ArticlePDF Available

On the depinning of a drop of partially wetting liquid on a rotating cylinder

Authors:

Abstract and Figures

We discuss the analogy of the behaviour of films and drops of liquid on a rotating horizontal cylinder on the one hand and substrates with regular one-dimensional wettability patterns on the other hand. Based on the similarity between the respective governing long-wave equations we show that a drop of partially wetting liquid on a rotating cylinder undergoes a depinning transition when the rotation speed is increased. The transition occurs via a sniper bifurcation as in a recently described scenario for drops depinning on heterogeneous substrates.
Content may be subject to copyright.
arXiv:1010.2920v1 [physics.flu-dyn] 14 Oct 2010
On the depinning of a drop of partially wetting liquid on a
rotating cylinder
Uwe Thiele
Department of Mathematical Sciences,
Loughborough University, Leicestershire LE11 3TU, UK
We discuss the analogy of the behaviour of films and drops of liquid on a rotating
horizontal cylinder on the one hand and substrates with regular one-dimensional
wettability patterns on the other hand. Based on the similarity between the
respective governing long-wave equations we show that a drop of partially wetting
liquid on a rotating cylinder undergoes a depinning transition when the rotation
speed is increased. The transition occurs via a sniper bifurcation as in a recently
described scenario for drops depinning on heterogeneous substrates.
The manuscript was accepted by the J. Fluid Mech. in October
2010 and will be published in the following months.
I. INTRODUCTION
Motivated by the question how much honey can be kept on a breakfast knife by rotating
the knife about its long axis, Moffatt studied the destabilisation of a film on the outside of a
rotating horizontal cylinder and the subsequent development of regular patterns of azimuthal
rings [29]. He also gives a lubrication equation that incorporates gravity and viscosity effects
but neglects surface tension and inertia. It is based on a model by [40] which includes
surface tension. Related questions had been studied before by [56] employing a rotating
horizontal cylinder which is at the bottom immersed in a liquid bath. i.e., in contrast to
[29] the amount of liquid on the cylinder is not an independent control parameter. The
related case of a liquid film covering the inner wall of a horizontal rotating cylinder was
first studied by [37] who discussed experiments performed with water employing an inviscid
model incorporating gravitational and centrifugal forces.
Since the early experimental results engineers and scientists alike have studied flows of
2
free-surface films of liquid on the inner or outer wall of resting or rotating horizontal cylinders
in a number of different settings. The studied systems are of high importance for several
coating and printing processes where rotating cylinders transport the coating material in the
form of liquid films. In the present contribution we will discuss the analogy between film flow
and drop motion on a rotating horizontal cylinder on the one hand and on heterogeneous
substrates with regular wettability patterns on the other hand. First, however, we will give
a short account of the seemingly disconnected fields.
The theoretical analyses given by [56] and [37] for the flow on/in a rotating cylinder
were based on linear stability considerations based on the full hydrodynamic equations for
momentum transport, i.e., the Navier-Stokes equations. Their studies were extended later
on, e.g., by [35] who considered inviscid film flow on the in- and outside of the cylinder
in a unified manner, and [45] whose analysis includes viscous film flow at small and large
Reynolds numbers.
However, most analyses of the non-linear behaviour are based on a long-wave or lubri-
cation approximation [33] valid for the case where the thickness of the liquid film h(θ, t) is
small as compared to the radius of the cylinder R. The resulting evolution equation for the
film thickness profile
th=θh3[˜α∂θ(θθ h+h)˜γcos(θ)] + h(1)
was first given by [40] and used by [29] (without capillarity effects, i.e., ˜α= 0). In the
dimensionless Eq. (1), ˜α= 3ε3σ/Rω ˜ηand ˜γ=ε2ρgR/3ω˜ηare the scaled dimensionless
surface-tension and gravity parameter, respectively [18]. It is important to note that both
depend on the angular velocity of the rotation ω, the dynamic viscosity ˜η, the cylinder
radius R, and the ratio εof mean film thickness ¯
hand R. The remaining material constants
are surface tension σand density ρ, and gis the gravitational acceleration. The angle θ
determines the position on the cylinder surface and is measured anti-clockwise starting at
the horizontal position on the right. The force ˜γcos(θ) corresponds to the component of
gravity parallel to the cylinder surface.
The long-wave equation (1) including surface tension effects has been studied in detail by,
e.g., [18, 41] and [23]. Note that for the parameter ranges where Eq. (1) applies, it is valid
without difference for films on the outside and inside of the cylinder (cf., e.g., [2, 18, 31]).
A preliminary comparison of the long-wave approach to results obtained with the Stokes
3
equation is given by [36]. Various extensions of Eq. (1) were developed. See, in particular,
[1] for a discussion of hydrostatic effects. Other higher order terms related to gravity, inertial
and centrifugal effects have been included [2, 8, 14, 15, 19, 25, 31, 32]. In particular, [31] and
[25] present a systematic derivation of several such higher order models and give also detailed
comparisons to a number of previous studies as e.g., [32] , [2] and [8]. The film stability
and non-linear evolution of a film on the outside of a non-isothermal horizontal cylinder was
considered without [43] and with [13] rotation. Most of the aforementioned studies consider
two-dimensional situations. The fully three-dimensionional problem is studied for the cases
without [55] and with [15, 30] rotation.
The second system that we would like to discuss are liquid drops and films on heteroge-
neous solid substrates as studied experimentally, e.g. by [46] and [42]. Although in principle
the heterogeneity may result from substrate topography or chemical heterogeneities we focus
here on the case of a smooth flat substrate with chemical heterogeneities, i.e., we assume
the substrate wettability depends on position. Horizontal heterogeneous substrates are often
considered in connection with micro-patterning of soft matter films via thin film dewetting
[3, 11, 26, 44]. Another application are free-surface liquid channels in microfluidics [16].
Often one employs thin film evolution equations obtained through a long-wave approxima-
tion to study the dynamics of dewetting for an initially flat film or to analyse steady drop
and ridge solutions and their stability on such patterned substrates (see, e.g., [24] and [49]).
Static structures may also be studied using a variational approach [10, 27].
More recently a model system was introduced to study the dynamics of drops on het-
erogeneous inclined substrates. The models apply also to such drops under other driving
forces along the substrate [52]. The corresponding dimensionless evolution equation for the
one-dimensional film thickness profile (i.e., describing a two-dimensional drop) is of the form
th=xh3[x(xxh+ Π(h, x))] + ˜µ,(2)
where ˜µ(h) represents the lateral driving force (µh3in the case of an inclined substrate, where
µis the inclination angle). The position-dependent disjoining pressure Π(h, x) models the
heterogeneous wettability of the substrate. One notices at once that Eqs. (2) and (1) are
rather similar: Identifying x=θ, ˜µ(h) = hand Π(h, x) = ˜αh ˜γsin(x), Eq. (1) may for
˜α= 1 be seen as a special case of Eq. (2). Here, we use the analogy to study drops or films
on a rotating cylinder. In particular, we show that drops on a rotating cylinder may show
4
rather involved depinning dynamics analogeous to the behaviour of drops on heterogeneous
substrates [5, 51, 52].
To illustrate the behaviour we choose the case of drops of a partially wetting liquid on
a substrate with a periodic array of localised hydrophobic wettability defects [52]. On a
horizontal substrate one finds steady drops that are positioned exactly between adjacent
defects, i.e., the drops stay on the most wettable part. At a small lateral driving force µthe
x→ −xsymmetry of the system is broken. However, the drops do not slide continuously
along the substrate as they do in the case of a substrate without defects [53]. Their advancing
contact line is blocked by the next hydrophobic defect, i.e., one finds steady pinned drops
at an upstream position close to the defect. As the driving force is increased, the drops are
pressed further against the defect. As a result the steady drops steepen. At a finite critical
value µcthe driving force allows the drop to overcome the adverse wettability gradient
and the drop depins from the defect and begins to slide. In a large part of the parameter
space spanned by the wettability contrast and the drop size the point of depinning at µc
corresponds to a Saddle-Node Infinite PERiod (sniper) bifurcation. Beyond µcthe drops
do not slide continuously down the incline. They rather perform a stick-slip motion from
defect to defect [51, 52]. However, in another region of the parameter space the depinning
can occur via a Hopf bifurcation, i.e., with a finite frequency. In particular, this happens in
the case of rather thick wetting layers.
