Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/236174044

Asymptotic behavior of a retracting two-dimensional fluid sheet

Article  in  Physics of Fluids · December 2011


DOI: 10.1063/1.3663577

CITATIONS READS

21 248

4 authors, including:

Leonardo Gordillo G. Gilou Agbaglah


University of Santiago, Chile University of Michigan
25 PUBLICATIONS   140 CITATIONS    7 PUBLICATIONS   290 CITATIONS   

SEE PROFILE SEE PROFILE

Christophe Josserand
French National Centre for Scientific Research
172 PUBLICATIONS   4,012 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Freezing drop impact View project

Capillary retraction of liquid filaments View project

All content following this page was uploaded by Leonardo Gordillo on 18 November 2014.

The user has requested enhancement of the downloaded file.


PHYSICS OF FLUIDS 23, 122101 (2011)

Asymptotic behavior of a retracting two-dimensional fluid sheet


Leonardo Gordillo,1 Gilou Agbaglah,2 Laurent Duchemin,3 and Christophe Josserand2
1
Departamento de Fı́sica, Facultad de Ciencias Fı́sicas y Matemáticas, Universidad de Chile, Casilla 487-3,
Santiago, Chile
2
Institut Jean Le Rond D’Alembert, CNRS & UPMC (Univ. Paris VI) UMR 7190 Case 162, 4 place Jussieu,
F-75252 Paris Cédex 05, France
3
IRPHE, Université d’Aix-Marseille I & II-CNRS 49 rue Joliot-Curie, BP 146, 13384 Marseille CEDEX,
France

(Received 3 August 2011; accepted 3 November 2011; published online 7 December 2011)
Two-dimensional (2D) capillary retraction of a viscous liquid film is studied using numerical and
analytical approaches for both diphasic and free surface flows. Full 2D Navier-Stokes equations are
integrated numerically for the diphasic case, while one-dimensional (1D) free surface model equa-
tions are used for free surface flows. No pinch-off is observed in the film in any of these cases. By
means of an asymptotic matching method on the 1D model, we derive an analytical expansion of
the film profile for large times. Our analysis shows that three regions with different timescales can
be identified during retraction: the rim, the film, and an intermediate domain connecting these two
regions. The numerical simulations performed on both models show good agreement with the
analytical results. Finally, we report the appearance of an instability in the diphasic retracting film
for small Ohnesorge number. We understand this as a Kelvin-Helmholtz instability arising due to
C 2011 American Institute of
the formation of a shear layer in the neck region during the retraction. V
Physics. [doi:10.1063/1.3663577]

I. INTRODUCTION behavior.14–18 For instance, the liquid viscosity affects the


timescale of the transitory regime leading to the Taylor-
A thin liquid sheet retracts due to the action of the sur-
Cullick velocity17 and it also influences drastically the trans-
face tension forming a growing rim at its retracting end as it
verse instability of the film.12,19,20 The film profile depend-
recedes. In 1959, Taylor and Cullick1,2 found simultaneously
ence on the liquid viscosity in the quasi-stationary regime at
that the retracting speed of a liquid film of uniform thickness
constant velocity is even more remarkable, as it has already
2e converges to a constant velocity UTC given by the balance
been seen in experiments and numerical simulations.14–17 For
between inertia and surface tension
instance, there is experimental evidence that a very viscous
rffiffiffiffiffi film does not form any rim during retraction and that the film
c
UTC ¼ ; apparently thickens as it retracts.14,15 Actually, numerical
qe
simulations of the thin film equations have shown that in this
where c is the surface tension and q is the liquid density. case, a very large retracting edge forms, which is thicker than
Numerical simulations (cf. Sec. II) show that, asymptoti- the film and the experimental film length.16 On the other
cally, the geometry of the rim tends to a circular shape, hand, when the viscous effects are small enough, a thin neck
whose radius evolves as the square-root of time1–3 as illus- connects the rim to the film of constant thickness. In that sit-
trated in Figure 1. uation, one can ask whether this neck can lead to the two-
Retraction dynamics is a crucial topic in many interface dimensional (2D) break-up of the film in the inviscid limit
dynamics problems such as curtain coating,4 atomization,5,6 for a large rim radius as postulated by Song and
and drop impact dynamics.7 In fact, many liquid sheet insta- Tryggvason.21
bilities commonly observed in nature are triggered by edge Therefore, the classical picture of a circular rim retract-
retraction. Such a mechanism is reported in liquid curtains, ing into a film of constant thickness has to be revised. In this
where transverse instabilities generated by retracting edges article, we investigate numerically and analytically the 2D
can exhibit spatio-temporal chaos.8,9 In liquid jets, disinte- film profile evolution in the large time limit, i.e., when the
gration of thin films leads to the formation of small droplets rim radius is much larger than the film thickness, so that the
and full atomization.10 In drop impact, the splashing and the film can be expanded into a quasi-stationary regime approxi-
corona instability are linked to the dynamics of a free liquid mation. In Sec. II, we present the problem describing and
edge.11–13 comparing the two numerical methods used to model the
Hence, a further analysis which goes beyond the classi- film dynamics, i.e., integration of the Navier-Stokes equation
cal Taylor-Cullick theory is of utmost importance for under- for the full system (liquid film and surrounding gas), and
standing film motion. The quasi-stationary Taylor-Cullick simulation of the one-dimensional (1D) system of equations
model is based on non-local mass and momentum conserva- obtained for thin films using the free surface boundary
tion without considering viscosity at all. In contrast, transient conditions and lubrication approximation. In Sec. III, we
dynamics and viscous effects can highly change the retraction present the film dynamics within the 1D model in the

1070-6631/2011/23(12)/122101/14/$30.00 23, 122101-1 C 2011 American Institute of Physics


V

Downloaded 07 Dec 2011 to 200.89.69.173. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions
122101-2 Gordillo et al. Phys. Fluids 23, 122101 (2011)

FIG. 1. (Color online) Evolution of the sheet interface and vorticity field for Z ¼ 0.14 at four different dimensionless times: (a) t* ¼ 0, (b) t* ¼ 10, (c) t* ¼ 20,
and (d) t* ¼ 30. The incompressible Navier-Stokes equations are solved for both fluids using a VOF method in GERRIS.22

quasi-stationary approximation. A mathematical analysis of Consequently, in this formulation, q and l are discontinuous
the 1D system of equations for large times shows that the scalar fields that account for the densities and p
viscosities
ffiffiffiffiffiffiffiffiffiffiffiffiffi of
time asymptotic solution can be separated into three different each fluid. If we define the capillary time s ¼ qL e3 =c, the
regions—film, neck and rim—with different length and time dynamics can then be rewritten in terms of dimensionless
scales. These regions can then be matched together in the variables
spatial domain in order to build a whole domain solution.
The details of the calculations can be found in Appendices A q~ð@t u þ u $ $uÞ ¼ &$p þ Z~
lDu þ jds n; (3)
and B. We would like to emphasize that obtaining such a $ $ u ¼ 0: (4)
quasi-analytical solution of the 2D film retraction opens the
way to improving the 3D linear stability analysis of liquid The length, time, velocity, pressure,
pffiffidensity,
ffiffiffiffiffiffiffiffiffi and viscosity
films, by taking into account the film profile shape in a more have been rescaled by e, s, UTC ¼ c=qL e, c/e, qL, and lL,
accurate way. Finally, the comparison between asymptotic respectively. Besides the density and viscosity ratios, the dy-
solutions and numerics is discussed in Sec. IV. namics depends only on one dimensionless number, the
Ohnesorge number Z, which compares viscous and capillary
II. DYNAMICS OF THE RETRACTING FILM effects

