Top Banner

of 16

1-s2.0-S0009250999000275-main

Jun 04, 2018

Download

Documents

Afshin Bi
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
  • 8/13/2019 1-s2.0-S0009250999000275-main

    1/16

    Chemical Engineering Science 54 (1999) 17591774

    Dynamics of drop formation in viscous flowsXiaoguang Zhang

    Air Products and Chemicals, Inc., 7201 Hamilton Boulevard, Allentown, PA 18195-1501, USA

    Received 22 April 1998; accepted 21 December 1998

    Abstract

    This paper presents numerical results of the dynamics of a viscous liquid drop that is being formed directly at the tip of a vertical,

    circular tube and breaks into an ambient, viscous fluid. A model is developed to predict the evolution of the drop shape and its

    breakup based on the volume-of-fluid/continuum-surface-force method, which is a solution algorithm for computing transient,

    two-dimensional, incompressible fluid flow with surface tension on free surfaces of general topology (Richards et al. (1995 Physics of

    Fluids, 257, 111145)). The full NavierStokes system is solved by using finite-difference formulation on a Eulerian mesh. The mesh is

    fixed in space, with the flow and interface moving through it to ensure accurate calculations of complex free surface flows and

    topology, including surface breakup and coalescence. The nonlinear dynamics of drop growth and breakup are simulated for

    describing and predicting the universal features of drop formation. The focus here is on dynamic effects of a quiescent or flowing

    ambient fluid on drop breakup and the subsequent generation of satellite droplets. The effects of finite inertial, capillary, viscous, and

    gravitational forces are accounted for in order to classify drastically different formation dynamics and to elucidate the fate of satellite

    droplets. The numerical predictions are compared with experimental measurements for a typical system of 2-ethyl-1-hexanol drops

    forming and breaking into quiescent water, and the results show excellent agreement. 1999 Elsevier Science Ltd. All rights

    reserved.

    Keywords: Drops; Satellites; Evolution; Deformation; Breakup; Dynamics; Free-surface flows; Volume-of-fluid; Continuum-surface

    force

    1. Introduction

    A common way of dispersing a liquid in an immiscible

    fluid is to flow it continuously through a nozzle or an

    orifice plate from which it emerges into the ambient fluid

    and breaks into drops. At low flow rates, the liquid being

    ejected emanates from the nozzle as discrete drops under

    its own weight. At high flow rates, the liquid is ejected

    from the nozzle as a jet that subsequently breaks up into

    small drops because of well-known Rayleigh instability

    (Clift et al., 1978). The formation of liquid drops from

    nozzles has long been a topic of interest because of its

    occurrence in a wide variety of engineering applications,

    such as distillation and extraction processes and spraying

    and emulsifying technologies, among others. The pre-

    vious applications have demonstrated the importance of

    a fundamental understanding of the dynamics of drop

    formation in designing and controlling these processes to

    obtain certain desired drop characteristics. The dynamicsof a viscous liquid drop forming from a capillary tube

    and breaking into a quiescent, ambient fluid that is

    inviscid and dynamically inactive has been extensively

    investigated in an earlier study by Zhang (1999). The

    evolution with time of surface profile and internal flow

    of the drop is simulated by using a numerical method

    based on algorithms of volume of fluid (VOF) (Hirt and

    Nichols, 1981) and continuum surface force (CSF)

    (Brackbill et al., 1992). In the study by Zhang (1999), the

    effects of finite inertial, capillary, viscous, and gravi-

    tational forces are accounted for in order to classify

    drastically different formation dynamics and, in particu-

    lar, to elucidate the feature of satellite generation. The

    present paper follows this earlier work very closely and

    extends the results to the situation where a viscous liquid

    drips from a circular capillary tube at low flow rates and

    breaks up as drops into another immiscible, viscous

    liquid. The focus here is on the dynamic effects of the

    ambient fluid on drop formation. Despite the fact that

    the overall process of drop formation may not be quali-

    tatively altered, the dynamic effects of an ambient fluidbring about an additional hydrodynamic force on the

    forming drop, which may change the surface profile of

    0009-2509/99/$see front matter 1999 Elsevier Science Ltd. All rights reserved.

