strong> adhesion<br />

Piau et al. [71] LLDPE Capillary Entrance <strong>in</strong>stabilities<br />

HDPE Capillary ”<br />

Kurtz [47, 49] LLDPE Capillary Pre-stress<strong>in</strong>g <strong>and</strong> critical exit velocity<br />

Ramamurthy [75] LLDPE Capillary Slip <strong>in</strong> the die l<strong>and</strong><br />

HDPE Capillary ”<br />

HDPE-LDPE Capillary ”<br />

Kalika <strong>and</strong> Denn [42] LLDPE Capillary Wall <strong>slip</strong><br />

Lim <strong>and</strong> Schowalter [52] PB Slit<br />

Moynihan et al. [64] LLDPE Capillary Pre-stress<strong>in</strong>g <strong>and</strong> critical exit velocity<br />

LLDPE Slit ”<br />

Piau et al. [72] PDMS Orifice High stresses at the die exit<br />

Wang et al. [100] LLDPE Capillary Coil-stretch transition at the exit<br />

El Kissi et al. [45] PDMS Capillary Rupture at the die exit<br />

PB Capillary ”<br />

LLDPE Capillary ”<br />

HDPE Capillary ”<br />

Venet <strong>and</strong> Vergnes [96] LLDPE Capillary<br />

Person <strong>and</strong> Denn [67] LLDPE Capillary Die l<strong>and</strong> <strong>slip</strong><br />

Ghanta et al. [26] LLDPE Capillary<br />

Barone <strong>and</strong> Wang [7] PB Slit Coil-stretch transition at the die exit<br />

Table 1: Summary <strong>of</strong> experimental results for melt fracture <strong>and</strong> sharksk<strong>in</strong>.<br />

1 S -sharksk<strong>in</strong>;MF - melt fracture.


5<br />

˙γ a [75, 42, 47, 100] for polyethylenes. For other <strong>polymer</strong>s, notably PVC [89], this is<br />

not always the case. Fig. 2 shows Ramamurthy’s flow curve for a LLDPE res<strong>in</strong> <strong>and</strong><br />

Figure 2: Flow curve for LLDPE.<br />

From Ramamurthy [75]<br />

the po<strong>in</strong>t marked OSMF <strong>in</strong>dicates the critical <strong>shear</strong> stress for the onset <strong>of</strong> sharksk<strong>in</strong><br />

behavior. Below this critical <strong>shear</strong> stress the extrudate is smooth. At this po<strong>in</strong>t,<br />

the flow curve is still monotonically <strong>in</strong>creas<strong>in</strong>g, <strong>and</strong> the flow curve does not flatten<br />

out until much larger <strong>shear</strong> rates <strong>and</strong> stresses are obta<strong>in</strong>ed. It should also be noted<br />

that this phenomenon occurs at very small Reynolds numbers, due to the very high<br />

viscosities <strong>of</strong> the <strong>polymer</strong> res<strong>in</strong>s.<br />

One <strong>of</strong> the most strik<strong>in</strong>g <strong>and</strong> suggestive features <strong>of</strong> sharksk<strong>in</strong> is the importance<br />

<strong>of</strong> the <strong>polymer</strong> melt/die <strong>wall</strong> <strong>in</strong>terface.<br />

Ramamurthy [75] was the first to show<br />

that the material <strong>of</strong> construction <strong>of</strong> the die <strong>wall</strong> affected the onset <strong>of</strong> sharksk<strong>in</strong>.<br />

He constructed dies out <strong>of</strong> many different metals, <strong>in</strong>clud<strong>in</strong>g steels, brasses, copper,<br />

<strong>and</strong> chromium, <strong>and</strong> then extruded LLDPE through these dies. Fig. 3 shows Ramamurthy’s<br />

results for a 1 MILLDPE res<strong>in</strong>. The important quantity is τ c1 ,whichis


6<br />

the critical <strong>shear</strong> stress for the onset <strong>of</strong> sharksk<strong>in</strong>. This quantity varies from metal<br />

to metal, from a low <strong>of</strong> 0.104 MPa for beryllium copper, to a high <strong>of</strong> 0.172 MPa for<br />

Figure 3: Critical stresses for LLDPE for dies constructed <strong>of</strong> different metals.<br />

From Ramamurthy [75]<br />

CDA 360 brass. Ramamurthy also found that the die composition had a pr<strong>of</strong>ound<br />

effect <strong>in</strong> film blow<strong>in</strong>g operations, where the use <strong>of</strong> brass dies entirely removed sharksk<strong>in</strong>,<br />

which appeared dur<strong>in</strong>g film blow<strong>in</strong>g with steel dies. This <strong>in</strong>consistency was the<br />

subject <strong>of</strong> recent work on extrusion through brass dies by Person <strong>and</strong> Denn [67] <strong>and</strong><br />

Ghanta et al. [26], which shows clearly the importance <strong>of</strong> surface preparation with<br />

brass dies, particularly the removal <strong>of</strong> the oxide layer which forms at the surface<br />

<strong>of</strong> the die. Removal <strong>of</strong> this layer <strong>in</strong> situ us<strong>in</strong>g an additive resulted <strong>in</strong> suppression<br />

<strong>of</strong> sharksk<strong>in</strong> distortions. Low surface energy surfaces have been studied by Piau et<br />

al. [71] for LLDPE <strong>and</strong> HDPE. Dies made from poly(tetrafluoroethylene) (PTFE)<br />

were found to greatly <strong>in</strong>crease the flow rates possible before sharksk<strong>in</strong> began over<br />

those <strong>in</strong> dies made <strong>of</strong> sta<strong>in</strong>less steel.<br />

Plac<strong>in</strong>g coat<strong>in</strong>gs on the die surface affects the appearance <strong>of</strong> melt fracture.<br />

Moynihan et al. [64] performed experiments where LLDPE was extruded through


7<br />

a slit die which could be selectively coated with fluoro<strong>polymer</strong> <strong>in</strong> the entrance <strong>and</strong><br />

exit regions. When the die was uncoated, sharksk<strong>in</strong> was evident at an apparent<br />

<strong>shear</strong> rate <strong>of</strong> 27 s −1 .<br />

Coat<strong>in</strong>g the entrance region or the exit region suppressed<br />

sharksk<strong>in</strong> up to the maximum flow rate they could obta<strong>in</strong>; coat<strong>in</strong>g both was not<br />

required to suppress the <strong>in</strong>stability. However, the coat<strong>in</strong>g at the entrance <strong>of</strong> the die<br />

may have detached <strong>and</strong> been transported to the end <strong>of</strong> the die. Such a mechanism<br />

was suggested by the authors <strong>and</strong> also reported by Wang et al. [100] <strong>in</strong> their experiments,<br />

<strong>in</strong> which they reported similar stabiliz<strong>in</strong>g <strong>effects</strong> when their dies were<br />

coated with a fluoro<strong>polymer</strong> known to discourage surface <strong>in</strong>teractions.<br />

Peel<strong>in</strong>g experiments have also demonstrated the importance <strong>of</strong> surface <strong>in</strong>teractions.<br />

Hill, Hasegawa, <strong>and</strong> Denn [38] performed peel<strong>in</strong>g tests for solid LLDPE from<br />

a variety <strong>of</strong> metal surfaces. They measured the propagation speed for the crack as<br />

a function <strong>of</strong> the force used to peel the sample. In addition to f<strong>in</strong>d<strong>in</strong>g differences<br />

<strong>in</strong> the forces <strong>and</strong> velocities for various surfaces, they also found that the forces <strong>and</strong><br />

velocities changed with ag<strong>in</strong>g time. From these results, <strong>and</strong> some theoretical analysis,<br />

they were able to estimate the critical stress for the onset <strong>of</strong> sharksk<strong>in</strong>. These<br />

values compare favorably to those found by Ramamurthy [75] for the same metals.<br />

Other process<strong>in</strong>g conditions, such as the temperature <strong>and</strong> pressure, as well as<br />

the molecular weight (MW) affect the onset <strong>and</strong> severity <strong>of</strong> sharksk<strong>in</strong> deformations.<br />

Dur<strong>in</strong>g sharksk<strong>in</strong>, the pressure <strong>in</strong> the die <strong>and</strong> the reservoir rema<strong>in</strong>s relatively constant,<br />

unlike the situation for bamboo distortion, dur<strong>in</strong>g which both the pressure<br />

<strong>and</strong> the mass flow rate oscillate. Increas<strong>in</strong>g the L/D ratio <strong>of</strong> the die decreases the<br />

severity <strong>of</strong> sharksk<strong>in</strong> [95, 34] as well as <strong>in</strong>creases the critical <strong>shear</strong> stress necessary


8<br />

for the <strong>in</strong>stability [42]. Kurtz [47, 49] reported that <strong>in</strong>creased process<strong>in</strong>g temperature<br />

had little effect on the critical <strong>shear</strong> stresses for sharksk<strong>in</strong> for his LLDPE<br />

samples. Of course, a higher <strong>shear</strong> rate was necessary to obta<strong>in</strong> the critical <strong>shear</strong><br />

stress at higher temperatures due to lower melt viscosities. Venet <strong>and</strong> Vergnes [96]<br />

also reported an <strong>in</strong>crease <strong>in</strong> the critical <strong>shear</strong> rate with <strong>in</strong>creas<strong>in</strong>g temperature, but<br />

did not report critical <strong>shear</strong> stresses. Kurtz [48] showed that <strong>in</strong>creas<strong>in</strong>g the molecular<br />

weight <strong>in</strong>creases the amplitude <strong>of</strong> sharksk<strong>in</strong> <strong>and</strong> lowers the frequency <strong>of</strong> the<br />

distortion <strong>in</strong> LLDPE. There is also evidence that the molecular weight distribution<br />

may have an <strong>in</strong>fluence on the flow phenomenon. Ajji et al. [4] partially fractionated<br />

their LLDPE samples to remove the lowest molecular weight components.<br />

This<br />

fractionation <strong>in</strong>creased the critical <strong>shear</strong> rates for <strong>in</strong>stability without significantly<br />

<strong>in</strong>creas<strong>in</strong>g the number or weight average molecular weights. It is unclear whether<br />

the viscosity was affected, preclud<strong>in</strong>g statements about the critical stresses for the<br />

<strong>in</strong>stability.<br />

A few general conclusions can be drawn from the experimental evidence. Sharksk<strong>in</strong><br />

is a short wavelength, wavy distortion conf<strong>in</strong>ed to the surface <strong>of</strong> the extrudate<br />

<strong>and</strong> observed at low Reynolds numbers. This <strong>in</strong>stability occurs on the <strong>in</strong>creas<strong>in</strong>g<br />

portion <strong>of</strong> the flow curve, always before the flatten<strong>in</strong>g <strong>of</strong> the flow curve associated<br />

with bamboo, or spurt, distortions. Most importantly, sharksk<strong>in</strong> is clearly <strong>in</strong>fluenced<br />

by the <strong>in</strong>terfacial <strong>in</strong>teractions between the <strong>polymer</strong> melt <strong>and</strong> the die <strong>wall</strong>, as<br />

adhesion promoters <strong>and</strong> coat<strong>in</strong>gs which reduce the <strong>polymer</strong>/die <strong>in</strong>teractions have<br />

been shown to suppress the onset <strong>of</strong> the <strong>in</strong>stability. In addition, experiments have<br />

clearly established that the surface <strong>in</strong>teractions which lead to sharksk<strong>in</strong> take place<br />

primarily near the die exit. This is not surpris<strong>in</strong>g s<strong>in</strong>ce the highest stresses, both


9<br />

<strong>shear</strong> <strong>and</strong> normal, are at the exit <strong>and</strong> the flow possesses a strong elongational component<br />

there. There are few explanations that can account for these observations,<br />

<strong>and</strong> these generally <strong>in</strong>voke the notion <strong>of</strong> <strong>wall</strong> <strong>slip</strong>.<br />

1.3 Wall Slip<br />

1.3.1 Experimental Evidence<br />

Wall <strong>slip</strong> is one <strong>of</strong> the few explanations for sharksk<strong>in</strong> that truly depends upon the<br />

<strong>in</strong>terfacial conditions <strong>in</strong> the die.<br />

The idea was proposed as far back as 1931 by<br />

Mooney [62], who also devised a way <strong>of</strong> calculat<strong>in</strong>g <strong>slip</strong> velocities from capillary<br />

experiments us<strong>in</strong>g different sized dies.<br />

The critical <strong>shear</strong> stress for the onset <strong>of</strong><br />

sharksk<strong>in</strong> has generally been observed to co<strong>in</strong>cide with or occur at a slightly higher<br />

<strong>shear</strong> stress [75, 95] than the critical <strong>shear</strong> stress for the onset <strong>of</strong> <strong>slip</strong>.<br />

Mackay<br />

<strong>and</strong> Henson [53] argued that the critical <strong>shear</strong> stress for <strong>slip</strong> is an experimental<br />

artifact because capillary measurements are not sensitive enough to measure <strong>slip</strong> at<br />

lower stress levels. There have also been arguments that the <strong>slip</strong> only takes place<br />

at the die exit [47, 49]. Either way, <strong>slip</strong> relieves some <strong>of</strong> the <strong>shear</strong> stress <strong>in</strong> the<br />

<strong>polymer</strong>, result<strong>in</strong>g <strong>in</strong> what appears as a lower viscosity. Therefore, <strong>slip</strong> would show<br />

up as <strong>in</strong>creased <strong>shear</strong> th<strong>in</strong>n<strong>in</strong>g <strong>of</strong> the sample, result<strong>in</strong>g <strong>in</strong> a slope change <strong>in</strong> the flow<br />

curve. Other explanations, such as cavitation, rupture at the die exit, constitutive<br />

<strong>in</strong>stabilities, <strong>and</strong> apparent <strong>slip</strong> generally do not depend upon the <strong>in</strong>terface conditions<br />

<strong>and</strong> cannot describe sharksk<strong>in</strong>.<br />

Slip measurement has typically fallen <strong>in</strong>to two broad categories, flow visualization<br />

<strong>and</strong> <strong>in</strong>direct measurement. Table 2 summarizes the experiments which have


10<br />

Author(s) Polymer Apparatus Technique<br />

Atwood <strong>and</strong> Schowalter [6] HDPE Slit die Hot film anemometry<br />

Lim <strong>and</strong> Schowalter [52] PB Slit die Hot film anemometry<br />

White et. al. [102] PBR Biconical rheometer<br />

Hatzikiriakos <strong>and</strong> Dealy [33] HDPE Slid<strong>in</strong>g plate rheometer Mooney analysis<br />

Hatzikiriakos <strong>and</strong> Dealy [34] HDPE Capillary rheometer Modified Mooney analysis<br />

Migler, et al. [58] PDMS Simple Shear Evanescent waves<br />

Archer et al. [5] PS Simple Shear Visualization <strong>of</strong> immersed<br />

spheres<br />

Wang et al. [100] LLDPE Capillary rheometer Moody analysis<br />

Person <strong>and</strong> Denn [67] LLDPE Slit die Moody analysis<br />

Table 2: Summary <strong>of</strong> measurements <strong>of</strong> <strong>slip</strong> velocities.<br />

been performed <strong>in</strong> both areas.<br />

These techniques have a variety <strong>of</strong> limitations<br />

<strong>and</strong> problems. The traditional Mooney technique fails for capillary dies if either<br />

the viscosity or the <strong>slip</strong> velocity depends upon the pressure or if there are significant<br />

temperature gradients. Hatzikiriakos <strong>and</strong> Dealy [34] have proposed a modified<br />

technique that corrects for the pressure <strong>effects</strong>. Immersed sphere visualization [5],<br />

while be<strong>in</strong>g able to measure <strong>slip</strong> velocities with<strong>in</strong> microns <strong>of</strong> the plate surface, is<br />

limited by the requirement <strong>of</strong> a transparent flow cell as well as the distortion to the<br />

flow field from the f<strong>in</strong>ite sized spheres. Evanescent wave techniques [58] also require<br />

a transparent flow cell, however, the technique is non-<strong>in</strong>vasive <strong>and</strong> <strong>slip</strong> velocities<br />

can be measured as close as 100 nm from the plate surface. Hot film anemometry<br />

has the dist<strong>in</strong>ct advantage <strong>of</strong> be<strong>in</strong>g a real time measur<strong>in</strong>g system for <strong>slip</strong>, as well as<br />

be<strong>in</strong>g non-<strong>in</strong>vasive <strong>and</strong> not requir<strong>in</strong>g any special properties <strong>of</strong> the flow cell <strong>and</strong> fluid.<br />

The probes themselves are usually brittle <strong>and</strong> have a difficult time withst<strong>and</strong><strong>in</strong>g the


11<br />

high stresses at the <strong>wall</strong>, <strong>and</strong> the technique also requires complicated data reduction<br />

to back out the <strong>slip</strong> velocities, as they are not measured directly. Other flow<br />

visualization techniques, such as NMR <strong>and</strong> X-ray visualization <strong>and</strong> laser-Doppler<br />

velocimetry (which are not listed <strong>in</strong> the table) are still quite limited by poor resolution<br />

near the <strong>wall</strong> <strong>and</strong> are not yet viable for measur<strong>in</strong>g <strong>slip</strong> velocities.<br />

Several experiments clearly show that the <strong>slip</strong> velocity <strong>and</strong> the critical <strong>shear</strong><br />

stress for <strong>slip</strong> depend upon the <strong>polymer</strong>/die <strong>wall</strong> <strong>in</strong>terface. Hatzikiriakos <strong>and</strong> Dealy<br />

[33] have shown that fluorocarbon coat<strong>in</strong>gs can both suppress <strong>and</strong> promote <strong>slip</strong> <strong>in</strong><br />

their slid<strong>in</strong>g plate rheometer. Fig. 4, taken from their paper, shows that the fluo-<br />

Figure 4: Effect <strong>of</strong> DFL coat<strong>in</strong>g on the flow curve <strong>of</strong> HDPE.<br />

rocarbon coat<strong>in</strong>g Dry Film Lube (DFL) reduces the <strong>slip</strong> velocity without affect<strong>in</strong>g<br />

the critical stress for the onset <strong>of</strong> <strong>slip</strong>, while a different fluorocarbon, Dynamar,<br />

illustrated <strong>in</strong> Fig. 5 promotes <strong>slip</strong> by reduc<strong>in</strong>g the critical <strong>shear</strong> stress necessary for<br />

<strong>slip</strong>. Both have been used <strong>in</strong>dustrially to suppress sharksk<strong>in</strong> distortion. White et al.


12<br />

Figure 5: Effect <strong>of</strong> Dynamar coat<strong>in</strong>g on the flow curve <strong>of</strong> HDPE.<br />

[102] measured <strong>slip</strong> velocities is a biconical rheometer <strong>in</strong> which rotors constructed<br />

<strong>of</strong> different metals were <strong>in</strong>serted. They found that <strong>slip</strong> velocities were smaller for<br />

rotors constructed <strong>of</strong> copper or brass than for steel rotors.<br />

Rotors coated with<br />

poly(tetrafluoroethylene) exhibited the highest <strong>slip</strong> velocities. This order<strong>in</strong>g is consistent<br />

with the observations <strong>of</strong> Ramamurthy [75]. Ramamurthy [76] argued that<br />

<strong>in</strong>creased adhesion was responsible for the removal <strong>of</strong> sharksk<strong>in</strong> <strong>in</strong> film blow<strong>in</strong>g<br />

when brass dies were used. The observation by Halley <strong>and</strong> Mackay [30] that the<br />

exit pressure is higher for flow through brass dies than for flow through steel dies is<br />

consistent with the idea that the <strong>slip</strong> velocity is lower on brass that steel. Halley <strong>and</strong><br />

Mackay analyzed the die surface after extrusion <strong>and</strong> reported dez<strong>in</strong>cification <strong>of</strong> the<br />

brass <strong>and</strong> the formation <strong>of</strong> a pitted, porous surface. This surface roughness would<br />

be expected to h<strong>in</strong>der <strong>slip</strong> at the surface [84]. However, Ghanta et al. [26] found<br />

that, with careful preparation <strong>of</strong> the brass surface, LLDPE actually <strong>slip</strong>s more on


13<br />

brass than on steel <strong>and</strong> that sharksk<strong>in</strong> is entirely elim<strong>in</strong>ated. More work is required<br />

to reconcile these experiments.<br />

Several experiments have been done to determ<strong>in</strong>e the <strong>slip</strong> behavior as a function<br />

<strong>of</strong> the pressure. These experiments generally agree <strong>and</strong> demonstrate that the <strong>slip</strong><br />

velocity decreases as the pressure <strong>in</strong>creases. White et al. [102] performed experiments<br />

<strong>in</strong> a biconical rheometer <strong>in</strong> which the ambient pressure could be arbitrarily<br />

set. They found that decreas<strong>in</strong>g the pressure caused an <strong>in</strong>crease <strong>in</strong> the <strong>slip</strong> velocity<br />

<strong>and</strong> that at high pressures <strong>slip</strong> was negligible. They tested a variety <strong>of</strong> elastomeric<br />

compounds with different coat<strong>in</strong>gs <strong>and</strong> metals for the rotors <strong>in</strong> the rheometer <strong>and</strong><br />

found the same general trends. This is the same general trend as the sharksk<strong>in</strong><br />

dependence on pressure.<br />

To summarize, experiments strongly suggest <strong>wall</strong> <strong>slip</strong> as a participant <strong>in</strong> the<br />

sharksk<strong>in</strong> phenomenon. Slip has been shown to exist, <strong>and</strong> it occurs at the same<br />

or slightly lower critical stress as does sharksk<strong>in</strong>. Slip clearly depends on the die<br />

<strong>wall</strong>/<strong>polymer</strong> melt <strong>in</strong>terface <strong>and</strong> has temperature <strong>and</strong> pressure behavior similar to<br />

sharksk<strong>in</strong>. It should be possible to take the microscopic mechanisms <strong>of</strong> <strong>slip</strong> <strong>and</strong><br />

develop macroscopic models which can be used to simulate die flow <strong>and</strong> analyzed to<br />

determ<strong>in</strong>e stability. Until recently, such formalisms have been lack<strong>in</strong>g, but ad hoc<br />

<strong>slip</strong> models have been proposed <strong>and</strong> analyzed. The predictions <strong>of</strong> these models are<br />

described below.<br />

1.3.2 Plane Shear Flow Analyses with Slip<br />

Most <strong>of</strong> the theoretical evidence is <strong>of</strong> the negative variety, as very few analyses have<br />

actually demonstrated any hydrodynamic <strong>in</strong>stability which could lead to sharksk<strong>in</strong>


14<br />

<strong>and</strong> melt fracture. Attempts have been made with various constitutive equations,<br />

a variety <strong>of</strong> <strong>boundary</strong> conditions, <strong>and</strong> even compressibility, but direct theoretical<br />

evidence <strong>of</strong> sharksk<strong>in</strong> is lack<strong>in</strong>g.<br />

Although the study <strong>of</strong> <strong>slip</strong> <strong>boundary</strong> conditions was motivated here by the observation<br />

that <strong>slip</strong> may be important <strong>in</strong> sharksk<strong>in</strong> formation, <strong>in</strong>itial work was motivated<br />

by the lack <strong>of</strong> hydrodynamic <strong>in</strong>stabilities for low Reynolds number flow with<br />

no-<strong>slip</strong> boundaries. These analyses are summarized <strong>in</strong> Table 3.<br />

The first results<br />

<strong>of</strong> note are those <strong>of</strong> Gorodtsov <strong>and</strong> Leonov [27], who studied the plane Couette flow<br />

<strong>of</strong> an upper convected Maxwell fluid us<strong>in</strong>g a l<strong>in</strong>ear stability analysis with no-<strong>slip</strong><br />

<strong>boundary</strong> conditions. They found analytically that the flow is always l<strong>in</strong>early stable<br />

<strong>in</strong> the absence <strong>of</strong> <strong>in</strong>ertia (Re = 0). This result was corroborated by Renardy <strong>and</strong><br />

Renardy [82], who also showed that the flow was stable when the Reynolds number<br />

was small. The flow only becomes unstable at large Re, which is unrealistic for<br />

<strong>polymer</strong> extrusion operations. This result is further supported by the work <strong>of</strong> Lee<br />

<strong>and</strong> F<strong>in</strong>layson [50]. Several authors have also studied pressure driven flow between<br />

parallel plates, as opposed to the plate driven flow above. Plane Poiseuille flow was<br />

first studied by Porteous <strong>and</strong> Denn [74] for an upper convected Maxwell fluid <strong>and</strong> no<strong>slip</strong><br />

<strong>boundary</strong> conditions. They found that at low Reynolds numbers the flow could<br />

become unstable. However, their analysis failed to account for the existence <strong>of</strong> spurious<br />

eigenvalues, an oversight later corrected by Ho <strong>and</strong> Denn [40]. These authors<br />

showed that plane Poiseuille flow was stable at zero <strong>and</strong> low Reynolds numbers <strong>and</strong><br />

the flow became unstable only at large Reynolds numbers. Lee <strong>and</strong> F<strong>in</strong>layson [50]<br />

later confirmed this result us<strong>in</strong>g a different numerical technique. F<strong>in</strong>ally, Lim <strong>and</strong><br />

Schowalter [51] have shown that the plane Poiseuille flow <strong>of</strong> a Giesekus liquid is also


15<br />

Author(s) Model Geometry A/C B.C. Result<br />

Gorodtsov <strong>and</strong> Leonov [27] UCM Plane Couette A No-<strong>slip</strong> Stability when Re =0.<br />

Tlapa <strong>and</strong> Bernste<strong>in</strong> [93] UCM Plane Poiseuille C No-<strong>slip</strong> Instability at large Re.<br />

Ho <strong>and</strong> Denn [40] UCM Plane Poiseuille C No-<strong>slip</strong> Stability at low Re.<br />

Lee <strong>and</strong> F<strong>in</strong>layson [50] UCM Plane Poiseuille C No-<strong>slip</strong> Flow is stable at low Re.<br />

Plane Couette C No-<strong>slip</strong> Flow is stable at low Re.<br />

Renardy <strong>and</strong> Renardy [82] UCM Plane Couette C/A No-<strong>slip</strong> Stability at zero <strong>and</strong> low Re.<br />

Lim <strong>and</strong> Schowalter [51] Giesekus Plane Poiseuille C No-<strong>slip</strong> Stability at low Re.<br />

Renardy [77] UCM Plane Couette A Memory <strong>slip</strong> Instability when Re = 0 <strong>and</strong> <strong>slip</strong><br />

is happen<strong>in</strong>g.<br />

Georgiou [25] UCM Plane Poiseuille A/C Non-l<strong>in</strong>ear <strong>slip</strong> Instability when the slope <strong>of</strong> the<br />

flow curve is negative.<br />

Shore et al. [87] UCM Plane Couette C Dynamic<strong>slip</strong> Non-monotonicflow curve.<br />

Black <strong>and</strong> Graham [13] UCM Plane Couette A K<strong>in</strong>etic <strong>slip</strong> Instability when Re =0aftera<br />

critical flow rate.<br />

Black <strong>and</strong> Graham [29] UCM,PTT Planar <strong>shear</strong> A/C Normal stress dependent <strong>slip</strong> Normal stress dependence required<br />

for <strong>in</strong>stability.<br />

Table 3: Summary <strong>of</strong> stability analyses <strong>of</strong> flow. A - analytical results, C - computational results.


16<br />

stable to short waves at low Reynolds numbers. These results, <strong>and</strong> the results for<br />

plane Couette flow, show clearly that there are no elastic <strong>in</strong>stabilities present when<br />

a reasonable constitutive equation <strong>and</strong> no-<strong>slip</strong> <strong>boundary</strong> conditions are used.<br />

The lack <strong>of</strong> low Reynolds number <strong>in</strong>stabilities has motivated a number <strong>of</strong> studies<br />

employ<strong>in</strong>g a variety <strong>of</strong> <strong>of</strong> <strong>slip</strong> <strong>boundary</strong> conditions. Pearson <strong>and</strong> Petrie [66] were<br />

among the first to analyze <strong>flows</strong> with <strong>slip</strong>, <strong>and</strong> they studied several algebraic <strong>slip</strong><br />

models <strong>in</strong> which the <strong>slip</strong> velocity depends upon the <strong>shear</strong> stress.<br />

They showed<br />

that the flow could be unstable only if the <strong>slip</strong> curve, which is a plot <strong>of</strong> the <strong>slip</strong><br />

velocity versus <strong>shear</strong> stress, was nonmonotonic, i.e.<br />

<strong>of</strong> the curve.<br />

du s<br />

dτ yx<br />

< 0, for some region<br />

Essentially, this is equivalent to a nonmonotonic flow curve, with<br />

the nonl<strong>in</strong>earity that causes the flow curve to be multi-valued transferred to the<br />

<strong>boundary</strong>. Renardy [77] used one <strong>of</strong> the models proposed by Pearson <strong>and</strong> Petrie,<br />

namely the memory <strong>slip</strong> model<br />

Du s<br />

Dt + λ su s = f(τ yx ), (1)<br />

where u s is the <strong>slip</strong> velocity, τ yx is the <strong>shear</strong> stress, <strong>and</strong> λ s is the relaxation time for<br />

<strong>slip</strong>. For this dynamic <strong>slip</strong> model, the <strong>slip</strong> velocity depends upon the <strong>shear</strong> stress<br />

<strong>and</strong> the <strong>shear</strong> stress history, through the <strong>in</strong>clusion <strong>of</strong> a convected derivative term,<br />

<strong>and</strong> the <strong>slip</strong> velocity is a monotonically <strong>in</strong>creas<strong>in</strong>g function <strong>of</strong> the <strong>shear</strong> stress. The<br />

upper convected Maxwell model was used to describe the bulk behavior, <strong>and</strong> <strong>slip</strong><br />

was assumed to be happen<strong>in</strong>g. He found analytically that with Re =0theflowwas<br />

unstable to short waves. However, the growth rate is proportional to the square<br />

root <strong>of</strong> the wavenumber, imply<strong>in</strong>g that the the growth rate is <strong>in</strong>f<strong>in</strong>ite for <strong>in</strong>f<strong>in</strong>itesimal<br />

waves. Therefore, this <strong>in</strong>stability is a Hadamard-type <strong>in</strong>stability result<strong>in</strong>g from


17<br />

ill-posedness <strong>of</strong> the problem at the <strong>boundary</strong>.<br />

Also, the model can predict a f<strong>in</strong>ite<br />

<strong>slip</strong> velocity even when the <strong>shear</strong> stress is zero. These unphysical properties<br />

br<strong>in</strong>g the model validity <strong>in</strong>to question [13]. Georgiou [25] used a <strong>slip</strong> model explicitly<br />

constructed to predict a nonmonotonic region <strong>in</strong> the <strong>slip</strong> curve to study the<br />

<strong>in</strong>compressible plane Couette flow <strong>of</strong> an Oldroyd-B fluid. He performed both an<br />

analytical stability analysis <strong>and</strong> numerical simulations <strong>of</strong> the flow. The results do<br />

predict <strong>in</strong>stability, but only when the slope <strong>of</strong> the <strong>slip</strong> curve is negative. F<strong>in</strong>ally,<br />

Shore et al. [87, 88] proposed a rather complex <strong>slip</strong> model which <strong>in</strong>corporates a first<br />

order phase transition near the <strong>wall</strong>. Aga<strong>in</strong>, the model only depends upon <strong>shear</strong><br />

stresses <strong>and</strong> <strong>in</strong>stability only occurs on the decreas<strong>in</strong>g branch <strong>of</strong> the <strong>slip</strong> curve with<br />

the most unstable wavenumber be<strong>in</strong>g zero. Both <strong>of</strong> the latter two analyses essentially<br />

confirm the results <strong>of</strong> Pearson <strong>and</strong> Petrie [66]. The obvious conclusion from<br />

these analyses is that <strong>shear</strong> stress dependent <strong>slip</strong> models cannot predict <strong>in</strong>stabilities<br />

consistent with experimental observations <strong>of</strong> sharksk<strong>in</strong>. The mechanisms for <strong>slip</strong><br />

<strong>and</strong> improved <strong>slip</strong> models which <strong>in</strong>corporate more <strong>of</strong> the essential physics <strong>of</strong> cha<strong>in</strong><br />

orientation <strong>and</strong> stretch<strong>in</strong>g are described <strong>in</strong> the follow<strong>in</strong>g chapter.