The aim of the present contribution is to elucidate under which conditions a similarly
intricate ‘depinning’ behaviour can be found for a pendant drop underneath a horizontal
cylinder when increasing the speed of rotation. In the following section II we will discuss
our scaling and introduce a long-wave model for a film of partially wetting liquid on a
rotating cylinder. It allows (even without gravitation) for the coexistence of drops and a
wetting layer. We discuss several physical situations it can be applied to. Next, section
III presents steady state solutions depending on the non-dimensional rotation velocity for
several thicknesses of the wetting layer and discusses the depinning scenario in the case of
the partially wetting liquid. Finally, we conclude and give an outlook in Section IV.
5
II. MODEL
Before we introduce our long-wave model we discuss our scales. In the non-dimensionali-
sation used in most works studying Eq. (1) the angular velocity of the cylinder rotation ω
is not reflected in a single dimensionless parameter but both dimensionless parameters are
inversely proportional to ω. Furthermore, the time scale depends on ω[18, 40]. This and
a further re-scaling of the equation for steady drop and film states, that absorbs the flow
rate into both dimensionless parameters and also into the thickness scale, implies that the
existence of two steady solutions for identical parameters (thin capillary film and pendant
drop with through-flow), shown analytically by [41] and numerically by [23], actually refers
to solutions of different liquid volume.
Such a scaling is not convenient for our study as we would like to investigate the change in
system behaviour for fixed liquid volume when increasing the rotation speed of the cylinder
from zero, i.e., ωshould only enter a single dimensionless parameter that will be our main
control parameter. A scaling fit for our purpose is the one based on the scales of the gravity-
driven drainage flow employed by [14].
Note also that in particular [23] and to a lesser extent [55] represent some of their solutions
in a way that misrepresents their findings and, in general, the concept of an equation in long-
wave approximation. Whenever one shows the solutions as a thickness profile on a cylinder
one has to pick a particular radius Rand smallness ratio ǫfor the representation (cf. Figs. 3,
5 and 9 below). This is a rather arbitrary choice as the equation in long-wave approximation
is strictly valid only in the limit ǫ0, ie. it is always only an approximation to a real
physical situation where ǫis finite.
In the present work we focus on solutions without axial variation, i.e., we consider the
two-dimensional physical situation depicted in Fig. 1. Following [14] we use ¯
h,R,ρg¯
h2,
ερg¯
h2, 3η¯
hρg and ρg¯
hto scale the coordinate zorthogonal to the cylinder surface,
the coordinate x=θR along the surface, the two velocity components vxand vz, time, and
pressure, respectively. The ratio of zand xscale is the smallness parameter ε=¯
h/R, i.e.,
the film height is scaled by ¯
h=εR. Here the angle θis defined in clockwise direction starting
at the upper vertical position, i.e., a pendant droplet underneath the resting horizontal
cylinder has its centre of mass at θ=π.
The evolution equation is derived from the Navier-Stokes equations with no-slip boundary
6
θ
ω
g
h
R
FIG. 1: Sketch of a drop of partially wetting liquid coexisting with a thick wetting layer on a
rotating cylinder.
conditions at the (rotating) cylinder, and tangent- and normal force equilibria at the free
surface using the long-wave approximation [14, 33, 40, 43]. Note that here we furthermore
assume that the ratio ¯
h/R and the physical equilibrium contact angle are both of O(ε)
(cf. Section IV). This gives
τh=θnh3θhBo1(θθh+h)cos(θ) + e
Π(h)i+e
ho,(3)
where τis the non-dimensional time, and
Bo = R3ρg
¯
and e
Ω = ηωR
ρ¯
h2g0 (4)
are an effective Bond number and a rotation number for a clock-wise rotation, respectively.
The latter corresponds to the ratio of the rotation velocity at the cylinder surface and the
drainage velocity. The function e
Π(h) is a non-dimensional disjoining pressure that accounts
for the wettability of the liquid film on the cylinder surface [9, 17, 21].
To make the analogy with the depinning droplet on a heterogeneous substrate more
obvious we introduce another timescale t=τ/Bo, disjoining pressure Π = Bo e
Π and rotation
number Ω = Bo e
Ω and obtain
th=θh3θ[θθh+hBo cos(θ) + Π(h)] + Ωh,(5)
with
Ω = ηωR
ε3σ.(6)
7
Defining the derivative
hf=Π(h)h+ Bo cos(θ) (7)
of a ‘local free energy’ f(h, θ) and identifying ˜µ= Ωh, Eq. (5) takes the form of Eq. (2)
and the Bond number takes the role of a ‘heterogeneity strength’. Note that the scaling
used is well adapted for discussing independently the influences of rotation and gravity in
the presence of capillarity. However, it does not allow for a study of the limit of vanishing
surface tension.
As disjoining pressure we choose here the combination of a long- and a short-range power
law Π(h) = Ha/h3(1 b/h3), where Ha is a dimensionless Hamacker constant. For b > 0
one has a precursor film or wetting layer thickness h0=b1/3[38, 39, 48]. This is valid only
for Ha<0, i.e., when the long-range interaction is destabilising. This corresponds to the
case of a partially wetting liquid. For Ha>0, a rather thick wetting layer can be stabilised
and the b/h6term is of no further relevance and could just as well be dropped.
Note that hfmay contain other terms beside the ones in Eq. (7). For a heated or cooled
cylinder a term (3/2) BoBi Ma log[h/(1 + Bi h)] + 1/(1 + Bi h)] would enter where Bi is the
Biot number and Ma an effective Marangoni number [34, 43, 50]. If the rotating cylinder
forms the inner electrode of a cylindrical capacitor, a term Vo/[h+ (dh)ǫ]2has to be
included for a dielectric liquid and DC voltage of non-dimensional strength Vo (ǫis the
electric permittivity and dthe distance between the two cylinders (see appendix of [52] and
[22]). The similarity in the behaviour of depinning drops under the influence of different
physical effects in the case that the effective pressure terms ‘look similar’ is discussed by
[52].
Although here we do not consider such other physical effects explicitly, we will explore
the effect of corresponding thick wetting layers by adapting the parameters of our disjoining
pressure accordingly. A thick wetting layer that coexists with droplets might, for instance,
result from the interplay of a destabilising thermal or electrical effect and a stabilising long-
range van der Waals interaction. We obtain adequate values of Ha and bfor our present
Π(h) by relating them to the (non-dimensional) thickness of the wetting layer h0and a static
macroscopic contact angle β0by [see, e.g., 17]
b=h3
0and Ha = 5
3β2
0h2
0.(8)
Note, that β0is the angle in long-wave scaling, i.e., a small physical equilibrium contact
8
angle βeq =εβ0corresponds to a long-wave contact angle β0of O(1). In the limit of a non-
rotating cylinder (i.e. Ω = 0) and assuming the disjoining pressure corresponds to complete
wetting (Ha>0, 1/h6term may be dropped), our equation (5) corresponds to the one given
by [43] for the isothermal case. There, however, first a viscous scaling is used, which they
transform in a second step – a rescaling of time and therefore velocities with the Galileo
number – into the ‘drainage scaling’ used here. Our equation corresponds also to that of
[18] when the disjoining pressure is added to their equation.
To analyse the system behaviour we determine in the following steady-state solutions
that correspond, e.g., to pendant droplets. The behaviour beyond the depinning thresh-
old is analysed using time-stepping algorithms. As we restrict ourselves to the two-
dimensional physical situation an explicit scheme for stiff equations suffices for the latter.
The steady-state solutions are obtained using the continuation techniques of the package
AUTO [12]. The steady and time-periodic solutions are characterized by their L2norm
||δh|| ≡ q(1/2π)R2π
0(h(θ)1)2and time-averaged L2norm
||δh|| ≡ q(1/2π T )RT
0R2π
0(h(θ)1)2dθdt, respectively. Tis the time period.
