Вы находитесь на странице: 1из 19

Surface tension effects on the motion of a free-falling liquid sheet

Gennaro Coppola, Fortunato De Rosa, and Luigi de Luca



Citation: Phys. Fluids 25, 062103 (2013); doi: 10.1063/1.4810751
View online: http://dx.doi.org/10.1063/1.4810751
View Table of Contents: http://pof.aip.org/resource/1/PHFLE6/v25/i6
Published by the AIP Publishing LLC.

Additional information on Phys. Fluids
Journal Homepage: http://pof.aip.org/
Journal Information: http://pof.aip.org/about/about_the_journal
Top downloads: http://pof.aip.org/features/most_downloaded
Information for Authors: http://pof.aip.org/authors
Downloaded 22 Jun 2013 to 147.127.100.21. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
PHYSICS OF FLUIDS 25, 062103 (2013)
Surface tension effects on the motion of a free-falling
liquid sheet
Gennaro Coppola, Fortunato De Rosa, and Luigi de Luca
Dipartimento di Ingegneria Industriale, Universit ` a di Napoli Federico II,
80125 Napoli, Italy
(Received 13 July 2012; accepted 7 May 2013; published online 21 June 2013)
The stationary motion of a liquid curtain falling under the effects of inertia, gravity,
and surface tension is analyzed. An original equation governing the streamwise dis-
tribution of thickness and velocity is derived by means of a Taylor expansion in the
lateral distance from the mean line of the sheet. Approximate solutions are obtained
by means of perturbation approaches involving the two parameters governing the
problem, namely, the slenderness ratio and the Weber number We. The numer-
ical procedure employed in order to integrate the non-linear equation is discussed
and a parametric study is presented, together with a comparison with the approxi-
mate asymptotic solutions valid for small and We.
C
2013 AIP Publishing LLC.
[http://dx.doi.org/10.1063/1.4810751]
I. INTRODUCTION
The planar two-dimensional liquid sheet ow, falling under the action of gravity, is a subject of
both industrial and academic interests. From the applications viewpoint, liquid sheets are employed
in a variety of technological processes, most notably on the curtain coating process
1
in which the
falling curtain is employed to deposit uniform liquid lms onto a solid moving substrate. From
the purely scientic viewpoint, the problem constitutes a prototypical conguration of spatially
developing system, which is able to sustain a variety of wave patterns. Within a wide bulk of
literature papers, of great interest is the historical work of Pritchard,
2
who experimentally observed
the free-surface class of ows arising when the uid is poured over the end of a at plate into a
reservoir below the plate. Pritchard
2
observed hundreds of different ow regimes, some of which
were steady, some nearly periodic in time, some others exhibiting almost chaotic behavior in time.
The effect of surface-active agents added to the working liquid, another important aspect of the
problem, was in particular studied by de Luca and Meola,
3
who quantied such effects in terms of
surface pressure rather than the standard surface tension coefcient.
Both the stationary ow motion of the falling sheet and its stability have been theoretically
studied in the past. Concerning the steady-state ow, which is the subject of the present paper, the
rst contribution addressing both experimentally and theoretically the determination of the velocity
eld is due to Brown,
4
who correlated the experimentally measured streamwise ow velocity with
that attained by a body freely falling under gravity (free-fall or Torricellis velocity). He found that,
for other than a small distance from the inlet, the ow velocity agrees with a modied free-fall
law in which the inlet squared velocity is decreased by a factor proportional to
2/3
, being the
kinematic viscosity. In an Appendix to Browns paper,
4
Taylor empirically derived a differential
equation governing the streamwise velocity inside the sheet. Taylors treatment assumes that the
streamwise velocity and the pressure inside the sheet are independent of transverse coordinate and
focuses on the effects of viscous stresses on the ow, neglecting surface tension. Taylors equation
was subsequently derived also by Clarke
5
as a leading order equation for an outer expansion in the
streamwise direction at low Reynolds numbers. Clarke
5
found also an analytical solution to Taylors
equation in terms of Airy functions.
The effects of the surface tension on the shape of the falling liquid curtain have been considered
in later contributions, mainly as an additional effect to viscosity. Aidun
6
used a perturbation approach
1070-6631/2013/25(6)/062103/18/$30.00 C
2013 AIP Publishing LLC 25, 062103-1
Downloaded 22 Jun 2013 to 147.127.100.21. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
062103-2 Coppola, De Rosa, and de Luca Phys. Fluids 25, 062103 (2013)
to derive an equation for the streamwise velocity, which includes the effects of both viscosity and
surface tension. Ramos
7
derived a variety of leading order non-stationary equations of the full
problem including viscosity and surface tension effects at low Reynolds numbers. His analysis is
based on an asymptotic treatment in which the slenderness ratio (dened as the ratio of the thickness
of the sheet at the inlet to a characteristic axial length) is considered as a small parameter. Finnicum
et al.
8
studied the steady conguration of a two-dimensional liquid sheet falling in gravity in the
particular case in which a pressure difference is applied to the curtain. Their treatment fully takes
into account surface tension effects for an inviscid sheet and uses the approximation that the ow is
locally of plug type.
In the context of the study of the stability of the falling curtain, some other basic ow solutions,
which include surface tension effects, have been proposed. de Luca and Costa
9
developed a two-
dimensional potential solution by introducing a multiple scale approach and a power series expansion
in terms of the slenderness parameter. They were able to obtain a second order correction of the
base gravity-inertial solution that takes into account both streamlines curvature and surface tension
effects. The same assumption of potential owhas been employed by Weinstein et al.,
10
who derived
approximate equations governing the time-dependent evolution of disturbances on a falling liquid
curtain subject to pressure disturbances. The treatment they developed includes surface tension
effects and is valid for slender two-dimensional sheets. As a rst step in their derivation, a steady-
state base ow is determined, in which, as assumed by de Luca and Costa,
9
the modications to the
free-fall solution are expressed as additional higher order terms in the asymptotic expansion in the
slenderness parameter.
In addition to the literature cited above, there is also a very large number of papers yielding
the steady ow eld of the liquid sheet by means of direct numerical simulation. Since the present
paper deals with the problem from a different perspective, the relevant numerical literature is not
reviewed here. The interested reader is referred to Refs. 11 and 12 (where different techniques were
employed) and to the literature cited therein.
The present paper presents a model governing the streamwise distribution of thickness and
velocity of a liquid curtain, essentially based on a Taylor expansion in the lateral distance from
the mean line of the sheet. The model focuses on the effects of surface tension while neglecting
viscous stresses, as well as the centrifugal effects produced by the curvature of the streamlines.
Although based on a truncated Taylor expansion in the transverse coordinate (which is equivalent to
the assumption of plug-type velocity and pressure distributions inside the sheet, and hence implicitly
assumes small values of the sheet thickness), the model accounts fully for the slenderness ratio as a
free parameter. Recent contributions to the falling sheet global stability (Schmid and Henningson
13
and Coppola and de Luca
14
) use the free-fall model, which is valid for slender sheets or in absence
of surface tension. Thus, an additional motivation of the present work is to furnish a solution to the
base ow of higher order than the free-fall one.
The paper is organized as follows. In Sec. II, the problem is set up and the approximate
one-dimensional equations governing the steady distribution of ow velocity and sheet thickness
along the streamwise direction are derived. Asymptotic approximations in the limits of small Weber
number and
2
are presented in Sec. III. The numerical treatment of the equation is discussed in
Sec. IV, together with some parametric results, while conclusions are presented in Sec.V.
II. THE MODEL
A. Governing equations
The governing equations describing the two-dimensional plane, steady evolution of an incom-
pressible and inviscid liquid sheet falling in the gravitational eld are (Euler equations)
u

