<<

Drop formation in viscous flows at a vertical capillary tube D. F. Zhang and H. A. Stone Division of Engineering & Applied Sciences, Harvard University, Cambridge, Massachusetts 02138 ͑Received 4 February 1997; accepted 14 April 1997͒ Drop formation at the tip of a vertical, circular capillary tube immersed in a second immiscible fluid is studied numerically for low-Reynolds-number flows using the boundary integral method. The evolution and breakup of the drop fluid is considered to assess the influences of the ratio ␭, the Bond number B, and the capillary number C for 10Ϫ2р␭р10, 10Ϫ2рC р1, and 0.1рBр5. For very small ␭, breakup occurs at shorter times, there is no detectable thread between the detaching drop and the remaining pendant fluid column, and thus no large satellite drops are formed. The distance to detachment increases monotonically with ␭ and changes substantially for ␭Ͼ1, but the volume of the primary drop varies only slightly with ␭. An additional application of the numerical investigation is to consider the effect of imposing a uniform flow in the ambient fluid ͓e.g., Oguz and Prosperetti, J. Fluid Mech. 257, 111 ͑1993͔͒, which is shown to lead to a smaller primary drop volume and a longer detachment length, as has been previously demonstrated primarily for high-Reynolds-number flows. © 1997 American Institute of Physics. ͓S1070-6631͑97͒01808-4͔

I. INTRODUCTION hibit deviations from experimental measurements with errors exceeding 20%.9,10 Drop formation at the tip of a capillary tube occurs in a Overall, two themes are common to the great majority of variety of engineering applications. as well as studies in this field. First, either -in- ͑i.e., are commonly dispersed into a second fluid phase. Fre- formation͒ or liquid-in-gas ͑e.g., drops in air͒ systems have quently cited applications include separation and extraction 1 2 been investigated, where the gas has negligible dynamical processes, spraying and ink-jet printing technologies, blood influence; the complete two-phase flow situation has been oxygenation,3 and the bubble departure process during boil- seldom studied. Second, the majority of these dynamical ing ͑e.g., Ref. 4͒. A summary of many modeling ideas is studies have focused on inertial effects.5 Hence we now sum- provided by Clift et al.5 from which it is clear that the ma- marize some recent studies of this drop formation problem jority of studies have concerned flows at high Reynolds num- with emphasis on investigations of viscous influences. bers. Here we study the low-Reynolds-number situation by 11 numerically investigating the detailed evolution of drop for- Wilson developed a quasi-one-dimensional flow model mation at the end of a capillary, continuing the simulations to determine the drop volume formed by dripping from a past breakup, in order to obtain insight into the formation of fluid-filled tube into a gaseous surrounding. The flow is as- the primary drop and the largest satellite drop. Buoyancy, sumed to be a Stokes flow and the unsteady extension of the interfacial tension effects, viscous effects in both fluid viscous thread as it sags under its own weight is analyzed. phases, and the effect of an external flow are all considered. This useful model, however, is unable to describe flow near When the flow rate is small, a static description of the the nozzle exit, as well as the end of the thread, and an shape of the drop is useful since a pendant droplet slowly infinite detachment length is predicted. The predicted pri- forms at the capillary tip and the drop detaches when a criti- mary drop volume is about 25% lower than Wilson’s own 6–8 cal volume is reached; the critical volume corresponds to experiments. The viscous flow limit was also studied re- a balance between interfacial tension and buoyancy. These cently by Wong et al.12 who investigated the formation of a analyses have been modified to include dynamical effects bubble from a submerged capillary in a viscous environment. and the primary focus of many studies has then been to pre- Numerical solutions in excellent agreement with their experi- dict the drop sizes as a function of the fluid properties, the ments were described. nozzle geometry, and the flow rate inside the nozzle ͑e.g., A one-dimensional model for jet-like flows ͑liquid into Refs. 5 and 9͒. Conceptually, from a simplified modeling gas͒ was presented recently by Eggers and DuPont.13 These point of view, the drop formation process is conveniently authors derived one-dimensional mass conservation and divided into two stages: The first ͑nearly static͒ stage corre- sponds to growth of the drop attached to the capillary, which axial momentum equations accounting for viscous effects, ends with a loss of equilibrium of forces, and the second inertia and capillarity by systematically approximating the stage corresponds to the necking and breaking of the drop. Navier–Stokes equations. These ideas were applied to the The final volume of the primary drop so formed is the sum of highly nonlinear problem of pinching ͑breakup͒ of a fluid 14 the volume of that portion of the static drop that breaks off at thread and the model predictions were shown to be in ex- the end of the first stage and the volume that flows into the cellent agreement with experiments focusing on the dynam- drop during the second, or pinching, stage. It should be noted ics near breakup.15–17 The breakup of the liquid thread shows that the predictions of these simplified models typically ex- a self-similar behavior during the final stages of the pinching

