Download as pdf or txt
Download as pdf or txt
You are on page 1of 4

PHYSICAL REVIEW LETTERS week ending

PRL 99, 104502 (2007) 7 SEPTEMBER 2007

Stability of a Jet in Confined Pressure-Driven Biphasic Flows at Low Reynolds Numbers


Pierre Guillot,1,* Annie Colin,1 Andrew S. Utada,2 and Armand Ajdari3,†
1
Rhodia Laboratoire du Futur, Unité mixte Rhodia-CNRS, Université Bordeaux I,
178 Avenue du Docteur Schweitzer, 33608 Pessac, France
2
School of Engineering and Applied Sciences, Harvard University, Cambridge, Massachusetts 02138, USA
3
Gulliver, UMR CNRS-ESPCI 7083, 10 rue Vauquelin, 75231 Paris Cedex 05, France
(Received 1 November 2006; published 7 September 2007)
Motivated by its importance for microfluidic applications, we study the stability of jets formed by
pressure-driven concentric biphasic flows in cylindrical capillaries. The specificity of this variant of the
classical Rayleigh-Plateau instability is the role of the geometry which imposes confinement and
Poiseuille flow profiles. We experimentally evidence a transition between situations where the flow takes
the form of a jet and regimes where drops are produced. We describe this as the transition from convective
to absolute instability, within a simple linear analysis using lubrication theory for flows at low Reynolds
number, and reach remarkable agreement with the data.

DOI: 10.1103/PhysRevLett.99.104502 PACS numbers: 47.15.Fe, 47.55.t, 47.61.Jd, 83.50.Ha

Since the seminal works of Plateau [1] and Rayleigh [2], ensures good alignment and centering [14]. Rc is in the
i.e., for more than a century, the breakup of a liquid jet 200–500 m range, whereas the radius of the tapered
injected in an immiscible liquid has been extensively in- orifice of the square tube is set between 20 and 50 m
vestigated because of its relevance for an immense field of using a pipette-puller set up. Syringe pumps are used to
practical and industrial applications ranging from chemical inject an inner fluid of viscosity i at a rate Qi in the square
processes to ink jet printing, through spray atomization, capillary and the outer fluid of viscosity e at a rate Qe
emulsification process, and polymer extrusion, to name a through the cylindrical capillary. This leads to coaxial
few. A major result is that inviscid cylindrical jets in air are injection at the tapered orifice.
unstable to disturbances of wavelength larger than the jet We observe flow patterns which vary significantly with
circumference. The interplay of interfacial stresses and operational (Qe , Qi ), geometrical (Rc ), and system pa-
inertial forces that rules jet breakup is rather complex rameters (i , e , surface tension ). Figure 2 displays
[3,4], and can be rationalized in terms of absolute and the typical outcome of an experiment where the flow rates
convective instabilities as determined from a linear analy- are varied for a given system (here the inner solution is a
sis [5]. An absolute instability corresponds to disturbances 50% in weight glycerine in water solution with i 
growing and propagating both downstream and upstream, 55 mPa:s and the outer one a silicone oil for which e 
leading after a transient to drops released at or close to the 235 mPa:s). A droplet regime is found for low Qi , with
injection nozzle. Increasing velocity sufficiently can make either droplets emitted periodically right at the nozzle —
the instability convective with perturbations that propagate symbol (open circle) —or non spherical pluglike droplets
downstream while they grow, allowing for a continuous resulting from the instability of an emerging oscillating jet
fluid thread to persist. Several experimental studies support (filled gray circle). Jets are found in the bottom right corner
this picture [6,7]. of Fig. 2 with different visual aspects: wavy jets with
In this Letter, we apply these concepts to analyze jet features that are convected downstream (open square),
stability in a somewhat different realm, namely, biphasic and for larger values of Qi , straight jets (filled square)
flows in microfluidic channels. In this context, dripping that persist throughout the cylindrical capillary. For large
and coflow regimes have been not only observed, but also values of the external flow rate Qe , we observe what we
exhaustively used for various applications [8–10]. Beyond call jetting: thin and rather straight jets (open diamond)
the importance of viscous forces (the Reynolds numbers that extend over some distance in the capillary tube before
are typically small to moderate), the essential contrast with
most of the previous studies which focused on unbounded
flows [6,11–13] is the major role of the microchannel walls
which induce parabolic flow profiles and strongly affect the
development of perturbations.
In our experiments, we generate a jet in a cylindrical
glass capillary of inner radius Rc , using as a nozzle a glass
capillary of square cross-section with a tapered end (see
Fig. 1). The outer dimension of this square capillary is very
close to the inner diameter of the cylindrical tube which FIG. 1. Flow geometry and notations used in the text.

