24
12 3 45efghi UNIVERSITY OF WALES SWANSEA REPORT SERIES A General Approach for True Transient Computation of Viscoelastic Flows by I. J. Keshtiban, H. Tammadon-Jahromi and M. F. Webster Report # CSR 10-2006

Transient CSR - Swanseamathematical structure, transient analytical solutions may be obtained in both velocity and extra-stress. This is true for example with constant shear viscosity

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Transient CSR - Swanseamathematical structure, transient analytical solutions may be obtained in both velocity and extra-stress. This is true for example with constant shear viscosity

12345efghi

UNIVERSITY OF WALES SWANSEA

REPORT SERIES

A General Approach for True Transient Computation of Viscoelastic Flows

by

I. J. Keshtiban, H. Tammadon-Jahromi and M. F. Webster

Report # CSR 10-2006

Page 2: Transient CSR - Swanseamathematical structure, transient analytical solutions may be obtained in both velocity and extra-stress. This is true for example with constant shear viscosity

1

A General Approach for True Transient Computation of Viscoelastic Flows

I. J. Keshtiban, H. Tammadon-Jahromi and M. F. Webster *

Institute of non-Newtonian Fluid Mechanics,

Department of Computer Science,

University of Wales Swansea,

Singleton Park, SA2 8PP,

Swansea, United Kingdom.

Abstract

This article considers the problem of solving true-transient flows in planar

contractions for Oldroyd-B and Pom-pom fluids. This is conducted through both

controlled force and controlled flow-rate configurations, with comparison therein,

considering consistency issues. To accomplish this task we employ a novel time-

dependent hybrid finite element/finite volume (fe/fv) algorithm. Such an approach

combines an incremental pressure-correction fe-treatment for mass and momentum

equations, with a cell-vertex fv-discretisation of the hyperbolic stress constitutive

equation. In particular, we highlight differences in dynamic flow structure evolution

within the field and its impact upon stress generation, as a consequence of the choice

of driving flow conditions. The advantages of smooth flow transitions are highlighted

under controlled-force configurations, in contrast to non-smooth states for controlled

flow-rate scenarios.

1. Introduction

The numerical simulation of truly transient complex flows of viscoelastic

materials is an important area of study in many industrial and natural processes. This

is due to the fact that the nature of common flows of complex materials are essentially

time-dependent. Hence, many important viscoelastic phenomena arise in the transient

context. To conduct true-transient simulations for viscoelastic flows, it is important to

develop an accurate and robust numerical procedure that can handle transient

boundary conditions and appropriate constitutive equations. Here, for Oldroyd-B and

Pom-pom fluids and planar rounded-corner contraction flows, we employ a novel

time-marching hybrid finite element/finite volume (fe/fv) algorithm. This combines

an incremental pressure-correction fe-treatment for mass conservation and momentum

transport equations, with a cell-vertex fv-discretisation of the hyperbolic stress

constitutive equation. As a consequence of the choice of driving flow conditions, we

highlight differences in dynamic flow structure evolution within the field, say in

vortex development, and indicate the impact this has upon stress generation. Both

controlled force and controlled flow-rate configurations are adopted to develop

transient inflow-outflow boundary conditions.

* Author for correspondence. Email: [email protected]

Page 3: Transient CSR - Swanseamathematical structure, transient analytical solutions may be obtained in both velocity and extra-stress. This is true for example with constant shear viscosity

2

The governing equations of motion of viscoelastic fluids encompass the continuity

and momentum equations, with a constitutive equation which reflects material

response under deformation. These space-time partial differential equations (PDE) are

non-linear and of mixed-type, parabolic-hyperbolic. In transient computations,

imposing a consistent set of boundary conditions (BCs) is essential to extract a

meaningful solution. Nevertheless, incorporating a consistent set of transient BCs is

often a non-trivial task, with a lack of analytical solutions for complex flows. In this

article we introduce a generalised approach through which we define consistent

transient BCs for viscoelastic flows independent of the frame of reference.

For viscoelastic flows, two main approaches have been deployed in the literature to

incorporate inflow and outflow type BCs. Typically, these are considered as simple

shear-flow conditions, reducing to one-dimensional ODEs for both velocity and

stress. In the first approach and for some constitutive equations of simple

mathematical structure, transient analytical solutions may be obtained in both velocity

and extra-stress. This is true for example with constant shear viscosity

Maxwell/Oldroyd-B models in a channel flow [1]. These solution profiles, or their

limiting steady-state form, may then be imposed at inlet for both velocity and stress,

and at outlet in velocity alone. This is the controlled flow-rate scenario commonly

encountered. The second approach is a controlled force-driven configuration, see

Bodart and Crochet [2] for the transient simulation of the falling sphere problem.

Force-control may be organized through the application of a controlled pressure-drop

across a flow (say impulsively), built into the problem via force-type BCs (boundary

integrals in a weak form solution). We note in passing that experimentally, it may

prove more difficult to generate controlled flow-rate driven conditions than those

under controlled force.

Under the controlled flow-rate instance, the application of steady-state BCs is often

inconsistent with the interior evolving solution. We proceed to demonstrate that the

standard practice of imposing steady-state BCs in contraction flows may result in

large predicted transient oscillations in pressure, which may deteriorate the accuracy

of the evolving solution. To rectify this position, we may consider the force-driven

setting and mimic more closely the actual physics of the flow. We note that many

flows arise under driving force conditions, such as under pressure-difference or

gravity. As such, a key aspect here is to impose a pressure-difference across the flow

as a boundary condition, and compute velocity and stress on boundaries. To this end,

we employ a boundary integral formulation on open boundaries to incorporate the

associated kinematics. Inlet flow conditions are considered as pure shear flow, in

such a manner so that convection terms vanish and stress can be computed without

upstream reference. This approach has several distinct advantages. It is independent of

the coordinate system and constitutive equation. More importantly, this approach is

consistent with the flow dynamics and is straightforward to reproduce, both

experimentally and numerically. In addition as we shall demonstrate, such an

approach does not result in large pressure oscillations, so that the flow is observed to

smoothly propagate.