18<br />

Chapter 2<br />

Model<strong>in</strong>g <strong>of</strong> Wall Slip<br />

2.1 Slip Mechanisms<br />

Two broad explanations exist for <strong>slip</strong> at the <strong>in</strong>terface between the melt <strong>and</strong> the die,<br />

as shown <strong>in</strong> Fig. 6. First, molecules anchored at the <strong>wall</strong> can desorb from the surface,<br />

with the rate <strong>of</strong> desorption dependent upon the strength <strong>of</strong> <strong>in</strong>teraction between the<br />

die <strong>wall</strong> <strong>and</strong> the <strong>polymer</strong> melt. The second method is disentanglement between a<br />

layer <strong>of</strong> <strong>polymer</strong> molecules adsorbed to the <strong>wall</strong> <strong>and</strong> the first bulk molecular layer.<br />

As the die surface is changed, the graft<strong>in</strong>g density is changed, thereby chang<strong>in</strong>g the<br />

force <strong>and</strong> hence, the critical stress, for the disentanglement process.<br />

Several observations suggest that desorption is important at the <strong>in</strong>terface. Halley<br />

<strong>and</strong> Mackay [30] reported dez<strong>in</strong>cification <strong>of</strong> a brass die <strong>and</strong> the resultant formation<br />

<strong>of</strong> a porous metal surface dur<strong>in</strong>g the extrusion <strong>of</strong> LLDPE. These alterations<br />

to the metal surface seem<strong>in</strong>gly cannot be expla<strong>in</strong>ed by a purely entanglement/disentanglement<br />

<strong>slip</strong> mechanism. Mackay <strong>and</strong> Henson [53] argued that, based


19<br />

0000000000000000000000<br />

1111111111111111111111<br />

0000000000000000000000<br />

1111111111111111111111<br />

0000000000000000000000<br />

1111111111111111111111<br />

0000000000000000000000<br />

1111111111111111111111<br />

0000000000000000000000<br />

1111111111111111111111<br />

0000000000000000000000<br />

1111111111111111111111<br />

00000000000000000000000<br />

11111111111111111111111<br />

00000000000000000000000<br />

11111111111111111111111<br />

00000000000000000000000<br />

11111111111111111111111<br />

00000000000000000000000<br />

11111111111111111111111<br />

00000000000000000000000<br />

11111111111111111111111<br />

00000000000000000000000<br />

11111111111111111111111<br />

(a)<br />

(b)<br />

0000000000000000000000<br />

1111111111111111111111<br />

0000000000000000000000<br />

1111111111111111111111<br />

0000000000000000000000<br />

1111111111111111111111<br />

0000000000000000000000<br />

1111111111111111111111<br />

0000000000000000000000<br />

1111111111111111111111<br />

0000000000000000000000<br />

1111111111111111111111<br />

0000000000000000000000<br />

1111111111111111111111<br />

Figure 6: Schematic <strong>of</strong> the two pr<strong>in</strong>cipal mechanisms for <strong>slip</strong>. The dark l<strong>in</strong>e is a<br />

cha<strong>in</strong> adsorbed to the surface. (a) - the adsorbed cha<strong>in</strong> desorbs <strong>and</strong> slides along<br />

the surface, (b) - the adsorbed cha<strong>in</strong> disentangles from the bulk cha<strong>in</strong>s, which then<br />

move along the <strong>in</strong>terface. The <strong>shear</strong> flow is to the right.<br />

on the activation energy for desorption for PS on steel, the force hold<strong>in</strong>g the cha<strong>in</strong>s<br />

to the surface was the same order <strong>of</strong> magnitude as the drag force on the cha<strong>in</strong>s, so<br />

that some fraction <strong>of</strong> the cha<strong>in</strong>s was detached from the surface. A similar argument<br />

was put forth by Yar<strong>in</strong> <strong>and</strong> Graham [106] with regard to the experimental data <strong>of</strong><br />

Piau <strong>and</strong> El Kissi [70] for LLDPE. Yar<strong>in</strong> <strong>and</strong> Graham concluded that, based on<br />

their proposed model for <strong>slip</strong>, which <strong>in</strong>cluded competition between desorption <strong>and</strong><br />

disentanglement, desorption for LLDPE was a faster process than disentanglement<br />

<strong>and</strong> desorption can lead to nonmonotonic <strong>slip</strong> curves.<br />

Several authors, <strong>in</strong>clud<strong>in</strong>g Brochard <strong>and</strong> de Gennes [16], Ajdari et al. [3], <strong>and</strong><br />

Mhetar <strong>and</strong> Archer [57, 56], have proposed scal<strong>in</strong>g theories for disentanglement.<br />

These theories are based on the paradigm that an anchored molecule which does<br />

not <strong>in</strong>teract with other anchored molecules can be modeled as a cha<strong>in</strong> pulled by<br />

one end through the bulk material. The conformation <strong>of</strong> the pulled, or probe, cha<strong>in</strong>


20<br />

changes as the pull<strong>in</strong>g force (F ) <strong>in</strong>creases, <strong>and</strong> the friction coefficient, ζ, canbe<br />

estimated for various regimes. The result is a curve <strong>of</strong> the pull<strong>in</strong>g force, F ,versus<br />

the cha<strong>in</strong> velocity, V = F/ζ. On a macroscopic level, F is related to the <strong>shear</strong><br />

stress at the <strong>wall</strong> <strong>and</strong> V is related to the <strong>slip</strong> velocity. There are generally several<br />

low <strong>slip</strong> regimes, followed by a regime where the probe cha<strong>in</strong> disentangles from the<br />

bulk cha<strong>in</strong>s.<br />

The entanglement/disentanglement idea was orig<strong>in</strong>ally proposed as an explanation<br />

for gross melt fracture. At the transition, there is sudden disentanglement<br />

throughout the die <strong>and</strong> the <strong>boundary</strong> condition shifts from no-<strong>slip</strong> to complete <strong>slip</strong>.<br />

Dur<strong>in</strong>g <strong>slip</strong>, the stresses relax <strong>and</strong> the cha<strong>in</strong>s reentangle.<br />

Hysteresis is expected<br />

due to the resultant cha<strong>in</strong> orientation <strong>and</strong> stretch<strong>in</strong>g. Wang <strong>and</strong> coworkers [100]<br />

have recently extended the idea <strong>in</strong> an attempt to expla<strong>in</strong> both the sharksk<strong>in</strong> behavior<br />

<strong>and</strong> the slope change <strong>in</strong> the flow curve. In their hypothesis, the well-known<br />

stress s<strong>in</strong>gularity at the die exit causes the critical stress for disentanglement to be<br />

exceeded locally <strong>in</strong> the exit region before it is exceeded throughout the die. The<br />

<strong>boundary</strong> condition oscillates between the stick <strong>and</strong> <strong>slip</strong> states at the die exit. Disentanglement<br />

releases the s<strong>in</strong>gular stress, allow<strong>in</strong>g the cha<strong>in</strong>s to reattach <strong>and</strong> start<br />

the oscillation over aga<strong>in</strong>.<br />

The predictions <strong>of</strong> these theories are <strong>in</strong> agreement with some experimental evidence<br />

[58] but are not conclusive. In particular, Drda <strong>and</strong> Wang [22] <strong>and</strong> Wang<br />

<strong>and</strong> Drda [99] argued that the temperature <strong>in</strong>dependent extrapolation length that<br />

they measured for HDPE was direct evidence for the entanglement/disentanglement<br />

transition. Yar<strong>in</strong> <strong>and</strong> Graham [106] demonstrated that desorption can lead to the


21<br />

same behavior. Wang et al. [100] showed that the sharksk<strong>in</strong> oscillation period correlated<br />

directly with the <strong>polymer</strong> relaxation time <strong>and</strong> argued that this also was<br />

conclusive evidence <strong>of</strong> disentanglement.<br />

However, the dynamics <strong>of</strong> reattachment<br />

<strong>and</strong> reentanglement are both expected to be dom<strong>in</strong>ated by relaxation <strong>of</strong> the <strong>polymer</strong><br />

cha<strong>in</strong> so that this observation is <strong>in</strong>conclusive at best. A f<strong>in</strong>al limitation is that<br />

the <strong>slip</strong> velocity is always associated only with the <strong>shear</strong> stress at the <strong>wall</strong>. This<br />

is not consistent with physical <strong>in</strong>tuition regard<strong>in</strong>g <strong>slip</strong>, where the orientation <strong>and</strong><br />

stretch<strong>in</strong>g <strong>of</strong> cha<strong>in</strong>s are important.<br />

2.2 Connections Between Normal Stresses <strong>and</strong> Slip<br />

Several factors <strong>and</strong> observations motivate the <strong>in</strong>clusion <strong>of</strong> normal stresses <strong>in</strong>to <strong>slip</strong><br />

models.<br />

First, as described <strong>in</strong> §1.3.2, <strong>shear</strong> stress dependent <strong>slip</strong> models have<br />

failed to predict <strong>in</strong>stabilities consistent with sharksk<strong>in</strong>. Second, from a more mesoscopic<br />

viewpo<strong>in</strong>t, the <strong>in</strong>teractions at the <strong>wall</strong> can be modeled as junction po<strong>in</strong>ts,<br />

similar to the junction po<strong>in</strong>ts <strong>in</strong> network theories for bulk constitutive behavior.<br />

The most successful network constitutive equations, the Marrucci [1] <strong>and</strong> Phan-<br />

Thien/Tanner [69, 68] models, were derived with the assumption that the lifetime<br />

<strong>of</strong> the network str<strong>and</strong>s depends on the stress <strong>in</strong> the fluid, specifically the normal<br />

stresses through the trace <strong>of</strong> the extra stress tensor. The situation near the die <strong>wall</strong><br />

should be similar, <strong>and</strong> this was exploited by Hatzikiriakos <strong>and</strong> Kalogerakis [35],<br />

who developed a stochastic network theory. Third, the normal stresses measure the<br />

elongation <strong>of</strong> the molecules, or equivalently, the tension <strong>in</strong> the cha<strong>in</strong>s, <strong>and</strong> higher


22<br />

tension should <strong>in</strong>crease both the rate <strong>of</strong> desorption <strong>and</strong> the rate <strong>of</strong> disentanglement,<br />

so that regardless <strong>of</strong> the mechanism for <strong>slip</strong>, the higher the normal stresses,<br />

the higher the <strong>slip</strong> velocity. This idea can also be applied to the scal<strong>in</strong>g theories<br />

for <strong>polymer</strong> disentanglement. In these scal<strong>in</strong>g theories, the friction force depends<br />

upon the number <strong>and</strong> lifetime <strong>of</strong> entanglements surround<strong>in</strong>g the probe cha<strong>in</strong>. It is<br />

generally assumed that the bulk material is at equilibrium <strong>and</strong> the probe cha<strong>in</strong> is<br />

be<strong>in</strong>g pulled through this quiescent material. However, dur<strong>in</strong>g extrusion, the bulk<br />

material is stressed, <strong>and</strong> the lifetime <strong>of</strong> entanglements depends on the stress. As<br />

a possible first correction the the scal<strong>in</strong>g theories, the friction coefficient should<br />

depend upon the normal stresses to account for the stress <strong>in</strong> the bulk material.<br />

Experimental observations also suggest a l<strong>in</strong>k between normal stresses <strong>and</strong> sharksk<strong>in</strong>.<br />

It is widely accepted that the normal stresses are the root cause <strong>of</strong> die swell.<br />

It has been observed (El Kissi <strong>and</strong> Piau [44]) that the amount <strong>of</strong> die swell decreases<br />

at the onset <strong>of</strong> sharksk<strong>in</strong>, <strong>in</strong>dicat<strong>in</strong>g that the normal stresses are reduced <strong>in</strong> the<br />

fluid, possibly due to <strong>slip</strong>.<br />

Perhaps the most direct experimental evidence for a<br />

normal stress dependence is the peel<strong>in</strong>g experiments <strong>of</strong> Hill et al. [38]. Based on<br />

st<strong>and</strong>ard theories <strong>of</strong> adhesion, they derived the follow<strong>in</strong>g condition for the critical<br />

normal stress difference for the onset <strong>of</strong> <strong>slip</strong> <strong>and</strong> <strong>in</strong>stability:<br />

N 1<br />

∼<br />

2Wa = , (2)<br />

fδ r<br />

where Wa is the work <strong>of</strong> adhesion, δ r is the thickness <strong>of</strong> a proposed rubbery region<br />

near the <strong>wall</strong> (c.f. V<strong>in</strong>ogradov <strong>and</strong> Ivanova [97, 98]), <strong>and</strong> f is the fractional recovery.<br />

f is between zero <strong>and</strong> one <strong>and</strong> is used to take <strong>in</strong>to account the fact that the rubbery<br />

region is constra<strong>in</strong>ed to rema<strong>in</strong> near the <strong>wall</strong> by the bulk fluid. The work <strong>of</strong> adhesion<br />

is related to the surface free energy <strong>of</strong> the <strong>polymer</strong> <strong>and</strong> the die <strong>wall</strong>. The free energy


23<br />

<strong>of</strong> the metals used to make dies is typically much larger than the surface free energy<br />

<strong>of</strong> the <strong>polymer</strong>s, so that Wa basically equals the free energy <strong>of</strong> the die <strong>wall</strong>. Several<br />

other approximations give f =0.2 <strong>and</strong>δ = R/4, where R is the radius <strong>of</strong> the<br />

capillary. Therefore, a very simple criterion exists for <strong>in</strong>stability <strong>and</strong> <strong>slip</strong>, namely<br />

N 1,c<br />

∼ =<br />

40Wa<br />

R , (3)<br />

For comparison with experimental data, most <strong>of</strong> which is reported <strong>in</strong> terms <strong>of</strong> a<br />

critical <strong>shear</strong> stress, the normal stress difference can be replaced by the recoverable<br />

<strong>shear</strong>, s R = N 1,w /τ yx . The recoverable <strong>shear</strong> is normally order unity at the onset <strong>of</strong><br />

<strong>in</strong>stability so<br />

τ yx,c = 40Wa<br />

R . (4)<br />

This criterion compares very well to the results <strong>of</strong> Ramamurthy [75], lend<strong>in</strong>g support<br />

to the conclusion that normal stresses <strong>in</strong>fluence melt fracture. Peel<strong>in</strong>g tests<br />

performed by Ajji et al. [4] also l<strong>in</strong>k the adhesion strength to the onset <strong>of</strong> sharksk<strong>in</strong><br />

melt fracture.<br />

Experiments have shown that <strong>slip</strong> is relevant to sharksk<strong>in</strong>, although its role <strong>in</strong><br />

sharksk<strong>in</strong> formation has not been theoretically elucidated. This may be due to the<br />

fact that current <strong>slip</strong> models have neglected normal stresses <strong>in</strong> their formulations,<br />

even though experimental evidence <strong>and</strong> physical <strong>in</strong>tuition suggests that they are<br />

important. Ideally, we would like to underst<strong>and</strong> the molecular dynamics <strong>and</strong> the<br />

fluid dynamics <strong>in</strong> a real die geometry dur<strong>in</strong>g extrusion. As a start<strong>in</strong>g po<strong>in</strong>t, <strong>and</strong><br />

to highlight the generality with which normal stress <strong>effects</strong> can be <strong>in</strong>cluded <strong>in</strong> <strong>slip</strong><br />

models, we derive simple <strong>slip</strong> models which <strong>in</strong>clude more <strong>of</strong> the essential physics.<br />

The models are <strong>in</strong>troduced below <strong>and</strong> analyzed <strong>in</strong> the next chapter.


24<br />

2.3 A Slip Model Based on Network Theory<br />

The network <strong>slip</strong> law to be used here can be motivated by a few simple, physical<br />

arguments. Consider the coarse gra<strong>in</strong>ed picture <strong>of</strong> <strong>polymer</strong> molecules <strong>in</strong>teract<strong>in</strong>g<br />

with the <strong>wall</strong> shown <strong>in</strong> Figure 7. The junction po<strong>in</strong>ts can be chemical bonds, <strong>in</strong>-<br />

Figure 7: K<strong>in</strong>etic <strong>slip</strong> model. Left: All <strong>polymer</strong> molecules are <strong>in</strong>teract<strong>in</strong>g with the<br />

<strong>wall</strong>. Right: A portion <strong>of</strong> the molecules are no longer <strong>in</strong>teract<strong>in</strong>g with the <strong>wall</strong>.<br />

termolecular forces, hydrodynamic <strong>in</strong>teractions, physical adherence, or molecular<br />

entanglements, all <strong>of</strong> which are lumped together <strong>in</strong> this analysis. X is a structural<br />

parameter that will describe the state <strong>of</strong> <strong>in</strong>teraction between the <strong>polymer</strong> <strong>and</strong> the<br />

<strong>wall</strong>. At equilibrium, X = 1, <strong>and</strong> all <strong>of</strong> the available sites for <strong>polymer</strong>-<strong>wall</strong> <strong>in</strong>teraction<br />

are filled. Dur<strong>in</strong>g flow, some <strong>of</strong> the str<strong>and</strong>s are lost due to break<strong>in</strong>g <strong>of</strong> the<br />

junction po<strong>in</strong>ts with the <strong>wall</strong>, 0


25<br />

through more str<strong>and</strong>s still attached to the <strong>wall</strong> <strong>and</strong> will experience greater drag.<br />

InEq. 5,τ yx ∗ is the <strong>shear</strong> stress at the <strong>wall</strong>, ɛ ∗ is a constant, <strong>and</strong> the superscript<br />

star ( ∗ ) implies that the variables have dimension. An overall <strong>slip</strong> velocity can be<br />

obta<strong>in</strong>ed by averag<strong>in</strong>g the velocity <strong>of</strong> the free segments <strong>and</strong> the velocity <strong>of</strong> the bound<br />

segments (which is zero) to get<br />

( ) 1 − X<br />

u ∗ s = ɛ ∗ τ ∗ w . (6)<br />

X<br />

To complete the <strong>slip</strong> model, an expression for X <strong>in</strong> terms <strong>of</strong> the flow variables is<br />

needed. A k<strong>in</strong>etic expression is used to describe the evolution <strong>of</strong> X as a function <strong>of</strong><br />

time <strong>and</strong> the stress,<br />

DX<br />

Dt = 1 λ s<br />

[(1 − X) − sXF (trτ )] , (7)<br />

where λ s is the relaxation time for <strong>slip</strong>, <strong>and</strong> s is a constant. The first term on the<br />

right h<strong>and</strong> side describes the attachment k<strong>in</strong>etics (i.e. the forward reaction), <strong>and</strong><br />

hence, λ s can be <strong>in</strong>terpreted as the reciprocal <strong>of</strong> the attachment rate constant. The<br />

attachment k<strong>in</strong>etics are proportional to the fraction <strong>of</strong> <strong>polymer</strong> molecules that are<br />

detached from the <strong>wall</strong>. The second term describes the detachment k<strong>in</strong>etics. The<br />

detachment rate is assumed to depend upon the <strong>polymer</strong> orientation <strong>and</strong> stretch<strong>in</strong>g<br />

at the <strong>wall</strong>, through the function F . The detachment k<strong>in</strong>etics are proportional to<br />

the fraction <strong>of</strong> bonded <strong>polymer</strong> molecules <strong>and</strong> s, which essentially is an equilibrium<br />

constant which gives the ratio between the attachment <strong>and</strong> detachment rate<br />

constants.<br />

Two ma<strong>in</strong> ideas are espoused <strong>in</strong> this k<strong>in</strong>etic law. One is that the rearrangement<br />

at the <strong>wall</strong> is not <strong>in</strong>stantaneous, but rather, has some f<strong>in</strong>ite relaxation time. In<br />

addition, attachment <strong>and</strong> detachment do not necessarily take place at the same


26<br />

rate. The second is that the breakage k<strong>in</strong>etics depend upon the normal stresses.<br />

For simple k<strong>in</strong>etic theory models <strong>of</strong> <strong>polymer</strong>s, specifically Hookean dumbbells, the<br />

trace <strong>of</strong> the stress tensor is proportional to the mean square end-to-end distance<br />

[10], <strong>and</strong> the detachment from the <strong>wall</strong> thus depends upon the conformation <strong>of</strong> the<br />

molecule. The trace <strong>of</strong> the stress tensor can also be thought <strong>of</strong> as a measure <strong>of</strong> the<br />

elastic stress <strong>in</strong> the molecules. Higher values <strong>of</strong> tr τ imply that the molecule feels<br />

more tension <strong>in</strong> the x- <strong>and</strong>y-directions.<br />

Several other comments are <strong>in</strong> order. First, all the F functions which have been<br />

considered are cont<strong>in</strong>uous <strong>and</strong> monotonically <strong>in</strong>creas<strong>in</strong>g. Therefore, the <strong>slip</strong> velocity<br />

will be a monotonic function <strong>of</strong> the stresses. Second, the parameters ɛ, λ s ,<strong>and</strong>s<br />

will depend upon the strength <strong>and</strong> type <strong>of</strong> <strong>in</strong>teractions between the <strong>polymer</strong> <strong>and</strong> the<br />

die <strong>wall</strong>. In terms <strong>of</strong> rate constants, λ s =1/k a <strong>and</strong> s = k d /k a ,wherek a is the rate<br />

constant for the formation <strong>of</strong> str<strong>and</strong>s <strong>and</strong> k d is the rate constant for the destruction<br />

<strong>of</strong> str<strong>and</strong>s.<br />

With<strong>in</strong> the entanglement/disentanglement mechanism, the rate<br />

constants k a <strong>and</strong> k d are the <strong>in</strong>verses <strong>of</strong> the relaxation times for entanglement <strong>and</strong><br />

disentanglement, respectively, at the surface. Accord<strong>in</strong>g to Ajdari et al. [3], these<br />

are the constra<strong>in</strong>t release times, k a ∝ k d ∝ 1/τ cr ∝ 1/(τ R N 5/2 ). τ R is the Rouse<br />

relaxation time, which is proportional to N 2 <strong>and</strong> has an Arrhenius temperature<br />

dependence (because it depends upon the monomer friction coefficient). Clearly,<br />

these k<strong>in</strong>etic constants only depend upon bulk properties <strong>and</strong> are <strong>in</strong>dependent <strong>of</strong><br />

any surface energetics or characteristics. The only place for surface properties to<br />

show up is <strong>in</strong> the <strong>slip</strong> coefficient, ɛ. The <strong>slip</strong> relation itself is just a statement <strong>of</strong><br />

Stokes’ law, so that ɛ is essentially the friction coefficient. The frictional force will<br />

depend upon the number <strong>of</strong> adsorbed cha<strong>in</strong>s that the bulk cha<strong>in</strong>s must slide past,


27<br />

i.e. ɛ =1/(n e ζ), where ζ is the monomeric friction coefficient <strong>and</strong> n e is the equilibrium<br />

number density <strong>of</strong> cha<strong>in</strong>s at the surface. The situation is somewhat different<br />

when the mechanism for <strong>slip</strong> is adsorption/desorption at the <strong>wall</strong>. In this case, the<br />

k<strong>in</strong>etic constants essentially follow the Arrhenius relation with activation energies<br />

that depend upon the work <strong>of</strong> adhesion between the <strong>polymer</strong> <strong>and</strong> the surface [37].<br />

The temperature dependence is the same as for the disentanglement case <strong>and</strong> the<br />

rates <strong>of</strong> adsorption <strong>and</strong> desorption <strong>in</strong>crease strongly with temperature. The work<br />

<strong>of</strong> adhesion is a function <strong>of</strong> the surface free energies, so that the k<strong>in</strong>etics <strong>of</strong> attachment/detachment<br />

depend <strong>in</strong>tr<strong>in</strong>sically on the <strong>in</strong>teractions between the <strong>polymer</strong> <strong>and</strong><br />

the solid surface. In addition, the <strong>slip</strong> coefficient depends on the surface coverage;<br />

but now, <strong>in</strong> general, the surface coverage depends on the k<strong>in</strong>etics at the surface<br />

<strong>and</strong> is a complicated function <strong>of</strong> the stress [106]. For the network <strong>slip</strong> model, this<br />

coefficient should depend only on the equilibrium number <strong>of</strong> adsorbed cha<strong>in</strong>s, n e .<br />

Molecular simulations (Bitsanis <strong>and</strong> various coworkers [65, 11]) <strong>and</strong> molecular theories<br />

[85, 86] suggest that the number <strong>of</strong> segment-surface contacts <strong>in</strong>creases as the<br />

square root <strong>of</strong> the molecular weight. The k<strong>in</strong>etics <strong>of</strong> this <strong>slip</strong> mechanism, however,<br />

do not explicitly depend upon the molecular weight <strong>of</strong> the <strong>polymer</strong>, as the k<strong>in</strong>etic<br />

constants basically describe monomer adsorption <strong>and</strong> desorption at the surface.<br />

2.4 Anisotropic Drag Slip Model<br />

Similar functional forms for the <strong>slip</strong> velocity can be obta<strong>in</strong>ed from more detailed<br />

arguments. Yar<strong>in</strong> <strong>and</strong> Graham [106] recently proposed a <strong>slip</strong> model obta<strong>in</strong>ed from<br />

k<strong>in</strong>etic theory arguments for a <strong>polymer</strong> represented by a bead-spr<strong>in</strong>g model. The


28<br />

conformation <strong>of</strong> a molecule anchored to the <strong>wall</strong> at the orig<strong>in</strong> is given by the balance<br />

between the drag forces due to the bead motions <strong>and</strong> the restor<strong>in</strong>g spr<strong>in</strong>g force,<br />

assumed Hookean,<br />

ζ ·<br />

[<br />

u s − dR ]<br />

dt<br />

− 3kT R =0, (8)<br />

a2 where R =(X, Y ) is the end-to-end vector <strong>of</strong> the dumbbell, ζ is the friction tensor,<br />

u s =(u s , 0) is the bulk <strong>slip</strong> velocity <strong>in</strong> the x-direction, k is the Boltzmann constant,<br />

T is the absolute temperature, <strong>and</strong> a is the characteristic coil size <strong>of</strong> the molecule at<br />

equilibrium. The molecular stretch<strong>in</strong>g is assumed strong enough to neglect Brownian<br />

forces along the cha<strong>in</strong>. Yar<strong>in</strong> <strong>and</strong> Graham assume that the friction tensor is<br />

isotropic, i.e., ζ = ζ o δ,whereδ is the unit tensor. If, <strong>in</strong>stead, the friction tensor is<br />

anisotropic <strong>and</strong> given by the Giesekus formula [10], ζ −1 = 1<br />

ζ 0<br />

(δ + ρτ ), then Eq. 8<br />

can be written <strong>in</strong> component form as<br />

u s − dX dt<br />

− 3kT<br />

a 2 ζ 0<br />

[X + ρτ xx X + ρτ yx Y ]=0<br />

dY<br />

dt + 3kT<br />

a 2 ζ 0<br />

[Y + ρτ yx X + ρτ yy Y ]=0<br />

. (9)<br />

X is related to the <strong>shear</strong> stress by Hooke’s law, τ yx /n = HX,whereH(= 3kT/a 2 )<br />

is the spr<strong>in</strong>g constant <strong>and</strong> n is the number density <strong>of</strong> anchored dumbbells. Yar<strong>in</strong><br />

<strong>and</strong> Graham consider the case where the surface coverage depends upon the <strong>shear</strong><br />

stress. For simplicity, it is assumed here that anchored molecules are permanently<br />

attached so that n is a constant, Eq. 9 can be solved at steady state to give<br />

[ (1 + ρτxx )(1 + ρτ yy ) − ρ 2 τ 2 ]<br />

yx<br />

u s = ɛ<br />

τ yx . (10)<br />

(1 + ρτ yy )<br />

where ɛ =1/(ζ 0 n). If ρ = 0, the quasi-steady approximation reduces the anisotropic<br />

drag model to a Navier <strong>slip</strong> relation. Eq. 10 predicts monotonically <strong>in</strong>creas<strong>in</strong>g <strong>slip</strong>


29<br />

<strong>and</strong> flow curves. This contrasts with the full Yar<strong>in</strong>-Graham model, which predicts<br />

non-monotonic curves. Tak<strong>in</strong>g n to be constant elim<strong>in</strong>ates the competition between<br />

surface coverage <strong>and</strong> molecular stretch<strong>in</strong>g responsible for the non-monotonicity.<br />

While the approximations used here are severe, our purpose <strong>in</strong> deriv<strong>in</strong>g this model<br />

is to illustrate that <strong>slip</strong> models which conta<strong>in</strong> normal stress dependencies can be expected<br />

to arise from more detailed treatments when the relevant physics is <strong>in</strong>cluded.<br />

Eq. 10 is similar to the static Black-Graham <strong>slip</strong> model, except that the function F<br />

here depends upon all the stress components, not just the normal stresses.


30<br />

Chapter 3<br />

Stability <strong>of</strong> Plane Shear Flow <strong>of</strong> a<br />

Polymer Melt with Slip †<br />

3.1 Relevance <strong>of</strong> Viscometric Flow to Die Exit Flow<br />

Die exit flow is the problem <strong>of</strong> primary importance <strong>in</strong> underst<strong>and</strong><strong>in</strong>g the <strong>in</strong>ception <strong>of</strong><br />

sharksk<strong>in</strong> deformation dur<strong>in</strong>g extrusion, however, this problem has several <strong>in</strong>herent<br />

complications which make analysis difficult. Fig. 8 shows a schematic <strong>of</strong> a die dur<strong>in</strong>g<br />

<strong>polymer</strong> melt extrusion. The melt swells as it leaves the die due to the elasticity <strong>of</strong><br />

the fluid, <strong>and</strong> the <strong>boundary</strong> is discont<strong>in</strong>uous, transition<strong>in</strong>g from solid boundaries to<br />

free surfaces. This leads to velocity pr<strong>of</strong>ile rearrangement, with the fully developed<br />

pressure driven flow far upstream from the corner evolv<strong>in</strong>g to plug flow downstream.<br />

The <strong>boundary</strong> s<strong>in</strong>gularity also causes the development <strong>of</strong> s<strong>in</strong>gular stresses as the<br />

corner is approached. These difficulties make tackl<strong>in</strong>g the full die exit flow stability<br />

† The results presented here were published <strong>in</strong> part <strong>in</strong> references [13] <strong>and</strong> [14].


31<br />

Figure 8: Typical die exit show<strong>in</strong>g the geometric s<strong>in</strong>gularity where the free surface<br />

beg<strong>in</strong>s <strong>and</strong> die swell. The fully developed upstream flow is paraboloidal <strong>and</strong> the<br />

downstream flow is plug.<br />

problem a formidable task.<br />

Fortunately, simple viscometric <strong>flows</strong> are relevant to die exit <strong>flows</strong>. This can<br />

be seen by exam<strong>in</strong><strong>in</strong>g flow around sharp corners. Dean <strong>and</strong> Montagnon [20] <strong>and</strong><br />

M<strong>of</strong>fatt [61] derived similarity solutions for Stokes’ flow around sharp corners, <strong>in</strong>clud<strong>in</strong>g<br />

symmetric solutions which are applicable to free surfaces downstream <strong>of</strong> the<br />

corner. Fig. 9(a) shows the case <strong>of</strong> a reentrant corner <strong>and</strong> Fig. 9(b) shows a free<br />

surface downstream <strong>of</strong> the corner. In both cases, the flow is viscometric near the<br />

θ = 3π 2<br />

<br />

V<br />

<br />

<br />

<br />

V<br />

θ = 0<br />

C<br />

<br />

θ = 0<br />

<br />

V<br />

θ<br />

C<br />

B<br />

θ = α<br />

(a)<br />

(b)<br />

Figure 9: Stress <strong>boundary</strong> layers <strong>in</strong> corner <strong>flows</strong>.<br />

solid surfaces (θ = 0 <strong>in</strong> (a) <strong>and</strong> (b) <strong>and</strong> θ = 3π 2<br />

<strong>in</strong> (a)), with ψ ∼ rρ θ 2 ,whereψ is<br />

the stream function, θ is the angular coord<strong>in</strong>ate <strong>in</strong> cyl<strong>in</strong>drical coord<strong>in</strong>ates, <strong>and</strong> ρ is<br />

the power law exponent. Near the free surface, however, the stream function adopts<br />

an elongational form, ψ ∼ r ρ (α − θ), where θ = α is the position <strong>of</strong> the free surface.<br />

Renardy [79] <strong>and</strong> H<strong>in</strong>ch [39] studied the flow <strong>of</strong> a upper convected Maxwell fluid


32<br />

around a reentrant corner <strong>and</strong> found that viscometric <strong>boundary</strong> layers exist near the<br />

solid surface for both Newtonian <strong>and</strong> viscoelastic k<strong>in</strong>ematics. These <strong>boundary</strong> layers,<br />

denoted by V <strong>in</strong> Fig. 9, are also present for a nonl<strong>in</strong>ear viscoelastic constitutive<br />

equations [81], although the <strong>boundary</strong> layer thickness is much larger for the Phan-<br />

Thien–Tanner model than the UCM fluid.<br />

The presence <strong>of</strong> this region suggests<br />

that a simplified analysis us<strong>in</strong>g model geometries may be useful <strong>in</strong> underst<strong>and</strong><strong>in</strong>g<br />

the dynamics <strong>of</strong> die exit <strong>flows</strong>.<br />

Therefore, planar <strong>shear</strong> <strong>flows</strong> are exam<strong>in</strong>ed here as a first attempt to underst<strong>and</strong><br />

how normal stresses <strong>in</strong>fluence <strong>in</strong>terfacial <strong>slip</strong> <strong>and</strong> surface distortions <strong>in</strong> <strong>polymer</strong><br />

process<strong>in</strong>g. Such an analysis is useful for extract<strong>in</strong>g the underly<strong>in</strong>g molecular<br />

mechanisms for <strong>in</strong>stability <strong>and</strong> for determ<strong>in</strong><strong>in</strong>g the qualitative potential <strong>of</strong> normal<br />

stress dependent <strong>slip</strong> models to predict <strong>in</strong>stabilities consistent with sharksk<strong>in</strong>. It is<br />

possible to analyze general <strong>slip</strong> models <strong>in</strong> planar, parallel <strong>shear</strong> <strong>flows</strong> if a simple constitutive<br />

equation, namely the upper convected Maxwell equation, is used. These<br />

results are presented first <strong>and</strong>, as shown below, several predictions are consistent<br />

with sharksk<strong>in</strong> deformations. Then, the specific <strong>slip</strong> models <strong>in</strong>troduced <strong>in</strong> Ch. 2 are<br />

analyzed <strong>in</strong> order to relax some <strong>of</strong> the <strong>in</strong>herent assumptions <strong>in</strong> the general analysis.<br />

Analytical solutions are still possible <strong>of</strong> the UCM equation is used, but to exam<strong>in</strong>e<br />

more complex constitutive behavior, numerical techniques must be employed.<br />

3.2 Formulation<br />

The model problem considered here is simple <strong>shear</strong> flow between two parallel plates,<br />

shown schematically <strong>in</strong> Fig. 10. The orig<strong>in</strong> is located at the bottom plate <strong>and</strong> the


33<br />

l<br />

01<br />

01<br />

01<br />

01<br />

01<br />

y<br />

01<br />

01<br />

x<br />

01<br />

0000 1111 000 111<br />

u *(y)<br />

b * u s *<br />

γ . * t<br />

Figure 10: Basic parallel <strong>shear</strong> flow geometry with <strong>slip</strong> at the solid surfaces. A<br />

Couette velocity pr<strong>of</strong>ile is shown on the left <strong>and</strong> a Poiseuille pr<strong>of</strong>ile on the right.<br />

For short wavelength perturbations, only the <strong>shear</strong> rate near the <strong>wall</strong> is important.<br />

flow can be either a Couette flow driven by the motion <strong>of</strong> the top plate, a Poiseuille<br />

flow driven by a pressure gradient, or some l<strong>in</strong>ear comb<strong>in</strong>ation <strong>of</strong> the two. The<br />

constitutive behavior <strong>of</strong> the fluid is described generally by the Phan-Thien–Tanner<br />

(PTT) network model without nonaff<strong>in</strong>e motion [69],<br />

τ (1) + 1<br />

We n<br />

(1 + µ tr τ )τ = ˙γ, (11)<br />

where ˙γ = ∇v +(∇v) T is the stra<strong>in</strong> rate tensor, v is the velocity vector, τ is the<br />

<strong>polymer</strong> extra stress tensor, µ describes the effect <strong>of</strong> τ on the creation <strong>and</strong> destruction<br />

rates <strong>of</strong> network junctions, <strong>and</strong> the the convected derivative, τ (1) ,isgivenby<br />

τ (1) = ∂τ + v ·∇τ −{τ · ∂t (∇v)+(∇v)T · τ }. The upper convected Maxwell (UCM)<br />

equation, Eq. 11 with µ = 0, is used for the majority <strong>of</strong> the results because <strong>of</strong> its<br />

simplicity even though it does not describe the <strong>shear</strong> th<strong>in</strong>n<strong>in</strong>g behavior <strong>of</strong> real <strong>polymer</strong><br />

melts. The full PTT model, which predicts <strong>shear</strong> th<strong>in</strong>n<strong>in</strong>g <strong>in</strong> the viscosity <strong>and</strong><br />

first normal stress coefficient, but has a zero second normal stress difference, is used<br />

to gauge the effect <strong>of</strong> nonl<strong>in</strong>ear viscoelasticity. In Eq. 11, the stresses have been