    P I I : S 0 0 0 9 - 2 5 0 9 ( 9 9 ) 0 0 0 2 7 - 5

  • 8/13/2019 1-s2.0-S0009250999000275-main

    2/16

    the drop and features of drop breakup and satellite

    creation. Moreover, an external flow of the ambient fluid

    has been found to lead to a smaller-volume breakoff drop

    and a longer drop length prior to its breakup (Oguz and

    Prosperetti, 1993; Zhang and Stone, 1997).

    During liquid dripping from a capillary tube, the vol-

    ume of a drop that emerges from the tube increases by the

    continuous addition of the drop liquid. When the volumeof the forming drop exceeds a critical value, the drop

    necks and a large portion of it starts to fall rapidly and

    eventually breaks off from the tube. The most interesting,

    yet complicated, phenomena in the drop formation are

    the evolution of a liquid thread that connects the detach-

    ing drop and the remainder of the liquid on the tube

    during the drop breakup and generation of satellite drop-

    lets subsequent to the thread breakup. A thorough know-

    ledge of the drop evolution and breakup is indispensable

    for complete predictions and designs of practical pro-

    cesses that involve interfacial contact and flows of two

    liquids.

    Drop formation from a nozzle or an orifice has been

    studied extensively from theoretical and experimental

    perspectives. Much of the previous theoretical, as well as

    experimental, work has been aimed at predicting the size

    of drops breaking from a nozzle as a function of fluid

    properties, nozzle geometry, and flow rate of the liquid

    through the nozzle (Clift et al., 1978; Kumar and Kuloor,

    1970). The theoretical analyses for this purpose have been

    based primarily on macroscopic force balances and have

    assumed that drop formation occurs in two stages. The

    first stage takes place with a pure static growth of thedrop, which ends with a loss of equilibrium of forces. The

    second stage corresponds to the necking and breaking of

    the drop from the nozzle (Scheele and Meister, 1968;

    Heertjes et al., 1971). These studies have concluded that

    the volume of the drops that are so formed depends on

    not only the nozzle size and liquid properties, as is the

    case with static pendant drop (Michael, 1981), but also

    the liquid flow rate through the nozzle. Although the

    previous studies differ slightly in their approaches to

    analyzing the second stage of drop formation, these sim-

    plified models can, at best, approximate reality when the

    liquid flow rate is vanishingly small and the predictions

    of the size of breakoff drops exhibit deviations from

    experimental measurements, with errors around 20%

    (Clift et al., 1978).

    Although the previously cited studies of drop forma-

    tion in the dripping mode have captured some of the

    gross features of the phenomenon, they have done little

    to elucidate the fundamental fluid mechanics of the pro-

    cess, viz., the evolution of the shape of the forming drop

    in particular, the development, extension, and breakup of

    the liquid thread and satellite generation. Complete

    simulations of drop formation involve the solution of theNavierStokes system with specified boundary condi-

    tions. Cram (1984) and Eggers and Dupont (1994) have

    derived and solved one-dimensional equations of mass

    and axial momentum conservations to simulate drop

    dripping. These approximate equations have been

    derived from the NavierStokes system by either

    (1) neglecting the radial component of the velocity

    and variation of the axial component of the velocity

    and pressure in the radial direction (Cram, 1984) or (2)

    extending the velocity and pressure variables in a Taylorseries in the radial direction and retaining only

    the lowest-order terms in these expansions (Eggers

    and Dupont, 1994). Notwithstanding the failure to

    profile actual velocity fields observed inside growing

    drops and to describe the dynamics of drop breakup

    (Schulkes, 1993), these models are surprisingly successful

    in predicting the evolution with time of the drop shapes,

    as made evident by the qualitative comparison of cal-

    culated shapes with experimental observations in a few

    specific situations when the liquid flow rate is vanishingly

    small.

