Vous êtes sur la page 1sur 11

Chemical Engineering Science 61 (2006) 3986 – 3996

www.elsevier.com/locate/ces

Breakup mode of an axisymmetric liquid jet injected into


another immiscible liquid
Shunji Homma a,∗ , Jiro Koga a , Shiro Matsumoto a , Museok Song b , Grétar Tryggvason c
a Department of Applied Chemistry, Saitama University, 255 Shimo-Okubo, Sakura-ku, Saitama 338-8570, Japan
b Department of Naval Architecture and Ocean Engineering, Hong Ik University, Jochiwon, Chungnam, 339-701, Korea
c Mechanical Engineering Department, Worcester Polytechnic Institute, 100 Institute Road, Worcester, MA 01609-2280, USA

Received 3 October 2005; received in revised form 19 January 2006; accepted 21 January 2006
Available online 10 March 2006

Abstract
The breakup of an axisymmetric liquid jet, injected vertically upward from a nozzle into another immiscible liquid, into droplets is studied
numerically. The unsteady motion of the interface separating two immiscible fluids is followed by solving the Navier–Stokes equations for
incompressible and Newtonian fluids in axisymmetric cylindrical coordinates with a Front-Tracking method. The evolution of the interface and
the specific surface area of the droplets are in good agreement with experimental results. Three breakup modes, dripping, jetting with uniform
droplets, and jetting with non-uniform droplets, are identified. The different modes are shown on a Weber number—the viscosity ratio map.
䉷 2006 Elsevier Ltd. All rights reserved.

Keywords: Drop; Front-Tracking; Jet breakup; Laminar flow; Multiphase flow; Simulation

1. Introduction system is still an attractive phenomenon in science and engi-


neering.
The formation of a liquid jet injected into another immisci- When a liquid is injected from an orifice into another medium
ble liquid and the breakup of the jet into droplets is of funda- several breakup modes are observed, as shown schematically
mental importance in industrial liquid–liquid contact processes, in Fig. 1 (Clift et al., 1978; Kitamura and Takahashi, 1986).
such as solvent extraction. The breakup of the jet increases For low injection velocities droplets are formed periodically at
the interfacial area, hence enhancing heat and/or mass trans- the orifice and no jet is observed. This breakup mode is called
fer, and sometimes chemical reactions. The jet formation and dripping. As the injection velocity is increased to the jetting ve-
breakup has therefore been studied extensively in order to pre- locity, ujet , jetting mode appears. The length of the jet increases
dict the length of the jet and the size of the droplets (Meister with the injection velocity up to umax , at which the jet reaches
and Scheele, 1969a, 1969b; Kitamura et al., 1982; Bright, 1985; its maximum length. Between ujet and umax , axisymmetric cap-
Das, 1997b). illary waves are observed and droplets break off from the tip
The formation of droplets by pinch off has recently generated of the jet. For velocities greater than umax , the jet breaks up
significant interest among researchers examining the scaling primarily by growth of asymmetric disturbances and the length
laws governing the late time evolution and how it depends on of the jet decreases. At still higher injection velocities, the jet
the governing parameters (Eggers, 1997). Studies of immiscible breaks up by atomization, where many non-uniform droplets
fluids include Lister and Stone (1998), Cohen et al. (1999), are formed near the orifice. Around umax or above, the flow
Zhang and Lister (1999), Longmire et al. (2001), Webster and and the motion of the interface are three dimensional, whereas
Longmire (2001), and Milosevic and Longmire (2002). Hence they can be axisymmetric up to umax .
the breakup of a liquid jet into droplets in immiscible liquids We investigate numerically the formation of a liquid jet in-
jected vertically upward into another immiscible liquid and the
∗ Corresponding author. Tel.: +81 48 858 3510; fax: +81 48 858 3510. breakup of the jet into droplets for axisymmetric flow regime
E-mail address: honma@apc.saitama-u.ac.jp (S. Homma). for ũ0 < umax , where ũ0 is the average injection velocity.
0009-2509/$ - see front matter 䉷 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2006.01.029
S. Homma et al. / Chemical Engineering Science 61 (2006) 3986 – 3996 3987

Dripping Jetting (Capillary Breakup) Atomization 2. Mathematical formulation and numerical method

As
2.1. Governing equations

ym
s
ave

me
cW

tric
A sketch of the problem studied here, the formation of an
etri

Wa
Jet Length

mm axisymmetric jet and its breakup into droplets, is shown in

ves
Fig. 2. An incompressible, Newtonian fluid of density j and
isy
Ax

viscosity j is injected vertically upward from a nozzle into a


second quiescent, immiscible, incompressible, Newtonian fluid
of density c and viscosity c , with the average injection ve-
Axisymmetric Flow locity ũ0 . The symbol R0 in Fig. 2 is the inner radius of the
3-Dimensional Flow nozzle.
u jet u max Here we use a single field form of the governing equations,
Injection (Nozzle) Velocity where two different fluids are treated as a single fluid with
variable physical properties. The continuity equation and the
Fig. 1. Schematic diagram of breakup modes when a liquid is injected into unsteady momentum equations for the whole computational
another medium. domain are:
∇ · u = 0, (1)
The breakup mode is our main concern in this paper, so that
the jet length and droplet size, scaling laws near pinch-off, and j
u + ∇ · uu = −∇P + Re−1 ∇ · (∇u + ∇uT )
other interesting issues such as the coalescence of droplets, are jt