34<br />

nondimensionalized by the <strong>shear</strong> modulus, G ∗ = η p /λ, whereη p is the zero-<strong>shear</strong><br />

rate viscosity <strong>and</strong> λ is the bulk <strong>polymer</strong> relaxation time, time <strong>and</strong> <strong>shear</strong> rate by the<br />

nom<strong>in</strong>al <strong>shear</strong> rate at the <strong>wall</strong>, <strong>and</strong> lengths by the gap width, l. More details on<br />

the nondimensionalization can be found <strong>in</strong> Appendix A.1. The asterisk <strong>in</strong>dicates<br />

dimensional quantities while unmarked quantities are dimensionless. Certa<strong>in</strong> variables,<br />

such as the relaxation time, λ <strong>and</strong> the viscosity, η p , always have dimension.<br />

The nom<strong>in</strong>al Weissenberg number, We n (= λ ˙γ n), ∗ is the dimensionless applied <strong>shear</strong><br />

rate at the <strong>wall</strong>. The <strong>shear</strong> rate <strong>in</strong> the fluid is less than the applied <strong>shear</strong> rate due to<br />

<strong>slip</strong>, so a second dimensionless <strong>shear</strong> rate, the true Weissenberg number, We t = λ ˙γ t<br />

∗<br />

must be def<strong>in</strong>ed, where ˙γ t ∗ is the actual <strong>shear</strong> rate <strong>in</strong> the fluid at the <strong>wall</strong>. The ratio<br />

<strong>of</strong> Weissenberg numbers appears frequently <strong>and</strong> is denoted as ˙γ = We t /We n . The<br />

flow is assumed to be <strong>in</strong>compressible <strong>and</strong> <strong>in</strong>ertialess (Re = 0), so that the equations<br />

<strong>of</strong> cont<strong>in</strong>uity <strong>and</strong> motion are, respectively,<br />

∇·v = 0 (12)<br />

∇·τ −∇p =0. (13)<br />

These equations have been nondimensionalized <strong>in</strong> the same manner as the constitutive<br />

equation (c.f. Appendix A.1).<br />

The tangential velocity <strong>boundary</strong> conditions are provided by <strong>slip</strong> models relat<strong>in</strong>g<br />

the <strong>slip</strong> velocity to the stress tensor.<br />

The <strong>slip</strong> models considered here have the<br />

general, nondimensional form<br />

u s = f(τ ,σ nn )τ yx , (14)<br />

where σ is the total stress tensor, σ = −δp + τ so that σ nn = n · σ · n is the total<br />

normal stress act<strong>in</strong>g on the <strong>wall</strong>, where n is the outward unit normal. This general


35<br />

form allows the <strong>slip</strong> velocity to depend upon molecular stretch<strong>in</strong>g <strong>and</strong> orientation,<br />

through the normal stresses, as well as the isotropic pressure <strong>in</strong> the system. The<br />

asymptotic results considered below are facilitated by rewrit<strong>in</strong>g the <strong>slip</strong> relation <strong>in</strong><br />

terms <strong>of</strong> the extrapolation length, b ∗ , def<strong>in</strong>ed <strong>in</strong> general as b ∗ = u ∗ s / ˙γ∗ t . As shown<br />

below, the steady state dimensionless <strong>shear</strong> stress at the <strong>wall</strong> is τ yx = We t /(1 +<br />

µτ xx )=We n ˙γ/(1 + µτ xx ), <strong>and</strong> therefore, b(= b ∗ /l ∗ ) is, from Eq. 14,<br />

b =<br />

¯fWe n<br />

1+µ¯τ xx<br />

, (15)<br />

where ¯f = f(¯τ , ¯σ nn ) <strong>and</strong> overbars denote steady state values.<br />

The stability <strong>of</strong> the flow is determ<strong>in</strong>ed by a l<strong>in</strong>ear stability analysis. Only stability<br />

with respect to two-dimensional disturbances is exam<strong>in</strong>ed, as Squire’s theorem<br />

holds directly for the UCM equation [93]. The basic equations are first solved for<br />

the unidirectional flow field, which for plane Couette flow is given by<br />

ū =˙γy +ū s ,<br />

¯τ yx =<br />

¯v =0,<br />

We t<br />

1+µ¯τ xx<br />

,<br />

¯τ yy =0,<br />

(16a)<br />

(16b)<br />

(16c)<br />

(16d)<br />

the pressure is arbitrary, <strong>and</strong> ¯τ xx is given by the only real solution to µ 2¯τ 3<br />

xx +2µ¯τ 2<br />

xx +<br />

¯τ xx −2We 2 t<br />

= 0. Stability is determ<strong>in</strong>ed by exam<strong>in</strong><strong>in</strong>g perturbations to the base flow.<br />

The flow variables, p – pressure, u – x-component <strong>of</strong> velocity, v – y-component <strong>of</strong><br />

velocity, τ xx – first normal stress, τ yx – <strong>shear</strong> stress, <strong>and</strong> τ yy – second normal stress,<br />

are written as a =(u, v, τ xx ,τ yx ,τ yy ,p)=ā + ã, with the perturbations given by<br />

thenormalmodeformã = â(y)e ikx(x−ct) + c.c. The overbar <strong>in</strong>dicates the base state


36<br />

values <strong>and</strong> the hat denotes the perturbation amplitudes. The basic equations are<br />

l<strong>in</strong>earized around the steady state, result<strong>in</strong>g <strong>in</strong> a generalized eigenvalue problem<br />

for the eigenvalues {c} <strong>and</strong> eigenvectors {â}. This procedure is shown <strong>in</strong> detail <strong>in</strong><br />

Appendix A.2 for planar <strong>shear</strong> flow <strong>of</strong> the PTT fluid, where the system <strong>of</strong> equations<br />

for the perturbations is derived <strong>and</strong> it is shown that this system can be reduced<br />

to one general stability equation, Eq. 89, which is analogous to the famous Orr-<br />

Sommerfeld equation from Newtonian fluid mechanics [21]. The eigenvalues are <strong>in</strong><br />

general complex, with the imag<strong>in</strong>ary part giv<strong>in</strong>g the exponential growth or decay<br />

rate <strong>of</strong> the perturbations. If an eigenvalue satisfies Im(c) > 0, perturbations grow<br />

<strong>and</strong> the flow is unstable. Analytical expressions for the eigenvalues can be obta<strong>in</strong>ed<br />

when the UCM equation (µ = 0) is used as the constitutive equation. For more<br />

complicated constitutive relations, numerical methods must be employed.<br />

3.3 Asymptotic Solutions with General Slip Models<br />

The stability analysis <strong>of</strong> planar <strong>shear</strong> flow <strong>of</strong> the UCM fluid <strong>in</strong> the asymptotic case<br />

We n ≫ 1<strong>and</strong>b ≪ 1 is discussed first. These assumptions permit exam<strong>in</strong>ation <strong>of</strong><br />

<strong>in</strong>stability without requir<strong>in</strong>g a particular model for <strong>slip</strong>. These results clarify the<br />

general roles <strong>of</strong> the elastic normal stresses <strong>and</strong> pressure <strong>in</strong> the stability <strong>of</strong> flow with<br />

<strong>slip</strong>.<br />

The approach taken here is motivated by the observation that, <strong>in</strong> the absence<br />

<strong>of</strong> <strong>slip</strong>, the growth rates <strong>of</strong> perturbations approach zero (from below) as We −1<br />

n<br />

as<br />

We n →∞[27]. So, by perturb<strong>in</strong>g 1/We n from zero, the growth rates are perturbed<br />

from zero, i.e. from marg<strong>in</strong>al stability. The small <strong>slip</strong> limit (b ≪ 1) allows the


37<br />

approximation We t = We n , elim<strong>in</strong>at<strong>in</strong>g the need to specify a particular expression<br />

for ū s . F<strong>in</strong>ally, the large wavenumber assumption implies that there are <strong>boundary</strong><br />

layers localized near the solid surface <strong>and</strong> it turns out that only the <strong>in</strong>ner, <strong>boundary</strong><br />

layer solutions need to be considered to determ<strong>in</strong>e the stability. As shown below, the<br />

condition k x ≫ 1/(bWe n ) is required for the asymptotic solutions to be uniformly<br />

valid <strong>and</strong>, therefore, the approximation is self-consistent.<br />

Several additional assumptions are made. First, f <strong>in</strong> Eq. 14 is taken to be an<br />

<strong>in</strong>creas<strong>in</strong>g function <strong>of</strong> each <strong>of</strong> the stress components, so that non-monotonicity is<br />

not <strong>in</strong>troduced <strong>in</strong>to the <strong>slip</strong> behavior or the flow curve, <strong>and</strong> <strong>in</strong>ertia is neglected. To<br />

f<strong>in</strong>d the <strong>boundary</strong> layer near the bottom plate the y-coord<strong>in</strong>ate is scaled with the<br />

wavenumber as ỹ = k x y. The base state velocity pr<strong>of</strong>ile is ū =ū s +ỹ/k x + O( 1<br />

k 2 x),<br />

the steady state <strong>slip</strong> velocity is ū s<br />

= b ˙γ = b, <strong>and</strong> the base state stresses are:<br />

¯τ xx =2We 2 n + O( 1<br />

k x<br />

), ¯τ yx = We n + O( 1<br />

k x<br />

), <strong>and</strong> ¯τ yy = 0. When the asymptotic scal<strong>in</strong>gs<br />

are applied to the general stability equation, Eq. 89, the result at lead<strong>in</strong>g order <strong>in</strong><br />

wavenumber is<br />

2 d2<br />

( ˜Q<br />

dỹ − ˜Q 2 − 2 ˜Q d d2<br />

+2)( 2 dỹ dỹ +2iWe d<br />

2 n<br />

dỹ − 1 − 2We 2 n )ˆv = 0 (17)<br />

where ˜Q =ỹ − k x˜c − i/We n . The four solutions to this equation are<br />

ˆv =(ỹ − k x˜c)e ±ỹ ,e ±(iWen+ √1+We 2 n )ỹ . (18)<br />

Only the two decay<strong>in</strong>g solutions can match the outer solution (ˆv =0)<strong>and</strong>are<br />

physically realistic.<br />

After application <strong>of</strong> the no-penetration <strong>boundary</strong> condition,<br />

ˆv = 0, the solution to Eq. 17 is<br />

ˆv = K 1 k x˜ce −(iWen+ √1+We 2 n )ỹ + k 1 (ỹ − k x˜c)e −ỹ , (19)


38<br />

where K 1 is a constant <strong>of</strong> <strong>in</strong>tegration which drops out later <strong>in</strong> the analysis. Eq. 19<br />

suggests a velocity <strong>boundary</strong> layer thickness <strong>of</strong> 1/k x , s<strong>in</strong>ce the second term on the<br />

right h<strong>and</strong> side dom<strong>in</strong>ates at large We n , <strong>and</strong> a stress <strong>boundary</strong> layer <strong>of</strong> thickness<br />

1/(k x We n ), because the first term on the right h<strong>and</strong> side dom<strong>in</strong>ates the expression<br />

for the velocity gradients, <strong>and</strong> hence the stresses, at large We n . These <strong>boundary</strong><br />

layer scal<strong>in</strong>gs are consistent with those reported by Renardy [80] <strong>and</strong> Graham [28].<br />

The solution given by Eq. 19 is used to derive expressions for the stress components,<br />

which are then substituted <strong>in</strong>to the <strong>slip</strong> model to obta<strong>in</strong> the eigenvalues. The <strong>slip</strong><br />

equation for the perturbations can be written <strong>in</strong> the l<strong>in</strong>earized form<br />

û s = bB<br />

We n<br />

(ˆτ yx + Rˆτ xx + Gˆτ yy − H ˆσ nn ), (20)<br />

where the coefficients are<br />

B =1+ ¯τyx<br />

¯f<br />

R =<br />

(<br />

∂f<br />

∂τ yx<br />

,<br />

)¯τ,¯σ nn<br />

∂f ¯τyx(<br />

∂τxx<br />

)<br />

¯f+( )¯τ,¯σnn<br />

∂f<br />

,<br />

∂τyx<br />

) ¯τ,¯σnn<br />

∂τyy<br />

) ¯τ,¯σnn<br />

¯f+( ∂f<br />

,<br />

∂τyx ¯τ,¯σnn<br />

∂f<br />

−¯τyx(<br />

∂σnn<br />

)<br />

¯f+( )¯τ,¯σnn<br />

∂f<br />

∂τyx ¯τ,¯σnn<br />

G = ¯τyx ( ∂f<br />

H =<br />

,<br />

(21)<br />

<strong>and</strong> the negative sign <strong>in</strong> front <strong>of</strong> H explicitly takes <strong>in</strong>to account the experimental observation<br />

that <strong>in</strong>creas<strong>in</strong>g pressure decreases the <strong>slip</strong> velocity(Kalika <strong>and</strong> Denn [42],<br />

Hatzikiriakos <strong>and</strong> Dealy [34], White et al. [102]). Exam<strong>in</strong>ation <strong>of</strong> the l<strong>in</strong>earized<br />

constitutive equation reveals that ˆτ yx = O(We n ), ˆτ xx = O(We 2 n ), <strong>and</strong> ˆτ yy = O(1)<br />

as We n →∞. Because ˆτ yy is relatively small, it is neglected <strong>in</strong> this analysis. The<br />

coefficients B <strong>and</strong> R are both positive, as one <strong>of</strong> the basic assumptions is that the<br />

<strong>slip</strong> curve is monotonically <strong>in</strong>creas<strong>in</strong>g.


39<br />

The first case <strong>of</strong> <strong>in</strong>terest is <strong>shear</strong> stress dependent <strong>slip</strong> case, i.e., R = H =0.<br />

To obta<strong>in</strong> large Weissenberg number expressions, B is assumed to be strictly O(1).<br />

This way, the <strong>slip</strong> velocity rema<strong>in</strong>s nonzero but bounded as We n →∞. There are<br />

three discrete eigenvalues, which to lead<strong>in</strong>g order <strong>in</strong> 1/We n are<br />

⎧ ( (<br />

1<br />

⎪⎨<br />

− i 1+ √ ))<br />

3 1<br />

2 2 k xWe n<br />

+ O( 1<br />

We<br />

( (<br />

n) 2<br />

˜c = 1<br />

− i 1 − √ ))<br />

3 1<br />

2 2 k xWe n<br />

+ O( 1 ) . (22)<br />

We 2 n<br />

⎪⎩<br />

1<br />

(1 − i) bB + O(<br />

k xWe n<br />

)<br />

Two roots scale as 1/We n ,onerootisO(1) <strong>and</strong> all <strong>of</strong> the eigenvalues are stable,<br />

i.e. Im(c) < 0. There is also a stable, cont<strong>in</strong>uous set <strong>of</strong> eigenvalues at ˜c =ỹ/k x −<br />

i/(k x We n ). In order for the second order terms to be asymptotically smaller than<br />

the first order terms, k x ≫ 1/(bWe n ). This condition holds for higher order terms<br />

as well, so that the assumption <strong>of</strong> large wavenumbers was justified. This analysis<br />

confirms the previous literature results [66] that <strong>shear</strong> stress dependent <strong>slip</strong> does<br />

not lead to <strong>in</strong>stability when the <strong>slip</strong> velocity is an <strong>in</strong>creas<strong>in</strong>g function <strong>of</strong> the <strong>wall</strong><br />

<strong>shear</strong> stress.<br />

To determ<strong>in</strong>e the effect <strong>of</strong> add<strong>in</strong>g normal stress <strong>and</strong> pressure dependencies, it is<br />

necessary to first determ<strong>in</strong>e the scal<strong>in</strong>gs necessary for the pressure <strong>and</strong> normal stress<br />

contributions to be the same order as the <strong>shear</strong> stress contribution, as this leads<br />

to the most <strong>in</strong>terest<strong>in</strong>g results. The total normal stress on the surface at y =0is<br />

σ nn = −σ yy =ˆp − ˆτ yy . An expression for the pressure perturbation can be obta<strong>in</strong>ed<br />

from the x-component <strong>of</strong> the equation <strong>of</strong> motion as<br />

ˆp =ˆτ xx − i<br />

k x<br />

dˆτ yx<br />

dy . (23)<br />

Explicitly <strong>in</strong>troduc<strong>in</strong>g the eigenvectors shows that, even though both dˆτ yx /dy <strong>and</strong>ˆτ xx<br />

are O(We 3 ), ˆp is O(We 2 ), due to cancelation, <strong>and</strong> is therefore, the same order as the


40<br />

<strong>shear</strong> stress. This suggests the scal<strong>in</strong>gs R = R 0 /We n <strong>and</strong> H = O(1), where R 0 is<br />

O(1). S<strong>in</strong>ce ˆτ yy ≪ ˆp, Eq. 20 can be written as<br />

u s = bB [<br />

ˆτ yx + R (<br />

0<br />

ˆτ xx − H ˆτ xx − i )]<br />

∂ˆτ yx<br />

. (24)<br />

We n We n k x ∂y<br />

Four discrete eigenvalues are found this time, with three <strong>of</strong> them be<strong>in</strong>g small <strong>in</strong><br />

magnitude (O(1/(k x We n ))) <strong>and</strong> one be<strong>in</strong>g large (O(1)).<br />

If H = 0, we f<strong>in</strong>d that<br />

normal stress dependent <strong>slip</strong> leads to well-posed <strong>in</strong>stability. In this case, the large<br />

root is found to be stable <strong>and</strong> one <strong>of</strong> the small roots becomes unstable at R 0 =<br />

0.1826 [13]. This criterion is <strong>in</strong>dependent <strong>of</strong> both k x <strong>and</strong> b. Thus, rewrit<strong>in</strong>g <strong>in</strong><br />

terms <strong>of</strong> the unscaled coefficients, <strong>in</strong>stability is predicted to occur if<br />

We n > 0.1826<br />

R , (25)<br />

for small R.<br />

The growth rate for the disturbances, k x Im(˜c) is bounded for all<br />

wavenumbers, so that the model is well-posed for this case, <strong>and</strong> the real part <strong>of</strong> ˜c<br />

for this root is O(1/We n ), so that the wave speed, Re(c) =ū s + Re(˜c), is essentially<br />

equal to the steady state <strong>slip</strong> velocity <strong>and</strong> the <strong>in</strong>stability is convected along with the<br />

base flow. As <strong>in</strong> the previous case, k x ≫ 1/(bWe n ). The dimensionless frequency,<br />

def<strong>in</strong>ed as ω = k x Re(c) =k x b+k x Re(˜c)isO(k x b), so that the dimensional frequency,<br />

ω ∗ = ˙γ nω ∗ ≃ k x b ˙γ n ∗ is proportional to the wavenumber <strong>and</strong> the critical <strong>shear</strong> rate<br />

at high We n . A picture <strong>of</strong> the destabiliz<strong>in</strong>g disturbance is shown <strong>in</strong> Fig. 11. The<br />

stream function <strong>and</strong> normal stress perturbations shown would be superimposed on<br />

the base state flow. The streaml<strong>in</strong>es show pairs <strong>of</strong> counter-rotat<strong>in</strong>g vortices, with<br />

the dotted l<strong>in</strong>es denot<strong>in</strong>g clockwise rotation <strong>and</strong> the solid l<strong>in</strong>es counterclockwise<br />

rotation. The tops <strong>of</strong> the vortices are cut <strong>of</strong>f on the graph <strong>in</strong> order to illustrate<br />

the stress <strong>boundary</strong> layer, which is much th<strong>in</strong>ner than the velocity <strong>boundary</strong> layer.


41<br />

1<br />

0.8<br />

y <br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 2 4 6 8 10 12<br />

x <br />

Figure 11: Snapshot <strong>of</strong> the destabiliz<strong>in</strong>g disturbance at the onset <strong>of</strong> <strong>in</strong>stability,<br />

show<strong>in</strong>g perturbation streaml<strong>in</strong>es overlaid on a density plot <strong>of</strong> the first normal<br />

stress perturbation <strong>in</strong> a coord<strong>in</strong>ate system mov<strong>in</strong>g with the wavespeed (≈ ū s ). The<br />

nom<strong>in</strong>al Weissenberg number is 20. The dashed l<strong>in</strong>es denote clockwise rotation <strong>and</strong><br />

the solid l<strong>in</strong>es counterclockwise rotation. The white regions <strong>in</strong> the density plot are<br />

regions <strong>of</strong> large, positive ˆτ xx <strong>and</strong> the black regions are regions <strong>of</strong> large, negative<br />

stress. The arrow denotes the base flow direction.<br />

These observations comprise the most important result <strong>of</strong> the analysis: the normal<br />

stress dependence <strong>of</strong> the <strong>slip</strong> velocity is necessary for <strong>in</strong>stability, <strong>and</strong> for B>0<strong>and</strong><br />

any value <strong>of</strong> R>0.1826/We n there is a f<strong>in</strong>ite We n above which the flow is unstable.<br />

If R 0 = 0 but H>0, then a result similar to the Coulomb friction result <strong>of</strong><br />

Renardy [78] is found, as the large root, which has growth rate proportional to<br />

k x , becomes unstable at H = 1. All three <strong>of</strong> the small roots are stable, <strong>and</strong>, <strong>in</strong><br />

addition, one <strong>of</strong> them collapses to the end <strong>of</strong> the cont<strong>in</strong>uous spectrum, which is<br />

given by the same relation, ˜c = ỹ/k x − i/(k x We n ), as before.<br />

In this case, the<br />

<strong>in</strong>stability is the result <strong>of</strong> ill-posedness at the <strong>boundary</strong>. Typically, the pressure<br />

dependence is assumed to follow an exponential form, i.e., f ∝ e −β∗ p ∗ , where the<br />

pressure coefficient is very small (β ∗ ≈ 10 −3 MPa −1 for LLDPE [67]). The stability


42<br />

H<br />

4 <br />

<br />

<br />

<br />

<br />

<br />

Unbounded growth rate<br />

3 <br />

<br />

<br />

<br />

<br />

<br />

2 <br />

<br />

<br />

<br />

<br />

<br />

1 <br />

<br />

<br />

<br />

Bounded growth rate<br />

<br />

<br />

0 <br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

R 0<br />

Figure 12: Stability diagram for pressure- <strong>and</strong> normal stress- dependent <strong>slip</strong>. The<br />

solid curve is the neutral curve for the small (bounded growth rate) root <strong>and</strong> the<br />

correspond<strong>in</strong>g unstable region <strong>in</strong> parameter space is <strong>in</strong>dicated by the forward hatch<strong>in</strong>g.<br />

The dashed curve is the neutral curve for the large (unbounded growth rate)<br />

root with its unstable region highlighted by the backward hatch<strong>in</strong>g. Both roots are<br />

unstable <strong>in</strong> the cross-hatched region.<br />

criterion for this model is We n,c =1/β ≫ 1, where β = β ∗ G ≈ 10 −4 for LLDPE. We<br />

therefore consider it highly unlikely that <strong>in</strong>teractions between elasticity <strong>and</strong> pressure<br />

dependent <strong>slip</strong> could be responsible for sharksk<strong>in</strong>, as this results <strong>in</strong> extremely large<br />

critical Weissenberg numbers for <strong>in</strong>stability.<br />

Much more complicated dynamics are found if H <strong>and</strong> R 0 are both nonzero. From<br />

the limit<strong>in</strong>g cases described above, <strong>in</strong>creas<strong>in</strong>g the pressure coefficient, H, stabilizes<br />

the small root <strong>and</strong> destabilizes the large root, while <strong>in</strong>creas<strong>in</strong>g the normal stress<br />

coefficient, R 0 , has exactly the opposite effect. This leads to regions <strong>in</strong> parameter<br />

space where both roots are stable, only one is unstable, or both are unstable, as<br />

shown <strong>in</strong> Fig. 12. Generally, for small values <strong>of</strong> the pressure coefficient, the small<br />

root is unstable at large enough We n <strong>and</strong> the growth rate is <strong>in</strong>dependent <strong>of</strong> k x , while<br />

the large root is unstable for larger values <strong>of</strong> the pressure coefficient at sufficiently


43<br />

large We n , <strong>and</strong> the growth rate is O(k x ). There is also an overlap region where both<br />

roots are unstable.<br />

This analysis leads to several stability predictions. These results clearly show<br />

that a <strong>slip</strong> model where the <strong>slip</strong> velocity depends solely on <strong>shear</strong> stresses will not<br />

lead to hydrodynamic <strong>in</strong>stabilities.<br />

However, an arbitrarily small, but nonzero,<br />

normal stress <strong>slip</strong> dependence does lead to short wave <strong>in</strong>stability if We n is large<br />

enough. In an extrusion die, velocity pr<strong>of</strong>ile rearrangement at the die exit leads to<br />

an extensional component <strong>of</strong> the velocity pr<strong>of</strong>ile there <strong>and</strong> large normal stresses <strong>in</strong><br />

the exit region. The <strong>in</strong>stability predicted here should then be manifest first <strong>in</strong> the<br />

exit region, which may expla<strong>in</strong> why many <strong>in</strong>vestigators have attributed sharksk<strong>in</strong><br />

to an exit effect.<br />

In fact, similar behavior is expected at an <strong>in</strong>terior <strong>boundary</strong><br />

s<strong>in</strong>gularity.<br />

Barone <strong>and</strong> Wang [7] coated the downstream half <strong>of</strong> a slit die with<br />

a fluoro<strong>polymer</strong> to prevent adhesion <strong>and</strong> promote <strong>slip</strong> <strong>and</strong> left the upstream half<br />

uncoated. They found evidence <strong>of</strong> <strong>in</strong>stability near the change <strong>in</strong> surface condition<br />

which decayed downstream.<br />

The <strong>boundary</strong> s<strong>in</strong>gularity results <strong>in</strong> an extensional<br />

velocity component which <strong>in</strong>creases the stresses, particularly the normal stresses,<br />

<strong>and</strong> triggers the <strong>in</strong>stability. Downstream from the s<strong>in</strong>gularity, where fluoro<strong>polymer</strong><br />

prevents adhesion <strong>and</strong> the <strong>slip</strong> velocity is therefore proportional to the <strong>shear</strong> stress,<br />

the flow is stable <strong>and</strong> the <strong>in</strong>stability decays.<br />

Although these results do not depend on the specific form <strong>of</strong> the <strong>slip</strong> model,<br />

there are a number <strong>of</strong> limitations to the analysis. To address these limitations, it is<br />

necessary to perform analytical <strong>and</strong> numerical analyses <strong>in</strong>corporat<strong>in</strong>g specific <strong>slip</strong><br />

models, nonl<strong>in</strong>ear viscoelasticity, <strong>and</strong> f<strong>in</strong>ite wavelength perturbations.


44<br />

3.4 Results with Specific Slip Models<br />

3.4.1 Analytical Results for the UCM Equation<br />

The analytical results discussed below are specialized to the UCM equation <strong>in</strong> plane<br />

Couette flow, for which the general stability equation simplifies to Eq. 90.<br />

The<br />

solutions to this equation were first given by Gorodtsov <strong>and</strong> Leonov [27] <strong>and</strong> used to<br />

determ<strong>in</strong>e the eigenvalues for no-<strong>slip</strong> boundaries. Here, the network <strong>and</strong> anisotropic<br />

drag <strong>slip</strong> models <strong>in</strong>troduced <strong>in</strong> Ch. 2 are employed <strong>in</strong> order to relax the assumptions<br />

<strong>in</strong>herent <strong>in</strong> the asymptotic results presented above.<br />

The network model with F =trτ is, <strong>in</strong> dimensionless form,<br />

( ) 1 − X<br />

u s = ɛ τ yx<br />

X<br />

(26)<br />

DX<br />

Dt = 1 [(1 − X) − s tr τ ]<br />

We s<br />

(27)<br />

where We s = λ s ˙γ ∗ n. Black [12] performed an asymptotic analysis for k x ≫ 1, ɛ ≪ 1,<br />

<strong>and</strong> We n ≫ 1 <strong>and</strong> computed the critical nom<strong>in</strong>al Weissenberg number as a function<br />

<strong>of</strong> We s <strong>and</strong> s, the results <strong>of</strong> which are reproduced <strong>in</strong> Fig. 13. Clearly, the critical<br />

values are <strong>in</strong>sensitive to the time dependence <strong>of</strong> the structural rearrangement at the<br />

<strong>wall</strong> unless the k<strong>in</strong>etics are very slow. As the rate <strong>of</strong> <strong>in</strong>terchange <strong>of</strong> material at the<br />

surface slows, the flow becomes more stable, s<strong>in</strong>ce, as We s becomes large, DX<br />

Dt<br />

→ 0,<br />

X → constant, <strong>and</strong> u s ∝ τ yx , which is a stable case, as shown previously. The<br />

rema<strong>in</strong>der <strong>of</strong> the calculations are presented for We s = 0, which simplifies the <strong>slip</strong><br />

model to u s = ɛs(tr τ )τ yx .<br />

Analytical results for the network model with the UCM equation, Eq. 11 with<br />

µ = 0, as the constitutive equation are considered first. Only the spatially decay<strong>in</strong>g


45<br />

We n,c<br />

100<br />

We s<br />

= 0.1<br />

We s<br />

= 1<br />

We s<br />

= 10<br />

We s<br />

= 100<br />

Unstable<br />

10<br />

Stable<br />

10 −9 10 −8 10 −7 10 −6 10 −5 10 −4 10 −3 10 −2 10 −1 10 0 10 1 10 2<br />

Figure 13: Critical Weissenberg number as a function <strong>of</strong> s for several values <strong>of</strong> We s<br />

for the network <strong>slip</strong> model.<br />

s<br />

solutions to Eq. 90 are reta<strong>in</strong>ed <strong>in</strong> the derivation <strong>of</strong> the eigenvalues, <strong>in</strong> effect ignor<strong>in</strong>g<br />

the presence <strong>of</strong> the upper <strong>boundary</strong>. This is a very good approximation for large k x ,<br />

due to the localized <strong>boundary</strong> layers. Application <strong>of</strong> the <strong>boundary</strong> conditions results<br />

<strong>in</strong> a fourth order polynomial for the eigenvalues which can be solved numerically.<br />

Fig. 14 shows neutral curves for plane Couette flow, plotted as the critical applied<br />

104<br />

We n,c<br />

103<br />

102<br />

Increas<strong>in</strong>g wavelength<br />

Unstable<br />

101<br />

Stable<br />

10 -13 10 -11 10 -9 10 -7 10 -5 10 -3<br />

Figure 14: Critical We n as a function <strong>of</strong> ɛs for the UCM equation <strong>and</strong> the network<br />

<strong>slip</strong> model, keep<strong>in</strong>g only the decay<strong>in</strong>g eigenfunctions. ( )-k x =10 5 ,<br />

(— ——)-k x = 100, <strong>and</strong> (– ––––)-k x =1.<br />

εs


46<br />

flow rate for <strong>in</strong>stability, We n,c , versus the <strong>slip</strong> coefficient, ɛs. As the limit <strong>of</strong> no<strong>slip</strong><br />

boundaries is approached, ɛs → 0, the critical Weissenberg number approaches<br />

<strong>in</strong>f<strong>in</strong>ity as We n,c ∼ (ɛs) −1/4 . At high enough values <strong>of</strong> ɛs, the flow aga<strong>in</strong> becomes<br />

unconditionally stable, approach<strong>in</strong>g the s<strong>in</strong>gularity as We n,c ∼ (2.06 × 10 −4 − ɛs) −1 .<br />

When ɛs 2.06 × 10 −4 , a high enough stress to cause the <strong>in</strong>stability cannot be<br />

generated <strong>in</strong> the fluid, due to the large <strong>slip</strong> velocities. The different l<strong>in</strong>es <strong>in</strong> each<br />

graph also show that the flow is more stable for longer wavelength disturbances.<br />

In fact, the most unstable disturbance is that correspond<strong>in</strong>g to k x →∞,<strong>and</strong>this<br />

disturbance also has the highest growth rate, as shown <strong>in</strong> Fig. 15. However, the<br />

0.0002<br />

Im(k x<br />

c)<br />

0.0001<br />

0<br />

−0.0001<br />

10 2 10 3 10 4 10 5 10 6<br />

k x<br />

Figure 15: Growth rate versus wavenumber for the UCM equation <strong>and</strong> the network<br />

<strong>slip</strong> model with the follow<strong>in</strong>g parameters: We n =11.62, We t = 11, <strong>and</strong> ɛs =10 −5 .<br />

growth rate rema<strong>in</strong>s bounded as k x →∞; therefore, the model is mathematically<br />

well-posed. A nonzero, but bounded growth rate as k x →∞is also found <strong>in</strong> the<br />

analysis <strong>of</strong> <strong>in</strong>terfacial <strong>in</strong>stability <strong>of</strong> viscoelastic two-layer flow [83]. Fig. 16 shows<br />

a typical analytical eigenvalue spectrum. A numerical spectrum is also shown <strong>and</strong><br />

will be discussed below. There are four discrete eigenvalues <strong>and</strong> a cont<strong>in</strong>uous set<br />

along the l<strong>in</strong>e c = ˙γy +ū s − i/(k x We n ). Only four discrete eigenvalues are found<br />

analytically, even though one would expect the spectrum to be symmetric across


47<br />

0.05<br />

0.00<br />

Im(c)<br />

−0.05<br />

−0.10<br />

−0.15<br />

−0.20<br />

0 0.2 0.4 0.6 0.8 1<br />

Re(c)<br />

Figure 16: Typical eigenvalue spectra obta<strong>in</strong>ed analytically <strong>and</strong> numerically. The<br />

+ symbols <strong>and</strong> the dashed l<strong>in</strong>e <strong>in</strong>dicate the analytical eigenvalues, while the circles<br />

<strong>and</strong> the solid l<strong>in</strong>e <strong>in</strong>dicate the numerical spectrum. The parameters are k x =5,<br />

ɛs =10 −5 , We n =17.341, We t = 15, <strong>and</strong> for the numerical results, N = 256.<br />

the centerl<strong>in</strong>e <strong>of</strong> the channel, y =0.5, because only the spatially decay<strong>in</strong>g solutions<br />

to the stream function were kept <strong>in</strong> deriv<strong>in</strong>g the eigenvalues. Replac<strong>in</strong>g F =trτ<br />

by F =2τ 2<br />

yx <strong>in</strong> the <strong>slip</strong> model leads to the same steady <strong>slip</strong> behavior but never to<br />

<strong>in</strong>stability, consistent with the asymptotic result that flow with only <strong>shear</strong> stress<br />

dependent <strong>slip</strong> is stable. F<strong>in</strong>ally, Fig. 17 shows neutral curves for plane Couette<br />

flow <strong>and</strong> the anisotropic drag <strong>slip</strong> model. Comparison <strong>of</strong> these curves with those for<br />

the network model shows that the stability results are virtually identical for these<br />

two models.<br />

3.4.2 Numerical Method<br />

To gauge the effect <strong>of</strong> nonl<strong>in</strong>ear viscoelasticity, numerical solutions to the PTT<br />

equation must be obta<strong>in</strong>ed. Chebyshev collocation is used to set up a generalized<br />

matrix eigenvalue problem.<br />

Chebyshev collocation falls under a general class <strong>of</strong><br />

numerical methods for <strong>boundary</strong> value problems known as the method <strong>of</strong> weighted


