You are on page 1of 31

J. Non-Newtonian Fluid Mech.

106 (2002) 29–59

Drop formation dynamics of constant


low-viscosity, elastic fluids
J.J. Cooper-White∗ , J.E. Fagan, V. Tirtaatmadja, D.R. Lester, D.V. Boger
Department of Chemical Engineering, Particulate Fluids Processing Centre,
University of Melbourne, 3010 Melbourne, Vic., Australia
Received 11 September 2001

Abstract
The dynamics of drop formation under gravity has been investigated as a function of elasticity using a set of
low-viscosity, ideal elastic fluids and an equivalent Newtonian glycerol–water solution. All solutions had the same
shear viscosity, equilibrium surface tension, and density, but differed greatly in elasticity. The minimum drop
radius in the early stages of drop formation (necking) was found to scale as expected from potential flow theory,
independent of the elasticity of the solutions. Thus, during this stage of drop formation when viscous force is still
weak, the dynamics are controlled by a balance between inertial and capillary forces, and there is no contribution
of elastic stresses of the polymer. However, upon formation of the pinch regions, there is a large variation in the
drop development to break-off observed between the various solutions. The elastic solutions formed secondary fluid
threads either side of a secondary drop from the necked region of fluid between the upper and lower pinches, which
were sustained for increasing amounts of time. The break-off lengths and times increase with increasing elasticity
of the solutions. Evolution of the filament length is, however, identical in shape and form for all of the polymer
solutions tested, regardless of differing elasticity. This de-coupling between filament growth rate and break-up time
(or equivalently, final filament length at break-up) is rationalised. A modified force balance to that of Jones and Rees
[48] is capable of correctly predicting the filament growth of these low-viscosity, elastic fluids in the absence of any
elastic contributions due to polymer extension within the elongating filament. The elongation of the necked region
of fluid (which becomes the filament) is dominated by the inertia of the drop, and is independent of the elasticity of
the solution. However, elasticity does strongly influence the resistance of the pinch regions to break-off, with rapid
necking resulting in extremely high rates of surface contraction on approach to the pinch point, initiating extension
of the polymer chains within the pinch regions. This de-coupling phenomenon is peculiar to low-viscosity, elastic
fluids as extension does not occur prior to the formation of the pinch points (i.e. just prior to break-up), as opposed
to the high viscosity counterparts in which extension of polymers in solution may occur even during necking. Once
steady-state extension of the polymers is achieved within the pinch at high extension rates, the thread undergoes
elasto-capillary break-up as the capillarity again overcomes the viscoelastic forces. The final length at detachment


Corresponding author. Tel.: +61-38344-4704; fax: +61-38344-6233.
E-mail address: jjcw@unimelb.edu.au (J.J. Cooper-White).

0377-0257/02/$ – see front matter © 2002 Elsevier Science B.V. All rights reserved.
PII: S 0 3 7 7 - 0 2 5 7 ( 0 2 ) 0 0 0 8 4 - 8
30 J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59

and time-to-break-off (relative to the equivalent Newtonian fluid) is shown to be linearly proportional to the longest
relaxation time of the fluid.
© 2002 Elsevier Science B.V. All rights reserved.

Keywords: Drop formation; Break-up; Elastic fluids; Filament extension; Pinch dynamics

1. Introduction

The dynamics and associated mechanisms of drop formation and fluid rupture from a nozzle under the
influence of gravity alone has seen much interest since the mid 1800s [1–3]. These early works established
that as a fluid column begins to exit a nozzle, the speed of the growth and shape of the emerging drop
are primarily controlled by the opposing forces of surface tension and gravity. While gravity causes the
fluid to exit the nozzle, surface tension strives to attain a minimum surface area, which initially slows the
fluid from exiting the nozzle. As gravity overcomes surface tension, the fluid emerges and surface tension
forces cause the formation of a neck, from which a primary filament and drop evolve. Inertia obviously
plays a major role in the development, elongation and breaking the drop. Even for highly viscous fluids,
the effects of inertia cannot be ignored as the filament becomes thin and the drop approaches break-up
[4,5]. Break-up of the drop from a suspended column of fluid attached to the primary thread is a result
of the propagation of surface oscillations, caused by previous drop detachment or other external sources,
along the length of this thread (the classic Rayleigh instability phenomenon [3]). As the drop detaches
and falls, surface tension creates a sphere to enclose the maximum volume of fluid in the minimum
surface area.
In the early 1970s, with the development of ink-jet printing technology, researchers began investigating
the dynamics of high-speed drop formation in more detail. Early mathematical models of continuous jet
streams of differing fluid properties (viscosity, surface tension and density) were presented (e.g. Bruce
[6]), and these ideas were also applied to drop-on-demand ink-jet technology. Focuses were on the
drop-on-demand production of drops of consistent volume and velocity, coping with the problem of
small satellite droplets which often follow the primary drop, and understanding the effects of variables
such as fluid viscosity and density, nozzle diameter and the thickness of the nozzle wall (e.g. Young [7]).
Peregrine et al. [8] re-sparked significant interest in this field with their qualitative discussion and
pictures of drop formation with low-viscosity Newtonian fluids up to and past the ‘pinch’ point. The
‘pinch’ point is reached after a period of rapid necking of the drop downstream of the nozzle, and marked
by the formation of a secondary thread or microthread, far thinner than the primary thread (formed from
the necked fluid), just prior to the final break-up event. Shi et al. [9] provided a computational analysis
of what Peregrine et al. [8] had discussed, and compared it with further experimental work. It was found
that after the initial pinch point or secondary thread formation, this secondary liquid thread could lead to
a series of smaller threads with still thinner diameters [9].
Shi et al. [9], Zhang and Basaran [10] and Wilkes et al. [11] studied the effect of viscosity on the drop
formation and break-up of Newtonian fluids. Zhang and Basaran [10] observed the thread diameter and
elongation length of the forming drop, and the volumes of primary and satellite droplets formed. The
effects of nozzle geometry, liquid flowrate, viscosity and surfactants were systematically investigated. In
comparing water with an 85% glycerol solution, Zhang and Basaran [10] found that increasing viscosity
increased the maximum length of the primary thread prior to break-up. This agrees with much earlier
J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59 31

experimental reports by Narasinga Rao et al. [12] and Edgerton et al. [13], and with the finite element
computations of Wilkes et al. [11]. It is now well understood that viscosity gives rise to a viscous pressure
in the thinning filament, which opposes the capillary pressure, dampening surface oscillations and, hence,
increasing the lifetime (and length) of the primary thread [10,14]. Additionally, Brenner et al. [15] found
that viscosity has a significant effect on the shape of the fluid interface at the point of attachment between
the drop and the filament, immediately before rupture.
As mentioned earlier, surface tension plays a considerable role in the drop formation process. Zhang
and Basaran [10] found that increasing the surface tension caused a drop to appear more spherical during
formation, and increased the volume of the drop and the length of the drop at detachment. However,
Badie and de Lange [16] report that surface tension has little effect on the volume of the primary drop
from a drop-on-demand ink-jet. The liquid flowrate and the nozzle geometry are also known to affect
drop formation, with larger drops being formed from larger nozzles and at higher flowrates.
Henderson et al. [17] found that there are two distinct pinch methods. Between the nozzle fluid and the
primary thread, the thread gradually necks to form an upper pinch. However, between the primary thread
and the drop, there is not gradual necking, but rather the distinct rapid formation of the lower pinch via a
secondary or microthread, as was seen by Peregrine et al. [8] and Shi et al. [9]. Both threads are subject
to Rayleigh instabilities and break in multiple locations, forming small secondary satellite droplets in
most cases. The pinch point between the primary thread and drop usually breaks first, followed swiftly by
break-up at the upper pinch point, which is enhanced by the capillary waves propagating up the primary
thread to the nozzle fluid; although the reverse has also been observed [9].
Numerous studies have provided simplified Navier–Stokes equations for incompressible flow with a
free surface in an effort to describe drop formation [18–21]. The effects of fluid flow within [21] and
surrounding [22] the forming drop have also been considered. While more recently, the solutions of the
complete Navier–Stokes equations to describe drop formation have been given by Wilkes et al. [11] and
Gueyffier et al. [23]. The finite element calculations of Wilkes et al. [11] were able to capture both the
gross features of the drop formation, including the limiting length of a drop at break-up, and its fine
features such as the secondary thread that forms from the main thread at certain fluid conditions.
For drop formation from a nozzle under gravity of viscoelastic fluids, the recent experimental results
of Amarouchene et al. [24] are very relevant. Although their work looked at low-viscosity, elastic fluids
similar to those used in this study, they only considered the behaviour of very high molecular weight
polymers (4 and 8 × 106 g/mol) in solution. Unfortunately, they also did not give any indication of the
magnitude of the relaxation time scale of their solutions. In addition, although not quantified by these
workers, such high molecular weights would generally result in the solutions being shear-thinning to
different extents. In contrast to the solutions of Amarouchene et al. [24], the fluids used here are a set of
well characterised and matched constant low-viscosity, elastic fluids [25], of equivalent density, surface
tension and constant shear viscosity, with the only difference being the addition of different molecular
weight polymers at dilute concentrations. The result of such well designed solutions is a set of fluids which
vary significantly only in their elastic properties and, hence, relaxation times. These elastic properties and
relaxation times are also quantified independently and related to the associated drop formation process.
Thus, the current paper provides a detailed insight into the effects of varying degrees of elasticity on
the dynamics of drop development, elongation and break-up in low-viscosity solutions. The variation in
drop length and minimum drop radius is investigated, and the similarities and differences amongst the
various solutions compared. It will be shown that the length at detachment and time-to-break-off of the
fluid thread can be correlated with the relaxation time. Finally, the dynamics of the drop evolution up to
32 J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59

break-up of the solutions with differing elasticity is discussed and the mechanism that helps sustain the
drop from break-up is proposed.

