58
Extensions of the Ferry shear wave model for active linear and nonlinear microrheology Sorin M. Mitran, M. Gregory Forest, Lingxing Yao, Brandon Lindley 1 , David B. Hill 2 The Virtual Lung Project, University of North Carolina, Chapel Hill, 27599-3250 June 15, 2007 1 Applied Mathematics Program, Department of Mathematics, University of North Carolina, Chapel Hill, 27599-3250 2 Cystic Fibrosis Center, University of North Carolina, Chapel Hill, 27599-7248

Extensions of the Ferry shear wave model for active linear

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Extensions of the Ferry shear wave model for

active linear and nonlinear microrheology

Sorin M. Mitran, M. Gregory Forest,

Lingxing Yao, Brandon Lindley1,

David B. Hill2

The Virtual Lung Project,

University of North Carolina,

Chapel Hill, 27599-3250

June 15, 2007

1Applied Mathematics Program, Department of Mathematics, University of NorthCarolina, Chapel Hill, 27599-3250

2Cystic Fibrosis Center, University of North Carolina, Chapel Hill, 27599-7248

Abstract

The classical oscillatory shear wave model of Ferry et al. [J. Polym. Sci..

2:593-611, (1947)] is extended for active linear and nonlinear microrheology.

In the Ferry protocol, oscillation and attenuation lengths of the shear wave

measured from strobe photographs determine storage and loss moduli at each

frequency of plate oscillation. Small, microliter, volumes typical in biology re-

quire modifications of experimental method and theory. Microbead tracking

replaces strobe photographs. Reflection from the top boundary yields counter-

propagating modes which are modeled here for linear and nonlinear viscoelastic

constitutive laws. Furthermore, bulk imposed strain is easily controlled, and we

explore the onset of normal stress generation and shear thinning using nonlinear

viscoelastic models. For this paper, we present the theory, exact linear and non-

linear solutions where possible, and simulation tools more generally. We then

illustrate errors in inverse characterization by application of the Ferry formu-

las, due to both suppression of wave reflection and nonlinearity, even if there

were no experimental error. This shear wave method presents an active and

nonlinear analog of two-point microrheology Crocker et al. [Phys. Rev. Lett.

85: 888 - 891 (2000)]. Nonlocal (spatially extended) deformations and stresses

are propagated through a small volume sample, on wavelengths long relative to

bead size. The setup is ideal for exploration of nonlinear threshold behavior.

Keywords: Viscoelastic shear wave; Upper convected Maxwell model; Small

sample effects; Biological fluids; Inverse characterization

1 Introduction

The unidirectional propagation of shear waves from an imposed oscillatory plate

is a classical technique for linear viscoelastic characterization of gels and poly-

mer solutions [1, 2, 3]. The experimental technique and associated modeling and

analysis are extensively discussed in the literature [4, 6, 7]. Linear viscoelastic

constitutive equations, coupled with momentum balance and boundary condi-

tions, yield an exact solution for a semi-infinite domain. In the original applica-

tions, strain-induced birefringence of synthetic polymers allows snapshots of the

propagating shear wave, from which oscillation and attenuation lengthscales are

related explicitly to frequency-dependent shear and loss moduli. These formulas

are recalled in Section 2.

Effects due to a finite depth in the direction of shear wave propagation

are negligible as long as the attenuation scale is much shorter than the gap

height, suppressing back propagation from the opposing surface; see [6], Chapter

5. The advantage of the Ferry design over various other devices carried over

from acoustics, which involve interpreting surface signals, lies in the ability to

strobe the entire shear wave as it progresses through the viscoelastic medium.

The method is therefore effective if the material is birefringent under strain,

and available in sufficient volumes. Linear response is ostensibly controlled

in the bulk by imposing sufficiently small amplitude displacements, yet the

possibility of large local strains near the source plate is evident even in the

classical snapshots. In our microrheology experiments with microliter samples,

strain-induced birefringent waves are replaced by microscopy and bead-tracking

techniques [9]. This method is appealing for various reasons:

1

• Similar to 2-point passive microrheology [13], bead time series are domi-

nated by non-local deformations and stresses rather than local bead-fluid

interactions. The signals are deterministic rather than stochastic.

• One-dimensional heterogeneity is easily modeled for general linear vis-

coelastic fluids, extending the classical viscous fluid analysis of Rayleigh

[6].

• It is straightforward to model one-dimensional heterogeneity with nonlin-

ear constitutive laws (cf. [7, 12], [14]), enabling an experimental-modeling

protocol for nonlinear inverse characterization at the low stress levels typ-

ical in human physiology. The analysis and/or numerical simulation of

the resulting system of quasilinear partial differential equations must be

developed for application to experimental data; these results are provided

in this paper.

• Selected models are capable of capturing and assessing two signature non-

linear phenomena, normal stress generation in shear and shear thinning,

which we implement to explore onset of nonlinear behavior versus plate-

induced strain.

An additional compelling feature of the Ferry shear wave experiment arises

from our interest in viscoelastic characterization of pulmonary liquids as part

of the Virtual Lung Project at UNC Chapel Hill. Namely, the oscillating plane

Couette cell mimics in vivo conditions of oscillatory transport of pulmonary

mucus by coordinated cilia and tidal breathing cycles. Thus, this method has

the potential to characterize linear and nonlinear viscoelastic properties under

conditions that approximate biological conditions. The shear wave experiment

controls oscillatory strain and frequency deformations at one fluid layer inter-

face; one can ramp up the bulk strain by controlling the amplitude of the lower

2

plate displacement relative to the height of the sample, while also monitor-

ing, with the modeling tools presented here, the local strain per wavelength

of the shear-induced deformation. Rather than rely upon refraction due to

strain-induced birefringence, Hill and Superfine [9] mix micron-scale, neutrally

buoyant beads and use particle-tracking to measure shear wave propagation fea-

tures. The model and experiment can also couple asymmetry in the imposed

plate cycle, which mimics asymmetry of the beat cycle of coordinated cilia. This

feature is easily incorporated into the framework presented here by modifying

the boundary conditions on the nonlinear solver algorithm, but that extension

is beyond the scope of the present paper. The numerical solver has also been

benchmarked in the small strain limit with exact formulas presented below, so

that this software along with inverse procedures can be used in general appli-

cations. An article focusing on the experimental protocols and applications to

benchmark viscoelastic fluids and biological liquids is in preparation [8].

This method for viscoelastic constitutive characterization holds special promise

for applications in other biological systems as well, where samples are often only

available in small volumes. The modeling and assessment of counterpropagat-

ing waves and nonlinearity versus imposed strain, especially in regard to inverse

characterization, are the primary contributions of this paper. We show in the

present paper that accuracy of the classical Ferry formulas varies significantly

depending on viscoelastic moduli versus frequency, imposed displacement am-

plitude, and sample height. For example, a conservative estimate of wave speeds

for 30μm thick samples of a viscoelastic material, based on reported ranges of

bulk moduli in pulmonary liquids, indicates multiple wave reflections within

each period of plate oscillation in 10 − 20Hz frequency range. From simulated

finite depth data for prescribed viscoelastic moduli, thereby removing any ex-

perimental error, we show that fitting to formulas based on unidirectional waves

3

can lead to enormous errors in inverse recovery of the moduli.

The key issue motivating this paper is whether the assumptions of the Ferry

model, linearity and a semi-infinite domain, are approximately satisfied under

experimental conditions with limited volumes. More precisely, we assess accu-

racy of the linear, semi-infinite assumption across a multiple parameter space:

storage and loss moduli versus frequency, gap height, and plate displacement

amplitude. The infinite depth assumption is surely not valid, so the first target

is to assess the order of errors in storage and loss moduli across some frequency

range, due to this effect alone. Next, we aim to characterize the threshold

strain (imposed by the lower plate displacement relative to sample height) at

which nonlinear effects become significant, and to assess measurable flow and

stress features triggered by nonlinearity. It is hypothesized that normal stresses

in shear, a standard signature of nonlinearity, play an important role in pul-

monary liquid transport by coordinated cilia and coughing events. A possible

scenario is that normal and shear stresses propagate from the mucus and peri-

ciliary liquid layer to the epithelium, causing cellular compressive and shear

strain, serving to regulate biochemical feedback. The role of shear thinning in

pulmonary transport is not known, so a method is desirable that can simply

ascertain whether a given sample shear thins under controlled strain conditions.

The remainder of this paper is organized as follows. Section 2 gives a sum-

mary of models, from linear–viscous to nonlinear–viscoelastic. Section 3 con-

tains analytical results, including exact solutions and formulas for linear and

nonlinear, semi-infinite and finite depth, models, as well as analysis of well-

posed boundary conditions for the nonlinear models. Section 4 contains numer-

ical studies of the two nonlinear models, their comparison to linear models and

one other, and a discussion of physical predictions.

4

2 Shear wave models

2.1 Physical parameters and non-dimensionalization

We consider a layer of incompressible fluid of height H , set into motion by

an oscillating lower flat plate, and whose upper surface remains flat and par-

allel to the bottom plate. The˜ superscript will be used to denote quantities

expressed in physical units. The bottom plate is set into harmonic motion

V (t) = V0 sin(ωt), as illustrated in Figure 1.