Note finally that Eq. (5) is the result of our choices for the relative order of magnitude of
the involved dimensionless numbers that is based on the choice of physical effects that shall
be discussed (cf. [33]). Once this is accepted, equation (5) is to O(1) asymptotically correct.
Its form as a conservation law implies that R2π
0hdθ = 2π. To O(1) this corresponds to mass
conservation. For a discussion of higher order corrections to this picture see [25].
III. RESULTS
A. Parameters
We base our estimate for realistic Bond and rotation numbers on two liquids typically
studied in the literature: (i) water at 25C as in [43] and a silicone oil with σ= 0.021N/m,
η= 1kg/ms and ρ= 1200kg/m3as in [14]. As experimentally feasible configurations we
assume angular rotation velocities ω= 0.1...10s1, cylinder of radii R= 103...102m
and smallness ratios ε= 0.01 ...0.1. It is only for the purpose of illustration of steady
droplets on a cylinder that we later use ε= 0.1 in selected figures.
For ω= 1s1and a smallness ratio ε= 0.1 we find for a cylinder of radius R= 102m for
9
material (i) Bo= 136 and Ω = 0.14, whereas material (ii) gives Bo= 561 and Ω = 476. For a
smaller cylinder radius of R= 103m one has (i) Bo= 1.36 and Ω = 0.014, and (ii) Bo= 5.61
and Ω = 47.6. Decreasing the smallness ratio to ε= 102increases all Bond numbers by
a factor 10 and all rotation numbers by a factor 103. Focusing on cylinder diameters in
the millimetre range, we mainly investigate Bo10. Note that experiments may also be
performed with larger cylinders. To keep the Bond number in the interesting range one could
decrease the relevant density by replacing the ambient gas by a second (immiscible) liquid.
A feasible experimental set-up could be a horizontal Taylor-Couette apparatus with rotating
inner cylinder. The theoretical framework would also need to be amended – replacing the
one-layer theory [Eq. (5)] by a closed two-layer model similar to [28].
Note, that silicone oils may have viscosities 2 orders of magnitude smaller or larger than
the chosen value and one is able to vary the velocity of rotation in a wide range. This
implies that there is some flexibility in the choice of Ω. We will find that for Bo= O(1) the
interesting range for the rotation number is Ω = O(1). For instance, with Bo= 1 one finds
that depinning occurs at Ω = 1.68.
As the present work aims at establishing the qualitative correspondence between the
behaviour of a depinning drop on a heterogeneous substrate and the one of a droplet on a
rotating cylinder we restrict our results to one partially wetting case (choosing h0= 0.1 and
β0= 2) and the completely wetting case (β0= 0). For comparison, selected results are also
shown for the contact angle β0= 1.
Note that here the scaling and therefore the discussion of the smallness parameter ǫis
based on ǫ=¯
h/R, i.e., we use the ratio of mean film thickness and cylinder radius. In the
case of pendant drop solutions one also has to keep the maximal film height hmax in mind
when discussing ǫas the scaling becomes questionable when hmax ¯
h. This is, however,
normally not the case. In the present work the ratio hmax/¯
his always below 3. Although
an inspection of the figures in [23] shows a ratio hmax /¯
hthat is not much larger, one notices
that there ǫitself (based on ¯
h) seems to be larger than one (see remark above Eq. (II)).
The same applies to Fig. 3 of [55]. There exists no such problem in the region where only
a thin wetting layer of thickness hmin covers the cylinder even if hmin ¯
h. Actually, the
long-wave approximation gets better there. In an approach based on matched asymptotics
one could further simplify the governing equation in this region (cf. [7]). However, such an
approach is not taken here as it is of limited use when considering parameter regions that
10
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
Bo
0
0.2
0.4
0.6
0.8
1
||δh||
1.0
2.0
β0
(i)
(ii)
(iii)
FIG. 2: Solutions on a horizontal cylinder without rotation (Ω = 0) are characterised by their
L2norm as a function of the Bond number for two different equilibrium contact angles β0and
h0= 0.1, ¯
h= 1.0.
show qualitative changes in solution behaviour, e.g., close to the sniper bifurcation discussed
below.
B. Partially wetting case
First, we consider the case of a horizontal cylinder without rotation (Ω = 0) and determine
steady drop and film solutions. Employing the Bond number as control parameter we
obtain families of steady state solutions for two selected values of the equilibrium contact
angle β0. Inspecting Fig. 2 one notices that above a critical value of the Bond number Boc
there exists only a single solution. It corresponds to a symmetric pendant drop located
underneath the cylinder. An example for Bo= 1 and β0= 2.0 is given as solid line in Fig. 3.
However, two additional steady solutions exist below Boc. One of them corresponds to an
unstable symmetric drop sitting on top of the cylinder (dotted line in Fig. 3). The other
one corresponds to an unstable solution that has two minima – a deep one underneath the
cylinder and a shallow one on top of it (dashed line in Fig. 3).
We find that Bocbecomes smaller with decreasing contact angle β0, i.e., the range of Bo
11
0
1
2
3
h(θ)
0 3.14 6.28
θ
-1
0
1
cos θ
(i)
(ii)
(iii)
a
b-10 0 10
x
-10
0
10
y
c
FIG. 3: Steady film thickness profiles on a horizontal cylinder without rotation (Ω = 0) for Bo= 1
and β0= 2.0 (corresponding to marks (i) to (iii) in Fig. 2). The remaining parameters are as in
Fig. 2. Panel (a) gives the profiles as h(θ), (b) gives the space-dependent part of hf(Eq. (7)),
and panel (c) shows the profiles on a cylinder. Note that for illustration purposes we have (rather
arbitrarily) assumed a radius R= 10, i.e. ε= 0.1 (cf. discussion in Section I II A). Line styles in
(c) correspond to those in (a). The cylinder surface is represented by the solid black line.
numbers where steady structures beside the pendant drop exist becomes smaller. The com-
pletely wetting case (β0= 0) is qualitatively different and is discussed below in Section III C.
Increasing β0above β0= 2 the critical value Bocstrongly increases, e.g., for β0= 10 one
finds Boc= 43.3.
Note, that the pendant drop underneath the cylinder [drop on top of cylinder] corresponds
on a horizontal heterogeneous substrate to the drop on the more [less] wettable region as
discussed by [49]. The limit of zero Bond number is analogue to the horizontal homogeneous
substrate.
Next, we increase the dimensionless angular velocity Ω for several selected Bond numbers
and determine the steady state solutions (Fig. 4). Focusing first on Bo= 1 we recognise at
Ω = 0 three solutions as marked by the vertical dotted line in Fig. 2. The curve of largest
norm corresponds to the pendant drop. As Ω increases from zero, its norm and amplitude
decrease slightly, whereas its centre of mass moves towards larger θ, it is ‘dragged along’ by
the rotation without changing its shape much. This can be well appreciated in Fig. 5 which
12
0 1 2 3 4 5
0
0.2
0.4
0.6
0.8
1
1.2
||δh||
0.5
0.75
1.0
2.5
5.0
10.0
Bo
FIG. 4: Solutions on a horizontal rotating cylinder as a function of the Rotation number Ω for
various Bond number Bo as given in the legend and contact angle β0= 2.0. The drop and film
profiles are characterised by their L2norm. The remaining parameters are h0= 0.1, and ¯
h= 1.0.
0
1
2
3
h(θ)
0.0
0.5
1.0
1.68
0 3.14 6.28
θ
-1
0
1
cos θ
a
b-10 0 10
x
-10
0
10
y
c
FIG. 5: Steady film thickness profiles on a rotating horizontal cylinder for Bond number Bo= 1
and β0= 2.0 for Rotation numbers Ω as given in the legend. All shown solutions are situated on
the stable upper branch in Fig. 4. The remaining parameters are as in Fig. 4. Panel (a) gives the
profiles as h(θ), (b) gives the space-dependent part of hf(Eq. (7)), and panel (c) illustrates the
profiles on a cylinder assuming a radius R= 10. Line styles in (c) correspond to those in (a).