A. Fluids equations l
Z ¼ pffiffiffiLffiffiffiffiffiffi :
qL ce
We consider the two dimensional dynamics of a thin
film of initial uniform thickness 2e of a liquid in the presence The dimensionless density q~ and viscosity l~ are both 1 in the
of a surrounding gas. The densities and dynamical viscosities liquid phase and qG/qL and lG/lL, respectively, in the gas
of the liquid and the gas are noted as (qL, qG) and (lL, lG), phase. As a consequence, the case of free surface flow can be
respectively. Both fluids can be considered as incompressi- easily accounted for by considering qG ¼ 0 and lG ¼ 0.
ble, thus, the dynamics is governed by the 2-D incompressi-
ble Navier-Stokes equations
B. Full numerical simulations
qð@t u þ u $ $uÞ ¼ &$p þ lDu þ cjds n; (1) Numerical integration of the set of Eqs. (3) and (4) is per-
r $ u ¼ 0; (2) formed using the GERRIS code,22 which uses a staggered-in-
time discretisation of the volume-fraction/density and pres-
where the velocity and pressure fields are noted u ¼ uex þ vey sure. The interface was tracked using the volume of fluid
and p, c being the liquid-gas surface tension. Capillary forces method (VOF) and an adaptive mesh refinement based on
are modeled by introducing the Dirac delta function ds at the quadtree decomposition (octree in 3D) is used. The combina-
interface, with n and j denoting the local normal direction tion of both techniques allows efficient computation of com-
and the local curvature of the interface, respectively. plex interfacial flows.23–25

Downloaded 07 Dec 2011 to 200.89.69.173. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions
122101-3 Asymptotic behavior of a retracting 2-D fluid sheet Phys. Fluids 23, 122101 (2011)

FIG. 2. (Color online) Spatial domain and initial condition for the retracting
sheet numerical simulation. The domain consists of four square subdomains.
The initial length of the sheet is 18e0 and the square subdomain length is
20e0.

As an initial condition, we considered a long free end


liquid film of thickness e0 ¼ 2e as illustrated in Figure 2. The
simulations were performed in a comoving frame of refer- FIG. 4. (Color online) Time evolution of the rim radius extracted from the
ence that recedes at the Taylor-Cullick velocity. This was direct numerical simulation of the capillary film retraction. The linear law
expected by the Taylor-Cullick regime is in very good agreement with the
achieved by imposing Dirichlet boundary conditions for the numerics.
velocity field on the left boundary and an outlet condition on
the right side. On the other hand, the mesh refinement is con-
trolled both by the interface position and the vorticity field, in dimensionless units. This is in good agreement with the
in order to maintain a good resolution around the interface as numerical results shown in Figure 4. However, we must note
well as in the region where the vorticity becomes strong. We that this result relies on the rim being an approximately cir-
will consider later an air/water like system by setting the cular profile, a feature that depends strongly on the Ohne-
density and viscosity ratios to qL/qG ¼ 850 and lL/lG ¼ 50, sorge number (see Figure 5). In particular, we observe a
respectively. The evolution of the film profile and the vortic- qualitative change in the film profile as the Ohnesorge varies.
ity field is shown in Figure 1 for Z ¼ 0.14. The surface For viscous liquid, i.e., high Ohnesorge number, the film
tension pulls back the rim and, after a few units of dimen- thickness decays monotonically in the upstream &x direc-
sionless time, the rim reaches the Taylor-Cullick velocity tion. By contrast, for capillary driven flows, i.e., low Ohne-
UTC, which corresponds to rest in the comoving frame. Then, sorge number, one can observe spatial oscillations, whose
a quasi-stationary regime sets up where the most of the dy- amplitude decreases with the distance to the rim. The profile
namics consists of a slow increase of the rim size. We will behind the rim seems to be quite stable after convergence to
refer to this asymptotic stage as the Taylor-Cullick regime. the Taylor-Cullick velocity, at least for the first two cases
As a first approach, following the arguments of Taylor (Z ¼ 0.7 and Z ¼ 0.14). Figure 5 also shows that for low
and Cullick,1,2,26 the rim can be considered circular (see Ohnesorge numbers, vorticity is mainly concentrated in
Figure 3) and a simple mass balance equation for the rim strong curvature points, especially in the first minimum
gives the evolution of the rim radius R(t), region behind the rim. Henceforth, we will refer to these
regions as the neck of the film, following Song and
d" #
2eUTC ¼ pRðtÞ2 : Tryggvason.21
dt
This equation can be easily integrated with our initial C. Thin film equation
conditions and yields
$ % When the effects of the surrounding gas can be
2 neglected, the Navier-Stokes equations with free surface
RðtÞ2 ¼ t þ 1;
p boundary conditions can be reduced to a set of coupled equa-
tions within the thin film approximations—or long wave-
length limit, where horizontal variations can be considered
much smaller than the vertical ones—as shown by Erneux
and Davis.27 The equations are written in terms of the local
thickness 2h(x,t) and the parallel velocity of the flow u(x,t),
in dimensionless form

@t h þ @x ðhuÞ ¼ 0; (5)
) ð *
& 2'
@t ðhuÞ þ @x hu ¼ @x 4Zh@x u þ h@x j dx : (6)

This set of equations, obtained within the lubrication approx-


imation, is valid both in a fixed reference frame or in a con-
stant velocity moving frame. Equation (5) is derived from
FIG. 3. (Color online) Fit between numerical rim for Z ¼ 0.14 at t ¼ 20 and local mass conservation in the film while Eq. (6), written
circular shape, (x & 0.199)2 þ y2 ¼ 0.025 here in a conservative form, corresponds to the local

Downloaded 07 Dec 2011 to 200.89.69.173. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions
122101-4 Gordillo et al. Phys. Fluids 23, 122101 (2011)

FIG. 5. (Color online) Sheet profile and vorticity field for three different Ohnesorge numbers at t ¼ 20: (a) Z ¼ 0.7, (b) Z ¼ 0.14, and (c) Z ¼ 0.028.

momentum balance. However, in order to cope with the tip kinematic condition at the tip and the symmetries of the
of the rim, the surface-tension term has to be amended by problem.
using the complete curvature j instead of its long wave- Numerical simulation of these 1D Eqs. (5) and (6) can
length limit j ' @ xxh. The integral associated with the sur- be performed using a finite difference method. Equations (5)
face tension term can then be expressed in a closed form and (6) are written in terms of u and A ¼ h2 and solved on a
ð staggered grid. A numerical issue arising in these kinds of
h@xx h þ ð@x hÞ2 þ1 equations is that the computational domain shrinks in time.
h@x j dx ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi3 :
In order to avoid having a time-dependent grid, we mapped
1 þ ð@x hÞ2 the domain into [0,1], rescaling x by the total length of the
sheet l(t). As a consequence, the boundary of the film is mov-
This model has been able to reproduce successfully 2D film _
ing in this frame and one needs to write an equation for lðtÞ.
retraction when the film is semi-infinite.16 The set of Eqs. (5) This equation simply says that the end point velocity, which
and (6) is well-posed by adding consistent initial conditions. _
can be extrapolated from the bulk velocity, equals lðtÞ. The
For the free retracting fluid sheet, this problem can also be new set of equations reads
seen as a kind of boundary value problem on a semi-infinite
domain which evolves with time. In the Taylor-Cullick ve- 1& '
_ @X A & 2A @X u;
locity moving frame the solutions satisfy a Dirichlet bound- @T A ¼ & u & lX (9)
l l
ary condition far from the tip
1& ' " #
hðx; tÞ ¼ 1
+
@T u ¼ & _ @X u þ 4Z @X A1=2 @X u
u & lX
at x ¼ &1; (7) l l2 A1=2
uðx; tÞ ¼ 1 1
þ @X j; (10)
l
and a singular Cauchy boundary condition at the film tip,
9 @T l ¼ uðlðtÞ; tÞ; (11)
hðx; tÞ ¼ 0 =
@x hðx; tÞ ¼ &1 at x ¼ x0 ðtÞ: (8) where X ¼ x/l(t) and T ¼ t.
; A Runge-Kutta method is used together with centered fi-
uðx; tÞ ¼ x_ 0
nite difference formulas to solve this set of equations with
The first boundary condition (7) comes from the unper- X 2 ½0; 1). The time evolution of the film retraction is shown
turbed geometry far away from the rim and the influx veloc- in Figure 6 for the same Ohnesorge numbers as in Figure 5.
ity on the reference frame at which the rim does not recede. The numerical solutions are robust when the domain size is
The second boundary condition (8) is imposed by the such that the boundary is far enough from the spatial

FIG. 6. (Color online) Sheet profile for three different Ohnesorge numbers at t ¼ 20, using the thin film approximation: (a) Z ¼ 0.7, (b) Z ¼ 0.14, and (c)
Z ¼ 0.028.