Figure 1: Schematic of oscillatory Couette flow and notation used to describethe deviatoric stress on an infinitesimal fluid element. The x-axis is parallel tothe plate motion, the y-axis is the cross-channel direction.

A reference viscosity η0 value is chosen for the fluid of interest, namely the

zero strain-rate viscosity of the fluid. The reference length A = V0/ω is the

5

maximum displacement of the driven plate; the reference time ω−1 is set by the

plate frequency; and the reference stress η0ω is the viscous plate shear stress.

With this choice, the non-dimensional velocity of the bottom plate is given by

vx(0, t) = V (t) = sin(t), where vx(y, t) denotes the velocity component of the

fluid in the direction of plate motion at height y. It is a property of the model

equations that with this scaling, we can focus on the 2-parameter space of

values of storage and loss moduli, and make general statements about errors in

recovering the moduli that depend only on the pair of values, independent of the

imposed frequency. We return later to the practical implications of this scaling

property. Typically, a distinguished material relaxation time λ is identified,

the ratio of a viscosity to an elastic modulus. The following dimensionless

parameters arise in the model equations:

• Reynolds number Re = ρωA2/η0

• Deborah number λ = λω

• Bulk shear strain γ = A/H.

By symmetry, the one-dimensional propagation of velocity and stress leads

to the following reductions in flow and stress variables, all functions only of t

and the height y between the two flat surfaces. The fluid pressure p and velocity

components are

p = p(y, t), vx = vx(y, t), vy = 0, vz = 0. (1)

From vx(y, t), we may construct transient particle displacements X(y, t) at each

height (y) in the layer,

X(y, t) =∫ t

0

vx(y, t′)dt′ +X(y, 0), (2)

6

for given initial conditions X(y, 0).

The fluid motion is determined from the non-dimensional momentum equa-

tion,∂v

∂t+ (v ·�)v =

1Re

�·(−pI + τ ), (3)

together with a constitutive law for the extra stress, τ , which likewise is assumed

to depend only on t and y,

τ = τ (y, t), with τxz = τyz = 0. (4)

To close the system for (p, vx, τ )(y, t), we require the constitutive law relating

τ and �v. A hierarchy of laws has been discussed in standard texts [6], [7, 12],

the elements of which we now recall.

2.2 Newtonian Viscous Fluid

For viscous fluids, the Navier-Stokes constitutive law is

τ = 2D, (5)

where D ≡ 12 (�v + �vT ), is the rate-of-strain tensor (the viscosity coefficient

does not appear due to the scaling, 2.1). From (1), it follows that

⎧⎪⎪⎨⎪⎪⎩τxy =

∂vx

∂y,

τxx = τyy = τxz = τyz = τzz ≡ 0.(6)

The momentum equation (3) reduces to

⎧⎪⎪⎨⎪⎪⎩∂vx

∂t=

1Re

∂2vx

∂y2,

∂p

∂y= 0 .

(7)

7

2.3 Linear Viscoelastic Fluids

The constitutive law of linear viscoelasticity (the Lodge model) is

τ = 2∫ t

−∞G(t− t′)D(y, t′)dt′, (8)

where G(t) is the non-dimensional shear relaxation modulus function. In phys-

ical units we have G = η0ωG. Combining (1) and (8), only shear stresses are

generated: ⎧⎪⎪⎨⎪⎪⎩τxy =

∫ t

−∞G(t− t′)

∂vx

∂y(y, t′)dt′,

τxx = τyy = τxz = τyz = τzz = 0.

(9)

The momentum equation (3) closes the system:

∂vx

∂t=

1Re

∂τxy

∂y, (10a)

∂p

∂y= 0. (10b)

From (9) and (10a), a closed equation for vx follows:

∂vx

∂t=

1Re

∫ t

−∞G(t− t′)

∂2vx

∂y2(y, t′)dt′, (11)

and then the shear stress is calculated from (9).

We note for future reference that if the shear modulus G(t) is a single-mode

Maxwell model, G(t) = (ηp/λ) e−t/λ (non-dimensional form), then the equation

(11) simplifies to a damped wave equation for vx (cf. [6, 7]),

∂2vx

∂t2− ηp

λ Re

∂2vx

∂y2+

∂vx

∂t= 0. (12)

Furthermore, it is straightforward to show that the shear stress τxy satisfies the

8

identical damped wave equation (12).

2.4 Upper Convected Maxwell and Giesekus Models of

Nonlinear Viscoelasticity

The standard procedure for lifting linear viscoelasticity to nonlinear regimes is

treated in classical texts [7, 12]. Either one pursues the K-BKZ class of integral

constitutive laws, generalizing (8), or one pursues differential laws, the choice

made here. This approach begins with a decomposition of the extra stress into

a finite sum,

τ =k∑

i=1

τ i, (13)

where each τ i obeys a Maxwell-type evolution equation, with nonlinearity through

the upper convected derivative as well as other nonlinear terms that have been

developed to capture specific nonlinear properties. We will explore a Giesekus

model below which extends the Upper Convected Maxwell model to include

shear thinning capability. For the remainder of this paper, we restrict to a

single Maxwell-type mode; the extension to any finite number of modes is nu-

merically straightforward, where the numerical cost scales proportional to the

number of modes.

The Maxwell-type evolution equation for the extra stress is

λ∇τ = 2ηpD− τ − a τ τ , (14)

where the upper convected derivative∇τ is defined by

∇τ =

∂τ

∂t+ (v · ∇) τ − (∇v)T τ − τ (∇v). (15)

In this type of model the additional polymeric stress τ is transported by the

9

velocity field v, decays exponentially at rate λ, is produced by the solvent fluid

rate of strain D (with scaling 2ηp/λ, ηp = ηp/η0) and is nonlinearly modified

through the Giesekus term (a/λ) τ τ . The Giesekus “mobility parameter” a

distinguishes various viscoelastic models: a = 0 reduces to the upper convected

Maxwell (UCM) model, while a = 0.5 is the Leonov model. In general, it is

assumed that 0 ≤ a ≤ 1, and the mobility parameter is to be fitted from data

along with λ.

For the oscillatory-plate problem the polymeric stress transport equations

(14) become

⎧⎪⎪⎪⎪⎪⎪⎪⎪⎨⎪⎪⎪⎪⎪⎪⎪⎪⎩

∂τxx

∂t− 2τxy

∂vx

∂y= − 1

λτxx − a

λ(τ2

xx + τ2xy)

∂τxy

∂t− τyy

∂vx

∂y=ηp

λ

∂vx

∂y− 1λτxy − a

λτxy (τxx + τyy)

∂τyy

∂t= − 1

λτyy − a

λ(τ2

xy + τ2yy)

τyz = τxz = τzz = 0 .

(16)

The momentum equation reduces to

∂vx

∂t=

1Re

∂τxy

∂y, (17a)

∂p

∂y=∂τyy

∂y. (17b)

The upper convected derivative,∇τ , and the shear thinning term τ τ , render

the stress transport equations nonlinear. As in the Newtonian fluid case, there

is no convective nonlinearity in the momentum equation itself. An important in-

dicator of nonlinearity is the shear-induced generation of a non-zero first normal

stress difference, N1 = τxx − τyy, an easily measured feature. The presence of a

normal stress gradient along the y-direction then generates a pressure gradient,

absent in the linear model. Indeed, neglecting all nonlinear terms from (16)

10

the two normal stress components τxx, τyy, obey identical ordinary differential

equations∂τxx

∂t+

1λτxx = 0,

∂τyy

∂t+

1λτyy = 0 . (18)

Thus, there is no possibility of generation of normal stresses in shear flow under

linear assumptions. Additionally, any initial normal stresses distributed across

the layer relax exponentially fast, and do not propagate nor couple to the shear

stress and shear wave. Finally, the Giesekus model captures shear thinning. All

nonlinear features are addressed in the context of the numerical results.

The coupled system (16) and (17a) can be written as an evolution equation,

∂q

∂t+A(q)

∂q

∂y= ψ(q), (19)

with

q =[vx τxx τxy τyy

]T,

A(q) =−

⎡⎢⎢⎢⎢⎢⎢⎢⎣

0 0 1/Re 0

2τxy 0 0 0

ηp/λ+τyy 0 0 0

0 0 0 0

⎤⎥⎥⎥⎥⎥⎥⎥⎦, ψ = − 1

λ

⎡⎢⎢⎢⎢⎢⎢⎢⎣

0

τxx + a (τ2xx + τ2

xy)

τxy+a τxy (τxx +τyy)

τyy + a (τ2xy + τ2

yy)

⎤⎥⎥⎥⎥⎥⎥⎥⎦,

together with the pressure equation (17b) and τxz = τyz = 0.

2.5 Boundary conditions

Boundary conditions are required at the bottom plate y = 0, which is experi-

mentally controlled, and at the top boundary y = H , which may or may not

be explicitly controlled. The system of partial differential equations (19) is hy-

perbolic if the normal stress τyy obeys an inequality (see Sect. 3.3.1). If the

11

inequality is violated, the model is no longer well-posed for the study of this

experiment, a common occurrence for Maxwell-type models (cf. [14]). Hence,

the characteristics of the system must be investigated to determine allowable

boundary conditions. The following physically motivated conditions are ana-

lyzed in Sec. 3.3.2.