2234 Phys. Fluids 9 (8), August 1997 1070-6631/97/9(8)/2234/9/$10.00 © 1997 American Institute of Physics

Downloaded¬14¬Feb¬2003¬to¬140.247.59.174.¬Redistribution¬subject¬to¬AIP¬license¬or¬copyright,¬see¬http://ojps.aip.org/phf/phfcr.jsp process where the flow near the exit has no influence. In particular, after the formation of a long thread with a nearly conical tip connected to an almost spherical drop, a remark- able series of smaller necks with thinner diameters were se- quentially spawned. Following breakup and formation of a large primary drop, there is recoil of the liquid thread, and then secondary necking and breakup, which leads to satellite drops. The purely flow viscous limit of the pinching process was studied by Papageorgiou18 ͑see also Ref. 19͒. This re- search, and many of its extensions, are described in a recent review article by Eggers.20 Recent investigations have utilized modern numerical methods to investigate the complete free-boundary problem in both the high- and low-Reynolds-number flow limits. For example, Oguz and Prosperetti21 ͑see also Day and Hinch22͒ studied dynamics of bubble growth and detachment from a submerged needle by assuming the flow was inviscid and irrotational. A boundary integral method was used and sev- eral simple, illustrative models of the detachment process were developed. The boundary-integral numerical results were in excellent agreement with published experiments23 as well as their own experiments. Oguz and Prosperetti also showed that bubbles growing when there is a liquid flowing parallel to the needle may detach with a considerably smaller radius than in a quiescent liquid ͑see also Clift et al.5͒. This significant effect motivated our investigation ͑Sec. III͒ of the analogous viscous flow problem. Also, in the spirit of Oguz and Prosperetti’s numerical investigation of the free- boundary problem, Wong et al.12 used a boundary integral method for low-Reynolds-number flows to study bubble de- tachment in a viscous fluid. Our work reported here thus combines elements of the above two investigations. FIG. 1. Schematic of drop formation at a vertical, circular capillary tube. It is important to note that the aforementioned studies pertain to two limits of the drop formation processes: one is smaller than the typical dimension over which the bulk flow a liquid flowing into an ambient gas and the other is a gas varies. For this axisymmetric flow, a cylindrical coordinate ejected into a liquid. Here we explore the details of drop system (r,z) is defined with the z axis coincident with that of formation at a capillary in a general two-fluid system for a the capillary tube, increasing in the direction of g, and the viscously dominated flow. Numerical results are based on the origin is placed at the center of the tube exit see Fig. 1 . boundary integral method for Stokes flows. The formation, ͑ ͒ In the low-Reynolds-number flow limit, the governing extension, and breakup of the drop fluid and, subsequently, equations for motion of the two fluids are (iϭ1,2) the generation of satellite drops, are investigated for a wide 2 range of fluid viscosity ratios. Other effects such as buoy- ٌ•TiϭϪٌpiϩ␮iٌ uiϩ␳igϭ0, ٌ•uiϭ0, ͑1͒ ancy, interfacial tension, and an external flow are also stud- where in fluid i, u is the velocity field, p is the pressure, ied. We describe the numerical formulation in Sec. II and i i and ␮ and ␳ are the fluid viscosity and density, respec- report numerical results in Sec. III. i i tively. In ͑1͒ the stress tensor T is defined to include the hydrostatic body force in order to define a divergence-free II. NUMERICAL FORMULATION field, The formation of a drop at the tip of a vertical, circular T TiϭϪ͑piϪ␳ig–x͒Iϩ␮iٌ͑uiϩٌ͑ui͒ ͒, ͑2͒ capillary tube of radius R0 is shown in Fig. 1. An incom- pressible, Newtonian fluid (iϭ1) flows, owing to a pressure where x is a position vector and g–xϭgz. gradient, with a constant flow rate Q into a second incom- In the present study, the drop fluid (x෈⍀1) is bounded pressible, Newtonian fluid (iϭ2). For simplicity, the tube by the fluid interface SD , the capillary tube inner wall ST1, wall is assumed to have zero thickness, which is physically and the tube inlet SI . The ambient fluid (x෈⍀2) is bounded reasonable since the wall thickness has been shown experi- internally by the fluid interface SD and the capillary tube mentally to have little influence on the drop formation outer wall ST2. In Fig. 1 these surfaces are represented by process.24 The ambient fluid may be quiescent or may be their traces in the (r,z) plane. assumed to be in a constant steady motion Uϱ far from the Inside the capillary tube, far from the tube exit, a Poi- capillary; the latter case is representative of configurations seuille flow is assumed and, thus, the velocity profile at the where the characteristic dimension of the capillary is much inlet SI is

Phys. Fluids, Vol. 9, No. 8, August 1997 D. F. Zhang and H. A. Stone 2235

Downloaded¬14¬Feb¬2003¬to¬140.247.59.174.¬Redistribution¬subject¬to¬AIP¬license¬or¬copyright,¬see¬http://ojps.aip.org/phf/phfcr.jsp 2Q r 2 r 2 ͑1ϩ␭͒ u1ϭ 2 1Ϫ ezϭ2V 1Ϫ ez , ͑3͒ u͑x͒ϭC outezϩ J–nٌ͓͑s–n͒ϪBz͔dSyϩ͑␭Ϫ1͒ ␲R ͫ ͩ R ͪ ͬ ͫ ͩ R ͪ ͬ 2 ͵S 0 0 0 D