2. Drop development, formation and break-up

2.1. Region approaching break-off or the ‘pinch’ region

2.1.1. Newtonian fluids


Experimental observations by previous investigators utilising Newtonian fluids, e.g. Shi et al. [9],
Kowalewski [26], have shown that free-surface shapes of a given fluid are very similar when approaching
the pinch point (i.e. within the pinch region), totally independent of the initial conditions, such as the
nozzle radius [5]. The dynamics of break-up are characteristic of the non-linear properties of the equations
of motion, and as the flow near the break-up point accelerates, only the fluid that is within (or near) the
pinch region is able to follow. The break-up is, thus, localised in both space and time. Consequently, the
singularity becomes independent of both the initial conditions and the experimental method [5], whether
it be a jetting, liquid bridge or dripping experiment (as performed in this study). Keller and Miksis [27]
first introduced the principle of self-similarity for free surface flows where the solution at different times
can be mapped onto itself by rescaling the axes. Similarity solutions of the Navier–Stokes equations of
motion for inviscid (infinite Reynolds number) [28,29], low-viscosity (finite Reynolds number) [15] and
highly viscous (zero Reynolds number) [20] Newtonian fluids have been found to adequately represent
the drop formation process approaching the pinch region in such flow regimes.
However, it has been shown that the dynamics of this drop formation can pass through a number of
these regimes as break-up is approached [30–32]. As discussed by Lister and Stone [31], depending
on the viscosity of the Newtonian fluids, the dynamics may initially fall in either the ‘viscous thread’
(viscosity dominated) or ‘potential flow’ (inertial dominated) regime. However, as the break-up singularity
is approached, both inertia and viscosity must be considered. This is the ‘inertial-viscous’ regime proposed
by Eggers [4], in which the dynamics is given by the balance between the capillary, inertial and the viscous
resistance of the fluid. Additionally, a fourth regime has been shown numerically to exist close to break-up
[31], the so-called ‘two-fluid Stokes flow’ regime, in which viscosity of the surrounding fluid, no matter
how small, must be accounted for, and the asymptotic balance is between the surface tension and viscous
stress of the two fluids.
Yildirim and Basaran [32] were able to predict numerically the transitions from symmetric flow for a
low-viscosity Newtonian fluid in the potential flow regime, where Rmin ∝ τ 2/3 , Rmin being the filament
radius minimum and τ the time-to-break-up, to asymmetric inertial-viscous flow, where Rmin ∝ τ, as
time-to-break-up approaches zero. They also predict Rmin ∝ τ for high viscosity Newtonian fluids down
to very short time, covering both the viscous thread and inertial-viscous flow regimes. Experimentally,
the results of Amarouchene et al. [24] with water confirm the expected potential flow scaling theory of
filament radius variation with time to the power 2/3, to within 0.1 ms of break-up, but no transition to the
linear dependence region was noted. This is possibly due to the very short internal time scale of water,
beyond both the temporal and spatial resolution limit of any present experimental technique.
For the viscous thread case, the variation of the minimum radius of the necking fluid can be written as
σ
Rmin = κ τ, (1)
ηs
J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59 33

where σ and ηs are the fluid surface tension and viscosity. The value of the proportionality factor κ
has been determined by Papageorgiou [20], using self-similarity, to be 0.0709 when inertial effects can
be neglected. In the case where inertial effects cannot be neglected, such as when the necking process
rapidly approaches break-off, Eggers [4,5] predicted a value of κ = 0.0304. The first value for κ =
0.0709 has been confirmed by McKinley and Tripathi [33] for a high viscosity Newtonian fluid using the
capillary break-up device. Additionally, Rothert et al. [34] clearly shows that, for a viscous Newtonian
glycerol–water solution (i.e. viscosity 99 mPa s), a transition exists at approximately 2 ms from the drop
break-off time; both these regions show linear dependence of radius on time-to-break-off, but with the
proportionality factor changing from 0.0709 to 0.0304 as break-off is approached. The results clearly
confirm the dynamic pathway for capillary break-up of viscous Newtonian fluids from the symmetric
viscous thread to the asymmetric inertial-viscous flow regime, as outlined by Lister and Stone [31].
However, no experimental evidence of the two-fluid Stokes flow regime shown by Lister and Stone [31]
has yet been found. As also pointed out by the authors, depending on the initial Reynolds number and
viscosity ratio of the fluids, the transition may proceed directly from the viscous thread to the two-fluid
Stokes flows, without going through the inertial-viscous regime.
As both low and high viscosity Newtonian fluids have been shown to pass through an asymmetric
scaling regime as break-up or the pinch region is approached, the behaviours observed very close to the
pinch region for highly viscous fluids are expected to be characteristic of lower viscosity fluids. In this
flow regime, the behaviour near the pinch region can be decoupled from the initial and macroscopic
conditions, and the appropriate length and time scales are the ‘inner’ scales, termed the viscous length, lv
(=η2s /ρσ), and the corresponding viscous time, tv (=η3s /ρσ 2 ), where ρ is the fluid density. For water the
length scale lv is almost of molecular level, and thus, the length of the secondary thread (or microthread)
at the pinch is not observable. However, self-similarity suggests that if the viscosity is doubled (for the
same density and surface tension values), break-off should look similar when viewed on length scales
four times as large and time scales eight times as long [4]. Such achievable variations in lv and tv have
been confirmed experimentally [9,11].
Kowalewski [26] described the transition to the asymmetric scaling regime, the timing of which is
dependent on the viscosity of the fluid and the initial conditions, as the formation of a secondary thread
from the original neck or primary thread. This newly formed secondary thread is highly asymmetric,
and has been observed to get longer with increasing viscosity in Newtonian fluids [9,26]. The viscous
length and time scales, lv and tv , provide an effective measure of the length and time of the region
where asymptotic self-similarity can be expected [5]. For a Newtonian liquid of viscosity of the order of
6 mPa s, as used in this study, the values of lv and tv are expected to be of the order: lv = 5.1 × 10−7 m and
tv = 4.9 × 10−8 s. These values are beyond the spatial and temporal resolution of any present measuring
technique, yet as we shall see later, the pinch regions exist for many milliseconds after their inception in
our low-viscosity elastic solutions, far beyond the time predicted by tv . This phenomenon must be due to
effects other than the fluid shear viscosity, such as the elasticity of these solutions.
Although pinch behaviour is self-similar near break-up, the exact form of the drop near the nozzle prior
to the pinch region exhibits a strong dependence on the viscosity of the fluid. Low-viscosity fluids, e.g.
water, all show a conical tip attached to a sharp front at the pinch, whilst for high viscosities (approximately
>80 times viscosity of water, e.g. 85% glycerol–water solution (see Shi et al. [9], Brenner et al. [15]),
long, thin threads are produced with a broader, smoother attachment. These long, thin, cylindrical threads
are characteristic of the final stages of filament development prior to the onset of pinching and are well
represented by the lubrication equations [5].
34 J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59

In the drop formation and break-up process, the relative magnitudes of the forces of gravity, inertia,
capillarity and viscosity vary greatly with fluid properties. Several terms are often used to compare the
relative magnitudes of these opposing forces. The Bond number, Bo, gives the ratio of the gravitational
to capillary effects, and is only important in the initial stages of drop formation: Bo = ρgR20 /σ, where
g is gravitational acceleration and R0 is the radius of the nozzle. The Reynolds number gives a balance
between the inertial to viscous forces and is given by: Re = ρR0 v/ηs , where v is the velocity of the
system. Edgerton et al. [13] and Eggers [5] have suggested that for jet break-up and drop formation
studies, the velocity of the system may be expressed as a function of only the intrinsic properties of
the fluid with v = lv /tv = σ/ηs as the process is self-similar, as discussed earlier, so that the intrinsic
Reynolds number may be expressed as: Re = ρR0 σ/η2s . Note, however, that this modified definition of
Re, which no longer includes an external velocity scale, is simply the inverse square of the Ohnesorge
number, and is the ratio of capillary time to the time for momentum transport within the system. The
final term, is the capillary time, which describes the relative importance of viscous and capillary forces:
tcap = ηs /(σ/R0 ). The choice of these three terms is made such that only the intrinsic properties of the
fluids and one physical length scale are involved; it is totally independent of the type of the capillary break-
up process, be it jetting, dripping or capillary break-up experiments, and hence, the rate of the process
considered. An external velocity scale is, thus, irrelevant on approach to break-up as the fluid determines
the resultant dynamics of this process. In addition, for low-viscosity fluids, the flow will be an inertially
dominated flow within the free surface and either reference to a high Reynolds number or low Ohne-
sorge number flow may be utilised interchangeably during the discussion of drop formation and break-up
dynamics.

2.1.2. Non-Newtonian fluids


Much of the previous work in the field of drop formation is on Newtonian fluids. However, the formation
and subsequent break-up of the fluid drops from a nozzle is significantly affected by fluid elasticity and
other complex fluid properties. Non-Newtonian fluids in drop formation have attracted limited attention so
far. Eggers [5] discussed numerous one-dimensional models, which incorporate the viscoelastic effects of
polymer addition on drop formation, jets and liquid bridges [35–40], but outlined the difficulty associated
with such simplified formalisations of this highly non-linear problem. Recently, Yildirim and Basaran
[32] provided a comparison between the 1D and 2D models for Newtonian as well as a shear-thinning
fluids, but no elastic effects have been included.
It is well known that the addition of a minute quantity of high molecular weight polymer to a Newtonian
solvent can significantly affect its flow behaviour, and hence, introduces an inherently non-linear behaviour
to an otherwise linear system [5]. In a drop or capillary break-up process, the most significant outcome
of polymer addition to low-viscosity Newtonian solvents on the flow dynamics is a stabilisation effect
[32,41–43]. Increased stability of the filament results from the significant increase in the viscous stress
due to extension of the polymer chains under extensional flows, typical during formation of the primary
filament, but also on the approach to and within the pinch region. The extensional thickening or strain
hardening of these solutions is believed to result in the classic ‘beads-on-a-string’ shape in polymeric
jets, as originally shown by Goldin et al. [44] and further investigated by others, including Mun et al.
[41], Christanti and Walker [42,43], who investigated the influence of polymers on jet stream break-up
utilising similar fluids as used in the present study. They found that both the polymer molecular weight and
polymer concentration affect the break-up dynamics, and showed that solutions with higher extensional
viscosity and relaxation time are more effective at retarding break-up.
J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59 35

Little detailed mechanistic insight as to the reasons why such phenomena exist in these dilute polymer
solutions has been provided by these experimental studies. However, the recent numerical works of Chang
et al. [45] may be one of the first to provide the necessary model development to allow interpretation of
the outcomes of these jetting studies, and for that matter, the outcomes of this present drop formation
study. Chang et al. [45] examined, with asymptotic analysis and numerical simulation, what they term the
‘iterated stretching’ dynamics of model viscoelastic jets, utilising the Oldroyd-B and FENE constitutive
equations. They propose that during jetting, three defined stages can be isolated from the formation
dynamics of this bead-filament process: stretching, elastic drainage and surface recoil.
Chang et al. [45] suggest that during jetting of a dilute polymer solution, it is the usual Rayleigh
instability that first stretches the local uniaxial extensional flow region near a minimum in the jet radius
(i.e. neck) into a primary filament (between two beads). The attainment of a ‘final’ radius, which was
shown to be proportional to the capillary and Weissenberg numbers of the fluid, signifies the creation of
an axisymmetric cylindrical filament between the beads. However, prior to attaining this final radius or
cylindrical filament, the formation of a stagnation point at the minimum of the neck can produce extension
of the polymers within solution at this minimum. This is only relevant if their characteristic relaxation
time is large enough for the process to invoke the ‘coil-stretch’ transition of the polymer, i.e. the Deborah
number is greater than 0.3–0.5. This minimum criterion was briefly discussed by Chang et al. [45], but
for their work they assumed that a strong enough flow exists to produce an elastic response in the fluid
in all stages. Such an assumption must of course be justified as relevant to the fluids in question prior to
using this proposed mechanistic model to interpret experimental outcomes. In any case, this extension of
the polymer will result in an increase in the elastic stress at the minimum radius of the filament, which
then retards the flow of the fluid from the jet minimum to the jet nodes. In this primary stage, such a flow
is driven by the azimuthal curvature differences between these two reference points. In the case when the
polymer relaxation time is far shorter than the stretching event, Chang et al. [45] noted that the excess
elastic stresses will never be triggered in the minimum radius and the azimuthal pressure gradient will
drive the jet towards pinch-off without a stretching event. Thus, this ‘stretching’ event is only applicable
when the relaxation time of the fluid is of a similar (or larger) magnitude to the process time of stretching.
This ‘stretching’ event ceases, however, when the capillary pressure increases sufficiently, as the min-
imum radius decreases, to balance the elastic stress. At this point, the liquid within the filament cannot
continue to drain towards the jet node as the gradient in the azimuthal curvature is lower than the level
of the elastic stress within the filament. Thus, Chang et al. [45] suggest that at the end of the ‘stretching’
event, the filament is near cylindrical and the strain rate has dropped to negligibly small values, such
that there is no flow out of the filament due to stretching. Prior to this point in time, the inception of a
pinch region at either end of the jet minimum is continually retarded by the growing elastic stress within
the developing filament. However, once the flow from the filament decreases significantly, to the now
developed beads at either end, two pinch regions (referred to in Chang et al. [45] as secondary necks)
form immediately at both ends of this now straight filament. On the inception of the pinch points, they
suggest that there is a change in the driving force for flow of fluid from the primary filament to the beads,
driven now not by stretching of the primary filament, but by capillary pressure differences between the
filament and the beads. Previous drainage theories assume a slender filament without pinch regions at the
boundary [37], and hence, fail to consider this change in driving force. This stage is termed the ‘elastic
drainage’ stage that follows the ‘stretching’ stage.
Chang et al.’s jetting simulations predict that once this cylindrical filament was formed, the previous
rapid reduction in the minimum radius of the jet with time during the formation of the primary filament
36 J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59