not discussed. −1 −1
Related studies are briefly summarized here: the first theoret- + Fr g + We nf (x − xf ) dAf . (2)
f
ical study of the breakup of a liquid jet in another immiscible
liquid was done by Tomotika (1935), who extended the linear Here, u is the velocity vector, P the pressure, and g the
stability analysis of Rayleigh (1879). Tomotika’s equation has gravity. Eqs. (1) and (2) have been non-dimensionalized by
often been used to estimate the jet length and the droplet size for the characteristic scales: the length 2R0 and the time 2R0 /u0 ,
real jets in the axisymmetric jetting mode (Meister and Scheele leading to the three non-dimensional numbers that characterize
1969a,b; Kitamura et al., 1982; Teng et al., 1995; Das, 1997b). the jet dynamics: the Reynolds number, Re = 2R0 ũ0 j /j ,
Meister and Scheele (1969a,b) carried out experiments and de- the Weber number, We = 2R0 ũ20 j /, and the Froude number,
veloped their own drop formation theory, covering dripping to Fr = ũ20 j /2R0 g(j − c ). Here g is the gravitational accel-
axisymmetric jetting. Kitamura et al. (1982) found experimen- eration and  is the interfacial tension. Note that g is always
tally that the Tomotika’s theory predicts well the size of the negative in this study and Fr is thus positive if a lighter fluid is
droplets when the surrounding fluid moves with the same speed injected into a heavier one (j − c < 0). The non-dimensional
as the jet. Bright (1985) performed inviscid stability analysis density, , and viscosity, , are normalized by the density and
that took into account the velocity profiles of both phases. Das viscosity of the jet, respectively:
(1997a,b) combined Tomotika’s theory with his own model of 
jet expansion and contraction to predict the size of the droplets. 1 in the jet,
,  =
Richards et al. (1993, 1994, 1995) carried out direct numeri- ,  in the continuous phase.
cal simulations of the formation of an axisymmetric jet and its
Thus,  is the density ratio,  = c /j , and  is the viscosity ra-
breakup into droplets in liquid–liquid systems, using the Vol-
tio,  = c /j . The interfacial tension is found using twice the
ume of Fluid method (VOF) (Hirt and Nichols, 1981). Apply-
mean curvature, , the unit normal, nf , and a delta function,
ing the same technique as Richards et al. (1993), Xiaoguang
(x − xf ) which is zero everywhere except at the interface, xf .
(1999) investigated numerically the formation of droplets from
Since it is assumed that the system is isothermal and the inter-
a nozzle in the dripping mode, focusing on the formation of
face is completely clean, interfacial tension gradients along the
satellite droplets. These studies have contributed significantly
interface due to the variation of temperature or surfactant con-
to our understanding of the formation of a jet and the breakup
centration are neglected. In order to advect the discontinuous
of the jet into droplets in immiscible liquids systems. How-
density and viscosity fields, we need two additional equations:
ever the breakup mode have not been discussed in detail unlike
liquid–air systems, e.g. Chaudhary and Maxworthy (1980). D
 = 0, (3)
The rest of this paper is organized as follows: The mathe- Dt
matical formulation and the numerical method are presented in
D
Section 2. In Section 3, the numerical method is validated by  = 0. (4)
grid refinement tests and comparison of the simulation results Dt
with experimental ones. In Section 4, the modes of the breakup These equations of state imply that the physical properties of
of a jet into droplets are presented and then the mechanism of each fluid remain constant since the fluids are immiscible and
the breakup of the jet in liquid–liquid systems is discussed. incompressible.
3988 S. Homma et al. / Chemical Engineering Science 61 (2006) 3986 – 3996