ϫ n K u dS ϩ J n T dS where ez is the unit vector along the axis and V is the average ͵ – – y ͵ – – 1 y 2 SD SI fluid velocity in the capillary (VϭQ/␲R0). The boundary conditions along the inner and outer tube ϩ␭ n K u1 dSyϩ J ͑n T2 walls, ST1 and ST2, are no slip: u1ϭ0, u2ϭ0. Along the fluid ͵ – – ͵ – – SI ST interface, SD , the velocity is continuous, u1ϭu2, and the stress jump is balanced by the density contrast and the inter- Ϫn–T1͒dSy for x෈SD , ͑7a͒ facial tension stress, which depends on the local curvature ,s–n of the interfaceٌ 0ϭC outezϩ ͵ J–nٌ͓͑s–n͒ϪBz͔dSyϩ͑␭Ϫ1͒ SD n–T2Ϫn–T1ϭ͓␥ٌ͑s–n͒Ϫ⌬␳g–x͔n, ͑4͒ ϫ͵ n–K–u dSyϩ͵ J–n–T1 dSyϩ␭͵ n–K–u1 dSy SD SI SI where ␥ is the constant interfacial tension, n is the unit nor- mal vector directed into the ambient fluid, ٌsϭ(IϪnn)–ٌ is ϩ͵ J–͑n–T2Ϫn–T1͒dSy, for x෈ST , ͑7b͒ the gradient operator tangent to the interface, and ST ⌬␳ϭ␳1Ϫ␳2. Along SD there is also a kinematic constraint, which can be expressed with the Lagrangian description of a where the kernels functions are labeled point xL as 1 I ͑xϪy͒͑xϪy͒ Jϭ ϩ , 8␲͉ͫxϪy͉ ͉xϪy͉3 ͬ dxL ͑8͒ ϭu͑x ͒ for x ෈S . ͑5͒ dt L L D 3 ͑xϪy͒͑xϪy͒͑xϪy͒ KϭϪ . 4␲ ͉xϪy͉5 The governing equations and boundary conditions are nondimensionalized by choosing the tube radius R0 as the We note that the capillary number appears in the dimension- length scale, and it is convenient to choose the velocity scale less form of the inlet velocity profile ͑integral over SI). For as ␥/␮. Accordingly, the scales for time and pressure, re- this axisymmetric flow, the surface integrals are simplified to spectively, are R0␮/␥ and ␥/R0. line integrals along the generating curve of the boundary by Three dimensionless parameters, a Bond number B,a performing the azimuthal integrations analytically.27 Details capillary number C , and a viscosity ratio ␭, describe the of the numerical implementation are given in the Appendix flow: and here we only note that although the drop was pinned at the edge of the tube, no was specified and so 2 was allowed to take any value consistent with the numerical ⌬␳gR0 ␮1 ␮V Bϭ , ␭ϭ , and C ϭ . ͑6͒ solution ͑e.g., Ref. 12͒. ␥ ␮2 ␥

The Bond number measures the relative importance of the III. RESULTS AND DISCUSSION buoyancy force to the interfacial tension force while the cap- illary number represents the relative importance of the vis- In this section the formation, extension, and breakup of cous force generated by the internal flow relative to the in- the drop fluid and, subsequently, the generation of satellite terfacial tension force. If a constant velocity Uϱ is imposed drops, are investigated. Calculations are performed by vary- in the fluid surrounding the capillary, the outer capillary ing one dimensionless parameter while keeping the other two Ϫ2 number C outϭ ␮Uϱ /␥ enters the problem description. parameters fixed and we have investigated 10 р␭р10, In order to solve this free-boundary problem, Stokes 10Ϫ2рC р1, and 0.1рBр5. We have chosen to consider equations are reformulated into a system of integral equa- situations where the flow rate Q is fixed ͑constant C ͒, which, tions. The numerical procedure is standard and the details owing to the dynamics of interfacial rearrangement, would, can be found elsewhere.25,26 In the present flow, along the in practice, require a time-varying pressure gradient. In com- 5 tube walls the velocities u1 (x෈ST1) and u2 (x෈ST2) are mon circumstances, this pressure change is not significant. identically zero, at the inlet SI the velocity distribution is For low-Reynolds-number flow situations, dimensional specified, and along the deforming fluid–fluid interface SD analysis implies, for example, that in the absence of an im- the stress jump condition (n–T2Ϫn–T1) is known ͓Eq. ͑4͔͒. posed flow in the external fluid the dimensionless volume The unknown quantities are then u on SD , n–T1 on ST1, and Vd of the primary drop formed is Vdϭ f (B,C ,␭). In the Ϫ1 8 n–T2 on ST2; in fact, only the difference (n–T2Ϫn–T1) ap- nearly static limit, C Ӷ1, then VdϰB for all ␭. However, pears at the ͑infinitely thin͒ tube wall, ST ͑ϭ ST1ϭST2). The although there have been many studies on bubble formation final form of the boundary integral equations are under high-Reynolds-number flow conditions, there appears

