You are on page 1of 17

Thin-liquid-film flow on a topographically patterned rotating cylinder

Akhilesh K. Sahu and Satish Kumar


Citation: Physics of Fluids 26, 042102 (2014); doi: 10.1063/1.4869208
View online: http://dx.doi.org/10.1063/1.4869208
View Table of Contents: http://scitation.aip.org/content/aip/journal/pof2/26/4?ver=pdfcov
Published by the AIP Publishing
Articles you may be interested in
Stability analysis of non-inertial thin film flow over a heterogeneously heated porous substrate
Phys. Fluids 28, 022104 (2016); 10.1063/1.4941306
Visualization of flow patterns past various objects in two-dimensional flow using soap film
Phys. Fluids 23, 091104 (2011); 10.1063/1.3640020
The flow of thin liquid films over spinning disks: Hydrodynamics and mass transfer
Phys. Fluids 17, 052102 (2005); 10.1063/1.1891814
Erratum: Steady free-surface thin film flows over topography [Phys. Fluids 12, 1889 (2000)]
Phys. Fluids 12, 3305 (2000); 10.1063/1.1321265
Steady free-surface thin film flows over topography
Phys. Fluids 12, 1889 (2000); 10.1063/1.870438

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 14.139.208.84 On: Wed, 13 Apr
2016 11:51:52

PHYSICS OF FLUIDS 26, 042102 (2014)

Thin-liquid-film flow on a topographically patterned


rotating cylinder
Akhilesh K. Sahu and Satish Kumara)
Department of Chemical Engineering and Materials Science, University of Minnesota,
Minneapolis, Minnesota 55455, USA
(Received 6 February 2013; accepted 2 March 2014; published online 3 April 2014)

The flow of thin liquid films on rotating surfaces is directly relevant to the coating
of discrete objects. To begin understanding how surface topography influences such
flows, we consider a model problem in which a thin liquid film flows over a rotating
cylinder patterned with a sinusoidal surface topography. Lubrication theory is applied
to develop a partial differential equation that governs the film thickness as a function
of time and the angular coordinate. Static situations are considered first in order to
determine the parameter regime in which the lubrication approximation is expected
to be valid. When gravitational forces are relatively weak, cylinder rotation leads to
the formation of droplets connected by very thin films. The number of droplets is
equal to the pattern frequency at low and high rotation rates, with the droplets located
at the pattern troughs at low rotation rates and the pattern crests at high rotation
rates. When gravitational forces become significant, the film thickness never reaches
a steady state, in contrast to the case of an unpatterned cylinder. The results of this
work clearly establish that the flow of thin liquid films on rotating surfaces can be
C 2014 AIP Publishing LLC.
very sensitive to the presence of surface topography. 
[http://dx.doi.org/10.1063/1.4869208]

I. INTRODUCTION

Many coating processes involve the flow of thin liquid films on non-flat discrete objects.
Examples of products where such flows occur during manufacturing include drug-eluting stents,
pacemakers, contact lenses, chocolates, cereals, and auto bodies. Despite the prevalence of coating
of non-flat objects, relatively little is known about its fundamentals. As a consequence, it is difficult
to design and optimize such processes. For example, conventional coating techniques for stents have
problems of webbing between struts, varied drug concentration on struts, and variability in drug
concentration from stent to stent.1 In addition, wastage of the drug leads to a significant increase in
product cost.
As the surfaces to be coated may be rotated in practice, the flow of thin liquid films on rotating
surfaces is directly relevant to the coating of discrete objects. A popular model problem for studying
the coating of discrete objects is the flow of a thin liquid film on the outside of a rotating cylinder.
Building on the early efforts of Pukhnachev2 and Moffatt,3 many researchers have examined various
aspects of such flows, as reviewed in the paper of Evans et al.4 Both two-dimensional flows and
flows with axial variations have been considered. Lubrication theory is often employed to develop a
model equation, although full two-dimensional simulations have been conducted as well.
One aspect of these flows that appears to have been overlooked is the influence of surface
topography. For planar substrates, the influence of surface topography on thin-liquid-film flow has
been extensively investigated in recent years, as reviewed by Craster and Matar.5 The topography
may be sinusoidal in nature, or may consist of sharp features. By leading to modifications in the
interface shape, the presence of topography can generate additional flows that greatly influence film

a) Electronic mail: kumar030@umn.edu

1070-6631/2014/26(4)/042102/16/$30.00

26, 042102-1


C 2014 AIP Publishing LLC

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 14.139.208.84 On: Wed, 13 Apr
2016 11:51:52

042102-2

A. K. Sahu and S. Kumar

Phys. Fluids 26, 042102 (2014)

FIG. 1. Rotating topographically patterned cylinder.

evolution. Because surfaces appearing in the applications mentioned above are never perfectly flat,
and in many cases may contain an imposed topographical pattern, there is a need to investigate
how surface topography influences the flow of thin liquid films on rotating surfaces. To begin
understanding this issue, we consider a model problem in which a thin liquid film flows over a
rotating cylinder patterned with a sinusoidal surface topography.
Our work builds on the work of Evans et al.,4 who applied lubrication theory to develop a
partial differential equation that governs the film thickness as a function of time and the angular
coordinate. As in that paper, only two-dimensional flows are considered here. In Sec. II, we extend
their model to account for the presence of surface topography. Another important difference is that
a rotating reference frame needs to be adopted for the present problem. In Sec. III, we first consider
static situations in order to determine the parameter regime in which the lubrication approximation
is expected to be valid. This is followed by an investigation of the rotating case. Finally, conclusions
are presented in Sec. IV.
II. MATHEMATICAL MODEL