0031-9007=07=99(10)=104502(4) 104502-1 © 2007 The American Physical Society


PHYSICAL REVIEW LETTERS week ending
PRL 99, 104502 (2007) 7 SEPTEMBER 2007

the two fluids. The important difference is that, as the


radius of the jet varies, so does, through the Laplace law,
the difference between the two pressure gradients:
 
r
@z Pi  Pe    @z 02i  @2z ri (1)
ri
with @2z ri and r
r02
i
the jet curvatures along and perpendicu-
i
lar to the flow. Mass conservation of the incompressible
fluids provides the closure of the system of equa-
tions: Qe z; t  Qi z; t  0 and @t r0i  ri 2 
@z Qe z; t.
Considering perturbations proportional to eikz!t , with
k  kr  iki and !  !r  i!i complex numbers, we
FIG. 2. Map of the flow behavior in the (Qi , Qe ) plane. The obtain the dispersion equation:
droplet regime comprises droplets smaller than the capillary
(open circle) and pluglike droplets confined by this capillary i Ka x3 Ex; k~  Fx; k~2  k~4 
(filled circle). Jets are observed in various forms: jets with visible ~
! (2)
peristaltic modulations convected downstream (open square),
x9 1  1   x5
wide straight jets (filled square) that are stable throughout the ~  !16
where we use the notations k~  r0i k, ! e Rc
,

5 cm long channel, and thin jets breaking into droplets at a 


well-defined location (open diamond). Parameters are Rc  Ex;   4x  8  41 x3  41  1x5 ;
275 m, inner viscosity i  55 mPa:s, outer viscosity e 
235 mPa:s, surface tension   24 mN=m. Fx;   x4 4  1  4 lnx  x6 8  41 
 x8 4  31  4  41  lnx;
breaking into droplets at a well-defined and reproducible
location. This ‘‘jet length’’ increases with Qi for a fixed and where we identify the three essential adimensional
value of Qe [15]. Similar ‘‘dynamic phase diagrams’’ have parameters that rule the behavior of the system: the
been reported for more complex microfluidic geometries viscosity ratio   ei , the degree of confinement of
r0
q
[16 –20]. the unperturbed jet x  Ric  11 1
with  
We now attempt to model these phenomena analytically, q 0 2

taking advantage of the simplicity and symmetry of the 1  1 QQei , and a capillary number Ka  @zP Rc . Note
present cylindrical geometry. The problem remains how- that Ka is a capillary number defined at the scale Rc and not
ever quite complex, even at the level of a linear stability at the jet scale r0i as in most studies of unbounded flows.
analysis, so we proceed with approximations [21]. We The jet would be linearly stable if all !r were negative,
neglect inertial effects (i.e., as the Reynolds number is which is not the case. With some !r positive, an initially
small in most of our experiments), and we use lubrication localized perturbation generates a growing distortion that
theory (i.e., we formally assume that the wavelengths of spreads in a domain bounded by two fronts moving at
the perturbations are long compared to the capillary radius) velocities v and v , selected because they correspond
which has been shown to be remarkably insightful in to maximal growth rate !r and extremal velocity of the
somewhat related situations [22]. The unperturbed refer- envelope of the perturbation v  !kir . These criteria read
  
ence state (Fig. 1) is easily described using Stokes equa- v  !kr , @!  @!r
@kr  0 and v  @ki [5,22], which selects, in a
r