    Qualitative features of the dynamics of the thread

    breakup have been explored experimentally by Peregrine

    et al. (1990) and, in more detail, by Shi et al. (1994). These

    studies show the details of the evolution with time of

    a liquid thread at the time preceding, at the instant of,

    and at the time following drop detachment. Peregrine et

    al. (1990) have documented photographic sequences of

    events occurring during breakup of the thread in their

    paper. Their results exhibit the process of double break-

    age of the liquid thread. Under the weight of a detaching

    drop underneath, the thread necks and breaks at its

    lower end, where the thread joins with the falling drop, toform a free primary drop. Because of unbalanced capil-

    lary forces on the thread after its first breakup, the thread

    recoils; secondary breakup then occurs at its upper end,

    leading to the generation of satellite droplets. These ex-

    perimental observations have been well modeled by

    Schulkes (1994), who numerically integrated a potential

    flow formulation of Eulers equations for the formation

    of inviscid drops. The primary contribution of Shi et al.

    (1994) is the demonstration of the breakup of drops with

    a large range of viscosities by experiment and computa-

    tion, using the one-dimensional model developed by Egg-

    ers and Dupont (1994). Shi and his coworkers show that

    liquid threads can spawn a series of smaller necks with

    ever-thinner diameters prior to breakup. In spite of their

    demonstration of these phenomena during drop detach-

    ment, the authors do not provide any quantitative evalu-

    ation of the process or systematic description of the roles

    of operating conditions on the drop formation and

    thread breakup. Recently, Zhang and Stone (1997) have

    developed a numerical model to simulate the formation

    of viscous drops from a capillary tube using a boundary

    integral method. These researchers consider the evolu-

    tion and breakup of a drop to assess quantitativelythe effects of viscous, buoyant, and capillary forces.

    In particular, they reveal the important role of the

    1760 X. Zhang/Chemical Engineering Science 54 (1999) 17591774

  • 8/13/2019 1-s2.0-S0009250999000275-main

    3/16

    viscous effects and imposed flows of an ambient fluid on

    the dynamics of drop deformation and breakup. In the

    low-Reynolds-number flow limit, this model cannot

    identify the importance of the inertial force, which has

    been found to be effective as the flow rate is increased

    (Zhang and Basaran, 1995; Zhang, 1999).

    It is noteworthy that Richards et al. (1995) have de-

    veloped a dynamic simulation method based on theVOF/CSF numerical technique to investigate the full

    transient of liquid drop and jet formation from startup to

    breakup. In their study, a transition from dripping to

    jetting of an emerging liquid is obviously identified when

    the liquid flow rate exceeds a critical value. Results of

    their numerical simulations show significantly more ac-

    curacy than previously simplified analyses in predicting

    the jet dynamics, including the jet evolution, velocity

    distribution, and volume of breakoff drops. The novel

    feature of this method is the use of a Eulerian volume-

    tracking approach designed to simulate flows with free

    interfaces of arbitrarily complex topology (e.g., merging

    and breakup), which are of particular interest to us in our

    research. However, Richards and coworkers have fo-

    cused, in their paper (Richards et al., 1995), on predicting

    the volume of breakoff drops under the conditions cor-

    responding to liquid flow rates near and above the

    formation of a jet with the Reynolds number exceeding

    400. A detailed understanding of liquid dripping into

    another immiscible, viscous fluid at the level of Zhang

    (1999) is still lacking.

    The major goal of this study, along with the sub-

    sequent study of numerical solutions, is to remedy theseand related inadequacies so as to complete our under-

    standing of drop formation. The transient NavierStokes

    equation has been solved for the axisymmetric free-

    boundary problem of a Newtonian liquid that is dripping

    vertically and breaking as drops into another immiscible

    Newtonian liquid. The VOF/CSF-based numerical ap-

    proximations used by Richards et al. (1995) have been

    extended to simulate the complete process of drop forma-

    tion from the time a drop emerges from the tube to its

    breakoff from the tube with continuous feeding of the

    drop liquid at a certain flow rate. In contrast to most of

    the previous studies, the only assumptions involved in

    the present numerical model are (1) Newtonian fluids of

    drop and ambient phases with constant physical proper-

    ties and (2) laminar flows. The special feature of

    VOF/CSF, which allows free surfaces to cross the com-

    putational mesh smoothly, ensures that the calculations

    pass the point of necking followed by natural breakup of

    drops without interruption, which is a major incentive

    for using the VOF/CSF method in this study. We pay

    particular attention to the dynamic effects of an ambient

    fluid on breakup of a liquid thread and generation of

    satellite droplets. The numerical results are comparedwith the available experimental data, which are obtained