13
11.5 22.5
0.94
0.96
0.98
1
1.02
||δh||
10-6 10-5 10-4 10-3 10-2 10-1 100
Ω−Ωsn
10-3
10-2
10-1
1/T
FIG. 6: Bifurcation diagram for the depinning transition for a drop on a rotating cylinder with
Bo= 1 and β0= 2.0. Shown are the L2norm for the branch of steady solutions as obtained by
continuation (solid line), selected steady solutions as obtained by direct integration in time (circles)
and the time-averaged L2norm for the unsteady solutions beyond the sniper depinning bifurcation
(triangles). The inset gives for the latter a log-log plot of the dependence of the inverse temporal
period on the distance from the bifurcation sn . The solid line indicates an exponent of 1/2.
gives selected profiles for the branch of pendant drop solutions. The branch terminates in a
saddle-node bifurcation at Ωsn = 1.68 where it annihilates with one of the unstable branches.
The third branch, i.e., the one of lowest norm, shows for increasing Ω an increase in the
norm, then it undergoes two saddle-node bifurcations at Ω = 1.313 and Ω = 1.329. At
larger Ω the norm decreases again. Note that the two saddle-node bifurcations are a first
sign of rather complicated re-connections which occur when one increases Bo from 1.0 to
2.5. The re-connections involve further (unstable) branches that are not shown in Fig. 4 and
result in the loop structure observed for Bo= 2.5 at about Ω = 1.5. A further indication for
the existence of additional branches is the observation that the value of the norm at Ω = 0
for Bo= 1.0 in Fig. 4 does not agree with the lowest norm at Bo= 1.0 in Fig. 2. As the
re-connections are not relevant for the depinning transition we will here not consider them
further.
At Bond numbers smaller than Bo= 1 the behaviour is qualitatively the same as described
14
(a) (b)
FIG. 7: Space-time plots illustrating the time evolution of co-rotating drops beyond depinning via
a sniper bifurcation (at >sn). Shown is one period in space and time (a) close to depinning
at Rotation number Ω = 1.8 with a temporal period of T= 14.8, and (b) far from depinning at
Ω=3.0 with T= 2.7. The Bond number is Bo= 1 and β0= 2.0.
for Bo= 1. However, the critical sn becomes smaller with decreasing Bo. Above, we have
discussed that the behaviour at Ω = 0 changes qualitatively when decreasing Bo beyond
Boc(Fig. 2). Although this also affects the behaviour at small Ω it has less influence on
the behaviour at larger Ω (Fig. 4). In particular, the saddle-node at Ωsn where the pendant
drop solution ceases to exist, persists for larger Bond number as does the related dynamic
behaviour (see below). Note that for Bo= 10 the saddle-node is at Ωsn 6.2 outside the
range of Fig. 4.
An interesting resulting question is what happens to a stable obliquely pendant droplet
at Ω <sn when the rotation velocity is further increased such that it is slightly larger than
sn? Does the solution approach the remaining steady solution? Simulations of Eq. (5)
show that this is not the case. For Ω >sn the droplet moves in a non-stationary way with
the rotating cylinder and corresponds to a space- and time-periodic solution. The resulting
bifurcation diagram for Bo= 1 is given in Fig. 6. It shows that the ‘new’ branch of space- and
time-periodic solutions (characterised by its time-averaged norm) emerges from the saddle
node bifurcation of the steady solutions at Ωsn. The behaviour close to and far from the
saddle-node bifurcation is illustrated in the space-time plots of Fig. 7. Panel (b) shows
that far away from the bifurcation the drop moves continuously with its velocity and shape
varying smoothly. It moves fastest when its maximum passes θπ/2 (velocity 4.6)
15
and slows down when passing θ3π/2 (velocity 1.2), i.e., respectively, when gravity
most strongly supports and hinders the motion driven by the cylinder rotation. The ratio
of largest to smallest velocity is about 4:1. When approaching the bifurcation from above,
the droplet still moves with the rotation but the time scales of the slow and the fast phase
become very different. For instance, at Ω = 1.8 (Fig. 7 (a)) the drop moves fastest when
passing θπ/2 (velocity 3.3). However, when the drop is situated at about θ= 3π/2
it barely moves (velocity 0.1) resulting in a velocity ratio of 33:1 between the fastest
and slowest phase. The resulting overall behaviour strongly resembles the stick-slip motion
discussed in the context of contact line motion on heterogeneous substrates [52]. Note that
it also resembles so-called sloshing modes found for a partially liquid-filled rotating cylinder
[19, 54].
The ratio of the velocities diverges when approaching the bifurcation point. In conse-
quence, the frequency 1/T of the periodic drop motion goes to zero. As shown in the inset of
Fig. 6, the dependence corresponds to a power law 1/T (Ω sn)1/2. This indicates that
the bifurcation at Ωsn is actually a Saddle-Node-Infinite-PERiod bifurcation (or ‘sniper’ for
short; see [47], and discussion by [52]).
The described behaviour is quite generic for the partially wetting case, i.e., it is found
in a wide range of parameters, in particular, Bond number, contact angle and wetting
layer thickness. However, the behaviour changes dramatically when strongly decreasing the
contact angle, i.e., when approaching the completely wetting case.
C. The completely wetting case
After having discussed the intricate depinning behaviour in the partially wetting case,
to emphasise the contrast, we consider briefly the completely wetting case (β0= 0). Note
that the physical setting is then identical to the one employed in most studies of the classic
Moffatt problem [29, 40]. However, the parametrisation used here is different and results
in Eq. (5) with Ha= 0. We analyse the system behaviour as above by studying the steady
state solutions without and with rotation.
In the case without rotation, the first difference one notices is that for fixed Bond number
only one solution exists. It corresponds to the pendant drop solution. With increasing Bo,
i.e., with increasing importance of gravity, its amplitude monotonically increases. Fixing
16
0 1 2 3 4 5
0
0.2
0.4
0.6
0.8
1
1.2
||δh||
1.0
2.5
5.0
10.0
Bo
FIG. 8: Solutions for a film of completely wetting fluid (β0= 0) on a horizontal rotating cylinder
as a function of the Rotation number Ω for various Bond numbers Bo as given in the legend. The
solutions are characterised by their L2norm. The remaining parameter is ¯
h= 1.0.
0
1
2
3
h(θ)
1e-10
0.5
1.0
2.5
5.0
0 3.14 6.28
θ
-1
0
1
cos θ
a
b-10 0 10
x
-10
0
10
y
c
FIG. 9: Steady film thickness profiles for the case of complete wetting with Bo= 1 for Rotation
numbers Ω as given in the legend. The remaining parameters are as in Fig. 8. Panel (a) gives the
profiles as h(θ), (b) gives the space-dependent part of hf(Eq. (7)), and panel (c) illustrates the
profiles on a cylinder assuming a radius R= 10. Line styles in (c) correspond to those in (a).
17
Bo and increasing the rotation velocity from zero, one finds a monotonic decrease of the
norm (Fig. 8) and height of the drop. Its maximum moves towards larger angular position
θ, i.e., the liquid is dragged along with the rotation. However, in the wetting case there
is no force besides gravity that can favour drops as compared to a flat film. As gravity
acts downwards it only favours pendant drops. Therefore, the drop does not survive as a
‘coherent structure’ when dragged upwards with the rotation: it is flattened and smeared
out. Typical profiles are given in Fig. 9. Note that we do not show the exact case without
rotation, but choose Ω = 1010 as then a dynamically created wetting layer still exists that
is numerically advantageous (cf. [50]).
When increasing Ω further than shown in Fig. 8 one finds that the thickness profiles
approaches a flat film. For instance, for Bo= 1 the norm decreases below 102at Ω 70.