2. Governing equations and flow problem

Incompressible viscoelastic flows are governed by equations for mass conversation

and momentum transport, with a constitutive law for stress. In non-dimensional form,

the balance equations under isothermal conditions may be represented as

Page 4: Transient CSR - Swanseamathematical structure, transient analytical solutions may be obtained in both velocity and extra-stress. This is true for example with constant shear viscosity

3

.( ) 0∇ =u 2.1

Re( . ) .( 2 ) ,pt

β∂

+ ∇ = ∇ + −∇∂u

u u dττττ 2.2

We have recourse to both Oldroyd-B and Pom-pom models, and hence, employ

the following generalised constitutive law. This is appropriate also for the Single

Extended Pom-pom (SXPP) model, which requires three additional

parameters ( ), ,qε λ for system entanglement, number of side-branch arms to the

back-bone segment, and stretch of the back-bone segment. This generalized equation

takes the form:

( ) ( ) ( )( )

( )1 We

f , We f , 1 2 11We

β αλ λ β

β

∇ − + + − + ⋅ = − −

I dτ τ τ τ τ ττ τ τ τ τ ττ τ τ τ τ ττ τ τ τ τ τ . 2.3

Here, d= †( ) / 2∇ ∇u + u denotes the rate of deformation and I the unit tensor.

Furthermore, τ∇

represents the upper-convected material derivative of ττττ , and f(λ,ττττ) a generalised function, viz.

†. ( . .( ),t

ττ τ τ τ∇ ∂= + ∇ − ∇ + ∇∂

u u) u 2.4

( ) ( )

( )( )

2

1

2

2 1 1 Wef , 1 e 1 tr

1 3

ν λ αλ

ε λ λ β−

= − + − ⋅ −

τ τ ττ τ ττ τ ττ τ τ . 2.5

In the above, ν is taken as ( )2 /q and the back-bone stretch parameter λ is identified

as:

( )( )1 We

1 tr3 1

λβ

= +−

ττττ . 2.6

The Pom-pom model has been derived to characterize the rheological behaviour of

polymer melts with long side-branches, such as applicable for low density

polyethylene [3,4]. The Oldroyd-B model corresponds to setting f(λ,ττττ)=1 and the anisotropy parameter, α=0.

The conventional definitions of group numbers, in Reynolds (Re) and Weissenberg

(We) numbers, with parameters β and ε are defined as:

ReULρµ

= , 0e bUWL

λ= ,

0 0

s

s bG

µβ

µ λ=

+, 0

0

s

b

λε

λ= .

In this notational form, λ0b and λ0s represent the orientation and backbone stretch

relaxation time-scales, respectively. Go is the linear relaxation modulus and parameter

ε is the ratio of stretch to orientation relaxation times. The parameter ν is found to be

inversely proportional to the number of side-branch arms to the molecular chain-

segment (q), Blackwell et al [5]. ρ is the fluid density and for contraction flows, the

characteristic velocity (U) and length (L) scales are taken as the downstream mean

velocity and channel height, respectively. µ is the fluid viscosity, that splits into

additive contributions from a polymeric, pµ , and Newtonian solvent component sµ .

To satisfy common convention, a ratio of solvent to total viscosity is taken as

1/ 9β = .

Page 5: Transient CSR - Swanseamathematical structure, transient analytical solutions may be obtained in both velocity and extra-stress. This is true for example with constant shear viscosity

4

In this study, the benchmark flow problem is creeping flow with a level of We =1

within a 4:1 planar, rounded-corner contraction geometry (Fig. 2). The length of

upstream and downstream sections are 27.5L and 49L, respectively. Three different

boundary condition settings are imposed at inlet and outlet flow stations. The first

equates to steady-state flow-rate (Q) control (fully-developed conditions at inlet and

outlet). The second setting is transient flow-rate ( tQ ) controlled, where inlet and

outlet velocities are ensured through a Waters and King transient (W&K) solution [1].

The third setting is that of force-controlled (∆p ) flow, where a fixed pressure-

difference is imposed in time across the domain (inlet-outlet). No-slip boundary

conditions are applied along the stationary walls under all three settings. Transient

inlet stress representation can be extracted in weak-form by solving the equivalent

ODE system, in situ, under shear flows conditions. In this study, the level of pressure-

difference for the force-controlled setting is considered to be that equivalent to the

established steady-state pressure-drop under Q-control. This facilitates equitable

comparison across both transient and steady-state solutions.

3. Solution algorithm

3.1 Numerical methods

During recent years, significant progress has been made in both developing new

constitutive equations for complex materials and the computation of complex

viscoelastic flows. Reliable, accurate and consistent steady-state solutions through

various formulations and numerical methods have been obtained for a number of

benchmark problems. In this manner, highly elastic solutions have been investigated.

Achievable level of Weissenberg number (We) has been elevated significantly.