2.5.1 Lower plate boundary condition

At y = 0, no-slip boundary conditions apply, so that

vx(y = 0, t) = V (t), (20)

with V (t) the imposed velocity of the lower plate. From (2) and for harmonic

plate motion V (t) = sin(t), the plate displacement is given by

X(0, t) =∫ t

0

V (s)ds +X(0, 0) = (1 − cos t) +X0, (21)

where X0 labels the horizontal plate coordinate at t = 0.

2.5.2 Top boundary condition

We distinguish two types of top (flat) interface conditions:

1. The top of the layer is in contact with a flat plate, which could be held sta-

tionary or allowed to move (in which case the velocity at the top boundary

is imposed)

vx(y = H, t) = VH(t). (22)

The stationary top plate condition is the most readily realized experimen-

tal condition when seeking inverse characterization.

2. The top of the layer is in contact with another fluid, in which case conti-

12

nuity of shear stress yields

τxy(y = H, t) = τxy(H, t) of the ambient fluid. (23)

The simplest case is to assume a passive ambient so that the shear stress

is identically zero.

We address the issue of normal stress boundary conditions in the character-

istic analysis below.

3 Analytical results

For viscous and linear viscoelastic fluids, with semi-infinite and finite domains,

the model equations above are exactly solvable by Fourier methods if one ignores

transients, and further solvable with transients by Laplace methods. The tradi-

tional approach (cf. Ferry [6]) is to suppress transients, leading to closed form

solutions whose experimentally measurable features (e.g. attenuation length

and shear wavelength) are explicitly related to material properties. Thus in-

verse characterization is given by simple formulas. We review these results now

with some extensions that appear to be absent in the literature. The numerical

solutions, naturally, can easily be used to model transients. We further show

that the nonlinear model for an upper convected Maxwell law is also exactly

solvable, and therefore explicit representations for normal stress generation in

shear are determined. The Giesekus model does not submit to a closed-form

solution.

3.1 Viscous fluid solution

The long-time solution for the viscous model (7) is widely available (e.g. [5])

and recalled for completeness as well as comparison with generalizations.

13

Semi-infinite domain solution. Upon positing a solution of the form

vx(y, t) = Im{vx(y)eit} (recall time is measured in units of ω−1, so the physical

frequency is ω/2π), the solution is given by

vx(y, t) = Im{e−δ0y eit}

= e−α0y sin(t− β0y)

which obeys the far-field boundary condition limy→∞ vx(y, t) = 0. We have in-

troduced δ0 = α0 + iβ0 =√i Re = (1 + i)

√Re/2. The Reynolds number,

Re = ρωA2/η0, determines both the penetration (1/α0) and oscillation length-

scales (2π/β0), through α0 = β0 =√Re/2. The velocity field produces a

particle displacement

X(y, t) = X(y, 0) + e−α0y [cos(βt) − cos(t− βy)]

The viscous shear stress is given by τxy = ∂vx/∂y,

τxy(y, t) = Im{−δ0 e−δ0yeit}

= sin(t− β0y + 5π/4)e−α0y√Re ,

showing a constant phase difference φ = 5π/4 between the shear stress and the

velocity.

Finite-depth channel with stationary top plate. The no-slip boundary

condition at the top of the channel reads vx(y = H, t) = 0. The solution is simple

vx(y, t) = Im{eit sinh [δ0 (H − y)]

sinh [δ0 H ]

}, (24)

but does not appear to have been written down in the literature.

14

3.2 Linear viscoelastic model solutions

We summarize the Fourier, frequency-locked solutions to (10a - 10b) with the

bottom plate forcing V (t) = sin(t), ignoring transients.

Semi-infinite domain solution [6]. If the loss modulus is non-zero, then

vx(y → ∞, t) = 0. (25)

A separable Fourier solution is posited, vx(y, t) = Im{vx(y)eit}. Then from (9)

and (10a) a simple ordinary differential equation for vx(y) is obtained,

d2vx

dy2− i Re

η∗vx = 0. (26)

where the non-dimensional complex viscosity η∗ =∫∞0 G(s) exp(−is)ds = η′ −

iη′′, has been introduced. For a Maxwell model with G(s) = (ηp/λ)e−s/λ we

obtain

η′ =ηp

1 + λ2, η′′ =

ηpλ

1 + λ2. (27)

The decay condition (25) selects one independent solution of (26), and the

velocity has the same form as the viscous solution:

vx(y, t) = Im{eite−δy}

= e−αy sin(t− βy) , (28)

where δ = α+ iβ =√i Re/η∗, (α, β ∈ R+),

α =

√Re/2|η∗|

√|η∗| − η′′, β =

√Re/2|η∗|

√|η∗| + η′′ . (29)

15

For the single mode Maxwell fluid

α2 =Re

2

√1 + λ2 − λ

ηp=

12c2

(√1λ2

+ 1 − 1

), (30)

β2 =Re

2

√1 + λ2 + λ

ηp=

12c2

(√1λ2

+ 1 + 1

), (31)

where the zero-stress shear wave speed c =√ηp/(λ Re) has been introduced.

Clearly, 0 < α ≤ β, with equality only in the viscous limit η′′ = 0, η∗ = 1

(λ = 0, ηp = 1) such that α = β =√Re/2. The greater the elastic component

η′′, the larger the difference between the oscillation wavelength 2π/β and the

attenuation length 1/α.

The transform of the complex modulus, G(ω) = G′ + iG′′, in our choice of

non-dimensional units, is related to the complex viscosity η(ω) by G = iη∗; in

dimensional units, G = iωη∗. The non-dimensional formulas relating storage

and loss moduli and the reciprocal lengthscales α and β are:

G′ =

(β2 − α2

)Re

(α2 + β2)2, G′′ =

2αβ Re

(α2 + β2)2. (32)

Re-expressing the above formulas in physical quantities recovers the formulas of

Ferry, Sawyer, and Ashworth [6]

G′ = ρω2

(β2 − α2

)(α2 + β2

)2 , G′′ = ρω2 2αβ(α2 + β2

)2 ,

used to infer G′(ω), G′′(ω) from experimental measurements of α = αA, β = βA.

The particle positions X(y, t) at height y are then given by

X(y, t) = X(y, 0) + e−αy [cos(βy) − cos(t− βy)] . (33)

16

The shear stress can be computed using the Lodge stress formula (8)

τxy(y, t) = Im{∫ t

−∞G(t− t′)

dvx

dy(y)eit′ dt′

}= Im

{η∗dvx

dy(y)eit

}.

For the velocity (28) we obtain,

τxy(y, t) = Im{−δη∗eite−δy}

= sin(t− βy + φ+ π)e−αy√|η∗|Re , (34)

and, of course, the normal stresses are identically zero for linear viscoelasticity,

τxx = τyy = τzz = 0. The phase shift of the tangential stress with respect

to the velocity is φ + π, with φ = arctan(β/α), 0 ≤ φ ≤ π/2. Note that

the stress required to achieve the claimed boundary condition on plate speed is

τxy(y = 0, t) = − sin(t+ φ)√|η∗|Re.

A comparison to the viscous solution is informative. The ratio of the viscous

to viscoelastic penetration lengths (α/α0) is shown in Fig. 2. When η′′ = 0

we obtain the expected increase in penetration depth as η′ increases – a more

viscous fluid leads to more momentum transfer from the oscillating plate. As

the fluid elasticity η′′ increases, the penetration depth ratio decreases. Some of

the momentum imparted to the fluid by the oscillating plate is now radiated

as an elastic shear wave, hence the penetration depth is smaller. The ratio of

viscid to viscoelastic wavelengths (β/β0) is also shown in Fig. 2. For viscoelastic

fluids with η′′ > 0 the wavelength reflects the competition between diffusive and

radiative momentum transfer. At a given η′ an initial increase in the elasticity

decreases the wavelength. In this range the motion is overdamped and the

decrease in penetration depth dominates the wavelength. At greater values of

η′′ the radiative momentum transfer (through shear waves) becomes dominant

and there is an increase in the wavelength. The maximum viscoelastic shear

stress is greater than the purely viscous stress by a factor of√|η∗| and exhibits

17

a different phase shift, φ+ π instead of 5π/4.

2 4 6 8 100.0

0.2

0.4

0.6

0.8

1.0

Η'

Α�Α0

Penetration Length Ratio

2 4 6 8 100.00.20.40.60.81.01.2

Η'

Β�Β0

Wavelength Ratio

�� � 2

�� � 2�� � 0

�� � 0

�� � 10

�� � 10

Figure 2: Comparison of viscoelastic penetration (1/α) and wave lengths (2π/β)to the viscous values (1/α0, 2π/β0)

Finite channel depth. When the fluid is bounded above at y = H by a

flat interface moving at velocity vH(t), a no-slip boundary condition yields

vx(y = H, t) = vH(t) .