is significantly retarded. They noted that the minimum radius, corresponding to that of the straight
filament, undergoes an inflection and now decreases at a much slower exponential rate, as seen in recent
experimental works, such as that of Amarouchene et al. [24]. Chang et al. [45] predict that it is this slow
elastic draining that results in the strain rate within the primary filament remaining constant while the
radius decreases exponentially with time, and the elastic stress increases exponentially with time, with a
long elastic relaxation time proportional to three times the relaxation time of the fluid.
However, this primary filament may not exist for long as Chang et al. [45] suggest that instabilities
convected from the bead can relieve the elastic tension at the secondary necks (pinch regions) at each end
of the straight filament during this slow elastic drainage and trigger recoil of the filament free surface at
these pinch points back towards the centre of the primary filament. Secondary filaments are then expected
to form at these pinch regions, starting a new ‘stretching’ event. This process also forms the secondary
or micro-bead between the two primary beads, and such iterated stretching, elastic drainage and recoil
are predicted to occur successively to generate high-generation filaments until finite-extensibility of the
polymers within these new filaments of very small radii allows capillary break-up to then proceed. We
will return to the work of Chang et al. [45] in Section 4, detailing how their outcomes translate to the
observed thread dynamics of low-viscosity, elastic fluids during drop formation.
For the high constant viscosity, elastic fluids, Entov and Hinch [46] were the first to provide an outline
of the various stages associated with the decrease in filament radius of a liquid bridge due to surface
tension-driven break-off, such as observed in the capillary break-up experiments [46,47]. The dynamics
of filament break-up can be divided into three distinct stages: the first is dominated by viscous stresses
attributable to the high viscosity Newtonian solvent; in the second stage the elastic stresses caused by the
extension of the polymer coils in solution dominate; and in the final stage the extension of the polymer
strands and the extensional stresses are assumed to reach a maximum, so the fluid can be regarded as
Newtonian with a higher viscosity and viscous stresses are again dominant. In the elastic stress region,
the filament radius is found to decrease exponentially with time and show an inverse relationship with
the relaxation time of the solution [47], according to
 1/3  
GR0 4 −t
Rmin = exp , (2)
σ 3λc

where λc is the characteristic relaxation time of the fluid, G = ηp /λc the elastic modulus, ηp the polymer
contribution to the solution viscosity, ηp = ηs − ηsolv , and R0 is the filament radius after the separation of
the end-plates in the capillary break-up experiments. When Eq. (2) is written in the dimensionless form
of Rmin /R0 , the prefactor (GR0 /σ) is known as the elasto-capillary number. Anna and McKinley [47] have
shown that within their capillary-thinning experiments with high viscosity, elastic solutions, the filament
necks slowly enough so that all of the elastic stress is carried by the longest relaxation mode of the polymer
and the above equation correlates well with the independently determined Zimm relaxation time of the
solutions. For the low-viscosity, elastic fluids, Christanti and Walker [43] also found the exponential time
dependence to be applicable in the jet break-up experiments. However, the relaxation times obtained
differ significantly from the Zimm relaxation times determined from intrinsic viscosity measurements.
This point will be examined in more detail when discussing our results.
Experimental and analytical treatments of drop formation and the subsequent stretching of the filament
formed has been carried out by Jones and Rees [48] and Jones et al. [49] on several dilute polymer
solutions containing both low and high viscosity Newtonian solvents, whose dimensionless numbers,
J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59 37

Table 1
Bond number, Reynolds number and capillary time scale for the solutions used in various drop formation and liquid bridge
studies
Bo Re tcap (s)

PEO solutions (current study) 0.64 3.4 × 10 3


1.9 × 10−4
FM9 solution [48] 0.26 6.0 × 103 6.7 × 10−5
M1 [49] 0.95 5.3 × 10−3 0.18
Polystyrene solutions [47] 0.60 5.0 × 10−5 1.6–2.2

Bo and Re, as well as tcap are also given in Table 1 (to be compared with the values for our fluids).
In their initial study, Jones and Rees [48] described an axial force balance involving gravity, surface
tension, fluid stress and inertial effects within the growing fluid filament. By incorporating the apparent
extensional viscosity within the viscous stress term in the force balance, they obtained an equation which
provided a value for the apparent extensional viscosity of the solutions during the process of filament
development and elongation up until break-up of the drop. The apparent extensional viscosity results
obtained, and the success of the applied methodology, varied greatly depending on the solutions tested.
For the low-viscosity fluid FM9 (an anti-misting fluid of 0.2 wt.% in kerosene, whose shear viscosity
is about 2 mPa s), the extensional viscosity obtained increased continually during drop elongation up
to approximately 25 000 times ηs . We believe that the analysis is incorrect as the authors have applied
the force balance to the whole of the drop formation process, including the rapid necking prior to the
formation of a distinct cylindrical filament. Due to the very low relaxation time of the fluid, compared to
the time of necking, as can be seen by the tcap (1), it is not expected that elastic contributions due to
extension of polymer chains would occur in the early stage of drop formation. Only when the fluid has
necked significantly to within the region where the primary thread is formed and thinning (e.g. Shi et al.
[9]), are elastic contributions expected to be significant (if at all in the case of these fluids!). However,
for this low-viscosity solution, the experiment of Jones and Rees [48] was terminated by this stage, due
most likely to lack of resolution of the image capturing technique. The application of the above force
balance over the initial drop formation stages is then deemed invalid because the force balance was
applied in the region where there is no extension of polymer chains and, hence, no extensional stresses
are likely. In addition, the inertia of the low-viscosity systems used must surely dominate much of the
drop development, as can be seen from the much larger intrinsic Re shown in Table 1 compared to that
of the higher viscosity fluid M1 of Jones et al. [49]. Within the force balance proposed by Jones and
Rees [48], this high inertial load can only be compensated by a very high extensional stress component
within the filament. If no extension is occurring over such regions of filament development, we quickly
understand why excessively high (unreal) values of extensional viscosity were obtained.
For the more viscous, constant viscosity, elastic fluid M1 (a dilute solution of polyisobutylene in
solvent of polybutene and kerosene whose ηs = 3 Pa s), the results obtained from similar drop formation
experiments show the expected extensional viscosity starting from approximately 3ηs at short times (Jones
et al. [49]). As the filament extended, the extensional viscosity was found to increase continually until the
drop break-off. For the higher viscosity fluid M1, where Re  1 and tcap = O(1), the fluid relaxation time
is expected to be of the same order as the rate of necking and subsequent filament thinning (extension)
process. Hence, during the formation and extension of the primary filament, elastic stress contributions
are observed at all times and inertial contributions of the drop are less significant. The application of the
38 J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59

above force balance is then valid, and could be used to determine the extensional properties of the fluid.
The extensional viscosity obtained for fluid M1 is of the right order of magnitude when compared to the
values obtained by several other techniques (see James and Walters [50]).
Of greater overall importance, however, is the fact that the application of such a force balance equation
assumes that the extension of the polymer coils in solution only occurs within the filament. This assumption
may be valid but only when the extension rates of the fluid within the filament are high enough to cause
extension of the polymers in the surrounding solvent (i.e. a strong flow exists). In an extensional flow
field, the dimensionless Deborah number, De = λε̇, where ε̇ is the extension rate in the process, must
exceed a value of approximately 0.3–0.5 for the polymer molecules to experience sufficient deformation to
transform from the equilibrium, coiled configuration to the stretched state, i.e. the “coil-stretch” transition.
This is the minimum criterion that must be satisfied prior to using this equation to estimate the extensional
stress contributions to the overall force balance during drop formation experiments.
The drop formation results of Amarouchene et al. [24] on dilute and semi-dilute solutions of a high
molecular weight polyethylene oxide (4×106 g/mol) in water are particularly relevant to this present study,
showing that the radius of the filaments formed decrease exponentially with time. From the recordable
variation in the minimum radius data with time of the drop, the authors observed that a critical high
extension rate, determined from the inflection in the radius curve, seemed to occur just prior to the
formation of the filament proper. Thereafter, the rate of extension during filament growth decreases by an
order of magnitude but the filament continued to be supported by the elasticity of the solution. As stated
by these researchers, such a high critical extension rate must exist in order to extend these particular
polymers in solution due to their low relaxation time, even though they did not quantify the applicable
relaxation time, nor provide reasoning as to why and how such rates come to exist.
In this paper, we will attempt to show that for the low-viscosity, elastic fluids, the extending fluid thread
within drop formation experiments with low-viscosity, elastic fluids gives a range of extension rates based
on filament length which are much too low for extension of the polymers to occur within the filament.
It is proposed that the enhanced stability of the primary thread formed from low-viscosity, elastic fluids,
which must exist to counteract the Rayleigh instability, is due to behaviour more akin to a film drainage
phenomenon, not as a result of filament extension.

3. Experimental method

3.1. Image capture and measurement of drop formation dynamics

Images of drop formation were taken using a high-speed drum camera (Fig. 1), as used by Crooks and
Boger [51] and Crooks et al. [52] to study drop impact on dry surfaces. Fluid was fed from a reservoir
via a capillary tube to a nozzle of 4.0 mm o.d., 2.0 mm i.d. Prior to image capture, several drops of fluid
were allowed to form to ensure a constant liquid flowrate.
The drum camera consists of a NIKON F66 camera, fitted with an AF Micro Nikkor 105 mm lens.
Film, which was attached to the perimeter of the 60 cm o.d. film-carrier wheel, was rotated through the
body of the camera. The film-carrier wheel was driven by a servo driver (Baldor P/L) and controlled by
PC (windows software). The light source was provided by a strobe operating at 1 kHz, capturing images
at 1000 frames/s. The strobe light was synchronised to the drop formation stage using a trigger signal
with a time delay, which was released when the previous drop passed a laser beam. This signal was then
J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59 39

Fig. 1. Schematic diagram of the equipment used for drop generation and image capture.

gated by the second pulse generator for a duration of about 120 ms, equivalent to one revolution of the
film-carrier wheel at 500 rpm. The gated signal was then delayed to account for the lag between drops.
The film used was black and white TMAX400 Kodak, and an F-stop of 8 or 11 was chosen to maximise
the depth of field of the images. There is a small variation in the film-carrier wheel speed, but this has little
effect as the strobe frequency determines the rate at which images are captured. The method provides a
high image resolution of 3700 × 2700 dpi. Accurate measurements of filament diameters to 20 ␮m were
made, and filaments as thin as 10 ␮m are visible on the film, although is beyond the resolution of the
image analysis and was not actually measured. However, the ability to detect the thin filaments down to
break-off point is important for determining the exact time at which the primary droplet detaches from
the filament.
To investigate the dynamics of drop formation, the dimensions of the drop as it forms at the nozzle,
elongates and finally detaches are analysed. The dimensions measured (shown in Fig. 2a) are: L, the drop
elongation length, measured from the nozzle tip to the lowest point of attached fluid; R0 , the nozzle outer
radius; and Rmin , the minimum thread or drop radius.