    using an ultra-high-speed motion analysis and video

    system. Section 2 presents the numerical model, which

    includes the problem definition and formulation as well

    as a brief discussion of the VOF/CSF algorithm. A more-

    detailed description of the VOF/CSF and its validation

    on several test problems can be found elsewhere

    (Richards, 1994; Hirt and Nichols, 1981; Brackbill et al.,

    1992). The approach used to solve the governing equa-

    tions and associated boundary conditions is alsodescribed in Section 2. Section 3 briefly describes the

    experimental apparatus and methods of data acquisition

    and analysis. Typical experiments with 2-ethyl-1-hexanol

    (2EH) drops forming in distilled water are performed,

    and the resulting data have been compared with the

    numerical simulation for different conditions. The com-

    putational results and analyses of the findings are

    the subject of Section 4. Most of the results are presented

    in dimensionless forms and show the importance of

    dynamic effects of an ambient fluid and inertial,

    viscous, capillary, and gravitational forces on drop

    formation. For illustration and verification purposes,

    comparisons of numerical simulations and experimental

    measurements for typical processes are also presented

    in Section 4. The concluding remarks are provided in

    Section 5.

    2. Mathematic formulation and numerical method

    The system of interest is an axisymmetric drop of an

    incompressible Newtonian liquid, density and viscosity

    , forming into an immiscible, incompressible New-tonian fluid, density#and viscosity, at the tip ofa circular cylindrical capillary tube. The drop liquid is

    injected at a constant flow rate, Q, as shown in Fig. 1. The

    tube has inner and outer radii, R and R, respectively,

    and its axis lies along the direction of the gravity vector,

    g; therefore, an axisymmetric free-boundary problem is

    imposed. The ambient fluid is contained in a cylindrical

    tank that has an inner radius, R, and is coaxial with the

    capillary tube. The surface tension,, of the liquidliquidinterface is spatially uniform and constant in time. It is

    convenient to define a cylindrical coordinate systemr,z, , whose origin lies at the center of the outlet planeof the tube and where rdenotes the radial coordinate; zis

    the axial coordinate measured in the opposite direction

    of gravity, g; and is the azimuthal angle. For the

    axisymmetric configuration of interest in this study, the

    problem is independent of the azimuthal coordinate.

    Isothermal, transient flows of liquids of the drop and

    ambient phases are governed by the NavierStokes

    system

    ) v"0, (1)

    v

    t# ) (vv)"!

    1

    p#

    1

    ) #g#

    1

    F

    , (2)

    X. Zhang/Chemical Engineering Science 54 (1999) 17591774 1761

  • 8/13/2019 1-s2.0-S0009250999000275-main

    4/16

    Fig. 1. Schematic of a drop forming from a capillary tube and breaking

    into an ambient fluid.

    where v is the velocity vector, t is time, p is the scalarpressure, andFdenotes body forces that may be present

    in the system. The viscous stress tensor, , is defined as

    follows for the Newtonian liquid:

    "[(v)#(v)]. (3)

    The transient NavierStokes system (Eqs. (1) and (2)) is

    solved by using a finite-difference formulation on a

    Eulerian mesh, which is fixed in space with the flow and

    interface moving through it. The basic algorithm is to

    break up a time descretization of the momentum equa-

    tion into two steps. In the first step, a velocity field is

    computed from incremental changes resulting from vis-

    cosity, advection, gravitational acceleration, and body

    forces. In the second step, the velocity field is projected

    onto a zero-divergence vector field, resulting in a single

    Poisson equation for the pressure field. The details of the

    overall solution scheme can be found in a study by Kothe

    and Mjolsness (1992).

    A nonconventional approach, referred to as the CSF

    method (Brackbill et al. 1992), is used to represent the

    effect of the surface tension at free surfaces. It interprets

    the surface tension as a continuous, three-dimensional

    effect across free surfaces and incorporates it as a localiz-ed volume force in the NavierStokes equation rather

    than as a boundary-value condition. The volume force,

    which is nonzero only within free surfaces, is given in the

    CSF model by

    F"F

    "F, (4)

    where is the surface tension, is the local free surfacecurvature, andF denotes a VOF volume function that is

    used to track the profile of free surfaces.Free surfaces are represented using the VOF technique

    pioneered by Hirt and Nichols (1981). The VOF method

    provides a means of following fluid regions through a

    Eulerian mesh of stationary cells and enables a finite-

    difference representation of free surfaces that are arbit-

    rarily oriented with respect to the computational mesh.