The Fourier method used here ignores transients and can only be used for top

plate velocities of the form vH(t) = VHeit with VH some complex constant

giving the amplitude of the top plate motion as well as its relative phase with

respect to the bottom plate forcing. The solution in this case is given by

vx(y, t) = Im{eit sinh [δ(H − y)]

sinh (δH)+ VHe

it sinh (δy)sinh (δH)

}. (35)

The most easily controlled experimental condition is to place a stationary solid

lid on the top interface in which case VH = 0,

vx(y, t) = Im{eit sinh [δ(H − y)]

sinh (δH)

}. (36)

18

The half-plane solution (28) is recovered in the limit H → ∞

limH→∞

sinh [δ(H − y)]sinh (δH)

= e−δy, limH→∞

sinh (δy)sinh (δH)

= 0, (37)

recalling that δ = α + iβ and α > 0. (This “textbook” solution does not ap-

pear to be in the literature; it is given here for explicit comparison with the

semi-infinite solution, and since it is the basis for finite-depth inverse character-

ization.) The stress is given by

τxy(y, t) = Im{−δ η∗ cosh[δ (H − y)]

sinh(δ H)eit

}

Finite channel depth, flat fluid-fluid boundary above. The boundary

condition in this case is continuity of tangential stress within the viscoelastic

fluid τyx(H, t) and the adjacent fluid τH(t). Again, because we are neglecting

transients, the adjacent fluid must exhibit a tangential stress dependence of the

form τH(t) = THeit to be consistent with the Fourier procedure. This leads to

dv0x

dy(y = H) =

TH

η∗, (38)

and the solution

vx(y, t) = Im{eit

[cosh [δ(H − y)]

cosh (δH)+TH

η∗δsinh(δy)cosh(δH)

]}. (39)

This solution is relevant if one solves the model equations for the fluid layer

above, or designs an experiment with a controlled dynamic shear stress at the

interface. (This solution likewise appears not to have been previously pub-

lished.)

It should be noted however that ensuring that the top fluid layer has the

simple harmonic tangential stress τH(t) = THeit is difficult to realize experi-

19

mentally. If the adjacent fluid exerts negligible stress upon the viscoelastic fluid

layer TH∼= 0, then the formula

vx(y, t) = Im{eit cosh [δ(H − y)]

cosh (δH)

},

is valid. Otherwise, it is to be expected that τH(t) contains other frequencies

besides the fundamental frequency used in the bottom plate forcing eit. Should

one attempt, irrespective of this observation, to use formula (39), then the

complex viscosity deduced from the experimental measurements will be affected

by aliasing of the additional frequencies present in τH(t) onto the fundamental

mode eit.

Linear model limitations. The constitutive law (8) is valid for small

strains

max |Dxy| = max∣∣∣∣∂u∂y∣∣∣∣ = max

∣∣∣∣ ∂∂y∫ t

0

vx(y, s)ds∣∣∣∣ 1 .

For the half-plane solution (28) the strain is given by

Dxy(y, t) = e−αy {α [cos(t− βy) − cos(βy)] + β [sin(t− βy) + sin(βy)]}

assuming X(y, 0) = 0. Physically we expect the maximum strain to be obtained

at y = 0 where the displacement is

X(y = 0, t) = −α (1 − cos t) − β sin t .

The extremum ofX(0, t) is obtained when cos t = α/√α2 + β2, sin t = −β/

√α2 + β2,

hence

max |Dxy| =√α2 + β2 − α .

20

Since√α2 + β2 =

√Re/|η∗|, the small strain condition becomes

√Re

|η∗|

(1 −√

|η∗| − η′′

2 |η∗|

) 1 .

The elastic limit of fluid behavior corresponds to η′′ � η′ for which the small

strain condition becomes

Re η′′ .

It is useful to revert to physical units to highlight the significance of these limits.

Recall that Re = ρωA2/η0 and G′ = ωη′′. The small strain condition in the

elastic limit then states that

ρ(ωA)2 G′ ,

or, physically, that the dynamic pressure induced by the oscillating plate must

be much less than the elastic modulus of the fluid. In the viscous limit η′ =

1, η′′ = 0, the small-strain condition becomes√Re 1 . In physical terms this

states that

A d =√ω/2ν ,

or that the amplitude A of the oscillating bottom plate must be much less than

the viscous penetration depth d. If viscous and elastic effects are roughly equal,

η′ ∼= η′′, then the small strain condition becomes

Re << min(η′, η′′) .

The coefficients (η′, η′′) obtained from fitting experimental data to the formulas

above should be checked against these limits to ascertain whether non-linear

effects are insignificant. In the finite-depth linear and nonlinear models, one

21

can calculate the local strain from the numerical solutions, which generically

will be maximal in the first half-wavelength of the shear deformation at the

lower plate.

3.3 Nonlinear model solution in special limits

3.3.1 Hyperbolic system structure

We now turn to a consideration of nonlinear effects by solving (19). If the

system obeys the linear limit, that will become evident from the solution. Our

focus is conditions where nonlinear effects are measurable. In general numerical

methods are required for a solution. We present the analysis needed for such

numerical methods. For the upper convected Maxwell model, quite conclusive

analytical results are attainable.

The matrix A from (19) has eigenvalues μ1 = μ2 = 0, μ3 = − c, μ4 = c with

c =√ηp/(λRe) + τyy/Re , (40)

denoting the velocity of propagation of shear waves transverse to the flow direc-

tion. If the fluid is elastic to some extent, then λ > 0 and finite. Furthermore

if τyy ≥ 0 then the shear wave velocity c is a real number. It is interesting

to note that for a given relaxation time λ there exists a critical normal stress

τyy = −ηp/λ at which c = 0, and for τyy < −ηp/λ, the system is no longer

hyperbolic, but elliptic. The associated right eigenvectors are

R =[r1 r2 r3 r4

]=

⎡⎢⎢⎢⎢⎢⎢⎢⎣

0 0 −c c

0 1 2τxy 2τxy

0 0 c2Re c2Re

1 0 0 0

⎤⎥⎥⎥⎥⎥⎥⎥⎦. (41)

22

The eigenvectors form a basis for real c = 0, hence the system (19) is hy-

perbolic. The structure of this system is particularly simple with two non-

propagating components, whose physical significance will be addressed below.

The multiplicity-two zero characteristic speed allows considerable analytical

progress.

Consider a local linearization in which we use average values (denoted by an

overbar) in the eigenvector expressions

R =

⎡⎢⎢⎢⎢⎢⎢⎢⎣

0 0 −c c

0 1 2τxy 2τxy

0 0 c2Re c2Re

1 0 0 0

⎤⎥⎥⎥⎥⎥⎥⎥⎦, R

−1 =1

2c2

⎡⎢⎢⎢⎢⎢⎢⎢⎣

0 0 0 2c2

0 2c2 −4τxy/Re 0

−c 0 1/Re 0

c 0 1/Re 0

⎤⎥⎥⎥⎥⎥⎥⎥⎦.

The characteristic variables of the problem are

w =

⎡⎢⎢⎢⎢⎢⎢⎢⎣

w1

w2

w3

w4

⎤⎥⎥⎥⎥⎥⎥⎥⎦= R

−1q =

⎡⎢⎢⎢⎢⎢⎢⎢⎣

τyy

τxx − 2τxy/(c2Re)τxy

−vx/2c+ τxy/(2c2Re)

vx/2c+ τxy/(2c2Re)

⎤⎥⎥⎥⎥⎥⎥⎥⎦,

and (19) can be rewritten in terms of these variables as

∂w

∂t+ Λ

∂w

∂y= φ (42)

with Λ = diag {μ1, μ2, μ3, μ4} and φ = R−1ψ,

φ = − 1λ

⎡⎢⎢⎢⎢⎢⎢⎢⎣

τyy + a(τ2xy + τ2

yy

)τxx + a

(τ2xy + τ2

xx

)− 4τxy σ

σ

σ

⎤⎥⎥⎥⎥⎥⎥⎥⎦, σ =

τxy

2c2Re(1 + a (τxx + τyy)) .

23

The forcing term φ can be rewritten as a function of the characteristic variables

φ = φ(w) by using the transformation,

q = Rw =

⎡⎢⎢⎢⎢⎢⎢⎢⎣

c(w4 − w3)

w2 + 2τxy(w3 + w4)

c2(w3 + w4)Re

w1

⎤⎥⎥⎥⎥⎥⎥⎥⎦. (43)

3.3.2 Characteristic boundary conditions

The hyperbolic structure of the problem indicates the compatible physical bound-

ary conditions that must be used in numerical simulations. The non-propagating

quantities w1 = τyy and w2 = τxx − 2τxy

ρc2 τxy can be freely specified initially on

the bottom and top boundaries. Subsequent values must be determined by

solving the ordinary differential equations at each boundary

∂w1

∂t= − 1

λ

[τyy + a

(τ2xy + τ2

yy

)], (44)

∂w2

∂t= − 1

λ

[τxx + a

(τ2xy + τ2

xx

)− 4τxy σ].

At the bottom plate (y = 0) only the w4 = − 12cvx + 1

2ρc2 τxy quantity can be

specified as a boundary condition since it is propagating into the fluid from the

adjacent domain. The quantity w3 = 12cvx + 1

2ρc2 τxy propagates from within

the fluid domain towards the boundary and must be determined as part of the

solution procedure. Conversely, at the top boundary (y = H) w3 can be given as

a boundary condition and w4 results from the solution procedure. Typically, we

wish to give boundary conditions in terms of the physical variables. The essence

of the above discussion is that we cannot impose both vx and τxy at boundaries.