+
v

= 0, (1)
u

+v

=
1

+ g, (2)
Downloaded 22 Jun 2013 to 147.127.100.21. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
062103-3 Coppola, De Rosa, and de Luca Phys. Fluids 25, 062103 (2013)
u

+v

=
1

. (3)
In these equations u

and v

are the (dimensional) vertical and transverse velocity components along


the cartesian coordinates x

and y

, respectively, p

and are pressure and (constant) density in the


liquid, and g is the gravitational acceleration. Kinematic and dynamic boundary conditions at the
two interfaces between liquid and gas are
v

= u

dY

dx

, (4)
p

= P
a


d
2
Y

dx
2
_
1 +
_
dY

dx

_
2
_
3/2
, (5)
where is the surface tension coefcient, Y

(x

) and Y

+
(x

) are the equations of, respectively, the


left and right boundaries of the sheet, P
a

and P
a
+
are the local values of the external (gas) pressure,
respectively, on the left and on the right side of the sheet, and the notation

(x) = lim
yY

(x)
(x, y)
is employed, denoting any of the variables u

, v

, and p

. We also dene the quantities H

and Y

as the thickness and the centreline position of the sheet,


H

(x

) = Y

+
(x

) Y

(x

),
Y

(x

) =
Y

+
(x

) +Y

(x

)
2
,
for which the following relation is easily derived:
Y

=
H

2
.
In Fig. 1, all the quantities introduced are illustrated in a sketch of the geometrical conguration
under study. Note that the external gas pressures P
a

may be, in more general unsteady situations,


local functions of the deections of the boundaries Y

of the sheet (and of their derivatives) on the


whole domain. In this paper, the pressure eld will be assumed constant on each side of the sheet.
The equations of motion (1)(5) can be conveniently rewritten in non-dimensional form by
employing the following reference quantities:
x
r
= U
2
in
/g, y
r
= H
in
, p
r
= U
2
in
,
u
r
= U
in
, v
r
= U
in
,
where U
in
and H
in
are, respectively, the average streamwise velocity and sheet thickness at the inlet
of the sheet, and = H
in
/x
r
= H
in
g/U
2
in
is the slenderness parameter. Equations (1)(5) are written
accordingly in nondimensional variables u, v, p, Y

, and P
a

as
u
x
+
v
y
= 0, (6)
u
u
x
+v
u
y
=
p
x
+1, (7)
u
v
x
+v
v
y
=
1

2
p
y
, (8)
v

= u

x
, (9)
Downloaded 22 Jun 2013 to 147.127.100.21. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
062103-4 Coppola, De Rosa, and de Luca Phys. Fluids 25, 062103 (2013)
y
x
Y
+
Y

P
a
+
P
a

Y
H
FIG. 1. Sketch of the geometrical conguration of the sheet.
p

= P
a

We
2
d
2
Y

dx
2
_
1 +
2
_
dY

dx
_
2
_
3/2
, (10)
where We = /U
2
in
H
in
is the Weber number. Note that the present denition of the Weber number
differs from that generally adopted (see, for example, Finnicum et al.
8
and de Luca and Costa
9
),
which considers the ratio of the inertial termto the surface tension contribution (calculated on the half
thickness of the sheet). The present denition is consistent, on the other hand, with that employed
by Weinstein et al.,
10
who obtained a slender solution briey reconsidered in Sec. III A. Note also
that the slenderness ratio can be interpreted as the reciprocal of the Froude number.
The system of equations (6)(10) constitutes a nonlinear system of partial differential equations
in the unknown two-dimensional velocity and pressure elds inside the sheet, as well as in the
unknown shape of the sheet itself. In order to derive simplied models for the streamwise variations
of the thickness, vertical velocity and mean line, several perturbation schemes have been employed
in previous works, mainly within the context of linear stability analysis and by considering the more
general unsteady equations. For examples of such approaches, the reader is referred to the works
by Ramos,
7
de Luca and Costa,
9
Weinstein et al.,
10
and Ramos.
15
The perturbation approaches
employed are usually parameter-type expansions in which a parameter contained in the equations of
motion is considered as small and a Maclaurin series expansion of the ow variables is accordingly
assumed. The form of the system (6)(10) strongly suggests, for example, an expansion in the small
parameter
2
, at xed values of We, which is valid for slender sheet. A straightforward application
of this perturbation scheme to the system (6)(10) (not reported here) gives results that completely
agree with the analogous results obtained by Weinstein et al.
10
(briey reconsidered in Sec. III A),
who employed a model based on potential ow.
In this paper, the problem is treated from a different perspective, since the focus is on a
different asymptotic treatment of the equations of motions, which considers the lateral coordinate
y as the small parameter. This type of asymptotic treatment can be classied as a coordinate-type
expansion and differs in some important mathematical aspects from the previously considered
Downloaded 22 Jun 2013 to 147.127.100.21. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
062103-5 Coppola, De Rosa, and de Luca Phys. Fluids 25, 062103 (2013)
parameter-type expansion. The distinction between coordinate and parameter expansions has been
quite widely discussed in the past (see, for example, Van Dyke
16
and Chang
17
) and basically consists
in a different mathematical structure of the resulting perturbed equations, due to the fact that in the
equations of motion derivatives with respect to parameters do not occur, while all the derivatives
are taken with respect to time and space coordinates. It could be noted that in our problem the
nondimensional variable y and the slenderness ratio are strictly related, and this in principle means
that small values of y imply small values of . However, the two approaches are not equivalent,
since some mathematical differences are evident. In our case, for example, one of the differences
between the two approaches will be that in the coordinate expansion the lowest order solution
fully takes into account the complete expression for the curvature, expressed in Eq. (10), while a
parameter expansion considers simplied expressions until rst order in
2
. An approach similar to
that employed in the present paper has been exploited also by Eggers and Dupont
18
for the wider
problem of viscous and unsteady round jet. In that paper, it is noted that the full expression for the
curvature is an important element in the predictive strength of the nonlinear unsteady model.
The asymptotic treatment is conducted by considering a procedure similar to that employed by
Mehring and Sirignano
19
for the description of the linear and non-linear evolution of an incompress-
ible planar liquid sheet in zero gravity. In this approach, the thickness of the sheet H is considered
as the small parameter. Any dependent variable u, v, or p is accordingly expanded as a power series
in terms of the lateral distance from the mean line
_
y Y
_
:
u(x, y) = U
0
(x) +U
1
(x)
_
y Y
_
+U
2
(x)
_
y Y
_
2
. . . ,
v(x, y) = V
0
(x) + V
1
(x)
_
y Y
_
+ V
2
(x)
_
y Y
_
2
. . . ,
p(x, y) = P
0
(x) + P
1
(x)
_
y Y
_
+ P
2
(x)
_
y Y
_
2
. . . .
(11)
Such power series expansions are valid for y (Y