IV. CONCLUSION
Based on the observation that the equations in long-wave approximation that govern film
flow and drop motion (i) on or in a rotating cylinder and (ii) on a heterogeneous substrate are
rather similar, we have explored whether the analogy can be exploited, i.e., whether it allows
results obtained for one system to be transferred to the other one. In particular, we have
found that indeed for drops on a rotating cylinder there exists a counterpart of the rather
involved depinning dynamics described recently for drops on heterogeneous substrates.
To study the effect we have introduced an alternative scaling. This has been necessary
because the commonly used scaling contains the angular velocity of the cylinder rotation in
both dimensionless parameters and in the time scale. Together with a further re-scaling of the
steady state equation involving the flow rate [23, 40] this does not allow for an investigation
of either the system behaviour when increasing the rotation speed of the cylinder from
zero or the existence of multiple solutions for fixed rotation speed and liquid volume. The
scaling employed here does allow for such studies as the rotation speed only enters a single
dimensionless parameter. Furthermore, it allows us to discuss the analogy between the two
systems of interest in a rather natural way. In particular, downward gravitation and rotation
speed for the drop on the rotating cylinder correspond to the heterogeneous wettability and
lateral driving force, respectively, for drops on heterogeneous substrates.
Guided by this analogy we have studied the behaviour of drops of partially wetting liquid
18
on the rotating cylinder. As a result it has been shown, how ‘switching on’ gravity (in-
creasing the Bond number) without rotation changes the solution behaviour dramatically as
the ‘gravitational heterogeneity’ along the cylinder surface effectively suppresses multidrop
solutions until only the single pendant droplet underneath the cylinder survives. We have
furthermore found that increasing the rotation from zero to small values, the drop is dragged
along to a stable equilibrium position where downwards gravity and upwards drag compen-
sate. At a critical speed, however, the retaining downwards force is too small and the drop
is dragged above the left horizontal position. It then continuously moves with the rotating
cylinder in a non-constant manner. Close to the transition the drop shows stick-slip motion.
An analysis of the time-dependent behaviour has shown that the frequency related to the
periodic motion goes to zero at the threshold following a power law with power 1/2. This
and the related bifurcation diagram show that the observed transition corresponds to depin-
ning via a sniper bifurcation. The behaviour found is rather generic for the case of partial
wetting. Although, here we have presented results for particular parameter values, it can
be observed in a wide range of parameters, in particular, Bond number, equilibrium contact
angle and wetting layer thickness. However, we have also found that the behaviour changes
dramatically when strongly decreasing the equilibrium contact angle, i.e., when approaching
the completely wetting case that is normally studied in the literature. When increasing the
velocity of rotation in this case we have seen a transition from pendant drops underneath
the cylinder to a nearly uniform film around the cylinder as [14].
The presented results indicate that it may be very fruitful to further explore the analogy
of films/drops on rotating cylinders and heterogeneous substrates. We expect that the
recent exploration of the three-dimensional case for depinning drops [5] and for the various
transitions between ridge, drop and rivulet states [4] based on tools for the continuation of
steady and stationary solutions of the two-dimensional thin film equation (physically three-
dimensional system) [6] is also relevant for the case of a rotating cylinder. A future study
could, e.g., elucidate the relation between the formation of azimuthal rings, pendant ridges
and sets of drops.
The present circular cylinder is an analogy to a sinusoidal wettability profile. Cylinders
with other cross sections, for instance, an elliptical cylinder [20] would correspond to more lo-
calised defects of the corresponding heterogeneous substrate. However, the analogue system
turns out to be more complicated as a rotating non-circular cylinder introduces the aspect
19
of a time-periodic non-harmonic forcing parallel to the substrate into the heterogeneous
substrate system.
Our equation (5) is derived under two conditions: (i) the mean (and also the maximal)
film thickness has to be small as compared to the radius of the cylinder; and (ii) the local
surface slope (e.g., the physical equilibrium contact angle) has to be small. Here we have
assumed that the two related smallness parameters are of the same order of magnitude. In
a next step one may introduce two different smallness parameters and discuss a number of
distinct limits in dependence of their ratio. Such a systematic asymptotic study would also
permit to discuss the influence of other effects on the depinning behaviour like, for instance,
the effects of the hydrostatic pressure [1, 2], centrifugal forces [14] and inertia [19, 25].
V. ACKNOWLEDGEMENTS
I acknowledge support by the EU via grant PITN-GA-2008-214919 (MULTIFLOW).
First steady state solutions where calculated by N. Barranger employing a predecessor of
the present model. I benefited from discussions with E. Knobloch and several group members
at Loughborough University as well as from the input of several referees and the editor.
[1] Acrivos, A. & Jin, B. 2004 Rimming flows within a rotating horizontal cylinder: asymptotic
analysis of the thin-film lubrication equations and stability of their solutions. J. Eng. Math.
50, 99–120.
[2] Ashmore, J., Hosoi, A. E. & Stone, H. A. 2003 The effect of surface tension on rimming
flows in a partially filled rotating cylinder. J. Fluid Mech. 479, 65–98.
[3] Bauer, C. & Dietrich, S. 2000 Phase diagram for morphological transitions of wetting
films on chemically structured substrates. Phys. Rev. E 61, 1664–1669.
[4] Beltrame, P., E., Knobloch, H¨
anggi, P. & Thiele, U. 2010 Rayleigh and depinning
instabilities of forced liquid ridges on heterogeneous substrates. Phys. Rev. E (submitted).
[5] Beltrame, P., H¨
anggi, P. & Thiele, U. 2009 Depinning of three-dimensional drops from
wettability defects. Europhys. Lett. 86, 24006.
20
[6] Beltrame, P. & Thiele, U. 2010 Time integration and steady-state continuation method
for lubrication equations. SIAM J. Appl. Dyn. Syst. 9, 484–518.
[7] Benilov, E. S., Benilov, M. S. & Kopteva, N. 2008 Steady rimming flows with surface
tension. J. Fluid Mech. 597, 91–118.
[8] Benilov, E. S. & O’Brien, S. B. G. 2005 Inertial instability of a liquid film inside a
rotating horizontal cylinder. Phys. Fluids 17, 052106.
[9] Bonn, D., Eggers, J., Indekeu, J., Meunier, J. & Rolley, E. 2009 Wetting and
spreading. Rev. Mod. Phys. 81, 739–805.
[10] Brandon, S. & Marmur, A. 1996 Simulation of contact angle hysteresis on chemically
heterogeneous surfaces. J. Colloid Interface Sci. 183, 351–355.
[11] Brusch, L., K¨
uhne, H., Thiele, U. & B¨
ar, M. 2002 Dewetting of thin films on hetero-
geneous substrates: Pinning vs. coarsening. Phys. Rev. E 66, 011602.
[12] Doedel, E. J., Paffenroth, R. C., Champneys, A. R., Fairgrieve, T. F.,
Kuznetsov, Y. A., Oldeman, B. E., Sandstede, B. & Wang, X. J. 1997 AUTO2000:
Continuation and bifurcation software for ordinary differential equations. Montreal: Concor-
dia University.
[13] Duffy, B. R. & Wilson, S. K. 2009 Large-Biot-number non-isothermal flow of a thin film
on a stationary or rotating cylinder. Eur. Phys. J.-Spec. Top. 166, 147–150.
[14] Evans, P. L., Schwartz, L. W. & Roy, R. V. 2004 Steady and unsteady solutions for
coating flow on a rotating horizontal cylinder: Two-dimensional theoretical and numerical
modeling. Phys. Fluids 16, 2742–2756.
[15] Evans, P. L., Schwartz, L. W. & Roy, R. V. 2005 Three-dimensional solutions for coating
flow on a rotating horizontal cylinder: Theory and experiment. Phys. Fluids 17, 072102.
[16] Gau, H., Herminghaus, S., Lenz, P. & Lipowsky, R. 1999 Liquid morphologies on
structured surfaces: From microchannels to microchips. Science 283, 46–49.
[17] de Gennes, P.-G. 1985 Wetting: Statics and dynamics. Rev. Mod. Phys. 57, 827–863.