We consider flow of a thin film of Newtonian liquid with constant viscosity , surface tension
, and density over a topographically patterned cylinder. The cylinder is rotating about its axis
at constant angular speed  in the anti-clockwise direction (Fig. 1). The film completely wets the
cylinder and is sufficiently thick so that intermolecular forces can be neglected, and we study twodimensional flow with no axial variations. The pattern over the cylinder is taken to be sinusoidal
with frequency and amplitude . The radius of the cylinder, R, is described by
R = Rmean + sin( ),

(1)

where Rmean is the mean radius. Various shapes of the cylinder for different values of and are
shown in Fig. 2. To retain periodicity, only integer values of are considered.
We use cylindrical polar co-ordinates (r, ) in a frame of reference rotating with the cylinder,
where r is the radial distance from the center and is the angle from a fixed radial line in the cylinder
(Fig. 1). The film thickness h( , t) is measured from the surface of the cylinder, and for convenience
a new radial co-ordinate z = r (Rmean + sin( )) is defined so that the solid surface and the free
surface can be located easily at z = 0 and z = h( , t), respectively. The fluid velocity is expressed as
u = wer + ue ,

(2)

where er and e are unit vectors in the r- and -directions, respectively.


In rotating flows, it is more desirable and easier to work with the governing equations in a
rotating frame of reference6
u
1

+ u u = p + 2 u  ( r) 2 u + g,
t

(3)

u = 0,

(4)

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 14.139.208.84 On: Wed, 13 Apr
2016 11:51:52

042102-3

A. K. Sahu and S. Kumar

Phys. Fluids 26, 042102 (2014)

(a)

(b)

(c)

(d)

R mean

R mean

R mean

FIG. 2. Patterned cylinders with different values of : (a) = 0, (b) = 8, (c) = 11, and (d) = 21.

where p is pressure,  = e y with e y being the unit vector along the cylinder axis, and
g = (g sin f )er + (g cos f )e with g being a constant characterizing gravitational acceleration.
The angular position in a stationary frame of reference ( f ) is related to the position in the rotating
frame by f = + t.
The liquid satisfies the no-slip and no-penetration conditions at the cylinder surface
u = w = 0 at z = 0.

(5)

At the free surface, we have tangential and normal stress balances


n e t = 0,

(6)

= .
p + (n e n)

(7)

Here, e is the rate-of-strain tensor, n and t are the unit vectors normal and tangent to free surface,
and is the free-surface curvature. The unit vectors and free-surface curvature are defined as


 
h
1
1
(8)
er +
cos( )
e ,
n =
N
r

t =

1
N


 

1
h
cos( ) +
er + e ,
r

(9)

 


 

h 2
1 1
1 2h
2
2
= 3
sin( )
2
1 + 2 cos( ) +
,
(10)
N
r
r

r
2


2
. At the free surface we also have the kinematic boundary
where N = 1 + r12 cos( ) + h

condition
DF
= 0,
(11)
Dt
where F = z h( , t). Following Ref. 4, we take the pressure outside the film to be zero without
loss of generality.
We are interested in flows where the characteristic film thickness (H) is much smaller than the
mean radius of the patterned cylinder (
= H/Rmean  1). The disparity between these two length

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 14.139.208.84 On: Wed, 13 Apr
2016 11:51:52

042102-4

A. K. Sahu and S. Kumar

Phys. Fluids 26, 042102 (2014)

scales is exploited by applying lubrication theory. We use H to scale the film thickness and modified
radial co-ordinate z. Following Evans et al.,4 we use the characteristic speed U = gH2 / and define
dimensionless quantities as
u
w
r
u = ,
w =
,
r =
,
U

U
Rmean
=

Rmean

p =

p
,
g H

t =

t
.
Rmean /U

(12)

After substituting these scalings and neglecting small terms of O(


2 ), Eqs. (3) and (4) reduce to (bars
dropped)


2w
MWt
p
= W 2 (1 +
z),
+
2 sin +
(13)

z
z

2


 



u
p p
2u
MWt
= 0,

cos( ) + 2 +

cos +

z
z
z

(14)

u
(r w) +

cos( ) = 0.
(15)
z

z
Here, the dimensionless parameters M and W are a non-dimensional viscosity and rotational velocity,
M=



, W =
.
3
g/Rmean
g Rmean

(16)

Note that the r-component of the equation of motion, Eq. (13), is not affected by the surface
topography. However, the -component, Eq. (14), has an additional pressure-gradient term due to
surface topography. In addition, as is clear from Eqs. (13) and (14), inertial effects do not appear due
to the scalings considered in this work.
The simplified forms of the boundary conditions are
u = w = 0,

(17)

u

u + O(
2 ) = 0,
z

(18)

=
p +
(n e n)

1
,

Bo

(19)

where Bo, e, , and n are


Bo =

2
g Rmean
,

(20)

w
,
z


w
1 u
u
+
,
=
cos( )
r
z
r

err =
e



w
r
u
w
+
cos( )
,

z r
r
z
 2

h
2
= 1
(h + sin( ))

sin( ) + O(
2 ),
2


h

cos( )
e + O(
2 ).
n = er +
r

er = er =

(21)
(22)
(23)

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 14.139.208.84 On: Wed, 13 Apr
2016 11:51:52

042102-5

A. K. Sahu and S. Kumar

Phys. Fluids 26, 042102 (2014)

Here, Bo is the Bond number. The components of the rate-of-strain tensor (e) and the curvature
() are scaled by U/Rmean and 1/Rmean , respectively. The kinematic boundary condition takes the
following form:
h
u h
u
+
+ cos( ) = w.
(24)
t
r
r
In order to solve the complicated governing equations (13)(15), considered with the boundary
conditions (17)(24), we apply the following asymptotic expansions:
u = u (0) +
u (1) + ,
w = w (0) +
w (1) + ,
p =
1 p0 + p (0) +
p (1) + .