, Y
+
) and hence the inequality

y Y(x)

H/2
holds. By substituting positions (11) into Eqs. (6)(8), one obtains at zeroth order in (y Y):
dU
0
dx
= V
1
+U
1
dY
dx
, (12)
U
0
dU
0
dx
+U
1
K
0
= 1
dP
0
dx
+ P
1
dY
dx
, (13)
U
0
dV
0
dx
+ V
1
K
0
=
1

2
P
1
, (14)
where K
0
=
_
V
0
U
0
dY/dx
_
. These equations have to be coupled with the interface conditions
(4) and (5) approximated at rst order:
V
0
U
0
dY

dx

_
V
1
U
1
dY

dx
_
H
2
= 0,
P
0
P
a

_
We
2
F

d
2
Y

dx
2
+ P
1
H
2
_
= 0,
where
F

=
_
_
_
1 +
2
_
_
_
dY
dx
_
2

dY
dx
dH
dx
+
1
4
_
dH
dx
_
2
_
_
_
_
_
3/2
. (15)
A convenient manipulation yields the equivalent set of interface equations:
U
0
dY
dx
= V
0
, (16)
U
0
dH
dx
=
_
V
1
U
1
dY
dx
_
H, (17)
Downloaded 22 Jun 2013 to 147.127.100.21. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
062103-6 Coppola, De Rosa, and de Luca Phys. Fluids 25, 062103 (2013)
P
0
= P
a

We
2
2
_
F
d
2
H
dx
2
+

F
d
2
Y
dx
2
_
, (18)
P
1
H =

P
a
We
2
_

F
2
d
2
H
dx
2
+2F
d
2
Y
dx
2
_
, (19)
where for the variables F

and P
a

the following denitions have been introduced:


=
+

,
=

+
+

2
.
The term K
0
in Eqs. (13) and (14) is null by virtue of Eq. (16), and eliminating V
1
, P
0
, and P
1
from
Eqs. (12)(14) with the aid of Eqs. (17)(19) gives
d(U
0
H)
dx
= 0, (20)
U
0
dU
0
dx
= 1 +

P
a
H
dY
dx
+
We
2
2
d
dx
_
F
d
2
H
dx
2
+

F
d
2
Y
dx
2
_

We
2
H
dY
dx
_

F
2
d
2
H
dx
2
+2F
d
2
Y
dx
2
_
, (21)
U
0
dV
0
dx
=

P
a

2
H
+
We
H
_

F
2
d
2
H
dx
2
+2F
d
2
Y
dx
2
_
. (22)
Equations (16) and (20)(22) constitute a closed system of non-linear ordinary differential equations
in the unknowns, U
0
, V
0
, H, and Y. The external pressure distribution has been assumed known
and constant on each side of the sheet and appears in these equations through the difference

P
a
= P
a
+
P
a

.
B. Unsymmetric conguration
The system of equations (16) and (20)(22) will be hereafter employed in order to derive the
equations governing the steady conguration of a falling liquid sheet for the case in which a pressure
difference between the two sides is applied. This problem was extensively discussed by Finnicum
et al.
8
(essentially under the same assumptions made here) and the reference to their paper constitutes
a basic comparison for the present modelling.
Let us rst rewrite the present system of equations in terms of the streamwise velocity V
s
, which
is related to U
0
and V
0
by V
2
s
= U
2
0
+
2
V
2
0
. By employing Eq. (16) in order to eliminate U
0
in
Eq. (22) and by summing the resulting equation to Eq. (21) we obtain
V
s
dV
s
dx
= 1 +
We
2
2
d
dx
_
F
d
2
H
dx
2
+

F
d
2
Y
dx
2
_
. (23)
A second equation can be derived by eliminating this time V
0
in Eq. (22) through Eq. (16), thus
obtaining
d
dx
_
U
0
dY
dx
_
We
2
_

F
2
d
2
H
dx
2
+2F
d
2
Y
dx
2
_
=

P
a
, (24)
where the relation U
0
H = 1 has been employed for the rst approximation of the volumetric ow
rate per unit width, which is a constant quantity on the grounds of the continuity equation (20).
Hereafter, we will furthermore make the assumption that the thickness of the sheet varies very
slowly along the vertical direction, as made by Finnicum et al.
8
in their derivation. This implies
that the rst and second derivatives of H can be neglected in the relationship (15) as well as in
Downloaded 22 Jun 2013 to 147.127.100.21. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
062103-7 Coppola, De Rosa, and de Luca Phys. Fluids 25, 062103 (2013)
Eqs. (23) and (24). From these assumptions, we have F =
_
1 +(dY/dx)
2
_
3/2
and