48<br />

100<br />

n,c<br />

We<br />

Unstable<br />

10<br />

Stable<br />

10 −10 10 −9 10 −8 10 −7 10 −6 10 −5 10 −4 10 −3<br />

Figure 17: Semi-analytical neutral curves for the plane Couette flow <strong>of</strong> the UCM<br />

fluid with the anisotropic drag <strong>slip</strong> model. ( )-k x = 100, (— ——)-k x =1.<br />

For both curves, ρ =0.1.<br />

ερ<br />

residuals [17]. For a general, l<strong>in</strong>ear, nonhomogeneous problem with operator L,<br />

Lu(ξ) =f(ξ), (28)<br />

a solution can be proposed <strong>in</strong> terms <strong>of</strong> the basis functions, φ i (ξ),<br />

u N (ξ) =<br />

N∑<br />

a i φ i (ξ), (29)<br />

i=0<br />

where u N is an approximation to the true solution. The residual, def<strong>in</strong>ed by<br />

R(ξ) =Lu N (ξ) − f(ξ), is a measure <strong>of</strong> the error <strong>of</strong> the approximation <strong>and</strong> should<br />

be m<strong>in</strong>imized <strong>in</strong> some fashion. The m<strong>in</strong>imization is accomplished by requir<strong>in</strong>g the<br />

<strong>in</strong>ner product <strong>of</strong> the residual with another set <strong>of</strong> functions, the trial functions ψ i (ξ),<br />

to be zero, i.e. 〈R(ξ),ψ i (ξ)〉 =0forallψ i , i =0,... ,N. After substitut<strong>in</strong>g the<br />

def<strong>in</strong>ition <strong>of</strong> the residual <strong>and</strong> rearrang<strong>in</strong>g, the MWR problem reduces to<br />

〈Lφ j ,ψ i 〉a j = 〈f,ψ i 〉. (30)


49<br />

Everyth<strong>in</strong>g <strong>in</strong> this equation is known, except for the vector <strong>of</strong> coefficients, a j .This<br />

is just a st<strong>and</strong>ard system <strong>of</strong> algebraic equations to solve. For a Chebyshev collocation<br />

technique, the basis functions are taken to be Chebyshev polynomials <strong>and</strong><br />

the trial functions to be Dirac delta functions. This forces the residual to be zero<br />

at the collocation po<strong>in</strong>ts. Solution <strong>of</strong> Eq. 30 then gives the values <strong>of</strong> the expansion<br />

coefficients.<br />

However, knowledge <strong>of</strong> the expansion coefficients is not required to discretize<br />

the differential equation. All <strong>of</strong> the <strong>in</strong>formation about the solution is conta<strong>in</strong>ed <strong>in</strong><br />

the values <strong>of</strong> u N at the collocation po<strong>in</strong>ts. To see this, let ξ i = ξ 0 ,... ,ξ N be the<br />

chosen set <strong>of</strong> collocation po<strong>in</strong>ts. Eq. 29 can be evaluated at all <strong>of</strong> the collocation<br />

po<strong>in</strong>ts to give a system <strong>of</strong> equations, written <strong>in</strong> matrix form as<br />

⎛ ⎞ ⎛<br />

⎞ ⎛ ⎞<br />

u N (ξ 0 )<br />

φ 0 (ξ 0 ) ··· φ N (ξ 0 )<br />

a 0<br />

⎜ .<br />

⎟<br />

⎝ ⎠ = uN i =<br />

⎜ . . .<br />

⎟ ⎜ .<br />

⎟<br />

⎝<br />

⎠ ⎝ ⎠ = A ija j . (31)<br />

u N (ξ N )<br />

φ 0 (ξ N ) ··· φ N (ξ N )<br />

The derivative <strong>of</strong> u can be approximated as (u N ) ′ = ∑ N<br />

i=0 a iφ ′ i (ξ). The derivative<br />

can be evaluated at all <strong>of</strong> the collocation po<strong>in</strong>ts <strong>and</strong> the system <strong>of</strong> equations written<br />

as (u N ) ′ i = A d ij<br />

a j . This relation can be comb<strong>in</strong>ed with Eq. 31 to give (u N ) ′ i =<br />

A dij A −1<br />

jk uN k . The matrix (D N) ik = A dij A −1<br />

jk<br />

a N<br />

is the Chebyshev derivative matrix<br />

operator. It depends only upon the basis functions <strong>and</strong> collocation po<strong>in</strong>ts used, not<br />

the expansion coefficients or the differential equation.<br />

Chebyshev collocation is employed <strong>in</strong> the <strong>in</strong>terval ξ =[−1, 1]. The most common<br />

choice for the collocation po<strong>in</strong>ts is the Gauss-Lobatto po<strong>in</strong>ts given by<br />

ξ j =cos jπ N . (32)


50<br />

For these po<strong>in</strong>ts, the derivative matrix takes on a very simple form<br />

⎧<br />

⎪⎨<br />

(D N ) ik =<br />

⎪⎩<br />

a i (−1) i+k<br />

a k (ξ i −ξ k )<br />

i ≠ k<br />

−ξ k<br />

2(1−ξ k 2 )<br />

1 ≤ i = k ≤ N − 1<br />

2N 2 +1<br />

6<br />

i = k =0<br />

− 2N 2 +1<br />

6<br />

i = k = N<br />

. (33)<br />

The range for the present problem is y =[0, 1] rather than [−1, 1]. The mapp<strong>in</strong>g<br />

y j → 1 2 (1 + ξ j) maps the typical <strong>in</strong>terval onto the computational <strong>in</strong>terval for this<br />

problem. Derivatives with respect to y are replaced by d/dy → (dξ/dy)d/dξ =<br />

2 d/dξ. Fortunately, this mapp<strong>in</strong>g only results <strong>in</strong> the derivative matrix be<strong>in</strong>g scaled<br />

by a factor <strong>of</strong> 2. This choice <strong>of</strong> collocation po<strong>in</strong>ts is also beneficial <strong>in</strong> that the po<strong>in</strong>ts<br />

are distributed nonuniformly <strong>in</strong> the region <strong>of</strong> <strong>in</strong>terest. Specifically, the po<strong>in</strong>ts are<br />

more densely placed near the <strong>wall</strong>s <strong>and</strong> more sparsely placed near the center <strong>of</strong> the<br />

channel. The disturbance eigenvectors are expected to be localized near the <strong>wall</strong>,<br />

so the po<strong>in</strong>ts are arranged precisely where they are needed for the calculations.<br />

The formulation <strong>of</strong> the numerical problem is slightly different than the formulation<br />

for the analytical results. Instead <strong>of</strong> be<strong>in</strong>g reduced to one general equation,<br />

the system <strong>of</strong> equations for the amplitude functions, Eqs. 87a -87f , is written<br />

<strong>in</strong> terms <strong>of</strong> four <strong>in</strong>dependent variables: the two normal stresses, the <strong>shear</strong> stress,<br />

<strong>and</strong> the stream function. Pressure is elim<strong>in</strong>ated by tak<strong>in</strong>g the curl <strong>of</strong> the equation<br />

<strong>of</strong> motion. There are two reasons for this choice. The operator M <strong>in</strong> Eq. 89 is<br />

extremely unwieldy, <strong>and</strong> more importantly, the general stability equation conta<strong>in</strong>s<br />

fourth order derivatives, which leads to an extremely stiff system <strong>of</strong> equations after<br />

discretization <strong>and</strong> makes reliable solutions impossible to obta<strong>in</strong>.


51<br />

The generalized eigenvalue problem for the PTT equation is<br />

ik x We n c Câ = Lâ (34)<br />

where â =(ˆψ, ˆτ xx , ˆτ yx , ˆτ yy ) <strong>and</strong> the operators L <strong>and</strong> C are shown <strong>in</strong> Appendix A.3.<br />

Discretization gives 4(N + 1) unknowns, which are the values <strong>of</strong> the <strong>in</strong>dependent<br />

variables at the collocation po<strong>in</strong>ts. The discretized eigenvalue problem is solved<br />

us<strong>in</strong>g a public doma<strong>in</strong> rout<strong>in</strong>e, namely the Transactions on Mathematical S<strong>of</strong>tware<br />

(TOMS) rout<strong>in</strong>e 535 [23].<br />

This rout<strong>in</strong>e returns all <strong>of</strong> the eigenvalues, allow<strong>in</strong>g<br />

the entire spectrum to be analyzed.<br />

To remove spurious eigenvalues, <strong>boundary</strong><br />

regularization, as discussed by Graham [28], is used. ‡<br />

3.4.3 Numerical Results<br />

The analytical solutions for the UCM fluid presented <strong>in</strong> §3.4.1 can be used to validate<br />

the numerical technique. Fig. 18 compares growth rate curves obta<strong>in</strong>ed numerically<br />

us<strong>in</strong>g the Chebyshev technique <strong>and</strong> analytically us<strong>in</strong>g the spatially decay<strong>in</strong>g solutions.<br />

The number <strong>of</strong> collocation po<strong>in</strong>ts, N, is 256. At large k x , numerical results<br />

are not presented, due to the fact that the cont<strong>in</strong>uous spectrum becomes spuriously<br />

unstable above k x ≈ 20. At moderate wavenumbers (k x ≈ 10), the eigenvectors are<br />

localized near the plates, the analytical solutions are still accurate, <strong>and</strong> the numerics<br />

agree with the analytical solution. At small k x , the curves diverge, as the eigenvectors<br />

are no longer localized enough for the analytical solution to be accurate. The<br />

neutral curve for k x = 10 is shown <strong>in</strong> Fig. 19 for the network model. The po<strong>in</strong>ts<br />

were obta<strong>in</strong>ed us<strong>in</strong>g the Chebyshev method while the solid curve is the analytical<br />

‡ Boundary regularization was used to remove spurious modes for the results published <strong>in</strong><br />

Ref. [14] <strong>and</strong>, therefore, was mentioned here. A much better method for elim<strong>in</strong>at<strong>in</strong>g spurious


52<br />

0.0015<br />

Im(k x<br />

c)<br />

−0.0005<br />

−0.0025<br />

−0.0045<br />

0 5 10 15 20<br />

k x<br />

Figure 18: Comparison <strong>of</strong> the growth rate curves predicted numerically <strong>and</strong> analytically<br />

for the UCM equation <strong>and</strong> the network <strong>slip</strong> model. We t = 15, We n =17.341,<br />

ɛs =10 −5 , <strong>and</strong> for the numerical solutions, N = 256.<br />

result. As expected, these curves are <strong>in</strong> excellent agreement, as is the predicted<br />

eigenvalue spectrum, also shown <strong>in</strong> Fig. 16. The numerical spectrum shows four<br />

pairs <strong>of</strong> eigenvalues, s<strong>in</strong>ce the numerical technique obta<strong>in</strong>s all <strong>of</strong> the solutions to the<br />

general stability equation, <strong>and</strong> the cont<strong>in</strong>uous spectrum appears as a r<strong>in</strong>g s<strong>in</strong>ce the<br />

eigenvectors correspond<strong>in</strong>g to the cont<strong>in</strong>uous eigenvalues are s<strong>in</strong>gular <strong>and</strong> poorly<br />

approximated by polynomials (c.f. Graham [28]). The poor approximation <strong>of</strong> the<br />

cont<strong>in</strong>uous spectrum leads to the spuriously unstable behavior <strong>of</strong> the cont<strong>in</strong>uous<br />

spectrum alluded to earlier for large k x . Other spurious behavior is possible, most<br />

notably, the r<strong>in</strong>g can be large enough <strong>in</strong> some <strong>in</strong>stances to overlap with <strong>and</strong> obscure<br />

the discrete eigenvalues.<br />

For the PTT constitutive equation, only numerical solutions are possible. We<br />

chose small values <strong>of</strong> µ(= 10 −4 , 10 −3 ) s<strong>in</strong>ce Phan-Thien <strong>and</strong> Tanner [69] reported<br />

good fits <strong>of</strong> experimental data for a low density polyethylene with µ =10 −3 <strong>and</strong><br />

pressure modes is to use a primitive variable formulation coupled with a staggered grid for pressure<br />

(c.f. §5.2.1). Use <strong>of</strong> a primitive variable formulation similar to the one <strong>in</strong> §5.2.1 gives identical<br />

results to the streamfunction formulation with <strong>boundary</strong> regularization for this problem.


53<br />

100<br />

We<br />

n,c<br />

Unstable<br />

Stable<br />

10<br />

10 −8 10 −7 10 −6 10 −5 10 −4 10 −3<br />

Figure 19: Numerical <strong>and</strong> analytical neutral curves for the UCM equation <strong>and</strong> the<br />

network <strong>slip</strong> model with k x = 10. For the numerical results N = 192.<br />

ε s<br />

we wish to exam<strong>in</strong>e small deviations from the UCM equation. Fig. 20 shows the<br />

neutral curves for k x = 1 <strong>and</strong> 10 when the network model is used as the <strong>slip</strong> relation.<br />

1000<br />

We n,c<br />

100<br />

Unstable<br />

Stable<br />

10<br />

10 −8 10 −7 10 −6 10 −5 10 −4<br />

Figure 20: Neutral curves for the PTT constitutive equation with the network model<br />

as the <strong>slip</strong> relation. ○ - k x =1,µ =10 −3 ; 3 - k x =1,µ =10 −4 ; △ - k x = 10,<br />

µ =10 −3 ; 2 - k x = 10, µ =10 −4 . For all curves, N = 192.<br />

ερ εs<br />

Increas<strong>in</strong>g µ <strong>in</strong>creases the slope <strong>of</strong> the power law region at small ɛs. At higher ɛs<br />

the s<strong>in</strong>gularity is still present, but the location depends upon µ. In fact, the curves<br />

cross <strong>and</strong> the curves for higher µ are less stable. At the <strong>shear</strong> rates required for


54<br />

<strong>in</strong>stability, <strong>shear</strong> th<strong>in</strong>n<strong>in</strong>g is just beg<strong>in</strong>n<strong>in</strong>g to set <strong>in</strong>, but small changes <strong>in</strong> the stress<br />

greatly affect the <strong>slip</strong> velocity, s<strong>in</strong>ce at steady state ū s ∝ ¯τ yx 3 . So, at a given ɛs<br />

<strong>and</strong> applied flow rate, the <strong>slip</strong> velocity for the PTT model is reduced relative to the<br />

<strong>slip</strong> velocity for the UCM model <strong>and</strong> this leads to higher <strong>shear</strong> rates <strong>and</strong> stresses<br />

<strong>in</strong> the fluid for the same nom<strong>in</strong>al We. Higher stresses can be ma<strong>in</strong>ta<strong>in</strong>ed <strong>in</strong> the<br />

fluid allow<strong>in</strong>g the critical stress to be reachable for larger ɛs <strong>and</strong> We n lead<strong>in</strong>g to less<br />

stability. F<strong>in</strong>ally, neutral curves for the PTT equation <strong>and</strong> the anisotropic drag <strong>slip</strong><br />

model are shown <strong>in</strong> Fig. 21. In contrast to the UCM case, comparison <strong>of</strong> Figs. 20<br />

<strong>and</strong> 21 reveals that there is a significant difference <strong>in</strong> the neutral curves when the<br />

100<br />

We n,c<br />

Unstable<br />

10<br />

Stable<br />

10 −8 10 −7 10 −6 10 −5 10 −4 10 −3 10 −2<br />

Figure 21: Numerical neutral curves for the PTT constitutive equation with the<br />

anisotropic drag <strong>slip</strong> model. ○ - k x =1,3 - k x = 10. For both curves, µ =10 −4 ,<br />

N = 192, <strong>and</strong> ρ =0.1.<br />

ερ<br />

<strong>slip</strong> model is changed.<br />

In particular, the location <strong>of</strong> the s<strong>in</strong>gularity at large ɛρ<br />

is shifted to higher value <strong>of</strong> ɛρ for the anisotropic drag <strong>slip</strong> model as opposed to<br />

the network model, dramatically destabiliz<strong>in</strong>g the flow at large values <strong>of</strong> the <strong>slip</strong><br />

coefficient.<br />

For comparison with the asymptotic results, as well as experimental evidence


55<br />

<strong>and</strong> other <strong>slip</strong> theories, the results here can be recast <strong>in</strong> terms <strong>of</strong> the extrapolation<br />

length, b = u s / ˙γ. The extrapolation length is given by Eq. 15 with ¯f = ɛs¯τ xx for<br />

the network model <strong>and</strong> ¯f = ɛ(1 + ρ¯τ xx − ρ 2¯τ yx 2 ) for the anisotropic drag model.<br />

Interest<strong>in</strong>gly, a master curve can be plotted as the true <strong>shear</strong> stress, i.e., the true<br />

Weissenberg number, versus the scaled wavenumber k x b, as shown <strong>in</strong> Fig. 22 for both<br />

the UCM <strong>and</strong> PTT constitutive equations. The two sets <strong>of</strong> po<strong>in</strong>ts are numerical<br />

results for the PTT constitutive equation obta<strong>in</strong>ed us<strong>in</strong>g the Chebyshev method<br />

1000<br />

τ * yx,c<br />

/G *<br />

100<br />

Unstable<br />

UCM; Network<br />

UCM; AD<br />

PTT; Network<br />

PTT; AD<br />

10<br />

Stable<br />

10 −3 10 −2 10 −1 10 0 10 1 10 2 10 3<br />

Figure 22: Master curve <strong>of</strong> k x b versus We t for the network model.<br />

k x<br />

b<br />

while the solid curves are the analytical results for the UCM equation on a halfplane.<br />

All <strong>of</strong> the data fall roughly onto one curve <strong>and</strong> predict a basically constant<br />

critical <strong>shear</strong> stress (recoverable <strong>shear</strong>) <strong>of</strong> ≈ 11 for k x b 1. All wavelengths shorter<br />

than the extrapolation length have essentially the same growth rate <strong>and</strong> wave speed<br />

(= ū s ) <strong>and</strong> the flow is more stable for disturbance wavelengths greater than the<br />

extrapolation length (k x b


56<br />

by molecular dynamics occurr<strong>in</strong>g on fast time scales that are not <strong>in</strong>cluded <strong>in</strong> the<br />

models. Other possibilities arise when consider<strong>in</strong>g more complicated constitutive<br />

behavior( both bulk <strong>and</strong> <strong>in</strong>terfacial), where physical mechanisms such as reptation<br />

<strong>and</strong> disentanglement are <strong>in</strong>cluded (Wang et al. [100], Mhetar <strong>and</strong> Archer [55], Ajdari<br />

et al. [3]). In any case, the longest wavelength observed would be on the order <strong>of</strong><br />

the extrapolation length (i.e., k x b =1).<br />

The solutions presented <strong>in</strong> this section qualitatively predict aspects <strong>of</strong> sharksk<strong>in</strong><br />

other than those already predicted by the asymptotic results. Most importantly,<br />

these solutions show that <strong>slip</strong> can be stabiliz<strong>in</strong>g or destabiliz<strong>in</strong>g <strong>in</strong> Couette flow, depend<strong>in</strong>g<br />

upon the magnitude <strong>of</strong> the <strong>slip</strong> velocity. For small <strong>slip</strong> velocities, <strong>in</strong>creas<strong>in</strong>g<br />

the magnitude <strong>of</strong> <strong>slip</strong> destabilizes the flow by decreas<strong>in</strong>g the critical stress for the<br />

<strong>in</strong>stability. At large enough values <strong>of</strong> the <strong>slip</strong> coefficient, the critical stress for <strong>in</strong>stability<br />

becomes a constant <strong>and</strong> further <strong>in</strong>creas<strong>in</strong>g the magnitude <strong>of</strong> <strong>slip</strong> stabilizes the<br />

flow by reduc<strong>in</strong>g the effective <strong>shear</strong> rate <strong>in</strong> the fluid. If the magnitude <strong>of</strong> <strong>slip</strong> is large<br />

enough, the critical stress required for <strong>in</strong>stability cannot be obta<strong>in</strong>ed <strong>and</strong> the flow<br />

becomes stable for all We. These results are robust with respect to the exact form<br />

<strong>of</strong> the <strong>slip</strong> model <strong>and</strong> to the constitutive equation used. Care must be exercised<br />

<strong>in</strong> extrapolat<strong>in</strong>g these conclusions to pressure driven flow. The <strong>shear</strong> stress at the<br />

<strong>wall</strong> is determ<strong>in</strong>ed solely by the imposed pressure drop, not the <strong>shear</strong> rate. It is<br />

always possible to achieve the critical stress, so it is not correct to associate the<br />

substantially <strong>in</strong>creased <strong>shear</strong> rate with flow stabilization.<br />

3.4.4 Comparison with experiment


57<br />

This analysis leads to several quantitative predictions which can be checked aga<strong>in</strong>st<br />

experimental evidence. First, the <strong>slip</strong> models considered here predict that the <strong>slip</strong><br />

velocity has a power law dependence on the <strong>shear</strong> stress <strong>in</strong> the base state, with a<br />

power law exponent <strong>of</strong> 3. This value agrees very well with published data for HDPE<br />

(Hatzikiriakos <strong>and</strong> Dealy [34, 33]). Values for LLDPE range anywhere from 4 to 6<br />

(Hill et al. [38]), so the predicted power law exponent is low. Second, <strong>in</strong>formation<br />

about the molecular weight <strong>and</strong> temperature dependence can be obta<strong>in</strong>ed.<br />

We<br />

f<strong>in</strong>d a critical true Weissenberg number We t,c for <strong>in</strong>stability that is constant when<br />

k x b 1. Thus the critical <strong>shear</strong> rate is<br />

˙γ ∗ t,c = We t,c<br />

λ . (35)<br />

Generally, the relaxation time strongly decreases with temperature, follow<strong>in</strong>g an<br />

Arrhenius relation, <strong>and</strong> strongly <strong>in</strong>creases with molecular weight (MW) [19], e.g.<br />

λ ∼ MW 3.4 for entangled melts, so that the critical <strong>shear</strong> rate <strong>in</strong>creases sharply<br />

with temperature <strong>and</strong> decreases with the molecular weight. Venet <strong>and</strong> Vergnes [96]<br />

have recently shown for LLDPE that the critical <strong>shear</strong> rate <strong>in</strong>creases as a function<br />

<strong>of</strong> temperature with a rate much faster than l<strong>in</strong>ear, <strong>and</strong> it is well known that<br />

higher molecular weight <strong>polymer</strong>s are more susceptible to sharksk<strong>in</strong> distortions.<br />

∗<br />

The predicted critical recoverable <strong>shear</strong>, ¯τ yx,c /G ∗ = We t,c , is a constant, as observed<br />

by Pomar et al. [73], who diluted LLDPE with octadecane to lower the modulus <strong>and</strong><br />

found that sharksk<strong>in</strong> set <strong>in</strong> at a recoverable <strong>shear</strong> <strong>of</strong> 1.73 for a wide range <strong>of</strong> <strong>polymer</strong><br />

weight fractions. S<strong>in</strong>ce G ∗ ∝ T , we also f<strong>in</strong>d that ¯τ ∗ yx,c /T is a constant at onset, <strong>in</strong><br />

agreement with the experiments <strong>of</strong> Venet <strong>and</strong> Vergnes [96] <strong>and</strong> Wang et al. [100]<br />

The longest wavelength observed will be for k x b ∼ = 1, for which the frequency <strong>of</strong><br />

the distortion is ω ∗ c = k ∗ xRe(c ∗ ) ∼ = k ∗ xu ∗ s = k ∗ xb ∗ ˙γ ∗ n = k x bWe n /λ ≈ 11/λ, whichscales


58<br />

as the reciprocal <strong>of</strong> the bulk relaxation time, as seen experimentally for LLDPE<br />

by Wang et al. [100] <strong>and</strong> Barone et al. [8] F<strong>in</strong>ally, the critical recoverable <strong>shear</strong><br />

predicted by the model is around 10. Several experiments have reported critical<br />

stresses for sharksk<strong>in</strong> <strong>in</strong> LLDPE around 0.1 MPa. For Ramamurthy’s [75] data, <strong>in</strong><br />

particular, this gives a recoverable <strong>shear</strong> <strong>of</strong> about 8, so the <strong>slip</strong> results are with<strong>in</strong><br />

the correct order <strong>of</strong> magnitude.<br />

3.5 Mechanism<br />

An energy analysis was performed <strong>in</strong> order to elucidate the cause <strong>of</strong> the <strong>in</strong>stability. It<br />

is desirable to have a molecular level picture <strong>of</strong> what is happen<strong>in</strong>g at the <strong>in</strong>terface to<br />

destabilize the flow. This has not been fully realized <strong>of</strong> yet, but the energy analysis<br />

po<strong>in</strong>ts to the important terms <strong>in</strong> the equations, <strong>and</strong> establishes criteria that any<br />

proposed mechanism must fulfill.<br />

In order to do an energy analysis, one wants to follow the evolution <strong>of</strong> energy or<br />

an energy-like, i.e. a real, positive def<strong>in</strong>ite, quantity. The L 2 norm <strong>of</strong> the solution is<br />

used here, <strong>and</strong> this quantity can be derived start<strong>in</strong>g from the l<strong>in</strong>earized component<br />

equations <strong>of</strong> the UCM constitutive equation for the asymptotic case, k x ∼ 1/ɛ ≫ 1,<br />

y = k x ỹ, We n = We t . For this scal<strong>in</strong>g, the critical We n is We n,c =10.669 for s =1.<br />

The scaled stress equations are<br />

∂τ xx<br />

+ū ∂τ xx<br />

∂t ∂x − 2(1 + ¯τ xx) ∂u<br />

∂x − 2¯τ ∂u<br />

yx<br />

∂y − 2τ ∂ū<br />

yx<br />

∂y + 1 τ xx =0,<br />

We n<br />

(36)<br />

∂τ yx<br />

+ū ∂τ yx<br />

∂t ∂x − τ ∂ū<br />

yy<br />

∂y − ∂u<br />

∂y − (¯τ xx +1) ∂v<br />

∂x + 1 τ yx =0,<br />

We n<br />

(37)<br />

∂τ yy<br />

+ū ∂τ yy<br />

∂t ∂x − 2¯τ ∂v<br />

yx<br />

∂x − 2∂v ∂y + 1 τ yy =0,<br />

We n<br />

(38)


59<br />

where the barred variables denote the base state <strong>and</strong> now the unmarked variables<br />

are the perturbations.<br />

The perturbations are written <strong>in</strong> normal mode form as<br />

a =ã +ã ∗ ,whereã =â(ỹ, t)e ikxx <strong>and</strong> the complex conjugate ã ∗ =â ∗ (ỹ, t)e −ikxx .<br />

The time dependence has not been explicitly <strong>in</strong>cluded yet. With these def<strong>in</strong>itions,<br />

Eq. 36 becomes<br />

∂<br />

∂t (˜τ xx +˜τ ∗ xx )+ik x ū(˜τ xx − ˜τ ∗ xx ) − 2(¯τ xx +1)ik x (ũ − ũ ∗ ) − 2¯τ yx (ũ ′ +ũ ′∗ ) −<br />

2(˜τ yx +˜τ ∗ yx )+ 1 (˜τ xx +˜τ ∗ xx )=0. (39)<br />

We n<br />

Note that all the terms <strong>in</strong> this equation are real. The next step is to multiply the<br />

entire equation by (˜τ xx +˜τ xx ∗ ) <strong>and</strong> then <strong>in</strong>tegrate over x <strong>and</strong> ỹ. This yields<br />

1 ∂<br />

2 ∂t 〈(˜τ xx +˜τ ∗ xx ) 2 〉 = −iū〈(˜τ xx − ˜τ ∗ xx )(˜τ xx +˜τ ∗ xx )〉 +<br />

2i(¯τ xx +1)〈(ũ − ũ ∗ )(˜τ xx +˜τ ∗ xx )〉− 1 〈(˜τ xx +˜τ ∗ xx ) 2 〉 +<br />

We n<br />

2¯τ yx 〈(ũ ′ +ũ ′∗ )(˜τ xx +˜τ xx ∗ )〉 +2〈(˜τ yx +˜τ yx ∗ )(˜τ xx +˜τ xx ∗ )〉, (40)<br />

where the brackets denote the double <strong>in</strong>tegration. This equation can be further<br />

simplified to give<br />

∂<br />

∂t 〈〈ˆτ xxˆτ ∗ xx 〉〉 =2i(¯τ xx +1)〈〈ûˆτ ∗ xx − û ∗˜τ xx 〉〉 − 2 〈〈ˆτ xxˆτ ∗ xx 〉〉 +<br />

We n<br />

2¯τ yx 〈〈ˆτ xx û ′∗ +ˆτ xx∗û ′ 〉〉 +2〈〈ˆτ xxˆτ yx ∗ +ˆτ xx∗ˆτ yx 〉〉, (41)<br />

where the double bracket <strong>in</strong>dicates <strong>in</strong>tegration over ỹ, <strong>and</strong> the <strong>in</strong>tegration over x<br />

has been implicitly performed. There are a couple <strong>of</strong> po<strong>in</strong>ts to notice. First, s<strong>in</strong>ce<br />

the goal is to determ<strong>in</strong>e the important terms at the onset <strong>of</strong> <strong>in</strong>stability, Im(c) is<br />

zero, <strong>and</strong> c is a purely real quantity. S<strong>in</strong>ce every term <strong>in</strong> Eq. 41 conta<strong>in</strong>s a variable<br />

multiplied by the conjugate <strong>of</strong> another, <strong>and</strong> the perturbations can be written as


60<br />

â(ỹ)e −ikxct , the time dependence will cancel out <strong>of</strong> every term, <strong>and</strong> each term will<br />

only be a function <strong>of</strong> ỹ. Therefore, only the amplitude functions that were calculated<br />

earlier are needed. Second, because the time dependence cancels out <strong>of</strong> the left h<strong>and</strong><br />

side <strong>of</strong> the equation at the onset <strong>of</strong> <strong>in</strong>stability, the time derivative will be zero, <strong>and</strong><br />

this fact provides a simple check on the calculations. The other two components<br />

<strong>of</strong> the constitutive law can be reduced the same way.<br />

When all three result<strong>in</strong>g<br />

equations are added together, the result is one equation for the time rate <strong>of</strong> change<br />

<strong>of</strong> an energy-like quantity<br />

∂<br />

∂t 〈〈ˆτ xxˆτ ∗ xx +ˆτ yxˆτ ∗ yx +ˆτ yyˆτ ∗ yy 〉〉 =2i(¯τ xx +1)〈〈ûˆτ ∗ xx − û ∗ˆτ xx 〉〉 +<br />

2¯τ yx ˙γ〈〈ˆτ xx û ′∗ +ˆτ xx∗û ′ 〉〉 +2〈〈ˆτ xxˆτ ∗ yx +ˆτ xx∗ˆτ yx 〉〉 −<br />

2<br />

〈〈ˆτ xxˆτ ∗ xx 〉〉 +(¯τ xx +1)i〈〈ˆvˆτ ∗ yx − ˆv ∗ˆτ yx 〉〉 +<br />

We n<br />

˙γ〈〈ˆτ yxˆτ ∗ yy +ˆτ yx∗ˆτ yy 〉〉 + 〈〈ˆτ yx û ′∗ +ˆτ yx∗û ′ 〉〉 −<br />

2<br />

〈〈ˆτ yxˆτ ∗ yx 〉〉 +2i¯τ yx 〈〈ˆvˆτ ∗ yy − ˆv ∗ˆτ yy 〉〉 +<br />

We n<br />

2〈〈ˆτ yyˆv ′∗ +ˆv ′ˆτ yy ∗ 〉〉 − 2<br />

We n<br />

〈〈ˆτ yyˆτ yy ∗ 〉〉. (42)<br />

A very similar equation can be derived from the exact forms <strong>of</strong> the component<br />

equations so that the analysis can be extended to the no-<strong>slip</strong> case (Gorodtsov <strong>and</strong><br />

Leonov solution). This equation is not shown for brevity. The parameters are chosen<br />

to correspond with the asymptotic result, We n =10.7, k x = 100. Table 4 shows the<br />

values <strong>of</strong> the terms <strong>in</strong> Eq. 42 obta<strong>in</strong>ed for both the destabiliz<strong>in</strong>g <strong>slip</strong> mode <strong>and</strong> the<br />

Gorodtsov-Leonov mode for the no-<strong>slip</strong> case as well as the constitutive component<br />

that the terms are from.<br />

Note that the terms for the <strong>slip</strong> case sum to zero,<br />

validat<strong>in</strong>g the calculations, but that the terms for the no <strong>slip</strong> case do not. This is<br />

because the <strong>slip</strong> case is calculated at the onset <strong>of</strong> <strong>in</strong>stability, where the Im(c) is zero,


61<br />

Term Slip No-<strong>slip</strong> (G-L)<br />

2i(¯τ xx +1)〈〈ûˆτ<br />

⎧⎪ ∗ xx − û ∗ˆτ xx 〉〉 11003.4 -3045.31<br />

⎨<br />

2¯τ yx 〈〈ˆτ xx û ′∗ +ˆτ xx∗û ′ 〉〉 17078.6 3814.9<br />

ˆτ xx<br />

2˙γ〈〈ˆτ xxˆτ ∗ yx +ˆτ xx∗ˆτ yx 〉〉 44857 45114.2<br />

⎪ ⎩<br />

−<br />

⎧⎪ 2<br />

De n<br />

〈〈ˆτ xxˆτ ∗ xx 〉〉 -72939 -101921<br />

i(¯τ xx +1)〈〈ˆvˆτ ∗ yx − ˆv ∗ˆτ yx 〉〉 237.622 -23.6812<br />

⎨<br />

˙γ〈〈ˆτ yxˆτ ∗ yy +ˆτ yx∗ˆτ yy 〉〉 4624.58 62.4942<br />

ˆτ yx<br />

〈〈ˆτ yx û ′∗ +ˆτ yx∗û ′ 〉〉 -90.645 6.61066<br />

⎪ ⎩<br />

−<br />

⎧ 2<br />

De n<br />

〈〈ˆτ yxˆτ ∗ yx 〉〉 -4771.55 -100.899<br />

⎪⎨ 2i¯τ yx 〈〈ˆvˆτ ∗ yy − ˆv ∗ˆτ yy 〉〉 20.6294 -.0066018<br />

ˆτ yy 2〈〈ˆτ<br />

⎪ yyˆv ′∗ +ˆv ′ˆτ ∗ yy 〉〉 19.2603 .204656<br />

⎩<br />

2<br />

De n<br />

〈〈ˆτ yyˆτ ∗ yy 〉〉 -39.8897 -.439934<br />

Table 4: Comparison <strong>of</strong> the results for <strong>slip</strong> <strong>and</strong> no-<strong>slip</strong>.<br />

but there is no po<strong>in</strong>t where the Im(c) is zero for the no-<strong>slip</strong> case, s<strong>in</strong>ce the flow with<br />

no <strong>slip</strong> is unconditionally stable. This means that the the time exponential does<br />

not cancel out <strong>of</strong> the terms <strong>in</strong> the no-<strong>slip</strong> case. Instead, every term is multiplied<br />

by the factor e ikx(c−c∗ )t . The calculations were done with the implicit assumption<br />

that t = 0, thereby dropp<strong>in</strong>g that factor out. By this method, the largest terms are<br />

the coupl<strong>in</strong>g terms between the base state velocity gradients <strong>and</strong> the perturbation<br />

stresses. However, the effect <strong>of</strong> this coupl<strong>in</strong>g is the same <strong>in</strong> both cases, even though<br />

destabiliz<strong>in</strong>g, <strong>and</strong> cannot cause a fundamental change <strong>in</strong> the stability. Also, the<br />

viscous terms (terms conta<strong>in</strong><strong>in</strong>g the reciprocal <strong>of</strong> the nom<strong>in</strong>al Deborah number) are<br />

very large <strong>in</strong> absolute value <strong>and</strong> stabiliz<strong>in</strong>g, but, aga<strong>in</strong> the behavior is the same <strong>in</strong><br />

both cases. The crucial terms are those describ<strong>in</strong>g the coupl<strong>in</strong>g between the base<br />

state stresses <strong>and</strong> the perturbation velocity gradients, ¯τ ·∇ˆv. This implies that