Despite this advance, computation of true-transient viscoelastic flows has received

scant attention in the literature. This may be due to the inherent difficulty of the

problem, posed by numerical discretisation and the application of boundary

conditions, and may be felt in terms of consistency and accuracy properties. It is well

known that the accurate determination of the transient in unsteady incompressible

viscous flow, based on a primitive variable formulation, is a challenging task. A

principal difficulty encountered here is due to the mixed-type of equations, arising

through the coupling of the momentum equation with the incompressibility equation,

and subsequently the treatment of the pressure term. Moreover, in the viscoelastic

context, introducing an additional hyperbolic extra-stress equation subset dramatically

increases the numerical tractability of the problem.

After setting boundary conditions, a choice of numerical solver must be made.

There are numerous ways to discretise the time-dependent incompressible flow

problem. A popular selection has been to employ pressure-correction or so-called

projection schemes, see Chorin [6] and Temam [7]. This method gained vast

popularity in solving incompressible flows. A main advantage of this approach lies in

size reduction, based on decoupling the computation of velocity and pressure, via

fractional-staging on each time-step. On each time-step, first a set of convection-

diffusion equations are solved for velocity, excluding (or guessing) the pressure in the

momentum equation. Then, the forward time-step pressure is updated by imposing the

continuity constraint through a Poisson equation. At a third stage, the computed

velocity from the first fractional-stage is projected onto a divergence-free subspace,

incorporating the updated pressure field. This strategy is computationally effective

when compared to that for handling coupled systems. Through the course of time, it

Page 6: Transient CSR - Swanseamathematical structure, transient analytical solutions may be obtained in both velocity and extra-stress. This is true for example with constant shear viscosity

5

has become apparent that projection schemes should be analyzed from a fully-discrete

point of view. This is due to the fact that fully-discrete systems (in both time and

space) do not retain all the properties of their continuous semi-discrete counterparts

(discretised in time only). Hence, one should be most careful when analyzing

numerical schemes based on solely semi-discrete formulations, in either projection-

form or coupled statement of the problem. This is the case, for example, in the work

of Xue et al. [8], claiming that their PISO and SIMPLST pressure-correction schemes

for viscoelastic flows have temporal orders of ( )3O t∆ for pressure and ( )4O t∆ for

velocity. However, even for simple Newtonian flows, such orders of temporal

accuracy may be practically unattainable. See, likewise the quoted second-order

theoretical pressure-correction scheme of Webster et al.[9], which slightly declines

from this level in practice.

In the computation of viscoelastic flows, increasing elasticity levels drastically

deteriorates scheme stability. To rectify this situation various stabilisation procedures

have been introduced, such as stress-splitting, say via the Elastic Viscous Spitting

Scheme (EVSS)[10], where the level of achievable We has been elevated

dramatically. Here, some elliptic properties are introduced into the system through

stress-splitting; yet, the question arises as to whether this adversely affects the

transient flow development. In addition, Belbelidia et al. [11], have demonstrated that

the differed-correction stabilisation term, appearing in DEVSS formulations [12],

provides a localized signature for the stress corner singularity in a 4:1 abrupt

contraction. This contribution (independent of time) can affect the solution through

such elusive flow features, as the appearance and formation of lip-vortices. However,

such stabilization does not affect the solution through the salient-corner vortex. In

addition, schemes such as DEVSS/DG have been shown to demonstrate temporal

instabilities in some smooth flows [13]. This may be attributed to the particular

upwinding characteristics of the DG component of the scheme. Furthermore,

improving the satisfaction of inf-sup conditions is found to have a significant impact

on numerical stability, see Belbelidia et al. [11]. In this regard, Matallah et al. [14]

have introduced a local recovery technique for velocity-gradients, which contrasts

against the global weighted-residual counterpart used within DEVSS. These authors

concluded that such local approaches are competitive to the widely used global

DEVSS-variants, and may be gainfully employed within the transient regime.

3.2 Hybrid fe/fv scheme

The general framework of the time-marching hybrid fe/fv scheme employed here

involves two distinct aspects. First, velocity and pressure are computed via a semi-

implicit incremental pressure-correction procedure with finite element spatial

discretization. Secondly, a finite volume based fluctuation distribution scheme is

adopted for computation of hyperbolic extra-stress equations. The algorithm consists

of a two-step Lax-Wendroff time-stepping procedure, extracted via a Taylor series

expansion in time. Here, first velocity and stress components are predicted to a half

time-step (Stage 1a), and then, updated over the full time-step (Stage 1b). To ensure

the incompressibility constraint, pressure at the forward time-step is derived from a

Poisson equation (Stage 2), with velocity corrected at a final stage (Stage 3) to satisfy

continuity. The three-stage algorithmic structure per time-step may be expressed in

semi-discrete form with new boundary integral terms highlighted, as follows [15,16]:

Page 7: Transient CSR - Swanseamathematical structure, transient analytical solutions may be obtained in both velocity and extra-stress. This is true for example with constant shear viscosity

6

Stage1a:

3.2.1

3.2.2

Stage1b:

3.2.3

3.2.4

Stage2:

3.2.5

Stage3:

3.2.6

Interpolation on the parent fe-triangular element for velocity-pressure is of

piecewise continous quadratic-linear (P2P1) form. The extra-stress equations have

hyperbolic character, which demand upwinding procedures for their effective spatial

discretization. Here, inappropriate procedures would adversely affect scheme

accuracy and stability. The resolution of stress boundary layers, which form on no-

slip solid boundaries, has proved a major obstacle to successful viscoelastic

computations at high We. The effect of numerical noise within and across such thin

stress boundary layers often poses severe discretisation and consequent convergence

difficulties. We have observed that upwinding in the form of SUPG suppresses cross-

stream propagation and consequently adjusts the evolution and dynamics of the stress

boundary layer. Xue et al. [8] have also observed that high-order upwinding in the

form of a QUICK-scheme may degrade temporal stability. In the course of time, good

performance characteristic in terms of accuracy and computation cost for hyperbolic

equations have been realized through upwinding procedures based on finite volume

dicretisation. In the light of this, the current work adopts a cell-vertex finite volume

approach based on fluctuation distribution constructs for the computation of extra-

stress. This algorithm has been discussed extensively in the literature and we refer the

reader to [9,15,16]. Nevertheless, for completeness and self-contained reading, here

we present a brief description below.