F = 0, and
Eqs. (23) and (24) reduce to
V
s
dV
s
dx
= 1, (25)
d
dx
_
U
0
dY
dx
_
2We
2
_
_
_
d
2
Y/dx
2
_
1 +
2
_
dY/dx
_
2
_
3/2
_
_
_ =

P
a
.
After some manipulations and by employing the relation U
0
= V
s
_
1 +
2
(dY/dx)
2
_
1/2
, we are left
with
d
dx
_
_
_
_
_
_
dY/dx
_
1 +
2
_
dY/dx
_
2
_
1/2
_
_
_
_
V
s
2We
2
_
_

_ =

P
a
,
which is the nondimensional expression of Eq. (7) of Finnicum et al.
8
Coupled with the free-fall
solution for the streamwise velocity (25), it yields a single equation for the unknown deection of
the mean location of the curtain. This equation shows the presence of a singularity at the location x
where the local Weber number We
x
= /U
in
H
in
V
s
(x) = 1/2 (of course this singularity does not
appear if the liquid leaves the slot fromwhich the sheet originates with We
x
< 1/2). Finnicumet al.
8
studied the properties of the singularity both theoretically and experimentally and found that it can
be removed (physically the singularity does not occur) if the centreline of the curtain assumes a
specic, non-zero, slope (associated to the applied pressure difference) as it leaves the slot.
In the following Sec. II C, we shall consider the problem of a nite thickness liquid sheet falling
symmetrically in the gravitational eld.
C. Symmetric conguration
In the remaining of this paper we will assume that the external ambient gas pressures on the
two sides of the sheet are uniform and equal to each other. A direct consequence of this assumption
is that the value of the difference of ambient gas pressure

P
a
is zero, together with the already
assumed condition that the streamwise variation of the mean of the two gas pressures (dP
a
/dx) is
zero. As a consequence, the undisturbed shape of the sheet is symmetric with respect to the x-axis,
and hence the centreline location of the sheet is identically null: Y(x) = 0. Moreover, the transverse
velocity V
0
(x) has to be again zero for symmetry reasons. Hence, the non-null unknowns remain
the streamwise velocity U
0
and the thickness H, which are functions of the streamwise coordinate x
because of the gravitational acceleration.
In summary, symmetry considerations give the following form for the ow variables:
U
0
= U
0
(x), H = H(x), (26)
V
0
= 0, Y = 0,

P
a
= 0, (27)

F = 0, F = F =
_
1 +

2
4
_
dH
dx
_
2
_
3/2
. (28)
Substitution of these equations in the nonlinear governing system (16) and (20)(22) yields the
differential equations for U
0
and H. Equations (16) and (22) are in fact identically satised by
Eqs. (26)(28), and Eqs. (20) and (21) give
UH = 1, (29)
U
dU
dx
= 1 +
We
2
2
d
dx
_
F
d
2
H
dx
2
_
, (30)
Downloaded 22 Jun 2013 to 147.127.100.21. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
062103-8 Coppola, De Rosa, and de Luca Phys. Fluids 25, 062103 (2013)
where F is given by Eq. (28). Note that in these last equations and in the subsequent ones, the zeroth
order velocity component U
0
is denoted simply by U.
A rst analytical solution can be obtained when = 0, for which the well known Torricellis
free-fall solution holds
U(x) =

1 +2x, (31)
H(x) =
1

1 +2x
. (32)
Thus, the solution of Eqs. (29) and (30) may be considered as a correction of the free-fall model
when surface tension force is taken into account.
Equations (29) and (30) can also be derived on a physical basis from an elementary application
of integral mass and momentum balances to an innitesimal portion of the sheet of extension dx.
If one assumes that the velocity and pressure proles inside the sheet are uniform in the transversal
direction, integral mass conservation immediately gives Eq. (29). Momentum conservation dictates
that inertia is balanced by pressure and gravity forces, and can be expressed for constant proles of
velocity U

and pressure P

by means of the Bernoullis equation as


d
dx

_
P

+
U
2
2
gx

_
= 0. (33)
The pressure P

is derived from the assumption of equilibrium with external pressure and surface
tension contribution:
P

= P
a

_
_
_
d
2
Y

+
/dx
2
_
1 +
_
dY

+
/dx

_
2
_
3/2
_
_
_.
Substitution of this last equation into Eq. (33), under the assumption of constant external pressure,
nally gives
U

dU

dx


d
dx

_
_
_
_
_
_
_
1
2
d
2
H

dx
2
_
1 +
_
1
2
dH

dx

_
2
_
3/2
_

_
g = 0,
which is the dimensional version of Eq. (30).
III. ASYMPTOTIC APPROXIMATIONS
In this section, some simple analytical solutions are derived by applying a standard asymptotic
procedure to Eqs. (29) and (30) for the cases
2
1, We = O(1), and We 1, = O(1). These
solutions will be compared to the full numerical solution in Sec. IV.
A. Slender-sheet approximation
The formof Eqs. (29) and (30) suggests a perturbative approach in the small parameter
2
, which
can be developed in order to describe the evolution of slender sheets (the sheet will be considered
slender for 1).
By assuming a power series expansion of the ow variables in the small parameter
2
,
U = U

0
+
2
U

1
+
4
U

2
+. . . ,
H = H

0
+
2
H

1
+
4
H

2
+. . . ,
Downloaded 22 Jun 2013 to 147.127.100.21. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
062103-9 Coppola, De Rosa, and de Luca Phys. Fluids 25, 062103 (2013)
one obtains a hierarchy of equations at different orders by grouping terms of same power of
2
. The
lowest order approximation gives
U

0
H

0
= 1, (34)
U

0
dU

0
dx
= 1, (35)
whose solution uncovers the non-dimensional free-fall behavior, (31) and (32). Hence, the free-fall
solution, which is an exact solution of the present model in the case of absence of surface tension,
is also the lowest order solution to a perturbation approach in the small parameter
2
, i.e., it is the
asymptotic solution valid for slender sheets. By considering terms of order
2
in the expansion of U
and H, one obtains the rst order system:
U