To do so would determine both w3 and w4 in violation of characteristic infor-

mation propagation directions. We list common cases of compatible boundary

24

conditions which inform experimental protocols for this class of models to be

applicable.

1. Velocity and normal stresses. At the bottom plate we specify vx and

propagation from within the channel as specified by the characteristic

equation∂w3

∂t− c

∂w3

∂y= −σ/λ, (45)

gives w3. This allows τxy = 2 Re c2(w3 + 1

2cvx

)to be computed. We

solve (44) to obtain w1, w2. Using (43) we then compute τxx, τyy thereby

obtaining all the variables within the q vector. At the top plate we proceed

analogously, by specifying vx but obtaining τxy from the w4 component

as τxy = 2 Re c2(w4 − 1

2cvx

). The w4 component is determined from

propagation within the fluid by solving

∂w4

∂t+ c

∂w4

∂y= −σ/λ . (46)

2. Tangential and normal stresses. At both boundaries we specify τxy. Solv-

ing (44) gives w1, w2 and hence τyy, τxx. At the bottom boundary vx is

obtained from the propagating w3 component by solving (45) and at the

top from the w4 component obtained by solving (46).

3. Non-reflective top boundary condition. To simulate an infinite domain

we can set the incoming characteristic at y = H � 1 to zero, w3 = 0.

Numerically, care must be taken so there are no spurious reflections of the

w4 component at the y = H boundary.

Figure 3 gives a graphical overview of the procedure used to set boundary

conditions.

25

Figure 3: Geometry of characteristics of system (19) near the boundaries, high-lighting characteristics coming from inside the computational domain (w3 aty = 0, and w4 at y = H) and non-propagating characteristic variables w1, w2.

3.3.3 Upper convected Maxwell (UCM) solution

After the general analysis presented above let us consider an exactly solvable

case. For the UCM model we have a = 0 and the system (19) becomes

⎧⎪⎪⎪⎪⎪⎪⎪⎪⎪⎨⎪⎪⎪⎪⎪⎪⎪⎪⎪⎩

∂vx

∂t− 1Re

∂τxy

∂y= 0

∂τxx

∂t− 2τxy

∂vx

∂y= − 1

λτxx

∂τxy

∂t−(ηp

λ+ τyy

) ∂vx

∂y= − 1

λτxy

∂τyy

∂t= − 1

λτyy

(47)

with an immediate solution for τyy(y, t) = τyy(y, 0) exp(−t/λ).

Assume that τyy(y, 0) = 0, either because this is the true initial state of the

system or that enough time has passed since an initial excitation to ensure the

decay term exp(−t/λ) has become negligibly small. At later times, τyy(y, t) = 0.

26

Consequently the shear wave propagation velocity is constant c =√ηp/(λRe)

and the above system (47) decouples into a 2 × 2 linear system for vx, τxy,

⎧⎪⎨⎪⎩∂vx

∂t− 1Re

∂τxy

∂y= 0

∂τxy

∂t− ηp

λ

∂vx

∂y= − 1

λτxy

, (48)

and a linear equation for τxx,

∂τxx

∂t+

1λτxx = −2τxy

∂vx

∂y. (49)

The system (48) describes the propagation of shear waves and is identical to the

telegrapher’s equation (12) derived earlier in the linear Maxwell limit. The shear

wave propagates and decays as in the linear model. The convective nonlinearity

of the UCM system appears as a source driving the τxx stress which in the linear

model is zero. For known vx(y, t), τxy(y, t) the solution is

τxx(y, t) = −2∫ t

0

e−(t−t′)/λτxy(y, t′)∂vx

∂y(y, t′) dt′ . (50)

Longitudinal momentum is lost when τxy and ∂vx/∂y are in phase, and gained

when they are of opposite phase.

The telegrapher’s equation (12) for vx (and also for τxy) can be solved by

27

the Riemann method to yield

vx(y, t) = 12e

−t/(2λ)u(y, t), (51)

u(y, t) =12

[U(y − ct) + U(y + ct)] +∫ y+ct

y−ct

[K(s, y, t)U(s) + L(s, y, t)V (s)] ds , (52)

K(s, y, t) = − t

8λ2c

J1 (r(s, y, t))r(s, y, t)

, L(s, y, t) =12cJ0 (r(s, y, t)) , (53)

r(s, y, t) =1

2λc

√(y + ct− s)(s− y + ct) . (54)

Note that the kernel K(s, y, t) in the integrand is nonsingular since

lims→y±ct

K(s, y, t) = − t

16λ2c. (55)

In the above, the initial value functions U(y) = u(y, 0), V (y) = ut(y, 0),

ut = ∂u/∂t, have been introduced. For y > 0 we have U(y) = 0, V (y) = 0.

The original problem has boundary conditions given by vx(0, t) = sin t and

limy→∞ vx(y, t) = 0. Through the same Riemann representation the boundary

conditions can be transformed into initial conditions for y < 0 by assuming

V (y) = 0 and solving the Fredholm integral equation of the second kind

U(y) = μ

∫ 0

y

M(s, y)U(s) ds+ F (y) , (56)

M(s, y) = − y√y2 − s2

J1

(1

2λc

√y2 − s2

), (57)

F (y) = −4e−y/(2λc) siny

c, μ =

12λc

. (58)

for U(y). The solution to this integral equation for c = 1, λ = 4 is shown in

Fig. 4. The solution exhibits exponential growth as y → −∞, which is to be

expected since the initial condition at some y < 0 is attenuated by viscosity

28

until it reaches the y = 0 boundary condition line. High accuracy is essential

therefore in numerical solution of (56), but this can be achieved by using a

Nystrom method and Gauss-Lobatto quadrature (for details see [11]).

The comparison in Fig. 6 of the telegrapher’s equation solution (51) to the

Ferry half-plane solution (28) (for c = 1, λ = 4), shows the expected behavior

with an initially large start-up error followed by asymptotic convergence of the

telegrapher’s equation solution to the Ferry solution (28). The error between

the two models can be characterized using (30). The shear wave speed for the

UCM model is c while that for the Ferry solution is 1/β. Hence the relative

error in shear wave speed is

εc = 1 −√

2λ√λ+

√λ2 + 1

. (59)

Note that the error is large for small λ, i.e. the shorter the elastic relaxation

time (or, equivalently, the more rigid the medium), the greater the error. The

spatial attenuation length is 1/α for the Ferry model and 2λc for the UCM

model. The relative error in attenuation length is therefore

εα =1√

2λ√√

1 + λ2 − λ− 1 , (60)

and again is larger for small λ. The very large error between the initial transient

which is correctly captured by the telegrapher’s equation solution and that of

the asymptotic in time Ferry solution is shown in Fig. 7.

As mentioned, the tangential stress τxy satisfies the same telegrapher’s equa-

tion. However the boundary values at y = 0 are not known initially. However,

we have transformed the half-plane boundary value problem into an initial value

problem on the entire y-axis an approach which now pays dividends since we

29

�8 �6 �4 �2 0

�10

�5

0

5

10

y

U�y�

�2.0 �1.5 �1.0 �0.5 0.0�6�4�2

0246

y

U�y�

�60 �50 �40 �30 �20 �10 0

�5000

0

5000

10 000

y

U�y�

Figure 4: Solution U(y) to (56) for λ = 4, c = 1. The solution gives initialvalues U(y) = u(y, 0) obtained by transformation of boundary condition aty = 0, u(0, t) = 2et/(2λ) sin t. A Nystrom method based upon Gauss-Lobattoqudrature with n = 251 nodes was used to discretize (56). Exponential growthis observed for y → −∞. Insets show sharp resolution near the y = 0 boundaryobtained by numerical method.

30

�8 �6 �4 �2 0

�10

�5

0

5

10

y

U�y�

�2.0 �1.5 �1.0 �0.5 0.0�6�4�2

0246

y

U�y�

�60 �50 �40 �30 �20 �10 0

�5000

0

5000

10 000

y

U�y�

Figure 5: Solution U(y) to (56) for λ = 4, c = 1. The solution gives initialvalues U(y) = u(y, 0) obtained by transformation of boundary condition aty = 0, u(0, t) = 2et/(2λ) sin t. A Nystrom method based upon Gauss-Lobattoqudrature with n = 251 nodes was used to discretize (56). Exponential growthis observed for y → −∞. Insets show sharp resolution near the y = 0 boundaryobtained by numerical method.

31

0 10 20 30 40 50 60

�0.015�0.010�0.005

0.0000.0050.0100.015

y

vtel x�

y��

vFx�

y�

0 10 20 30 40 50 60

�0.4

�0.2

0.0

0.2

0.4

0.6

0.8

y

v x�y�

Figure 6: Comparison at t = 60 of telgrapher’s equation solution (51, dots) toFerry solution (28, line) for λ = 4, c = 1. The inset shows the absolute error inthe velocity vx(y, 60)

32

can apply the first equation in (48) everywhere to obtain

∂τxy

∂y=Re

2e−t/(2λ)

(∂u

∂t− 1

2λu

). (61)

This is an ordinary differential equation along constant t lines and we know that

τxy(y → ∞, t) = 0. Hence τxy can be computed as

τxy(y, t) =Re

2e−t/(2λ)

∫ y

y+ct

[∂u

∂t(s, t) − 1

2λu(s, t)

]ds , (62)

where we have taken into account that u = 0, ∂u/∂t = 0 for s > y + ct.