(25)

We substitute (25) in Eqs. (13)(24) and then equate equal powers of


to yield a sequence of linear
problems. Subsequently, we solve these problems for leading-order solutions u (0) , w (0) , and p(0) and
first-order correction u(1) as a function of h. Only u(1) is of interest so corrections w (1) and p(1) are
not calculated. In order to retain surface-tension effects, we are considering O(
) terms along with
leading-order terms. Note that p0 is a constant, and its value obtained from Eqs. (19) and (22) is
Bo1 .
At leading order, the governing equations are


MWt
p (0)
= W 2 ,
sin +
(26)

2


2 u (0)
MWt
= 0,

cos

+
2

(27)

w(0)
u (0)
u (0)
+

cos( ) = 0.
z

(28)

u (0) = w (0) = 0,

(29)

u (0)
= 0,
z

(30)

At z = 0, we have
and at z = h, we have

p (0) =

2h
1
(h + 2 + sin( ) 2 sin( )).
Bo

Integrating (26)(28) and using (29)(31) gives leading-order solutions p(0) , u(0) , and w (0)



1
2h
MWt
(0)
2
(z

h)

(h
+
p = W sin +
+ sin( ) 2 sin( )),

2
Bo
2



 2
MWt
z
hz cos +
,
u (0) =
2

2




 3

1 h 2
MWt
z
MWt
hz 2
w (0) =
+
sin +
cos

z
6
2

2
2

2
 2



z
MWt
+
cos( ).
hz cos +
2

(31)

(32)

(33)

(34)

Note that the leading-order component of the velocity along the cylinder surface (u(0) ) has no surface
topography component and it is similar to the expression obtained by Evans et al.4 for a smooth
cylinder.

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 14.139.208.84 On: Wed, 13 Apr
2016 11:51:52

042102-6

A. K. Sahu and S. Kumar

Phys. Fluids 26, 042102 (2014)

At the next order the governing equations are


2 w (0)
p (1)
+
= W 2 z,
z
z 2

(35)

p (0)
p (0)
2 u (1)
u (0)
+
cos( ) +
= 0,
+

z
z 2
z

(36)

(w(0) z) w (1)
u (1)
u (1)
+
+

cos( ) = 0.
z
z

(37)

u (1) = w (1) = 0,

(38)

At z = 0, we have
and at z = h, we have
u (1)
u (0) = 0.
z
Integrating (36) twice and using (38) and (39) gives
 3


z
MWt
3 2
2

u (1) = cos +

z
+
hz
h

2
3
2


 2
3h
1 h
z
+ 3 + cos( ) 3 cos( )
hz

Bo

2
 

 2


h
z
MWt
W 2 sin +
+
2
cos(
)

hz
.

(39)

(40)

Combining the continuity equation, kinematic condition, and Leibnizs rule yields the relationship
(1 +
h)

h
+
t

u (0) +
u (1) dz = 0.

(41)

Substituting Eqs. (33) and (40) into Eq. (41), we obtain a nonlinear evolution equation for the film
thickness




 4


h3
MWt
h
MWt
h
+
=
cos +

cos

+
[1 +
(h + sin( ))]
t
3

2
 3 

h
h
3h

3
+ 3 + cos( ) cos( )

3Bo

 
 3


h
MWt
h
W 2 sin +
+
2
cos(
)
. (42)

Following Evans et al.,4 to simplify the numerical scheme we dropped small terms of order

and rescaled all lengths including film thickness with Rmean . With these changes Eq. (42) becomes




 h
h3  2
h3
h3
h
3h
h
=
cos ( + M W t)
+ 3
W sin ( + M W t)
+
t
3
3Bo



 h3  2


h3 
3

cos( ) cos( )
W sin ( + M W t) 2 cos( ) .

3Bo
3
(43)
This simplified equation describes the behavior of a thin film on a rotating patterned cylinder in a
rotating frame of reference when surface tension, gravity, viscosity, and centrifugal forces are all
important. An analogous equation in a stationary frame of reference can be obtained by using the

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 14.139.208.84 On: Wed, 13 Apr
2016 11:51:52

042102-7

A. K. Sahu and S. Kumar

Phys. Fluids 26, 042102 (2014)

non-dimensional angular position in the fixed frame, f = +


find derivatives. The equation obtained is

MWt
,

and applying the chain rule to

 3



 h
h
h3  2
h
h
h

h3
3h
+
= M W
cos f
W sin f
+
+ 3
t
f
f 3
3Bo f
3
f
f




h3 
cos ( ) 3 cos ( (f M W t)) +

f
3Bo
 3


h  2

W sin f 2 sin ( (f M W t)) .