tion. The pressure gradients in the two fluids are for this i
dimensional form v ~  !~ r =k~i ,
unidirectional flow constant and equal @z Pe  @z Pi 
@z P0 , with the Laplace law relating the local pressures: Ka x3 Ex;  C1 Fx; 
P0i  P0e  r0 where r0i the radius of the inner jet (Fig. 1). ~  
v (3)
i
x9 1  1   x5
We perform a linear stability analysis of this solution, p q
considering only small z-dependent cylindrically symmet- where C1  518 7 p7241.
ric perturbations ri z; t. Within the lubrication approxi- As v is always positive, the nature of the instability is
mation (i.e., assuming that the interface is flat and that the set by the sign of v : if negative, growing perturbations
flow is uunidirectional), this leads to a description in also travel backwards and the jet is absolutely unstable,
terms of the resulting perturbations Qe z; t, Qi z; t, while for v > 0, the instability is convective as all
@z Pe z; t, @z Pi z; t, which are locally related by rela- perturbations are convected downstream. We reach a
tions similar to those describing the unperturbed flow. In rather simple analytical prediction for the transition:
particular, we still enforce no slip boundary condition at ~ Ka; x;   0, plotted on Fig. 3 in the (x, Ka) plane
v
the solid liquid interface as well as continuity of the describing operational conditions, for various values of the
velocity and of tangential stress at the interface between system dependent viscosity ratio   i =e .
104502-2
PHYSICAL REVIEW LETTERS week ending
PRL 99, 104502 (2007) 7 SEPTEMBER 2007
3
102 What are the main features predicted by this model?
101 convective (jet) Comparing systems, decreasing   i =e increases the
100 ‘‘droplet’’ regime at the expense of the ‘‘jet’’ regime. For a
λ↓
Ka
10−1 given system (a given ), one moves from droplets to jets,
10−2 either by increasing the capillary number Ka (i.e., the
10−3 absolute (droplets) normalized pressure drop) or the degree of confinement
10 x  r0i =Rc . This is physically sound, as increasing Ka
0 0.2 0.4 x 0.6 0.8 1
corresponds to convecting away the perturbations faster,
FIG. 3. Dynamic behavior diagram in the (x, Ka) plane. The while increasing the confinement x results in slowing down
lines are the predictions v~ Ka; x;   0 for the absolute- the rate of development of the perturbations due to the
convective transition, for viscosity ratios  equal to 10, 1, 0.1, proximity of the walls.
0.01, and 0.001 (bottom to top). Above the lines, the jet is In order to provide a directly useful guide, we turn to a
convectively unstable, whereas below the lines, the jet is abso- representation in the operational plane Qi , Qe , in the Fig. 4
lutely unstable. Regions below the lines correspond to droplets using characteristic values of our experiments. Noticeably,
region.
these plots predict a ‘‘reentrant’’ behavior jet ! drops !
jet upon increase of the external flow rate Qe with all other
We suggest that this plot describes the dynamic behavior parameters fixed. This relies on the physics mentioned
of our system, with the transition separating dripping above. At low Qe , increasing Qe mostly reduces the strong
(absolute instability) from jets (convective instability). confinement of the jet, allowing the instability to develop
Indeed, in the convective case, growing disturbances are faster and thus promoting the formation of droplets. At
simultaneously convected downstream and a continuous high Qe , confinement is not significant, and raising Qe
jet can persist in the system over some distance (which primarily increases the jet velocity and the downstream
does not preclude the formation of droplets downstream).
This encompasses the three types of jets observed in Fig. 2.
By contrast, in the absolute instability regime, no jet is 5 a1) a2)
stable, as any perturbation generates oscillations that grow 10
and travel backwards to invade the whole capillary. This
Q (µL/h)

4
corresponds to the droplets and plugs regimes displayed on 10
Fig. 2.
3
10
e

6 a) b)
10
10
5 • 10
2
Qe (µL/h)

4
10 η↓ Γ↑ b1) b2)
3 e 5
10 droplets droplets 10
2
10 jet jet
Q (µL/h)