Comparisons of the UCM model τxy with those from the Ferry solution are

shown in Fig. 8. Notice that the differences in τxy are much greater than those

in vx.

In conclusion we obtain an exact solution of the nonlinear UCM model ex-

pressed as a convolution integral. The solution explicitly captures the nonlinear

phenomenon of normal stress generation in shear, which one computes inde-

pendently from a linear evolution of the velocity and shear stress, uncoupled

from the normal stresses. This behavior has experimental implications, in that

nonlinear behavior cannot be detected by monitoring velocity or shear stress.

In order to ascertain nonlinear behavior experimentally, the first normal stress

difference, which reduces to τxx, has to be measured. To test the accuracy

of the UCM model, those measurements can be compared to the model re-

sults. In the simulations below, we present normal stresses generated from the

boundary-controlled plate motion; the model predicts bulk strain thresholds to

reach a certain normal stress level for given UCM parameters.

33

0 2 4 6 8 10 12 14�0.15

�0.10

�0.05

0.00

0.05

y

vtel x�

y��

vFx�

y�

0 2 4 6 8 10 12 14

0.0

0.2

0.4

0.6

y

v x�y�

Figure 7: As in Fig. 6 but for c = 1, λ = 0.25. At small λ the Ferry steady statesolution exhibits very large errors with respect to the true initial transient.

0 10 20 30 40 50 60�0.4

�0.2

0.0

0.2

0.4

y

Τ xy

Figure 8: Comparison of τxy from Ferry solution (34, line) with that from thetelgrapher’s equation solution (62, dots) .

34

4 High resolution numerical solution

4.1 Numerical algorithm

For the more general nonlinear constitutive laws of Giesekus type with shear-

thinning (a = 0), closed-form solutions do not exist, so numerical methods are

employed. A primary purpose of this study is to provide a basis for inverse

characterization of nonlinear viscoelastic properties. Hence it is desirable to

minimize any artificial dissipation and dispersion that might arise in a numerical

computation, which is the baseline for matching with experimental data through

parameter fitting. For this purpose, we apply a high-resolution algorithm [15] for

hyperbolic PDEs as implemented in the Clawpack [17] or Bearclaw packages

[16]. We present the salient points of the numerical method here for complete-

ness. It is a finite volume, Godunov-type method in which the jumps between

adjacent finite volume cell averages are propagated in accordance with local

wave speeds. Consider an uniform discretization of the interval [0, H ] with step

size h. The cell center coordinates are yj = (j−1/2)h for j = 1, 2, ...,m, and the

cell edge coordinates are yj−1/2 = (j − 1)h, j = 1, 2, ...,m+ 1, with h = H/m.

The cell finite volume average is

Qnj =

1h

∫ yj+1/2

yj−1/2

q(y, tn) dy .

The evolution equation (19) can be rewritten as

qt = (A + B)q, A = −A(q)∂

∂x, B = ψ(q), (63)

35

with A the convective operator and B the source term operator. Equation (63)

is split into two stages using Strang splitting,

q(t+ Δt) = e(A+B)Δtq(t) ∼= eBΔt/2eAΔteBΔt/2q(t).

The source term operator part of the splitting is qt = Bq, a system of ODE’s

which is advanced in time using a second-order Runge-Kutta scheme. For the

convective part, we first represent jumps between adjacent cells as propagating

waves. The jump at the j − 1/2 interface ΔQnj−1/2, is decomposed on the

eigenbasis R = R((Qnj +Qn

j−1)/2)

ΔQnj−1/2 = Qn

j −Qnj−1 =

4∑l=3

αlj−1/2r

lj−1/2,

W lj−1/2 = αl

j−1/2rlj−1/2, l = 3, 4, nw = 2.

Note that only the r3, r4 eigenmodes from (41) are propagating, hence nw = 2

waves W3,4j−1/2 are required. The α coefficients required are

α3j−1/2 =

ΔQ3,j−1/2

2c2j−1/2Re− ΔQ1,j−1/2

2cj−1/2, α4

j−1/2 =ΔQ3,j−1/2

2c2j−1/2Re+

ΔQ1,j−1/2

2cj−1/2,

with Q1, Q3 the 1, 3 components of Q. Cell average values are updated by

Qn+1j = Qn

j − Δth

(W4

j−1/2 + W3j+1/2

),

along with second-order corrections (for complete details, see [15]). The method

is adaptive and second order in space and time. In the spirit of Literate Pro-

gramming [18], all software details are available both within the Bearclaw

package and on a webpage [19].

36

4.2 Validation

To ensure accuracy of the method, numerical convergence studies and compar-

isons with the known analytical results (36), (51) were carried out. A typical

numerical convergence result is presented in Table I. The overall scheme exhibits

an asymptotic convergence rate of 1.6.

my 20 40 80 160

lg h -1.30 -1.60 -1.90 -2.20

lg ε -0.85 -1.20 -1.58 -2.06Table I. Numerical convergence results. The relative error ε is that in the

velocity vx at y = H/2 computed with my subintervals and at time t = 12.96.

The exact solution is taken to be vx computed with my = 320 subintervals.

(Re = 1, λ = 2, H = 1, VH = 0)

Turning now to comparisons with the previous analytical results, formula

(36) is valid for a linear viscoelastic fluid after transients have died out. For

small values of the boundary-controlled bulk strain parameter γ, the numerical

results obtained by solving the nonlinear system (19) for a = 0 should be close

to those from the linear case (36). This is indeed verified computationally and

validates the numerical approach. Fig. 9 shows velocity profiles for a layer of

thickness H = 16, hence γ = 1/16 1. In order to make such a comparison the

parameters of the linear model must be related to those of the nonlinear model.

In the nonlinear UCM model one chooses a Reynolds number Re, a polymer

viscosity ηp and the Deborah number λ. We obtain the δ=√i Re/η∗ parameter

in the linear solution (36) by using formulas (27) to compute η′, η′′.

There is the possibility of a phase error due to transient effects. To allow for

this the comparison should be made at t � 1. The time history of the velocity

at the layer midpoint y = H/2 is shown in an inset on Fig. 9 for λ = 1,

Re = 1. After roughly 20 bottom plate cycles the velocity shows periodic

37

behavior. A time slice of the vx(y) dependence is chosen (tn = 129.912 in Fig.

9). To eliminate any remnant phase error due to transients, the time at which

to evaluate vx using (36) is determined by solving the equation

Im{eit sinh [δ(H − y1)]

sinh (δH)

}= vn

x (y1) (64)

with vnx (y1) denoting the velocity determined in the numerical procedure at

time tn. The phase error goes to zero as time increases, but slowly. For the

parameters above, the solution of (64) is t = 129.928 ∼= tn. The error between

the numerical solution and the analytical result is shown in an inset on Fig. 9.

5 10 15y

0.0010.0020.0030.0040.0050.006

err

0 20 40 60 80 100 120

�0.02�0.01

0.000.010.020.030.04

t

vx

0 5 10 15

�0.8

�0.6

�0.4

�0.2

0.0

0.2

y

v x

Figure 9: Comparison of numerical result for vx(y, 130) (dots) to analyticalprediction by (36). Insets show absolute error at t = 130 and time history ofvx(H/2, t) to indicate time needed for transients to pass.

38

4.3 Experiment design guidelines from numerical results

An important aspect in the design of an experiment for viscoelastic character-

ization is to ensure controllable boundary conditions and to qualitatively un-

derstand the variation of the flow variables across the channel height H under

various boundary conditions. We now apply the numerical algorithm presented

above to show the effect of some experimentally relevant boundary conditions.

We investigate the problem of inverse viscoelastic characterization using the

numerical algorithm in the following section. In order to assess the effect of

boundary conditions, fluid parameters are kept constant at H = 10, ηp = 1,

λ = 4, Re = 1, a = 0 and τyy = 0. Results are presented in Fig. 10-11 by show-

ing superimposed snapshots of vx(y, tn), τxx(y, tn), τxy(y, tn) at times tn = nΔt,

n = 0, 1, ..., 10, Δt = 0.8. Also shown are subsequent time positions of tracer

beads initially placed at X(y, 0) = 0. Since τyy = 0 in these simulations the

normal stress difference is directly given by τxx, N1 = τxx − τyy = τxx.

Open-cell experiment. In an open-cell experimental setup the viscoelastic

fluid is typically in contact with the surrounding air which is assumed to exert

a negligible tangential stress upon the viscoelastic layer. The top boundary at

y = 0 is assumed to be flat. Hence the boundary condition τxy(H, t) = 0 is set.

Results for are shown in Fig. 10. Note that the nonlinear interaction terms in

the UCM model lead to large values of τxx, larger than those for τxy near the

plate.