(44)
f
3
This equation reduces to the evolution equation presented by Evans et al.4 if written for a smooth
cylinder ( = 0). We note that similar equations can be derived for the case of a liquid film on the
inside of a rotating cylinder.7, 8
The behavior of the film is determined by the parameters , , Bo, M, W , and
. The maximum
values of and that are consistent with the lubrication approximation are discussed in Sec. III. As
did Evans et al.,4 we assume that W/M is no larger than O(1) in order to neglect Coriolis forces.
Typical values of W , M, and Bo can be found in Ref. 4. The value of
serves to set the initial film
thickness, h 0 =
Rmean .
For a fixed set of problem parameters, numerical solution of Eq. (43) provides the film thickness
around the cylinder at a given instant of time in a reference frame rotating with the cylinder. To
obtain the interface shape in a stationary frame at a given time, the term M W t needs to be added to
the rotating angular coordinate of the numerical solution of Eq. (43).
We followed the finite-difference-based semi-implicit numerical method of Evans et al.4, 9 to
numerically solve the nonlinear evolution equation (43). The equation is discretized in an explicit
manner with the exception of the highest (fourth-order) derivative term. This semi-implicit discretization allows larger time steps to be taken while maintaining numerical stability. A uniform
distribution of 400 grid points around the cylinder is used. The use of a larger number of grid points
did not produce significant changes in the film thickness. Further details about the numerical method
can be found in Refs. 4 and 9.

III. RESULTS AND DISCUSSION


A. Stationary cylinder

In order to determine the parameter regime in which the lubrication approximation is expected
to valid, we first consider drainage of an initially uniform film over a stationary cylinder. Physically,
liquid inside the film drains off the exterior of the substrate under the influence of gravity and is
restrained by surface tension and viscous forces. The liquid flow primarily depends on the Bond
number (Bo), with the film draining at large Bond number but forming a static and symmetric
drop beneath the cylinder at small Bond number. The dynamics of the evolving film are governed
by Eq. (43) with zero rotational velocity (W = 0). By comparing the final drop shape from the
dynamic calculation (which uses an approximate curvature based on lubrication theory) with the
drop shape obtained from a static calculation that uses the exact curvature, we can determine when
the lubrication approximation is expected to hold.
The free-surface shape of a static drop at low Bond number can be obtained by a shooting
method.4 Because we are considering static situations, the scalings chosen in Sec. II are no longer
appropriate and it is convenient to revert to dimensional variables. (Alternatively, the equations could
be rescaled.) Following Evans et al.,4 we define zero altitude (x = 0) at the lowest point of the drop
and measure positive altitude x upward. The dimensional static pressure at any point inside the drop
at altitude x is written as
p(x) = p0 gx,

(45)

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 14.139.208.84 On: Wed, 13 Apr
2016 11:51:52

042102-8

A. K. Sahu and S. Kumar

Phys. Fluids 26, 042102 (2014)

where p0 is the pressure in the liquid at the lowest point of the drop. At equilibrium, the static
pressure at altitude x is equal to the capillary pressure ( ) just inside the drop surface at the same
altitude.
On replacing the static pressure by the capillary pressure, the above equation becomes
(x) = 0 gx,

(46)

where (x) and 0 are the free-surface curvatures at x and zero altitude. The altitude of the free
surface x( ) for drop height H0 is (Rmean + sin( ))(1 + sin ) + h sin + H0 . The dimensional
free-surface curvature () is described by




2
1 
1 1
1 + 2 ( cos( ) + h )2 2 h 2 sin( ) ,
= 3
(47)
r
r
N r

where r = Rmean + sin( ) + h is the radius of the drop, N = 1 + r12 ( cos( ) + h )2 , and
-subscripts denote derivatives. The corresponding dimensional form of the approximate curvature
(i.e., the form used in lubrication theory) is
approx =

1
Rmean

h + sin( ) h 2 sin( )

.
2
2
Rmean
Rmean

(48)

By substituting Eq. (47) or (48) into Eq. (46) we obtain a second-order ordinary differential
equation (ODE) for the radius of the drop or film thickness.
The ODE can be made non-dimensional
again by scaling all lengths using the capillary length lc = /g. This yields


 2 3/2
 2 
r
r
r = r 1 + 2
r 2 (0 (Rmean + sin( )) r sin H0 ) 1 +
r
r
(full curvature),

(49)

h =Rmean (h + sin( ))
2
(0 (Rmean + sin( )) (Rmean + sin( ) + h) sin H0 )
Rmean

(approximate curvature).

(50)

These equations are integrated (for fixed values of H0 , , and ) by using a fourth-order RungeKutta shooting method from = 3 /2 where r = 0 and r = Rmean + sin( ) + H0 to an a
priori unknown point l at which r = Rmean + sin( ). A spatial step of 2 /10 000 was found to
be sufficient to give accurate results. For given , Rmean , and H0 , the initial curvature is set by the
requirement that the free surface meets the substrate at zero contact angle. The value of 0 is varied
until the slope of the free surface obtained from integration is equal to the slope of the substrate at
the meeting point and r = Rmean + sin( ) (i.e., the drop radius equals the cylinder radius) for the
full-curvature case or h = 0 for the approximate-curvature case. If 0 is not chosen properly, the
free surface obtained from integration either does not meet the solid substrate or meets it at non-zero
contact angle.
To facilitate comparison of solutions from the dynamic and static models, we calculate the initial
uniform film thickness (h0 ) which on drainage yields a steady-state drop of the same area as a static
drop. The non-dimensional area of the liquid contained in a static drop is
l
 2

r (Rmean + sin( ))2 d.
(51)
A=
3/2

The required initial uniform film thickness is


h0
= 1 +
Rmean


1+

A
.
2
Rmean

(52)

This expression is exact for both smooth and patterned cylinders. It is used as an initial condition in
the dynamic simulation.