1 4
10 10
6 c) d)
10

5 3
ηi↓ 10
e

10
Qe (µL/h)

4
10 R↑ 2
3 c droplets
10 droplets 10
2
10 jet jet 2
Qi3(µL/h)
4 5 2
Qi3(µL/h)
4 5
1 10 10 10 10 10 10 10 10
10
1 2 3 4 5 1 2 3 4 5
10 10Qi (µL/h)
10 10 10 10 10Qi (µL/h) 10 10 10 FIG. 5. Experimental data (symbols) and theoretical predic-
tions (continuous line) displaying the effect on the dynamical
FIG. 4. Dynamic behavior in the (Qi , Qe ) plane from our behavior of an increase of the capillary radius Rc [a1 ! a2]
simple model. Above the line, the jet is convectively unstable, and a decrease of the surface tension  [b1 ! b2]. Gray
whereas below the lines it is absolutely unstable and droplets are symbols (open gray circle and filled gray circle) correspond to
expected. We vary independently each parameter around the droplets and the black ones (open square, filled square, and open
conditions of Fig. 2 chosen as reference:   24 mN=m, Rc  diamond) to jets (see Fig. 2). For (a1) and (a2),   24 mN=m,
275 m, e  235 mPa:s and i  55 mPa:s. (a) From top to e  0:235 Pa:s and i  0:055 Pa:s, with Rc  275 m for
bottom, e  23:5, 235 and 2350 mPa:s. (b) From top to bot- (a1) and Rc  435 m for (a2). For (b1) and (b2), Rc 
tom,   50, 24 and 5 mN=m. (c) From top to bottom, Rc  275 m, e  3 mPa:s and i  1 mPa:s, with  
350, 275 and 175 m. (d) From top to bottom, i  5:5, 55 and 12 mN=m in (b1), and a decreased value using surfactants  
550 mPa:s. The gray shading indicates where the average 0:12 mN=m in (b2). The lines are obtained without adjustable
Reynolds number is larger than unity [24]. parameters.
104502-3
PHYSICAL REVIEW LETTERS week ending
PRL 99, 104502 (2007) 7 SEPTEMBER 2007
1
10 ation. We have identified the relevant adimensional quan-
0 tities that control the dynamic behavior and obtained
10
Ka
−1
analytical predictions for the locus of the transition as a
10 function of all the involved physical and operational pa-
−2 rameters. Both agree quantitatively with our experiments,
10
−3 yielding a predictive synthetic map of the dynamic behav-
10 ior and of the resulting flow patterns. Returning to micro-
0 0.2 0.4 x 0.6 0.8 1
fluidic applications, we expect this mapping to be robust
FIG. 6. Flow behavior in the (x, Ka) plane for a given value of and relevant for more commonly used microchannels,
the viscosity ratio   0:23. x is calculated from the first order which have cross-sections of lower symmetry (square,
solution using the experimental values of the flow rates and rectangular, . . .). The corresponding experimental and
viscosities. The data (symbols as in Figs. 2 and 5) correspond to theoretical investigation is under way.
two radii and two surface tensions. The line is the theoretical The authors gratefully acknowledge support from the
prediction of our linear analysis for the droplets/jet transition, Région Aquitaine, and warmly thank D. A. Weitz for his
with no adjustable parameters.
interest and hospitality and M. Joanicot for his constant
support.
convection of any developing perturbation, which favors a
continuous jet. These plots also clarify the influence of the
various parameters. For given flow rates, droplet formation
is promoted by large surface tension, wide capillaries, and *pierre.guillot@eu.rhodia.com
large internal viscosity, while the influence of the external †
Corresponding author: armand.ajdari@espci.fr
viscosity depends on the flow rates considered [see [1] J. Plateau, Statique Experimentale et Theorique des
Fig. 4(a)]. Liquides soumis aux Seules Forces Moleculaires
We now quantitatively compare our predictions to ex- (Gauthier-Villars, Paris, 1873), pp. 450– 495.
perimental data obtained for three surface tensions, two [2] Lord Rayleigh, Proc. London Math. Soc. s1-10, 4 (1878).