62<br />

velocity perturbations further stretch <strong>and</strong> orient cha<strong>in</strong>s, which leads to <strong>in</strong>creased<br />

<strong>slip</strong>, lead<strong>in</strong>g to build up <strong>of</strong> the <strong>in</strong>stability.<br />

The same coupl<strong>in</strong>g behavior is seen <strong>in</strong> a phase shift between the perturbation<br />

<strong>slip</strong> velocity <strong>and</strong> the perturbation stresses, as shown <strong>in</strong> Fig. 23. Such a phase shift<br />

is impossible <strong>in</strong> the absence <strong>of</strong> a τ xx<br />

1.5<br />

1.0<br />

dependence <strong>of</strong> the <strong>slip</strong> coefficient unless a<br />

^<br />

u^ ^<br />

τ yx<br />

τ xx<br />

0.5<br />

τ ^<br />

xx<br />

, ^τ yx<br />

, u ^<br />

0.0<br />

−0.5<br />

−1.0<br />

−1.5<br />

0 10 20 30<br />

t<br />

Figure 23: Phase shift between the <strong>slip</strong> velocity <strong>and</strong> the <strong>shear</strong> <strong>and</strong> normal stress<br />

components at the critical po<strong>in</strong>t. ( )-ˆτ xx ,(–––––)-ˆτ yx ,<strong>and</strong>(———)<br />

-û. The parameter values are k x = 10, ɛs =10 −5 , We n,c =14.77, We t,c =13.36, <strong>and</strong><br />

Re(c) =0.04944, <strong>and</strong> the amplitudes have been normalized to unity.<br />

“memory <strong>slip</strong>” model like Eq. 1 is used. Instability is never observed <strong>in</strong> the absence<br />

<strong>of</strong> a phase shift (R 0 = 0 <strong>in</strong> the asymptotic analysis). The fact that the memory<br />

<strong>slip</strong> model also leads to <strong>in</strong>stabilities suggests that the phase shift is an essential<br />

component <strong>of</strong> the mechanism.<br />

3.6 Summary <strong>of</strong> the Melt Analysis<br />

Slip models which <strong>in</strong>clude a normal stress dependence <strong>of</strong> the <strong>slip</strong> velocity lead to<br />

novel shortwave hydrodynamic <strong>in</strong>stabilities <strong>in</strong> <strong>in</strong>ertialess, viscoelastic parallel <strong>shear</strong>


63<br />

<strong>flows</strong>. Analysis <strong>of</strong> <strong>flows</strong> with these <strong>slip</strong> models makes several predictions consistent<br />

with experimental observations <strong>of</strong> sharksk<strong>in</strong>.<br />

Most significantly, this model predicts<br />

a short wavelength <strong>in</strong>stability at zero Reynolds number that is localized near<br />

the bound<strong>in</strong>g surfaces. This <strong>in</strong>stability is found even though the flow curves are<br />

monotonic <strong>in</strong>creas<strong>in</strong>g <strong>and</strong> the model is mathematically well-posed at all wavenumbers.<br />

Both viscoelasticity <strong>and</strong> normal stress dependent <strong>slip</strong> are necessary for the<br />

<strong>in</strong>stability to occur. The onset conditions depend upon bulk <strong>polymer</strong> properties,<br />

<strong>in</strong> particular the relaxation time, as well as the <strong>in</strong>terfacial ones, although the predictions<br />

are not sensitive to the functional form used to describe <strong>slip</strong>. The model<br />

predicts that <strong>slip</strong> can be stabiliz<strong>in</strong>g or destabiliz<strong>in</strong>g <strong>in</strong> Couette flow, depend<strong>in</strong>g upon<br />

the flow parameters. The neutral curve goes through a m<strong>in</strong>imum as the degree <strong>of</strong><br />

<strong>slip</strong> <strong>in</strong>creases from zero, after which the critical Weissenberg number rapidly <strong>in</strong>creases<br />

<strong>and</strong> the neutral curve becomes s<strong>in</strong>gular. Beyond this po<strong>in</strong>t the <strong>slip</strong> velocity<br />

is so large that high enough stresses to cause the <strong>in</strong>stability cannot be generated<br />

<strong>in</strong> the fluid <strong>and</strong> Couette flow is unconditionally stable.<br />

Replott<strong>in</strong>g the neutral<br />

curves as the critical true Weissenberg number (the dimensionless <strong>shear</strong> stress) versus<br />

the wavenumber scaled with the extrapolation length yielded a master curve.<br />

The critical Weissenberg number is on the order <strong>of</strong> 10, which is the same order<br />

<strong>of</strong> magnitude as observed experimentally, <strong>and</strong> predictions <strong>of</strong> the critical <strong>shear</strong> rate<br />

<strong>and</strong> molecular weight <strong>and</strong> temperature dependencies are <strong>in</strong> qualitative agreement<br />

with experiments. The results are robust with respect to the actual forms <strong>of</strong> the<br />

<strong>in</strong>terfacial <strong>and</strong> bulk constitutive relations employed. F<strong>in</strong>ally, the mechanism must<br />

be tied to the phase shift between the perturbation <strong>slip</strong> velocity <strong>and</strong> the perturbation<br />

<strong>shear</strong> stress, as this phase shift is only present <strong>in</strong> normal stress dependent <strong>slip</strong>


64<br />

models or memory <strong>slip</strong> models. This analysis is the only analysis which predicts<br />

hydrodynamic <strong>in</strong>stabilities consistent with sharksk<strong>in</strong>, <strong>and</strong> highlights the importance<br />

<strong>of</strong> <strong>in</strong>clud<strong>in</strong>g cha<strong>in</strong> stretch<strong>in</strong>g <strong>and</strong> orientation <strong>effects</strong> <strong>in</strong> models for <strong>slip</strong>.<br />

A primary limitation to the results presented here is the model geometry used<br />

<strong>in</strong> the analysis, which does not approximate the real flow situation dur<strong>in</strong>g extrusion.<br />

The question naturally arises <strong>of</strong> whether similar elastic <strong>in</strong>stabilities are seen<br />

<strong>in</strong> viscometric <strong>flows</strong>. One possible answer was provided <strong>in</strong> the context <strong>of</strong> entangled<br />

<strong>polymer</strong> solutions by Mhetar <strong>and</strong> Archer [56], who observed enhanced concentration<br />

fluctuations <strong>in</strong> a plane Couette cell. The <strong>slip</strong> behavior <strong>of</strong> the solution clearly <strong>in</strong>fluences<br />

the onset <strong>of</strong> <strong>in</strong>stability <strong>in</strong> this system, but the analysis is more complicated<br />

due to coupl<strong>in</strong>g between the <strong>polymer</strong> stress <strong>and</strong> the concentration <strong>in</strong> these systems.


65<br />

Chapter 4<br />

Concentration Fluctuations <strong>in</strong><br />

Semidilute Polymer Solutions<br />

A recent experimental study by Mhetar <strong>and</strong> Archer [56] has demonstrated several<br />

novel features <strong>of</strong> <strong>shear</strong> enhanced concentration fluctuations <strong>in</strong> entangled <strong>polymer</strong><br />

solutions.<br />

They studied <strong>flows</strong> <strong>of</strong> semidilute solutions <strong>of</strong> polystyrene (PS) <strong>in</strong> diethylphthalate<br />

(DEP), to which tracer particles were added to measure the velocity<br />

pr<strong>of</strong>ile, us<strong>in</strong>g a planar Couette cell. Two features <strong>of</strong> the flow <strong>of</strong> these solutions st<strong>and</strong><br />

out. First, <strong>slip</strong> occurs at the <strong>polymer</strong>/solid <strong>in</strong>terface at low <strong>shear</strong> stresses where no<br />

turbidity was observed. Enhanced concentration fluctuations were observed at large<br />

<strong>shear</strong> stresses, where the <strong>slip</strong> velocity was a strong function <strong>of</strong> the <strong>shear</strong> stress, as<br />

shown <strong>in</strong> Fig. 24 for a 20 wt% solution. The extrapolation, or <strong>slip</strong>, length measured<br />

for these solutions was found to be on the order <strong>of</strong> 10 µm, which is the same as the<br />

length scale for the fluctuations themselves. Second, the concentration fluctuation<br />

enhancement started near the boundaries <strong>and</strong> modification <strong>of</strong> the surface to <strong>in</strong>crease


66<br />

Figure 24: Enhanced concentration fluctuations <strong>in</strong> a semidilute solution <strong>of</strong><br />

polystyrene <strong>in</strong> diethylphthalate at a <strong>shear</strong> rate <strong>of</strong> 3.5 s −1 . Repr<strong>in</strong>ted from Mhetar<br />

<strong>and</strong> Archer [56].<br />

<strong>slip</strong> delayed the onset on fluctuations to much larger <strong>shear</strong> stresses. These observations<br />

strongly suggest that <strong>slip</strong> plays a role <strong>in</strong> the formation <strong>and</strong> development <strong>of</strong><br />

these fluctuations.<br />

Shear enhanced fluctuations <strong>in</strong> semidilute polystyrene solutions have been studied<br />

previously, typically via light scatter<strong>in</strong>g measurements. There are two typical<br />

experimental configurations, which differ primarily <strong>in</strong> the direction <strong>of</strong> illum<strong>in</strong>ation.<br />

The two possibilities are shown schematically <strong>in</strong> Fig. 25. Hashimoto <strong>and</strong> coworkers<br />

[31, 32, 63, 46] have performed a series <strong>of</strong> experiments us<strong>in</strong>g solutions <strong>of</strong> PS<br />

dissolved <strong>in</strong> dioctylphthalate (DOP) <strong>shear</strong>ed <strong>in</strong> a cone <strong>and</strong> plate rheometer [31]<br />

through which laser light was sh<strong>in</strong>ed along the velocity gradient direction.<br />

The<br />

average scatter<strong>in</strong>g <strong>in</strong>tensities <strong>in</strong> the flow <strong>and</strong> normal directions were computed <strong>and</strong><br />

a typical result is shown <strong>in</strong> Fig. 26. At low <strong>shear</strong> rates, ˙γ n ∗ , the scatter<strong>in</strong>g <strong>in</strong>tensity<br />

was low <strong>and</strong> uniform. At ˙γ ∗ n > ˙γ ∗ c , the scatter<strong>in</strong>g <strong>in</strong>tensity <strong>in</strong>creased dramatically


67<br />

(a)<br />

Flow direction<br />

(b)<br />

y<br />

z<br />

x<br />

Figure 25: Schematic <strong>of</strong> typical light scatter<strong>in</strong>g experiments. The dashed arrows<br />

show the two possible directions for illum<strong>in</strong>ation: (a), the velocity gradient direction,<br />

which samples the near surface regions, <strong>and</strong> (b), the neutral direction, which only<br />

measures scatter<strong>in</strong>g <strong>in</strong> the bulk.<br />

Figure 26: Scatter<strong>in</strong>g <strong>in</strong>tensity as a function <strong>of</strong> <strong>shear</strong> rate normalized by the equilibrium<br />

<strong>in</strong>tensity for a 6wt% PS solution. The molecular weight <strong>of</strong> the PS was 5.48<br />

×10 6 . Reproduced from Kume et al. [46].


68<br />

Figure 27: Scatter<strong>in</strong>g pattern as a function <strong>of</strong> the <strong>shear</strong> rate for the same solution<br />

as <strong>in</strong> the previous figure. In (a), the anisotropic peak is just visible. As the <strong>shear</strong><br />

rate <strong>in</strong>creases, the <strong>in</strong>tensity <strong>in</strong>creases <strong>and</strong> the peak shape becomes more def<strong>in</strong>ed.<br />

Reproduced from Hashimoto <strong>and</strong> Kume [32].<br />

<strong>and</strong> the scatter<strong>in</strong>g pattern became anisotropic, as shown <strong>in</strong> Fig. 27, with a broad<br />

scatter<strong>in</strong>g peak <strong>in</strong> the flow direction <strong>and</strong> a narrow dark streak perpendicular to<br />

the flow direction. This “butterfly” pattern <strong>in</strong>dicates the formation <strong>of</strong> flow structures<br />

which are elongated roughly perpendicular to the flow direction so that the<br />

wavevector <strong>of</strong> the fluctuations is roughly oriented <strong>in</strong> the flow direction. The length<br />

scale for these structures can be estimated from the patterns <strong>and</strong> is on the order <strong>of</strong><br />

10 µm. The critical <strong>shear</strong> rate for the formation <strong>of</strong> these anisotropic structures is<br />

˙γ c ∗ ≈ 1/λ, whereλ is the <strong>polymer</strong> relaxation time. This gives a critical We n (≡ λ ˙γ ∗ )<br />

<strong>of</strong> about 1.<br />

Scatter<strong>in</strong>g measurements have also been performed with illum<strong>in</strong>ation along the<br />

neutral, or vorticity, direction, with somewhat different results. Wu et al. [105] used<br />

a circular Couette device with Pyrex cyl<strong>in</strong>ders, <strong>and</strong> observed anisotropic butterfly<br />

patterns. The angle <strong>of</strong> maximum scatter<strong>in</strong>g at low <strong>shear</strong> rates was not <strong>in</strong> the flow<br />

direction; <strong>in</strong>stead, the scatter<strong>in</strong>g pattern was tilted at an angle <strong>of</strong> about 40 ◦ .This<br />

pattern was already visible at a <strong>shear</strong> rate <strong>of</strong> 0.4s −1 (We n ∼ 0.2). The scatter<strong>in</strong>g<br />

peak rotated clockwise (towards the flow direction) <strong>and</strong> the magnitude <strong>of</strong> the<br />

wavevector decreased as the <strong>shear</strong> rate was <strong>in</strong>creased. The <strong>in</strong>itial wavenumber for


69<br />

this peak was ∼ 10µm −1 , giv<strong>in</strong>g a wavelength for the underly<strong>in</strong>g waves <strong>of</strong> ∼ 1µm.<br />

At high <strong>shear</strong> rates, the pattern had rotated so that the peak was below the x-axis.<br />

Similar behavior was reported by Wirtz [104] for a slightly more concentrated solution,<br />

although the measured wavenumber was much smaller than that observed<br />

by Wu et al. [105]. In addition, Wirtz reported that the peak <strong>in</strong>itially moved away<br />

from the orig<strong>in</strong> at low <strong>shear</strong> rates. This behavior was not reported by Wu et al.,<br />

nor is it predicted by any <strong>of</strong> the theories described below.<br />

The sem<strong>in</strong>al theory on a purely hydrodynamic mechanism <strong>of</strong> enhanced concentration<br />

fluctuations was <strong>in</strong>itially proposed by Helf<strong>and</strong> <strong>and</strong> Fredrickson [36] (HF)<br />

<strong>and</strong> later extended by several authors [59, 60, 41, 91]. This mechanism is based on<br />

the idea that variations <strong>in</strong> <strong>polymer</strong> stress can drive a flux <strong>of</strong> <strong>polymer</strong> molecules<br />

– <strong>in</strong>deed, several authors [41, 60, 54] have developed simple, two-fluid models<br />

which demonstrate quite generally that the <strong>polymer</strong> flux, j, for a dilute solution<br />

is j = −D tr ∇n +(D tr /k B T )∇ ·τ ,wheren is the <strong>polymer</strong> number density, τ is<br />

the <strong>polymer</strong> extra stress tensor, D tr<br />

is the <strong>polymer</strong> translational diffusivity, k B<br />

is the Boltzmann constant, <strong>and</strong> T is the temperature.<br />

Furthermore, a detailed<br />

k<strong>in</strong>etic theory derivation [18] leads to a very similar expression for the flux.<br />

A<br />

physical description <strong>of</strong> the basic mechanism was elucidated by Milner [59] <strong>and</strong> Ji<br />

<strong>and</strong> Helf<strong>and</strong> [41] <strong>and</strong> is shown <strong>in</strong> Fig. 28, which shows a r<strong>and</strong>om fluctuation with<br />

wave vector k. The regions <strong>of</strong> higher concentration (the dark regions) have larger<br />

stresses. If the angle <strong>of</strong> the wave vector with respect to the x-axis, β, is <strong>in</strong> the second<br />

quadrant, as <strong>in</strong> Fig. 28(a), the <strong>shear</strong> stresses set up net forces on the molecules<br />

such that the molecules are driven from the regions <strong>of</strong> high concentration to regions<br />

<strong>of</strong> low concentration, thereby <strong>in</strong>creas<strong>in</strong>g the effective rate <strong>of</strong> diffusion. For β <strong>in</strong> the


70<br />

k<br />

k<br />

(a)<br />

(b)<br />

Figure 28: Physical picture <strong>of</strong> the HF hydrodynamic mechanism for enhanced concentration<br />

fluctuations. The wave vector <strong>of</strong> the fluctuation is such that <strong>in</strong>: (a)<br />

the net force on molecules pushes molecules from regions <strong>of</strong> high concentration to<br />

regions <strong>of</strong> low concentration; (b) the net force on molecules pulls molecules <strong>in</strong>to<br />

regions <strong>of</strong> high concentration from regions <strong>of</strong> low concentration. Adapted from Ji<br />

<strong>and</strong> Helf<strong>and</strong> [41]<br />

first quadrant, as <strong>in</strong> Fig. 28(b), the result<strong>in</strong>g net force on molecules pulls molecules<br />

from low concentration regions <strong>in</strong>to higher concentration regions. This enhances<br />

fluctuations by decreas<strong>in</strong>g the effective rate <strong>of</strong> diffusion or by overcom<strong>in</strong>g diffusion<br />

altogether. HF predicted the maximum angle <strong>of</strong> scatter<strong>in</strong>g to be ∼ 41 ◦ at low <strong>shear</strong><br />

rates, which is close to the value measured by Wu et al. [105]. If fluctuations are<br />

short enough, diffusion is faster than stress relaxation <strong>and</strong> the HF mechanism is suppressed<br />

[59]. A scatter<strong>in</strong>g peak occurs when the diffusion <strong>and</strong> relaxation times are<br />

equal, which occurs at a wavenumber <strong>of</strong> k =(D tr λ) −1/2 . This gives a characteristic<br />

length scale for fluctuations <strong>of</strong> √ D tr λ.<br />

This length scale was observed <strong>in</strong> the earlier light scatter<strong>in</strong>g experiments at<br />

low <strong>shear</strong> rates. For the experiments <strong>of</strong> Hashimoto <strong>and</strong> coworkers, the relaxation<br />

time can be estimated as the reciprocal <strong>of</strong> the <strong>shear</strong> rate for the onset <strong>of</strong> <strong>shear</strong><br />

th<strong>in</strong>n<strong>in</strong>g [46] <strong>and</strong> us<strong>in</strong>g the diffusivity measured by Wu et al. [105] for a similar<br />

system gives a predicted wavenumber <strong>of</strong> k ≈ 0.7µm −1 , which is close to the measured


71<br />

value. This length scale was also observed by Wu et al. [105], who found a scatter<strong>in</strong>g<br />

peak at k x ≈ 10µm −1 , which is <strong>in</strong> good agreement with their predicted value <strong>of</strong><br />

9.4µm −1 .<br />

Even though both <strong>of</strong> these experiments show peaks at the requisite length scale,<br />

qualitative differences exist <strong>in</strong> the <strong>shear</strong> rate dependence <strong>of</strong> the results. In particular,<br />

Hashimoto <strong>and</strong> coworkers, who measured the <strong>in</strong>tensity essentially averaged over<br />

a volume conta<strong>in</strong><strong>in</strong>g both near surface regions, found that the <strong>in</strong>tensity <strong>in</strong>creased<br />

markedly above a critical <strong>shear</strong> rate, which was roughly the reciprocal <strong>of</strong> the <strong>polymer</strong><br />

relaxation time.<br />

If fact, for a factor <strong>of</strong> ten <strong>in</strong>crease <strong>in</strong> the <strong>shear</strong> rate above the<br />

critical value, the scatter<strong>in</strong>g <strong>in</strong>tensity <strong>in</strong>creased one hundred fold. Conversely, Wu<br />

et al., who measured the scatter<strong>in</strong>g <strong>in</strong> the bulk, completely miss<strong>in</strong>g the surface<br />

regions, found a much smaller <strong>in</strong>crease <strong>in</strong> <strong>in</strong>tensity as the <strong>shear</strong> rate was <strong>in</strong>creased,<br />

approximately an 11-fold <strong>in</strong>crease <strong>in</strong> scatter<strong>in</strong>g <strong>in</strong>tensity for a 25-fold <strong>in</strong>crease <strong>in</strong><br />

<strong>shear</strong> rate.<br />

This difference is not surpris<strong>in</strong>g, as fluid mechanics results suggest that the dynamics<br />

near surfaces should be different than dynamics <strong>in</strong> the bulk. Gorodtsov <strong>and</strong><br />

Leonov [27] analyzed the plane Couette flow <strong>of</strong> a UCM fluid <strong>and</strong> found that, while<br />

always stable <strong>in</strong> the absence <strong>of</strong> <strong>in</strong>ertia, the slowest decay<strong>in</strong>g modes are localized<br />

near the bound<strong>in</strong>g surfaces. As a result, one would expect the formation <strong>of</strong> <strong>boundary</strong><br />

localized structures due to r<strong>and</strong>om concentration fluctuations <strong>in</strong> the <strong>polymer</strong><br />

solution. In addition, <strong>slip</strong> leads to <strong>boundary</strong>-localized <strong>in</strong>stability <strong>in</strong> simple <strong>shear</strong><br />

<strong>flows</strong>, as shown <strong>in</strong> Ch. 3. Taken <strong>in</strong> whole, the experiments suggest that the basic<br />

HF mechanism, while be<strong>in</strong>g able to expla<strong>in</strong> bulk behavior, only partially accounts<br />

for the physics lead<strong>in</strong>g to enhancement <strong>of</strong> fluctuations near surfaces. The goal <strong>of</strong>


72<br />

the next chapter is to analyze the coupl<strong>in</strong>g between concentration, stress, <strong>and</strong> <strong>slip</strong><br />

to elucidate the <strong>effects</strong> <strong>of</strong> boundaries on the dynamics.


73<br />

Chapter 5<br />

Concentration Fluctuations <strong>and</strong> Flow<br />

Instabilities <strong>in</strong> Sheared Polymer<br />

Solutions †<br />

5.1 Formulation<br />

The experimental results outl<strong>in</strong>ed <strong>in</strong> Ch. 4 clearly po<strong>in</strong>t to the importance <strong>of</strong><br />

<strong>polymer</strong>-surface <strong>in</strong>teractions <strong>in</strong> determ<strong>in</strong><strong>in</strong>g the evolution <strong>of</strong> the flow <strong>of</strong> <strong>polymer</strong><br />

solutions. In particular, the experiments <strong>of</strong> Mhetar <strong>and</strong> Archer [56] explicitly l<strong>in</strong>k<br />

<strong>wall</strong> <strong>slip</strong> <strong>and</strong> the formation <strong>of</strong> enhanced fluctuations <strong>in</strong> simple <strong>shear</strong>, by demonstrat<strong>in</strong>g<br />

that: 1) substantial <strong>slip</strong> takes place at <strong>shear</strong> stresses <strong>and</strong> rates below those for<br />

which enhanced fluctuations appear; 2) fluctuations are <strong>in</strong>itiated near the surfaces;<br />

<strong>and</strong> 3) surface treatments can delay the onset <strong>of</strong> enhanced fluctuations to larger flow<br />

† The results <strong>in</strong> this chapter have been submitted for publication <strong>in</strong> Phys. Rev. Lett.


74<br />

01<br />

01<br />

y 01<br />

01<br />

z 01<br />

01<br />

000 111x<br />

u<br />

01<br />

b<br />

*<br />

s *<br />

γ . *<br />

l *<br />

Figure 29: Plane Couette geometry show<strong>in</strong>g <strong>slip</strong> between the solution <strong>and</strong> solid<br />

surface. The flow is <strong>in</strong> the x-direction, y is the gradient direction <strong>and</strong> z is the<br />

neutral, or vorticity, direction. l ∗ is the gap width, ū ∗ s is the <strong>slip</strong> velocity, ˙γ ∗ is the<br />

true <strong>shear</strong> rate <strong>in</strong> the fluid, <strong>and</strong> b ∗ =ū ∗ s / ˙γ∗ is the extrapolation length.<br />

rates. These facts constitute the primary motivation for the analyses presented <strong>in</strong><br />

this chapter.<br />

Given that the majority <strong>of</strong> the experiments were performed <strong>in</strong> simple <strong>shear</strong><br />

geometries, the l<strong>in</strong>k between <strong>slip</strong> <strong>and</strong> the formation <strong>of</strong> enhanced concentration fluctuations<br />

is studied <strong>in</strong> the plane Couette geometry shown <strong>in</strong> Fig. 29. The x-direction<br />

is the flow direction, the y-direction is the gradient direction, <strong>and</strong> the z-direction<br />

is the neutral, or vorticity, direction. The constitutive behavior <strong>of</strong> the solution is<br />

described by a two-fluid theory derived by Beris <strong>and</strong> Mavrantzas [54] that has been<br />

specialized to Hookean dumbbells [9]. Nondimensionaliz<strong>in</strong>g lengths with the characteristic<br />

length √ D tr λ, the stresses with n ∗ 0 kT, wheren∗ 0 is the average concentration<br />

<strong>of</strong> the solution, the velocities with √ D tr /λ, time with λ, <strong>and</strong> concentration with n ∗ 0


75<br />

gives the follow<strong>in</strong>g model equations<br />

−∇p + ∇·τ + S ∇ 2 v =0,<br />

τ (1) + τ −∇ 2 (τ + nδ)+δ Dn<br />

Dt = n ˙γ,<br />

Dn<br />

Dt = ∇2 n −∇∇: τ ,<br />

∇·v =0,<br />

(43a)<br />

(43b)<br />

(43c)<br />

(43d)<br />

where τ is the <strong>polymer</strong> extra stress tensor, n is the <strong>polymer</strong> concentration, λ is<br />

the <strong>polymer</strong> relaxation time, D tr is the translational diffusivity <strong>of</strong> a dumbbell, <strong>and</strong><br />

S = η s /η p = η s /(n ∗ 0kTλ) is the ratio <strong>of</strong> the solvent viscosity to the <strong>polymer</strong> viscosity<br />

at the average concentration <strong>of</strong> the solution, n ∗ 0. The true <strong>shear</strong> rate <strong>in</strong> the fluid has<br />

been used for the nondimensionalization, as opposed to the applied <strong>shear</strong> rate used<br />

<strong>in</strong> Ch. 3 for melts, to take advantage <strong>of</strong> the fact that the master curve reported<br />

<strong>in</strong> Fig. 22 <strong>in</strong> §3.4.3 relates the true Weissenberg number to the wavenumber <strong>and</strong><br />

to facilitate the solution technique outl<strong>in</strong>ed below.<br />

Eq. 43b is the <strong>polymer</strong> constitutive<br />

equation <strong>and</strong> is a generalization <strong>of</strong> the upper convected Maxwell (UCM)<br />

equation [10] which <strong>in</strong>cludes coupl<strong>in</strong>g between the concentration <strong>and</strong> stress as well<br />

as a stress diffusion term. In most cases, this term is negligible [9], but here, the<br />

flow structures <strong>of</strong> <strong>in</strong>terest have a length scale set by the diffusivity. Eq. 43c describes<br />

the <strong>polymer</strong> concentration <strong>and</strong> explicitly <strong>in</strong>cludes the coupl<strong>in</strong>g <strong>of</strong> stress <strong>and</strong><br />

concentration gradients. Eqs. 43a <strong>and</strong> 43d describe conservation <strong>of</strong> momentum <strong>and</strong><br />

mass. This model is similar to those studied previously [36, 59, 60, 41], although<br />

only Sun et al. [91] have <strong>in</strong>cluded the diffusive term <strong>in</strong> Eq. 43b. Furthermore, all<br />

earlier works have restricted themselves to bulk behavior <strong>and</strong> We(≡ λ ˙γ ∗ ) < 1.


76<br />

Specification <strong>of</strong> suitable <strong>boundary</strong> conditions for the velocity, stress, <strong>and</strong> concentration<br />

completes the model.<br />

The concentration satisfies the typical no-flux<br />

<strong>boundary</strong> condition at the surfaces,<br />

n · j n = ∂n<br />

∂y − ∂τ yx<br />

∂x − ∂τ yy<br />

∂y − ∂τ zy<br />

∂z<br />

=0, (44)<br />

where n is the outward unit normal, which states that mass cannot be lost through<br />

the boundaries. The presence <strong>of</strong> the stress diffusion term <strong>in</strong> the constitutive equation<br />

necessitates a stress <strong>boundary</strong> condition, which can be derived by rewrit<strong>in</strong>g Eq. 43b<br />

can be rewritten <strong>in</strong> terms <strong>of</strong> the conformation tensor us<strong>in</strong>g the Kramers’ form <strong>of</strong><br />

the stress tensor [10], τ = 〈QQ〉−nδ, as<br />

∂〈QQ〉<br />

∂t<br />

= −∇ · j QQ + 〈QQ〉·∇v +(∇v) T ·〈QQ〉−〈QQ〉 + nδ (45)<br />

where Q is the orientation vector <strong>of</strong> a dumbbell, j QQ = v ·〈QQ〉−∇〈QQ〉 is the<br />

conformation flux, <strong>and</strong> the rema<strong>in</strong>der <strong>of</strong> the terms on the right h<strong>and</strong> side can be<br />

thought <strong>of</strong> as volume source terms. Assum<strong>in</strong>g that there are no sources or s<strong>in</strong>ks for<br />

the conformation tensor at the <strong>boundary</strong> (i.e. the <strong>wall</strong>s are neutral) gives a no-flux<br />

<strong>boundary</strong> condition for the conformation tensor:<br />

∂〈QQ〉<br />

∂y<br />

= ∂τ<br />

∂y + ∂n δ =0. (46)<br />

∂y<br />

All <strong>of</strong> the computations reported below use this <strong>boundary</strong> condition. Computations<br />

performed with the classical UCM equation as the stress <strong>boundary</strong> condition [92]<br />

showed no qualitative difference <strong>in</strong> the results. F<strong>in</strong>ally, the last pieces to be specified<br />

are the velocity <strong>boundary</strong> conditions. The normal component <strong>of</strong> velocity is given<br />

by the no-penetration condition, v = 0, at both surfaces <strong>and</strong> the tangential velocity<br />

component is given by a <strong>slip</strong> <strong>boundary</strong> condition at the <strong>in</strong>terface.<br />

In Ch. 3 it


77<br />

was shown that a <strong>slip</strong> model where the <strong>slip</strong> velocity depends on normal stresses<br />

is required <strong>in</strong> order to f<strong>in</strong>d <strong>in</strong>stability <strong>in</strong> <strong>polymer</strong> melts.<br />

Even though the <strong>slip</strong><br />

mechanisms <strong>in</strong> entangled <strong>polymer</strong> solutions are expected to be similar to those <strong>in</strong><br />

melts, a Navier <strong>slip</strong> <strong>boundary</strong> condition is used here, where the <strong>slip</strong> velocity is<br />

proportional to the <strong>shear</strong> stress, u s = ɛτ yx (= b at steady state). This serves as a<br />

po<strong>in</strong>t <strong>of</strong> reference for comparisons to the melt results <strong>and</strong> a prelude to <strong>in</strong>clud<strong>in</strong>g<br />

more complicated <strong>slip</strong> relations.<br />

5.2 Stability Analysis<br />

The dynamics <strong>of</strong> concentration fluctuations were studied <strong>in</strong>itially by employ<strong>in</strong>g a<br />

l<strong>in</strong>ear stability analysis similar to that <strong>in</strong> Ch. 3.<br />

The steady state solution to<br />

Eqs. 43a -43d us<strong>in</strong>g Eq. 46 as the <strong>boundary</strong> condition is one dimensional <strong>and</strong> given<br />

by<br />

ā = {ū, ¯v, ¯w, ¯τ xx , ¯τ yx , ¯τ zx , ¯τ yy , ¯τ zy , ¯τ zz , ¯n, p} (47)<br />

= { We(y +ū s ), 0, 0, 2We 2 ,We,0, 0, 0, 0, 1, 1 } ,<br />

with y ∈ [0, √ Pe/We], where the Péclet number Pe =˙γ ∗ l ∗2 /D tr . Small perturbations<br />

are added to this base solution, a = ā(y)+ã(x, y, z, t), <strong>and</strong> the basic equations<br />

are l<strong>in</strong>earized with only the terms lead<strong>in</strong>g order <strong>in</strong> the perturbation reta<strong>in</strong>ed. The<br />

result<strong>in</strong>g set <strong>of</strong> 11 equations can be written <strong>in</strong> matrix notation as<br />

C S ∂ã<br />

∂t = −LS ã (48)<br />

where C S is a constant matrix with zero rows correspond<strong>in</strong>g to the momentum <strong>and</strong><br />

constitutive equations <strong>and</strong> L S is a cont<strong>in</strong>uous spatial differential operator. These


78<br />

operators are shown for completeness <strong>in</strong> Appendix B. The perturbations take the<br />

normal mode form<br />

ã(x, y, z, t) =â(y)e ik · x e −iσt + c.c. (49)<br />

where k =(k x , 0,k z ) is the wavevector <strong>of</strong> the disturbance <strong>in</strong> the x-z plane. Substitution<br />

<strong>of</strong> Eq. 49 <strong>in</strong>to Eq. 48 yields a generalized eigenvalue problem for the eigenvalues<br />

{σ} with associated eigenfunctions {â(y)}. The eigenvalues are, <strong>in</strong> general, complex<br />