2236 Phys. Fluids, Vol. 9, No. 8, August 1997 D. F. Zhang and H. A. Stone

Downloaded¬14¬Feb¬2003¬to¬140.247.59.174.¬Redistribution¬subject¬to¬AIP¬license¬or¬copyright,¬see¬http://ojps.aip.org/phf/phfcr.jsp FIG. 3. A time sequence of the interface shapes for ␭ϭ0.1, Bϭ0.5, and FIG. 2. Variation of the minimum dimensionless drop radius, Rmin , and the C ϭ0.1. After separation of the primary and satellite drops from the main corresponding dimensionless length, Lmin , with time; ␭ϭ0.1, Bϭ0.5, and fluid column, the drops are neglected in further calculations of the pendant C ϭ0.1. Some drop shapes are also shown. drop shape. to be little specific information available about drop forma- tion in viscous fluid flows5 so our results for a wide range of critical value, predicted approximately by static stability C , ␭, and C out essentially are graphical forms of the dimen- analyses applicable to low capillary number flows, where the sionless function f , or its analogue for other properties, for drop begins to break away ͑e.g., Middleman;8 see also Fig. such parameters as the dimensionless drop volume, breaking 9͒. time, and fluid column length at breakup. In order to distinguish the flow characteristics at differ- ent times, the drop formation process is generally divided A. Typical case into ͑at least͒ two stages, the latter stage after the drop begins We first consider a typical case ␭ϭ0.1, Bϭ0.5, and to break away and is much shorter than the former. As a C ϭ0.1. Figure 2 shows the variation of Rmin and Lmin with result of the rapid flow in the second stage, the drop stretches time, where Rmin denotes the minimum dimensionless drop and a neck subsequently forms (tϷ40). A thread then devel- radius and Lmin is the axial (z) distance from the capillary ops with a diameter that decreases rapidly with time. The tube exit (zϭ0͒ to Rmin . These two quantities characterize shape of the thread is not symmetric about its horizontal the formation of a neck which leads to breakup at a time centerline. The thread is connected to a nearly spherical pri- tb . Some drop shapes have also been included in Fig. 2. In mary drop at its lower end, where large curvatures develop the present numerical simulation, the drop is assumed to rup- and a local interfacial-tension-driven flow leads to breakup. ture, forming the primary drop, where Rminр0.005, which At tϭ50.7, thread breakoff is about to occur with appears to be a reasonable numerical criterion for breakup LminϷ3.16. We note that the portion of the drop below the when focus is on the primary drop since any additional de- breaking point takes a spherical shape while the upper part crease in the neck radius happens quickly and further com- approximates a cone ͑e.g., Peregrine et al.17͒. These shape putations of the solution become difficult because of the features near pinching are common to most, if not all, ex- large velocity and curvature gradients near the pinch point amples of drop breakup and so from now on we focus on ͑see, for example, Eggers13 and Papageorgiou18͒. In fact, for features of the shape and flow specific to drops formed at a a purely viscous internal ͑Stokes͒ flow, inertial effects be- capillary tube. come important as ͑rapid͒ pinching occurs,18 so locally the In order to demonstrate the entire drop formation pro- low-Reynolds-number approximation becomes invalid. With cess, in Fig. 3 a time sequence is shown ͑using the same an outer fluid present, however, the pinching dynamics are parameters as in Fig. 2͒ including breakup of a drop, forma- slowed and the low-Reynolds-number approximation can re- tion of a satellite drop, and return of the fluid interface to a main valid all the way to breakup.28 blob-like shape similar to the initial shape. Growth, exten- In Fig. 2 we observe that at early times, the drop volume sion, necking, and breakup of the drop can be clearly seen increases by the continuous addition of fluid from the capil- prior to tϭ50.7. Immediately after the thread breaks at its lary tube and the interface shape transforms slowly from the lower end, its free end is retracted by interfacial tension. The hemispherical initial shape to a pear-shaped surface. During thread breaks again at its upper end, resulting in the produc- this period, the buoyancy force acting on the drop, which is tion of a small satellite droplet, as shown in Fig. 3 at proportional to the drop volume, is not large enough to over- tϭ51.2. Satellite drop formation depends primarily on the come the interfacial tension force, and the drop remains at- shape of the thread when it is about to break for the first time tached to the capillary tube. Figure 2 shows that Rminϭ1 and and so depends upon the initial and flow conditions ͑i.e., ␭, Lminϭ0 for tϽ30.2, after which the drop volume reaches a C , and B͒; the volumes of the primary and satellite drops

Phys. Fluids, Vol. 9, No. 8, August 1997 D. F. Zhang and H. A. Stone 2237

Downloaded¬14¬Feb¬2003¬to¬140.247.59.174.¬Redistribution¬subject¬to¬AIP¬license¬or¬copyright,¬see¬http://ojps.aip.org/phf/phfcr.jsp FIG. 4. The breaking length Lb versus ␭ for Bϭ0.5 and C ϭ0.1. Four shapes of the fluid interfaces at breakup are included for ␭ϭ10Ϫ2,10Ϫ1,1, and 10, respectively. The drop shape shown for ␭ϭ10 is displayed with compressed horizontal and vertical scales since very long fluid columns develop for ␭ӷ1. are studied in the remaining sections. After a complete FIG. 5. The fluid interfaces for ␭ϭ10Ϫ2,10Ϫ1, and 1 at the instant of the breakup process, leading to the formation of primary and primary satellite drop detachment with Bϭ0.5, C ϭ0.1. The viscosity ratio satellite drops, the remaining drop fluid which hangs on the ␭ and the satellite drop formation time t*ϭtϪtb are labeled on the figures. tube continues to deform and subsequently, another similar drop formation process occurs. Careful examination shows that the interface shape at tϭ53.7 is almost identical to that role when threads form as breakup occurs. As a result, the at tϭ10.0 ͑see Fig. 2͒. breaking distance Lb increases monotonically with ␭ and rather dramatically for ␭Ͼ1. As indicated in the introduction, long, narrow threads B. Effect of the viscosity ratio ␭ which connect the falling drop and the remaining fluid col- The viscosity ratio ␭ plays an important role during the umn have been experimentally and theoretically observed in dynamical processes of necking and breaking. Figure 4 previous studies of high-viscosity dripping flows,11,13,15,24 shows the breaking length Lb versus ␭ for Bϭ0.5 and with much recent interest given to the dynamics in the neigh- 19,20 C ϭ0.1, where Lb is the dimensionless axial (z) location of borhood of the pinch point. The consistency of the the breaking point. Four interface shapes, just prior to brea- present results with these previous studies is not surprising koff, have also been included. The interface shapes for since in fact these previous studies are an asymptotic limit ␭ϭ10Ϫ2,10Ϫ1, 1 are displayed with the same scale while (␭ ϱ) of the two-fluid flow. We note that numerical accu- that for ␭ϭ10 is presented with a much smaller scale owing racy→ in the calculations presented here is difficult to preserve to the very long fluid column that is formed for this large for the very extended interface shapes characteristic of viscosity ratio. Clearly, as ␭ varies, there are different shapes ␭ӷ1, as large numbers of node points must be used, and this at breakup and although, as we shall see, the primary drop requirement limited our calculations to ␭р10. Also, detailed volume changes only a little with ␭ for the range of C and investigations of the dynamics near pinching shows that, in B studied, the viscosity ratio has a significant effect on the fact, inertia eventually becomes important if ␭ϭϱ but can formation of satellites. For ␭ ϭ10Ϫ2 breakup occurs at an remain insignificant for the finite ␭ case.28 early time, there is no detectable thread between the detach- The viscosity ratio also influences formation of satellite ing drop and the remaining pendant drop, and so no large drops. Figure 5 depicts the interfaces for ␭ϭ10Ϫ2,10Ϫ1,1at satellite drops are expected. This interface shape, calculated the instant of the satellite drop detachment (Bϭ0.5 and for a low value of the capillary number, is similar to that C ϭ0.1). Satellites form for ␭ϭ10Ϫ1 and 1 but, for calculated for bubbles in Stokes flows by Wong et al.12 and ␭ϭ10Ϫ2, there exists no visible thread, thus no satellite drop observed experimentally by Longuet-Higgins et al.23 and ex- is expected. For ␭ϭ10Ϫ1, a thin thread evolves and the perimentally and numerically by Oguz and Prosperetti21 dur- breakup of this thread occurs soon after formation of the ing the high-Reynolds-number bubble formation process primary drop, (t*ϭtϪtbϭ0.55), which generates a very Ϫ3 from a needle (␭ 0); dynamics ͑viscous or inertial͒ only small satellite (Vsϭ5.6ϫ10 ). In contrast, relatively slow play a significant→ role at late times near pinching. As ␭ is satellite drop formation occurs for ␭ϭ1(t*ϭ9.75) and the Ϫ2 increased, a fluid thread develops and its length increases satellite drop (Vsϭ8ϫ10 ) is larger than that for owing primarily to the difficulty of fluid squeezing out axi- ␭ϭ10Ϫ1, a result which is due to the existence of the longer ally along the thread. Thus viscous effects play a significant thread in the case of the more viscous drop fluid.