Downloaded 07 Dec 2011 to 200.89.69.173. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions
122101-5 Asymptotic behavior of a retracting 2-D fluid sheet Phys. Fluids 23, 122101 (2011)

oscillations behind the rim. Numerical convergence of the Because of the momentum balance between the inertial and
solution has been reached by changing the initial domain the surface-tension term at &1 coming from the Taylor-
size (l(0)). Cullick frame of reference, the constant of integration of
A good qualitative agreement is found between the two Eq. (15) given by boundary conditions at &1 is zero.
ð0Þ
numerical methods. In particular, we observe that the film ge- At z ! &1, hf ðzÞ tends to unity and nonlinear terms
ometry changes similarly as the Ohnesorge number varies. can be suppressed. The equation then behaves as a second
Before we perform more quantitative comparisons, we investi- order linear differential equation. By expanding the solution
gate in Sec. III the main features of the solution of this model ð0Þ
in this limit, i.e., hf ðzÞ ¼ 1 þ ! ek$z , we obtain
using analytic expansions in the asymptotic large time limit. pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
k ¼ 2Z6 4Z2 & 1. We find thus that the decay length scales
as 4Z for large Ohnesorge number and as (2Z)&1 for low
III. ASYMPTOTIC EXPANSION ones. Moreover, the solution presents spatial oscillations if
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
We are thus interested in finding an asymptotic expansion Z < Z0 ¼ 0.5 with wavenumber ki ¼ 1 & 4Z2 which con-
for the thin film model of Eqs. (5) and (6) for large times. In verge to unity for small Ohnesorge numbers, in quantitative
this regime, the rim is already well formed so that we can con- agreement with our numerical results.
sider that its radius is much larger than the film thickness Numerical integration of Eq. (16) reveals the appear-
ðRðtÞ * eÞ and that its receding velocity is only slightly dif- ance of two types of derivative singularities at finite z, cor-
ferent from the Taylor-Cullick value. In such a limit, by responding to a positive and a negative infinite slope of the
choosing a proper set of length scales for x, y, and u, time interface, respectively, as shown in Figure 7. In fact, such
derivatives can be neglected at least at the leading order, singularities appear because steady finite solutions of Eq.
when other terms containing spatial derivative may become (5) and (6) cannot be supported in the whole domain.
dominant. In order to obtain a solution valid in the whole Indeed, they would violate mass conservation since there is
domain at dominant and higher orders, we will analyze the a net flux of mass from infinity. The phase portrait of Eq.
nonlinear system given by Eqs. (5) and (6) at different self- (16) shows as well the existence of a manifold that behaves
similar scales. The existence of different length scales implies as the separatrix between the two types of singular solu-
the appearance of several regions, which should be matched tions. This separatrix plays a crucial role since it corre-
in space in order to generate a unique whole domain solution. sponds formally to an infinite mass rim, the only one that is
consistent with the mass conservation for the steady film
A. Far-field solution evolution. Therefore, in the large time asymptotic limit, the
rim should match this separatrix at the leading order. This
We expect, as observed in our numerical simulations,
curve is well defined in the whole domain and its asymptot-
that in the co-moving frame, the far-field flow remains
ical behavior can be found by balancing the viscous and the
unperturbed when the rim retracts. Thus, we pose solutions
surface-tension terms of Eq. (16). At leading order, the sep-
of the form
aratrix is quadratic and it can be expanded around z ! 1,
ð0Þ ð 1Þ
hf ðx; tÞ ¼ hf ðzÞ þ hf ðz; tÞ; (12)
ð0Þ ð 1Þ
uf ðx; tÞ ¼ uf ðzÞ þ uf ðz; tÞ; (13)

where z : x & xf accounts for the translational invariance of


ð1Þ ð1Þ
the equations. The fields hf ðz; tÞ and uf ðz; tÞ are supposed
to be small corrections of the dominant terms. Replacement
of the Ansatz (12) and (13) into Eqs. (5) and (6), yields at
zeroth order

@z ðhuÞ ¼ 0; (14)
2 3
& ' 6 h@zz h þ ð@z hÞ2 þ17
@z hu2 ¼ @z 44Zh@z u þ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 3 5: (15)
1 þ ð@ z h Þ2

We have omitted subscripts and superscript symbols for sim-


plifying notations. The first equation can be easily integrated
using the boundary conditions at infinity (7). Furthermore,
the momentum equation (15) can be integrated after substitu-
FIG. 7. (Color online) The two families of singular solutions of Eq. (16) can
tion of u in terms of h, yielding be obtained by choosing different initial conditions when performing numer-
ical integration. Both families develop derivative singularities for finite x but
1 þ 4Z@z h h@zz h þ ð@z hÞ2 þ1 differing sign diverging slopes: The upper left family for positive (red
¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 3 : (16) online) and the lower right family for negative (blue online). The separatrix
h
1 þ ð@z hÞ2 (black dashed curve) is the unique function that can be defined in the whole
domain.

Downloaded 07 Dec 2011 to 200.89.69.173. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions
122101-6 Gordillo et al. Phys. Fluids 23, 122101 (2011)

1 1 1
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ h&1 & aða & 2Þh: (18)
2 2 2
1 þ ð@z hÞ