3.2. Test solutions

The low-viscosity, elastic fluids tested were dilute solutions of polyethylene oxide of molecular weight
ranging from 8000 to 1 × 106 g/mol (supplied by Aldrich) in glycerol and water mixture. A constant shear
viscosity of approximately 5.8 mPa s and an equilibrium surface tension of approximately 61.2 mN/m
at 21 ◦ C was maintained by varying the solvent and polymer concentrations whilst keeping below the
polymer crossover concentration. The properties of the solutions were determined by methods used by
40 J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59

Fig. 2. Drop dimensions used for (a) image analysis; and (b) idealised model analysis.

Crooks and Boger [51] and Crooks et al. [52]. These properties, together with the density and Rouse
relaxation time, are given in Table 2. In preparing the solutions, de-ionised water was first warmed to
about 30–40 ◦ C prior to slow polymer addition under agitation to avoid agglomeration. The solutions
were mixed by gently rolling for at least 24 h before adding glycerol, and then rolled for a further 24 h
prior to use in the experiments. All the solutions were used within 1 week of preparation in order to
minimise bacterial degradation. The dynamic surface tension (measured using a Kruss Maximum Bubble
Pressure Tensiometer) and Trouton ratio (measured using a Rheometric RFX extensional rheometer) of the
solutions are shown in Figs. 3 and 4, respectively. The dynamic surface tension of the solutions containing
PEO showed differences at short times but attained the same equilibrium value of approximately 61 mN/m.
This value is somewhat lower than the value for the 50% glycerol solution, but it will be shown later that
this difference does not affect the outcome of the study. The Trouton ratio is the ratio of the extensional
viscosity, ηE , to shear viscosity, ηs , Tr = ηE /␩s . For a Newtonian fluid, the Trouton ratio is a constant
value of 3, but for elastic fluids the ratio can be much higher, as seen in Fig. 4 for the three highest
molecular weight PEO solutions.

Table 2
Fluid properties for test fluids (measured at 21 ◦ C)
PEO PEO Glycerol Static surface Shear viscosity Density Rouse relaxation
(g/mol) (% (w/w)) (%) tension (mN/m) (mPa s) (kg/m3 ) time (s)
– – 50 70.21 5.9 1126.4 –
8000 5 26 60.91 5.98 1070.5 1.10 × 10−7
300000 0.21 36.3 61.35 5.65 1088.5 7.21 × 10−5
600000 0.17 34 60.77 5.77 1082.8 2.04 × 10−4
1000000 0.1 37 61.69 5.63 1090.1 4.87 × 10−4
J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59 41

Fig. 3. Surface tension of the test fluids against surface age at 21 ◦ C.

The values of the Bo, intrinsic Re and tcap of all the test fluids are the same and are given in Table 1.
The table also contains the values for various other solutions, including the low-viscosity FM9 solution
of Jones and Rees [48]; the high viscosity fluid M1 of Jones et al. [49]; and the viscous polystyrene-based
Boger fluids of Anna and McKinley [47]. The PEO solutions and the FM9 solution are both low-viscosity
fluids, so have similar values for Bo, intrinsic Re and tcap , and we expect similar drop formation dynamics
as they are dominated by capillary and inertial forces. Conversely, the solutions of Jones et al. [49] and
Anna and McKinley [47] are strongly controlled by viscosity.

Fig. 4. Trouton ratio against apparent extension rate at 21 ◦ C.


42 J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59

4. Results and discussion

Fig. 5a–d show the time evolution of the drop formation experiments for the 50% glycerol–water
solution, the 300 000, 600 000 and the 1 000 000 molecular weight PEO solutions. The time scale shown
is the “relative time” or time-to-break-off (td − t), which is the time difference between the point at which
the extending drop detaches or undergoes break-off from the nozzle fluid or fluid thread, the detachment
time, td , and the actual time t of interest. Hence, time-to-break-off (td − t) is large at the beginning of the
experiment and approaches zero as the drop approaches the break-off point. Although the drop dynamics,
i.e. drop profile and overall length of the fluid drop during the formation and break-up, of the four solutions
are similar in each of the sequences shown side-by-side, the “relative time” scale is different for each.
Also given in the sequence is the “shifted” time scale, which is the same for all solutions, the significance
of which will be discussed shortly.
Fig. 5a shows the formation of the drop as the Newtonian fluid emerges from the nozzle. Features
include rapid conical necking, the approach to pinching at the base of the neck to form the drop (the
inception of the lower pinch point), the break-off at the drop (at td − t = 0), the formation of the upper
pinch point (at td − t = −1), and finally break-off at the nozzle fluid. This fluid exhibits many of the
features previously observed of low-viscosity Newtonian fluids at low flowrates by a number of researchers
[8–11]. These researchers have investigated glycerol–water mixtures of 0–100% glycerol. They noted that
as the viscosity was increased (i.e. as the glycerol content was increased), these pinch regions at the base
and top of the neck were sustained for longer time (due to viscous pressure damping capillary waves),
creating the primary (cylindrical, in the case of near 100% glycerol) thread. Thereafter, they observed the
formation of a microthread between the primary thread and drop just prior to break-up. This microthread
is not visible for the glycerol–water solution tested here, but is expected to exist albeit for microseconds
time period, based on self-similarity principles.
Fig. 5b–d show clearly that for the 300 000, 600 000 and the 1 000 000 PEO solutions, although the
necking process is similarly followed by the rapid formation of the lower and then upper pinch points,
now it is sustained and there is subsequent elongation of the necked fluid or primary ‘thread’. We can see
the formation of a secondary thread (or microthread) eventually from the lower pinch point in all PEO
solutions, something that is absent from the Newtonian solution of similar shear viscosity. The shape of
the primary ‘thread’ so produced, however, is not cylindrical, as observed for the much higher viscosity
Newtonian fluids, but is composed of an oval volume of fluid effectively crimped at both ends by the
pinch regions. In the case of the 300 000 and 600 000 PEO solutions, the oval volume soon attains a
spherical shape (the secondary drop), while the upper pinch point evolves into an elongating secondary
thread, which is of a larger dimension than the lower secondary thread formed from the lower pinch.
Interestingly, in the case of the 1 000 000 PEO solution, no such secondary drop exists, and instead we
observe a near conical primary thread of decreasing radius joining the upper pinch point to the lower
one, and more akin to the filament found formed in the higher molecular weight PEO solutions by
Amarouchene et al. [24]. The ‘threads’, regardless of whether a secondary drop exists or not, all finally
undergo break-off at the drop from the fluid thread at the lower pinch point (not shown here).
The non-dimensionalised minimum drop radius and length measurements of all the solutions investi-
gated are shown in Figs. 6a and 7a, respectively, as a function of the relative time (td − t). In all cases,
the radius decreases rapidly during the necking process, i.e. down to Rmin /R0 ≈ 0.07, at which point
the drop either detaches from the main body of the fluid, as for the Newtonian fluid and 800 000 PEO
solution, or for the three higher molecular weight PEO solutions, the ‘thread’ formed from the original
J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59 43

Fig. 5. Photo sequence of (a) 50% Glycerol–water; (b) 300 000; (c) 600 000; and (d) 1 000 000 PEO solutions during various
stages of drop development, elongation and break-up. Time is measured in millisecond.
44 J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59

Fig. 6. (a) Dimensionless minimum radius against relative time scale (td − t); and (b) dimensionless minimum radius curves
shifted, using a time scale of relative time minus lag time (td − t − tl ), to create a master curve for the dimensionless minimum
radius with time.
J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59 45

Fig. 7. (a) Dimensionless elongation length against relative time scale (td − t); and (b) dimensionless elongation length curves
shifted, using a time scale of relative time minus lag time (td − t − tl ), to create a master curve for the dimensionless elongation
length with time. The insert shows the dimensionless length increase exponentially with time from onset of pinch point.
46 J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59

Table 3
Lag time tl , relative to the reference glycerol–water solution, and drop detachment length Ld (inclusive of nozzle fluid, filament
and drop) for all solutions
Solution Lag time, tl (ms) Ld (m)

Glycerol–water – 0.0047
8000 PEO 0 0.0047
300000 PEO 11 0.0056
600000 PEO 22 0.0079
1000000 PEO 55 0.0142

neck is evident as the measured Rmin continues to exist, decreasing at a lower rate than the previous
necking process. As the radius of the thread formed soon approaches the resolution of the measuring
technique (≈20 ␮m), only a few points are shown in Fig. 6a post the formation of pinch region. However,
the existence of the thread is still clearly discernable and, hence, the overall length of the drop can be
measured up to the break-off point. The length of the three higher MW solutions is seen to grow very
rapidly once formed until break-off at zero relative time, with the final length increases with increasing
MW of the PEO in solutions, as shown in Fig. 7a. While for the Newtonian solution, the inception of the
lower pinch and subsequent break-off occurs in less that 1 ms of the end of the necking process, for the
solution of 1 000 000 PEO, the filament exists for up to 55 ms after the inception of the lower and then
upper pinch, with the other intermediate molecular weight solutions existing for shorter and shorter times
as the molecular weight decreases.
As seen in Fig. 6a, the approach towards the drop necking process appears similar in all the solutions
studied, with a similar rapid reduction in Rmin . In Figs. 6b and 7b, the minimum radius and filament length
of all the solutions have been shifted along the time axis by an amount called the ‘lag time’, tl , such that
all curves now coincide with the Newtonian glycerol–water solution. This lag time is characteristic of
each solution and their values are given in Table 3. All the curves are now shifted onto a master curve,
with the drop formation for the solutions now all starting at the same arbitrary shifted time (td − t − tl ).
The insert in Fig. 6b shows that for all the solutions studied, regardless of the polymer molecular weight
and, hence, elasticity, the initial necking is very similar to the Newtonian solution until the arbitrary time
(td − t − tl ) → 0. That is, Rmin ∝ (td − t − tl )2/3 to approximately 5 ms of the onset of the pinch, as
expected for a potential flow scaling theory for case of Re 1. This appears to be the case even for the
elastic polymer solutions studied here. Due to lack of resolution of the measuring technique, we cannot
ascertain at this stage the transition from the potential flow to the inertial-viscous thread regime for our
low-viscosity (high Re) Newtonian solution as it approaches break-off. For the elastic solutions after the
inception of the pinch points, however, the behaviours obviously depart from the Newtonian case, as
‘threads’ are formed at the expense of the neck fluid. In this case, the transition is most likely from the
potential flow to an inertial-viscoelastic thread regime, where the pinch is sustained for long period of
time thereafter, depending on some viscoelastic parameters, such as the relaxation time of the solutions.
Fig. 3 shows that the maximum difference in surface tension, equilibrium or dynamic, for any so-
lution used in this investigation is between that of the glycerol–water and 8000 PEO solutions, i.e. of
approximately 8 mN/m. Apart from the difference in the surface tension of these two solutions, there is
no measurable difference in either the shear or extensional rheology, although the 8000 PEO solution
shows a measurable relaxation time, its value is at least two orders of magnitude lower than the other
J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59 47