<strong>and</strong> if Im(σ) > 0 then disturbances will grow <strong>and</strong> the flow is unstable. Three<br />

dimensional disturbances are considered because Squire’s theorem has not been<br />

demonstrated for this model.<br />

5.2.1 Numerical Method<br />

The Chebyshev collocation technique outl<strong>in</strong>ed <strong>in</strong> §3.4.2 is <strong>in</strong>sufficient for this problem<br />

for several reasons. First, the streamfunction formulation becomes extremely<br />

stiff because the solvent term <strong>in</strong> the equation <strong>of</strong> motion leads to fourth derivatives<br />

<strong>of</strong> the stream function (c.f. §3.4.2). Second, the concentration equation leads to<br />

an additional strip <strong>of</strong> cont<strong>in</strong>uous spectrum, when D tr = 0, which <strong>in</strong>terferes with<br />

the calculation <strong>of</strong> the discrete modes. For D tr ≠ 0, the govern<strong>in</strong>g equations are no<br />

longer s<strong>in</strong>gular <strong>and</strong> all <strong>of</strong> the eigenvalues are discrete. However, there are sets <strong>of</strong><br />

discrete eigenvalues which correspond to the cont<strong>in</strong>uous modes <strong>of</strong> the D tr =0case<br />

<strong>and</strong> are similarly difficult to resolve. This will be more clear below.<br />

A modified Chebyshev collocation scheme, employ<strong>in</strong>g a three dimensional, primitive<br />

variable formulation, was developed to solve the generalized eigenvalue problem<br />

result<strong>in</strong>g from Eq. 48. The results <strong>in</strong> §3.4.3 show that, for the case <strong>of</strong> the PTT fluid


79<br />

(<strong>and</strong>, therefore, the UCM fluid), the eigenfunctions are localized near the boundaries<br />

<strong>and</strong> decay away from the surfaces for large wavenumbers. It was also shown that<br />

accurate approximations were obta<strong>in</strong>ed analytically by reta<strong>in</strong><strong>in</strong>g only the decay<strong>in</strong>g<br />

solutions to the general stability equation. This can be accomplished numerically by<br />

solv<strong>in</strong>g the the govern<strong>in</strong>g equations on a semi-<strong>in</strong>f<strong>in</strong>ite doma<strong>in</strong> (SID) <strong>and</strong> requir<strong>in</strong>g<br />

the eigenfunctions to decay to zero as y →∞.Wemapy onto the computational<br />

coord<strong>in</strong>ate, ξ, us<strong>in</strong>g[17,15]<br />

y = L<br />

( ) 1+ξ<br />

, (50)<br />

1 − ξ<br />

with {y, ξ |y ∈ [0, ∞),ξ ∈ [−1, 1]}. As discussed by Canuto et al. [17], this mapp<strong>in</strong>g<br />

is more robust than an exponential mapp<strong>in</strong>g <strong>and</strong> less sensitive to the value <strong>of</strong> the<br />

map parameter, L. The map parameter is a measure <strong>of</strong> the degree <strong>of</strong> stretch<strong>in</strong>g <strong>of</strong><br />

the coord<strong>in</strong>ates, i.e., itisthevalue<strong>of</strong>y which is mapped to the center <strong>of</strong> the computational<br />

doma<strong>in</strong>. Trial <strong>and</strong> error yielded an optimum value <strong>of</strong> the map parameter<br />

<strong>of</strong> L =0.1 for this problem. All <strong>of</strong> the results presented below use this value. This<br />

technique is accurate for disturbances with wavenumber much larger than the reciprocal<br />

<strong>of</strong> the gap width, kx ∗ ≫ 1/l ∗ , as for the analytical solutions <strong>of</strong> §3.4.1, a<br />

condition satisfied by the dom<strong>in</strong>ant wavenumber measured experimentally [56].<br />

Spurious pressure modes <strong>in</strong> primitive variable formulations are well known <strong>and</strong><br />

described <strong>in</strong> detail by Canuto et al. [17]. These modes can be elim<strong>in</strong>ated by exp<strong>and</strong><strong>in</strong>g<br />

pressure on a lower order grid. To balance the number <strong>of</strong> equations <strong>and</strong><br />

unknowns, the cont<strong>in</strong>uity equation is evaluated on the staggered grid. This requires<br />

extrapolation <strong>of</strong> the pressure from the staggered to regular grid for evaluation <strong>of</strong><br />

the momentum equation <strong>and</strong> <strong>in</strong>terpolation <strong>of</strong> the velocity components from the


80<br />

regular to staggered grid for the evaluation <strong>of</strong> the cont<strong>in</strong>uity equation. The extrapolation<br />

formula is derived from the expansion for pressure <strong>in</strong> terms <strong>of</strong> Chebyshev<br />

polynomials is<br />

p(ξ) =<br />

with the expansion coefficients given by<br />

where<br />

b i = 2<br />

πc i<br />

∫ 1<br />

−1<br />

N−1<br />

∑<br />

i=0<br />

b i T i (ξ) (51)<br />

p(ξ)T i (ξ)w(ξ) dξ, (52)<br />

⎧<br />

⎪⎨ 2 j =0<br />

c j =<br />

⎪⎩ 1 j ≥ 1<br />

The staggered grid po<strong>in</strong>ts, given by, ξ s i =cos<br />

(<br />

(i+<br />

1<br />

2 )π<br />

N<br />

. (53)<br />

)<br />

,withi =0,... ,N−1, are all<br />

<strong>in</strong> the <strong>in</strong>terior <strong>of</strong> the doma<strong>in</strong>, so this <strong>in</strong>tegral can be evaluated by Gauss quadrature<br />

to give<br />

b i = 2<br />

∑<br />

( j(k +<br />

1<br />

p(ξk s )cos )π )<br />

2<br />

. (54)<br />

N<br />

Nc i<br />

N−1<br />

k=0<br />

This can be put back <strong>in</strong>to the expansion for p <strong>and</strong> this polynomial evaluated at the<br />

regular grid po<strong>in</strong>ts to get<br />

p(ξ i )= 2 N<br />

∑ ∑<br />

N−1 N−1<br />

j=0<br />

k=0<br />

1<br />

c j<br />

p(ξ s k)cos<br />

( j(k +<br />

1<br />

)π 2<br />

N<br />

)<br />

cos<br />

( ) ijπ<br />

N<br />

(55)<br />

This gives the values <strong>of</strong> pressure at the regular grid po<strong>in</strong>ts <strong>in</strong> terms <strong>of</strong> the pressure<br />

values at the staggered grid po<strong>in</strong>ts.<br />

To get the <strong>in</strong>terpolation function for<br />

the velocities, we start with the expansion <strong>in</strong> terms <strong>of</strong> Chebyshev polynomials,


81<br />

v = ∑ N<br />

i=0 a iT i (ξ), <strong>and</strong> solve for the expansion coefficients us<strong>in</strong>g Gauss-Lobatto<br />

quadrature on the regular grid, s<strong>in</strong>ce the doma<strong>in</strong> <strong>in</strong>cludes the <strong>boundary</strong> po<strong>in</strong>ts,<br />

a i = 2<br />

Nc i<br />

N<br />

∑<br />

j=0<br />

( )<br />

1 jkπ<br />

v(ξ j )cos , (56)<br />

¯c j N<br />

where<br />

⎧<br />

⎪⎨ 2 j =0,N<br />

¯c j =<br />

⎪⎩ 1 1 ≤ j ≤ N − 1<br />

. (57)<br />

Putt<strong>in</strong>g this back <strong>in</strong>to the expansion <strong>and</strong> evaluat<strong>in</strong>g it at the staggered grid po<strong>in</strong>ts<br />

gives the <strong>in</strong>terpolation function<br />

v(ξ s i )= 2 N<br />

N∑<br />

N∑<br />

k=0 j=0<br />

1<br />

c k¯c j<br />

v(ξ j )cos<br />

( k(i +<br />

1<br />

)π 2<br />

N<br />

)<br />

cos<br />

( ) kjπ<br />

, (58)<br />

N<br />

Eqs. 55 <strong>and</strong> 58 can be written <strong>in</strong> matrix form <strong>and</strong> are implemented via matrix<br />

multiplication, as was done previously with the Chebyshev derivative operator.<br />

Code Verification<br />

The code was benchmarked aga<strong>in</strong>st 2D results by sett<strong>in</strong>g k z = 0 <strong>and</strong> dropp<strong>in</strong>g the<br />

stress diffusion terms <strong>in</strong> the constitutive equation. L<strong>in</strong>eariz<strong>in</strong>g around the steady<br />

state solution <strong>and</strong> <strong>in</strong>troduc<strong>in</strong>g the normal mode form given <strong>in</strong> Eq. 49 gives the


82<br />

follow<strong>in</strong>g reduced system <strong>of</strong> equations for 2D disturbances<br />

−ik x ˆp + ik xˆτ xx +ˆτ yx ′ + S ( −kxû 2 +û′′) =0<br />

−ˆp ′ + ik xˆτ yx +ˆτ yy ′ + S ( −k xˆv 2 +ˆv′′) =0<br />

Qˆτ xx − 2ik x (¯τ xx +1)û − 2¯τ yx û ′ − (iσ − ik x ū)ˆn − 2ˆτ yx =0<br />

Qˆτ yx − ik x (¯τ xx +1)ˆv − û ′ − ˆn − ˆτ yy =0<br />

Qˆτ yy − 2ik x¯τ yxˆv − 2ˆv ′ − (iσ − ik x ū)ˆn =0<br />

(59a)<br />

(59b)<br />

(59c)<br />

(59d)<br />

(59e)<br />

−ˆn ′′ +(−iσ + ik x ū + k 2 x)ˆn + [ −k 2 xˆτ xx +2ik xˆτ ′ yx +ˆτ ′′<br />

yy]<br />

=0 (59f )<br />

ik x û +ˆv ′ =0<br />

(59g)<br />

where Q =1−iσ + ik x ū <strong>and</strong> ū = We(y +ū s ). The constitutive equations, Eqs. 59c -<br />

59e can be solved for the stress components <strong>in</strong> terms <strong>of</strong> velocities <strong>and</strong> ˆn. After elim<strong>in</strong>at<strong>in</strong>g<br />

pressure from the momentum equations, Eqs. 59a <strong>and</strong> 59b, these expressions<br />

for the stresses can be put <strong>in</strong>to the momentum equation to give an equation relat<strong>in</strong>g<br />

the velocities <strong>and</strong> concentration. After us<strong>in</strong>g the cont<strong>in</strong>uity equation to elim<strong>in</strong>ate<br />

û <strong>in</strong> favor <strong>of</strong> ˆv, a general stability equation can be derived, which is<br />

[Q 2 D 2 − k 2 xQ 2 − 2ik x QD − 2k 2 x][D 2 +2ik x¯τ yx D − k 2 x(¯τ xx +1)]ˆv+<br />

SQ 3 (D 2 − k 2 x )2ˆv − ik x [Q 2 D 2 − 2ik x D + k 2 x Q2 ]ˆn =0.<br />

(60)<br />

This equation conta<strong>in</strong>s three pieces: 1) the first operator is the <strong>polymer</strong> (UCM)<br />

term, 2) the second term is the solvent term, <strong>and</strong> 3) the last term is a concentration<br />

term due to the coupl<strong>in</strong>g between stress <strong>and</strong> concentration. It is not clear from<br />

this equation where the cont<strong>in</strong>uous spectra should be found, as the concentration<br />

dependence <strong>of</strong> the velocities is unknown. So, Eq. 59f , is simplified by substitut<strong>in</strong>g


83<br />

for the stresses to give<br />

[ ( )] σ − kx ū<br />

1 − i<br />

ˆn ′′ +<br />

Q<br />

[<br />

[<br />

i(σ − k x ū) − kx<br />

2 1 − i<br />

( )]]<br />

σ − kx ū<br />

ˆn = 0 (61)<br />

Q<br />

This equation is decoupled from the stability equation, due to cancelation <strong>of</strong> the<br />

stress terms. One set <strong>of</strong> cont<strong>in</strong>uous eigenvalues will be where this equation is s<strong>in</strong>gular,<br />

i.e.,<br />

σ = k x (y +ū s ) − i 2<br />

(62)<br />

S<strong>in</strong>ce ˆn can be determ<strong>in</strong>ed from Eq. 61, the stability equation is just the Oldroyd-<br />

B stability equation with forc<strong>in</strong>g. Therefore, we also expect strips <strong>of</strong> cont<strong>in</strong>uous<br />

spectra where the Oldroyd-B stability problem is s<strong>in</strong>gular [43, 103], specifically<br />

⎧<br />

⎪⎨ k x (y +ū s ) − i<br />

σ =<br />

. (63)<br />

⎪⎩ k x (y +ū s ) − i(S+1)<br />

S<br />

All three strips <strong>of</strong> cont<strong>in</strong>uous spectra are stable. Theoretically, we could also determ<strong>in</strong>e<br />

some <strong>of</strong> the discrete eigenvalues by solv<strong>in</strong>g Eq. 61 <strong>and</strong> apply<strong>in</strong>g no-flux<br />

<strong>boundary</strong> conditions. However, there is no fortuitous cancelation after substitut<strong>in</strong>g<br />

for the stresses <strong>in</strong> the <strong>boundary</strong> condition <strong>and</strong> consequently, the no-flux <strong>boundary</strong><br />

condition conta<strong>in</strong>s velocities which must be determ<strong>in</strong>ed by solv<strong>in</strong>g the stability<br />

equation. It would appear then that all <strong>of</strong> the discrete eigenvalues are affected by the<br />

coupl<strong>in</strong>g between velocity <strong>and</strong> concentration <strong>and</strong> obta<strong>in</strong><strong>in</strong>g analytical expressions<br />

for any <strong>of</strong> them would be difficult. Nevertheless, the locations <strong>of</strong> the cont<strong>in</strong>uous<br />

spectra are useful checks on the code.<br />

Fig. 30 shows the eigenvalue spectrum for 2D disturbances with the stress diffusion<br />

terms dropped. Three strips <strong>of</strong> cont<strong>in</strong>uous spectrum exist <strong>and</strong> their locations


84<br />

40.0<br />

Inset<br />

Im(σ)<br />

−20.0<br />

−80.0<br />

−140.0<br />

−200.0<br />

−260.0<br />

−320.0<br />

0.5<br />

0.0<br />

−0.5<br />

−1.0<br />

−1.5<br />

−2.0<br />

−2.5<br />

Slip mode<br />

Slip mode<br />

Concentration mode<br />

G−L mode<br />

−3.0<br />

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0<br />

0.0 4.0 8.0 12.0 16.0 20.0 24.0 28.0 32.0<br />

Re(σ)<br />

Figure 30: Eigenvalue spectrum for 2D disturbances with the stress diffusion term<br />

dropped. The parameters are k x =0.3467, We =11.9,b=0.1442,S =0.01,N = 96.<br />

The eigenvalues which determ<strong>in</strong>e the stability are designated by circles <strong>and</strong> are on<br />

the left side <strong>of</strong> the spectrum. The “tail” <strong>of</strong> eigenvalues extend<strong>in</strong>g to the right are<br />

nearly cont<strong>in</strong>uous modes which are not resolved.<br />

are <strong>in</strong> excellent agreement with the above predictions. There are 4 discrete modes<br />

near the strip <strong>of</strong> cont<strong>in</strong>uous spectra due to the concentration equation: two <strong>slip</strong><br />

modes; a mode correspond<strong>in</strong>g to the classical Gorodtsov-Leonov eigenvalue; <strong>and</strong> a<br />

concentration mode. Note that the cont<strong>in</strong>uous strip due to the concentration equation<br />

is less stable than its UCM counterpart, <strong>and</strong> is therefore expected to <strong>in</strong>terfere<br />

more strongly with the calculation <strong>of</strong> the discrete modes lead<strong>in</strong>g to <strong>in</strong>stability.<br />

The SID technique is very effective at resolv<strong>in</strong>g the discrete <strong>and</strong> cont<strong>in</strong>uous<br />

modes, as shown <strong>in</strong> Fig. 31, which compares the SID technique with the st<strong>and</strong>ard<br />

collocation method on a bounded doma<strong>in</strong>, with focus on the portion <strong>of</strong> the spectrum<br />

near the ends <strong>of</strong> the UCM <strong>and</strong> concentration strips <strong>of</strong> cont<strong>in</strong>uous spectra. Only the<br />

<strong>slip</strong> mode is observed for the st<strong>and</strong>ard technique with N = 192; the concentration<br />

mode is absorbed <strong>in</strong>to the cont<strong>in</strong>uous strip due to the concentration equation. N<br />

must be <strong>in</strong>creased to 280 before the second mode is resolved. The SID technique


85<br />

0.5<br />

0.0<br />

Slip mode<br />

Concentration mode<br />

Im(σ)<br />

−0.5<br />

−1.0<br />

−1.5<br />

2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0<br />

Re(σ)<br />

Figure 31: Comparison <strong>of</strong> the st<strong>and</strong>ard <strong>and</strong> SID techniques. The figure does not<br />

show the entire spectrum; <strong>in</strong>stead, the focus is on the least stable portion. The<br />

parameters are: We = 19.64, k x = 10, b = 0.02, Pe = 2000. 3 – St<strong>and</strong>ard:<br />

N = 192, ○ – St<strong>and</strong>ard: N = 280, 2 –SID:N = 192, △ –SID:N = 96, ▽ –<br />

DD: N b = 75, N c = 43. Note that Pe has no mean<strong>in</strong>g for the SID simulations <strong>and</strong><br />

simply sets the location <strong>of</strong> the top plate for the rema<strong>in</strong><strong>in</strong>g techniques. The stress<br />

diffusion terms have dropped as described <strong>in</strong> the ma<strong>in</strong> text.<br />

with N = 192 clearly resolves the mode <strong>and</strong> is superior to the st<strong>and</strong>ard scheme with<br />

N = 280. In fact, the SID scheme with N = 96 is capable <strong>of</strong> resolv<strong>in</strong>g the mode<br />

as well as the st<strong>and</strong>ard scheme with N = 280. This gives an enormous sav<strong>in</strong>gs <strong>in</strong><br />

computer time <strong>and</strong> storage.<br />

Also shown are the results for a doma<strong>in</strong> decomposition (DD) technique. The<br />

key to improv<strong>in</strong>g the resolution <strong>of</strong> the spectrum near the endpo<strong>in</strong>ts is improv<strong>in</strong>g<br />

the underly<strong>in</strong>g spatial discretization <strong>of</strong> the govern<strong>in</strong>g equations near the boundaries.<br />

To do this without <strong>in</strong>creas<strong>in</strong>g the overall number <strong>of</strong> collocation po<strong>in</strong>ts, the<br />

doma<strong>in</strong> is broken up <strong>in</strong>to three subdoma<strong>in</strong>s, one near each <strong>boundary</strong> <strong>and</strong> one <strong>in</strong> the<br />

center <strong>of</strong> the channel <strong>in</strong> which the number <strong>of</strong> collocation po<strong>in</strong>ts can be varied <strong>in</strong>dependently.<br />

The <strong>boundary</strong> doma<strong>in</strong>s are much smaller than the central doma<strong>in</strong>, <strong>in</strong><br />

fact, the <strong>boundary</strong> doma<strong>in</strong> size scales as 1/k x , s<strong>in</strong>ce this is the size <strong>of</strong> the <strong>boundary</strong>


86<br />

layer that the perturbations are conf<strong>in</strong>ed to. Match<strong>in</strong>g conditions are applied at the<br />

<strong>in</strong>tersection <strong>of</strong> each doma<strong>in</strong>. Souvaliotis <strong>and</strong> Beris [90] reported only need<strong>in</strong>g C 0<br />

cont<strong>in</strong>uity (piecewise cont<strong>in</strong>uity) between doma<strong>in</strong>s for their flow calculations, but<br />

here C 1 cont<strong>in</strong>uity (cont<strong>in</strong>uous first derivatives) for each variable had to be imposed<br />

to get mean<strong>in</strong>gful spectra. Fig. 31 also shows the spectrum for doma<strong>in</strong> decomposition<br />

for <strong>boundary</strong> doma<strong>in</strong>s with N b = 75 for the number <strong>of</strong> collocation po<strong>in</strong>ts,<br />

<strong>and</strong> a central doma<strong>in</strong> with N c = 43 for the number <strong>of</strong> collocation po<strong>in</strong>ts, giv<strong>in</strong>g a<br />

total number <strong>of</strong> 193 collocation po<strong>in</strong>ts (which is the same as a s<strong>in</strong>gle doma<strong>in</strong> with<br />

N = 192). The resolution <strong>of</strong> the concentration mode is about the same as for the<br />

s<strong>in</strong>gle doma<strong>in</strong> problem with N = 280, so this technique is an improvement <strong>in</strong> that it<br />

does save time <strong>and</strong> storage over the st<strong>and</strong>ard scheme, but not nearly as much as the<br />

SID technique. Note that the doma<strong>in</strong> decomposition technique could be comb<strong>in</strong>ed<br />

with the SID technique, but the improvement is expected to be slight at best, as<br />

essentially all <strong>of</strong> the collocation po<strong>in</strong>ts <strong>in</strong> the SID technique are already resolv<strong>in</strong>g<br />

the <strong>boundary</strong> layer.<br />

5.2.2 Stability Results<br />

Fig. 32 shows a portion <strong>of</strong> a typical eigenvalue spectrum obta<strong>in</strong>ed us<strong>in</strong>g the SID<br />

technique.<br />

The “tail” <strong>of</strong> eigenvalues extend<strong>in</strong>g to the right is composed <strong>of</strong> the<br />

nearly cont<strong>in</strong>uous modes mentioned above. These modes are not mesh-resolved at<br />

N = 96, <strong>and</strong> <strong>in</strong> fact, part <strong>of</strong> the tail has Re(σ) > 0, spuriously <strong>in</strong>dicat<strong>in</strong>g <strong>in</strong>stability.<br />

The modes on the left designated with circles are the modes which determ<strong>in</strong>e the<br />

stability <strong>and</strong> are resolved for this number <strong>of</strong> collocation po<strong>in</strong>ts.<br />

The coupl<strong>in</strong>g <strong>of</strong> <strong>slip</strong> to concentration <strong>and</strong> stress leads to novel hydrodynamic


87<br />

15.0<br />

10.0<br />

5.0<br />

Im(σ)<br />

0.0<br />

−5.0<br />

−10.0<br />

−15.0<br />

−20.0<br />

0.0 100.0 200.0 300.0 400.0 500.0<br />

Re(σ)<br />

Figure 32: A typical eigenvalue obta<strong>in</strong>ed us<strong>in</strong>g the SID technique. The parameters<br />

are We = 10, k x =1,b =1,S =0.01, <strong>and</strong> N = 96.<br />

<strong>in</strong>stabilities <strong>in</strong> plane Couette flow. Note that if D tr is set to zero <strong>in</strong> the dimensional<br />

equations, the model reduces to the Oldroyd-B model, <strong>and</strong> plane Couette flow with<br />

either the no-<strong>slip</strong> [27] or Navier <strong>slip</strong> [66, 13, 14] <strong>boundary</strong> condition is stable for<br />

all Weissenberg numbers. Also, if concentration variations are permitted but <strong>slip</strong><br />

is not, by sett<strong>in</strong>g b = 0, the flow is aga<strong>in</strong> stable. Fig. 33 shows neutral curves with<br />

<strong>slip</strong>, plotted as We c vs. k x for two dimensional disturbances, k z =0,withb fixed.<br />

At low values <strong>of</strong> k x , power law behavior is observed. For b(= b ∗ / √ D tr λ)=1,the<br />

critical wavenumber is k x ≈ 0.8, so that the length scales for <strong>slip</strong> <strong>and</strong> the <strong>in</strong>stability<br />

are the same. Careful exam<strong>in</strong>ation reveals that at low wavenumbers (those most<br />

easily observed by microscopy), the critical We first decreases <strong>and</strong> then <strong>in</strong>creases<br />

with <strong>in</strong>creas<strong>in</strong>g b – the <strong>in</strong>creas<strong>in</strong>g section <strong>of</strong> the curve is <strong>in</strong> agreement with the<br />

observations <strong>of</strong> Mhetar <strong>and</strong> Archer [56] that treat<strong>in</strong>g the surface to <strong>in</strong>crease <strong>slip</strong><br />

delayed the onset <strong>of</strong> fluctuations. Note that there is a transition from one mode<br />

<strong>of</strong> <strong>in</strong>stability to another, lead<strong>in</strong>g to multiple m<strong>in</strong>ima <strong>in</strong> the neutral curve, as most<br />

clearly seen <strong>in</strong> the curve for b = 2.<br />

For larger values <strong>of</strong> b, the neutral curves


88<br />

100<br />

Unstable<br />

We c<br />

10<br />

b = 1<br />

b = 2<br />

b = 4<br />

b = 8<br />

Stable<br />

b = 10<br />

1<br />

10 −1 10 0 10 1<br />

Figure 33: Neutral curves for k z =0<strong>and</strong>S =10 −2 for various values <strong>of</strong> b. These<br />

curves were computed us<strong>in</strong>g the SID technique with N = 96.<br />

k x<br />

nearly lie on top <strong>of</strong> one another, due to localization <strong>of</strong> the waves with respect to<br />

the <strong>slip</strong> length. Typical three dimensional neutral curves are shown <strong>in</strong> Fig. 34 for<br />

both unstable modes, where clearly the most dangerous modes are two dimensional.<br />

Only two dimensional results are considered further.<br />

Fig. 35 shows a density plot <strong>of</strong> the unstable eigenvector for the concentration.<br />

The concentration perturbation is oriented <strong>in</strong> the same direction as enhanced fluctuations<br />

with<strong>in</strong> the HF mechanism described by Ji <strong>and</strong> Helf<strong>and</strong> [41]. In agreement<br />

with the observations <strong>of</strong> Mhetar <strong>and</strong> Archer [56], perturbations are localized near<br />

the surfaces <strong>and</strong> are observed at large We. As a result, one would expect scatter<strong>in</strong>g<br />

to be more pronounced <strong>in</strong> the <strong>boundary</strong> regions as opposed to the bulk. One<br />

concern is that for Rouse cha<strong>in</strong>s the characteristic length √ D tr λ is proportional to<br />

the end-to-end distance <strong>of</strong> the cha<strong>in</strong>, i.e., this is a molecular length scale. Note<br />

that the validity <strong>of</strong> the model becomes suspect for fluctuation sizes on the order<br />

<strong>of</strong> molecular lengths because it does not capture dynamics on these length scales.<br />

However, as the characteristic length is experimentally much larger than molecular


89<br />

We c<br />

11.0<br />

10.0<br />

Unstable<br />

b = 2 (Mode 1)<br />

b = 6 (Mode 2)<br />

9.0<br />

Stable<br />

8.0<br />

0.00 0.10 0.20 0.30 0.40 0.50<br />

k z<br />

Figure 34: Neutral curves for three dimensional disturbances at the given values <strong>of</strong><br />

k x <strong>and</strong> b. These curves were computed us<strong>in</strong>g the SID technique with N =96,S =<br />

10 −2 .<br />

length scales, the model is valid until k x ≫ 1.<br />

5.3 Brownian Fluctuations<br />

The stability results are suggestive, but do not address the dynamics <strong>of</strong> r<strong>and</strong>om,<br />

thermal fluctuations on the system. Typical theoretical calculations <strong>in</strong>volve determ<strong>in</strong><strong>in</strong>g<br />

the l<strong>in</strong>ear response <strong>of</strong> the system to r<strong>and</strong>om forc<strong>in</strong>g throughout the doma<strong>in</strong>,<br />

then comput<strong>in</strong>g the concentration correlation function. Instead <strong>of</strong> <strong>in</strong>troduc<strong>in</strong>g fluctuations<br />

<strong>in</strong> multiple directions, the normal mode form <strong>of</strong> the l<strong>in</strong>earized variables is<br />

used. This is equivalent to Fourier transform<strong>in</strong>g the govern<strong>in</strong>g equations <strong>in</strong> the x-<br />

<strong>and</strong> z-directions. The correlation function is then determ<strong>in</strong>ed for a given wavenumber<br />

<strong>in</strong> each direction. In addition, we only consider r<strong>and</strong>om concentration fluctuations<br />

<strong>and</strong> do not force the other variables. This assumption is not unprecedented,<br />

as Helf<strong>and</strong> <strong>and</strong> Fredrickson [36] only considered velocity <strong>and</strong> concentration fluctuations<br />

<strong>and</strong> neglected variations <strong>in</strong> stress, while Ji <strong>and</strong> Helf<strong>and</strong> [41] demonstrated that


90<br />

5<br />

4<br />

3<br />

y<br />

2<br />

1<br />

0<br />

0 Πk x 2Πk x 3Πk x 4Πk x<br />

x<br />

Figure 35: Unstable eigenfunction for the concentration. The parameters are k x =<br />

0.4, b = 10, We = 10, S =0.01, N = 96.<br />

velocity is a fast variable as well. The l<strong>in</strong>earized concentration equation becomes<br />

∂ˆn<br />

∂t + ¯v ·∇ˆn = ∇2ˆn −∇∇: ˆτ + ∇·w(x,t) (64)<br />

where w a the r<strong>and</strong>om component <strong>of</strong> the <strong>polymer</strong> mass flux satisfy<strong>in</strong>g [24]<br />

〈w(x,t) w(x ′ ,t ′ )〉 =2n ∗ 0(D tr λ) 3 2 δ δ(x − x ′ )δ(t − t ′ ) (65)<br />

〈w(x,t)〉 =0. (66)<br />

In order to satisfy conservation <strong>of</strong> mass, n · w = 0 on the boundaries, where n<br />

is the outward unit normal, <strong>and</strong> the forc<strong>in</strong>g was weighted appropriately so that<br />

the magnitude was uniform throughout the doma<strong>in</strong>. Time <strong>in</strong>tegration is performed<br />

us<strong>in</strong>g an implicit Euler scheme<br />

(<br />

C S +∆tL S) â n+1 = C S â n +∆t ∇·w n (x,t). (67)<br />

with arbitrary <strong>in</strong>itial condition. Here, ˆn(y, 0) = cos(6y), ˆτ (y, 0) = 0. Comparison <strong>of</strong><br />

the time <strong>in</strong>tegration results without noise with the SID stability results confirmed<br />

convergence as ∆t → 0.<br />

For accurate results, the time step had to be small;<br />

∆t =0.001 for all <strong>of</strong> the simulations below.