Depending on the choice of a, the constant of integration, the


solutions can exist in an infinite domain or be restricted to a
finite one. This can be understood through the analysis of the
critical points of Eq. (18). For 0 < a < 2, two critical points
with opposite concavities can be found, giving rise to period-
ical solutions. In the case of a > 2, only one critical point has
physical sense (h > 0) and due to its positive concavity, the
solution develops an infinite slope for z finite. The critical
case occurs when a ¼ 2, with a solution given by
1
h+ ðzÞ ¼ lim h0 ðzÞ ¼ cosh ð2zÞ:
Z¼0 2
Ignoring the fact that the unique solution that can satisfy the
FIG. 8. (Color online) Steady envelope solutions arising from Eq. (16) for
different Ohnesorge numbers. Equation (16) was integrated numerically in
boundary condition at z !&1 occurs when a ¼ 1 due to the
the &x direction starting from a point near the separatrix for x * 1. Oscilla- loss of the dynamic term @ zh, the solution h*(z) can be
tions appear for Z < 0.5. regarded as the separatrix of Eq. (16) for Z ¼ 0. Thus, its
minimum defines the maximal lower bound for the neck
$ % thickness for any given Z > 0.
ð0Þ 3 3 3
lim h ¼ z2 & z ln z þ z
z!1 f 32Z 28Z 28Z
3
þ ln2 z þ Oð1Þ: (17) B. Rim solutions
98
It is imperative to look for a different kind of solution
The rise of this manifold is very important for calculations as
that supports a mass influx into the rim. Inspired by the nu-
it is the signature of the existence of another spatial region in
merical simulations and by the translational invariance of
which different expansions and scalings are required. In
lubrication equations, we seek a quasi self-similar solution
addition, we observe that this manifold exhibits a finite
of the form
neck for small enough Ohnesorge number, i.e., Z < 0.5 (see
Figure 8), which shows that no pinch-off can occur in the h i
pure 2D dynamics of capillary retracting films. Indeed, even hr ðx; tÞ ¼ R hðr0Þ ðfÞ þ hðr1Þ ðf; tÞ ; (19)
in the zero Ohnesorge number limit, we observe that the h i
minimal film thickness tends to one half (see Figure 9). This ur ðx; tÞ ¼ R&1 uðr0Þ ðfÞ þ uðr1Þ ðf; tÞ ; (20)
can be proved by setting Z ¼ 0 in Eq. (16). The resulting
equation can be integrated after multiplying it by h&2@ zh, where f : R&1(x & xr) : R&1z. The free spatial origin of the
ð1Þ
yielding solution xr has been introduced. The correction fields hr ,
ð1Þ
ur are small and R * 1 depend weakly on time. Theoreti-
cally, we can understand this scaling as a consequence of the
appearance of a natural length scale in the rim given by the
radius R. The fact that ur scales with R&1 is related with the
persistence of the flux quantity hu ¼ 1 into this region as this
quantity cannot depend on the R length scale.
At zeroth order, Eqs. (5) and (6) become, respectively
(subscripts and superscripts omitted again),
_ & RRf@
RRh _ f h þ @f ðhuÞ ¼ 0; (21)
2 3
2
6h@ff h þ ð@f hÞ þ17
@f 4 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi3 5 ¼ 0: (22)
1 þ ð@ f h Þ2

The second equation is decoupled from the first one and also
independent of the choice of R. By contrast, the continuity
equation depends on the quantity RR. _ This equation gives
FIG. 9. (Color online) When the Ohnesorge number is Z < 12, a neck is
formed behind the rim. The figure shows the neck thickness dependence on
physical and suitable solutions for our problem only if
the Ohnesorge number. Thickness values were obtained by numerical inte- RR_ ¼ Oð1Þ in time, due to a least degeneracy argument.28
gration of Eq. (16), starting from a far point on the envelope (17). This distinguished limit breaks the invariance f ! &f and

Downloaded 07 Dec 2011 to 200.89.69.173. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions
122101-7 Asymptotic behavior of a retracting 2-D fluid sheet Phys. Fluids 23, 122101 (2011)

allows solutions with a net flux at one side and no flux at the
other one.
This system is also invariant under the two following set
of transformations: h ! R0h, f !f/R0, u ! R0u and f
!f & f0, u ! u þ RRf _ 0 . The second equation can in fact
easily be integrated
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
hðfÞ ¼ R20 & ðf & f0 Þ2 : (23)

In terms of z and t, we can then write


qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
hðz; tÞ ¼ R2 & ðz & RÞ2 ; (24)

where f0 has been fixed to R0 and the constant R0 has been


absorbed into R. This solution captures the dynamics
observed in the growing rim and its circular shape. We
remark that the power law dependence of R(t) appears natu-
1 FIG. 10. (Color online) The solution of Eq. (31) (black dashed curve)
rally, as RR_ ¼ Oð1Þ implies R / t2 . Moreover, the result is unveils the shape of the intermediate region that links the steady solution
independent of the Ohnesorge number because, at this order, with the growing rim. The asymptotic behavior when x ! 0 should coin-
the momentum is dominated by the surface tension term cide with the envelope of Eq. (16) for x ! 1, while the behavior when x
! 1 should match the solution of the growing rim when t1=6 / x / t1=2 ,
only. However, its validity is restricted to a finite domain i.e., f ! 0.
0 , z , 2R. While the right side z - 2R gives the correct
description of the rim tip, it does not represent the physical 1 1
where n . t&6 ðx & xm Þ . t&6 z and hð1Þ ð1Þ
m and um are correction
solution on the left side when z ! 0 where the rim should terms. Indeed, the substitution of Eqs. (27) and (28) into
connect to the film. The rim solution exhibits a singular Eqs. (5) and (6) gives at leading order the following set of
behavior for z ! 0 that can be expanded equations (subscripts and superscripts have been omitted
1 1 1
" 1 3# again),
limþ hðr0Þ ðzÞ ¼ ð2qÞ2 t 4 z2 þ O t&4 z2 : (25)
z!0
@n ðhuÞ ¼ 0; (29)
" #
Here, we have expressed our result in terms of q, where h@nn h þ ð@n hÞ2
q2 ¼ 2RR._ The value of q can be found by imposing the @n þ 4Z@n ðh@n uÞ ¼ 0: (30)
j@n hj3
boundary condition hu ¼ 1 at z ¼ 0, which is equivalent to
enforcing the mass conservation. Straightforward calcula- The continuity equation can be integrated using the boundary
tions give condition hu ¼ 1 at n ! 0 as the flux should match the far-
rffiffiffi field incoming flow. Replacing u by means of h in the mo-
2 mentum balance equation gives us, as well, an equation which
q¼ : (26)
p can be integrated again. Straightforward calculations provide
It can easily be seen that this rim solution cannot be matched $ %
@n h h@nn h þ ð@n hÞ2
to the film separatrix solution obtained above at the dominant 4Z ¼ : (31)
order. Therefore, an intermediate region with a different h j@n hj3
self-similar scaling is required in order to match the two
The constant of integration is zero because of the boundary
regions.
conditions imposed at n ! 1 and deduced from the match-
C. Intermediate region
ing with the rim solution. In this limit, the function must
1
match the rim solution so hðm0Þ - n2 . Equation (31) is separa-
The intermediate region will have to match for z ! þ1 ble and it can be written as
with the rim solution (that scales in the intermediate region
1 1 ð $ %1
like t4 z2 ) and for z ! 0 with the separatrix (that scales there 8Z &3 8Z &3 2
like z2). We look for self-similar solutions of the form h a þ h dh ¼ n & n0 ;
3 3
hðz; tÞ ¼ ta f ðtzb Þ that meet the time dependence of the match-
ing, that is, a & 12 b ¼ 14 and a & 2b ¼ 0. Hence, the unique where a is a constant of integration. The integral can be
self-similar scaling, which allows a matching between the expressed in terms of hypergeometric functions after a
far-field and the rim solutions, is change of variables. Then, the solution of the equations
h i acquires a closed form
1
hm ðx; tÞ ¼ tþ3 hð0Þ
m ð nÞ þ hð1Þ
m ð n; tÞ ; (27) "$ %1 #
h i &1 3 2n
1 ð0Þ hðnÞ ¼ aN ; (32)
um ðx; tÞ ¼ t&3 um0 ðnÞ þ uð1Þ m ðn; tÞ ; (28) 8Z a12

Downloaded 07 Dec 2011 to 200.89.69.173. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions
122101-8 Gordillo et al. Phys. Fluids 23, 122101 (2011)

where h(0) ¼ 0 has been imposed. N&1 represents the inverse


function of
$ %
1 1 1 7 3
NðgÞ ¼ 2g 2 F1 & ; ; ; &g ;
2 (33)
2 6 6

or equivalently,
2 $ %
23 p2 1 2 1 1
NðgÞ ¼ , & '-3 þ g2 2 F1 & ; & ; ; &g&3 : (34)
3 C 23 2 3 2 3