1
H

0
+U

0
H

1
= 0, (36)
U

1
dU

0
dx
+U

0
dU

1
dx
=
We
2
d
dx
_
d
2
H

0
dx
2
_
. (37)
By employing Eqs. (34) and (35), Eq. (37) reduces to
U

1
U

0
+U

0
dU

1
dx
=
We
2
15
U
7
0
, (38)
where U

0
is the known function of x: U

0
=

1 +2x. Equation (38) is a linear non-homogeneous


rst order ordinary differential equation, which can be easily integrated by standard methods to give
U

1
=
3
2
We
U
6
0

3
2
We
U

0
, (39)
which represents the rst order correction to the streamwise velocity U. The analogous correction
for the sheet thickness H, via Eq. (36), is
H

1
=
3
2
We
U
8
0
+
3
2
We
U
3
0
. (40)
These equations are consistent with Eqs. (13c) and (13e) of Weinstein et al.
10
(which are also the
asymptotic solutions at rst order in
2
of Eqs. (6)(10)), which in our notation read
H
1
U

0
+
2
_
5
24U
7
0

3We
2U
8
0

5 36We
24U
3
0
_
+ O(
4
), (41)
U U

0
+
2
_
y
2
2U
3
0
+
3We
2U
6
0

1
4U
5
0
+
5 36We
24U

0
_
+ O(
4
). (42)
The asymptotic solution at rst order in
2
of Eqs. (6)(10) gives also an expression for the pressure
P
0
which is given by
P
0
P
a
+
2
_

y
2
U
2
0
+
1
4U
4
0

3We
2U
5
0
_
+ O(
4
). (43)
By comparing Eq. (42) with the analogous expression for the velocity U, Eq. (39), one notes
immediately that the asymptotic solution to the full system includes additional terms that do not
scale with the Weber number. The term y
2
/2U
3
0
of Eq. (42) here vanishes because the dependence
of the axial velocity on the spanwise coordinate y within the sheet is not considered in our model.
The terms 1/4U
5
0
and 5/24U

0
, linked to the curvature of the streamlines inside the sheet, do not
appear in the present development that, although including surface tension effects, considers straight
Downloaded 22 Jun 2013 to 147.127.100.21. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
062103-10 Coppola, De Rosa, and de Luca Phys. Fluids 25, 062103 (2013)
streamlines. Hence, the simple model (29) and (30) gives asymptotic solutions at small
2
, which
agree with the asymptotic predictions given by the full model, apart from the obvious absence of
terms linked to the transverse variation of ow quantities inside the sheet.
B. Low-Weber number approximation
By assuming the analogous power series expansion of the ow variables in the small parameter
We,
U = U
w
0
+WeU
w
1
+We
2
U
w
2
+. . . ,
H = H
w
0
+WeH
w
1
+We
2
H
w
2
+. . . ,
one can easily generate a hierarchy of equations at different orders in We. The lowest order solution
is again the free-fall solution, while at rst order one obtains
U
w
1
H
w
0
+U
w
0
H
w
1
= 0,
U
w
1
dU
w
0
dx
+U
w
0
dU
w
1
dx
=

2
2
d
dx
_
_
d
2
H
w
0
dx
2
_
1 +

2
4
_
dH
w
0
dx
_
2
_
3/2
_
_
.
The rst order problem can be formulated in terms of U
w
1
and U
w
0
alone,
dU
w
1
dx
+
U
w
1
U
w2
0
=
1
U
w
0
dG
dx
,
where
G =

2
2
_
_
1 +

2
4
1
U
w6
0
_
3/2
3
U
w5
0
_
.
This equation can be again integrated by standard methods in order to give
U
w
1
=
3
2
2U
0
_
1
U
w5
0
_
1 +

2
4
1
U
w6
0
_
3/2

_
1 +

2
4
_
3/2
_
,
from which we obtain the following expression for H
w
1
:
H
w
1
=
3
2
2U
3
0
_
_
1 +

2
4
_
3/2

1
U
w5
0
_
1 +

2
4
1
U
w6
0
_
3/2
_
.
IV. NUMERICAL TREATMENT AND RESULTS
Once the streamwise velocity Uhas been expressed in terms of the thickness Hthrough Eq. (29),
Eq. (30) constitutes a third order (nonlinear) ordinary differential equation for the variable H(x).
The equation can be easily rewritten as the following system of three rst order ordinary differential
equations:
dH
dx
= t, (44)
dt
dx
= s, (45)
ds
dx
=
3
2
t s
2
4 +
2
t
2

2
We
2
_
1 +
t
H
3
__
1 +
1
4

2
t
2
_
3/2
, (46)
Downloaded 22 Jun 2013 to 147.127.100.21. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
062103-11 Coppola, De Rosa, and de Luca Phys. Fluids 25, 062103 (2013)
which requires three boundary conditions to be solved. At the inlet, the thickness of the sheet is xed
by the condition H(0) =1. The values for H

and H

(primes denote differentiation with respect to x)


are unknown. In fact, the physical interpretation of Eq. (30) allows one to dene the non-dimensional
pressure inside the sheet as
P =
We
2
2
_
F
d
2
H
dx
2
_
, (47)
where F is given by Eq. (28). Assigning H

and H

at x = 0 would imply a complete specication


of the pressure at the inlet of the ow. This value, however, is unknown because it results from the
equilibrium between inertia and surface tension, which is the physical principle leading to Eq. (30)
itself.
Additional boundary conditions should be assigned as asymptotic values at the other boundary
of the domain, namely, at innity, where H and its derivatives vanish. From a numerical point of
view, the asymptotic boundary conditions have to be applied at a nite station L, which should be
sufciently far from the inlet in order to give results independent on the extension of the domain. It is
worth pointing out explicitly that hereafter the dimensionless values of L will be expressed in length
units U
2
in
/g. A straightforward imposition of the asymptotic values H

= H

= 0 at a nite station L
results in unphysical solutions for intermediate values of L and slow convergence as L is increased.
In order to improve the convergence of the solution as L tends to innity, a different choice for the
far-eld boundary conditions can be made by considering the expected asymptotic behavior of H

and H

as x tends to innity. It can be obtained by noting that as x increases the curvature of the sheet
rapidly falls to values for which surface tension forces are negligible and the sheet assumes a shape
which resembles increasingly the free-fall conguration. This represents of course an asymptotic
behaviour of the solution which cannot give the precise values of H

and H

at a xed station L;
however, we found that enforcing the values of H

and H

obtained by the non-dimensional form


of the free-fall solution, (31) and (32), dramatically improves the convergence of the solution as L is
increased. As a conclusion, the boundary conditions we assign to the system (44)(46) are
H(0) = 1, H