2238 Phys. Fluids, Vol. 9, No. 8, August 1997 D. F. Zhang and H. A. Stone

Downloaded¬14¬Feb¬2003¬to¬140.247.59.174.¬Redistribution¬subject¬to¬AIP¬license¬or¬copyright,¬see¬http://ojps.aip.org/phf/phfcr.jsp FIG. 6. The dimensionless breaking time tb and the dimensionless volume Ϫ2 of the primary drop Vb as a function of ␭ for Bϭ0.5 and C ϭ5ϫ10 , 10Ϫ1, and 5ϫ10Ϫ1, respectively.

Figures 6 summarizes the dimensionless breaking times (tb) and the dimensionless volumes of the primary drop (Vb). The Bond number is maintained constant (Bϭ0.5) and three capillary numbers (C ϭ0.05,0.1, and 0.5) are con- FIG. 8. Variation of tb as a function of the capillary number C ; Bϭ0.5 and sidered, corresponding to an increasing pressure difference ␭ϭ0.1. driving larger flow rates through the tube. We note that Vb varies only a little as ␭ varies, especially for the small cap- 2 1/3 Rdϭ(3␥/16R0␳g) , where the released drop is assumed to illary number cases, e.g., C ϭ0.05, in spite of the different Ϫ1/3 be a sphere; nondimensionalization gives Rd /R0ϰB or dynamics indicated in Fig. 4, and so static predictions, de- Ϫ1 pendent on the Bond number, for the primary drop volume VbϰB . The small deviation of the exponent n from unity will be useful. Viscous stresses are more important as the is attributed to the small but finite capillary number as well capillary number increases, and correspond to shorter break- as the viscous dynamics at later times. ing times and larger primary drop volumes. Also, the viscos- Referring again to Fig. 7, we note that for larger Bond ity ratio influences the flow during the latter stages, which numbers, e.g., Bϭ5, substantial translation of the drop oc- curs on a short time scale which effectively leads to forma- terminate in breakup with different breaking lengths Lb ͑Fig. 4͒ and breaking times t ͑Fig. 6͒. tion of narrow, tapered threads connecting the drop and fluid b column. Further calculations show that the dynamics of the subsequent satellite formation for varying are also differ- C. Effect of the Bond number B B ent because the thread shape is substantially changed. For The effect of the Bond number is demonstrated in Fig. 7 small B ͑ϭ 0.1͒, a small satellite relative to the primary drop for 0.1рBр5 with ␭ϭ0.1 and C ϭ0.1. Three interface develops after only a short time (t*ϭ0.1) while, in contrast, shapes at breakup are shown for Bϭ0.1, 1, and 5, respec- for large B ͑ϭ 5͒ a relatively large satellite forms from the tively. The primary drop volume Vb decreases nearly linearly narrow, tapered thread, but the satellite formation time is Ϫn with B, and our numerical results indicate that VbϰB longer (t*ϭ2.04). with nϷ0.90, which is similar to theoretical results based 8 upon a static analysis for which the critical drop radius is D. Effect of the capillary number C Figure 8 shows the ͑large͒ variation of the breaking time tb as a function of the capillary number C with Bϭ0.5 and ␭ϭ0.1. We observe that tb decreases significantly with in- creasing C for C Ͻ0.2. However, for C Ͼ0.2, decreases in the breaking time become gradual and tb is nearly constant for C Ͼ0.75. For C Ͻ0.2 the drop grows slowly until it reaches the critical volume and the subsequent necking and breakup processes are relatively fast. In these cases, tb is essentially determined by the time necessary for the drop fluid volume to reach the critical volume for detachment from the capillary tube. For large C , accumulation of the drop fluid is fast and the critical volume is reached at earlier times. The accumulation time no longer determines the breaking time and, instead, the time for the subsequent neck- ing and breakup process is most significant. We note that the necking and breaking process for large C is complicated because a large amount of the drop fluid exits the tube during the second stage and the fluid interface keeps extending and FIG. 7. Variation of the primary drop volume Vb with B; ␭ϭ0.1, C ϭ0.1. The interface shapes at breakup for Bϭ0.1, 1, and 5 are also deforming. This breakup process is somewhat similar to that shown. of a ‘‘jetting’’ flow, as shown in Fig. 8 for C ϭ1. For such