91<br />

Generally, time <strong>in</strong>tegration on a bounded doma<strong>in</strong> provided the best results.<br />

The SID technique is very good at approximat<strong>in</strong>g the eigenvalues near the surfaces,<br />

however, the price paid is very poor resolution <strong>of</strong> the cont<strong>in</strong>uous modes further away<br />

from the surface. These modes can be spuriously unstable, mak<strong>in</strong>g time <strong>in</strong>tegration<br />

impossible.<br />

To use the bounded doma<strong>in</strong>, the location <strong>of</strong> the top plate must be<br />

specified, <strong>and</strong> is chosen to satisfy the SID constra<strong>in</strong>t that the wavenumber be much<br />

larger than the reciprocal <strong>of</strong> the gap width, <strong>in</strong> dimensionless terms, k x ≫ √ We/Pe.<br />

The primary quantity <strong>of</strong> <strong>in</strong>terest is the spatial correlation function, def<strong>in</strong>ed <strong>in</strong><br />

general for a given x-wavenumber k x as:<br />

〈g(x; k x )g(x ′ ; k x )〉 =2Re〈ĝ(y; k x )ĝ ∗ (y ′ ; k x )〉 cos(k x (x − x ′ )) −<br />

2Im〈ĝ(y; k x )ĝ ∗ (y ′ ; k x )〉 s<strong>in</strong>(k x (x − x ′ )). (68)<br />

To confirm that the noise is be<strong>in</strong>g generated correctly, we first compute the spatial<br />

correlation for the forc<strong>in</strong>g by sett<strong>in</strong>g g(x; k x )=∇·w(x; k x )=f(x; k x ). Figs. 36(a)<br />

<strong>and</strong> 36(b) show the real <strong>and</strong> imag<strong>in</strong>ary parts <strong>of</strong> the correlation, respectively, <strong>and</strong><br />

demonstrate clearly that the noise is real <strong>and</strong> correlated <strong>in</strong> y as ∇ 2 δ(y − y ′ ), as<br />

anticipated. The jaggedness evident <strong>in</strong> Fig. 36(a) is a plott<strong>in</strong>g artifact due to the<br />

fact that the noise is evaluated at discrete po<strong>in</strong>ts <strong>in</strong> y.<br />

The scatter<strong>in</strong>g is determ<strong>in</strong>ed by the concentration spatial correlation function,<br />

g(x; k x )=ñ(x; k x ). Fig. 37 shows the concentration correlation function at equilibrium.<br />

Fluctuations are very strongly correlated near the surface <strong>and</strong> the magnitude<br />

drops <strong>of</strong>f to a value O(1) <strong>in</strong> the bulk.<br />

The imag<strong>in</strong>ary part <strong>of</strong> the correlation is<br />

zero, so that fluctuations decay without a directional bias, as is <strong>in</strong>tuitively expected<br />

when there is no <strong>shear</strong>. The length scale given by the width <strong>of</strong> the correlation is


92<br />

Re<br />

.04<br />

0.02<br />

0<br />

0<br />

25<br />

50<br />

y<br />

75<br />

(a)<br />

0<br />

100<br />

75<br />

50<br />

y’<br />

25<br />

Im<br />

5⋆10 19<br />

0<br />

5⋆10 19 0 25<br />

50<br />

y<br />

(b)<br />

75<br />

0<br />

100<br />

75<br />

50<br />

y’<br />

25<br />

Figure 36: The noise correlation function. (a) real part, (b) imag<strong>in</strong>ary part. The<br />

noise is δ correlated <strong>in</strong> y <strong>and</strong> the imag<strong>in</strong>ary part <strong>of</strong> the correlation is zero, confirm<strong>in</strong>g<br />

that the noise is be<strong>in</strong>g generated correctly.<br />

Re<br />

⋆10 5<br />

2⋆10 5<br />

0<br />

0<br />

1<br />

2<br />

y<br />

3<br />

4<br />

5 0 1<br />

2<br />

5<br />

4<br />

3<br />

y’<br />

Figure 37: Correlation function at equilibrium.<br />

O( √ D tr λ). Figure 38 shows a series <strong>of</strong> correlation functions for <strong>in</strong>creas<strong>in</strong>g Weissenberg<br />

number. The real part <strong>of</strong> the correlation rema<strong>in</strong>s essentially unchanged, but<br />

the imag<strong>in</strong>ary part is nonzero, <strong>in</strong>dicat<strong>in</strong>g that the fluctuations are enhanced <strong>in</strong> a<br />

particular direction. The direction <strong>of</strong> enhancement can be determ<strong>in</strong>ed by rewrit<strong>in</strong>g<br />

the correlation function as a s<strong>in</strong>gle s<strong>in</strong>usoid<br />

〈ñ(x; k x )ñ(x ′ ; k x )〉 = √ Re〈ˆn(y; k x )ˆn ∗ (y ′ ; k x )〉 2 +Im〈ˆn(y; k x )ˆn ∗ (y ′ ; k x )〉 2 ×<br />

s<strong>in</strong> [(x − x ′ ) − φ] , (69)


93<br />

where φ is the phase angle, def<strong>in</strong>ed as<br />

φ =tan −1 (−Im〈ˆn(y; k x )ˆn ∗ (y ′ ; k x )〉/Re〈ˆn(y; k x )ˆn ∗ (y ′ ; k x )〉) , (70)<br />

<strong>and</strong> comput<strong>in</strong>g the phase angles from the correlation functions. In all cases shown<br />

<strong>in</strong> Fig. 38, the real part <strong>of</strong> the correlation is positive, so it suffices to consider<br />

only the imag<strong>in</strong>ary part to get the general idea. Consider Fig.38(f), which is the<br />

imag<strong>in</strong>ary part <strong>of</strong> the concentration correlation for We = 5, <strong>and</strong> the <strong>boundary</strong> po<strong>in</strong>t<br />

y ′ =0. Forpo<strong>in</strong>tsy <strong>in</strong> the <strong>in</strong>terior, y>y<strong>and</strong> Im〈ˆn(y; k x )ˆn ∗ (y ′ ; k x )〉 is negative.<br />

This gives a positive phase shift, φ>0. Therefore, the angle <strong>of</strong> positive correlation<br />

is positive. The wavevector <strong>of</strong> the enhanced fluctuation is normal to the correlation<br />

angle <strong>and</strong> po<strong>in</strong>ts <strong>in</strong>to the fourth quadrant. This is <strong>in</strong>deed the structure observed<br />

if the concentration pr<strong>of</strong>ile is plotted, as shown <strong>in</strong> Fig. 39(c). In general, if the<br />

correlation is negative for y ′ >y, then the orientation is <strong>in</strong> the fourth quadrant<br />

<strong>and</strong> if the correlation is positive, then the orientation is <strong>in</strong> the first quadrant. For<br />

We =0.1, the orientation is <strong>in</strong> the first quadrant, except for perhaps a very small<br />

region near the surface <strong>and</strong> that for We = 1 the orientation changes away from<br />

the surface, from the first to the fourth quadrant.<br />

This <strong>in</strong>terpretation is borne<br />

out look<strong>in</strong>g at the concentration pr<strong>of</strong>iles <strong>in</strong> Fig. 39. Evidently, the wavevector <strong>of</strong><br />

the enhanced fluctuation rotates as the <strong>shear</strong> rate <strong>in</strong>creases. This rotation <strong>of</strong> the<br />

wavevector has been predicted previously for the bulk by Ji <strong>and</strong> Helf<strong>and</strong> [41] <strong>and</strong><br />

Milner [60] <strong>and</strong> observed experimentally by Wu et al. [105], also for the bulk. The<br />

results presented here suggest that fluctuations near the surface are enhanced by<br />

mechanisms analogous to those <strong>in</strong> the bulk, but with larger magnitudes because<br />

<strong>of</strong> the vic<strong>in</strong>ity to the no-flux <strong>boundary</strong>. Results for the no-<strong>slip</strong> case are virtually<br />

identical – the response <strong>of</strong> the flow to Brownian noise is quite <strong>in</strong>sensitive to the


94<br />

presence <strong>of</strong> <strong>slip</strong>.<br />

In other words, the modes excited by the Brownian noise are<br />

those that are common to both the <strong>slip</strong> <strong>and</strong> no-<strong>slip</strong> cases, <strong>and</strong> <strong>in</strong> fact, most <strong>of</strong> the<br />

spectrum is changed only trivially by <strong>slip</strong>, as shown <strong>in</strong> Fig. 40. Therefore, even <strong>in</strong><br />

the absence <strong>of</strong> flow <strong>in</strong>stability, (i.e. when We O(1) <strong>and</strong>/or <strong>slip</strong> is absent), the<br />

near surface regions may make a nontrivial contribution to the scatter<strong>in</strong>g signal,<br />

particularly its anisotropy under flow.<br />

5.4 Summary<br />

Recent experimental evidence suggests that <strong>slip</strong> at the <strong>polymer</strong>/solid <strong>in</strong>terface may<br />

play a role <strong>in</strong> the formation <strong>and</strong> development <strong>of</strong> enhanced concentration fluctuations<br />

<strong>in</strong> semidilute <strong>polymer</strong> solutions. In particular, the <strong>in</strong>itiation <strong>of</strong> enhanced concentration<br />

fluctuations was directly observed to happen at the <strong>in</strong>terface <strong>and</strong> modify<strong>in</strong>g<br />

the surface to <strong>in</strong>crease <strong>slip</strong> delayed the onset <strong>of</strong> enhancement to much higher <strong>shear</strong><br />

rates. Previous theoretical treatments have focused on the behavior <strong>in</strong> the x-y plane<br />

<strong>and</strong> have successfully expla<strong>in</strong>ed most <strong>of</strong> the behavior observed. Slip leads to a new<br />

class <strong>of</strong> viscoelastic flow <strong>in</strong>stabilities which result from the <strong>in</strong>teraction <strong>of</strong> <strong>slip</strong> with<br />

stress <strong>and</strong> concentration. The critical wavenumber agrees with those observed <strong>in</strong><br />

experiments so that the relevant length scale for the <strong>in</strong>stability is √ D tr λ.Thecritical<br />

We is O(10) for b ∼ O(1), which is with<strong>in</strong> the correct order <strong>of</strong> magnitude. Time<br />

<strong>in</strong>tegration <strong>of</strong> the govern<strong>in</strong>g equations with r<strong>and</strong>om concentration fluctuations distributed<br />

throughout the doma<strong>in</strong> shows that Brownian fluctuations are selectively<br />

<strong>and</strong> dramatically enhanced near the surface, with a <strong>boundary</strong> layer size consistent<br />

with experiment. The local wavevector for the enhanced fluctuation rotates as the


95<br />

<strong>shear</strong> rate <strong>in</strong>creases, as predicted <strong>and</strong> observed for the bulk. Overall, these results<br />

are consistent with experimental observations <strong>and</strong> underscore two po<strong>in</strong>ts regard<strong>in</strong>g<br />

the flow behavior <strong>of</strong> <strong>polymer</strong>ic liquids: (1) the dist<strong>in</strong>ctness <strong>and</strong> importance <strong>of</strong> the<br />

dynamics <strong>of</strong> flow<strong>in</strong>g <strong>polymer</strong>s near boundaries, even at the cont<strong>in</strong>uum level, <strong>and</strong><br />

(2) the importance <strong>of</strong> coupl<strong>in</strong>gs between various phenomena for the dynamics <strong>and</strong><br />

stability <strong>of</strong> these <strong>flows</strong>.


96<br />

Re<br />

⋆10 5<br />

2⋆10 5 0<br />

0<br />

1<br />

2<br />

y<br />

3<br />

(a)<br />

4<br />

5 0 1<br />

5<br />

4<br />

3<br />

2 y’<br />

Im<br />

200<br />

0<br />

200<br />

0<br />

1<br />

2<br />

y<br />

3<br />

(d)<br />

4<br />

5 0 1<br />

2<br />

5<br />

4<br />

3<br />

y’<br />

Re<br />

⋆10 5<br />

2⋆10 5<br />

0<br />

0<br />

1<br />

2<br />

y<br />

3<br />

(b)<br />

4<br />

5 0 1<br />

5<br />

4<br />

3<br />

2 y’<br />

Im<br />

1000<br />

0<br />

1000<br />

0<br />

1<br />

2<br />

y<br />

3<br />

(e)<br />

4<br />

5 0 1<br />

5<br />

4<br />

3<br />

2 y’<br />

Re<br />

⋆10 5<br />

2⋆10 5 0<br />

0<br />

1<br />

2<br />

y<br />

3<br />

(c)<br />

4<br />

5 0 1<br />

5<br />

4<br />

3<br />

2 y’<br />

Im<br />

5000<br />

0<br />

5000<br />

0<br />

1<br />

2<br />

y<br />

3<br />

(f)<br />

4<br />

5 0 1<br />

5<br />

4<br />

3<br />

2 y’<br />

Figure 38: Series <strong>of</strong> concentration correlation functions for <strong>in</strong>creas<strong>in</strong>g We. Pictures<br />

(a) <strong>and</strong> (d) are for We =0.1, (b) <strong>and</strong> (e) are for We = 1, <strong>and</strong> (c) <strong>and</strong> (f) are<br />

for We = 5. As the Weissenberg number <strong>in</strong>creases, the wavevector <strong>of</strong> the enhanced<br />

fluctuation rotates from the first quadrant to the fourth. The rema<strong>in</strong><strong>in</strong>g parameters<br />

are k x =1,S =0.01, N = 192.


97<br />

y<br />

0 Πk 2Πk 3Πk 4Πk<br />

x<br />

(a)<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

0 Πk x 2Πk x 3Πk x 4Πk x<br />

(b)<br />

y<br />

0 Πk x 2Πk x 3Πk x 4Πk x<br />

x<br />

(c)<br />

Figure 39: Series <strong>of</strong> snapshots <strong>of</strong> typical concentration pr<strong>of</strong>iles as We is <strong>in</strong>creased:<br />

(a) - We =0.1; (b) - We =1;<strong>and</strong>(c)-We = 5. Note change <strong>in</strong> orientation, from<br />

the first quadrant to the fourth, as We is <strong>in</strong>creased.


98<br />

0.5<br />

0.0<br />

Im(σ)<br />

−0.5<br />

−1.0<br />

−1.5<br />

−2.0<br />

0.0 25.0 50.0 75.0 100.0<br />

Re(σ) − b<br />

Figure 40: Eigenvalue spectra for the <strong>slip</strong> <strong>and</strong> no-<strong>slip</strong> cases <strong>in</strong> the bounded flow<br />

doma<strong>in</strong>. This figure only shows the least stable portions <strong>of</strong> the spectra. 3 - b =1,<br />

○ - b = 0. The rema<strong>in</strong><strong>in</strong>g parameters are We = 10, k x =1,S =0.01,N = 128.


99<br />

Chapter 6<br />

Conclud<strong>in</strong>g Remarks<br />

The results <strong>in</strong> Chs. 3 <strong>and</strong> 5 highlight the importance <strong>of</strong> <strong>polymer</strong> – surface <strong>in</strong>teractions,<br />

<strong>wall</strong> <strong>slip</strong> <strong>in</strong> particular, on the macroscopic flow behavior <strong>of</strong> <strong>polymer</strong> melts<br />

<strong>and</strong> solutions. It has been demonstrated that well-posed <strong>slip</strong> models can lead to<br />

<strong>in</strong>stabilities <strong>in</strong> <strong>polymer</strong> melt <strong>flows</strong>. In particular, <strong>in</strong>clusion <strong>of</strong> cha<strong>in</strong> orientation <strong>and</strong><br />

stretch<strong>in</strong>g at the surface <strong>in</strong> <strong>slip</strong> models is crucial for predict<strong>in</strong>g <strong>in</strong>stabilities consistent<br />

with experimental observations. Slip couples to stress <strong>and</strong> concentration <strong>in</strong><br />

<strong>polymer</strong> solutions as well, lead<strong>in</strong>g to hydrodynamic <strong>in</strong>stability <strong>in</strong> these systems,<br />

although this is not the only effect <strong>of</strong> surfaces on solutions. Indeed, r<strong>and</strong>om fluctuations<br />

are selectively <strong>and</strong> dramatically enhanced near the surfaces, even <strong>in</strong> the<br />

absence <strong>of</strong> <strong>slip</strong> <strong>and</strong> even at equilibrium. This type <strong>of</strong> surface effect may be important<br />

<strong>in</strong> other systems driven by r<strong>and</strong>om fluctuations, particularly phase separat<strong>in</strong>g <strong>polymer</strong><br />

solutions <strong>and</strong> blends, where the surface may affect the onset <strong>of</strong> phase separation<br />

as well as the morphology.<br />

To conclude, a recent experimental result for sharksk<strong>in</strong> is highly suggestive <strong>of</strong>


100<br />

possible future directions for this work <strong>and</strong> highlights some <strong>of</strong> the synergies between<br />

<strong>flows</strong> <strong>of</strong> melts <strong>and</strong> entangled solutions.<br />

Barone <strong>and</strong> Wang [7] studied extrusion<br />

<strong>of</strong> polybutadiene through a transparent quartz slit die. They chemically<br />

treated the downstream half <strong>of</strong> the die, by coat<strong>in</strong>g it with a polysiloxane.<br />

This<br />

enhanced <strong>wall</strong> <strong>slip</strong> <strong>and</strong> created an <strong>in</strong>ternal <strong>boundary</strong> s<strong>in</strong>gularity. Instability was<br />

observed us<strong>in</strong>g birefr<strong>in</strong>gence <strong>in</strong> the vic<strong>in</strong>ity <strong>of</strong> the s<strong>in</strong>gularity which decayed downstream.<br />

The period <strong>of</strong> oscillation was similar to the period <strong>of</strong> oscillation at the<br />

die exit for an uncoated die <strong>in</strong> which the extrudate exhibited sharksk<strong>in</strong>. Wang <strong>and</strong><br />

Plucktaveesak [101] observed a similar effect due to an <strong>in</strong>ternal <strong>boundary</strong> s<strong>in</strong>gularity<br />

dur<strong>in</strong>g HDPE extrusion. Mhetar <strong>and</strong> Archer [56] used quartz plates for their plane<br />

Couette cell <strong>and</strong> a similar siloxane treatment was effective <strong>in</strong> promot<strong>in</strong>g <strong>slip</strong> for<br />

polystyrene solutions. The velocity pr<strong>of</strong>ile measurement technique <strong>of</strong> Mhetar <strong>and</strong><br />

Archer could be adapted to Barone <strong>and</strong> Wang’s system to reveal the <strong>slip</strong> behavior<br />

near the <strong>in</strong>ternal s<strong>in</strong>gularity <strong>and</strong>/or Mhetar <strong>and</strong> Archer’s Couette cell could be used<br />

to reproduce Barone <strong>and</strong> Wang’s experiment. From a theoretical po<strong>in</strong>t <strong>of</strong> view, these<br />

bounded systems are the next step up <strong>in</strong> complexity from the viscometric analyses<br />

reported here, <strong>and</strong> have the advantage <strong>of</strong> be<strong>in</strong>g much more amenable to analysis<br />

than the full die exit problem, as there are no free surfaces <strong>and</strong> the attendant die<br />

swell to consider. Even so, these geometries capture the critical issues <strong>of</strong> velocity<br />

pr<strong>of</strong>ile rearrangement downstream <strong>of</strong> the s<strong>in</strong>gularity <strong>and</strong> extensional velocity components<br />

upstream. Underst<strong>and</strong><strong>in</strong>g the role <strong>of</strong> <strong>boundary</strong> s<strong>in</strong>gularities <strong>and</strong> velocity<br />

pr<strong>of</strong>ile rearrangement will prove crucial to underst<strong>and</strong><strong>in</strong>g sharksk<strong>in</strong> dynamics.


101<br />

Nomenclature<br />

The follow<strong>in</strong>g table summarizes the notation used <strong>in</strong> this work. In general, starred<br />

quantities have dimension <strong>and</strong> unstarred ones have been nondimensionalized. Certa<strong>in</strong><br />

variables, namely the ones typically used to nondimensionalize the others, are<br />

always dimensional, even though they are not denoted with the asterisk. These are<br />

noted <strong>in</strong> the description.<br />

Variable Description Location(s)<br />

b, b ∗ Extrapolation length §3.2<br />

B L<strong>in</strong>earization parameter §3.2<br />

c, c ∗ Eigenvalue describ<strong>in</strong>g the time dependence §3.2<br />

δ r Rubbery region thickness §2.1<br />

D tr Translational diffusivity §3.2, §5.1<br />

ɛ, ɛ ∗ Slip coefficient §2.3<br />

f Fractional recovery §2.1<br />

f n R<strong>and</strong>om forc<strong>in</strong>g §5.2<br />

F Drag force §2.1<br />

G L<strong>in</strong>earization parameter §3.2<br />

Cont<strong>in</strong>ued on next page


102<br />

Cont<strong>in</strong>ued from previous page<br />

Variable Description Location<br />

G ∗ Shear modulus §3.2<br />

˙γ<br />

Ratio <strong>of</strong> true <strong>and</strong> nom<strong>in</strong>al Weissenberg<br />

numbers<br />

§3.2<br />

˙γ n ∗ Nom<strong>in</strong>al <strong>shear</strong> rate §3.2<br />

˙γ t ∗ True <strong>shear</strong> rate §3.2<br />

η p Polymer viscosity (dimensional) §3.2<br />

η s Solvent viscosity (dimensional) §5.1<br />

H L<strong>in</strong>earization parameter §3.2<br />

k Fluctuation orientation vector §4<br />

k Magnitude <strong>of</strong> k §4<br />

k x<br />

k z<br />

Perturbation wavenumber <strong>in</strong> the flow direction<br />

Perturbation wavenumber <strong>in</strong> the neutral<br />

direction<br />

§3.2, §5.2<br />

§5.2<br />

l Gap width (dimensional) §3.2, §5.1<br />

L Map parameter §5.2<br />

λ Polymer relaxation time (dimensional) §3.2,§5.1<br />

λ s Relaxation time for <strong>slip</strong> (dimensional) §1.3.1<br />

µ Phan Thien - Tanner model parameter §3.2<br />

n Polymer concentration §5.1<br />

n Outward unit normal §5.2<br />

Cont<strong>in</strong>ued on next page


103<br />

Cont<strong>in</strong>ued from previous page<br />

Variable Description Location<br />

N 1 First normal stress difference §2.1<br />

Q Dumbbell orientation vector §5.1<br />

R Radius (dimensional) §2.1<br />

R L<strong>in</strong>earization parameter §3.2<br />

ρ Density (dimensional) §A.1<br />

ρ Anisotropic drag <strong>slip</strong> model parameter §2.4<br />

Re Reynolds number §1.3.1<br />

s<br />

S<br />

Equilibrium constant for the k<strong>in</strong>etic expression<br />

Ratio <strong>of</strong> solvent viscosity to <strong>polymer</strong> viscosity<br />

§2.3<br />

§3.2<br />

s R Recoverable <strong>shear</strong> §2.1<br />

t, t ∗ Time §1.3.1<br />

T Temperature (dimensional) 2.4<br />

τ , τ ∗ Polymer extra stress tensor §3.2, §5.1<br />

ˆτ Amplitude <strong>of</strong> the extra stress perturbation §3.2, §5.1<br />

˜τ Extra stress perturbation §3.2, §5.1<br />

û Velocity §3.2<br />

u ∗ f Velocity <strong>of</strong> the free segments §2.3<br />

u s , u ∗ s Slip velocity §1.3.1<br />

v Velocity §3.2<br />

Cont<strong>in</strong>ued on next page


104<br />

Cont<strong>in</strong>ued from previous page<br />

Variable Description Location<br />

V Characteristic velocity (dimensional) §2.1<br />

Wa Work <strong>of</strong> adhesion §2.1<br />

We Weissenberg number §4<br />

We n Nom<strong>in</strong>al Weissenberg number §3.2<br />

We t True Weissenberg number §3.2<br />

X Structural parameter/Bond<strong>in</strong>g fraction §2.3<br />

ỹ Scaled y coord<strong>in</strong>ate §3.3<br />

ζ, ζ ∗ Friction coefficient §2.1<br />

ζ, ζ ∗ Friction tensor §2.4


105<br />

Appendix A<br />

Basic Melt Equations<br />

A.1 Nondimensionalization<br />

The equations to be nondimensionalized are the equation <strong>of</strong> motion <strong>and</strong> the constitutive<br />

equation. These equations are<br />

ρ Dv∗<br />

Dt = ∗ ∇∗ · τ ∗ −∇ ∗ P ∗ , (71)<br />

(1 + µ G tr τ ∗ )τ ∗ + λτ ∗ ∗ (1) = η (<br />

p ∇ ∗ v ∗ +(∇ ∗ v ∗ ) ) T . (72)<br />

The follow<strong>in</strong>g relations are used to nondimensionalize the variables:<br />

v = v∗<br />

˙γ n ∗ l,<br />

(73)<br />

t =˙γ nt ∗ ∗ , (74)<br />

τ = τ ∗<br />

G , ∗ (75)<br />

p = P ∗<br />

G , ∗ (76)<br />

∇ = l∇ ∗ . (77)


106<br />

where ˙γ ∗ n l is the characteristic velocity, ˙γ∗ n<br />

is the applied <strong>shear</strong> rate at the <strong>wall</strong>, G∗<br />

is the <strong>shear</strong> modulus <strong>and</strong> l is the gap width. When these def<strong>in</strong>itions are substituted<br />

<strong>in</strong>to Eqs. 71 <strong>and</strong> 72, the result is<br />

Dv<br />

Dt =<br />

G∗<br />

(∇·τ −∇P ) , (78)<br />

ρ ˙γ n ∗ 2 l2 τ (1) + G∗<br />

(1 + µtr τ )τ = ∇v +(∇v) T (79)<br />

η p ˙γ n<br />

∗<br />

Recall that G ∗ = η p /λ. With this def<strong>in</strong>ition, the constant on the left h<strong>and</strong> side <strong>of</strong><br />

the equation <strong>of</strong> motion can be written as:<br />

( )( )<br />

G<br />

ρ ˙γ = ηp 1<br />

n ∗ 2l2<br />

ρ ˙γ n ∗ 2l<br />

˙γ n ∗λ . (80)<br />

S<strong>in</strong>ce We n =˙γ n ∗λ <strong>and</strong> Re = ρ ˙γ∗ nl ∗<br />

, the equations can f<strong>in</strong>ally be written <strong>in</strong> dimensionless<br />

form as:<br />

η p<br />

Dv<br />

ReWe n =(∇·τ −∇p) ,<br />

Dt<br />

(81)<br />

τ (1) + 1 (1 + µ tr τ )τ = ∇v +(∇v) T<br />

We n<br />

(82)<br />

(83)<br />

The equation <strong>of</strong> cont<strong>in</strong>uity for an <strong>in</strong>compressible fluid is simply<br />

∇·v = 0 (84)<br />

For all <strong>of</strong> the calculations <strong>in</strong> this thesis, Re =0.<br />

A.2 Derivation <strong>of</strong> the General Stability Equation


107<br />

Equations 81, 82, <strong>and</strong> 84 can be written <strong>in</strong> component form for the two dimensional<br />

problem with Re =0as:<br />

∂τ yx<br />

∂t<br />

+ u ∂τ xx<br />

∂x + v ∂τ xx<br />

∂y − 2(τ xx +1) ∂u<br />

∂x − 2τ ∂u<br />

yx<br />

∂y +<br />

1<br />

τ xx (1 + µτ xx + µτ yy )=0,<br />

We n<br />

+ u ∂τ yx<br />

∂x + v ∂τ yx<br />

∂y − τ ∂u<br />

yy<br />

∂y − ∂u<br />

∂y − (τ xx +1) ∂v<br />

∂x +<br />

1<br />

τ yx (1 + µτ xx + µτ yy )=0,<br />

We n<br />

∂τ xx<br />

∂t<br />

∂τ yy<br />

∂t<br />

+ u ∂τ yy<br />

∂x + v ∂τ yy<br />

∂y − 2τ ∂v<br />

yy<br />

∂y − 2τ ∂v<br />

yx<br />

∂x − 2∂v ∂y +<br />

1<br />

τ yy (1 + µτ xx + µτ yy )=0,<br />

We n<br />

∂τ xx<br />

∂x + ∂τ yx<br />

∂y − ∂p<br />

∂x =0,<br />

∂τ yx<br />

∂x + ∂τ yy<br />

∂y − ∂p<br />

∂y =0,<br />

∂u<br />

∂x + ∂v<br />

∂y<br />

(85a)<br />

(85b)<br />

(85c)<br />

(85d)<br />

(85e)<br />

=0, (85f )<br />

The next step is to l<strong>in</strong>earize this system <strong>of</strong> equations.<br />

This is accomplished by<br />

assum<strong>in</strong>g that each variable is equal to its steady state value plus some small perturbation,<br />

then reta<strong>in</strong><strong>in</strong>g only the terms that are l<strong>in</strong>ear <strong>in</strong> the perturbations. Each<br />

variable has the form a =ā + δã where the bar denotes the steady state value, the<br />

tilde <strong>in</strong>dicates the perturbation, <strong>and</strong> δ is a small parameter used to keep track <strong>of</strong> the<br />

l<strong>in</strong>ear terms. Collect<strong>in</strong>g the O(1) terms yields the system to be solved for the steady<br />

state solution. The O(δ) system is the l<strong>in</strong>ear system for the perturbations. Note<br />

that the steady state values <strong>of</strong> the y-component <strong>of</strong> velocity <strong>and</strong> the second normal<br />

stress are identically zero, <strong>and</strong> there are no gradients <strong>of</strong> the base state variables <strong>in</strong>


108<br />

the x-direction. F<strong>in</strong>ally, the l<strong>in</strong>earized system <strong>of</strong> equations is<br />

∂˜τ yx<br />

∂t<br />

∂˜τ xx<br />

∂t<br />

+ū ∂˜τ xx<br />

∂x +ṽ d¯τ xx<br />

dy − 2(¯τ xx +1) ∂ũ<br />

∂x − 2¯τ ∂ũ<br />

yx<br />

∂y − 2˜τ ∂ū<br />

yx<br />

∂y +<br />

1<br />

(1 + 2µ¯τ xx )˜τ xx + µ ¯τ xx˜τ yy =0,<br />

We n We n<br />

+ū ∂˜τ yx<br />

∂x +ṽ d¯τ xx<br />

dy − ˜τ yy<br />

∂˜τ yy<br />

∂t<br />

∂ū<br />

∂y − ∂ũ<br />

∂y − (¯τ xx +1) ∂ṽ<br />

∂x + µ ¯τ yx˜τ xx +<br />

We n<br />

1<br />

(1 + µ¯τ xx )˜τ yx + µ ¯τ yx˜τ yy =0,<br />

We n We n<br />

+ū ∂˜τ yy<br />

∂x − 2¯τ ∂ṽ<br />

yx<br />

∂x − 2∂ṽ ∂y + 1 We (1 + µ¯τ xx)˜τ yy =0,<br />

∂˜τ xx<br />

∂x + ∂˜τ yx<br />

∂y − ∂ ˜p<br />

∂x =0,<br />

∂˜τ yx<br />

∂x + ∂˜τ yy<br />

∂y − ∂ ˜p<br />

∂y =0,<br />

∂ũ<br />

∂x + ∂ṽ<br />

∂y<br />

(86a)<br />

(86b)<br />

(86c)<br />

(86d)<br />

(86e)<br />

=0, (86f )<br />

The last step is to <strong>in</strong>troduce a form for the perturbations.<br />

For a normal mode<br />

analysis, it is a assumed that each variable can be written as â(y)e ikx(x−ct) +c.c.,<br />

where â(y) is the amplitude <strong>and</strong> is a function only <strong>of</strong> y, k x is the wavenumber, <strong>and</strong><br />

c is the eigenvalue for the time dependence <strong>of</strong> the disturbance. After substitut<strong>in</strong>g<br />

the perturbations <strong>in</strong>to Eqs. 86a -86f , the system <strong>of</strong> equations to be solved for the


109<br />

perturbations is<br />

(−ik x c + ik x ū + 1 + 2µ ¯τ xx )ˆτ xx − 2ik x (¯τ xx +1)û − 2¯τ yx û ′ − 2ū ′ˆτ yx +<br />

We n We n<br />

µ<br />

¯τ xxˆτ yy =0,(87a)<br />

We n<br />

(−ik x c + ik x ū + 1 + µ ¯τ xx )ˆτ yx − ik xˆv − û ′ − ū ′ˆτ yy + µ (ˆτ xx +ˆτ yy )=0,(87b)<br />

We n We n We n<br />

(−ik x c + ik x ū + 1 + µ ¯τ xx )ˆτ yy − 2ik x¯τ yxˆv − 2ˆv ′ =0,(87c)<br />

We n We n<br />

ik xˆτ xx +ˆτ ′ yx − ik xˆp =0,(87d)<br />

ik xˆτ yx +ˆτ ′ yy − ˆp′ =0,(87e)<br />

ik x û +ˆv ′ =0,(87f )<br />

The ′ denotes differentiation with respect to y.<br />

The system <strong>of</strong> equations def<strong>in</strong>ed by Eqs. 87a through 87f can be reduced to<br />

a s<strong>in</strong>gle equation for the stream function. S<strong>in</strong>ce the stream function perturbation<br />

is assumed to have the same form as the other perturbations, reduc<strong>in</strong>g to a s<strong>in</strong>gle<br />

equation for the stream function is equivalent to reduc<strong>in</strong>g to a s<strong>in</strong>gle equation for<br />

ˆv. The first step <strong>in</strong> the reduction is to comb<strong>in</strong>e Equations 87d <strong>and</strong> 87e to elim<strong>in</strong>ate<br />

pressure <strong>and</strong> get a s<strong>in</strong>gle equation relat<strong>in</strong>g just the stress components:<br />

ik xˆτ ′ xx + k 2 xˆτ yx +ˆτ ′′<br />

yx − ik xˆτ ′ yy =0. (88)<br />

The three constitutive relations can then be solved for the stresses <strong>and</strong> substituted<br />

<strong>in</strong>to Eq. 88. After not<strong>in</strong>g that û = ik xˆv ′ (from the equation <strong>of</strong> cont<strong>in</strong>uity) <strong>and</strong>


110<br />

do<strong>in</strong>g some substantial rearrangement, the general stability equation is<br />

[<br />

(Q 2 D 2 − k 2 xQ 2 +2(ū ′ ) 2 − 2Qū ′ D)(D 2 +2ik x¯τ yx D − k 2 x − k 2 x¯τ xx )+<br />

kxQū 2 ′′ (¯τ xx +1)+ik x (ik x Q¯τ xx ′ +2Q¯τ yx ′′ − 4¯τ yx ū ′′ )(QD − ū ′ ) − 3Qū ′′ +<br />

4ū ′ ū ′′ ′′′<br />

D + ik x Q<br />

2¯τ yx − 2Qū ′′′ D − 2ik x Q¯τ yx ū ′′′ +3ik x Q¯τ yx(QD ′ 2 − ū ′′ ) −<br />

ik x¯τ ′ yx(k 2 xQ 2 +4ū ′ QD − 2(ū ′ ) 2 ) ] ˆv + µM(¯τ,µ)ˆv =0, (89)<br />

where Q = −c +ū −<br />

i<br />

k xWe n<br />

, D is the ord<strong>in</strong>ary derivative with respect to y, <strong>and</strong>M is<br />

a complicated differential operator. For µ = 0, <strong>and</strong> plane Couette flow (i.e. the base<br />

state gradients <strong>of</strong> stress with respect to y <strong>and</strong> second order <strong>and</strong> higher derivatives<br />

<strong>of</strong> the base state velocity are zero), this reduces to<br />

(Q 2 D 2 − k 2 x Q2 +2(ū ′ ) 2 − 2Qū ′ D)(D 2 +2ik x¯τ yx D − k 2 x − k2 x¯τ xx)ˆv, (90)<br />

which is identical to the general equation derived by Gorodtsov <strong>and</strong> Leonov [27].<br />

A.3 PTT Matrix Eigenvalue Problem<br />

The generalized eigenvalue problem that is obta<strong>in</strong>ed when the PTT constitutive<br />

equation is used is


111<br />

L =<br />

⎡<br />

⎢ 0 iα d<br />

dy<br />

⎢⎢⎢⎢⎢⎢⎢⎣<br />

d 2<br />

+ α 2 −iα d<br />

dy 2 dy<br />

−2We n¯τyx d2 − 2iαWen(1 + ¯τxx) d dy 2 dy<br />

1+iαWen d ¯ψ dy<br />

+2µ¯τxx −2We n d2 ¯ψ dy 2 µ¯τxx<br />

⎤<br />

⎥ ⎥⎥⎥⎥⎥⎥⎥⎦<br />

−α 2 We n(1 + ¯τxx) − We n d2<br />

dy 2 µ¯τyx 1+iαWen d +<br />

dy<br />

µ¯τxx −We n d2 +<br />

dy 2 µ¯τyx<br />

−2α 2 We n¯τyx +2iαWen d<br />

dy<br />

0 0<br />

1+iαWen d +<br />

dy<br />

µ¯τxx<br />

⎡<br />

⎤<br />

(91)<br />

⎢ 0 ⎢⎢⎢⎢⎢⎢⎢⎣ 0 0 0 0 1 0 0<br />

C =<br />

0 0 1 0<br />

0 0 0 1


112<br />

Appendix B<br />

3D Stability Operators<br />

The three dimensional operators for the <strong>polymer</strong> solution model are, with ∇ 2 =<br />

d 2<br />

dx 2 + d2<br />

dy 2 + d2<br />

dz 2<br />

<strong>and</strong> ¯v ·∇=ū d<br />

dx +¯v d<br />

dy +¯w d dz ,<br />

⎡<br />

C S =<br />

⎢<br />

⎣<br />

0 0 0 0 0 0 0 0 0 0 0<br />

0 0 0 0 0 0 0 0 0 0 0<br />

0 0 0 0 0 0 0 0 0 0 0<br />

0 0 0 1 0 0 0 0 0 1 0<br />

0 0 0 0 1 0 0 0 0 0 0<br />

0 0 0 0 0 1 0 0 0 0 0<br />

0 0 0 0 0 0 1 0 0 1 0<br />

0 0 0 0 0 0 0 1 0 0 0<br />

0 0 0 0 0 0 0 0 1 1 0<br />

0 0 0 0 0 0 0 0 0 1 0<br />

0 0 0 0 0 0 0 0 0 0 0<br />

⎤<br />

⎥<br />

⎦<br />

(93)