[18] Hinch, E. J. & Kelmanson, M. A. 2003 On the decay and drift of free-surface perturbations
in viscous thin-film flow exterior to a rotating cylinder. Proc. R. Soc. London Ser. A-Math.
Phys. Eng. Sci. 459, 1193–1213.
[19] Hosoi, A. E. & Mahadevan, L. 1999 Axial instability of a free-surface front in a partially
filled horizontal rotating cylinder. Phys. Fluids 11, 97–106.
21
[20] Hunt, R. 2008 Numerical solution of the free-surface viscous flow on a horizontal rotating
elliptical cylinder. Numer. Meth. Part Differ. Equ. 24, 1094–1114.
[21] Israelachvili, J. N. 1992 Intermolecular and Surface Forces. London: Academic Press.
[22] John, K., H¨
anggi, P. & Thiele, U. 2008 Ratchet-driven fluid transport in bounded two-
layer films of immiscible liquids. Soft Matter 4, 1183–1195.
[23] Karabut, E. A. 2007 Two regimes of liquid film flow on a rotating cylinder. J. Appl. Mech.
Tech. Phys. 48, 55–64.
[24] Kargupta, K. & Sharma, A. 2003 Mesopatterning of thin liquid films by templating on
chemically patterned complex substrates. Langmuir 19, 5153–5163.
[25] Kelmanson, M. A. 2009 On inertial effects in the Moffatt-Pukhnachov coating-flow problem.
J. Fluid Mech. 633, 327–353.
[26] Konnur, R., Kargupta, K. & Sharma, A. 2000 Instability and morphology of thin liquid
films on chemically heterogeneous substrates. Phys. Rev. Lett. 84, 931–934.
[27] Lenz, P. & Lipowsky, R. 2000 Stability of droplets and channels on homogeneous and
structured surfaces. Eur. Phys. J. E 1, 249–262.
[28] Merkt, D., Pototsky, A., Bestehorn, M. & Thiele, U. 2005 Long-wave theory of
bounded two-layer films with a free liquid-liquid interface: Short- and long-time evolution.
Phys. Fluids 17, 064104.
[29] Moffatt, H. K. 1977 Behavior of a viscous film on outer surface of a rotating cylinder. J.
Mec. 16, 651–673.
[30] Noakes, C. J., King, J. R. & Riley, D. S. 2005 On three-dimensional stability of a
uniform, rigidly rotating film on a rotating cylinder. Q. J. Mech. Appl. Math. 58, 229–256.
[31] Noakes, C. J., King, J. R. & Riley, D. S. 2006 On the development of rational ap-
proximations incorporating inertial effects in coating and rimming flows: a multiple-scales
approach. Q. J. Mech. Appl. Math. 59, 163–190.
[32] O’Brien, S. B. G. 2002 Linear stability of rimming flow. Q. Appl. Math. 60, 201–211.
[33] Oron, A., Davis, S. H. & Bankoff, S. G. 1997 Long-scale evolution of thin liquid films.
Rev. Mod. Phys. 69, 931–980.
[34] Oron, A. & Rosenau, P. 1992 Formation of patterns induced by thermocapillarity and
gravity. J. Physique II France 2, 131–146.
[35] Pedley, T. J. 1967 The stability of rotating flows with a cylindrical free surface. J. Fluid
22
Mech. 30, 127–147.
[36] Peterson, R. C., Jimack, P. K. & Kelmanson, M. A. 2001 On the stability of viscous
free-surface flow supported by a rotating cylinder. Proc. R. Soc. London Ser. A-Math. Phys.
Eng. Sci. 457, 1427–1445.
[37] Phillips, O. M. 1960 Centrifugal waves. J. Fluid Mech. 7, 340–352.
[38] Pismen, L. M. 2001 Nonlocal diffuse interface theory of thin films and the moving contact
line. Phys. Rev. E 64, 021603.
[39] Pismen, L. M. & Thiele, U. 2006 Asymptotic theory for a moving droplet driven by a
wettability gradient. Phys. Fluids 18, 042104.
[40] Pukhnachev, V. V. 1977 The liquid film motion on the rotating cylinder surface at the
presence of gravity. Prikl. Mekh. Teor. Fiz. 3, 77–88.
[41] Pukhnachov, V. V. 2005 On the equation of a rotating film. Sib. Math. J. 46, 913–924.
[42] Qu´
er´
e, D., Azzopardi, M. J. & Delattre, L. 1998 Drops at rest on a tilted plane.
Langmuir 14, 2213–2216.
[43] Reisfeld, B. & Bankoff, S. G. 1992 Nonisothermal flow of a liquid-film on a horizontal
cylinder. J. Fluid Mech. 236, 167–196.
[44] Rockford, L., Liu, Y., Mansky, P., Russell, T. P., Yoon, M. & Mochrie, S. G. J.
1999 Polymers on nanoperiodic, heterogeneous surfaces. Phys. Rev. Lett. 82, 2602–2605.
[45] Ruschak, K. J. & Scriven, L. E. 1976 Rimming flow of liquid in a rotating horizontal
cylinder. J. Fluid Mech. 76, 113–125.
[46] Schwartz, L. W. & Garoff, S. 1985 Contact-angle hysteresis on heterogeneous surfaces.
Langmuir 1, 219–230.
[47] Strogatz, S. H. 1994 Nonlinear Dynamics and Chaos. Addison-Wesley.
[48] Thiele, U. 2010 Thin film evolution equations from (evaporating) dewetting liquid layers to
epitaxial growth. J. Phys.: Condens. Matter 22, 084019.
[49] Thiele, U., Brusch, L., Bestehorn, M. & B¨
ar, M. 2003 Modelling thin-film dewetting
on structured substrates and templates: Bifurcation analysis and numerical simulations. Eur.
Phys. J. E 11, 255–271.
[50] Thiele, U. & Knobloch, E. 2004 Thin liquid films on a slightly inclined heated plate.
Physica D 190, 213–248.
[51] Thiele, U. & Knobloch, E. 2006 Driven drops on heterogeneous substrates: Onset of
23
sliding motion. Phys. Rev. Lett. 97, 204501.
[52] Thiele, U. & Knobloch, E. 2006 On the depinning of a driven drop on a heterogeneous
substrate. New J. Phys. 8, 313, 1–37.
[53] Thiele, U., Velarde, M. G., Neuffer, K., Bestehorn, M. & Pomeau, Y. 2001 Sliding
drops in the diffuse interface model coupled to hydrodynamics. Phys. Rev. E 64, 061601.
[54] Thoroddsen, S. T. & Mahadevan, L. 1997 Experimental study of coating flows in a
partially-filled horizontally rotating cylinder. Exp. Fluids 23, 1–13.
[55] Weidner, D. E., Schwartz, L. W. & Eres, M. H. 1997 Simulation of coating layer
evolution and drop formation on horizontal cylinders. J. Colloid Interface Sci. 187, 243–258.
[56] Yih, C. S. & Kingman, J. F. C. 1960 Instability of a rotating liquid film with a free surface.
Proc R Soc London Ser A-Math 258, 63–89.
... For a well wetting liquid studied in Reisfeld & Bankoff (1992), these forces favor a thick film, and the disjoining pressure can be described as Π(h) = −A/h 3 with a positive Hamaker constant A > 0. In contrast, for a dewetting liquid, the purely destabilizing intermolecular forces modeled by Π(h) = A/h 3 can lead to finite-time rupture in the film thickness. For partially wetting liquids, a combination of stabilizing and destabilizing molecular interactions are involved and the disjoining pressure takes the form Π(h) = −A/h 3 (1−b/h) (Thiele (2011)). For an extensive review of this topic, we refer the readers to de Gennes (1985); Bonn et al. (2009) and Israelachvili (2011). ...
... The role of the intermolecular forces in slowly withdrawing a thin fiber out of a bath of wetting liquid has been studied in Quéré et al. (1989) and Quéré (1999). The dynamics of non-isothermal liquid film with van der Waals interactions on a horizontal cylinder was considered in Reisfeld & Bankoff (1992) and Thiele (2011). To the best of our knowledge, the stabilization of the coating film in dynamics of liquid flowing down vertical fibers has not been discussed in the literature. ...