Phys. Fluids, Vol. 9, No. 8, August 1997 D. F. Zhang and H. A. Stone 2239

Downloaded¬14¬Feb¬2003¬to¬140.247.59.174.¬Redistribution¬subject¬to¬AIP¬license¬or¬copyright,¬see¬http://ojps.aip.org/phf/phfcr.jsp FIG. 9. Variation of Lb and Vb with C for ␭ϭ0.1 and three different Bond numbers, Bϭ0.1, 0.5, and 1.0.

large C flow conditions, the breakup time tb primarily de- pends on the necking and breaking process in which the flow in the vicinity of the tube exit has a little influence: an ap- proximately constant tb is thus expected. Figure 9 reports Vb and Lb as functions of the capillary number for 10Ϫ2рC р1, ␭ϭ0.1, and Bϭ0.1, 0.5, 1.0. Both Lb and Vb increase with C . Since static analysis predicts Ϫ1 VbϰB , this scaling is used in Fig. 9, and is seen to be useful for collapsing some of the data. In the cases studied, for fixed ␭, the breaking length is only weakly dependent on B for a given C . As a final remark, when C ӷ1 and the shape remains nearly spherical, the dimensionless drop volume scales as 3/4 12 Vdϰ(C /B) as described by Wong et al. ͑see also Ref. FIG. 10. Interface shapes at breakup for different C and C out ; Bϭ0.5 and 5͒, who show that this result is in good agreement with their Ϫ2 Ϫ1 3 ␭ϭ0.1. The capillary numbers C , from top to bottom, are 10 ,10 , and numerical calculations for ␭ϭ0 provided / Ͼ10 .Onthe Ϫ2 C B 1, and the outer capillary numbers C out , from left to right, are 0, 10 , other hand, for ␭ӷ1, long fluid columns form and Wilson’s 10Ϫ1, and 1. 11 1/2 one-dimensional model predicts Vdϭͱ3␲(C /B) ϩ2.4/B. To study these asymptotic limits in detail would Ϫ2 Ϫ1 require more node points than used in our numerical simula- left to right, are C outϭ0,10 ,10 , and 1. In each frame the tions reported in the rest of the paper. dimensionless breakup time tb is indicated. It is evident that imposing a constant flow on the ambient fluid can effectively influence the drop formation process, as indicated by Clift E. Effect of an external uniform flow et al.5 and quantified here for the viscous flow limit. As com- The above results summarize the response for a quies- pared to the earlier cases of no externally imposed flow, with cent ambient fluid, Uϱϭ0. In this last section, drop forma- a constant Uϱ , or finite C out ,asmaller primary drop, a tion is studied for a uniform steady flow Uϱ . The motivation longer breaking length Lb , and a shorter breaking time re- for this flow configuration is to explore the possibility of sult, whereas, as discussed earlier, a larger C leads to a controlling the drop size and the drop detachment rate. This larger primary drop and a longer breaking length. For a given idea was investigated numerically by Oguz and Prosperetti21 C , the drag force on the drop fluid, which increases with in a study of bubble growth and detachment from a needle increasing C out , tends to stretch the fluid column, and the for the case of irrotational flow conditions and earlier re- earlier breakup time therefore leads to the formation of a 5 search was summarized by Clift et al. It has been observed smaller primary drop. Figure 11 shows the variation of Vb as that the external flow typically leads to the formation of a function of C out (Bϭ0.5, ␭ϭ0.1, and C ϭ0.1). smaller drops, which provides a useful control parameter In Fig. 10, it is observed that large values of C out not since in the absence of flow the drop radius is proportional to only alter the primary drop size and the breaking length, but the one-third power of the capillary tube radius. We now also change the shapes of the interfaces dramatically ͑e.g., Ϫ2 Ϫ1 consider the low-Reynolds-number flow limit and summa- C outϭ1 with C ϭ10 and 10 ). Thus we expect that the rize the change in drop volume, breakoff length, and outer flow also influences the formation of satellite drops. breakoff time as the external flow velocity (C out) is in- The influence of an applied flow on drop formation with creased. different Bond numbers B is considered in Fig. 12, where Figure 10 shows the interface shapes at breakup for dif- the dimensionless drop volumes Vb at C outϭ0 and 0.2 are ferent C and C out with other parameters fixed (Bϭ0.5 and compared for ␭ϭ0.1 and 0.1рBр5. Significant differences ␭ϭ0.1). All of the results are shown with the same scale so in Vb are observed for small B (Ͻ0.1), but the differences that the drop size and the breakup length can be compared are much smaller for BϾO(1). There are now two forces directly. The capillary numbers from top to bottom are trying to move the drop fluid away from the tube, one is the C ϭ10Ϫ2,10Ϫ1, and 1, and the outer capillary numbers, from buoyancy force ͑B͒ and the other is the viscous drag force

2240 Phys. Fluids, Vol. 9, No. 8, August 1997 D. F. Zhang and H. A. Stone

Downloaded¬14¬Feb¬2003¬to¬140.247.59.174.¬Redistribution¬subject¬to¬AIP¬license¬or¬copyright,¬see¬http://ojps.aip.org/phf/phfcr.jsp FIG. 11. Variation of Vb with C out for Bϭ0.5, ␭ϭ0.1, and C ϭ0.1.