viscosity ratios, and two capillary radii Rc (Fig. 5). Clearly, [3] S. P. Lin et al., Annu. Rev. Fluid Mech. 30, 85 (1998).
given the approximations involved, our simple model de- [4] J. Eggers, Rev. Mod. Phys. 69, 865 (1997).
scribes very well the experimental data with no adjustable [5] W. van Saarloos, Phys. Rev. A 37, 211 (1988); 39, 6367
(1989).
parameters, and appears as a powerful predictive tool. A
[6] J. M. Gordillo et al., J. Fluid Mech. 448, 23 (2001).
certain level of disagreement is expected for weak confine- [7] I. Vihinen et al., Phys. Fluids 9, 3117 (1997).
ment (small Qi large Qe ) as the lubrication approximation [8] B. Zheng et al., J. Am. Chem. Soc. 125, 11170 (2003).
is formally invalid in this case. Our model indeed over- [9] M. Joanicot and A. Ajdari, Science 309, 887 (2005).
estimates Ka at the transition for vanishing x, as demon- [10] P. Guillot et al., Langmuir 22, 6438 (2006).
strated by comparison to the exact result of Gañán-Calvo [11] A. Sevilla et al., Phys. Fluids 17, 018105 (2005).
[23] for unbounded creeping flows. Inertial effects may [12] A. M. Gañán-Calvo et al., Phys. Rev. Lett. 96, 124504
also slightly alter the picture for the largest outer flow rates (2006).
(see the shaded areas in Fig. 4). We also report in Fig. 6 all [13] A. M. Gañán-Calvo et al., J. Fluid Mech. 553, 75 (2006).
the experimental data obtained for a given viscosity ratio , [14] A. S. Utada et al., Science 308, 537 (2005).
but for various surface tensions, various capillary radii, and [15] We consider that we have a jet whenever there is a
continuous jet section, wavy or not, the length of which
various flow rates. This shows the relevance of our descrip-
is longer than 3 times the eventual droplet size (if any).
tion in terms of the adimensional variables Ka and x. An [16] T. Thorsen et al., Phys. Rev. Lett. 86, 4163 (2001).
additional interest of this mapping onto the (Ka, x) plane [17] C. Cramer et al., Chem. Eng. Sci. 59, 3045 (2004).
of this large set of data is that it collapses relatively well the [18] J. D. Tice et al., Anal. Chim. Acta 507, 73 (2004).
different types of flows observed within the droplets and [19] S. L. Anna et al., Appl. Phys. Lett. 82, 364 (2003).
jets regimes as can be seen by the grouping of the symbols. [20] P. Guillot and A. Colin, Phys. Rev. E 72, 066301 (2005).
In short, droplets correspond to small capillary number Ka [21] This is equivalent to a computation of the linear response
and small confinement ratio x, while plugs require larger in ri through a second order expansion in powers of kRc
values of x. At higher values of Ka, increasing the con- of the velocity field, where k is the axial wavelength
finement ratio x shifts the behavior from ‘‘jetting’’ (with of the perturbation, with an additional approximation.
emission of droplets at a large finite distance from the Specifically, in the calculation of the OkRc 2  term, we
neglect the pressure gradient created by the zero order in
nozzle) to stable jets.
kRc velocity compared to that generated by the interface
In conclusion, we have studied a variant of the Rayleigh curvature along z. This is valid for the data discussed here
Plateau problem, i.e., the stability of pressure-driven con- (discussion postponed for a forthcoming paper).
centric jets in a cylindrical confining geometry at low [22] T. R. Powers et al., Phys. Rev. Lett. 78, 2555 (1997).
Reynolds numbers. We have demonstrated that the nature [23] A. M. Gañán-Calvo, Phys. Rev. E 75, 027301 (2007).
of the instability at the linear level (absolute or convective) [24] We use the simple estimate Re  Q2e =e 
controls whether drops or a jet are obtained in this situ- Q2i =i =Rc Qe  Qi  and   1000 kg=m3 .

104502-4

You might also like