Article
Recent experiments on thin films flowing down a vertical fibre with varying nozzle diameters present a wealth of new dynamics that illustrate the need for more advanced theory. We present a detailed analysis using a full lubrication model that includes slip boundary conditions, nonlinear curvature terms and a film stabilization term. This study brings to focus the presence of a stable liquid layer playing an important role in the full dynamics. We propose a combination of these physical effects to explain the observed velocity and stability of travelling droplets in the experiments and their transition to isolated droplets. This is also supported by stability analysis of the travelling wave solution of the model.
... For a well wetting liquid studied in Reisfeld & Bankoff (1992), these forces favor a thick film, and the disjoining pressure can be described as Π(h) = −A/h 3 with a positive Hamaker constant A > 0. In contrast, for a dewetting liquid, the purely destabilizing intermolecular forces modeled by Π(h) = A/h 3 can lead to finite-time rupture in the film thickness. For partially wetting liquids, a combination of stabilizing and destabilizing molecular interactions are involved and the disjoining pressure takes the form Π(h) = −A/h 3 (1−b/h) (Thiele (2011)). For an extensive review of this topic, we refer the readers to de Gennes (1985); Bonn et al. (2009) and Israelachvili (2011). ...
... The role of the intermolecular forces in slowly withdrawing a thin fiber out of a bath of wetting liquid has been studied in Quéré et al. (1989) and Quéré (1999). The dynamics of non-isothermal liquid film with van der Waals interactions on a horizontal cylinder was considered in Reisfeld & Bankoff (1992) and Thiele (2011). To the best of our knowledge, the stabilization of the coating film in dynamics of liquid flowing down vertical fibers has not been discussed in the literature. ...
Preprint
Full-text available
Recent experiments of thin films flowing down a vertical fiber with varying nozzle diameters present a wealth of new dynamics that illustrate the need for more advanced theory. We present a detailed analysis using a full lubrication model that includes slip boundary conditions, nonlinear curvature terms, and a film stabilization term. This study brings to focus the presence of a stable liquid layer playing an important role in the full dynamics. We propose a combination of these physical effects to explain the observed velocity and stability of traveling droplets in the experiments and their transition to isolated droplets. This is also supported by stability analysis of the traveling wave solution of the model.
... is motivated by the form of disjoining pressure widely used in lubrication equations [7,26] to describe the wetting behaviour of a liquid on a solid substrate, and the scaling parameter A > 0 is typically selected based on a stable liquid layer in the coating film dynamics. Numerical investigations against experimental results in [10] show that compared with previous studies, the combined physical effects better describe the propagation speed and the stability transition of the moving droplets. ...
Article
Existence of non-negative weak solutions is shown for a full curvature thin-film model of a liquid thin film flowing down a vertical fibre. The proof is based on the application of a priori estimates derived for energy-entropy functionals. Long-time behaviour of these weak solutions is analysed and, under some additional constraints for the model parameters and initial values, convergence towards a travelling wave solution is obtained. Numerical studies of energy minimisers and travelling waves are presented to illustrate analytical results.
... In this paper, we numerically and experimentally study the free surface flow between vertical rotating cylinders. Some characteristics of the free surface flow around a rotating cylinder were the targets of the study of viscous fingerings [1], [2]. After the pioneering and milestone study conducted about one hundred years ago [3], a series of modified flows with no main axial flow has been carried out on the flows between rotating cylinders with finite lengths [4]. ...
... A full treatment of the problem requires a model that incorporates the subtle interplay of gravity, surface tension, viscous effects, inertia (including centrifugation), rotational effects and potentially even dewetting effects (Thiele 2011;Lin et al. 2016) (all other models, including the ones derived herein, implicitly assume complete wetting). As a result, since the pioneering work of Moffatt and Pukhnachev, many authors have extended this coating flow problem. ...
Article
Full-text available
Reduced-order modelling of thick inertial flows around rotating cylinders - Volume 898 - Alexander W. Wray, Radu Cimpeanu
... Many scientific and industrial problems involve the flow of thin liquid films (see, for instance [1][2][3][4][5][6][7][8]). Thin film technology is used extensively in many applications including microelectronics, optics, magnetism, hard and corrosion resistant coatings, biotechnology, micro-mechanics, lasers, and medicine. ...
Article
Full-text available
The motion of a thin suspended film of an incompressible fluid under the effect of an external electric field is studied. The effects of the interfacial Maxwell stress, surface tension and intermolecular forces are studied, in which the forces are included in the Navier-Stokes equations. The perturbation technique is used to solve the given model. The obtained results show that, the fluid moves in a rotating pattern and the fluid particles move along the streamlines with different velocities. The free boundaries show good agreement with experimental data. In addition, the stability criteria are examined in the present model.
... Also, in this chapter, we consider the dynamics for some generalisation of (1.11) and (1.8), namely, h t + (|h| n (a 0 h xxx + a 1 h x + a 2 w x (x, t))) x = 0, (1.12) where n > 0. The understanding of this dynamics is important for industrial printing processes, where rotating cylinder transports the coating material in the form of liquid paint. The rotating thin fluid film can exhibit variety of different behaviour including: interesting pattern formations ("shark teeth" and "duck bill" patterns), fluid curtains, hydroplaning drops, and frontal avalanches [159,161,170]. As a result, the coating flow has been the subject of continuous study since the pioneering work of Moffatt [116]. ...
... Many scientific and industrial problems connect to the flow of thin liquid films (see, for instance [1][2][3][4][5][6][7][8]). Thin film technology is used extensively in many applications including microelectronics, optics, magnetic, hard and corrosion resistant coatings, biotechnology, micro-mechanics, laser, medicine, etc. ...
Preprint
Full-text available
The motion of a thin suspended film of an incompressible fluid under the effect of an external electric field is studied. The effect of the interfacial Maxwell stress, surface tension and inter-molecular forces is studied, in which the forces are included in the Navier-Stokes equations. The perturbation technique is used to solve for a given model. The obtained results show that, the fluid moving in a rotating patterns and the fluid particles move along the streamlines with different velocities. The free boundaries show good results in comparison with experiments data. In addition, the stability criteria are examined in the preset model.
Article
Full-text available
Motivated by strategies for targeted microfluidic transport of droplets, we investigate how sessile droplets can be steered toward a preferred direction using travelling waves in substrate wettability or deformations of...
Article
The phase field approach is applied to numerically simulate the detachment of an isolated, wall-bound 2D pendant drop suspended in a fluid in a simple shear flow. The model has been previously employed to simulate several two-phase flow phenomena, assuming that the system consists of a regular, partially miscible mixture, with the drop and the continuous phase being in thermodynamic equilibrium with each other. In addition, it is assumed that the two phases are separated by an interfacial region having a non-zero characteristic thickness [Formula: see text], i.e., the interface is diffuse. In the creeping flow regime, the problem is described in terms of three non-dimensional numbers: the fluidity number [Formula: see text] as the ratio between capillary and viscous fluxes, the Bond number [Formula: see text] as the ratio between external and capillary forces, and the Peclet number [Formula: see text] as a non-dimensional shear rate. We find that, at large fluidity numbers and for small droplets (i.e., for [Formula: see text]), the onset of the drop detachment can be described in terms of a master curve, with the critical macroscopic Bond number [Formula: see text] decreasing monotonously with [Formula: see text] for five drop sizes in the micrometer range.
Article
Full-text available
A model for the evolution of a thin liquid coating on a horizontal cylinder is presented. The cylinder rotates about its axis, carrying liquid around its circumference. For a viscous coating, this leads to formation of a relatively thick coating where the cylinder surface moves upward. The model is based on lubrication theory, as the coating is thin compared to the cylinder radius, and includes the effects of cylinder rotation, gravity, surface tension, and flow along the cylinder axis. A two-dimensional numerical scheme based on finite differences is produced, for investigation of the case when axial flow is neglected. This numerical scheme is validated in appropriate limiting cases. Coating cross sections are obtained over a range of cylinder rotation rates, for realistic parameter values. These show a transition from pendant drops hanging beneath the cylinder to a nearly uniform coating wrapped around it as rotation rate is increased.