Both expressions have been included because the series


of the hypergeometric function 2F1 converges only inside the
unitary circle. Therefore, each expression will be useful for
matching the intermediate solution either with the far field or
with the growing rim regions, respectively.
Expanding the function N into a series of g, and invert-
ing it in terms of N, we get, at leading order, in terms of the FIG. 11. (Color online) Maximum slope evolution from 1D numerical simu-
lations for four different Ohnesorge numbers. The curves were rescaled by
physical variables, 1
Z3 following the analytical prediction. The asymptotic behavior is consistent
with the asymptotic analysis of the 1D equations. The constant l, which
$ %1 $ %1
3a3 4 1 1 2Za5 4 5 &1 defines the vertical shift of the curve, was obtained from the asymptotic
lim hðm0Þ ðzÞ ¼ t 4 z2 & A 0 t12 z 2 analysis and not by curve fitting.
z!1 2Z 3
$ 3 %1=2 " 3# D. Asymptotic matching
Za 1
þ t2 z&1 þ O z&2 ; (35)
6 The three solutions that we have obtained can be matched
when the free coefficients satisfy the following relations:
3 2 & ' $ %13
lim hðm0Þ ðzÞ ¼ z þ O a&3 Z&4 t&1 z8 ; (36) 16Z
z!0 32Z a¼ ; xr ¼ xm ¼ xf ¼ x0 ;
2
3p
23 p2
where A0 ¼ 3 is the constant term of Eq. (34) (see Fig-
3½Cð23Þ) as is shown in Figure 10. Indeed, as the lubrication equations
ure 10). are invariant under translations, we can freely choose x0 ¼ 0.
The previous results show that the intermediate region is With the previous results, we can build a zeroth order
always driven by viscous and surface-tension terms. Viscos- non-uniform solution of lubrication equations (5) and (6)
2
ity controls the bridge length, which scales with Z3 (see Sec. with proper boundary conditions (7) and (8),
III D for the a value) and expands in time following a 16th " 1# 1 " 1#
1 ð 0Þ
power law. hðx; tÞ ¼ t2 hðr0Þ xt&2 þ t3 hðm0Þ xt&6 þ hf ð xÞ
While the far-field region of the film can be easily
3 2 1 1 1
identified in the interface as the unperturbed zone, the rim & x & ð2qÞ2 t4 x2 þ h1 ðx; tÞ; (38)
region is pointed out as the swelling zone close to the tip. 32Z
In a similar way, the intermediate region can be recog- where the perturbative field h1(x,t) comes from the sum of
nized as the zone where the slope maximum takes place. ð1Þ 1 1 ð 1Þ ð0Þ
the perturbative fields hf , t3 hðm1Þ , and t2 hr . The function hr
This maximum is due to the change of curvature orienta- comes from Eq. (23) after substituting f0 ¼ &R0 and R0 ¼ q,
tion required between the far-field and the rim region. In while hð0Þ
m is defined by Eq. (32) after applying matching
fact, the maximum slope of the interface at dominant order ð0Þ
conditions. On the other hand, the function hf comes from
is given by the solution of Eq. (16) imposing (17).
. + Non-uniformity of the solution is due to the multiscale
@ 1 1
max hðx; tÞ ¼ lZ&3 t6 : (37) nature of the problem. A hierarchy of equations can be built
x @x up and the errors for each region follow different time scales
The constant l can be found directly from Eqs. (32) and since it corresponds to different spatial time-dependent scales.
(33). Figure 11 shows the results from 1D numerical simula- It is also important to notice that the convergence is very
tions for the evolution of the film maximum slope for four slow—the next term on the far-field solution behaves as x ln x.
1

different Ohnesorge numbers. The collapse of the four curves This asymptotic tail adds an error of order t2 ln t into the rim
into a single one for large times is proof of the existence of region. Therefore, a higher-order analysis is strongly required.
this region. It has to be noticed, here, that because of the
rapid variation of the profile near the neck, the lubrication
E. Higher-order analysis
approximation is in general not valid anymore in this region.
This discrepancy is expected to be more relevant for small The calculation of higher-order terms calculations is
Ohnesorge numbers where inertia might become important. complex. In the intermediate region, the integrals are very

Downloaded 07 Dec 2011 to 200.89.69.173. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions
122101-9 Asymptotic behavior of a retracting 2-D fluid sheet Phys. Fluids 23, 122101 (2011)

details). In Figure 13, we show the asymptotic evolution of


the interfaces obtained within this approach for several
Ohnesorge numbers. In order to validate these solutions and,
particularly, to estimate their accuracy, we require compari-
sons between these interface profiles and numerical simula-
tions obtained with both the 1D model and with the full
Navier-Stokes equations.

IV. COMPARISON AND DISCUSSION


FIG. 12. (Color online) Asymptotic matching between the rim and the inter-
mediate regions for Z ¼ 1.0 and t ¼ 102 to 802 time units. The rim swells as Since the previous analytical results correspond to an
the flow is injected at x ¼ 0. The region near the origin barely changes dur- expansion in the large time asymptotic regime, one expects
ing the motion. the agreement—between numerical simulations and
theory—to increase with time. Hence, we performed numeri-
hard to obtain analytically because they are related to non- cal simulations for both the full diphasic Navier-Stokes
tabulated functions. However, some results can still be equations (3) and (4) and the 1D film equations (5) and (6) in
deduced by numerical integration. order to compare the interface profiles with the asymptotic
In Appendix A, we show how a hierarchy of linear equa- analysis developed in Sec. III for different times. We con-
tions for higher-order terms in each region can be found. sider the two Ohnesorge numbers Z ¼ 0.14 and Z ¼ 0.7,
Nonlinearities appear as forcing terms that couple each equa- investigating the two configurations with and without a neck,
tion with lower-order terms. The tip position, which can also respectively. Figure 14 shows the different profiles at three
be expanded asymptotically in time, adds extra coupling different times showing a reasonable qualitative agreement
terms in the equations. between the numerics and the asymptotic analysis. In partic-
The first non-trivial solution of the hierarchy
& 'that fol- ular, the agreement is particularly good (quantitatively)
lows the zeroth order solution appears for O t&1=3 correc- between the asymptotic expansion of the 1D film equations
tions in the rim region. The inclusion of this correction and the numerics of the same equations, while some differ-
suppresses an asymptotic tail error which adds an error that ences appear for the full Navier-Stokes equations. In addi-
1
grew as t6 in the rim region. In Figure 12, we plot the inter- tion, we want to emphasize that the asymptotic expansion
face at different times for Z ¼ 1.0 using this higher-order limit, developed in Sec. III, gives a good analytical descrip-
term. The far-field region has not yet been included. Solu- tion of the film solution within the 1D long wavelength
tions are very similar to those obtained near the rim by approximation. Moreover, the discrepancy between the 1D
Sünderhauf et al.17 model and the full Navier-Stokes 2D equations can be
It can be shown& that 'the next term of the expansion in explained both by the long-wave approximation of the 1D
the rim region is O t&1=2 . At this level, the first correction equations which is not valid for j@x hj * 0 (i.e., in particular
for the tip position is obtained (see Appendix B for further near the neck as discussed in Sec. II C), but also because of
the free surface boundary conditions—Navier-Stokes equa-
tions account also for the gas dynamics.
In fact, we remark that in the diphasic case, a dynamical
instability develops for large times and low Ohnesorge num-
bers, i.e., Z , 0.018 6 0.003, as was already reported by two
of us.25,29 An example of such an instability is depicted in
Figure 15(c). Such an instability is thus associated with the
presence of the surrounding gas and it is related to the motion
of a liquid rim within this gas.30,31 We argue here that the
appearance of this destabilizing mechanism can be understood
as the onset of a Kelvin-Helmholtz instability due to the strong
shear layer created by the liquid retraction surrounded by the
still gas. Indeed, when the gas is neglected, as is in the case for
the 1D film equation for free surface flows, the development of
the instability is completely suppressed in numerical simula-
tions. The Kelvin-Helmoltz instability is in fact often invoked
to explain atomization of liquid sheets, although in such cases
only a small rim has been formed.5,10,32–35 The growth rate of
the Kelvin-Helmholtz instability for a diphasic viscous flow
without gravity is given by
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
FIG. 13. (Color online) Asymptotic solutions for four different Ohnesorge qL qG
x¼ kDU & "k2 ;
numbers, Z ¼ 0.1, 0.5, 1.0, and 2.0, in the comoving frame. The figures ðqL þ qG Þ2
show the evolution between t ¼ 102 and t ¼ 802 in dimensionless time units.