(Richards et al., 1993, 1994, 1995; Xiaoguang, 1999), here we


use the Front-Tracking method. Both algorithms have technical
advantage and disadvantage, but both should give the same
results. We will demonstrate the accuracy of the Front-Tracking
method in the next section. A detailed description of the Front-
Tracking method used here can be found elsewhere (Unverdi
and Tryggvason, 1992; Tryggvason et al., 2001). Here, we only
describe the specific modification needed for simulations of the
formation of a jet and its breakup into droplets.
When a jet or an existing droplet breaks up, the interface has
to be divided into two separate interfaces. In the simulations
presented here, we used a simple rupture model where the
interface is divided if it comes within a very small distance
s from the axis of symmetry. In this problem, most of the
breakup occurs there. Although the distance s may depend on
the flow conditions, it was chosen on an ad hoc basis as half of
the smallest grid interval because, as will be seen in the next
section, the jet breakup and the subsequent droplet formation
does not depend strongly on s, when the grid resolution is
high enough. As stated earlier, coalescence of two droplets and
the reunion of a jet with the droplet once broken off from the
Fig. 2. The geometry of the breakup of an axisymmetric jet into droplets. jet, are not accounted for here.
Experimental observations show that the jet surface does not
detach from the lip of the nozzle, hence the interface remains
In the present study, the flow is assumed to be axisymmetric attached to the lip in this study. As in the simulation by Richards
and, as shown in Fig. 2, cylindrical. Axisymmetric coordinates et al. (1993), the contact angle at the lip is not specified.
are therefore used. At t = 0, the continuous phase is at rest, The nozzle is imbedded as a solid object in the computa-
so that the initial velocity is u = v = 0, where u and v denote tional domain. The pressure in the solid is adjusted so that the
the radial and axial components of the velocity, respectively. velocities on the surface become zero. Thus the no-slip bound-
Symmetry boundary conditions are used at the r = 0 axis. The ary conditions at the surface of the nozzle are implicitly sat-
outflow boundary conditions at the top (z = L) are u = 0, and isfied. This technique has already been applied to the problem
jv/jz=0. This boundary condition prevents re-circulation near of molten metal solidification (Che et al., 2004). One of the
the top. If the jet does not breakup and it reaches the top bound- advantages of this technique is that the geometry of the nozzle
ary, the equation of continuity (1) is no longer satisfied and the can be changed without rebuilding the grid structure, and thus
divergence of u becomes very large. Therefore a sufficiently the numerical code remains simple.
long domain, L, is necessary to minimize the influence of the
outflow boundary. It is also necessary to stop the calculation 3. Validation of numerical method
far before the top of the jet or the first droplet reaches the top
of the domain. At the bottom of the nozzle, we apply inflow Preliminary simulations were carried out to select the ratio
boundary conditions, taken to be a fully developed parabolic between R0 and R, the distance between the axis and the right
velocity distribution given by: u = 0, and v = 2ũ0 {1 − (r/R0 )2 }. wall, such that the wall does not affect the flow in any serious
Full slip wall condition is applied on the right wall and on the way. This ratio was determined to be 1:10 in almost all cases.
bottom wall outside the nozzle in order to reduce the effect The aspect ratio of the domain, R/L, must, however, be changed
of the wall boundaries. For the surface of the nozzle, no-slip with the flow conditions. We chose it to be 1/6, 1/16, and 1/24,
boundary conditions are given implicitly rather than specify- depending on the location of the breakup.
ing the velocity zero. A detailed explanation of the boundary Grid refinement tests were also carried out. Fig. 3 shows the
condition at the nozzle will be given in the next section. time dependent length of the jet, Lj for Re = 160, We = 8,
Fr = 32, and  = 1 for several grid resolutions. The aspect ratio
2.2. Numerical methods of the domain is 1/16 in all cases. The breakup times, when Lj
changes abruptly, for 64 × 1024 are very different from those
When the single field form of the governing equations is for the 96 × 1536 and the finer grids, where the breakup times
used, an interface capturing or tracking algorithm is required are almost identical.
to identify the interface between the two fluids. The VOF (Hirt Fig. 4 shows a comparison of the interfacial shape computed
and Nichols, 1981) and the Front-Tracking method (Unverdi using a 128 × 2048 and a 256 × 4096 grid. The condition is the
and Tryggvason, 1992) are examples of successful interface same as the previous case. The shape for the 128 × 2048 grid is
capturing and tracking algorithms. Although the VOF has been exactly the same as that for the 256×4096 grid, so that 128 grid
used earlier for the formation of droplets and jets from a nozzle points are used for the r direction in this study. Note that the
S. Homma et al. / Chemical Engineering Science 61 (2006) 3986 – 3996 3989

20 condition holds the assumption in linear stability theory and it


64x 1024 usually disagrees especially when the continuous phase is at
96x 1536
rest and gravity works to the jet.
128x 2048
16 256x 4096 Tables 1 and 2 summarize, respectively, the systems and the
conditions examined. For all conditions, the dispersed phase
(jet and droplets) is an organic material and is lighter than the
12 continuous phase of a water solution. Hence, the Froude number
is always positive as mentioned in the previous section. The
Lj /(2R0)