Stationary top plate. A closed-cell experimental setup is advantageous

in ensuring isolation of the sample and surrounding medium. Typical results

are shown in Fig. 11 when the top plate is kept at zero velocity vx(H, t) =

0. An interesting observation is that the velocity and stress profiles are very

similar to those obtained in the open-cell experiment. The main difference is the

much smaller range of motion exhibited by tracer beads when the top plate is

39

stationary by comparison to the open-cell experimental setup. If signal-to-noise

ratio of tracer bead positions becomes a concern at low strain thresholds, then

the open-cell setup is preferable.

5 Implications for inverse viscoelastic character-

ization

The theoretical results presented above provide new formulas in the linear finite

depth case, nonlinear UCM infinite depth case, and accurate numerical simula-

tion algorithms for the nonlinear finite depth case. Especially for experimental

measurements involving small samples, in which of necessity the depth is com-

parable to shear wavelengths, it is crucial to employ the finite depth formulas

deduced above rather than trying to fit to the infinite depth shear wave (28).

In subsequent work we shall use theses new tools to deduce viscoelastic param-

eters from bead tracking data in the presence of experimental noise. For this

paper we want to illustrate the effects that would arise from attempting to fit

measurements to inappropriate models even if there were no experimental error

whatsoever.

Instead of experimentally measured data we shall use bead positions gener-

ated from numerical simulations of a specific viscoelastic model. The goal then

is to infer linear and nonlinear viscoelastic properties from this “experimental

data set”, and to assess the magnitude of errors made as well as the dominant

source of error. Typically, least-squares procedures are employed to fit mea-

sured bead trajectories to theoretical formulas in order to obtain the complex

viscosity η∗(ω) (or complex shear modulus G∗(ω)). We adopt this procedure to

provide fair measurements of relative error.

If the same model is used both to generate the data and to carry out pa-

40

VX

Y

-0.5 0 0.5

2

4

6

8

10

TAUXX

Y

0 0.2 0.4 0.6 0.8 1 1.2

2

4

6

8

10

TAUXY

Y

-0.4 -0.2 0 0.2 0.4

2

4

6

8

10

xBEAD

Y

0 0.5 1 1.5

2

4

6

8

10

Figure 10: Numerical results for zero top shear stress, τxy(H, t) = 0, boundarycondition.

41

TAUXX

Y

0 0.2 0.4 0.6 0.8 1 1.2

2

4

6

8

10

TAUXY

Y

-0.4 -0.2 0 0.2 0.4

2

4

6

8

10

xBEAD

Y

0 0.5 1 1.5

2

4

6

8

10

VX

Y

-0.5 0 0.5

2

4

6

8

10

Figure 11: Numerical results for stationary top plate boundary condition,vx(H, t) = 0.

42

rameter fitting, then errors arise only from generation of model solutions or the

numerical least-squares procedure. We use this as a check to ensure the proce-

dure is accurate. Further errors in viscoelastic characterization arise if the data

is fit to a different model, e.g. if the Ferry infinite depth formulas are used to

characterize measurements made on a linear viscoelastic fluid in a finite-depth

cell. We present errors that arise from finite depth and nonlinear effects.

One unusual feature of these effects in the present non-dimensional formula-

tion is that there is no frequency dependence. This is because we have restricted

attention to a single-mode Maxwell type fluid.

5.1 Finite depth effects, linear viscoelastic model

Consider a viscoelastic fluid of depth H set into motion from rest by harmonic

oscillation of a bottom plate and upon which is placed a stationary top plate

(vH = 0). We wish to ascertain the error made in deriving viscoelastic param-

eters (η∗ or G∗) by fitting observed velocity profiles to the Ferry semi-infinite

solution (28), rather than the correct, finite-depth solution given in (36). Con-

sider the following best-case experimental scenario: measurements are made

at a large number of points, say m = 40, across the fluid depth and at small

time increments, Δt = 2π/n (e.g. n = 36 measurements per plate oscillation

period). The bead positions generated by the finite-depth solution (36) with

complex viscosity η∗H =η′H − η′′H are used to construct a least-squares fit to the

Ferry tracer bead position profile from whence another estimate of the complex

viscosity η∗F = η′F − iη′′F is obtained. The least squares sum is

S(η′F , η

′′F ) =

m∑j=1

n∑k=1

e2(yj , tn, δF , δH), δF =√i Re/η∗F , δH =

√i Re/η∗H ,

e(y, t,H, δF , δH) =∫ t

0

Im{eis

[e−δF y − sinh [δH(H − y)]

sinh (δHH)

]}ds (65)

43

The error function e(y, t,H, δF , δH) is presented in Fig. 12-14. As expected

when H increases the bead position error decreases (Fig. 12). Recall the link

between complex viscosity and Maxwell fluid parameters, η′ = ηp/(1 + λ2),

η′′ = ηpλ/(1 + λ2). We can increase the magnitude of the complex viscosity η∗

by increasing ηp and vary the phase of η∗ by varying λ but setting ηp = 1 + λ2

always. When the absolute value of the complex viscosity increases, the error

increases and furthermore becomes quasi-linear over the depth (Fig. 13). This

makes the identification of parameters ill conditioned. As the relaxation time

of the fluid λ becomes larger, the magnitude of the error remains essentially the

same Fig. 14. At small λ the identification of η∗ will be more difficult since

there are fewer oscillations of the shear wave within the depth H (Fig. 14).

0 1 2 3 4 5�3

�2

�1

0

1

2

y

e

0 1 2 3 4 5

�0.3

�0.2

�0.1

0.0

0.1

0.2

y

e

0 1 2 3 4 5

0.000

0.005

0.010

y

e

0 1 2 3 4 5

�0.0002

0.0000

0.0002

0.0004

0.0006

y

e

Figure 12: Error in bead positions between finite depth (36) and infinite depth(28) solutions (Re = 1, η′ = η′′ = 1). Top row H = 1, 5. Bottom row H =10, 15.

Snapshots of the error in identification of η∗ as presented above are useful,

44

0 1 2 3 4 5�6. � 10�10

�4. � 10�10

�2. � 10�10

0

2. � 10�10

y

e

0 1 2 3 4 5

�0.0015

�0.0010

�0.0005

0.0000

0.0005

0.0010

y

e

0 1 2 3 4 5

�0.05

0.00

0.05

0.10

0.15

0.20

y

e

0 1 2 3 4 5

�0.2

�0.1

0.0

0.1

0.2

0.3

0.4

y

e

Figure 13: As in Fig. 12. Constant H = 10. Top row η′ = η′′ = 0.05, 0.5.Bottom row η′ = η′′ = 5, 50.

but limited to single values of the complex viscosity. To obtain an overall picture

of the error made in identification of viscoelastic parameters we introduce the

complex mapping δF (δH), δF : C → C. This provides a measurement of the

error in fitting the true finite-depth solution to the semi-infinite Ferry solution,

and can also serve to correct the Ferry δF value. It is preferable to use δ rather

than the complex viscosity η∗ for two reasons: 1. δ = α+iβ is directly related to

experimentally measured quantities, the penetration depth 1/α and the shear

wavelength 2π/β; 2. The inversion η∗ = i Re/δ amplifies errors made when δ

is small and hence relatively small errors in δ give large changes in η∗.

A useful technique, common in the studies of conformal mappings, is to

present the δF (δH) mapping as a deformation of the αH = j Δα =constant and

βH = kΔβ =constant lines (j, k are integers and Δα, Δβ are the spacings with

which we sample δH space). If no error is made, an initially Cartesian (αH , βH)

45

0 1 2 3 4 5�2. � 10�7

�1. � 10�7

0

1. � 10�7

2. � 10�7

3. � 10�7

4. � 10�7

y

e

0 1 2 3 4 5

�5. � 10�7

�4. � 10�7

�3. � 10�7

�2. � 10�7

�1. � 10�7

0

y

e

0 1 2 3 4 5

�2. � 10�7

0

2. � 10�7

4. � 10�7

6. � 10�7

y

e

0 1 2 3 4 5

�4. � 10�7

�2. � 10�7

0

2. � 10�7

4. � 10�7

6. � 10�7

y

e

Figure 14: As in Fig. 12. Constant H = 10. Top row λ = 0.5, 1. Bottom rowλ = 5, 10. ηp = 1 + λ2 in all cases.

grid would be transformed into itself identically. The deviation from a Cartesian

map of the (αF , βF ) lines serves as an indication of the error made in using the

Ferry formula in the least squares fit to obtain the complex viscosity. To cover

as wide of range of values as possible the mappings are presented in logarithmic

coordinates. Typical results for H = 2, 5 are presented in figures 15, 16. The

purely viscous case α = β =√Re/2 is shown for reference. As expected errors

decrease as α increases - the shear wave is more attenuated and the boundary

condition at H has less influence. Note however that large values of α in bead

tracking experiments imply small bead displacements, which are difficult to

track. For smaller values of α the wavelength 2π/β is poorly identified. The

region of poor identification is smaller for smaller H as expected.

For certain regions it is even possible to use the δF (δH) mapping to correct

an erroneous estimate made by fitting to the Ferry semi-infinite solution! This

46

would be of interest in reinterpreting archived experimental data. Fig. 17 shows

a sequence of values obtained by fitting to the infinite depth formula and arrows

to the correct values (α, β) for that particular fluid.