Downloaded 07 Dec 2011 to 200.89.69.173. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions
122101-10 Gordillo et al. Phys. Fluids 23, 122101 (2011)

FIG. 14. (Color online) Comparison between the numerical simulations and the analytical approach of the evolution of the film thickness for two different
Ohnesorge numbers: Z ¼ 0.7 (1) and Z ¼ 0.14 (2). The initial instability behind the rim and the rim curvature and size are practically the same in the three cases.
From top to bottom in the legends, the curves correspond to: Navier-Stokes simulations (black line), thin film approximation, light gray line (blue online), and
asymptotic expansion, gray line (red online). The insets show a zoom for the neck region. Simulation times are shown in the legend of each of the figures.

sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
where k is the wavenumber (k ¼ 2p/k, with k the wave- 1 qL qG
length), " ¼p lLffi/q kmax ¼ :
ffiffiffiffiLffiffiffiffiffiffi is the kinematic viscosity and 2Z ðqL þ qG Þ2
DU ’ UTC ¼ c=qL e. In the dimensionless units introduced
above, we obtain pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Taking kmax > ki ¼ 1 & 4Z2 , the wavenumber of the spa-
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi tial oscillation of the liquid film behind the rim, we obtain
qL qG
x¼ k & Zk2 : that the Kelvin-Helmoltz instability develops for
ðqL þ qG Þ2
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
The Kelvin-Helmoltz instability develops thus for 1 qG
qffiffiffiffiffiffiffiqffiffiffiffiffiffiffi Z < Zlim ¼ :
k < k0 ¼ Z1 ðq qþq L G
Þ2
, with the most unstable wave-number 2 qL þ qG
L G

FIG. 15. (Color online) Sheet profile and vorticity


field for three different Ohnesorge numbers at
t ¼ 10.6: (a) Z ¼ 0.14, (b) Z ¼ 0.03, and (c)
Z ¼ 0.005.

Downloaded 07 Dec 2011 to 200.89.69.173. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions
122101-11 Asymptotic behavior of a retracting 2-D fluid sheet Phys. Fluids 23, 122101 (2011)

For the air-water system studied in our simulations, we ) " #2 *32


1
obtain Zlim ¼ 0.01714 in very good agreement with our nu- Xr ðfÞ ¼ ð 0Þ
1þ @f hðr0Þ :
merical estimate Z , 0.018 6 0.003. Finally, we would like hr
to emphasize that even in these unstable cases, no sheet
The kernel space of the linear operator is 2-D and can be
break-up is observed.
built up using the solutions of the leading order equation
V. CONCLUSION
found in Eq. (23) and is given by
( )
ð0Þ ð0Þ ð0Þ ð0Þ
In this article, we have studied the 2D capillary retrac- q & h r @f h r q þ hr @f hr
tion phenomenon of a viscous fluid film. Two different ker Lr ¼ vr ¼ ð 0Þ
; wr ¼ ð 0Þ
:
hr hr
scenarios have been considered: a film surrounded by a
viscous gas and a film with free surface boundary condi- The non-homogeneous linear differential equation can be
tions. For the first case, we have solved numerically the solved using the method of variation of parameters. The
diphasic full 2D Navier-Stokes equations (3) and (4) using Wronskian of this set of solutions is
GERRIS.22 For the second case, we have studied numeri-
cally and analytically the thin film 1D model given by Wr ðfÞ ¼ &2q hðr0Þ &2 :
Eqs. (5) and (6). An analytic expansion of the interface is
obtained for this model using a matched asymptotic Thus the general solution of the ordinary differential equa-
method in the large time regime. We report a good agree- tion (ODE) can be obtained from
ment between these analytical results and numerical simu- $ð %
lation for the diphasic and the free surface flow cases. It is ð nÞ 1 2 ð0Þ &2 ð nÞ
hr ðfÞ ¼ q hr wr pr df vr
quite remarkable that our analysis shows that no pinch-off 2
$ð %
can occur in the 2D retraction. Moreover, the retracting 1 2 ð0Þ &2 ðnÞ
& q hr vr pr df wr
film profile exhibits three well separated domains: the rim, 2
the film, and an intermediate region, which connects them. & qPðrnÞ þ AðrnÞ vr þ BðrnÞ wr :
The three regions have different length scalings for large
times (t1/2, 1, and t1/6, respectively). Finally, we have con- It can be noticed that the solution space has an extra dimen-
sistently explained the destabilizing mechanism arising at ðnÞ
sion given by a constant function associated with Pr . Any-
low Ohnesorge number for diphasic systems in terms of a way, this subspace does not interact directly, by means of
Kelvin-Helmoltz instability due to the shear between the integrals, with the lower order terms.
liquid and the surrounding gas. Similarly, one can find that for the intermediate region
the hierarchy is
ACKNOWLEDGMENTS
Lm hðmnÞ ¼ Xm ðnÞpðmnÞ ðnÞ; (A2)
We want to thank Stéphane Zaleski, Peter Mason, and
colleagues from the Institut Jean Le Rond D’Alembert at where again pðnÞ ðiÞ ðiÞ
m depends on n through hm and um with
UPMC for valuable discussions. L.G. and C.J. acknowledge i ¼ 0…(n & 1) and
the SCAT projet. L.G. also thanks the financial support of 1 " #3
CONICYT through its project ACT 127 and C.J., the support Xm ðnÞ ¼ ð0Þ @n hðm0Þ :
hm
of the Agence Nationale de la Recherche through its
Grant DEFORMATION ANR-09-JCJC-0022-01 and of the Now, the 2-D kernel can also be constructed from solutions
Programme Émergence(s) of the Ville de Paris. of the zeroth order problem and can be chosen to be
. ) *+
a ð 0Þ 1 ð 0Þ 1 ð 0Þ
ker Lm ¼ vm ¼ @n hm ; wm ¼ hm & n@n hm :
APPENDIX A: HIERARCHY OF EQUATIONS q a 2

Let us suppose that the corrective terms in Eqs. (12), Straightforward calculations show that the Wronskian of the
(13), (19), (20), (27), and (28) can be expanded in a power differential operator is
series of t. A hierarchy of equations can then be written for 3p h ð0Þ i2 h ð0Þ i4
each region. For each order, a linear differential operator Wm ðnÞ ¼ h @n hm :
4q m
acting on the n-th corrective term may be balanced with non-
linearities coming from lower-order terms. The free coeffi- By taking these results into account, a closed solution can be
cients can then be fixed by matching leading terms with found explicitly
those coming from other regions. $ð h i&1 %
4q
For the rim region, the hierarchy adopts the form hðmnÞ ðnÞ ¼ & hðm0Þ 3 @n hðm0Þ wm pðmnÞ dn vm
3p
$ð h i&1 %
Lr hðrnÞ ¼ Xr ðfÞ pðrnÞ ðfÞ; (A1) 4q
þ hðm0Þ 3 @n hðm0Þ vm pðmnÞ dn wm
ðnÞ ðiÞ ðiÞ 3p
where pr depends on f through hr and ur with
i ¼ 0…(n & 1) and þ AðmnÞ vm þ BðmnÞ wm þ PðmnÞ pm þ QðmnÞ #m :

Downloaded 07 Dec 2011 to 200.89.69.173. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions
122101-12 Gordillo et al. Phys. Fluids 23, 122101 (2011)

which shows that viscosity and inertia effects are incorpo-


rated at this order through the forcing term. Plugging the pre-
vious expression into Eq. (A1) and solving the integrals, one
can find that the solution near f ¼ 0 behaves as

2Z &1 &1 1 1 1
lim hðr1=2Þ ðfÞ ¼ q f & pffiffi q2 f&2 ln f
f!0 3 4 2
h pffiffi 1 ð1=2Þ i &1
þ b þ 2q2 Br f 2þc
" 1 #
þ O f2 ln f ; (B1)

where b and c are constant terms that depend on Z and c1/2.