3.3. Finite volume fluctuation distribution scheme

Cell-vertex fv-scheme applied to extra-stress are based upon an upwinding

technique (fluctuation distribution), that distributes control volume residuals to

provide nodal solution updates. Concisely, by rewriting the extra-stress equation in

non-conservative form, with flux ( ,τu.∇=R ) and absorbing remaining terms under

the source (Q), one may obtain:

( ) ( ) ( )1/2 1 122Re, , ,2 2n n T n n n n n n nsu u D p p Ip Ip ndt

µν ν τ ν ν τ

µ+ − −

Γ

− = ∇ + + ∇ − + + − ⋅ Γ ∆

( ) ( ) ( ) ( ) ( )( )

1/212 We

2 1 f , f , 11

n

n n TWeWe u u u

t We

β ατ τ β τ τ τ λ λ

β+

− − = − − ⋅∇ − ⋅∇ −∇ ⋅ − − − + ⋅ ∆ −

d Iτ τ τ τ ττ τ τ τ ττ τ τ τ ττ τ τ τ τ

( ) ( ) ( )* 1/2 1 1/2 122Re, , ,2 2

n T n n T n n n n nsu u D p p Ip Ip ndt

µν ν τ ν ν τ

µ+ − + −

Γ

− = ∇ + + ∇ − + + − ⋅ Γ ∆

( ) ( ) ( ) ( ) ( )( )

1/2

11 We

2 1 f , f , 11

n

n n TWeWe u u u

t We

β ατ τ β τ τ τ λ λ

β

+

+ −

− = − − ⋅∇ − ⋅∇ −∇ ⋅ − − − + ⋅ ∆ − d Iτ τ τ τ ττ τ τ τ ττ τ τ τ ττ τ τ τ τ

( )( ) ( )( ) ( )1 * 11, ,T n n n nq p p q u q p p n d

tθ ρ+ +

Γ

∇ ∇ − = − ∇ ⋅ + − ⋅ Γ∆ ∫

( )( ) ( )( ) ( )1 1 1 1Re, ,n n T n n n nu u p p p p n d

tν ρ θ ν ν+ + ∗ + +

Γ

− = ∇ ∇ − + − ⋅ Γ∆ ∫

Page 8: Transient CSR - Swanseamathematical structure, transient analytical solutions may be obtained in both velocity and extra-stress. This is true for example with constant shear viscosity

7

t

∂+ =

∂ττττ

R Q . (3.3.1)

We consider each scalar stress component, τ, acting on an arbitrary

volumel

l

Ω = Ω∑ , whose variation is controlled through corresponding components

of fluctuation of the flux vector (R) and the source term (Q),

l l l

d Rd Qdt

τΩ Ω Ω

∂Ω = Ω+ Ω

∂ ∫ ∫ ∫ . (3.3.2)

The objective is to evaluate these flux and source variations over each finite volume

triangle, with distribution to its three vertices according to the preferred strategy. The

nodal update for a given node (l) is obtained by accumulating the contributions from

its control volumeΩl, composed of all fv-triangles surrounding node (l). The flux and

source residuals may be evaluated over different control volumes associated with a

given node (l) within the fv-cell T; namely, over the fv-triangle T, (RT,QT), and its

subtended median dual cell form, (Rmdc, Qmdc), providing scheme variant CT3 (with

mdc), formulated in [9,17] Fig. 1 provides detailed interconnectivity of the fv-

triangular cells (iT ) about the sample node (l), the blue-shaded zone of mdc, the

parent triangular fe-cell, and the fluctuation distribution (fv-upwinding) parameters

( T

iα ).

4. Numerical results

Fig. 3a charts the centerline-velocity against time, monitored under the three

different boundary condition settings for the Oldroyd-B model. Here, we observe

three distinct evolution patterns before the same limiting steady-state solutions are

established. Transient flow-rate controlled (W&K) [1], setting (2), shows pronounced

overshoots/undershoots, see also [9,17] for more details. Force-controlled flow,

setting (3), displays a smooth flow evolution pattern with delayed

overshoots/undershoots. In this case, smaller magnitudes in peak-values are observed

in comparison to the transient flow-rate controlled instance. We also report in Fig. 3b,

the transient velocity profile monitored at a sample point around the contraction.

Here, similar trends in overshoot/undershoot and relative peak-magnitudes are

discerned as those for velocity inlet, under respective BC settings (2) and (3), transient

flow rate and pressure controlled. Under the steady-state flow-rate control BC setting,

although the inlet-velocity profile is fixed, an undershoot/overshoot is observed for

the velocity point around the contraction. In addition, we present in Fig. 3c and 3d the

wall stress profiles for τxx at inlet and at the sample point around the contraction,

respectively. Here, large stress peaks are now observed for the sample point around

the contraction (O(19) units) in comparison to the stress-peaks at inlet (O(0.15) units).

Overshoot/undershoot patterns are evident with BC setting (2) and (3) both at inlet

and around the contraction. In contrast, for the steady-state Q-setting, a monotonic

increase in stress at the sample point around the contraction is prominent before

reaching its plateau.