It is apparent that very large errors are made when trying to deduce vis-

coelastic parameters from bead tracking data if the fit is carried out on the

infinite depth formula rather than the correct, finite depth model.

�1.0 �0.5 0.0 0.5 1.0�1.0

�0.5

0.0

0.5

1.0

lg�ΑH�

lg�ΒH�

Figure 15: The mapping δF (δH) for H = 2, Re = 1. The dot represents thepurely viscous case α = β =

√1/2. Large errors in viscoelastic characterization

arise for α < 10−0.25.

47

�1.0 �0.5 0.0 0.5 1.0�1.0

�0.5

0.0

0.5

1.0

lg�ΑH�

lg�ΒH�

Figure 16: As above for H = 5.

48

�0.5 �0.4 �0.3 �0.2 �0.1 0.0�1.0

�0.9

�0.8

�0.7

�0.6

�0.5

lg�ΑH�

lg�ΒH�

Figure 17: A plot of δF (δH) for H = 5, Re = 1 with arrows indicating how tocorrect erroneous (αF , βF ) values to the true viscoelastic fluid values (αH , βH),assuming linear viscoelasticity.

49

5.2 Onset on non-linear effects

We now consider how non-linear effects influence viscoelastic characterization.

Velocity profiles for a viscoelastic fluid placed between an oscillatory bottom

plate and a stationary top plate are generated by numerically solving system

(19). The numerically-generated velocity profiles contain non-linear effects from

(19). In order to quantify the onset of nonlinearity, we shall do a least-squares

fit of the numerical velocity data to the linear finite-depth formula (36) to obtain

the parameters (α, β) characterizing the shear wave. We present typical results

of the fitting procedure, highlighting the role of the bulk strain γ = 1/H and

Giesekus parameter a which controls shear thinning. At low values of γ (e.g.

γ < 0.1) the fitting procedure reproduces the numerical parameters used in the

simulations – this was used to validate the least squares algorithm used.

Moderate bulk strain, UCM fluid (γ = 0.25, a = 0). Results are pre-

sented in Fig. 18. At this moderate bulk strain value the fitting procedure

works very well for small Deborah numbers λ (λ ≤ 1). As the Deborah num-

ber is increased the results become markedly worse with errors in parameter

identification greater than 100% at λ = 16.

Large bulk strain, UCM fluid (γ = 0.5, a = 0). When the bulk strain

is large, the fitting procedure fails at all values of λ (Fig. 19).

Moderate bulk strain, Giesekus fluid (γ = 0.25, a = 0.5). For moderate

bulk strains on a shear-thinning fluid we notice a systematic overestimate of α

even while β is captured rather well (Fig. 20). This is the expected behavior.

The penetration depth of wave propagating into the fluid 1/α is underestimated

by the linear viscoelastic model hence the α values are higher.

50

��

�� � � � � � � � �

��

� � � � � � � � � �

��

���� �

� �� �

��������

���

��������

������

��

Λ�1

Λ�2

Λ�4

Λ�16H�4a�0

ΒH

2�Π0.6

ΒH

2�Π1.3

ΒH

2�Π1.9

ΒH

2�Π2.6

ΒH

2�Π3.2

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

1

2

3

4

5

6

Α

Β

Figure 18: Comparison of shear wave identified by fitting to the linear, finitedepth formula (36), (circles) to parameters used to generate data numericall(squares). Bulk strain γ = 1/H = 0.25.

51

��

�� � � � � � � � �

� �����

��

����

��

����� �

����

����

�����

����

��������

���

��

���

Λ�1Λ�2

Λ�4

Λ�16

H�2a�0

ΒH

2�Π0.6

ΒH

2�Π1.3

ΒH

2�Π1.9

ΒH

2�Π2.6

ΒH

2�Π3.2

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

1

2

3

4

5

Α

Β

Figure 19: As above, but for higher strain γ = 1/H = 0.5. Nonlinear effects areso strong that fitting to linear model leads to meaningless results.

52

��

��������

��

��

��

��

��

������

��

��

��

Λ�1

Λ�2Λ�4

H�4a�0.5

ΒH

2�Π0.6

ΒH

2�Π1.3

ΒH

2�Π1.9

0.2 0.4 0.6 0.8 1.00.5

1.0

1.5

2.0

2.5

3.0

Α

Β

Figure 20: As above for a fluid with shear thining. a = 0.5 (Leonov model),γ = 1/H = 0.25 bulk strain.

53

6 Conclusions

For purposes of inverse characterization experiments in low volume samples of

biological fluids (such as pulmonary airway surface liquids, our motivation for

this work), we have extended the Ferry shear wave analysis to include finite-

depth within linear viscoelasticity, and then to include nonlinearity with con-

stitutive laws of Maxwell type. The linear models are exactly solvable, allowing

for an explicit measure of errors in the inverse characterization of a known vis-

coelastic material due solely to finite depth effects. At frequencies and bulk

strains typical for conditions encountered by the airway surface liquid in the

lung, errors in the recovery of storage and loss moduli are enormous when fit-

ting to inappropriate theoretical models. For viscoelastic fluids that can safely

be assumed to be behave linearly, fitting to the newly derived finite depth for-

mula (36), which accounts for counterpropagating shear waves and stress modes,

is shown to work well. If the fluid behaves nonlinearly as controlled bulk strain

from the driven plate is ramped up, either due to convective nonlinearity or

shear thinning, then inverse recovery of viscoelastic parameters by fitting to the

extended finite-depth linear formulas induce significant errors.

These modeling tools have been developed in context with a new device, a

micro, parallel-plate rheometer, which allows both types of velocity controls on

the opposing flat interface to the driven plate. The present paper has illustrated

recovery of linear and nonlinear viscoelastic parameters as well as predictions

of onset of nonlinear effects at threshold values of the imposed strain from

the driven plate. In a future article, we will apply these modeling tools to

experiments on biological liquids.

Acknowledgements. This work originated with and has benefited from dis-

cussions with colleagues involved in the University of North Carolina Virtual

Lung Project. The research effort was sponsored by NIH grant R01-HL077546-

54

01A2, NSF Research Training Grant DMS-0502266, NSF-DMS 0604891, and

ARO 47089-MS-SR.

References

[1] John D. Ferry, Studies of the Mechanical Properties of Substances of High

Molecular Weight I. A Photoelastic Method for Study of Transverse Vibra-

tions in Gels, Rev. Sci. Inst., 12, 79-82, (1941)

[2] John D. Ferry, W.M. Sawyer, and J. N. Ashworth, Behavior of Concentrated

Polymer Solutions under Periodic Stresses, J. Polym. Sci. 2:593-611, (1947)

[3] F.T. Adler, W.M. Sawyer, John D. Ferry, Propagation of transverse waves

in viscoelastic media, J. Appl. Phys., 20, 1036-1041 (1949)

[4] G.B. Thurston, J.L. Schrag, Shear wave propagation in a Birefringent vis-

coelastic medium, J. Appl. Phys. 35, 144, 1964.

[5] Hermann Schlichting, Boundary layer theory, McGraw-Hill, New York,

1960.

[6] J.D. Ferry, Viscoelastic properties of polymers, John Wiley, New York, 1980.

[7] R.B. Bird, R.C. Armstrong, O. Hassager, Dynamics of Polymeric Liquids,

Fluid Mechanics, Vol. 1, Wiley, New York, 1987.

[8] D. B. Hill, B. Lindley, M.G. Forest, S.M. Mitran, Microrheology of DNA

solutions from bead tracking of shear waves, UNC preprint, 2007.

[9] D.B. Hill, R. Superfine, Microbead rheology, (In preparation).

[10] H. Matsui, S. Randell, S. Peretti, C. W. Davis, and R. C. Boucher, Coor-

dinated clearance of periciliary liquid and mucus from airway surfaces, J.

Clin. Invest., 102(6), 1125-1131, 1998.

55

[11] S. M. Mitran, Highly accurate solutions of the non-linear one-dimensional

viscoelastic shear wave problem for Giesekus type fluids, (in preparation)

[12] R.G. Larson, Constitutive Equations for Polymer Melts and Solutions, But-

terworths, Guildford, UK, 1988.

[13] J.C. Crocker, M.T. Valentine, E.R. Weeks, T. Gisler, P.D. Kaplan, A.G.

Yodh, and D.A. Wietz, Phys. Rev. Lett. 85, 888, 2000.

[14] M. Gregory Forest, Qi Wang, Dynamics of slender viscoelastic free jets,

SIAM J. Appl. Math., 54(4), 996-1032,(1994).

[15] R. J. LeVeque, Wave propagation algorithms for multi-dimensional hyper-

bolic systems, J. Comp. Phys., 131, 327-353, 1997.

[16] S. Mitran, BEARCLAW - A software frame-

work multiphysics numerical simulation,

http://www.amath.unc.edu/Faculty/mitran/bearclaw.html.

[17] R. J. LeVeque, CLAWPACK - Conservation Law Package,

http://www.amath.washington.edu/~claw.

[18] D.E. Knuth, Literate Programming, Stanford, CSLI Lecture Notes, no. 27,

1992.

[19] http://www.amath.unc.edu/Faculty/mitran/research/viscelas/shear1d.xhtml

56