dynamics of the interfacial shape are tested using cases 1 (Song


et al., 1999) and 2 (Bright, 1985), highlighting the dripping and
8
jetting modes, respectively. For case 1, the volumes of primary
and satellite droplets are also compared. Specific surface area
is tested using cases 3–20 (Meister and Scheele, 1969a).
4
Fig. 5 shows the unsteady formation of kerosene droplets
in water for the dripping mode. The left-hand side of each
frame shows the simulation result and the right half shows an
0 experimental picture taken by a high-speed video camera of
0 2 4 6 8
t 400 frame/s with 480 ×128 pixel resolution (Song et al., 1999).
A domed interface appears at the early stage (t = 0), and then a
Fig. 3. Grid refinement test for jet length (Re=160, We=8, Fr=32, and  =1). neck is formed (t = 0.6). A cusp is observed behind the primary
droplet just after its break off (t = 0.68). The cusp moves back
toward the nozzle due to the strong interfacial tension at the
tip of the cusp. This portion bounces back again, and a small
satellite droplet is produced (t = 0.7). The droplets rise due to
4
buoyancy and inertial forces (t = 0.78.1.2). Although the lo-
cation of the satellite droplet is slightly different (t = 0.7 and
0.78), overall the motion of the interface is in excellent agree-
ment between the simulation and the experiment. As shown in
3 Table 3, the sizes of both the primary and the satellite droplets
obtained from the simulation are also in good agreement with
the observed sizes in the experiments. Although a satellite for-
256×4096

128×2048

mation is out of the focus in this study, we briefly discuss the


2 relation between the size of the satellite droplet and the pinch
off criterion s. Since the size of the grid mesh is 367 m in
this case and the radius of the experimental satellite droplet is
740 m, s (≈ 184 m) is about 25% of the radius of the satel-
1
lite. Thus it may not be enough to discuss the details of the
formation of the satellite. It is, however, reasonable to discuss
the mode of the breakup of the jet, because s is about 4% of
the radius of the nozzle (4.7 mm).
Fig. 6 shows the simulated shapes of the jet and the droplets
0
for the jetting mode at two different times, as well as a tracing
Fig. 4. Grid refinement test for the shape of jet and droplets (Re = 160, of a jet and droplet shapes from the experimental pictures of
We = 8, Fr = 32, and  = 1). Bright (1985). The simulated jet compares well with the exper-
imental jet. Since the wavelengths of the disturbances in our
simulation and the experiment are almost identical, the vol-
breakup criterion s, which is changed depending on the grid ume of the droplets should agree between the simulation and
resolution, does not affect the jet breakup and the subsequent the experiment. The droplet shapes, on the other hand, differ
droplet formation. slightly. The second and the third droplets in the simulation have
In order to validate the numerical method, several computa- ellipsoidal and round shapes, respectively, while both the first
tions were compared with experiments (Meister and Scheele, and the second droplets in the experiment have a pear like
1969a; Bright, 1985; Song et al., 1999). We, however, did shape. It is, however, difficult to compare the droplet shapes
not compare our computations with linear stability theory exactly, since the motion of the droplets is unsteady.
(Tomotika, 1935). As pointed out by Kitamura and Takahashi Fig. 7 shows the simulated motion of the droplets broken
(1986) and Richards et al. (1994, 1995), the wavelength pre- off from the jet for case 2. The left-hand side of each frame
dicted by linear stability theory is in good agreement with shows the velocity vectors and the right-hand side shows the
the one observed in experiments only when the experimental vorticity contours. Two waves at the tip of the jet breakup
3990 S. Homma et al. / Chemical Engineering Science 61 (2006) 3986 – 3996

Table 1
Systems used in comparison of numerical and experimental results

System Jet and droplets j j Continuous c c 


no. (Dispersed phase) (kg/m3 ) (mPa s) phase (kg/m3 ) (mPa s) (mN/m)

1 Kerosene 890 2.30 Water 992 1.04 36.5


2 n-Decane 732 0.990 Water 999 1.10 22.5
3 70% Paraffin oil 822 6.71 Water 996 0.958 44.8
30% Heptane
4 80% Paraffin oil 843 15.7 Water 996 0.958 45.4
20% Heptane
5 90% Paraffin oil 865 35.3 Water 996 0.958 44.4
10% Heptane
6 Heptane 683 0.393 Water 996 0.958 36.2
7 Heptane 683 0.393 50% Glycerine 1140 6.95 26.0
50% Water
8 Heptane 683 0.393 70% Glycerine 1190 21.9 26.0
30% Water

Table 2
Conditions used in comparison of numerical and experimental results

Case no. System u0 2R0 Re We Fr  = c / j  = c /j


(cm/s) (mm) (dimensionless) (dimensionless) (dimensionless) (dimensionless) (dimensionless)

1 1 0.600 9.40 21.8 0.00825 0.00341 0.45 1.23


2 2 44.0 1.60 521 10.1 33.9 1.11 1.36
3 3 10.0 2.54 31.1 0.466 1.90 0.143 1.21
4 3 20.0 2.54 62.2 1.86 7.59 0.143 1.21
5 3 30.0 2.54 93.3 4.19 17.1 0.143 1.21
6 3 40.0 2.54 124 7.46 30.4 0.143 1.21
7 3 50.0 2.54 156 11.7 47.4 0.143 1.21
8 4 10.0 2.54 13.6 0.472 2.21 0.061 1.18
9 4 20.0 2.54 27.3 1.89 8.85 0.061 1.18
10 4 30.0 2.54 40.9 4.24 19.9 0.061 1.18
11 4 40.0 2.54 54.6 7.55 35.4 0.061 1.18
12 5 10.0 2.54 6.22 0.495 2.65 0.0271 1.15
13 5 20.0 2.54 12.4 1.98 10.6 0.0271 1.15
14 5 30.0 2.54 18.7 4.45 23.9 0.0271 1.15
15 6 33.0 0.813 467 1.67 29.9 2.44 1.46
16 6 39.7 0.813 561 2.42 43.1 2.44 1.46
17 6 45.1 0.813 637 3.11 55.6 2.44 1.46
18 6 51.1 0.813 721 4.00 71.4 2.44 1.46
19 7 20.0 0.813 283 0.854 7.45 17.7 1.67
20 8 20.0 0.813 283 0.854 6.76 55.7 1.74