exerted by the ambient fluid (C out). The buoyancy force dominates for large Bond numbers, such as Bϭ5, which results in a rapid formation of the small primary drop (t ϭ5.32). In contrast, for a small Bϭ0.1, the drag force b FIG. 13. Interface shapes for C ϱϭ0,0.2 with Bϭ0, ␭ϭ0.1, and C ϭ0.1. exerted by the imposing flow is most significant. Figure 13 provides a further view of interface shapes IV. CONCLUDING REMARKS with Bϭ0, which is a limit of obvious relevance to micro- We have studied numerically the dynamics of drop for- gravity applications. Two outer capillary numbers, C outϭ0 mation from a capillary tube for two-phase low-Reynolds- and 0.2, are examined for ␭ϭ0.1 and C ϭ0.1. For C outϭ0, as expected, the volume of the drop fluid increases continu- number flows. The emphasis has been on determining the ously, the drop fluid remains attached to the tube, and the volume of the primary drop as a function of the Bond num- interface maintains a spherical shape due to interfacial ten- ber, capillary number, and the viscosity ratio. Known sion. With an external flow, however, viscous stresses are asymptotic limits have been summarized in the text and the exerted on the interface by the ambient fluid, and leads to figures reported here thus represent graphically, in the spirit of the compendium of Clift et al.5 ͑Chap. 12͒, the complete drop breakup at tbϭ440.7. It is interesting to note that a comparison of different simulations indicates that the inter- dependence of the drop volume, breakup length, and breakup time as a function of the dimensionless parameters ͑which face shape for Bϭ0, C outϭ0.2 ͑Fig. 13͒ is similar to that typically are outside the region where the asymptotic results near breakup for Bϭ0.1, C outϭ0 ͑Fig. 12͒. are valid͒. The effect of an external flow, known to generally lead to smaller drop sizes, is here quantified for viscously dominated flows.

ACKNOWLEDGMENTS We gratefully acknowledge financial support from the Petroleum Research Fund ͑Grant No. 28690-AC9͒. We also thank C. Maldarelli for supplying preprints of his research group’s related studies. We thank M. P. Brenner for helpful conversations related to the generic problem of drop forma- tion at capillaries.

APPENDIX In this appendix we provide the details of the numerical implementation for the solution of the integral equations from which the boundary velocity is determined. Given an initial shape for the interface, Eqs. ͑7a͒ and ͑7b͒ can be solved numerically by approximating the integral equations by a linear system of equations. Each boundary, S or S , is defined by a set of discrete boundary nodes at FIG. 12. Comparison of Vb for C outϭ0 and C outϭ0.2 with ␭ϭ0.1, D T 0.1рBр5. which the velocity is calculated. A hemispherical cap is as-