4.1 Development of vortex structure

Fig. 4 demonstrates the transient evolution of streamlines from on early stage of

simulation (t=0.001 units) towards the steady-state solution (t=10 units) utilising the

Oldroyd-B model.

Page 9: Transient CSR - Swanseamathematical structure, transient analytical solutions may be obtained in both velocity and extra-stress. This is true for example with constant shear viscosity

8

4.1.1 Early time development stage

We observed that the flow is more developed in the initial stages (t=0.001 and

t=0.01) for the steady-state Q-controlled case in comparison to other instances. At

these early times, both force-controlled flow and transient Q-controlled states exhibit

similar trends. The first signs of salient-corner vortex activity appear at time t=0.1,

under steady-state Q-controlled. At this point, no salient-corner vortex is observed for

the other two transient settings. This is delayed to time, t=0.3, whereupon this feature

becomes visible for both transient flow-rate and force-controlled flows. A growth

trend in both salient-corner vortex intensity and cell-size now sets in by t=0.5 and

t=0.6. From rest and up to this stage of t=0.6, steady-state Q-controlled and force-

controlled BC-settings display the most rapid and slowest rates of evolutionary

development in velocity, respectively.

4.1.2 Mid-term development stage

Upon arriving at time of t=0.8, the first departure in shape and salient vortex is

evident. Here, a reverse-flow region begins to emerge along the upstream-wall in the

transient Qt-controlled case. This corresponds to half-way between the first overshoot

and undershoot in the centerline velocity, shown in Fig 3a, see Webster et al. [9]. At

this juncture, the salient-corner vortex increases once more, under both force (∆p )-controlled and steady-state Q-controlled settings, in comparison to their earlier times.

At t=1.0, this reversed flow intensifies. A change from convex to concave shape is

visible at t=0.8 and t=1.0 under steady-state Q-controlled form.

We observe an eye-like secondary vortex for the transient Qt-controlled instance

at t=2.0 (corresponding to half-way between the first undershoot and second

overshoot in centerline velocity, Fig. 3a). At this juncture, vortex enhancement is

prominent for the force-controlled setting, corresponding to a position after the first

overshoot in centerline velocity in Fig. 3a. Further vortex enhancement is detectable

at t=3.0 under the force-controlled setting, where it reaches its largest salient-corner

cell-size and intensity. At this moment, a clear concave shape is apparent with the

same cell-size and intensity under both transient settings. Note that, centreline inlet

velocity at t=3 practically coincides for these two situations.

4.1.3 Long-term development stage

At time of t=4.0 and beyond a different phase is entered where vortex reduction

is recognizable under all three instances. Here, transient Qt-controlled and steady-state

Q-controlled cases attract the smallest and largest vortex size, respectively. From t=5

onwards (to t=10), salient-corner vortices are practically identical having reached their

limiting state under the steady-state Q-controlled evolution. The convex structure of

the separation line adjusts to a linear shape under the transient Q-controlled BC from

t=5.0 to t=6.0. This variation is completely contrary to that under the force-controlled

instance, where linear separation-line shape alters to convex form. A further

difference is observed in vortex activity for these two types of boundary condition

settings between t=5 to t=6. That is, vortex intensity and cell-size decreases under

transient Qt-control, yet increase under force-control. Finally, identical vortex

intensity and cell-size with convex shapes are gathered with all three BC setting from

t=7 to t=10, as shown in Fig. 4.

Page 10: Transient CSR - Swanseamathematical structure, transient analytical solutions may be obtained in both velocity and extra-stress. This is true for example with constant shear viscosity

9

4.2 Trends in velocity, stress and pressure fields

The transient development of the velocity field is illustrated in Fig. 5 from rest up

to t=10 at representative phases in-between. Under steady-state flow rate setting, a

conspicuous inconsistency at outlet and inlet with interior flow field is revealed from

the outset and over the early stage of evolution (indicated by arrows). This irregularity

continues up to time of t=1.0, due to the independence of the description of boundary

conditions from that of interior core flow fields. Similar behaviour is apparent under

the transient Qt-controlled setting, starting from t=0.2 until t=2.0. Evolution of the

flow field under forced-driven setting is observed to be much smoother than for the

other two flow-rate controlled alternatives. Finally, identical steady-state solutions are

established with all three BC settings at the limiting time of t=10 (Fig. 5).

In Fig. 6, the corresponding transient development of the stress field is displayed.

Here again, force-driven flow evolution is much smoother than under the other two

flow-rate-controlled instances. There are three observations to make here. First, the

stress develops in time subject to a time-lag, when compared to the evolving

kinematic fields. Second, the strong kinematic undershoot/overshoot present in the

transient Qt-controlled instance, is reflected throughout the stress states. Third, the

force-driven scenario contains a time-lag within both velocity and stress emerging

solutions, over their counterpart flow-rate controlled cousins. There is ample evidence

here for preference towards the smoother solution development of force-driven flows,

devoid of oscillatory patterns even around the complex corner flow region.

Fig. 7 illustrates the transient evolution of pressure fields during the early stages

of the simulation. Now, for direct-comparison, fields of displayed superimposed

within a single plot. Extremely large oscillations between positive and negative values

in pressure (O(106)) are observed under the steady-state Q-driven setting. As a

consequence, the velocity field also displays large oscillations. The transient Qt-

control setting demonstrates smoother oscillations in developing pressure, of the order

of O(103), whilst only mild variation is detected under the force-driven setting of

O(102) see Fig. 7.

Incremental continuation in We-solutions is a common practice in the simulation

of highly-elastic flows, in order to reach elevated levels of steady-state We-solutions.