higher molecular weight PEO solutions. The 8000 PEO solution can, thus, be taken as being virtually
a Newtonian fluid, and this is further confirmed by its drop formation dynamics, which is observed to
be identical to that of the Newtonian glycerol–water solution as shown in Figs. 6a and 7a. Hence, the
effect of the difference in the surface tension is negligible in this case. The observed differences in drop
dynamics shown in Figs. 6 and 7 for all solutions of higher molecular weight PEO are, thus, expected
to be due primarily to elasticity. We recognise, however, that although we observe no difference in drop
dynamics between solutions which differ by 8 mN/m up to the inception of the lower and upper pinch
points, larger differences in dynamic surface tension will most likely have a significant effect. So, whilst
the assumption of constant surface tension made in all equations used here is sufficient for this study, it
may be invalid in further investigations. Research is currently underway in our laboratory to determine
the effects of large variations in dynamic surface tension on drop dynamics.
Thus far we have established that the drop formation dynamics of all the solutions used, including
those with elastic component, follow the same potential flow scaling theory as the drop approaches the
pinch region, as predicted for low-viscosity Newtonian fluids [29]. Once the pinch is formed, the elastic
properties of the solutions of higher molecular weight PEO clearly manifest themselves in sustaining the
pinch region, stopping it from immediate break-off, and instead, causing the formation and subsequent
elongation of the necked fluid into secondary threads and drops. Referring back to Fig. 5, we note that
once the lower and upper pinches are formed, the length scales of the drop and nozzle fluid are both
essentially set for each fluid. The measured increase in the total drop length in Figs. 7a and b is, thus, a
direct reflection of the additional elongation of the necked fluid, or secondary drop and threads formed
in each of the elastic solutions. As shown in the master curve of Fig. 7b, after the inception of pinch
point, i.e. L(t)/R0 > 4.7 at (td − t − tl ) < 0, the thread elongation for the 300 000, 600 000 and 1 000 000
PEO solutions seems to follow the same growth curve, with the final thread length and time-to-break-off
increasing with molecular weight. In addition, the maximum thread elongation rate or ‘terminal’ drop
velocity just prior to break-off (shown by the gradient of the curves in Fig. 7a at (td −t) → 0) also increases
with increasing maximum elongation length, and, hence, molecular weight of the PEO in solution.
From the recent work of Anna and McKinley [47] and Stelter et al. [53], it has been shown that for
constant (high to low) viscosity, elastic fluids undergoing capillary break-up process, the fluid exhibits
an exponential time dependence of the thread diameter, as given in Eq. (2). Other low-viscosity poly-
meric solutions in both jet break-up and drop formation experiments have also been shown to show
this exponential dependency [32,43]. This is the inertial-viscoelastic flow regime alluded to earlier for
flow of viscoelastic solutions approaching break-off. For our solutions, due to lack of resolution of the
thread diameter measurements, it was not possible to check if this exponential dependency also holds.
However, we can at least gain some insight into the expected form of the length of the filament with
time if we consider a related filament stretching experiment (see, for example, Anna et al. [54]), in which
the exponential decay in radius is accompanied by an exponential increase in length of the filament,
if the stretching filament can be assumed to be a cylindrical column throughout the entire elongation
process volume is conserved. Then Eq. (2) can be rewritten in terms of the time evolution of filament
length as
   
σ 2/3 2t
L(t) = L0 exp , (3)
GR0 3λc
where L0 is the initial length of the filament. If we consider this relation in terms of our study of the
drop formation process, which is analogous to an elasto-capillary break-up experiment, but now with
48 J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59

Fig. 8. The Rouse relaxation time can be found from the limiting drop elongation length Ld0 , or alternatively, from the linear
relationship between Rouse relaxation time and the lag time, tl (insert).

significant inertia, we note immediately that such a relationship with the relaxation time is not evident
from the experimental results obtained. Referring to the insert on Fig. 7b, although the thread length after
the onset of lower and upper pinch follows an exponential increase with time, the slopes of the various
elastic solutions are essentially the same and, hence, independent of the relaxation time of the fluids. This
is in contrast to the results of Christanti and Walker [43], who found a dependence of the slope of the
exponential radius–time curves of their fluid jets on the polymer relaxation time. We will discuss this
point further in the Section 4.1.
As seen in Fig. 7a and b, the length of the fluid thread at break-off or detachment and the time-to-break-off
increases as the molecular weight of the polymer, and, hence, the relaxation time of the solution, increases.
The lag time, tl , and the detachment thread length, Ld0 , of the various elastic solutions are plotted as a
function of the Rouse relaxation time, λ, in Fig. 8. The detachment thread length is the total length of
the extended drop at the point of the drop detaching or breaking-off from the thread or necked region,
Ld , relative to that of an equivalent Newtonian fluid of equal shear viscosity and surface tension, LdN ,
i.e. Ld0 = Ld − LdN . From the figure, it is apparent that both quantities show a linear relationship with
relaxation time of the solutions. Therefore, although the growth rate of fluid thread is found to be inde-
pendent of relaxation time, and hence, elasticity, of the solutions, the final length of the thread is not. The
dimensionless detachment length can be written as a function of the relaxation time, with a proportionality
constant relating the fluid surface tension and polymer viscosity as
   
Ld0 λR σ σ
=a =a , (4)
R0 ηp R0 GR0
J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59 49

where a = 0.126 for this set of fluids. It is anticipated that the value of the constant a in the above
linear equation is a function of other parameters which were kept constant in this study, e.g. density
and liquid flowrate. Further investigation is required to provide more detailed understanding of this re-
lationship. However, of greater interest at this point is the fact that the term in brackets is the inverse of
the elasto-capillary number coined by Anna and McKinley [47]. It seems then that, although the rate of
increase of the elastic fluid thread formed is not controlled or affected by the relaxation time of the fluids,
the final thread length at detachment is similarly dependent on the balance of elastic and capillary forces,
i.e. we still have elasto-capillary break-up at the pinch point. Is it possible then that the growth of the
thread and the break-up may be considered as two separable events?

4.1. Dynamics of thread growth

As discussed earlier, Jones and Rees [48] and Christanti and Walker [43] have investigated the drop
formation and break-up of low-viscosity, elastic fluids, by either gravity or forced jetting, and found that
polymer extension, causing an increase in the elastic stress within their fluids, is due to stretching or
elongation of the filament or thread formed. We will attempt to show that for the low-viscosity, elastic
fluids used here and also by Christanti and Walker [43], as well as those of Jones and Rees [48], the
longest relaxation times (Rouse or Zimm relaxation time), calculated based on intrinsic viscosity data,
are too low, and hence, the extensional flow field occurring within the growing fluid thread is insufficient
to invoke the ‘coil-stretch’ transition of the polymer in the solutions.
Prior to the onset of the (upper and lower) pinch regions, the drop formation process for all solutions
tested here is dominated by inertial and capillary forces, with the elastic property of these solutions having
a minimal impact. No such cylindrical filament as observed in high-viscosity, elastic fluids is formed for
our fluids. This is a direct result of the short relaxation times of the fluids used in our study compared to
the time of the primary process of stretching (which can be seen visually in Fig. 5 to occur), as described
previously within our discussion of Chang et al.’s numerical analysis. This produces the similarity in
the behaviour of all solutions, regardless of elasticity, prior to the inception of the pinch regions, as all
the solutions have the same surface tension, density and shear viscosity. Thereafter, the Newtonian 50%
glycerol solution undergoes immediate break-up, yet the lower pinch region continues to exist for the
elastic solutions.
That is, at and beyond the onset of the pinch region, i.e. (td − t − tl ) ≤ 0 in Figs. 6b and 7b, the
elastic stresses become important; this is manifested by the prevention of immediate break-up of the
primary fluid drop and sustaining the necked fluid into long thin threads of various morphologies.
One would, thus, perceive that the elastic contribution comes from the extension of polymers in so-
lution within the thinning and/or elongating filament, as seen in the capillary-thinning and filament-
stretching experiments, and in previous drop formation experiments, utilising high-viscosity, elastic
fluids.
The extension rate within the elongating filament may be calculated on the basis of length or diameter,
as
2 dRmin d ln(Rmin )
ε̇ = − =2 , (5)
Rmin dt dt
1 dl d ln(l)
ε̇ = = , (6)
l dt dt
50 J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59

where l is the length of the filament. By assuming that the filament between the nozzle fluid and drop is
cylindrical and, more importantly, there is conservation of volume during the stretching event, the above
equations can be used interchangeably to calculate the extension rates within the elongating primary
thread of our drop formation experiments. For the high-viscosity, elastic fluid M1 [49], with relaxation
time at least three orders of magnitude higher than the solutions used here, a comparison of the estimated
extension rates by both Eqs. (5) and (6) shows that reasonable agreement was achieved, and thus, the
assumption of a conservation of mass within the elongating filament is reasonable. The extensional
viscosity of the fluid obtained is also comparable to those from other techniques [50]. It, thus, seems
that for the more viscous elastic fluids undergoing drop formation and subsequent primary thread growth
prior to break-up, the relaxation time of the solution is sufficiently high that the coil-stretch transition of
the polymer molecules is invoked by the extensional flow within the stretching filament.
In the case of drop formation experiments with low constant viscosity elastic fluids, in which inertial
contributions are significant throughout the whole process, the situation may be different from that of
Jones et al.’s fluid M1 case. The extension rate of the elongating fluid thread for our low-viscosity elastic
solutions can be calculated from the semi-log plot of the thread length versus time from the onset of pinch
and is found to be of the order of 23 s−1 (by taking the time from onset of the pinch as zero time). In this
current study, even in the case of our PEO solution of the highest molecular weight (i.e. 1 × 106 g/mol),
the fluid relaxation time is of the order of 10−4 s. This suggests that in order that elastic stresses can be
manifested, an extension rate (i.e. filament extension rate) of the order of 103 s−1 or higher must exist for
molecular extension to occur. This is a very high rate and is definitely not invoked within the elongating
filaments of the low-viscosity, elastic fluids used.
Christanti and Walker [42,43] have used a set of near identical low-viscosity, elastic fluids as ours in
their perturbed jetting experiments, in which higher extension rates than possible in the drop formation
under gravity, have been invoked. The relaxation times of the fluids were calculated using λ = 2/3ε̇,
where ε̇ was estimated from the recorded filament radius reduction with time. These relaxation times were
approximately five times higher than the Zimm values calculated based on intrinsic viscosity measure-
ments. If the Deborah number is calculated from the Zimm relaxation times and the recorded extension
rates, a strong flow does not exist under filament extension scenarios. We, thus, suggest that the relaxation
time estimates of Christanti and Walker [43] are high purely due to an incorrect assumption: it is not the
extension rates imposed during filament stretching that are responsible for the initiation of extension of
the polymers in solution, as we shall soon explain.
Thus, although the physical behaviour of the elastic solutions in drop formation convinces us of the
dominance of elastic effects within the time scale of the process, in sustaining the pinch and preventing it
from immediate break-off as compared to Newtonian fluid of similar viscosity, it has been shown that the
extension of polymer molecules cannot be initiated within the elongating filament. The question arises
as to where such extension can occur? There is only one location in which such extension rates can be
sufficiently high for the initiation of extension of the polymers to occur—the pinch regions, as has also
been alluded to inadvertently by Amarouchene et al. [24] and in the analysis of Chang et al. [45]. Although
not stated by Amarouchene et al. [24] but extractable from their minimum radius versus time plot, high
critical extension rates were invoked with the formation of the ‘pinch’ region in their low-viscosity,
elastic fluids. Such a ‘pinch’ region was not seen explicitly within the work of Amarouchene et al. [24]
due to the very high molecular weight PEO used (we have noted similar absence of a well defined ‘pinch’
region with PEO solutions of 2 × 106 g/mol), but is clearly seen within this work for all polymer solutions
investigated.
J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59 51