Min — — 6.22 0.00825 0.00341 0.0271 1.15


Max — — 721 11.7 71.4 55.7 1.74

into two bullet shape droplets [(a) → (b)]. Positive vortici- sponds to the volume-surface mean diameter, d32 , is defined as
ties from the top to middle along the interface of the droplets,  n
cause the deformation of the droplets [(b) → (c)]; The big- n 
Ss = Si Vi (=1/d32 ) , (5)
ger droplet (top) becomes a pear like shape and the smaller
i=1 i=1
one (bottom) becomes a shell like shape. These droplets even-
tually become spherical shapes through the oscillatory motion where n is the number of the resulting droplets, Si the surface
between oblate and prolate shapes (d). In this way, it is difficult area of ith droplet, and Vi the volume of ith droplet. Although
to compare the shape of the droplets with the shape from the the error is relatively large for larger specific surface area, the
still picture. agreement between the computations and experiments is rea-
Fig. 8 shows a comparison of the specific surface area be- sonably good and the error is always less than 20%.
tween the computational results and the experiments of Meister Finally, we note the limitation of our numerical method. In
and Scheele (1969a). The specific surface area, which corre- the bottom of the Table 2, the maximum and the minimum
S. Homma et al. / Chemical Engineering Science 61 (2006) 3986 – 3996 3991

Fig. 6. Comparison of the simulated shape of a jet and droplets with the
experiments for jetting mode (case 2: Re = 521, We = 10.1, Fr = 33.9, and
 = 1.11).

4. Results and discussion

We examined over 80 combinations of the non-dimensional


numbers, selecting existing liquid–liquid systems as shown in
Table 1. The size of the nozzle and the injection velocity were
varied to adjust the non-dimensional numbers. In addition, a few
“virtual” liquid–liquid systems were selected to study different
viscosity ratios. Several computational results highlighting the
Fig. 5. Comparison of the simulated interfacial shape with the experiments different breakup modes are shown in Fig. 9.
for dripping mode (case 1: Re = 19.8, We = 7.5 × 10−3 , Fr = 1.73 × 10−3 , Fig. 9(a) shows the time dependent motion of the interface
and  = 0.45). The unit of the time shown under each frame is second and for a typical dripping mode (Re=377, We=1.28, Fr=8.18, and
the time is adjusted as the first frame is t = 0.
 = 6.67). While no satellite droplets are observed, the motion
is almost identical with the experimental results as shown in
Fig. 5.
Table 3 Fig. 9(b)–(d) show the time dependent motion of the interface
Comparison of numerical and experimental droplet sizes for dripping mode
(Case 1: Re = 19.8, We = 7.5 × 10−3 , Fr = 1.73 × 10−3 , and  = 0.45)
for jetting mode. A jet forms and capillary waves are observed
in these cases, but they are slightly different from each other.
Experimental (mm3 ) Numerical (mm3 ) Error (%) For Re = 402, We = 2.11, Fr = 27.5, and  = 1.79, (Fig. 9(b)),
Primary droplet 482 477 1.04 uniform droplets are produced and the breakup occurs regularly.
Satellite droplet 1.70 1.75 2.94 Similar shapes of the jet and the droplets to the previous case
are observed in Re = 288, We = 3.24, Fr = 52.9, and  = 1.00
(Fig. 9 (c)). However, the resulting droplets are not uniform.
Two nodes of the wave sometime become one droplet in this
values of the non-dimensional numbers are shown. The ap- case. This is a multiple-node breakup identified experimentally
propriate condition must be at least within the range between by Meister and Scheele (1969a) and Das (1997a). For Re=425,
the maximums and the minimums. Special attention is nec- We = 1.70, Fr = 8.33, and  = 100, (Fig. 9(d)), a jet does not
essary for the Froude number, whose sign can be positive or break up due to the size limitation of the computational domain.
negative depending on the direction of the injection or the Since a jet is always unstable and it eventually breaks up into
choice of liquids. In this study, we limit the positive Froude droplets, the jet breaks if the domain size is long enough in
number cases, where a lighter fluid is injected vertically up- z-direction. It is unknown whether uniform size droplets are
ward from a nozzle on Earth. Note therefore that the results produced or not in this case.
presented in this study are valid only for the positive Froude As shown in Fig. 9, four different breakup modes are ob-
numbers. served: (I) dripping, (II) jetting with uniform droplets, (III)
3992 S. Homma et al. / Chemical Engineering Science 61 (2006) 3986 – 3996

Fig. 8. Comparison of the specific surface area between the simulation and
the experiments (Meister and Scheele, 1969a).