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 14.139.208.84 On: Wed, 13 Apr
2016 11:51:52

042102-9

A. K. Sahu and S. Kumar

Phys. Fluids 26, 042102 (2014)

(a)

Dynamic model
Static model (approx. curvature)
Static model (full curvature)

(b)

8.5

8.5

10

10

11.5

11.5

13

13
3

1.5

1.5

1.5

1.5

1.5

1.5

(c)

(d)

8.5

10

11.5

13
2

FIG. 3. Comparison of dynamic and static model solutions for a patterned cylinder with (a) = 11, = 0.01Rmean ,
Rmean = 10lc , H0 = 0.5lc , (b) = 11, = 0.01Rmean , Rmean = 10lc , H0 = 1lc , (c) = 11, = 0.01Rmean , Rmean = lc , H0
= 0.5lc , (d) = 21, = 0.01Rmean , Rmean = 10lc , H0 = 0.5lc . Here, , Rmean , and lc are dimensional.

Figure 3 shows a comparison between static and dynamic solutions at different pattern frequencies for different mean cylinder radii (Rmean ) and drop heights (H0 ). To facilitate comparison between
the solutions obtained from the static and dynamic models, we express all lengths in terms of lc and
Rmean . The dimensional amplitude of the sinusoidal pattern () is set to 0.01Rmean . Simulations of
the dynamic model are performed with non-dimensional rotation speed W = 0, and Bo = 100 and
1. Here, Bo is obtained from the mean cylinder radius Rmean = 10lc and lc .
For = 11, Figs. 3(a) and 3(b) show that the static solution using the full curvature and
approximate curvature agree well for small drop height or low film thickness, but agreement becomes worse with an increase in drop height or film thickness. This indicates that the curvature
approximation begins to fail as film thickness increases. In both figures, the dynamic solution is in
good agreement with the approximate-curvature static solution. In Fig. 3(c), for a relatively thick
film (H0 = 0.5Rmean ), the dynamic solution and approximate-curvature static solution do not agree
well with the full-curvature static solution. This suggests that simulations with large drop height or
film thickness are beyond the capability of the present dynamic model. For = 21 (Fig. 3(d)), the
maximum film thickness is very small as in Fig. 3(a) but agreement between all three results is poor
due to the increased value of . At high frequencies, the pattern wavelength (2 Rmean /) becomes
comparable to the film thickness (h) and this imposes the restriction H  Rmean for the lubrication
approximation to be valid.
We have also performed calculations similar to those described above to examine the influence
of , although we do not show these for brevity. The results of these runs indicate that the lubrication
approximation will be valid for 0.01Rmean . So, based on the above observations, we present
results for 11 and 0.01Rmean since in this range the lubrication approximation is expected

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 14.139.208.84 On: Wed, 13 Apr
2016 11:51:52

042102-10

A. K. Sahu and S. Kumar

Phys. Fluids 26, 042102 (2014)

1.5

1.5

(a)

(b)

S = 50, = 0

0.5

0.5

0.5

0.5

1.5

S = 30, = 11

1.5
1.5

0.5

0.5

1.5

1.5

1.5

0.5

0.5

1.5

1.5

(c)

(d)

S = 50, = 11

0.5

0.5

0.5

0.5

1.5

S = 80, = 11

1.5
1.5

0.5

0.5

1.5

1.5

0.5

0.5

1.5

FIG. 4. Steady-state film profiles for rotation in absence of gravity. (a) Smooth cylinder with S = 50, (b) = 11 and S = 30;
drops form over troughs, (c) = 11 and S = 50, and (d) = 11 and S = 80; drops form over crests. For these calculations,
= 0.01. (For ease of visualization, the film thickness shown is five times larger than the actual value.)

to yield accurate predictions. Finally, we note that for a fixed mass, we have not seen evidence of
there being more than one stationary solution.

B. Rotating cylinder

We first study film evolution around a rapidly spinning cylinder. To perform simulations in this
regime, we neglect the gravitational term in Eq. (43). Because Eq. (43) was non-dimensionalized
using gravity, it is convenient to revert to the dimensional form of Eq. (43) (alternatively, the equation
could be rescaled)


h
 3
=
h h + h + Sh + cos( ) 3 cos( ) + S2 cos( ) ,
4
t
3Rmean
(53)
3
/ is the Weber number. Evans et al.4 performed a linear stability analysis
where S = 2 Rmean
for a smooth cylinder. They found that a film becomes unstable with a growth rate and cutoff wave
number given by
gr = ki +


h 30 
(1 + S) k 2 k 4 ,
4
3R

kcutoff =  1 + S,

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 14.139.208.84 On: Wed, 13 Apr
2016 11:51:52

042102-11

A. K. Sahu and S. Kumar

Phys. Fluids 26, 042102 (2014)

TABLE I. Critical values of S for = 0.01 and different values of .