113<br />

L S =<br />

⎡<br />

SWe∇ 2 d<br />

d<br />

0 0<br />

dx<br />

dy<br />

⎢ ⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣ 0 SWe∇ 2 d<br />

0 0<br />

dx<br />

0<br />

[<br />

0 SWe∇ 2 0 0<br />

−2We(1 − ξ) (¯τxx +¯n) d<br />

( )<br />

] dx<br />

We(1 − ξ) d¯τxx<br />

dy + d¯n<br />

We(1 − ξ)(¯v ·∇)<br />

0<br />

dy<br />

−(1 − ξ)(∇ 2 −2We(1 − ξ) dū<br />

− 1)<br />

dy<br />

+¯τyx d dy +¯τzx d [ dz<br />

[<br />

−We(1 − ξ) d<br />

dx +<br />

]<br />

−We(1 − ξ) (¯τxx d<br />

dx ] 0 0<br />

¯τyx d<br />

+¯n) d<br />

(¯τyy +¯n) d<br />

dy +¯τzy d dz<br />

−We(1 − ξ)<br />

[<br />

¯τzx d<br />

dx<br />

+¯τzy d dy +(¯τzz +¯n) d dz<br />

0<br />

0<br />

+¯τyx d<br />

dy +¯τzx d dz − d¯τyx<br />

dy<br />

] We(1 − ξ) d¯τzx<br />

dy<br />

−2We(1 − ξ)<br />

[<br />

¯τyx d<br />

dx +<br />

(¯τyy +¯n) d<br />

dy + τzy d dz − 1 2<br />

[<br />

d¯τyy<br />

dy<br />

−We(1 − ξ) ¯τzx d<br />

dx +¯τzy d ] dy<br />

+(¯τzz +¯n) d dz − d¯τzy<br />

dy<br />

(<br />

0 We(1 − ξ) d¯τzz<br />

dy + d¯n<br />

dy<br />

−We(1 − ξ)<br />

[<br />

+¯τyx d dy +¯τzx d dz<br />

We(1 − ξ)(¯v ·∇)<br />

−(1 − ξ)(∇ 2 − 1)<br />

(¯τxx +¯n) d<br />

] dx<br />

0 −We(1 − ξ) d ¯w<br />

] 0 0 0<br />

−We(1 − ξ)<br />

[<br />

¯τyx d<br />

dx<br />

+¯τyy d dy +(¯τzy +¯n) d [ dz<br />

) −2We(1 − ξ)<br />

¯τzx d<br />

dx<br />

+¯τzy d dy +(¯τzz +¯n) d dz<br />

] 0 0<br />

] 0 0<br />

0 We(1 − ξ) d¯n<br />

0 (1− ξ) d2<br />

dy dx 2 2(1 − ξ) d2<br />

dx dy<br />

d<br />

dx<br />

d<br />

dy<br />

d<br />

0 0<br />

dz<br />

d<br />

0 0 0 0 − d<br />

dz dx<br />

d<br />

d<br />

0<br />

0 0 − d<br />

dy<br />

dz dy<br />

d<br />

d<br />

d<br />

0<br />

0 − d dx dy<br />

dz dz<br />

0 0 0 0<br />

dy<br />

We(1 − ξ)(¯v ·∇)<br />

−(1 − ξ)∇ 2 0<br />

0 −We(1 − ξ) d<br />

0 0 −We(1 − ξ) dū 0<br />

dy dy<br />

We(1 − ξ)(¯v ·∇)<br />

−(1 − ξ)(∇ 2 0 −We(1 − ξ) dū<br />

0 0 0<br />

− 1)<br />

dy<br />

0<br />

We(1 − ξ)(¯v ·∇)<br />

−(1 − ξ)(∇ 2 +1)<br />

0 −We(1 − ξ) d ¯w<br />

dy<br />

0 0<br />

We(1 − ξ)(¯v ·∇) − d¯v<br />

dy<br />

−(1 − ξ)(∇ 2 − 1)<br />

0 0 −2We(1 − ξ) d ¯w<br />

dy<br />

We(1 − ξ)(¯v ·∇)<br />

−(1 − ξ)(∇ 2 − 1)<br />

0 −We(1 − ξ) d ¯w 0<br />

dy<br />

We(1 − ξ)(¯v ·∇)<br />

−(1 − ξ)(∇ 2 − 1)<br />

We(1 − ξ)(¯v ·∇)<br />

−(1 − ξ)∇ 2 0<br />

0<br />

We(1 − ξ)(¯v ·∇)<br />

2(1 − ξ) d2<br />

(1 − ξ) d2<br />

dx dz dy 2 2(1 − ξ) d2<br />

(1 − ξ) d2<br />

dz dy dz 2<br />

−(1 − ξ)∇ 2 0<br />

0 0 0 0 0 0<br />

···<br />

⎤<br />

⎥ ⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦<br />

(94)


114<br />

Bibliography<br />

[1] D. Acierno, F. P. L. Mantia, G. Marrucci, <strong>and</strong> G. Titomanlio, “A non-l<strong>in</strong>ear<br />

viscoelastic model with structure dependent relaxation times I: Basic formulation,”<br />

J. Non-Newtonian Fluid Mech., 1:125–146 (1976).<br />

[2] J.-F. Agassant, P. Avenas, J.-P. Sergent, <strong>and</strong> P. J. Carreau, Polymer Process<strong>in</strong>g,<br />

Pr<strong>in</strong>ciples <strong>and</strong> Model<strong>in</strong>g, Hanser, 1991.<br />

[3] A. Ajdari, F. Brochard-Wyart, P.-G. de Gennes, L. Leibler, J.-L. Viovy, <strong>and</strong><br />

M. Rub<strong>in</strong>ste<strong>in</strong>, “Slippage <strong>of</strong> an entangled <strong>polymer</strong> melt on a grafted surface,”<br />

Physica A, 204:17–39 (1994).<br />

[4] A. Ajji, H. P. Schreiber, <strong>and</strong> D. Duchesne, “Flow defects <strong>in</strong> l<strong>in</strong>ear low density<br />

polyethylene process<strong>in</strong>g: Instrumental detection <strong>and</strong> molecular weight<br />

dependence,” Polymer Eng<strong>in</strong>eer<strong>in</strong>g <strong>and</strong> Science, 33(23):1524–1531 (1993).<br />

[5] L. A. Archer, R. G. Larson, <strong>and</strong> Y. L. Chen, “Direct measurements <strong>of</strong> <strong>slip</strong> <strong>in</strong><br />

<strong>shear</strong>ed <strong>polymer</strong> solutions,” J. Fluid Mech., 301:133–151 (1995).<br />

[6] B. T. Atwood <strong>and</strong> W. R. Schowalter, “Measurements <strong>of</strong> <strong>slip</strong> at the <strong>wall</strong> dur<strong>in</strong>g<br />

flow <strong>of</strong> high density polyethylene through a rectangular conduit,” Rheol. Acta,<br />

28:134–146 (1989).<br />

[7] J. Barone <strong>and</strong> S.-Q. Wang, “Flow birefr<strong>in</strong>gence study <strong>of</strong> sharksk<strong>in</strong> <strong>and</strong> stress<br />

relaxation <strong>in</strong> polybutadiene melts,” Rheol. Acta, 38(5):404–414 (1999).


115<br />

[8] J. R. Barone, N. Plucktaveesak, <strong>and</strong> S. Q. Wang, “Interfacial molecular <strong>in</strong>stability<br />

mechanism for sharksk<strong>in</strong> phenomenon <strong>in</strong> capillary extrusion <strong>of</strong> l<strong>in</strong>ear<br />

polyethylenes,” J. Rheol., 42(2):813–832 (1998).<br />

[9] A. N. Beris <strong>and</strong> V. G. Mavrantzas, “On the compatability between various<br />

macroscopic formalisms for the concentration <strong>and</strong> the flow <strong>of</strong> dilute <strong>polymer</strong><br />

solutions,” J. Rheol., 38(5):1235–1250 (1994).<br />

[10] R. B. Bird, C. F. Curtiss, R. C. Armstrong, <strong>and</strong> O. Hassager, Dynamics <strong>of</strong><br />

Polymeric Liquids, volume 2, Wiley - Interscience, New York, 1987.<br />

[11] I. A. Bitsanis <strong>and</strong> G. ten Br<strong>in</strong>ke, “A lattice Monte Carlo study <strong>of</strong> long cha<strong>in</strong><br />

conformations at solid-<strong>polymer</strong> melt <strong>in</strong>terfaces,” J. Chem. Phys., 99(4):3100–<br />

3111 (1993).<br />

[12] W. B. Black. “A k<strong>in</strong>etic <strong>slip</strong> model for the prediction <strong>of</strong> <strong>polymer</strong> melt flow<br />

<strong>in</strong>stabilities,” . Master’s thesis, University <strong>of</strong> Wiscons<strong>in</strong>-Madison, 1997.<br />

[13] W. B. Black <strong>and</strong> M. D. Graham, “Wall-<strong>slip</strong> <strong>and</strong> <strong>polymer</strong>-melt flow <strong>in</strong>stability,”<br />

Phys. Rev. Lett., 77(5):956–959 (1996).<br />

[14] W. B. Black <strong>and</strong> M. D. Graham, “Effect <strong>of</strong> <strong>wall</strong> <strong>slip</strong> on the stability <strong>of</strong> viscoelastic<br />

plane <strong>shear</strong> flow,” Phys. Fluids, 11(7):1749–1756 (1999).<br />

[15] J. P. Boyd, “Orthogonal rational functions on a semi-<strong>in</strong>f<strong>in</strong>ite <strong>in</strong>terval,” J.<br />

Comp. Phys., 70:63–88 (1987).<br />

[16] F. Brochard <strong>and</strong> P. G. de Gennes, “Shear-dependent <strong>slip</strong>page at a <strong>polymer</strong>/solid<br />

<strong>in</strong>terface,” Langmuir, 8:3033–3037 (1992).<br />

[17] C. Canuto, M. Y. Hussa<strong>in</strong>i, A. Quarteroni, <strong>and</strong> T. A. Zang, Spectral Methods<br />

<strong>in</strong> Fluid Dynamics, Spr<strong>in</strong>ger-Verlag, Berl<strong>in</strong>, 1988.


116<br />

[18] C. F. Curtiss <strong>and</strong> R. B. Bird, “Diffusion-stress relations <strong>in</strong> <strong>polymer</strong> mixtures,”<br />

J. Chem. Phys., 111(22):10362 – 10370 (1999).<br />

[19] J. Dealy <strong>and</strong> K. Wissbrun, Melt Rheology <strong>and</strong> its Role <strong>in</strong> Plastics Process<strong>in</strong>g:<br />

Theory <strong>and</strong> Applications, Van Nostr<strong>and</strong> Re<strong>in</strong>hold, New York, 1990.<br />

[20] W. R. Dean <strong>and</strong> P. E. Montagnon, “On the steady motion <strong>of</strong> viscous liquid<br />

<strong>in</strong> a corner,” Proc. Cambridge. Philos. Soc., 45:389 (1949).<br />

[21] P. G. Draz<strong>in</strong> <strong>and</strong> W. H. Reid, Hydrodynamic Stability, Cambridge University<br />

Press, 1981.<br />

[22] P. A. Drda <strong>and</strong> S.-Q. Wang, “Stick-<strong>slip</strong> transition at <strong>polymer</strong> melt/solid <strong>in</strong>terfaces,”<br />

Phys. Rev. Let., 75(14):2698–2701 (1995).<br />

[23] B. S. Garbow, “The QZ algorithm to solve the generalized eigenvalue problem<br />

for complex matrices,” ACM Transactions on Mathematical S<strong>of</strong>tware,<br />

4(4):404–410 (1978).<br />

[24] C. W. Gard<strong>in</strong>er, H<strong>and</strong>book <strong>of</strong> Stochastic Methods, Spr<strong>in</strong>ger-Verlag, Berl<strong>in</strong>,<br />

1985.<br />

[25] G. C. Georgiou, “On the stability <strong>of</strong> the <strong>shear</strong> flow <strong>of</strong> a viscoelastic fluid with<br />

<strong>slip</strong> along the fixed <strong>wall</strong>,” Rheol. Acta, 35:39–47 (1996).<br />

[26] V. G. Ghanta, B. J. Riise, <strong>and</strong> M. M. Denn, “Disappearance <strong>of</strong> extrusion<br />

<strong>in</strong>stabilities <strong>in</strong> brass capillary dies,” J. Rheol., 43(2):435–442 (1999).<br />

[27] V. A. Gorodtsov <strong>and</strong> A. I. Leonov, “On a l<strong>in</strong>ear stability <strong>of</strong> a plane parallel<br />

Couette flow <strong>of</strong> viscoelastic fluid,” J. Appl. Math. Mech., 31(2):310–319<br />

(1967).


117<br />

[28] M. D. Graham, “Effect <strong>of</strong> axial flow on viscoelastic Taylor-Couette <strong>in</strong>stability,”<br />

J. Fluid Mech., 360:341–374 (1998).<br />

[29] M. D. Graham, “The sharksk<strong>in</strong> <strong>in</strong>stability <strong>of</strong> <strong>polymer</strong> melt <strong>flows</strong>,” Chaos,<br />

9(1):154–163 (1999).<br />

[30] P. J. Halley <strong>and</strong> M. E. Mackay, “The effect <strong>of</strong> metals on the process<strong>in</strong>g <strong>of</strong><br />

LLDPE through a slit die,” J. Rheol., 38(1):41–51 (1994).<br />

[31] T. Hashimoto <strong>and</strong> K. Fujioka, “Shear enhanced concentration fluctuations <strong>in</strong><br />

<strong>polymer</strong> solutions as observed by flow light scatter<strong>in</strong>g,” J. Phys. Soc. Jpn.,<br />

60(2):356–359 (1991).<br />

[32] T. Hashimoto <strong>and</strong> T. Kume, ““Butterfly” light scatter<strong>in</strong>g patterns <strong>in</strong> <strong>shear</strong>enhanced<br />

concentration fluctuations <strong>in</strong> <strong>polymer</strong> solutions <strong>and</strong> anomaly at high<br />

<strong>shear</strong> rates,” J. Phys. Soc. Jpn., 61(6):1839–1843 (1992).<br />

[33] S. G. Hatzikiriakos <strong>and</strong> J. M. Dealy, “Wall <strong>slip</strong> <strong>of</strong> molten high density<br />

polyethylene. I. Slid<strong>in</strong>g plate rheometer studies,” J. Rheol., 35(4):497–523<br />

(1991).<br />

[34] S. G. Hatzikiriakos <strong>and</strong> J. M. Dealy, “Wall <strong>slip</strong> <strong>of</strong> molten high density<br />

polyethylene. II. Capillary rheometer studies,” J. Rheol., 36(4):703–741<br />

(1992).<br />

[35] S. G. Hatzikiriakos <strong>and</strong> N. Kalogerakis, “A dynamic <strong>slip</strong> velocity model for<br />

molten <strong>polymer</strong>s based on a network k<strong>in</strong>etic theory,” Rheol. Acta, 33:38–47<br />

(1994).<br />

[36] E. Helf<strong>and</strong> <strong>and</strong> G. H. Fredrickson, “Large fluctuations <strong>in</strong> <strong>polymer</strong> solutions<br />

under <strong>shear</strong>,” Phys. Rev. Lett., 62(21):2468–2471 (1989).


118<br />

[37] D. A. Hill, “Wall <strong>slip</strong> <strong>in</strong> <strong>polymer</strong> melts: A pseudo-chemical model,” J. Rheol.,<br />

42(3):581–601 (1998).<br />

[38] D. A. Hill, T. Hasegawa, <strong>and</strong> M. M. Denn, “On the apparent relation between<br />

adhesive failure <strong>and</strong> melt fracture,” J. Rheol., 34(6):891–918 (1990).<br />

[39] E. J. H<strong>in</strong>ch, “The flow <strong>of</strong> an Oldroyd fluid around a sharp corner,” J. Non-<br />

Newtonian Fluid Mech., 50:161–171 (1993).<br />

[40] T. C. Ho <strong>and</strong> M. M. Denn, “Stability <strong>of</strong> plane Poiseuille flow <strong>of</strong> a highly elastic<br />

liquid,” J. Non-Newtonian Fluid Mech., 3:179–195 (1978).<br />

[41] H. Ji <strong>and</strong> E. Helf<strong>and</strong>, “Concentration fluctuations <strong>in</strong> <strong>shear</strong>ed <strong>polymer</strong> solutions,”<br />

Macromolecules, 28(11):3869–3880 (1995).<br />

[42] D. S. Kalika <strong>and</strong> M. M. Denn, “Wall <strong>slip</strong> <strong>and</strong> extrudate distortion <strong>in</strong> l<strong>in</strong>ear<br />

low-density polyethylene,” J. Rheol., 31(8):815–834 (1987).<br />

[43] R. A. Keiller, “Numerical <strong>in</strong>stability <strong>of</strong> time dependent <strong>flows</strong>,” J. Non-<br />

Newtonian Fluid Mech., 43:229–246 (1992).<br />

[44] N. E. Kissi <strong>and</strong> J. M. Piau, “The different capillary flow regimes <strong>of</strong> entangled<br />

polydimethylsiloxane <strong>polymer</strong>s: Macroscopic <strong>slip</strong> at the <strong>wall</strong>, hysteresis, <strong>and</strong><br />

cork flow,” J. Non-Newtonian Fluid Mech., 37:55–94 (1990).<br />

[45] N. E. Kissi, J.-M. Piau, <strong>and</strong> F. Toussa<strong>in</strong>t, “Sharksk<strong>in</strong> <strong>and</strong> crack<strong>in</strong>g <strong>of</strong> <strong>polymer</strong><br />

melt extrudates,” J. Non-Newtonian Fluid Mech., 68(2-3):271–290 (1997).<br />

[46] T. Kume, T. Hashimoto, T. Takahashi, <strong>and</strong> G. G. Fuller, “Rheo-optical studies<br />

<strong>of</strong> <strong>shear</strong> <strong>in</strong>duced structures <strong>in</strong> semi-dilute polystyrene solutions,” Macromolecules,<br />

30(23):7232–7236 (1997).


119<br />

[47] S. J. Kurtz, “Die geometry solutions to sharksk<strong>in</strong> melt fracture,” In Advances<br />

<strong>in</strong> Rheology, Proceed<strong>in</strong>gs <strong>of</strong> the IX International Congress on Rheology, volume<br />

3, pages 399–407, 1984.<br />

[48] S. J. Kurtz, “The dynamics <strong>of</strong> sharksk<strong>in</strong> melt fracture <strong>in</strong> LLDPE,” In Proceed<strong>in</strong>gs<br />

<strong>of</strong> the 7th PPS Meet<strong>in</strong>g, pages 54–55, 1991.<br />

[49] S. J. Kurtz, “The dynamics <strong>of</strong> sharksk<strong>in</strong> melt fracture: Effect <strong>of</strong> the die geometry,”<br />

In Theoretical <strong>and</strong> Applied Rheology, Proceed<strong>in</strong>gs <strong>of</strong> the XI International<br />

Congress on Rheology, pages 377–379, 1992.<br />

[50] K.-C. Lee <strong>and</strong> B. A. F<strong>in</strong>layson, “Stability <strong>of</strong> plane Poiseuille <strong>and</strong> Couette flow<br />

<strong>of</strong> a Maxwell fluid,” J. Non-Newtonian Fluid Mech., 21:65–78 (1986).<br />

[51] F. J. Lim <strong>and</strong> W. R. Schowalter, “Pseudo-spectral analysis <strong>of</strong> the stability<br />

<strong>of</strong> pressure driven flow <strong>of</strong> a Giesekus fluid between parallel planes,” J. Non-<br />

Newtonian Fluid Mech., 26:135–142 (1987).<br />

[52] F. J. Lim <strong>and</strong> W. R. Schowalter, “Wall <strong>slip</strong> <strong>of</strong> narrow molecular weight distribution<br />

polybutadienes,” J. Rheol., 33(8):1359–1382 (1989).<br />

[53] M. E. Mackay <strong>and</strong> D. J. Henson, “The effect <strong>of</strong> molecular mass <strong>and</strong> temperature<br />

on the <strong>slip</strong> <strong>of</strong> polystyrene melts at low stress levels,” J. Rheol.,<br />

42(6):1505–1517 (1998).<br />

[54] V. G. Mavrantzas <strong>and</strong> A. N. Beris, “Model<strong>in</strong>g <strong>of</strong> the rheology <strong>and</strong> flow-<strong>in</strong>duced<br />

concentration changes <strong>in</strong> <strong>polymer</strong> solutions,” Phys. Rev. Lett., 69(2):273–276<br />

(1992).<br />

[55] V. Mhetar <strong>and</strong> L. A. Archer, “Molecular orig<strong>in</strong> <strong>of</strong> <strong>polymer</strong> melt flow <strong>in</strong>stabilities,”<br />

In Conference Proceed<strong>in</strong>gs/ANTEC, pages 1876–1880, 1997.


120<br />

[56] V. Mhetar <strong>and</strong> L. A. Archer, “Slip <strong>in</strong> entangled <strong>polymer</strong> solutions,” Macromolecules,<br />

31(19):6639–6649 (1998).<br />

[57] V. R. Mhetar <strong>and</strong> L. A. Archer, “Secondary flow <strong>of</strong> entangled <strong>polymer</strong> fluids<br />

<strong>in</strong> plane Couette <strong>shear</strong>,” J. Rheol., 40(4):549–571 (1996).<br />

[58] K. B. Migler, H. Hervet, <strong>and</strong> L. Leger, “Slip transition <strong>of</strong> a <strong>polymer</strong> melt<br />

under <strong>shear</strong> stress,” Phys. Rev. Let., 70(3):287–290 (1993).<br />

[59] S. T. Milner, “Hydrodynamics <strong>of</strong> semidilute <strong>polymer</strong> solutions,” Phys. Rev.<br />

Lett., 66(11):1477–1480 (1991).<br />

[60] S. T. Milner, “Dynamical theory <strong>of</strong> concentration fluctuations <strong>in</strong> <strong>polymer</strong><br />

solutions under <strong>shear</strong>,” Phys. Rev. E, 48(5):3674–3691 (1993).<br />

[61] H. K. M<strong>of</strong>fatt, “Viscous <strong>and</strong> resistive eddies near a sharp corner,” J. Fluid<br />

Mech., 18:1–18 (1964).<br />

[62] M. Mooney, “Explicit formulas for <strong>slip</strong> <strong>and</strong> fluidity,” J. Rheol., 2:210 (1931).<br />

[63] E. Moses, T. Kume, <strong>and</strong> T. Hashimoto, “Shear microscopy <strong>of</strong> the “butterfly<br />

pattern” <strong>in</strong> <strong>polymer</strong> mixtures,” Phys. Rev. Lett., 72(13):2037–2040 (1994).<br />

[64] R. H. Moynihan, D. G. Baird, <strong>and</strong> R. Ramanathan, “Additional observations<br />

on the surface melt fracture behavior <strong>of</strong> l<strong>in</strong>ear low density polyethylene,” J.<br />

Non-Newtonian Fluid Mech., 36:255–263 (1990).<br />

[65] C. Pan <strong>and</strong> I. Bitsanis, “Structure, conformation, <strong>and</strong> dynamics <strong>of</strong> <strong>polymer</strong><br />

cha<strong>in</strong>s at solid-melt <strong>in</strong>terfaces,” Makromol. Chem.,Macromol. Symp., 65:211–<br />

222 (1993).<br />

[66] J. R. A. Pearson <strong>and</strong> C. J. S. Petrie, “On the melt flow <strong>in</strong>stability <strong>of</strong> extruded<br />

<strong>polymer</strong>s,” In R. E. Wetton <strong>and</strong> R. W. Whorlow, editors, Polymer Systems:


121<br />

Deformation <strong>and</strong> Flow, Proceed<strong>in</strong>gs <strong>of</strong> the 1966 Annual Conference <strong>of</strong> the<br />

British Society <strong>of</strong> Rheology, pages 163–187, London, 1968. Macmillan.<br />

[67] T. J. Person <strong>and</strong> M. M. Denn, “The effect <strong>of</strong> die materials <strong>and</strong> pressuredependent<br />

<strong>slip</strong> on the extrusion <strong>of</strong> l<strong>in</strong>ear low-density polyethylenes,” J. Rheol.,<br />

41(2):249–265 (1997).<br />

[68] N. Phan-Thien, “A nonl<strong>in</strong>ear network viscoelastic model,” J. Rheol.,<br />

22(3):259–283 (1978).<br />

[69] N. Phan-Thien <strong>and</strong> R. I. Tanner, “A new constitutive equation derived from<br />

network theory,” J. Non-Newtonian Fluid Mech., 2:353–365 (1977).<br />

[70] J. M. Piau <strong>and</strong> N. E. Kissi, “Measurement <strong>and</strong> modell<strong>in</strong>g <strong>of</strong> friction <strong>in</strong> <strong>polymer</strong><br />

melts dur<strong>in</strong>g macroscopic <strong>slip</strong> at the <strong>wall</strong>,” J. Non-Newtonian Fluid Mech.,<br />

54:121–142 (1994).<br />

[71] J.-M. Piau, N. E. Kissi, F. Toussa<strong>in</strong>t, <strong>and</strong> A. Mezghani, “Distortions <strong>of</strong> <strong>polymer</strong><br />

extrudates <strong>and</strong> their elim<strong>in</strong>ation us<strong>in</strong>g <strong>slip</strong>pery surfaces,” Rheol. Acta.,<br />

34:40–57 (1995).<br />

[72] J.-M. Piau, N. E. Kissi, <strong>and</strong> B. Tremblay, “Influence <strong>of</strong> upstream <strong>in</strong>stabilities<br />

<strong>and</strong> <strong>wall</strong> <strong>slip</strong> on melt fracture <strong>and</strong> sharksk<strong>in</strong> phenomena dur<strong>in</strong>g silicones<br />

extrusion through orifice dies,” J. Non-Newtonian Fluid Mech., 34:145–180<br />

(1990).<br />

[73] G. Pomar, S. J. Muller, <strong>and</strong> M. M. Denn, “Extrudate distortions <strong>in</strong> l<strong>in</strong>ear lowdensity<br />

polyethylene solutions <strong>and</strong> melts,” J. Non-Newtonian Fluid Mech.,<br />

54:143–151 (1994).<br />

[74] K. Porteous <strong>and</strong> M. M. Denn, “L<strong>in</strong>ear stability <strong>of</strong> plane Poiseuille flow <strong>of</strong><br />

viscoelastic liquids,” Trans. Soc. Rheol., 16(2):205–308 (1972).


122<br />

[75] A. V. Ramamurthy, “Wall <strong>slip</strong> <strong>in</strong> viscous fluids <strong>and</strong> <strong>in</strong>fluence <strong>of</strong> materials <strong>of</strong><br />

construction,” J. Rheol., 30(2):337–357 (1986).<br />

[76] A. V. Ramamurthy, “Extrudate irregularities <strong>and</strong> the <strong>polymer</strong>-metal <strong>in</strong>terface<br />

connection,” In Proceed<strong>in</strong>gs from the Xth International Congress on Rheology,<br />

pages 85–90, 1988.<br />

[77] M. Renardy, “Short wave <strong>in</strong>stabilities result<strong>in</strong>g from memory <strong>slip</strong>,” J. Non-<br />

Newtonian Fluid Mech., 35:73–76 (1990).<br />

[78] M. Renardy, “Ill-posedness at the <strong>boundary</strong> for elastic solids slid<strong>in</strong>g under<br />

Coulomb friction,” J. Elasticity, 27:281–287 (1992).<br />

[79] M. Renardy, “On the stability <strong>of</strong> parallel <strong>shear</strong> flow <strong>of</strong> an oldroyd-b fluid,”<br />

Differential <strong>and</strong> Integral Equations, 6(3):481–489 (1993).<br />

[80] M. Renardy, “High Weissenberg number <strong>boundary</strong> layers for the upper convected<br />

Maxwell fluid,” J. Non-Newtonian Fluid Mech., 68:125—132 (1997).<br />

[81] M. Renardy, “Re-entrant corner behavior <strong>of</strong> the PTT fluid,” J. Non-<br />

Newtonian Fluid Mech., 69:99–104 (1997).<br />

[82] M. Renardy <strong>and</strong> Y. Renardy, “L<strong>in</strong>ear stability <strong>of</strong> plane Couette flow <strong>of</strong> an<br />

upper convected Maxwell fluid,” J. Non-Newtonian Fluid Mech., 22:22–33<br />

(1986).<br />

[83] Y. Renardy, “Stability <strong>of</strong> the <strong>in</strong>terface <strong>in</strong> two layer Couette flow <strong>of</strong> upper convected<br />

Maxwell liquids,” J. Non-Newtonian Fluid Mech., 28:99–115 (1988).<br />

[84] S. Richardson, “On the no-<strong>slip</strong> <strong>boundary</strong> condition,” J. Fluid Mech., 59:707–<br />

719 (1973).


123<br />

[85] J. M. H. M. Scheutjens <strong>and</strong> G. J. Fleer, “Statistical theory <strong>of</strong> the adsorption<br />

<strong>of</strong> <strong>in</strong>teract<strong>in</strong>g cha<strong>in</strong> molecules. 1. Partition function, segment density distribution,<br />

<strong>and</strong> adsorption isotherms,” J. Phys. Chem., 83(12):1619–1635 (1979).<br />

[86] J. M. H. M. Scheutjens <strong>and</strong> G. J. Fleer, “Statistical theory <strong>of</strong> the adsorption<br />

<strong>of</strong> <strong>in</strong>teract<strong>in</strong>g cha<strong>in</strong> molecules. 2. Tra<strong>in</strong>, loop, <strong>and</strong> tail size distribution,” J.<br />

Phys. Chem., 84(2):178–190 (1980).<br />

[87] J. D. Shore, D. Ronis, L. Piché, <strong>and</strong> M. Grant, “Model for melt fracture <strong>in</strong>stabilities<br />

<strong>in</strong> the capillary flow <strong>of</strong> <strong>polymer</strong> melts,” Phys. Rev. Lett., 77(4):655–658<br />

(1996).<br />

[88] J. D. Shore, D. Ronis, L. Piché, <strong>and</strong> M. Grant, “Theory <strong>of</strong> melt fracture<br />

<strong>in</strong>stabilities <strong>in</strong> the capillary flow <strong>of</strong> <strong>polymer</strong> melts,” Phys. Rev. E, 55(3):2976–<br />

2992 (1997).<br />

[89] C. L. Sieglaff, “Rheological properties <strong>of</strong> polyv<strong>in</strong>yl chloride,” S. P. E. Transactions,<br />

4:129–138 (1964).<br />

[90] A. Souvaliotis <strong>and</strong> A. N. Beris, “Spectral collocation / doma<strong>in</strong> decomposition<br />

method for viscoelastic flow simulations <strong>in</strong> model porous geometries,”<br />

Comput. Methods Appl. Mech. Engrg., 129:9–28 (1996).<br />

[91] T. Sun, A. C. Balazs, <strong>and</strong> D. Jasnow, “Dynamics <strong>of</strong> phase separation <strong>in</strong><br />

<strong>polymer</strong> solutions under <strong>shear</strong> flow,” Rhys. Rev. E, 55(6):6344–6347 (1997).<br />

[92] R. Sureshkumar <strong>and</strong> A. N. Beris, “Effect <strong>of</strong> artificial stress diffusivity on the<br />

stability <strong>of</strong> numerical calculations <strong>and</strong> the flow dynamics <strong>of</strong> time-dependent<br />

viscoelastic <strong>flows</strong>,” J. Non-Newtonian Fluid Mech., 60(1):53–80 (1995).<br />

[93] G. Tlapa <strong>and</strong> B. Bernste<strong>in</strong>, “Stability <strong>of</strong> a relaxation-type viscoelastic fluid<br />

with slight elasticity,” Phys. Fluids, 13(3):565–568 (1971).


124<br />

[94] J. P. Tordella, “Fracture <strong>in</strong> the extrusion <strong>of</strong> amorphous <strong>polymer</strong>s through<br />

capillaries,” J. App. Phys., 27(5):454–458 (1956).<br />

[95] C. Tzoganakis, B. C. Price, <strong>and</strong> S. G. Hatzikiriakos, “Fractal analysis <strong>of</strong> the<br />

sharksk<strong>in</strong> phenomenon <strong>in</strong> <strong>polymer</strong> melt extrusion,” J. Rheol., 37(2):355–365<br />

(1993).<br />

[96] C. Venet <strong>and</strong> B. Vergnes, “Experimental characterization <strong>of</strong> sharksk<strong>in</strong> <strong>in</strong><br />

polyethylenes,” J. Rheol., 41(4):873–892 (1997).<br />

[97] G. V. V<strong>in</strong>ogradov <strong>and</strong> L. I. Ivanova, “Viscous properties <strong>of</strong> <strong>polymer</strong> melts<br />

<strong>and</strong> elastomers exemplified by ethylene-propylene co<strong>polymer</strong>,” Rheol. Acta,<br />

6:209–222 (1967).<br />

[98] G. V. V<strong>in</strong>ogradov <strong>and</strong> L. I. Ivanova, “Wall <strong>slip</strong>page <strong>and</strong> elastic turbulence <strong>of</strong><br />

<strong>polymer</strong>s <strong>in</strong> the rubbery state,” Rheol. Acta, 7:243–254 (1967).<br />

[99] S.-Q. Wang <strong>and</strong> P. A. Drda, “Superfluid-like stick-<strong>slip</strong> transition <strong>in</strong> capillary<br />

flow <strong>of</strong> l<strong>in</strong>ear polyethylene melts: I. General features,” Macromolecules,<br />

29:2627–2632 (1996).<br />

[100] S.-Q. Wang, P. A. Drda, <strong>and</strong> Y.-W. Inn, “Explor<strong>in</strong>g molecular orig<strong>in</strong>s <strong>of</strong><br />

sharksk<strong>in</strong>, partial <strong>slip</strong>, <strong>and</strong> slope change <strong>in</strong> flow curves <strong>of</strong> l<strong>in</strong>ear low density<br />

polyethylene,” J. Rheol., 40(5):875–898 (1996).<br />

[101] S. Q. Wang <strong>and</strong> N. Plucktaveesak, “Self-oscillations <strong>in</strong> capillary flow <strong>of</strong> entangled<br />

<strong>polymer</strong>s,” J. Rheol., 43(2):453–460 (1999).<br />

[102] J. L. White, M. H. Han, N. Nakajima, <strong>and</strong> R. Brzoskowski, “The <strong>in</strong>fluence<br />

<strong>of</strong> materials <strong>of</strong> construction on biconical rotor <strong>and</strong> capillary measurements <strong>of</strong><br />

<strong>shear</strong> viscosity <strong>of</strong> rubber <strong>and</strong> its compounds <strong>and</strong> considerations <strong>of</strong> <strong>slip</strong>page,”<br />

J. Rheol., 35(1):167–189 (1991).


125<br />

[103] H. J. Wilson, M. Renardy, <strong>and</strong> Y. Renardy, “Structure <strong>of</strong> the spectrum <strong>in</strong><br />

zero Reynolds number <strong>shear</strong> flow <strong>of</strong> the UCM <strong>and</strong> Oldroyd-B liquids,” J.<br />

Non-Newtonian Fluid Mech., 80:251–268 (1999).<br />

[104] D. Wirtz, “Flow-<strong>in</strong>duced scatter<strong>in</strong>g peak <strong>in</strong> the structure factor <strong>of</strong> <strong>polymer</strong><br />

solutions,” Phys. Rev. E, 50(3):1755–1758 (1994).<br />

[105] X.-L. Wu, D. J. P<strong>in</strong>e, <strong>and</strong> P. K. Dixon, “Enhanced concentration fluctuations<br />

<strong>in</strong> <strong>polymer</strong> solutions under <strong>shear</strong> flow,” Phys. Rev. Lett., 66(18):2408–2411<br />

(1991).<br />

[106] A. L. Yar<strong>in</strong> <strong>and</strong> M. D. Graham, “A model for <strong>slip</strong> at <strong>polymer</strong>/solid <strong>in</strong>terfaces,”<br />

J. Rheol., 42(6):1491–1504 (1998).

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

Saved successfully!

Ooh no, something went wrong!