(L) = H

T
(L), H

(L) = H

T
(L), (48)
where the subscript T stands for the free-fall (Torricellis) solution H
T
(x) = (1 +2x)
1/2
and L is
chosen in such a way that convergence on global parameters of the sheet is achieved.
The system of Eqs. (44)(46), together with boundary conditions (48), constitutes a nonlinear
boundary value problem which can be integrated by means of several techniques. In this paper, we
have employed a shooting procedure which consists in imposing a tentative value for H(L) and the
two free-fall values for H

and H

in x =L. The shooting procedure is then initiated by adjusting the


value H(L) in such a way that the correct value H(0) = 1 is achieved when the system is integrated
backward in space. By employing a bisection procedure, a typical value of n = 30 iterations is
required in order to obtain an absolute error of order of 10
10
on H(0) when assuming the free-fall
value H
T
(L) as initial guess for H(L). Once the solution has been obtained for a xed value of L, the
calculation is repeated with greater values for L until convergence of the solution is reached.
In Fig. 2, a representative set of curves (on which the bisection procedure has been applied)
depicting the value H(0) as a function of the guess value H(L) is plotted. These curves have been
computed by integrating Eqs. (44)(46) (equipped with asymptotic conditions H

(L) = H

T
(L) and
H

(L) = H

T
(L)) starting from different initial guesses for H(L) and recording the corresponding
values H(0), for the case L = 20,
2
= 1, and for various Weber numbers. The bisection procedure
starts fromthe point at the intersection between a certain curve and the vertical axis at H(L)/H
T
(L) =
1 and proceeds until the correct value of H at x = L, given by the intersection of the curve with the
horizontal line at H(0) = 1, is reached.
The convergence history obtained for increasing values of L (where boundary conditions are
enforced) is reported in Fig. 3 for a typical case corresponding to
2
= 0.25 and We = 0.5. The
global parameters on which the convergence check is made are the values of H

and P at x = 0.
The pressure P is calculated from the computed values of H by employing Eq. (47). The two curves
plotted in each graph are relative to the two different choices for the far-eld boundary conditions:
dashed line refers to the homogeneous conditions H

(L) = H

(L) = 0, while continuous line is


Downloaded 22 Jun 2013 to 147.127.100.21. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
062103-12 Coppola, De Rosa, and de Luca Phys. Fluids 25, 062103 (2013)
0.98 0.99 1 1.01 1.02
0.5
1
1.5
2
H(L)/H
T
(L)
H
(
0
)
We =0.01
0.2
0.4
0.6
0.8
1.0
We =0.01
0.2
0.4
0.6
0.8
1.0
We =0.01
0.2
0.4
0.6
0.8
1.0
We =0.01
0.2
0.4
0.6
0.8
1.0
We =0.01
0.2
0.4
0.6
0.8
1.0
We =0.01
0.2
0.4
0.6
0.8
1.0
FIG. 2. Values of H(0) as a function of the guess value H(L) for various We numbers.
2
= 1, L = 20.
obtained by imposing the free-fall values for H

and H

(Eq. (48)). It should be noted that both the


initial slope of the sheet and the pressure at x = 0 are strongly affected by the location L at which
the asymptotic boundary conditions are imposed, at least for moderate values of L. Moreover, it is
evident that the imposition of the free-fall boundary conditions strongly improves the convergence
rate as L is increased.
The comparison among the convergence histories achieved for different values of We with
the two free-fall boundary conditions is reported in Fig. 4 for
2
= 0.25. In order to facilitate the
1 2 3 4 5 6 7
0.85
0.8
0.75
L
H
(
0
)
(a)
1 2 3 4 5 6 7
0.12
0.1
0.08
0.06
L
P
(
0
)
(b)
FIG. 3. Convergence history for
2
= 0.25, We = 0.5. Continuous lines (red) refer to boundary conditions of Eq. (48),
dashed lines refer to H

(L) = H

(L) = 0.
Downloaded 22 Jun 2013 to 147.127.100.21. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
062103-13 Coppola, De Rosa, and de Luca Phys. Fluids 25, 062103 (2013)
1 2 3 4 5 6 7
0.98
0.99
1
1.01
1.02
L
H
(
0
)
( ) ( ) (a)
1 2 3 4 5 6 7
0.98
0.99
1
1.01
1.02
L
P
(
0
)
( ) ( ) (b)
FIG. 4. Normalized derivative of thickness H (a), and normalized pressure P (b), at x = 0 for different locations L at which
the free-fall boundary condition is applied for
2
= 0.25. We = 0.1 (solid line), We = 0.5 (dashed-dotted line), We = 1
(dashed line).
comparison between the curves obtained with different values of We, both of the two quantities
H

(0) and P(0) have been normalized with respect to the asymptotic values they assume at large
values of L. These normalized quantities have been denoted by H

(0) and P(0). Figure 4 shows that


the sensitivity of the solution to the location at which the articial boundary condition is assigned
is quite strong for small values of L, and the convergence toward the asymptotic values is always
nonmonotonic. The oscillation amplitudes are stronger for higher values of the Weber number and
a similar behavior has been detected by varying the slenderness parameter (the resulting plots are
not shown here). This is expected, since the free-fall solution is recovered in both limits of We and
tending to zero, and hence the articial boundary conditions are increasingly more accurate with
decreasing one of these two parameters. The relative amplitude of the oscillations on the global
parameters here displayed typically reduces, in the case of free-fall asymptotic boundary conditions,
to values of the order of 10
4
for L = 20. Thus, the free-fall asymptotic boundary conditions have
been enforced at L = 20 in all the numerical calculations reported hereafter.
The shape of the sheet interface and the pressure distribution inside the liquid are represented
for different values of We at xed
2
in Fig. 5 and for different values of
2
at xed We in Fig. 6. The
free-fall solution is plotted for comparison as dashed lines. Both the external boundary of the sheet
and the pressure distribution inside it converge to the free-fall solutions as one of the parameters,
2
or We, is reduced by considering the other one xed. The effect of surface tension on the shape of
the sheet always results in an increase of the local thickness, i.e., of reducing the local streamwise
velocity as compared to the free-fall solution. This effect can be seen also as a stretching of the
sheet interface corresponding to a reduction of the surface of the sheet between the inlet and the
generic station x. Figure 6 shows that the inuence of
2
on the sheet shape is apparently the same as
that of We, but this is clearly unphysical, since gravity tends to thin the sheet. In fact, the expected
thinning is masked by the dimensionless x scale that varies linearly with . In the inset of Fig. 6(a),
the sheet shape is reported as a function of the scaled streamwise distance x/ = x