Max value of S for


drops over troughs

Min value of S for


drops over crests

2
4
6
8
11

1
5
10
16
30

2
10
30
40
78

where x is the greatest integer less than x. Perturbations with wave numbers larger than kcutoff are
damped and those with k < kcutoff grow in amplitude. The fastest-growing wave number is


1+S
kmax =
.
(54)
2
The instability arises from centrifugal forces and leads to formation of drops around the circumference of the cylinder. For patterned cylinders, a film of uniform thickness is not a steady base state,
so a standard linear stability analysis cannot be applied to develop a dispersion relation. We instead
examine the case of patterned cylinders with numerical simulations as described below.
1.5E-05

0.015

Q smooth

(a)

Q topography
Solid boundary

0.01

5E-06

0.005

1E-05

-5E-06

-0.005

-1E-05

-0.01

-1.5E-05

5.4

5.6

5.8

-0.015

0.015

1.5E-05

1E-05

0.01

5E-06

0.005

(b)

-5E-06

-0.005

-1E-05

-0.01

-1.5E-05

5.4

5.6

5.8

-0.015

FIG. 5. Contributions to liquid flow rate after drop development is complete (a) S = 30 and (b) S = 80.

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 14.139.208.84 On: Wed, 13 Apr
2016 11:51:52

042102-12

A. K. Sahu and S. Kumar

6E-06

Phys. Fluids 26, 042102 (2014)

6E-06

Q smooth
Q topography

(a)
4E-06

Trough

(b)
4E-06

Trough

2E-06

2E-06

Time
0

-4E-06

-6E-06
6E-06

-6E-06
6E-06

4E-06

4E-06

2E-06

2E-06

-2E-06

-4E-06

-2E-06

-4E-06

-4E-06

-6E-06

-6E-06

6E-06

6E-06

4E-06

4E-06

2E-06

2E-06

-2E-06

-2E-06

-2E-06

-2E-06

-4E-06

-4E-06

-6E-06

5.3

5.4

5.5

5.6

5.7

5.8

-6E-06

5.3

5.4

5.5

5.6

5.7

5.8

FIG. 6. Contributions to liquid flow rate during the transient states (a) S = 30 and (b) S = 80.

For numerical simulations, it is convenient to use the dimensionless version of Eq. (53), where
we scale time by Rmean / . Thus, the non-dimensional form of Eq. (53) is

h
1  3
=
h h + h + Sh + cos( ) 3 cos( ) + S2 cos( ) .
(55)
t
3
We performed simulations with = 0.01 and two initial film conditions: (i) uniform thickness
h( , 0) = h0 and (ii) constant radius r (, 0) = Rmean + h 0 . Both conditions result in the same steadystate film profiles. For a smooth cylinder with S = 50, the fastest growing wave number is five and
Fig. 4(a) shows that five drops develop around the circumference of the cylinder. Our results show
that similar phenomena occur in the case of a patterned cylinder but the number of drops is equal
to the frequency of the sinusoidal pattern, , at low and high rotation rates . However, for a fixed
value of , where the drops are positioned depends upon the value of S. Steady-state film profiles in
rotating coordinates are shown in Figs. 4(b)4(d) for = 11 with S = 30, 50, and 80. Figure 4(b)
shows that drops develop over the troughs for S = 30. When the initial film is of uniform thickness,
higher capillary pressure at the crests drives flow toward the troughs and drops develop over the
troughs. If the initial film is thick over the troughs and thin over the crests (constant radius film),
higher centrifugal forces on the liquid over the troughs draws liquid from surrounding areas and
produces drops on the troughs.
A further increase in S ends the equality between the number of drops and and results in a
steady-state profile shown in Fig. 4(c). The equality occurs again if S is higher than a critical value,
as seen in Fig. 4(d) for S = 80. Furthermore, at these higher values of S the position of the drops
shifts from the troughs to the crests. Critical values of S where drop development on troughs ends
and development over crests begin are given in Table I. The range of S where drops shift from the
troughs to the crests increases with an increase in .
In order to investigate the mechanism for the shifting of drops at high values of S, we examine
the different contributions to the liquid flow rate during the film evolution and at steady state. In
Eq. (53), the first two terms are contributions due to surface tension and the third term is due to

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 14.139.208.84 On: Wed, 13 Apr
2016 11:51:52

042102-13

A. K. Sahu and S. Kumar

Phys. Fluids 26, 042102 (2014)

1.5

1.5

(a)

(b)

=1e-02

0.5

0.5

0.5

0.5

1.5

=1e-03

1.5
1.5

0.5

0.5

1.5

1.5

0.5

0.5

1.5

1.5

(c)

=1e-04

0.5

0.5

1.5
1.5

0.5

0.5

1.5

FIG. 7. Effect of pattern amplitude on steady-state film profiles for a cylinder with = 11, S = 100 and (a) = 1 102 ,
(b) = 1 103 , and (c) = 1 104 .