In this approach, first a steady-state solution for the problem is extract at a low level

of We. Subsequently, this solution is considered as the initial condition for the next

We-increment solution, and the process is repeated to one step higher in We-level.

This scenario is continued until a maximum attainable We is reached for the problem

in hand. Here, we illuminate the rational behind the success of this procedure. To this

end and under Qt-controlled setting, we select the first step We=0.1, a second step to

We=0.5, and a final step to We=1.0. Fig. 4 demonstrates the contrasting position

between solutions extracted via this incremental continuation approach and that with

its true-transient counterpart. Fig. 4a provides the plot for the kinematics with the

incremental procedure (red lines), which is much smoother than under the transient

procedure. Increasing We through this incremental continuation process prevents

large oscillations appearing in the stress field. Generation of large values in stress is

believed to be a major source of numerical failure in the computation of high We-

solutions. In addition, in the incremental scenario, after extracting the solution for the

Page 11: Transient CSR - Swanseamathematical structure, transient analytical solutions may be obtained in both velocity and extra-stress. This is true for example with constant shear viscosity

10

first step We, the velocity and stress field has not changed dramatically by increasing

one further step in We. As such, this distinguishes the various forms of kinematic

response in overshoot/undershoot phenomena at high We. For true-transient

computations, overshoots in both velocity and stress of this form may stimulate a high

We-region of local nature, and thus, disturb the stability of the scheme. Hence, it is

not surprising that continuation in We (a false-transient procedure) tends to be more

successful at reaching higher We steady-states than does the Waters and King tQ -

controlled setting.

4.3 SXPP model

We continue the above approach of force-driven BCs to illustrate application in

the wider context involving the SXXP-model, which generally lacks analytical BC

provision. Following article [18], and as a typical case, computations have been

performed using a single set of parameters, that is q=2, ε=1/3, α=0.15 and β=1/9. Fig.

9 (right) displays the evolution of velocity and stress (τxx) at a sample point around the

contraction for this SXPP instance, using both forced-controlled BC and transient Qt-

controlled BC (W&K). Encouragingly, identical steady-state solutions are once again

extracted for both cases, yet the development of the evolving solutions are quite

different. For the forced-driven instance, we observe a monotonic increase both in

velocity and primary stress component, τxx. On the contrary under transient Qt-

controlled setting, overshoot/undershoots in the solution are present before being

gradually dampened out. Again, as shown in Fig. 9 (left), practically identical steady-

state solution are represented in both velocity and (τxx) stress component.

5. Conclusion

In this article a new and general approach for determining transient and consistent

BCs for viscoelastic flows has been developed. This approach is independent of frame

of reference, dimension of the problem (2D or 3D) and complexity of the constitutive

equation. The current approach has been contrasted with more conventional forms of

boundary condition definition, through comparison of transient and steady-state

solutions for the test case of a 4:1 rounded-corner contraction flow problem. Three

different boundary condition settings have been imposed at inlet and outlet flow

stations. The first equates to steady-state flow-rate control. The second setting is

transient flow-rate controlled, where inlet and outlet velocities are ensured through a

Waters and King transient solution. The third setting is that of force-controlled flow,

where a fixed pressure-difference is imposed in time across the domain (inlet-outlet).

We have demonstrated that boundary conditions have a significant impact on the

transient evolution of the flow.

Indeed, some transient flow phenomena reported in the literature may prove to be

solely due to the type of boundary condition imposed, and unphysical thereby. The

overshoot/undershoot kinematics observed under any particular flow setting are linked

to the boundary conditions considered. It is also demonstrated that under flow-rate

controlled alternatives, large oscillations are reflected in pressure that consequently

degrade the accuracy of the computed velocity and stress fields. Incremental

continuation through We is a common practice in achieving high We-solutions. Here,

we elaborate that this practice disconnects the dynamics between velocity and stress,

which prevents highly-elastic regions from developing in the evolving true-transient

solution. In addition, we demonstrate some of the practical difficulties that may arise

in the computation of transient viscoelastic flows. The evolution of force-controlled

Page 12: Transient CSR - Swanseamathematical structure, transient analytical solutions may be obtained in both velocity and extra-stress. This is true for example with constant shear viscosity

11

flow fields are displayed as much smoother patterns than for their counterpart flow-

rate controlled equivalents. This may prove of significant advantage in the drive

towards accurate transient solution determination.

Page 13: Transient CSR - Swanseamathematical structure, transient analytical solutions may be obtained in both velocity and extra-stress. This is true for example with constant shear viscosity

12

REFERENCES

[1] N.D. Waters and M.J. King, Unsteady flow of an elastico-viscous liquid,

Rheology Acta 9 (1970) 345-355.

[2] C. Bodart and M.J. Crochet, The time-dependent flow of a viscoelastic fluid

around a sphere, Journal of Non-Newtonian Fluid Mechanics 54 (1994) 303.

[3] N.J. Inkson, T.C.B. McLeish, O.G. Harlen and D.J. Groves, Predicting low

density polyethylene melt rheology in elongational and shear flows with "pom-

pom" constitutive equations, Journal of Rheology 43 (1999) 873-896.

[4] T.C.B. MacLeish and R.G. Larson, Molecular constitutive equations for a class

of branched polymers: The pom-pom polymer, Journal of Rheology 41 (1998)

81-110.

[5] R.J. Blackwell, T.C.B. McLeish and O.G. Harlen, Molecular drag-strain

coupling in branched polymer melts, Journal of Rheology 44 (2000) 121-136.

[6] A.J. Chorin, A numerical method for solving incompressible viscous flow

problems, Journal of Computational Physics 2 (1967) 12-26.