/H
in
and the
intuitive behavior is clearly illustrated. The pressure inside the sheet is always reduced with respect
to the free-fall isobaric solution, this being a direct consequence of the equilibrium of forces at the
interface between the liquid and the gas when surface tension is taken into account (Eq. (5)).
Downloaded 22 Jun 2013 to 147.127.100.21. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
062103-14 Coppola, De Rosa, and de Luca Phys. Fluids 25, 062103 (2013)
0.3 0.4 0.5
2
1.8
1.6
1.4
1.2
1
0.8
0.6
0.4
0.2
0
Y
+
x
(a) ( ) ( ) ( ) ( )
We
0.2 0.1 0
2
1.8
1.6
1.4
1.2
1
0.8
0.6
0.4
0.2
0
P
(b) ( ) ( ) ( ) ( )
We
FIG. 5. Sheet prole (a) and axis pressure distribution (b) for
2
= 0.25 and We = 0.1, 0.5, 1, 1.5, 2. Dashed line (red) is
the free-fall solution.
Figures 7 and 8 present a parametric study of the variation of the slope of the sheet boundary
and of the pressure at the inlet section as
2
and We are varied. The variation of We at selected values
of
2
, reported in Fig. 7, shows that the effect of the surface tension on the values of P and Y

+
at the
inlet of the sheet is more pronounced for higher values of
2
. Moreover, the slope of these curves at
0.3 0.4 0.5
2
1.8
1.6
1.4
1.2
1
0.8
0.6
0.4
0.2
0
Y
+
x
(a) ( ) ( ) ( ) ( )

2
0.2 0.1 0
2
1.8
1.6
1.4
1.2
1
0.8
0.6
0.4
0.2
0
P
(b) ( ) ( ) ( ) ( )

2
0.3 0.4 0.5
2
1.5
1
0.5
0
Y
+
x
/

2
FIG. 6. Sheet prole (a) and axis pressure distribution (b) for We = 1 and
2
= 0.05, 0.1, 0.2, 0.3, 0.4. Dashed line (red) is
the free-fall solution. In the inset, the sheet shape is reported as a function of the scaled streamwise distance x/.
Downloaded 22 Jun 2013 to 147.127.100.21. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
062103-15 Coppola, De Rosa, and de Luca Phys. Fluids 25, 062103 (2013)
0 0.2 0.4 0.6 0.8 1
0.2
0.1
0
P
(
0
)
(a)

2
( )

2
( )

2
( )

2
( )

2
0 0.2 0.4 0.6 0.8 1
0.55
0.5
0.45
0.4
0.35
0.3
Y
+
(
0
)
We
(b)
2
( )
2
( )
2
( )
2
( )
2
FIG. 7. Variation with We of pressure P (a) and dY
+
/dx (b) at inlet location x = 0 for
2
= 0.05, 0.1, 0.2, 0.3, 0.4. Dashed
line (red) is the free-fall solution.
We = 0 is not zero, thus indicating that signicant changes in the global features of the sheet with
respect to the free-fall solution are expected when a small amount of surface tension is considered
in the model. The effects of a variation of
2
at selected values of We is reported in Fig. 8. Here,
again, it is shown that at higher values of We the changes in the global properties of the sheet with
respect to the free-fall solution are more pronounced as
2
is increased.
0 0.1 0.2 0.3 0.4
0.2
0.1
0
0.1
P
(
0
)
(a)
We
( )
We
( )
We
( )
We
( )
We
0 0.1 0.2 0.3 0.4
0.55
0.5
0.45
0.4
0.35
0.3

2
Y
+
(
0
)
We
(b)
We
( )
We
( )
We
( )
We
( )
FIG. 8. Variation with
2
of pressure P (a) and dY
+
/dx (b) at x = 0 for We = 0.05, 0.2, 0.4, 0.6, 0.8, 1. Dashed line (red) is
the free-fall solution.
Downloaded 22 Jun 2013 to 147.127.100.21. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
062103-16 Coppola, De Rosa, and de Luca Phys. Fluids 25, 062103 (2013)
0 0.05 0.1 0.15 0.2
0.02
0.01
0
P
(
0
)
(a)
0 0.05 0.1 0.15 0.2
0.04
0.02
0
P
(
0
)
(b)
We
FIG. 9. Variation with We of inlet liquid pressure P for
2
= 0.0625 (a) and
2
= 0.25 (b). Continuous thick line is the
numerical solution, continuous thin line (red) corresponds to the zeroth order asymptotic solution for H, dashed-dotted line
corresponds to the rst order asymptotic solution for H.
Figures 9 and 10 report a comparison between the numerically determined inlet pressure in-
side the sheet with respect to We and
2
based on the numerical integration of Eq. (30) and the
corresponding prediction given by the asymptotic formulas derived in Sec. III. In all cases pres-
sure P is determined trough Eq. (47) in which the derivatives of H are expressed by means of
the numerical solution of the system (44)(46) as in Figs. 7 and 8 (continuous thick lines), or by
differentiating the approximate formulas for H obtained in the asymptotic expansion with respect
to
2
or We. Concerning the analytical solutions, we computed for comparison both the solution
corresponding to the zeroth order approximations H = H

0
or H = H
w
0
(continuous thin lines in
Figs. 9 and 10), and the solution corresponding to the rst order approximation, H = H

0
+
2
H

1
or
H = H
w
0
+WeH
w
1
(dashed-dotted lines). Figure 9 shows that the rst order approximation in terms
of We number expansion, although more accurate than that of the zeroth order at low values of We,
quickly deviates form the numerical solution as We is increased. This effect is more pronounced
for higher values of
2
. In the case presented in Fig. 9(b), for instance, (
2
= 0.25) pressure P
obtained by employing the rst order analytical approximation for H in terms of powers of We is a
reasonable approximation to the numerical solution for We 0.05. For We 0.1, on the other hand,
the rst order approximation for H gives pressure values which deviate from the numerical solution,
nally leading to completely unphysical results for We 0.166, at which the pressure becomes
positive.
A similar trend is displayed in Fig. 10, in which the variation of P(0) with
2
numerically
computed is compared to the zeroth and rst-order analytical approximation in terms of powers of