rotation. The rest of the terms within the square brackets are contributions due to surface topography
(Q topography ). Since the surface-topography contribution would be zero for a smooth cylinder we
refer to the combined contribution from the first three terms as the smooth-cylinder contribution
(Q smooth ). These two flow contributions at steady state between two consecutive crests on the cylinder
are shown in Fig. 5 for = 11, and S = 30 and 80. Figures 5(a) and 5(b) show that Q topography and
Q smooth are equal and opposite in direction at all positions and hence there is no net flow at steady
state as expected.
During the transient period |Q topography | is always higher than |Q smooth | (Fig. 6), and this results
in net flow in the same direction as that of the topography contribution (Q topography ). For S = 30
(Figs. 5(a) and 6(a)), the topography contribution is positive on the left of the trough ( = 5.57) and
negative on the right of the trough. The positive topography contribution on the left of the trough
drives flow in the positive -direction (toward the right in the figure) and similarly the negative
topography contribution on the right drives flow in the negative -direction (toward the left in the
figure). Consequently, liquid over the crests flows toward the troughs and drops develop on the
troughs connected by very thin film on the crests. At high S values, the signs of Q topography and
Q smooth get interchanged, as seen in Figs. 5(b) and 6(b) for S = 80. In addition, the region where
the flow rate contributions are non-zero shifts relative to the position of the trough. Therefore, liquid
over the troughs flows toward the crests, leading to the development of drops on the crests connected
by very thin films on the troughs.
In all simulations presented above, the dimensional amplitude is kept constant at 0.01Rmean .
Figure 7 shows steady-state film profiles for a cylinder with = 11, S = 100, and different values
of dimensionless amplitude = 1 102 , 1 103 , and 1 104 . It can be seenhere that the
number of drops for = 1 102 is equal to 11 () whereas it is reduced to 7 ([

S+1
])
2

like a

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 14.139.208.84 On: Wed, 13 Apr
2016 11:51:52

042102-14

A. K. Sahu and S. Kumar

Phys. Fluids 26, 042102 (2014)

TABLE II. Lowest value of for different and S to invoke topography


effects.

S = 20

S = 100

6
7
8
9
10
11

1 102
1 102
1 102
1 102
1 102
1 102

1 1015
1 1014
1 1013
1 109
1 104
1 104

4
3
smooth
 cylinder for = 1 10 . At intermediate = 1 10 , the number of drops is neither
nor [ S+1
]. The lowest values of to show topography effects for different values of and S are
2
presented in Table II. The data indicate that the topography need only have a very small amplitude
to affect the film dynamics, especially at high rotation rates and pattern wavelengths (low ). Thus,
the drop formation process may be very sensitive to the presence of surface topography.
We have also performed some additional calculations using different initial conditions. For
brevity, we simply summarize the results here. If an initial condition is used in which the number
of droplets is different from the pattern frequency, then the final number of droplets is still equal to
the pattern frequency, with the droplets located at the troughs at low rotation rates and at the crests
at high rotation rates. This was observed to be true whether the droplets were initially located at the
troughs or crests. We have also observed that if the droplets are initially of unequal mass, then in
the final state they are also of unequal mass, but their number and location in the final state behave

1.5

1.5

(a)

(b)

W = 0.002

0.5

0.5

0.5

0.5

W = 0.004

1.5

1.5
1.5

0.5

0.5

1.5

1.5

1.5

0.5

0.5

1.5

1.5

(c)

(d)

W = 0.008

0.5

0.5

0.5

0.5

1.5

W = 0.016

1.5
1.5

0.5

0.5

1.5

1.5

0.5

0.5

1.5

FIG. 8. Unsteady film profile at t = 1.5T for (a) W = 0.002, (b) W = 0.004, (c) W = 0.008, and (d) W = 0.016. Here,
= 11 and = 0.01.

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 14.139.208.84 On: Wed, 13 Apr
2016 11:51:52

042102-15

A. K. Sahu and S. Kumar

Phys. Fluids 26, 042102 (2014)

0.011

(a)
0.01

h(3/2)

0.009
0.008
0.007
0.006
0.005
0.004
0.0e+00

3.0e+05

6.0e+05

9.0e+05

time
0.011

(b)
0.01

h(3/2)

0.009
0.008
0.007
0.006
0.005
0.004
0.0e+00

3.0e+05

6.0e+05

9.0e+05

time

FIG. 9. Oscillations in film thickness with time at = 3 /2 for W = 0.008 and (a) = 11 and (b) = 8. Here, = 0.01.

in the way described above. We were not able to determine whether the droplets eventually reach a
state where they are all of equal size, presumably because the thin films connecting them lead to a
very slow mass redistribution.
Next, we consider the effect of gravity with all of the other important effects in Eq. (43).
Following Evans et al.,4 we fix M = 0.007, Bo = 100, and h 0 = 0.007Rmean (dimensional), and
focus on variation in the film profile with a change in rotation speed W and pattern frequency
. For a smooth cylinder, the film thickness reaches a steady state within a few revolutions when
W 0.014. At higher speeds traveling waves appear in the film. However, these waves get stabilized
by surface tension and the film thickness eventually reaches a steady state. Our results show that
for a patterned cylinder, the film thickness never reaches a steady state. The non-dimensional times
for one revolution (T) at W = 0.002, 0.004, 0.008, and 0.016 are 448 800, 224 400, 112 200
and 56 100, respectively. The unsteady film profiles at t = 1.5T for a cylinder with = 11 and
= 0.01Rmean for W = 0.002, 0.004, 0.008, and 0.016 in a stationary frame are shown in Fig. 8.
For W = 0.002, 0.004, and 0.008, a drop forms quickly in the first revolution on the upward moving
side (3 /2 < 2 ) and continuously jiggles due to the pattern over the surface. On an increase
in W from 0.002 to 0.008, the drop diminishes in size and moves toward the upward moving side,
as seen in Figs. 8(a)8(c).
Figure 8(d) shows that at higher speed, transport of liquid around the cylinder makes the film
thickness nearly uniform. For W = 0.016 (Fig. 8(d)), the dimensional film thickness at = /2
and 3 /2 in a stationary frame varies over time from 0.0056Rmean to 0.0084Rmean and 0.0054Rmean
to 0.0091Rmean , respectively. Since the time-averaged thickness is of the same order of magnitude
for all , the film appears nearly uniform at higher speeds. For a smooth cylinder rotating at the
same speed, film thickness at both positions varies from 0.00672Rmean to 0.00674Rmean after 100
revolutions. The variations in film thickness with time at = 3 /2 (in a stationary frame) for
W = 0.008, = 0.01, and = 8 and 11 are presented in Fig. 9. It is seen that the amplitude and