Article
Full-text available
Macroscopic thin liquid films are entities that are important in biophysics, physics, and engineering, as well as in natural settings. They can be composed of common liquids such as water or oil, rheologically complex materials such as polymers solutions or melts, or complex mixtures of phases or components. When the films are subjected to the action of various mechanical, thermal, or structural factors, they display interesting dynamic phenomena such as wave propagation, wave steepening, and development of chaotic responses. Such films can display rupture phenomena creating holes, spreading of fronts, and the development of fingers. In this review a unified mathematical theory is presented that takes advantage of the disparity of the length scales and is based on the asymptotic procedure of reduction of the full set of governing equations and boundary conditions to a simplified, highly nonlinear, evolution equation or to a set of equations. As a result of this long-wave theory, a mathematical system is obtained that does not have the mathematical complexity of the original free-boundary problem but does preserve many of the important features of its physics. The basics of the long-wave theory are explained. If, in addition, the Reynolds number of the flow is not too large, the analogy with Reynolds's theory of lubrication can be drawn. A general nonlinear evolution equation or equations are then derived and various particular cases are considered. Each case contains a discussion of the linear stability properties of the base-state solutions and of the nonlinear spatiotemporal evolution of the interface (and other scalar variables, such as temperature or solute concentration). The cases reducing to a single highly nonlinear evolution equation are first examined. These include: (a) films with constant interfacial shear stress and constant surface tension, (b) films with constant surface tension and gravity only, (c) films with van der Waals (long-range molecular) forces and constant surface tension only, (d) films with thermocapillarity, surface tension, and body force only, (e) films with temperature-dependent physical properties, (f) evaporating/condensing films, (g) films on a thick substrate, (h) films on a horizontal cylinder, and (i) films on a rotating disc. The dynamics of the films with a spatial dependence of the base-state solution are then studied. These include the examples of nonuniform temperature or heat flux at liquid-solid boundaries. Problems which reduce to a set of nonlinear evolution equations are considered next. Those include (a) the dynamics of free liquid films, (b) bounded films with interfacial viscosity, and (c) dynamics of soluble and insoluble surfactants in bounded and free films. The spreading of drops on a solid surface and moving contact lines, including effects of heat and mass transport and van der Waals attractions, are then addressed. Several related topics such as falling films and sheets and Hele-Shaw flows are also briefly discussed. The results discussed give motivation for the development of careful experiments which can be used to test the theories and exhibit new phenomena.
Article
Full-text available
Wetting phenomena are ubiquitous in nature and technology. A solid substrate exposed to the environment is almost invariably covered by a layer of fluid material. In this review, the surface forces that lead to wetting are considered, and the equilibrium surface coverage of a substrate in contact with a drop of liquid. Depending on the nature of the surface forces involved, different scenarios for wetting phase transitions are possible; recent progress allows us to relate the critical exponents directly to the nature of the surface forces which lead to the different wetting scenarios. Thermal fluctuation effects, which can be greatly enhanced for wetting of geometrically or chemically structured substrates, and are much stronger in colloidal suspensions, modify the adsorption singularities. Macroscopic descriptions and microscopic theories have been developed to understand and predict wetting behavior relevant to microfluidics and nanofluidics applications. Then the dynamics of wetting is examined. A drop, placed on a substrate which it wets, spreads out to form a film. Conversely, a nonwetted substrate previously covered by a film dewets upon an appropriate change of system parameters. The hydrodynamics of both wetting and dewetting is influenced by the presence of the three-phase contact line separating ``wet'' regions from those that are either dry or covered by a microscopic film only. Recent theoretical, experimental, and numerical progress in the description of moving contact line dynamics are reviewed, and its relation to the thermodynamics of wetting is explored. In addition, recent progress on rough surfaces is surveyed. The anchoring of contact lines and contact angle hysteresis are explored resulting from surface inhomogeneities. Further, new ways to mold wetting characteristics according to technological constraints are discussed, for example, the use of patterned surfaces, surfactants, or complex fluids.
Book
This reference describes the role of various intermolecular and interparticle forces in determining the properties of simple systems such as gases, liquids and solids, with a special focus on more complex colloidal, polymeric and biological systems. The book provides a thorough foundation in theories and concepts of intermolecular forces, allowing researchers and students to recognize which forces are important in any particular system, as well as how to control these forces. This third edition is expanded into three sections and contains five new chapters over the previous edition.
Article
A liquid drop on a tilted plate can remain at rest, provided that an hysteresis for the contact angle exists. We examine in this paper the case of small contact angle hysteresis, for which the drops are found to be nearly spherical caps, despite the slope. We derive in particular the condition (both in angle and hysteresis) under which the drop remains stuck on the surface: a simple model is proposed and compared with experimental data.
Article
We consider the case of rimming flow where a thin film of viscous liquid coats the walls of a cylinder whose axis is horizontal and which is rotating with constant angular velocity. It has been experimentally established that such flows are often unstable, and that the liquid often segregates into “rings” along the length of the tube. Using a leading-order lubrication theory, we utilise recently established steady solutions [the author and E. G. Gath, Phys. Fluids 10, 1-3 (1998)], which in some instances contain shocks, to examine the linear stability of the flow when subjected to two-dimensional disturbances. All solutions are shown to be at least neutrally stable. We suggest that further investigations should include higher-order (small) effects, and that the origin of the observed instabilities lies in these terms.
Article
We study the linear stability of rigidly rotating films on a rotating circular cylinder in the absence of gravity. Three-dimensional disturbances are examined, particularly in the limit in which the liquid volume fraction is small, it being assumed throughout that there is a continuous film coating the surface of the cylinder. The most unstable mode for thin film flows on the inside surface of a cylinder, that is, rimming flows, is shown to be purely axial (ring instability), and the critical wave number is identified for given values of the reduced Reynolds number and reciprocal Weber number S. For thin film flows on the outside surface of a cylinder, that is, coating flows, the critical wave number (that with maximal growth rate) is also identified for small reduced Reynolds numbers. For S > 1, the disturbance is purely axial (ring instability). For 0 < S < 1, however, the most unstable disturbance is shown to be either a purely axial or a purely azimuthal mode. Numerical results suggest that a stripe instability (azimuthal mode) is likely to occur. While agreeing with some experimental evidence, this result conflicts with results from others, suggesting that effects ignored in this study, gravity for example, may play a significant role in mode selection.
Article
Herein we establish a relationship between controlled nanoscale surface interactions and subsequent macromolecular ordering. Striped surfaces of oxide and metal are generated over large areas, where the stripe width is comparable to the size of a polymer molecule. Homopolymers are found to anisotropically dewet such surfaces, while incompatible polymer mixtures phase separate at the substrate/polymer interface on a molecular level. An orientation of lamellar block copolymer microdomains normal to the surface is found, when substrate and polymer length scales are commensurate. {copyright} {ital 1999} {ital The American Physical Society}
Article
The centripetal acceleration of a rotating liquid film is tantamount to a centrifugal force, which tends to cause the liquid film to form rings around the circular cylinder to which it is attached. The stabilizing factors are surface tension and, presumably, viscosity. But it is shown in this paper that instability occurs even for large values of the surface-tension parameter and at small Reynolds numbers. The critical wave number is shown to depend predominantly on the surface tension. Its dependence on the Reynolds number, R, is slight if R is small, and nil if R is large. The effect of viscosity is therefore essentially to slow down the rate of amplification of the unstable disturbances. The analysis is carried out for both large and small Reynolds numbers, for various ratios of film thickness to cylinder radius, and for various surface tension parameters. (The calculation for intermediate Reynolds numbers turns out to be unnecessary for the purpose of comparison with the experiments obtained. Enough information is provided by the calculations performed for practical applications.) Numerical results are given. Comparison of results obtained from 65 experiments with pure glycerine, water + glycerine mixture, and water with the analytical results shows satisfactory agreement.