Figs. 10–12 are examples of the diagrams: Re–log(j /c ),


Fr–log(j /c ), and We–log(j /c ). Other combinations of
Re–Fr, Re–We, and Fr–We, are not shown here, because the
modes are scattered all over the diagrams and no clear maps
are obtained. Note that the logarithmic scale is used for the
viscosity ratio in order to be symmetric with respect to the
point of  = 1 [log(j /c ) = 0].
Fig. 10 shows the breakup modes on the Re–log(j /c )
diagram. The mode shifts from dripping to jetting as the
Reynolds number increases. This is as expected because the
higher Reynolds number means higher inertial force, which
helps the formation of the jet. When the viscosity of the
continuous phase is higher than that of the dispersed phase
[log(j /c ) < 0], there are relatively clear borders between
the modes (I) and (II), and between the modes (II) and (III).
When the viscosity of the continuous phase is lower than that
of the dispersed phase [log(j /c ) > 0], on the other hand, the
borders are not clear and the mode shifts from (I) to (III) even
at low Re numbers (∼ 50). Hence the modes cannot be clearly
distinguished from this diagram.
Fig. 11 shows the breakup modes on the Fr–log(j /c )
Fig. 7. The simulated shape of droplets, velocity vectors, and vorticity contours diagram. The mode shifts from dripping to jetting as the
(case 2: Re = 521, We = 10.1, Fr = 33.9, and  = 1.11): (a) t = 0.102 (s); (b) Froude number increases, if the viscosity ratio is constant.
t = 0.108 (s); (c) t = 0.120 (s) and (d) t = 0.140 (s).
Since the large Froude number means a high inertial force
relative to the buoyancy force, the jet begins to form as the
Froude number increases. However, there are no clear borders
jetting with non-uniform droplets, and (IV) jetting without between the modes, and the modes cannot be categorized by
breakup. The mode (IV) is observed only in this computational this diagram.
study and will be classified into mode (II) or (III), so that the When the Weber number and the viscosity ratio are selected
modes (I), (II), and (III) are physically meaningful. as the axes, the breakup modes are easily identified as shown
In order to know when these modes appear, we plot all cases in Fig. 12. Mode (I) appears when the Weber number is low;
examined onto the diagrams whose abscissa and ordinate is dripping is not observed for We > 2. In this mode, interfacial
a combination of two independent non-dimensional numbers. tension is dominant over inertial force, so that the jet end tends
S. Homma et al. / Chemical Engineering Science 61 (2006) 3986 – 3996 3993

1000

800

600

400

200

0
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
µ µ

Fig. 10. The breakup mode on the diagram of Reynolds number and viscosity
ratio.

µ µ

Fig. 11. The breakup mode on the diagram of Froude number and viscosity
ratio.

of the non-linear term (convection term) in Eq. (2) to the jet


instability is not so serious in this mode that the hypothesis
of linear stability theory holds and consequently the jet breaks
up into uniform droplets whose size is almost equivalent to
the volume between the nodes of the fastest growing capillary
wave. As the Weber number increases further, the mode shifts
from (II) to (III). The effect of the inertial force begins to
contribute further to the instability of the jet in mode (III). As
Fig. 9. The motion of the interface: (a) dripping (Re = 377, We = 1.28, a result, non-uniform droplets are produced.
Fr=8.18, and  =6.67); (b) jetting with uniform droplets (Re=402, We=2.11,
When the continuous phase is more viscous than the dis-
Fr = 27.5, and  = 1.79); (c) jetting with non-uniform droplets (Re = 288,
We = 3.24, Fr = 52.9, and  = 1.00) and (d) jetting without breakup (Re = 425, persed phase [log(j /c ) < 0], mode (II) shifts to mode (IV) as
We = 1.70, Fr = 8.33, and  = 100). the Weber number increases. As we stated earlier, the finite size
of the computational domain prevented us from seeing breakup
to form a big spherical bulb. Mode (II) is observed when the and determining whether the jet will form uniform [mode (II)]
Weber number is moderate. Since inertial force begins to be or non-uniform [mode (III)] droplets. According to the lin-
important in this mode, a jet forms. However, the contribution ear stability theory for an infinite liquid column suspended in
3994 S. Homma et al. / Chemical Engineering Science 61 (2006) 3986 – 3996