Phys. Fluids, Vol. 9, No. 8, August 1997 D. F. Zhang and H. A. Stone 2241

Downloaded¬14¬Feb¬2003¬to¬140.247.59.174.¬Redistribution¬subject¬to¬AIP¬license¬or¬copyright,¬see¬http://ojps.aip.org/phf/phfcr.jsp sumed to be the initial shape of the drop at the end of the 3K. M. Sutherland, D. T. Derek, and L. S. Gordon, ‘‘Independent control of capillary and is represented by 15–20 points. With increases blood gas PO2 and PCO2 in a bubble oxygenator,’’ Clin. Phys. Physiol. of the interface length, more points are added so as to pre- Meas. 9,97͑1988͒. 4 serve the numerical representation of the interface and the W. M. Sluyter, P. C. Slooten, C. A. Coprajj, and A. K. Chesters, ‘‘The departure size of pool-boiling bubbles from artificial cavities at moderate node points are uniformly redistributed along the interface and high pressure,’’ Int. J. 17, 153 ͑1991͒. every 5–10 time steps. For the calculations reported here the 5R. Clift, J. R. Grace, and M. E. Weber, Bubbles, Drops, and maximum number of node points is 100, which has been ͑Academic, New York, 1978͒. shown to be sufficient by performing several numerical 6R. Kumar and N. R. Kuloor, ‘‘The formation of bubbles and drops,’’ Adv. simulations using more points. To resolve carefully smaller Chem. Eng. 8, 256 ͑1970͒. 7D. H. Michael, ‘‘Meniscus stability,’’ Annu. Rev. Fluid Mech. 13, 189 satellite drops than those studied here, more node points are ͑1981͒. necessary. Cubic splines are used for a continuous interpola- 8S. Middleman, Modeling Axisymmetric Flows: Dynamics of Films, Jets, tion of the interface shape where the spline parameter is the and Drops ͑Academic, New York, 1995͒. arclength s along the interface measured from the corner of 9G. F. Scheele and B. J. Meister, ‘‘Drop formation at low velocities in the tube exit at zϭ0 ͓i.e., r(s), z(s)]. Along the capillary liquid-liquid systems: Part I. Prediction of drop volume,’’ AIChE. J. 14,9 tube wall S (rϭ1), the boundary is discretized using ͑1968͒. T 10P. M. Heertjes, L. H. de Nie, and H. J. de Vries, ‘‘Drop formation in r(z), a fixed number of points (20) are distributed along the liquid-liquid systems I. Prediction of drop volumes at moderate speed of capillary for a distance of 10 tube diameters and, beyond this formation,’’ Chem. Eng. Sci. 26, 441 ͑1971͒. point, the integral is truncated. For the unknown quantities 11S. D. Wilson, ‘‘The slow dripping of a viscous fluid,’’ J. Fluid Mech. 190, velocities and stress jumps , a linear piecewise interpolation 561 ͑1988͒. ͑ ͒ 12 in terms of the spline parameter between adjacent nodes is H. Wong, D. Rumschitzki, and C. Maldarelli, ‘‘Theory and experiment on the low expansion, contraction and detachment of a used. bubble pinned at the tip of a submerged capillary’’ ͑unpublished͒. The numerical integrations are performed with Gaussian 13J. Eggers and T. F. DuPont, ‘‘Drop formation in a one-dimensional ap- quadratures using the IMSL Math/library routines. The abso- proximation of the Navier-Stokes equation,’’ J. Fluid Mech. 262, 205 lute and relative error limits for the numerical integrations ͑1994͒. 14 are chosen as 10Ϫ5 and 10Ϫ7, respectively. Special care is J. Eggers, ‘‘Theory of drop formation,’’ Phys. Fluids 7, 941 ͑1995͒. 15X. D. Shi, M. P. Brenner, and S. R. Nagel, ‘‘A cascade of structure in a required to handle the logarithmic singularity as y x.We → drop falling from a faucet,’’ Science 265, 219 ͑1994͒. subtract the logarithmic behavior of the singularity from the 16M.P. Brenner, J. Eggers, K. Joseph, S. R. Nagel, and X. D. Shi, ‘‘Break- integrands and so reduce the integrands into a regular part down of scaling in droplet fission at high Reynolds numbers,’’ Phys. Flu- and a singular part, whose contributions are computed sepa- ids 9, 1573 ͑1997͒. 17 rately ͑using different codes in the IMSL library͒. For most D. H. Peregrine, G. Shoker, and A. Symon, ‘‘The bifurcation of liquid bridges,’’ J. Fluid Mech. 212,25͑1990͒. of the cases, the singular part ͉xϪy͉ is set to be 1% of the 18D. Papageorgiou, ‘‘Drop formation in a one-dimensional approximation of distance of the adjacent nodes. The resulting matrix equation the Navier-Stokes equation,’’ J. Fluid Mech. 262, 205 ͑1994͒. is solved by iterative refinement. Once the linear system of 19M. P. Brenner, J. R. Lister, and H. A. Stone, ‘‘Pinching threads, singu- equations is solved to obtain the interfacial velocities, the larities and the number 0.0304 . . . ,’’ Phys. Fluids 8, 2827 ͑1996͒. 20 interface location is updated by solving the kinematic condi- J. Eggers, ‘‘Nonlinear dynamics and breakup of flows,’’ Rev. Mod. Phys. 69, 865 ͑1997͒. tion ͓Eq. ͑5͔͒ using an explicit Euler method. 21H. N. Oguz and A. Prosperetti, ‘‘Dynamics of bubble growth and detach- Numerical accuracy is assessed by increasing the bound- ment from a needle,’’ J. Fluid Mech. 257, 111 ͑1993͒. ary discretization and by monitoring the change of the drop 22R. Day and J. Hinch, ‘‘Boundary integral simulation of a finite inviscid fluid volume. Changing the boundary discretization is per- jet,’’ in Proceedings of CFD95, Banff, Canada, edited by P. Thibault and formed by increasing the node number, N, while keeping D. Bergeron 2, 375 ͑1995͒. 23M. S. Longuet-Higgins, B. R. Kerman, and K. Lunde, ‘‘The release of air ⌬t/⌬s fixed, where ⌬t is the time step and ⌬s, proportional bubbles from an underwater nozzle,’’ J. Fluid Mech. 230, 365 ͑1991͒. to 1/N, is representative of the node spacing. The volume of 24X. Zhang and O. A. Basaran, ‘‘An experimental study of dynamics of drop the drop at the end of the computations is calculated numeri- formation’’ Phys. Fluids 7, 1184 ͑1995͒. cally and compared with the sum of the initial volume and 25C. Pozrikidis, Boundary Integral and Singularity Methods for Linearized the fluid volume that flowed from the capillary tube during Viscous Flow ͑Cambridge University Press, Cambridge, England, 1992͒. 26J. Tanzosh, M. Manga, and H. A. Stone, ‘‘Boundary integral methods for the computing time. The volume differences are always viscous free-boundary problems: Deformation of single and multiple fluid- within 2% of the initial volume. fluid interfaces,’’ Proceedings of Boundary Element Technologies VII, ed- ited by C. A. Brebbia and M. S. Ingber ͑Computational Mechanics, 1W. J. Heideger and M. W. Wright, ‘‘Liquid extraction during drop forma- Southampton, 1992͒,p.19. 27 tion: Effect of formation time,’’ AIChE. J. 32, 1372 ͑1986͒. S. H. Lee and L. G. Leal, ‘‘The motion of a sphere in the presence of a 2T. W. Shield, D. B. Bogy, and F. E. Talke, ‘‘Drop formation by DOD deformable interface,’’ J. Colloid. Interface Sci. 87,81͑1982͒. ink-jet nozzles: A comparison of experiment and numerical simulation,’’ 28J. R. Lister, M. P. Brenner, and H. A. Stone, Presentation at the 1996 IBM J. Res. Dev. 31,96͑1987͒. APS-DFD Meeting, Syracuse, NY ͑unpublished͒.

2242 Phys. Fluids, Vol. 9, No. 8, August 1997 D. F. Zhang and H. A. Stone

Downloaded¬14¬Feb¬2003¬to¬140.247.59.174.¬Redistribution¬subject¬to¬AIP¬license¬or¬copyright,¬see¬http://ojps.aip.org/phf/phfcr.jsp