    The scalar function, F, is defined as the fractional volume

    of the drop fluid in the respective cells of the computa-

    tional mesh and is given by

    F(x,z,t)"

    1, in the drop,

    '0,(1, at the free surface,

    0, in the ambient fluid.

    (5)

    The free surface position in the respective cells is divided

    from its neighboring cells and is governed by

    F

    t#(v ))F"0. (6)

    This equation states that the volume function, F, moves

    with the fluid and provides the information necessary to

    reconstruct the free surfaces.The NavierStokes system (Eqs. (1) and (2)) is solved

    for drop formation beginning at the instant at which the

    free surface of the drop is flat, situated at the tip of the

    tube with the entire system being at rest att)0, subject

    to the following boundary conditions. The three-phase

    contact line, where the drop liquid, the ambient fluid, and

    the solid surface meet, remains pinned to the sharp inner

    edge of the tube surface for all times t*0:

    r

    R"1, at

    z

    R"0. (7)

    In contrast to liquidgas systems, when a liquid drop

    forms from a tube into another immiscible liquid (the

    focus of the present paper), whether the contact line pins

    on the inner edge or the outer edge depends on the

    relative wettabilities of the two liquids on the flat surface

    of the tube (Berg, 1993). For simplicity, the contact line is

    assumed to pin to the inner edge of the tube and a tube

    having a wall sufficiently thin is used in this study. Never-

    theless, the simulations can also be applied for systems in

    which the contact line pins to the outer edge of the tube

    as long as the tube wall is sufficiently thin that its effectscan be neglected (Zhang, 1999). The fixed contact line

    eliminates the troublesome determination of the contact

    1762 X. Zhang/Chemical Engineering Science 54 (1999) 17591774

  • 8/13/2019 1-s2.0-S0009250999000275-main

    5/16

    angle in this problem. Far upstream of the tube outlet,

    the flow inside the tube is fully developed,

    v"0, v

    "2

    Q

    R 1!r

    R

    , (8)as

    z

    RP!R, for 0)

    r

    R)1,

    where v

    and v

    are the radial and axial components of

    the velocity of the drop liquid, respectively. In computa-

    tions reported in this paper, the tube length has been set

    to be twice of the inner radius of the tube. Greater lengths

    have been tested, and the calculated results have shown

    no noticeable difference. The ambient fluid may be qui-

    escent or may be set in motion by continuously supplying

    the ambient fluid at a constant flow rate,Q. In the latter

    case, a fully developed flow in the annulus far upstream of

    the outer container is imposed (Bird et al., 1960),

    v"0,

    v"2

    Q

    (R!R

    )

    1!r

    R#

    1!(R

    /R)

    ln(R/R

    )

    ln(r/R), (9)

    as z

    RP!R, for

    R

    R)

    r

    R)

    R

    R,

    wherev

    and v

    are the radial and axial components of

    the velocity of the ambient fluid, respectively. Along

    surfaces where the liquid and solid come in contact, theliquid obeys the conditions of no slip and no penetration:

    v"0, at z

    R"0, for 1)

    r

    R)

    R

    R, (10)

    at r

    R"1 and,

    r

    R"

    R

    R, for!R(

    z

    R)0,

    at r

    R"

    R

    R, for !R(

    z

    R(R.

    A continuative outflow boundary condition is used at

    the top of the computational domain. The length in#zdirection is set to be sufficiently longer than the estimated

    drop breakoff length (i.e., the length of a drop at instant

    of its breakup). Greater lengths have been tested, and the

    calculated results have shown no noticeable difference in

    features of drop breakup and satellite generation.

    Nondimensionalizing the governing NavierStokes

    system and boundary conditions, using R as the length

    scale and the average velocity of liquid inside the tube

    u"Q/Ras the velocity scale, yields three dimension-

    less parameters that describe the fluid mechanics of drop

    formation:

    Re"u

    R

    , Ca"

    u

    and G"

    Rg

    .