[7] R. Temam, Sur l'approximation de la solution de Navier-Stokes par la méthode

des pas fractionnaires, Archiv. Ration. Mech. Anal. 32 (1969) 377-385.

[8] S.C. Xue, R.I. Tanner and N. Phan-Thien, Numerical modelling of transient

viscoelastic flows, Journal of Non-Newtonian Fluid Mechanics 123 (2004) 33.

[9] M.F. Webster, H.R. Tamaddon-Jahromi and M. Aboubacar, Transient

viscoelastic flows in planar contractions, Journal of Non-Newtonian Fluid

Mechanics 118 (2004) 83.

[10] D. Rajagopalan, R.C. Armstrong and R.A. Brown, Calculation of steady

viscoelastic flow using a multimode Maxwell model: application of the

explicitly elliptic momentum equation (EEME) formulation, Journal of Non-

Newtonian Fluid Mechanics 36 (1990) 135-157.

[11] F. Belblidia, I.J. Keshtiban and M.F. Webster, Stabilised computations for

viscoelastic flows under compressible implementations, Journal of Non-

Newtonian Fluid Mechanics 134 (2006) 56.

[12] R. Guénette and M. Fortin, A new mixed finite element method for computing

viscoelastic flows, Journal of Non-Newtonian Fluid Mechanics 60 (1995) 27-

52.

[13] A.C.B. Bogaerds, M.A. Hulsen, G.W.M. Peters and F.P.T. Baaijens, Time

dependent finite element analysis of the linear stability of viscoelastic flows

with interfaces, Journal of Non-Newtonian Fluid Mechanics 116 (2003) 33.

[14] H. Matallah, P. Townsend and M.F. Webster, Recovery and stress-splitting

schemes for viscoelastic flows, Journal of Non-Newtonian Fluid Mechanics 75

(1998) 139-166.

[15] M. Aboubacar and M.F. Webster, A cell-vertex finite volume/element method

on triangles for abrupt contraction viscoelastic flows, Journal of Non-Newtonian

Fluid Mechanics 98 (2001) 83-106.

[16] P. Wapperom and M.F. Webster, A second-order hybrid finite-element/volume

method for viscoelastic flows, Journal of Non-Newtonian Fluid Mechanics 79

(1998) 405-431.

Page 14: Transient CSR - Swanseamathematical structure, transient analytical solutions may be obtained in both velocity and extra-stress. This is true for example with constant shear viscosity

13

[17] M.F. Webster, H.R. Tamaddon-Jahromi and M. Aboubacar, Time-dependent

algorithm for viscoelastic flow-finite element/volume schemes, Numerical

Methods for Partial Differential Equations 121 (2005) 272-296.

[18] J.P. Aguayo, H.R. Tamaddon-Jahromi and M.F. Webster, Extensional response

of the pom-pom model through planar contraction flows for branched polymer

melts, Journal of Non-Newtonian Fluid Mechanics 134 (2006) 105.

Page 15: Transient CSR - Swanseamathematical structure, transient analytical solutions may be obtained in both velocity and extra-stress. This is true for example with constant shear viscosity

14

fe-cell

nodes (τ)

fv-cell

midside nodes (u, τ)

vertex nodes (p, u, τ)

l(mdc)

T1

T2

T3

T4

T5

T6

αi

T

j

lk

j

k

Fig. 1. Spatial discretisation: fe-cell with four fv-subcells and fv control volume for

node l with median-dual-cell (shaded)

Fig. 2. Planar 4:1 rounded-corner domain: section of contraction-flow mesh

Page 16: Transient CSR - Swanseamathematical structure, transient analytical solutions may be obtained in both velocity and extra-stress. This is true for example with constant shear viscosity

15

c)

τxx, U

x

Ux

τxx

X

d)

Time

Ux

0 2 4 6 8 10 120

0.2

0.4

0.6

0.8

1

1.2

Steady-state Q-control

Transient Qt-control (W&K)

Force ∆p-control

Transient Centreline-Velocity, Inlet

(1)

(3)

(2)

Time

Ux

0 2 4 6 8 10 120

0.2

0.4

0.6

Steady-state Q-control

Transient Qt-control (W&K)

Force ∆p-control

Transient Around Contraction-Velocity

(2)

(1)

(3)

Time

τ xx

0 2 4 6 8 10 120

0.05

0.1

0.15

0.2

Steady-state Q-control

Transient Qt-control (W&K)

Force ∆p -control

Transient Wall-Stress, Inlet

(1)

1

(2)

(3)

Time

τ xx

0 2 4 6 8 10 120

3

6

9

12

15

18

Steady-state Q-control

Transient Qt-control (W&K)

Force ∆p -control

Transient Around Contraction-Stress

(3)

(2)

(1)

Fig. 3. Transient development of velocity and stress (τxx), Oldroyd-B, We=1

a)

b)

Page 17: Transient CSR - Swanseamathematical structure, transient analytical solutions may be obtained in both velocity and extra-stress. This is true for example with constant shear viscosity

16

Fig. 4. Transient streamlines, 4:1 contraction, Oldroyd-B, We=1

t=0.001

Steady-state Q-control

t=0.6

t=0.8

Transient Qt-control

Force ∆p-control

t=0.2

t=0.5

t=0.3

t=0.1

t=0.01

Page 18: Transient CSR - Swanseamathematical structure, transient analytical solutions may be obtained in both velocity and extra-stress. This is true for example with constant shear viscosity