From close examination of the dynamics during drop development and approach to break-off
(see Fig. 5), we have observed that on the inception of the lower pinch point at the bottom of the
neck, and soon thereafter, the upper pinch point at the mid-region of the original neck, the lifetime of the
necked region of fluid from the nozzle is extended. This necked fluid normally constitutes the secondary
drop in an equivalent Newtonian fluid, as break-up at both of these pinch points is near immediate. In
the case of the low constant viscosity elastic fluids, it now simply forms a reservoir of fluid. In fact,
throughout the whole drop development and elongation, as long as the pinch points exist, the secondary
threads formed from this reservoir of fluid exist. We propose that on the inception of the pinch points
at the boundaries between the necked fluid (or primary thread) and the nozzle fluid and primary drop,
regions of high surface curvature are created and very large flows out of the pinch points are produced,
resulting in significant extension of the polymers at these newly created stagnation zones. The inception
of these pinch points, thus, initiates extension of the polymer coils as high enough rates will exist to incept
a strong flow prior to any filament extension, creating an immediate localised increase in elastic stresses
that subsequently dampen the surface waves and retards break-off.
Once the secondary threads are produced from these pinch points, new pinch regions may form at
the boundaries of the secondary thread and secondary drop, and similarly, between the secondary thread
and primary drop, as suggested by Chang et al. [45]. This allows slow film drainage to occur from the
secondary threads due to the relatively high Laplace pressure differentials, which is further assisted by
the squeeze flow invoked by the falling drop at the bottom of the lower pinch. The difference in Laplace
pressure between the thread fluid and the drop or nozzle fluid is of the order of 2000 Pa (i.e. an order of
magnitude higher than either the drop or nozzle fluid) and increases as the thread radius decreases. With
the increasing flow through the newly formed pinch points comes continual enhanced extension of the
polymers in solution. This increased extensional stress, thus, continues to sustain the pinch point near the
primary drop, secondary drop and the nozzle fluid, by dampening any growing surface waves approaching
the pinch regions. The secondary threads must also be sustained along their length by extension of the
polymers within them, even at the low extension rates calculated from the length change with time. This
behaviour still requires further insight. In any case, break-off at the lower pinch is, thus, possible only
when the polymers have attained a steady-state extension and, hence, its maximum elastic or extensional
stress. Once this maximum stress is attained, surface tension forces, now at much higher values due to
the much smaller filament radius, will regain dominance and eventually result in the break-up of the
primary drop. From the above argument, it is obvious that the profile of filament growth is expected to
be independent of the fluids elastic properties, while the final length of the filament and time at break-up
is proportional to the relaxation time of the polymer solutions. The reasons for the differences in the
morphology of the thread across the range of solutions investigated, including the presence of secondary
beads between two threads for the 300 000 and 600 000 PEO solutions and the absence of one in the case
of the 1 000 000 PEO solution, have not yet been explored. We will consider these now.
Although the time frames over which the three events (stretching, elastic drainage and recoil) occur for
our fluids are not well represented by Chang et al.’s model, and the chosen definitions of the Weissenberg
and capillary numbers are not directly relevant to our fluids, we believe that their mechanistic interpretation
of filament dynamics is very relevant to our experimental results and directly supports our claims of
extension occurring at the pinch regions and not within the thread. In our fluids, there is no initial or
primary ‘stretching’ region, just rapid necking towards impending break-up. Thereafter, for the elastic
PEO solutions, we do not observe a characteristic slow elastic drainage region from a cylindrical primary
filament described by Chang et al. [45], but instead a skip straight into the formation of the lower pinch
52 J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59

which in turn becomes an asymmetric secondary thread at the base of a columnar-like primary thread
(see Fig. 5b–d). Since the neck profile is asymmetric about the minimum during pinching, asymmetric
stretching occurs initially that evolves later into a straight filament [36]. Whilst the first secondary thread
is forming, the upper pinch region forms. Recoil of the filament free surface from both lower and upper
pinch points to the centre of the necked region to form the secondary drop then occurs quite rapidly
(see Fig. 5b and c) in the case of the 300 000 and 600 000 PEO solutions. The transition of the upper
pinch region to the secondary thread is, however, believed to be more representative of a characteristic
recoil of the free surface of the necked fluid towards the centre of the primary thread as described by
Chang et al. [45]. Interestingly, this recoil is not visible in the case of the 1 000 000 PEO solution. From
this point onwards, the development of the secondary threads is proposed to be representative of slow
elastic drainage of the solution from two more (smaller) pinch points at either end of the lower and upper
secondary threads (i.e. at the primary drop, secondary bead and nozzle fluid). This is the start of the
iterated elastic drainage, recoil and stretching process for our fluids, i.e. we bypass the first stretching and
primary filament elastic drainage stages due to the low viscosity of the solvent. The process is continued
until the polymers, flowing through the smallest microthread formed adjacent to the primary drop in some
later elastic drainage event (beyond resolution of the experiment), undergo full extension. This puts a
halt to the increasing elastic stress and allows capillary break-up to occur at the pinch region between the
primary drop and the most recently formed microthread. The change in the length of the overall ‘thread’
is, thus, believed to be unaffected by such dynamics. The localised high extensional stresses within the
pinch regions are expected to contribute little to the axial force balance of the drop system, resulting only
in sustaining the pinch and causing the thread to continue to exist.
However, since the same extension rate profile will be followed by all fluids (a result of all fluids
subscribing to iterated stretching, elastic drainage and recoil), lower molecular weight polymers will,
thus, reach a steady-state extension at an earlier time than a higher molecular weight polymer. Break-off
will then occur earlier and at shorter overall length for the lower molecular weight solutions, as noted
from our experiments. This is why break-off length and time is linearly proportional to the relaxation time
of the fluids and the length growth is not, a result of the dynamics of thread growth, i.e. stretch, elastic
drainage and recoil, being self-similar. The extensional response of each of the fluids within the pinch
regions only affects the timing of the break-off event, due to the resistance imposed by an extensional
response of the free surface boundary within the pinch region. This is the only reason why the filament
continues to exist beyond that observed for the equivalent Newtonian fluid. The elasto-capillary number,
representing the final balance of elastic versus surface tension forces, thus, scales to the maximum length
of the filament at break-off. The break-off length and time, and elongation of the fluid thread or filament
are then effectively separable events. Whether these solutions are investigated within the dripping or
jetting regime, the inertia of the system will still dominate the elongation behaviour of the filaments
between the drops and the extension of the polymers in the pinch regions will ensure that the thread
between primary drops (inclusive of secondary beads) continues to exist.

4.2. Predicting the growth of the thread

The significant contributions of inertia to this particular flow scenario, coupled with the low viscosity
of the solutions, results in a separation between the elongation of the overall thread and the extensional
flow within the pinch regions. It is only within the pinch that increases in extensional stress are primarily
initiated. Neglecting the morphological changes to the necked fluid or primary thread after the inception
J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59 53

of the lower and upper pinch points, the time evolution of the length of this necked fluid volume is, thus,
proposed to be unaffected by the elastic properties of the solutions. This is due to the fact that the drop
falling under gravity deforms the necked region or thread of fluid as if it were purely Newtonian (i.e. a
shear viscosity of 6 mPa s). If this hypothesis is correct, we should be able to predict this behaviour by
simply modifying the axial force balance originally proposed by Jones and Rees [48] to represent such a
separation.
We begin by proposing a vertical force balance within an elongating thread between the nozzle fluid
and the primary drop. For this preliminary analysis, we choose to constrain the geometry of the thread,
assuming that it is cylindrical in shape and has two pinch regions, at the nozzle fluid (top) and primary
drop (bottom), throughout the whole elongation process (refer to Fig. 2b). We recognise that there will be
some fluid flow through both upper and lower pinch regions (i.e. film drainage), but due to the lower pinch
forming before the upper pinch, this pinch region will always have the smallest radius at any time and,
hence, determines the point at which maximum extension occurs within the thread. The lower pinch, thus,
dictates the break-off time of the thread and the maximum length at detachment. Therefore, the following
analysis only considers the flow albeit small through the lower pinch region into the primary drop. The
force balance still consists of contributions from inertia, gravity, surface tension and extensional stress
as given by Jones and Rees [48]. However, from the results of our experimental investigation described
earlier, it is observed that the extension rates are too small for the extension of the polymer to occur
within the growing thread, and consequently the apparent elongational viscosity term is not a variable to
be determined as originally proposed by Jones and Rees [48], but instead taken as being a fixed value of
three times the shear viscosity, ηs :

ml̈ = mg − 2πσr1 − 3ηs πr12 . (7)
l
Here, m represents the mass of the system, including the primary drop and filament, σ the surface
tension and r1 and l represent the radius and length of the filament, respectively. The last term in Eq. (7)
is the elastic stress term of the extension rate, l̇/ l, times the extensional viscosity of the fluid. Following
the method of Jones and Rees [48], a hydrodynamic pressure balance for the system can also be derived.
Neglecting bi-axial extension associated with growth of the droplet and introducing a pressure drop, P,
associated with flow through the lower pinch region into the droplet of radius r2 . (Flow through the upper
pinch point was not considered within this preliminary analysis, and in any case, the lowest pinch region
(i.e. at the primary drop) is expected to see the highest extension rates due to it being the smallest in
size at any time t, and hence, will determine the time at which steady-state extension is achieved.) This
balance takes the form
σ 2σ m
= + 2 (g − l̈) − P. (8)
r1 r2 πr1
For the purpose of this exercise, it is assumed that the pinch takes the form of a tapered contraction
of constant lower radius rn and vertical angle α. The pressure drop P comprises of two contributions,
a regular shear resistance associated with contraction flows and an additional resistance arising from
extensional stress [55] within the pinch, which is expected to be significant in this system due to extension
of the polymer molecules going through the pinch. We have neglected the influence of extensional stress in
the pinch upon the pressure drop in this preliminary analysis, as we cannot accurately measure extension
rates in this region at this stage. In any case, extension of the polymer molecules within the pinch will
54 J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59

only serve to decrease the flow through the contraction during elongation of the filament, not change the
pressure drop P, as this pressure drop is effectively controlled by the difference in Laplace pressures
between the thinning, elongating thread and the drop. The contribution of extensional stress due to the
presence of polymers within the pinch, thus, does not represent a new contribution to the mechanism of
elongation within the system. However, as alluded to earlier, it is critical to predicting the final break-up
event.
A standard correlation for tapered contraction flows [56] is used for the shear component of the pressure
drop, which when combined with Eqs. (7) and (8) yields the pressure balance through the pinch of radius
rn . In this balance, the volumetric flowrate through the lower pinch is expressed as the change in volume
of the droplet of radius r2 according to
 