computationally determined boundaries between the different


modes are shown. The open square (case 1) shows where Song
et al. (1999) found a dripping mode and the black square
(case 2) marks the location where Bright (1985) found jetting.
Both fall in the right regions on the diagram. Black circles
with error bars show where a jet starts to form according
to the experimental study of Scheele and Meister (1968a,b).
The Weber numbers are computed using the jetting velocity
(ujet ). While there is some scatter, overall the experimental
results agree reasonably well with our numerical ones. There
are two major disagreements: A [log(j /c ) = 0.36] and
B [−1.8 < log(j /c ) < −1.2]. For A, Scheele and Meister
(1969a,b) used a butyl alcohol–water system, which was more
miscible ( = 1.8 mN/m) than other systems assumed in this
study ( = 20 ∼ 50 mN/m), whereas completely immiscible
system was presumed in our simulations. Thus this can be a
µ µ
reason of the disagreement for A. On the other hand, we have
not yet found any reason of the disagreement for B.
Fig. 12. The breakup mode on the diagram of Weber number and viscosity Although only the cases for positive Froude numbers have
ratio. been examined, we remark on the mode for negative Froude
numbers. According to the experimental study by Webster and
Longmire (2001), where a dispersed phase is injected verti-
cally downward into a continuous phase, a jet is observed
when Re = 34 and j /c = 0.152 [log(j /c ) < − 0.82]. The
Weber number is calculated to be 0.187 from the experimen-
tal condition and the value is much lower than the bound-
(III) or (IV) ary between jetting and dripping for positive Froude numbers
(Fig. 13). This disagreement results from the difference of the
experimental setup from the setup in this computational study;
since the downward gravity works on the jet in addition to the
downward inertial force by injection, a jet forms even when
(II) the Weber number is low for negative Froude numbers. There
are similar disagreements in other experimental studies for neg-
ative Froude numbers (Das, 1997a; Milosevic and Longmire,
2002). Therefore the result in this study is limited to the posi-
(I) tive Froude numbers.

µ µ 5. Conclusions
Fig. 13. Experimental results on the diagram of Weber number and viscosity
Direct numerical simulations, solving the axisymmetric
ratio.
Navier–Stokes equations with a Front-Tracking method were
conducted to study the formation of a liquid jet and its breakup
another immiscible fluid (Tomotika, 1935), the growth rate of into droplets in another immiscible liquid. The simulation
disturbances decreases with increasing viscosity of the external method was validated by grid refinement tests and comparison
fluid. Hence, the jet is stabilized by a viscous continuous phase with experimental results. The motion of the interface and the
and breakup is therefore not observed in our simulations, due specific surface area are in good agreement between the simu-
to the finite size of the computational domain. lations and experiments when a dispersed liquid is injected ver-
Mode (IV) overlaps with mode (III) when the continuous tically upward into a continuous phase. Three breakup modes,
phase is less viscous than the dispersed phase [log(j /c ) > 0]. dripping, jetting with uniform droplets, and jetting with non-
Because of the same reason mentioned above, we could not uniform droplets, are identified and well mapped on a diagram
determine whether the jet will form uniform droplets or non- using the Weber number and the viscosity ratio as axes.
uniform ones. However, we expect that the mode (IV) in this
region will be mode (III).
Fig. 13 shows experimental results(Scheele and Meister Notation
1968a,b; Bright, 1985; Song et al., 1999) plotted on the same
diagram as used for the computational results in Fig. 12. Af area of the interface
The computational results are omitted in the figure but the d32 volume-surface mean diameter
S. Homma et al. / Chemical Engineering Science 61 (2006) 3986 – 3996 3995