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 14.139.208.84 On: Wed, 13 Apr
2016 11:51:52

042102-16

A. K. Sahu and S. Kumar

Phys. Fluids 26, 042102 (2014)

the dominant frequency of the oscillations in film thickness increase with an increase in the pattern
frequency. We also performed a few simulations for W = O(1). Similar to the zero-gravity case,
drops begin to develop around the circumference of cylinder at high rotation rates.
IV. CONCLUSIONS

To begin addressing the general issue of how surface topography influences the flow of thin liquid
films on rotating surfaces, we have considered a model problem in which a thin liquid film flows over
a rotating cylinder patterned with a sinusoidal surface topography. When the cylinder is stationary,
calculations of static interface shapes provide guidance regarding the validity of the lubrication
approximation. When the cylinder is rotating, the behavior of the film is strongly influenced by the
strength of gravitational forces.
When gravitational forces are relatively weak, cylinder rotation leads to the formation of droplets
connected by very thin films. Remarkably, the number of droplets is equal to the pattern frequency at
low and high rotation rates, with the droplets located at the pattern troughs at low rotation rates and
the pattern crests at high rotation rates. Analysis of the flow rates in the films reveals how surface
topography generates additional flows that drive this shift. The larger the rotation rate, the smaller
the topographical features need to be in order to observe this behavior. When gravitational forces
become significant, the film thickness never reaches a steady state, in contrast to the case of an
unpatterned cylinder.
The results of this work demonstrate that the flow of thin liquid films on rotating surfaces can be
very sensitive to the presence of surface topography. Thus, surface topography will be an important
factor to account for when developing accurate models for the coating of discrete objects, particularly
those that are rotated at high speeds. In addition, it will be important to account for other complex
phenomena that are also important including solvent evaporation, non-Newtonian rheology, inertia,10
and more complicated geometries.1114 Our results also raise the important question of how surface
topography will influence (or could possibly be designed to influence) complex three-dimensional
effects such as the formation of rings, drops, and fingers.15
ACKNOWLEDGMENTS

This work was supported through the Industrial Partnership for Research in Interfacial and
Materials Engineering of the University of Minnesota.
1 Y.

M. Chen, U.S. patent 8,236,369 B2 (7 August 2012).


Pukhnachev, Motion of a liquid film on the surface of a rotating cylinder in a gravitational field, J. Appl. Mech. Tech.
Phys. 18, 344351 (1977).
3 H. K. Moffatt, Behaviour of viscous film on the outer surface of a rotating cylinder, J. Mech. 16, 651673 (1977).
4 P. L. Evans, L. W. Schwartz, and R. V. Roy, Steady and unsteady solutions for coating flow on a rotating horizontal
cylinder: Two-dimensional theoretical and numerical modeling, Phys. Fluids 16, 27422756 (2004).
5 R. V. Craster and O. K. Matar, Dynamics and stability of thin liquid films, Rev. Mod. Phys. 81, 11311198 (2009).
6 G. K. Batchelor, An Introduction to Fluid Dynamics (Cambridge University Press, Cambridge, UK, 1967).
7 S. B. G. OBrien, A mechanism for linear instability in two-dimensional rimming flow, Q. Appl. Math. 60, 283299
(2002).
8 E. S. Benilov, N. Kopteva, and S. B. G. OBrien, Does surface tension stabilize liquid films inside a rotating horizontal
cylinder, Q. J. Mech. Appl. Math. 58, 185200 (2005).
9 P. L. Evans, Mathematical and numerical investigation of coating flows, Ph.D. thesis, University of Delaware, 2000.
10 M. Chugunova and R. M. Taranets, Qualitative analysis of coating flows on a rotating horizontal cylinder, Int. J. Differ.
Eq. 2012, 570283 (2012).
11 R. V. Roy, A. J. Roberts, and M. E. Simpson, A lubrication model of coating flows over a curved substrate in space,
J. Fluid Mech. 454, 235261 (2002).
12 A. J. Roberts and Z. Li, An accurate and comprehensive model of thin fluid flows with inertia on curved substrates,
J. Fluid Mech. 553, 3373 (2006).
13 R. Hunt, Numerical solution of the free-surface viscous flow on a horizontal rotating elliptical cylinder, Numer. Methods
Partial Differ. Equ. 24, 10941114 (2008).
14 R. J. Braun, R. Usha, G. B. McFadden, T. A. Driscoll, L. P. Cook, and P. E. King-Smith, Thin film dynamics on a prolate
spheroid with application to the cornea, J. Eng. Math. 73, 121138 (2011).
15 P. L. Evans, L. W. Schwartz, and R. V. Roy, Three-dimensional solutions for coating flow on a rotating horizontal cylinder:
Theory and experiment, Phys. Fluids 17, 072102 (2005).
2 V.

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 14.139.208.84 On: Wed, 13 Apr
2016 11:51:52

You might also like