σ 2σ l̇ 8 ηs 1 1
+ + 3ηs − − r2 ṙ2 = 0. (9)
r1 r2 l 3 tan(α/2) rn3 r13 2
Conservation of volume within the filament–droplet system forms a third equation to close the r1 , r2 , l
system of coupled ordinary differential equations, that is,
4πr22 ṙ2 + πr12 l̇ + 2πr1 ṙ1 l = 0. (10)
We propose that it is these three Eqs. (7), (9) and (10), which include the influence of shear flow
through the pinch region, that can effectively describe elongation of the fluid thread up to break-off. This
analysis obviously cannot predict the development of the thread morphology, but given that we believe
that these dynamics have little effect on the length development of the thread, such an analysis will suffice.
Eqs. (7), (9) and (10) were solved numerically using the package MathematicaTM to a relative accuracy
of 10−6 using an adaptive time-step fourth order Runge–Kutta method up to the time of break-off.
As we cannot accurately measure the extension rates in the pinch, we have not attempted to predict
the dynamics of the pinch region due to the extension of the polymers within it. Such dynamics may
include widening of the pinch region due to the enhanced extensional stress at the free surface boundary,
as predicted and observed by Shaqfeh et al. [57] during coating flows with viscous elastic fluids. The
flow scenario in the particular counter-rotating two-roll coating flow investigated by Shaqfeh et al. [57]
is very similar to the flow within the pinch region as described here, excepting that within the roll coating
scenario, flow of the fluid within the free surface boundary is perpetuated by the rotation of the roll. In
our case, such a flow is perpetuated by the combination of the initial Laplace pressure difference and the
squeeze flow created by the high inertia of the descending drop. This inertial driven elongation of the
filament may counteract such expansion of the free surface boundary (i.e. pinch region) in our situation,
as alluded to by Chang et al. [45] when discussing the formation of an asymmetric extending microthread
upon inception of the pinch region at the base of the primary cylindrical thread in jetting. In any case,
such dynamics require further experimental investigation in order to truly understand the status of the
pinch region during this particular flow scenario. This present analysis is only aimed at capturing the
filament growth dynamics up to the final break-up event.
The inception of the two primary pinch points in our experiments is set to represent time zero in our
calculations. The system was initialised with a necked region (or filament) of length 1.28 mm and radius
110 ␮m (assumed cylindrical), a droplet radius of 1.0 mm and a lower pinch of radius 10 ␮m (all these
quantities are as observed from the image at the onset of the pinch), and an estimated contraction angle
of 10◦ (as per the simulation and experimental work of Shi et al. [9] for similar low-viscosity Newtonian
fluids). This time interval after inception of the pinch points was chosen so as to provide a value for the
J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59 55

Fig. 9. Actual (master curve) and predicted filament length, and radius (insert), with time calculated using Eqs. (7), (9) and (10).
Predictions for variations of 10% in drop radius are included to show the sensitivity of this inertially driven process to drop size.

initial growth rate of the fluid thread, estimated as 87 mm/s from the differentiation of a second order
polynomial fit to the experimental data. The surface tension is set at 0.061 N/m, the density at 1080 kg/m3
and the system mass (drop plus thread) was determined by calculating the initial volume from these
starting parameter values.
Upon assessing the magnitude of the terms in the vertical force balance (Eq. (7)) via simulation and
experimental data, the contributions of the inertial (10−5 N), gravitational (10−5 N) and surface forces
(10−5 N) dominate over the extensional viscous term (10−8 N), which is noted to contribute little to the
dynamics of the system. The pressure balance (Eq. (9)) also exhibits similar behaviour, in which the
contraction pressure drop and filament Laplace pressure are of similar magnitude (103 Pa). They are
larger than the droplet Laplace pressure (102 Pa) and much larger than the viscous term (10−1 Pa). These
results are in contrast with the high viscosity solution problems of Jones et al. [49], where extension of
the filament is a rate-limiting phenomenon, and illustrate the importance of flow from the pinch region
upon the filament evolution. The measured (master curve) and predicted filament lengths for the 1 000 000
PEO solution are shown in Fig. 9; both illustrate the classical filament growth prior to pinch break-off,
where the final predicted filament length is 11.67 mm after 55 ms as compared to the measured length
of 11.34 mm. The shape of the predicted radius reduction with time is also shown as an insert in Fig. 9,
together with the available measured Rmin data for the 1 000 000 PEO solution, showing that it is expected
to be a non-linear decreasing function of time.
As the data is subject to some inaccuracies due to measurement resolution, a limited sensitivity analysis
has been performed on selected parameters used in the prediction. The initial droplet radius exhibits the
greatest influence, as this has a cubic dependence on the majority of the mass (the thread contributes
56 J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59

relatively little), and it in turn modifies the dominant gravitational and inertial terms, whilst the change in
droplet surface force due to changes in the radius is small. The predicted thread length for r2 of +10 and
−10% is also depicted in Fig. 9, illustrating the strong influence of this parameter. Significant accuracy
is, thus, required when measuring the drop radius, with our present accuracy being ±5%. Also of interest
is the influence of the pinch radius, rn , and the taper angle, α, as these are simply estimated and subject to
the largest uncertainties. For pinch radii of 1 and 15 ␮m the predicted final filament lengths are 11.57 and
11.93 mm, respectively, and for taper angles of 5 and 20◦ the predicted lengths are 11.62 and 11.78 mm,
respectively. The variations in the final filament length, thus, demonstrate that the predictions are fairly
insensitive to either the pinch radius or taper angle. Finally, the initial extension rate which has been
estimated by differentiating a fit of the data is also subject to some inaccuracy, and variations of +10 and
−10% give final filament lengths of 12.33 and 11.00 mm, respectively, which are also insignificant. Hence,
the major contributing factor is the drop radius, providing significant mass at the base of the filament,
which effectively overrides any contributions of extensional stress (due to the presence of polymers in
solution) to the axial force balance.
The agreement between prediction and experiment in Fig. 9 is very good considering inaccuracies
associated with the estimate of lower pinch radius and deferral of the influence of extensional stress (axial
and radial) in this region. The extensional stress developed within the pinch will increase the localised
viscosity at the free surface boundary, retarding break-up, and at the same time, slowing flow through the
pinch, which in turn slows the evolution of the thread morphology. It does not, however, substantially
change the mechanics of the thread elongation before break-off, which are essentially governed by the
high inertia of the suspended drop. This is reflected in the analytical prediction.
Such agreement suggests that filament length growth rate is not the relevant measure of the extension
rate which invokes effective elastic stresses within the fluids. If volume was conserved during such fil-
ament growth of these low-viscosity, elastic fluids, the rate determined from the radius reduction would
be equivalent and also too low to provide extension of the polymers in solution. It is, thus, highly
likely that any extension is the result of large internal flows within the threads produced between the
nozzle fluid, secondary drop and the primary drop as a result of high surface tension forces at small
radii and high inertia. Further work is currently underway to assess the relative interaction of inertia
and surface tension as the minimum radius of the thread decreases on approach to the pinch point or
brake-up.

5. Conclusion

The effect of fluid elasticity on drop dynamics during drop formation at a nozzle under gravity has been
investigated using a set of constant low-viscosity, elastic fluids and compared directly with an equivalent
Newtonian glycerol–water solution. All solutions have the same shear viscosity, equilibrium surface
tension, and density, but differ greatly in elasticity. It was found that the early stages of drop formation,
i.e. necking, were the same for all solutions, regardless of the elasticity, up to the formation of a lower
pinch region. Inertial and capillary forces are responsible for perpetuating rapid necking of the drop, with
the rate of radius reduction prior to the formation of the lower pinch point being given by the potential
flow scaling theory predicted by Day et al. [29], with Rmin ∝ τ 2/3 . We conclude that prior to the onset of
this pinch point, there is no contribution of elastic stresses in such fluids, i.e. the solutions behave purely
as inertia-dominated Newtonian fluids.
J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59 57

However, once the Newtonian drop had formed the lower pinch region, it underwent break-up at this
point after less than 1 ms. An upper pinch region then rapidly formed and again underwent break-up in
less than 1 ms, resulting in the formation of a secondary drop. For the elastic solutions, both the lower
and upper pinch regions were formed but were sustained, resulting in a necked region of fluid between
the nozzle fluid and the primary drop. Under the influence of inertia this volume of necked fluid was
quickly converted into thin secondary threads and secondary drops, which varied slightly in morphology
depending on the elasticity of the fluid. The threads remained for up to 55 ms for the solution of highest
molecular weight PEO of 1 × 106 g/mol, before undergoing break-up at the lower pinch point. The length
of the drops at the point of detachment and the associated delay time-to-break-up (when compared to the
Newtonian fluid) increased with increasing fluid elasticity, both showing linear dependence on the Rouse
relaxation time of the fluids. The evolution in length of the thread with time, however, was identical in
shape and form for all polymer solutions tested, regardless of differing elasticity. These results show
conclusively that there is no extension of the polymers via filament stretching within the growing thread,
as the extension rates based on length are far too low to invoke such a response from these fluids. Instead,
the extensional response is believed to be initiated by high extension rates resulting from rapid flow of the
fluid from the pinch regions into the droplet and nozzle fluid. Only then is continued polymer extension
believed to be possible at lower rates within the secondary threads created from these pinch regions.
The development of the morphology of the threads are well described, in part, by the mechanistic
numerical model provided by Chang et al. [45]. We believe that our experimental results show conclusively
that thread elongation in drop formation is composed of iterated stretching, elastic drainage and recoil
events. The inception of and times at which such stages exist are, however, very dependent on the
magnitude of the relaxation time of the solutions under investigation. For example, the low-viscosity,
elastic solutions used in this investigation show no initial stretching or primary filament elastic drainage
events, just immediate recoil upon formation of the pinch points. A secondary stretching event is initiated
on this recoil of the free surface forming secondary threads on either side of a secondary bead. No
secondary bead was observed for the highest molecular weight solution, however. Elastic drainage of
the secondary threads is then believed to occur, producing the noted inflection of the minimum radius
with time.
Drop development, elongation and break-up in such ideal low-viscosity, elastic fluids is, thus, believed
to be separable into two distinct physical contributions to the overall event: elongation of the necked
fluid volume is purely a result of drop inertia, whilst the resistance of the pinch regions to break-off
is perpetuated by the extension of the polymers in solution whilst undergoing film drainage from these
pinch regions, causing an increase in elastic stresses. Prior to reaching steady-state extensional stresses,
the transient increases in stress during extensional flow within the pinch regions combat capillarity, thus,
retarding break-up. But once the steady state in extension has been reached, the fluid thread finally
ruptured or underwent elasto-capillary break-up, as dictated by the balance between the final capillary
and the viscous stresses, at the smallest pinch region nearest the primary drop. The lower the molecular
weight of the polymer in solution, the shorter the time that maximum extension is achieved, and hence,
the shorter the filament length at final break-up. Elongation of the thread is not affected grossly by the
extensional stress increase within the pinch regions, but growing purely as though it was a fluid without
elasticity. This is why the break-off length is linearly dependent on the Rouse relaxation time and the
growth rate of the thread length is not.
Such a claim has been substantiated by the results of our idealised model in which we have treated
these elastic fluids as purely Newtonian but with slow drainage of the thread fluid through the lower pinch
58 J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59

region into the drop. We provide a set of equations describing a modified force balance, which does not
include any extensional viscosity contribution from the polymers in solution due to filament stretching.
On this basis, we are able to correctly predict the magnitude and shape of the length growth curve of the
filament of these low-viscosity, elastic fluids.