Fr Froude number; Fr = ũ20 j /2R0 g(j − c ) References


g gravitational acceleration
g gravity (vector) Bright, A., 1985. Minimum drop volume in liquid jet breakup. Chemical
Engineering Research and Design 63, 59–66.
L axial length of the domain
Chaudhary, K.C., Maxworthy, T., 1980. The nonlinear capillary instability of
Lj length of the jet a liquid jet. Part 3. Experiments on satellite drop formation and control.
n the number of droplets Journal of Fluid Mechanics 96, 287–297.
nf unit normal at the interface Che, J., Ceccio, S.L., Tryggvason, G., 2004. Computations of structures
P pressure formed by the solidification of impinging molten metal drops. Applied
r radial coordinate Mathematical Modelling 11, 127–144.
Clift, R., Grace, J.R., Weber, M.E., 1978. Bubbles, Drops and Particles.
R radial length of the domain
Academic Press, New York.
R0 inner radius of the nozzle Cohen, I., Brenner, M.P., Eggers, J., Nagel, S.R., 1999. Two fluid drop
Re Reynolds number; Re = 2R0 ũ0 j /j snap-off problem: experiments and theory. Physical Review Letters 83,
Si surface area of ith drop 1147–1150.
Ss specific surface area Das, T.K., 1997a. Droplet formation with single and multiple nodes from a
liquid jet in immiscible liquids. Atomization and Sprays 7, 407–415.
t time
Das, T.K., 1997b. Prediction of jet breakup length in liquid–liquid systems
u radial component of the velocity using the Rayleigh–Tomotika analysis. Atomization and Sprays 7,
ũ0 average injection velocity 549–559.
ujet jetting velocity Eggers, J., 1997. Nonlinear dynamics and breakup of free-surface flows.
umax injection velocity where maximum breakup Reviews of Modern Physics 69, 865–929.
length is observed Hirt, C.W., Nichols, B.D., 1981. Volume of fluid (VOF) method for the
dynamics of free boundaries. Journal of Computational Physics 39,
u velocity vector 201–225.
v axial component of the velocity Kitamura, Y., Takahashi, T., 1986. Stability of jets in liquid–liquid systems.
Vi volume of ith drop Encyclopedia of Fluid Mechanics, vol. 2. Gulf, Houston.
We Weber number; We = 2R0 ũ20 j / Kitamura, Y., Mishima, H., Takahashi, T., 1982. Stability of jets in
x position vector liquid–liquid systems. Canadian Journal of Chemical Engineering 60,
723–731.
xf position of the interface
Lister, J.R., Stone, H.A., 1998. Capillary breakup of a viscous thread
z axial coordinate surrounding by another viscous fluid. Physics of Fluids 10, 2758–2764.
Longmire, E.K., Norman, T.L., Gefroh, D.L., 2001. Dynamics of pinch-off in
Greek letters
liquid/liquid jets with surface tension. International Journal of Multiphase
 delta function Flow 27, 1735–1752.
s breakup criterion of the interface Meister, B.J., Scheele, G.F., 1969a. Drop formation from cylindrical jets in
immiscible liquid systems. A.I.Ch.E. Journal 15, 700–706.
 density ratio;  = c /j Meister, B.J., Scheele, G.F., 1969b. Prediction of jet length in immiscible
 curvature liquid systems. A.I.Ch.E. Journal 15, 689–699.
 viscosity ratio;  = c /j Milosevic, I.N., Longmire, E.K., 2002. Pinch-off modes and satellite formation
 viscosity in liquid/liquid jet systems. International Journal of Multiphase Flow 28,
c viscosity of the continuous phase 1853–1869.
Rayleigh, J.W.S., 1879. On the instability of jets. Proceedings of the London
j viscosity of the dispersed phase (jet) Mathematical Society 10, 4–13.
 density Richards, J.R., Beris, A.N., Lenhoff, A.M., 1993. Steady laminar flow of
c density of the continuous phase liquid–liquid jets at high Reynolds numbers. Physics of Fluids A 5,
j density of the dispersed phase (jet) 1703–1717.
Richards, J.R., Lenhoff, A.M., Beris, A.N., 1994. Dynamic breakup of
 interfacial tension
liquid–liquid jets. Physics of Fluids 6, 2640–2655.
Superscript Richards, J.R., Beris, A.N., Lenhoff, A.M., 1995. Drop formation in
liquid–liquid systems before and after jetting. Physics of Fluids 7,
T transpose 2617–2630.
Scheele, G.F., Meister, B.J., 1968a. Drop formation at low velocities in
Subscript liquid–liquid systems: Part I. Prediction of drop volume. A.I.Ch.E. Journal
14, 9–15.
f interface (front) Scheele, G.F., Meister, B.J., 1968b. Drop formation at low velocities in
liquid–liquid systems: Part II. Prediction of jetting velocity. A.I.Ch.E.
Journal 14, 15–19.
Song, M., Homma, S., Hong, K., 1999. Formation of oil drops discharged
underwater. Proceedings of the Ninth International Offshore and Polar
Acknowledgments Engineering Conference, vol. I. pp. 390–396.
Teng, H., Kinoshita, C.M., Masutani, S.M., 1995. Prediction of droplet size
from breakup of cylindrical liquid jets. International Journal of Multiphase
S. H. was funded by a Japan Society for the Promotion of
Flow 21, 129–136.
Science Postdoctoral Fellowships for Research Abroad. M. S. Tomotika, S., 1935. On the instability of a cylindrical thread of a viscous
was supported by ASERC, KOSEF under R11-2002-104- liquid surrounded by another viscous fluid. Proceedings of the Royal
02005-0. Society of London A 150, 322–337.
3996 S. Homma et al. / Chemical Engineering Science 61 (2006) 3986 – 3996

Tryggvason, G., Bunner, B., Esmaeeli, A., Juric, D., Al-Rawahi, N., Tauber, Webster, D.R., Longmire, E.K., 2001. Jet pinch-off and drop formation in
W., Han, J., Nas, S., Jan, Y., 2001. A front-tracking method for the immiscible liquid–liquid systems. Experiments in Fluids 30, 47–56.
computations of multiphase flow. Journal of Computational Physics 169, Xiaoguang, Z., 1999. Dynamics of drop formation in viscous flows. Chemical
708–759. Engineering Science 54, 1759–1774.
Unverdi, S.O., Tryggvason, G., 1992. A front tracking method for viscous, Zhang, W.W., Lister, J.R., 1999. Similarity solutions for capillary pinch-off
incompressible, multi-fluid flows. Journal of Computational Physics 100, in fluids of differing viscosity. Physical Review Letters 83, 1151–1154.
25–37.

Vous aimerez peut-être aussi