ð1=2Þ
The coefficient Ar was fixed to zero as the term associated
with wr(f) cannot fulfill the boundary condition when f ¼ 2q.
When Eq. (B1) is written in the physical space, one finds
that the first term matches the third term of expansion (35).
FIG. 16. (Color online) Solution space associated with the hierarchy linear
operator of the intermediate region. The functions vm and wm are related to
By contrast, matching the second term needs further calcula-
the translational and rim-size invariance, respectively. tions. If any matching is possible, & this' term would corre-
spond to the first of the O t&1=6 corrections in the
The functions pm and #m come from the constants of integra- ð1=2Þ
intermediate region. Moreover, the constant Br would be
tion of the mass and momentum conservation equations. They
fixed by matching
& the
' third term of Eq. (B1) with the next
also increase the solution space dimension by two, as was
term of the O t&1=6 corrections.
expected, without direct interaction with nonlinearities. The
After some straightforward calculations on this order of
four functions of the solution space are plotted in Figure 16.
the intermediate region hierarchy, we find that the forcing
The boundary condition at the tip may also pose some
term on Eq. (A2) is
problems because of the inherent divergence of the deriva-
tive. To avoid them, we promote the parameter x0 as a weak 1
function of time. As a result, the tip position will play some pðm1=6Þ ¼ ð0Þ
;
ðnÞ hm
role in the hierarchy equations inducing extra terms for pr
and pðnÞ
m in Eqs. (A1) and (A2). which comes from inertial forces. If we replace the last
The first non-trivial solution of the hierarchy
& 'that fol- expression in Eq. (A2), we can obtain the following order
lows the zeroth order solution appears for O t&1=3 correc- correction in terms of the variable g:
ð1=3Þ
tions in the rim region. In this case, pr ¼ 0 in Eq. (A1)
and this linear equation become homogeneous. Boundary 2q
conditions impose a solution of the form hðm1=6Þ ðnÞ ¼ ½UðgÞvm ðgÞ & HðgÞ wm ðgÞ)
3a
hðr1=3Þ ðfÞ ¼ Bðr1=3Þ vr ðfÞ: þ Aðm1=6Þ vm ðgÞ þ Bðm1=6Þ wm ðgÞ
ð1=3Þ
The constant Br can be obtained by matching the first
term of hr
ð1=3Þ
around f ¼ 0 with the second term of hð0Þ which depends on hð0Þ ð0Þ
m through g ¼ hm =a. The terms con-
m ð1=6Þ ð1=6Þ
around n ¼ 1 in the physical space. This yields taining Pm and Qm constants have been omitted as it
can be shown that the coefficients are equal to zero to ensure
$ %1
pZa5 4 the matching.
Bðr1=3Þ ¼ &A0 : The functions U(g) and H(g) are related to the integrals
12
ðg
& '1
APPENDIX B: HIGHER-ORDER TERMS UðgÞ ¼ t&3 1 þ t&3 2 dt;
ðg ) *
& The' next term of the expansion in the rim region is &2
& &3
' 1 NðtÞ
O t&1=2 . The effect of the tip motion has been added by HðgÞ ¼ t 1 þ t t& dt;
2 N0 ðtÞ
expanding P the tip position into a series, where
x_ 0 t ¼ & cd td with d , 12. The existence of the first term, which can be expanded analytically into series around g ¼ 1
c12 which plays a role at this order,& was 'suggested by Song and g ¼ 0þ. Thus, we can calculate the asymptotic behavior
and Tryggvason,21 who reported O t&1=2 corrections in the ð1=6Þ
of hm when g ! 1,
neck velocity on their numerical simulations.
$ pffiffi %
Thus, the forcing term in Eq. (A1) can be written as ð1=6Þ 3 ð1=6Þ 3pq q
follows: lim hm ðgÞ ¼ Bm & g & g&1 ln g
g!1 4 7aA0 2a
ð " # $ pffiffi %
ð1=2Þ ð0Þ ð0Þ 1 ð1=6Þ A0 ð1=6Þ 2 3qp q &1
pr ¼ c1=2 hr ur & f@f hðr0Þ uðr0Þ df þ Am & Bm þ & g
2 2 21a 4a
& &2 '
þhðr0Þ uðr0Þ 2 & 4Zhðr0Þ @f uðr0Þ ; þO g ;

Downloaded 07 Dec 2011 to 200.89.69.173. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions
122101-13 Asymptotic behavior of a retracting 2-D fluid sheet Phys. Fluids 23, 122101 (2011)

ð1=6Þ
and the asymptotic behavior of hm when g ! 0, From the former of these expansions, we can already notice
that the matching with the outer region is possible only if
$ % 4pq
2q 1 4q 1 Bðm1=6Þ ¼ pffiffi :
lim hðm1=6Þ ðgÞ ¼ & g2 ln g þ Aðm1=6Þ & g2
g!0 7a 49a 7 3aA0
" 7 # Furthermore, since we know the expansion of g in terms of
þ O g2 ln g :
n, the last result implies that

) *
1 1 &12 1 1 1 2Za 8a 4a ð1=6Þ &12 & '
lim hðm1=6Þ ðnÞ
¼ & pffiffi q n ln n þ pffiffi q ln 2 þ d þ Am
2 2 n þ O n&1 ; (B2)
n!1 4 2 4 2 2 3e 3q q
" $ %12 #
ð1=6Þ 3 3 32Za 3 ð1=6Þ
& '
lim hm ðnÞ ¼ & n ln n þ ln 2=7 þ Am n þ O n7 ln n : (B3)
n!0 28Z 56Z 3e 32Za

where d is a constant depending on Z. When these expan- remanent time logarithmic tails coming from the &secondary'
&1=6
sions are written back into the physical plane, both dominant terms of ' and (B3). The corrections are O t
Eqs. (B2)
& &1=2 ln t
orders match, respectively, the secondary terms of expan- and O t ln t , respectively, and are related to the transla-
ð1=6Þ
sions (17) and (B1). Besides, the constant coefficients Am tional invariance of the solution. This intermediate-order
ð1=2Þ
and Br from Eqs. (B1)–(B3), which have not been fixed matching led to homogeneous equations in the rim and inter-
yet, should meet the following set of two linear equations: mediate regions that can be solved directly.
$ %12 At this point, it seems that there is not any condition for
3 32Za 3 3 the tip position correction c12 at this order. However, a careful
ln 2=7 þ Aðm1=6Þ ¼ ;
56Z 3e 32Za 28Z analysis of the matching conditions shows that the Oð1Þ cor-
1 2Za 8a 4a pffiffi 1 rection in Eq. (B1), c, becomes dominant when the far-field
ln 2 þ d þ Aðm1=6Þ ¼ b þ 2q2 Bðr1=2Þ ; and intermediate regions are matched. In order to annihilate
2 3e 3q q
this term, the first correction of the tip position should satisfy
in order to ensure the asymptotic matching in the whole do- $ %
main. It is quite remarkable that the conditions that deter- 3 2
c1=2 ¼ q 1 & pZ : (B4)
mine the coefficients arrive from the matching of secondary 4 9
terms instead of the dominant ones.
ð1=6Þ ð1=2Þ
After solving Am and Br from the previous set of Taking all of these considerations into account, our second
equations, the second order solution can be built up. Indeed, order asymptotical approach is finally written in terms of
two extra terms should be added to suppress the effect of the z ¼ x þ 2c12t1/2 and given by