2
. In this plot a further curve is reported, which is the pressure variation with
2
as predicted by the
parameter-type expansion of the full equations of motions (6)(10) by considering terms up to rst
order in
2
. This pressure is calculated with the formula (43) of Sec. III A. The curve is denoted by
a dashed line and is barely distinguishable from the thin continuous line, relative to the zeroth order
asymptotic treatment of our Eqs. (29) and (30).
For the case of Fig. 10(b) (We = 0.5), the solution corresponding to rst order approxima-
tion for H is quite accurate for 0.1 but strongly deviates from the numerical solution for
0.2, and becomes unphysical for 0.248, at the pressure at which the inlet becomes
positive.
Downloaded 22 Jun 2013 to 147.127.100.21. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
062103-17 Coppola, De Rosa, and de Luca Phys. Fluids 25, 062103 (2013)
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08
0.02
0.01
0
P
(
0
)
(a)
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08
0.03
0.02
0.01
0
P
(
0
)
(b)

2
FIG. 10. Variation with
2
of inlet liquid pressure Pfor We = 0.1 (a) and We = 0.5 (b). Continuous thick line is the numerical
solution, continuous thin line (red) corresponds to the zeroth order asymptotic solution for H, dashed-dotted line corresponds
to the rst order asymptotic solution for H. The dashed line (barely distinguishable from the thin continuous line) is the
pressure calculated with the formula (43) of Sec. III A.
V. CONCLUSIONS
The effects of surface tension on the stationary motion of a liquid sheet falling in the gravitational
eld have been discussed under the assumptions of inviscid owand negligible friction of the external
ambient gas. A quite general model governing the streamwise distribution of thickness as well as
pressure and velocity elds inside the sheet has been derived by means of a power series expansion
in terms of lateral displacement fromthe mean line of the sheet. The lower order truncated expansion
leads to plug-type velocity and pressure distributions and, consequently, the relevant quantities are
functions of the streamwise coordinate only. The resulting one-dimensional model exhibits the
slenderness ratio and the Weber number We as parameters. The validity of the present model has
been assessed at rst by comparing the resulting simplied equations in the unsymmetrical case,
in which a pressure difference between the two sides of the sheet is applied, to previous existing
models.
The case of symmetric conguration has been treated separately in more detail and a novel
non-linear set of governing equations has been derived for this case. These equations constitute a
simple model in which the effects of surface tension on the stationary shape of the sheet are taken
into account by considering the complete expression for the curvature. Approximate solutions have
been obtained by means of perturbation approaches, which are asymptotically valid for small We and
parameters. The numerical procedure employed to integrate the equation, based on a shooting-type
method, has been accurately discussed in connection with its convergence properties. The results
are consistent with the expected behavior that surface tension tends to stretch the sheet interface,
and thus to reduce the local streamwise velocity as compared to the Torricellis free-fall solution,
and that gravity thins the sheet.
1
S. J. Weinstein and K. J. Ruschak, Coating ows, Annu. Rev. Fluid Mech. 36, 29 (2004).
2
W. G. Pritchard, Instability and chaotic behaviour in a free-surface ow, J. Fluid Mech. 165, 1 (1986).
3
L. de Luca and C. Meola, Surfactant effects on the dynamics of a thin liquid sheet, J. Fluid Mech. 300, 71 (1995).
4
D. R. Brown, A study of the behaviour of a thin sheet of moving liquid, J. Fluid Mech. 10, 297 (1961).
5
N. S. Clarke, Two-dimensional ow under gravity in a jet of viscous liquid, J. Fluid Mech. 31, 481 (1968).
6
C. K. Aidun, Mechanics of a free-surface liquid lm ow, Trans. ASME, J. Appl. Mech. 54, 951 (1987).
Downloaded 22 Jun 2013 to 147.127.100.21. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
062103-18 Coppola, De Rosa, and de Luca Phys. Fluids 25, 062103 (2013)
7
J. I. Ramos, Planar liquid sheets at low Reynolds numbers, Int. J. Numer. Methods Fluids 22, 961 (1996).
8
D. S. Finnicum, S. J. Weinstein, and K. J. Ruschak, The effect of applied pressure on the shape of a two-dimensional
liquid curtain falling under the inuence of gravity, J. Fluid Mech. 255, 647 (1993).
9
L. de Luca and M. Costa, Instability of a spatially developing liquid sheet, J. Fluid Mech. 331, 127 (1997).
10
S. J. Weinstein, A. Clarke, A. G. Moon, and E. A. Simister, Time-dependent equations governing the shape of a two-
dimensional liquid curtain, Part 1: Theory, Phys. Fluids 9, 3625 (1997).
11
L. de Luca and M. Costa, Two-dimensional ow of a liquid sheet under gravity, Comput. Fluids 24, 401 (1995).
12
G. Rocco, G. Coppola, and L. de Luca, The VOF method applied to the numerical simulation of a 2D liquid jet under
gravity, WIT Trans. Eng. Sci. 69, 207 (2010).
13
P. J. Schmid and D. S. Henningson, On the stability of a falling liquid curtain, J. Fluid Mech. 463, 163 (2002).
14
G. Coppola and L. de Luca, On transient growth oscillations in linear models, Phys. Fluids 18, 078104 (2006).
15
J. I. Ramos, Asymptotic analysis and stability of inviscid liquid sheets, J. Math. Anal. Appl. 250, 512 (2000).
16
M. Van Dyke, Perturbation Methods in Fluid Mechanics (The Parabolic Press, Stanford, CA, 1975).
17
I. D. Chang, Navier-Stokes solutions at large distance from a nite body, J. Math. Mech. 10, 811 (1961).
18
J. Eggers and T. F. Dupont, Drop formation in a one-dimensional approximation of the NavierStokes equation, J. Fluid
Mech. 262, 205 (1994).
19
C. Mehring and W. A. Sirignano, Nonlinear capillary wave distortion and disintegration of thin planar liquid sheets, J.
Fluid Mech. 388, 69 (1999).
Downloaded 22 Jun 2013 to 147.127.100.21. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions

Вам также может понравиться