Acknowledgements

The non-Newtonian fluid mechanics program at the University of Melbourne has been funded by a
Special Investigator Grant and in part by the Particulate Fluids Processing Centre. We would like to thank
the reviewers for their insightful and detailed comments.

References

[1] C. Guthrie, On drops, Proc. R. Soc. London 13 (1864) 444.


[2] T. Tate, On the magnitude of a drop of liquid formed under different circumstances, Philos. Mag. 27 (1864) 176.
[3] L. Rayleigh, Investigations in capillarity, Philos. Mag. 48 (1899) 321.
[4] J. Eggers, Universal pinching of 3D axisymmetric free-surface flow, Phys. Rev. Lett. 71 (1993) 3458.
[5] J. Eggers, Nonlinear dynamics and break-up of free-surface flows, Rev. Mod. Phys. 69 (1997) 865.
[6] C.A. Bruce, Dependence of ink-jet dynamics on fluid characteristics, IBM J. Res. Dev. 20 (1976) 258.
[7] B.W. Young, Pesticide Formulations and Application Systems, Vol. 5 (13), ASTM STP, 1986.
[8] D.H. Peregrine, G. Shoker, A. Simon, The bifurcation of liquid bridges, J. Fluid Mech. 212 (1990) 25.
[9] X.D. Shi, M.P. Brenner, S.R. Nagel, A cascade of structure in a drop falling from a faucet, Science 265 (1994) 219.
[10] X. Zhang, O.A. Basaran, An experimental study of dynamics of drop formation, Phys. Fluids 7 (1995) 1184.
[11] E.D. Wilkes, S.D. Phillips, O.A. Basaran, Computational and experimental analysis of dynamics of drop formation, Phys.
Fluids 11 (1999) 3577.
[12] E.V.L. Narasinga Rao, R. Kumar, N.R. Kuloor, Drop formation in liquid–liquid systems, Chem. Eng. Sci. 21 (1966) 867.
[13] H.E. Edgerton, E.A. Hauser, W.B. Tucker, Studies in drop formation as revealed by the high-speed camera, J. Phys. Chem.
41 (1937) 1017.
[14] E.D. Wilkes, O.A. Basaran, Forced oscillations of pendant (sessile) drops, Phys. Fluids 9 (1997) 1512.
[15] M.P. Brenner, J. Eggers, K. Joseph, S.R. Nagel, X.D. Shi, Break-down of scaling in droplet fission at high Reynolds numbers,
Phys. Fluids 9 (1997) 1573.
[16] R. Badie, D.F. de Lange, Mechanism of drop constriction in an drop-on-demand ink-jet system, Proc. R. Soc. London A
453 (1997) 2573.
[17] D.M. Henderson, W.G. Pritchard, L.B. Smolka, On the pinch-off of a pendant drop of viscous fluid, Phys. Fluids 9 (1997)
3188.
[18] S.E. Bechtel, M.G. Forest, K.J. Lin, Closure to all order to a 1D models for slender viscoelastic jets: an integral theory for
axisymmetric, torsionless flow, Stability Appl. Anal. Cont. Media 2 (1992) 59.
[19] J. Eggers, T.F. Dupont, Drop formation in a one-dimensional approximation of the Navier–Stokes equation, J. Fluid Mech.
262 (1994) 205.
[20] D.T. Papageorgiou, On the break-up of viscous liquid thread, Phys. Fluids 7 (1995) 1529.
[21] R.M.S.M. Schulkes, The evolution and bifurcation of a pendant drop, J. Fluid Mech. 278 (1994) 83.
[22] D.F. Zhang, H.A. Stone, Drop formation in viscous flows at a vertical capillary tube, Phys. Fluids 9 (1997) 2234.
[23] D. Gueyffier, J. Li, A. Nadim, R. Scardovelli, S. Zaleski, Volume-of-fluid interface tracking with smoothed surface stress
methods for three-dimensional flows, J. Comput. Phys. 152 (1999) 423.
[24] Y. Amarouchene, D. Bonn, J. Meunier, H. Kellay, Inhibition of the finite-time singularity during droplet fission of a polymeric
fluid, Phys. Rev. Lett. 86 (2001) 3558.
[25] D.V. Boger, A highly elastic constant-viscosity fluid, J. Non-Newt. Fluid Mech. 3 (1977/1978) 87.
[26] T.A. Kowalewski, On the separation of droplets from a liquid jet, Fluid Dyn. Res. 17 (1996) 121.
J.J. Cooper-White et al. / J. Non-Newtonian Fluid Mech. 106 (2002) 29–59 59

[27] J.B. Keller, M.J. Miksis, Surface tension driven flows, SIAM J. Appl. Math. 43 (1983) 268.
[28] L. Ting, J.B. Keller, Slender jets and thin sheets with surface tension, SIAM J. Appl. Math. 50 (1990) 1533.
[29] R.F. Day, E.J. Hinch, J.R. Lister, Self-similar capillary pinchoff of an inviscid fluid, Phys. Rev. Lett. 80 (1998) 704.
[30] P.K. Notz, A.U. Chen, O.A. Basaran, Satellite drops: unexpected dynamics and change of scaling during pinch-off, Phys.
Fluids 13 (2001) 549.
[31] J.R Lister, H.A. Stone, Capillary break-up of a viscous thread surrounded by another viscous fluid, Phys. Fluids 10 (1998)
2758.
[32] O.E. Yildirim, O.A. Basaran, Deformation and break-up of stretching bridges of Newtonian and shear-thinning liquids:
comparison of one- and two-dimensional models, Chem. Eng. Sci. 56 (2001) 211.
[33] G.H. McKinley, A. Tripathi, How to extract the Newtonian viscosity from capillary break-up measurements in a filament
rheometer, J. Rheol. 44 (2000) 653.
[34] A. Rothert, R. Richter, I. Rehberg, Transition from symmetric to asymmetric scaling function before drop pinch-off, Phys.
Rev. Lett. 87 (2001) 084501-1.
[35] M. Renardy, A quasilinear parabolic equation describing the elongation of thin filaments of polymeric liquids, SIAM J.
Math. Anal. 13 (1982) 226.
[36] M. Renardy, Some comments on the surface-tension driven break-up (or the lack of it) of viscoelastic jets, J. Non-Newt.
Fluid. Mech. 51 (1994) 97.
[37] M. Renardy, A numerical study of the asymptotic evolution and break-up of Newtonian and viscoelastic jets, J. Non-Newt.
Fluid. Mech. 59 (1995) 267.
[38] M.G. Forest, Q. Wang, Change-of-type behavior in viscoelastic slender jet models, J. Theor. Comp. Fluid Dyn. 2 (1990) 1.
[39] M.G. Forest, Q. Wang, Dynamics of slender viscoelastic free jets, SIAM J. Appl. Math. 54 (1994) 996.
[40] A.L., Yarin, Free Liquid Jets and Films: Hydrodynamics and Rheology, Wiley, New York, 1993.
[41] P.R. Mun, J.A. Byars, D.V. Boger, The effect of polymer concentration and molecular weight on the break-up of laminar
capillary jets, J. Non-Newt. Fluid Mech. 74 (1998) 285.
[42] Y. Christanti, L.M. Walker, Surface tension driven jet break up of strain-hardening polymer solutions, J. Non-Newt. Fluid
Mech. 100 (2001) 9.
[43] Y. Christanti, L.M. Walker, Effect of fluid relaxation time of dilute polymer solutions on jet break up due to a forced
disturbance, J. Rheol. 46 (2002) 733.
[44] M.J. Goldin, J. Yerushalmi, R. Pfeffer, R. Shinnar, Break-up of laminar capillary jet of a viscoelastic fluid, J. Fluid Mech.
38 (1969) 689.
[45] H.-C. Chang, E.A. Demekhin, E. Kalaidin, Iterated stretching of viscoelastic jets, Phys. Fluids 11 (1999) 1717.
[46] V.M. Entov, E.J. Hinch, Effect of a spectrum of relaxation times on the capillary thinning of a filament of elastic liquid, J.
Non-Newt. Fluid Mech. 72 (1997) 31.
[47] S.L. Anna, G.H. McKinley, Elasto-capillary thinning and break-up of model elastic liquids, J. Rheol. 45 (2001) 115.
[48] W.M. Jones, I.J. Rees, The stringiness of dilute polymer solutions, J. Non-Newt. Fluid Mech. 11 (1982) 257.
[49] W.M. Jones, N.E. Hudson, J. Ferguson, The extensional properties of M1 obtained from the stretching of a filament by a
falling pendant drop, J. Non-Newt. Fluid Mech. 35 (1990) 263.
[50] D. James, K. Walters, in: A.A. Collyer (Ed.), Techniques in Rheological Measurement, Elsevier, Amsterdam, 1993.
[51] R.C. Crooks, D.V. Boger, The influence of fluid elasticity on drops impacting on dry surfaces, J. Rheol. 44 (4) (2000) 973.
[52] R.C. Crooks, J. Cooper-White, D.V. Boger, The role of dynamic surface tension and elasticity on the dynamics of drop
impact, Chem. Eng. Sci. 56 (2001) 5575.
[53] M. Stelter, G. Brenn, A.L. Yarin, R.P. Singh, F. Durst, Validation and application of a novel elongation device for polymer
solutions, J. Rheol. 44 (2000) 595.
[54] S.L. Anna, G.H. McKinley, D.A. Nguyen, T. Sridhar, S.J. Muller, J. Huang, D.F. James, An interlaboratory comparison of
measurements from filament-stretching rheometers using common test fluids, J. Rheol. 45 (2001) 83.
[55] M. Yao, S.H. Spiegelberg, G.H. McKinley, Dynamics of weakly strain-hardening fluids in filament stretching devices, J.
Non-Newt. Fluid Mech. 89 (2000) 1.
[56] R.H. Perry, D. Green, Chemical Engineers’ Handbook, 6th Edition, McGraw-Hill, Singapore, 1984, pp. 5–36.
[57] E.S.G. Shaqfeh, A.G. Lee, G. Bhatara, B. Khomami, The free surface displacement and coating of a polymeric solution—a
combined finite element and experimental study, in: Proceedings of the 3rd Pacific Rim Conference on Rheology, Vancouver,
BC, Canada, 2001, p. 46.

You might also like