17

Fig. 4. (Continued) Transient streamlines, 4:1 contraction, Oldroyd-B, We=1

Steady-state Q-control Transient Qt-control Force ∆p-control

t=1.0

t=2.0

t=4.0

t=3.0

t=5.0

t=6.0

t=7.0

t=10.0

Page 19: Transient CSR - Swanseamathematical structure, transient analytical solutions may be obtained in both velocity and extra-stress. This is true for example with constant shear viscosity

18

Steady-state Q-control Transient Qt-control Force ∆p-control Steady-state Q-control Transient Qt-control Force ∆p-control

Fig. 5. Transient development of velocity field, 4:1 contraction, Oldroyd-B, We=1

t=2.0

t=1.0

t=0.01

t=0.2

t=0.3

t=0.5

t=0.7

t=10.0

t=2.0

t=1.0

t=0.01

t=0.2

t=0.3

t=0.5

t=0.7

Page 20: Transient CSR - Swanseamathematical structure, transient analytical solutions may be obtained in both velocity and extra-stress. This is true for example with constant shear viscosity

19

Steady-state Qt-control Transient Q-control Force ∆p--control

Fig. 6. Transient development of stress field, 4:1 contraction, Oldroyd-B, We=1

t=10.0

t=3.0

t=2.0

t=1.0

t=0.5

t=0.3

t=0.2

t=0.01

Page 21: Transient CSR - Swanseamathematical structure, transient analytical solutions may be obtained in both velocity and extra-stress. This is true for example with constant shear viscosity

20

Fig. 7. Transient development of pressure field, 4:1 contraction, Oldroyd-B, We=1

Steady-state Q-control Transient Qt-control

Force ∆p-control

Heavy oscillation Mild oscillation Medium oscillation Pmax = 1.1*10

6

Pmin = -8.5*105

Pmax = 1.3*103

Pmin = 0.0

Pmax = 2.3*102

Pmin = 0.0

tn=7∆t

tn=6∆t

tn=4∆t t

n=2∆t

tn=3∆t

tn=5∆t

tn=1∆t

tn=8∆t

SteadyStateSolution

tn=1∆t

tn=3∆t

tn=5∆t

tn=7∆t

SteadyStateSolution

tn=2∆t

tn=4∆tt

n=6∆tt

n=8∆t

tn=1∆t

tn=9∆t

tn=8∆t t

n=2∆tt

n=4∆tt

n=6∆t

tn=5∆t

tn=3∆t

SteadyStateSolutionSteady-state solution

Steady-state solution Steady-state solution

Page 22: Transient CSR - Swanseamathematical structure, transient analytical solutions may be obtained in both velocity and extra-stress. This is true for example with constant shear viscosity

21

a)

Time

Ux

0 1 2 3 4 5 6 7 8 9 10 11 12 130

0.2

0.4

0.6

0.8

1

1.2

Transient Qt-Control(W&K)

W&K, We_incrementation

Transient Centreline-Velocity, Inlet, We=1

We=0.5We=0.1 We=1.0

b)

Time

τ xx

0 1 2 3 4 5 6 7 8 9 10 11 12 130

0.02

0.04

0.06

0.08

0.1

Transient Qt-Control (W&K)

W&K, We-incrementation

Transient Wall-Stress, Inlet, We=1

We=0.1 We=0.5 We=1.0

Fig 8. Transient development of velocity and stress (τxx), Oldroyd-B

Page 23: Transient CSR - Swanseamathematical structure, transient analytical solutions may be obtained in both velocity and extra-stress. This is true for example with constant shear viscosity

22

Time

Txx

0 5 100

1

2

3

4

5

6

7

8 Transient Qt-control

Force ∆p-control

Evolution of velocity aroundcontraction

Time

U-velocity

0 5 100

0.1

0.2

0.3

0.4

0.5

0.6

Transient Qt-control

Force ∆p-control

Evolution of velocity aroundcontraction

Txx

3.6

3.2

2.8

2.4

2

1.6

1.2

0.8

0.4

0.02

-0.4

Transient Qt-control

Stress τxx

Txx

3.6

3.2

2.8

2.4

2

1.6

1.2

0.8

0.4

0.02

-0.4

Force ∆p -control

Stress τxx

U

1.4

1.3

1.2

1.1

1

0.9

0.8

0.7

0.6

0.5

0.4

0.3

0.2

0.1

0

Transient Qt-control

Velocity Ux

U

1.4

1.3

1.2

1.1

1

0.9

0.8

0.7

0.6

0.5

0.4

0.3

0.2

0.1

0

Force ∆p -control

Velocity Ux

Fig. 9. Transient development of velocity and stress (τxx), SXPP, We=1

Sample point

Sample point

Page 24: Transient CSR - Swanseamathematical structure, transient analytical solutions may be obtained in both velocity and extra-stress. This is true for example with constant shear viscosity

23

Captions

Fig. 1. Spatial discretisation: fe-cell with four fv-subcells and fv control volume for

node l with median-dual-cell (shaded)

Fig. 2. Fig. 2. Planar 4:1 rounded-corner domain: section of contraction-flow mesh

Fig. 3. Transient development of velocity and stress (τxx), Oldroyd-B, We=1

Fig. 4. Transient streamlines, 4:1 contraction, Oldroyd-B, We=1

Fig. 5. Transient development of velocity field, 4:1 contraction, Oldroyd-B, We=1

Fig. 6. Transient development of stress field, 4:1 contraction, Oldroyd-B, We=1

Fig. 7. Transient development of pressure field, 4:1 contraction, Oldroyd-B, We=1

Fig. 8. Transient development of velocity and stress (τxx), Oldroyd-B

Fig. 9. Transient development of velocity and stress (τxx), SXPP, We=1