" 1# 1 " 1# 1 " 1# 1 " 1#


ð 0Þ 1
hðz; tÞ ¼ hf ðzÞ þ t3 hðm0Þ zt&6 þ t6 hðm1=6Þ zt&6 þ t2 hðr0Þ zt&2 þ t6 hðr1=3Þ zt&2
" 1# $ %1
5 4
ð1=2Þ &2 3 2 3 3 1 1 1 2Za 5 1
þ hr zt & z þ z ln z & z & ð2qÞ2 t4 z2 þ A0 t12 z&2
32Z 28Z 28Z 3
2Z &1 1 &1 1 11 1 " pffiffi 1 #1 1
& q t2 z þ pffiffi q2 t4 z&2 ln z þ b þ 2q2 Bðr1=2Þ t4 z&2
3 4 2
$ %12 " 1# 5 " 1#
1 2a 1 1 1 1 1
& t6 ln t vm zt&6 & ln t wr zt&2 & ln t z & pffiffiffi t4 ln t z&2 :
14 3Z 56 56Z 16 p

1 6
G. I. Taylor, “The dynamics of thin sheets of fluid III. Disintegration of J. Eggers and E. Villermaux, “Physics of liquid jets,” Rep. Prog. Phys. 71,
fluid sheets,” Proc. Roy. Soc. London A, 253, 313 (1959). 036601 (2008).
2 7
F. E. C. Cullick, “Comments on a ruptured soap film,” J. Appl. Phys. 31, M. Rein, “Phenomena of liquid drop impact on solid and liquid surfaces,”
1128 (1960). Fluid Dyn. Res. 12, 61 (1993).
3 8
J. Keller and M. Miksis, “Surface tension driven flows,” SIAM J. Appl. L. Limat, P. Jenffer, B. Dagens, E. Touron, M. Fermigier, and J. Wesfreid,
Math. 43, 268 (1983). “Gravitational instabilities of thin liquid layers: Dynamics of pattern
4
K. Miyamoto and Y. Katagiri, Curtain Coating, edited by S. Kistler and P. selection,” Physica D 61, 166 (1992).
9
Schweizer (Chapman and Hall, London, 1997). X. Hu and A. Jacobi, “The intertube falling film: Part1—flow characteris-
5
N. Bremond, C. Clanet, and E. Villermaux, “Atomization of undulating tics, mode transitions and hysteresis,” Trans. ASME, Ser. C: J. Heat Trans-
liquid sheets,” J. Fluid Mech. 585, 421 (2007). fer 118, 616 (1996).

Downloaded 07 Dec 2011 to 200.89.69.173. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions
122101-14 Gordillo et al. Phys. Fluids 23, 122101 (2011)

10 23
E. Villermaux and C. Clanet, “Life of a flapping liquid sheet,” J. Fluid S. Popinet, “Gerris: A tree-based adaptive solver for the incompressible
Mech. 462, 341 (2002). euler equations in complex geometries,” J. Comput. Phys. 190, 572
11
C. Josserand and S. Zaleski, “Droplet splashing on a thin liquid film,” (2003).
24
Phys. Fluids 15, 1650 (2003). S. Popinet, “An accurate adaptive solver for surface-tension-driven interfa-
12
L. V. Zhang, P. Brunet, J. Eggers, and R. D. Deegan, “Wavelength selec- cial flows,” J. Comput. Phys. 228, 5838 (2009).
25
tion in the crown splash,” Phys. Fluids 22, 122105 (2010). G. Agbaglah, S. Delaux, D. Fuster, J. Hoepffner, C. Josserand, S. Popinet,
13
E. Villermaux and B. Bossa, “Drop fragmentation on impact,” J. Fluid P. Ray, R. Scardovelli, and S. Zaleski, “Parallel simulation of multiphase
Mech. 668, 412 (2011). flows using octree adaptivity and the volume-of-fluid method,” C.R. Méca-
14
G. Debrégeas, P. Martin, and F. Brochard-Wyart, “Viscous bursting of sus- nique 339, 194 (2011).
26
pended films,” Phys. Rev. Lett. 75, 3886 (1995). J. Keller, A. King, and L. Ting, “Blob formation,” Phys. Fluids 7, 226 (1995).
15 27
G. Debrégeas, P. de Gennes, and F. Brochard-Wyart, “The life and death T. Erneux and S. Davis, “Nonlinear rupture of free films,” Phys. Fluids 5,
of bare viscous bubbles,” Science 279, 1704 (1998). 1117 (1993).
16 28
M. Brenner and D. Gueyffier, “On the bursting of viscous films,” Phys. M. V. Dyke, Perturbation Methods in Fluid Mechanics (The Parabolic,
Fluids 11, 737 (1999). Stanford, CA, 1975).
17 29
G. Sünderhauf, H. Raszillier, and F. Durst, “The retraction of the edge of a D. Fuster, G. Agbaglah, C. Josserand, S. Popinet, and S. Zaleski,
planar liquid sheet,” Phys. Fluids 14, 198 (2002). “Numerical simulation of droplets, bubbles and waves: state of the art,”
18
R. Krechetnikov, “Stability of liquid sheet edges,” Phys. Fluids 22, Fluid Dyn. Res. 41, 065001 (2009).
30
092101 (2010). É. Reyssat and D. Quéré, “Bursting of a fluid film in a viscous environ-
19
J. M. Fullana and S. Zaleski, “Stability of a growing end-rim in a liquid ment,” Europhys. Lett. 76, 236 (2006).
31
sheet of uniform thickness,” Phys. Fluids 11, 952 (1999). J. B. Bostwick and P. H. Steen, “Stability of constrained cylindrical interfa-
20
I. Roisman, K. Horvat, and C. Tropea, “Spray impact: Rim transverse ces and the torus lift of Plateau-Rayleigh,” J. Fluid Mech. 647, 201 (2010).
32
instability initiating fingering and splash, and description of a secondary P. Yecko and S. Zaleski, “Transient growth in two-phase mixing layers,”
spray,” Phys. Fluids 18, 102104 (2006). J. Fluid Mech. 528, 43 (2005).
21 33
M. Song and G. Tryggvason, “The formation of thick borders on an ini- T. Boeck and S. Zaleski, “Viscous versus inviscid instability of two-phase mix-
tially stationary fluid sheet,” Phys. Fluids 11, 2487 (1999). ing layers with continuous velocity profile,” Phys. Fluids 17, 032106 (2005).
22 34
S. Popinet, Gerris Flow Solver (supported by the National Institute H. Lhuissier and E. Villermaux, “Soap films burst like flapping flags,”
of Water and Atmospheric Research, Auckland, NZ, and Institut Jean Phys. Rev. Lett. 103, 054501 (2009).
35
Le Rond d’Alembert, Paris, FR, as at October 2011, available at http:// E. Villermaux and B. Bossa, “Drop fragmentation on impact,” J. Fluid
gfs.sourceforge.net/). Mech. 668, 412 (2011).

View publication stats Downloaded 07 Dec 2011 to 200.89.69.173. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

You might also like