    The Reynolds number, Re, measures the importance of

    inertial forces relative to viscous forces; the capillary

    number, Ca, measures the importance of viscous forces

    relative to surface tension forces; and the gravitational

    Bond number, G, measures the importance of gravi-

    tational forces relative to surface tension forces. The

    viscosity ratio of the drop liquid and ambient fluid, , is

    introduced to account for the dynamic effects of theambient fluid. The wall effects of the ambient fluid con-

    tainer are described by the ratio of the inner radii of the

    container and the capillary tube,R/R. If an axial flow is

    imposed in the ambient fluid surrounding the capillary

    tube with a average velocity u"Q

    /(R

    !R

    ), a di

    mensionless characteristic velocity,u/u

    , enters the prob-

    lem description.

    The numerical solutions for the axisymmetric free-

    boundary problem of drop formation are closely fol-

    lowed the previous studies (Richards, 1994; Zhang, 1999).

    For a comprehensive description and analysis of the

    numerical method, the reader is referred to Richards

    (1994). The VOF/CSF algorithms are used to solve the

    governing NavierStokes system and the boundary con-

    ditions on a Eulerian rectangular, staggered mesh in

    cylindrical geometry. In order to achieve large cell-wise

    resolution of the free surface change and high computa-

    tional efficiency, a finer local mesh is used around the

    drop region, whereas a coarser mesh is used in the ambi-

    ent region. The complete process of drop formation is

    simulated from the time a drop emerges from the tube

    until its detachment from the tube, with continuous feed-

    ing of the drop liquid at a specified flow rate.

    3. Experimental approach

    The experiments have been designed to obtain quanti-

    tative information on the dynamics of drop formation of

    a typical liquidliquid system of 2-ethyl-1-hexanol (2EH)

    drops forming and breaking into distilled water in order

    to compare the measurements with and verify the predic-

    tions of the numerical model presented earlier. In the

    experiments, attention is particularly paid to the evolu-

    tion with time of the shape of a growing drop as it necks

    and then breaks up and also to the creation of satellite

    droplets subsequent to drop breakup.

    The apparatus used to form drops has been described

    elsewhere (Zhang and Basaran, 1995). It consists essen-

    tially of a fine-capillary tube through which the liquid

    used to form drops is delivered at a constant volumetric

    flow rate by means of a liquid syringe pump (ATI Orion

    M361). A liquid drop is formed at or emerges directly

    from the tip of the tube. The capillary tubes used in the

    experiments are chosen to have sufficiently large ratios of

    inner and outer radii, with R/R'0.8 that the walleffects of the tubes can be neglected (cf. Zhang, 1999). The

    capillary tubes are submerged in a water container

    X. Zhang/Chemical Engineering Science 54 (1999) 17591774 1763

  • 8/13/2019 1-s2.0-S0009250999000275-main

    6/16

    Fig. 2. Time sequence of shapes of a 2EH drop forming in quiescent water from a tube ofR"0.16 cm at the liquid flow rateQ"5 ml/min, wheret is

    the time of the drop formation measured from the instant that the drop emerges from the tube.

    having a sufficiently large inner radius as compared with

    the tube radius with their ratio R/R"62.5 so that the

    wall effects of the container can be neglected.

    An ultra-high-speed video camera by Kodak (Ektapro

    Electronic Memory Motion Analyzer Model EM1012),

    which is used for continuously capturing images of the

    drop formation process, and the associated hardware for

    recording, storing, and analyzing drop-shape data areessential to the experimental study. The camera system is

    composed of an intensified imager, which can record

    1000 full images or 12,000 partial images per second and

    allows rapid and accurate determination of the loci of

    instantaneous interface profiles from which the various

    measures of drops are evaluated. The entire apparatus is

    placed on a vibration isolation table from Newport.

    The drop liquid is 2-ethyl-1-hexanol with the viscosity

    and density of 0.089 g cm sand 0.83 g cm, respec-

    tively; and the ambient liquid is distilled water which

    has a viscosity and density of 0.01 g cm s and

    1.0 g cm, respectively. The interfacial tension is meas-

    ured to be 13.2 g s(Harris and Byers, 1989). All experi-

    ments are performed at room temperature of 22$0.5C.

    In a typical run, a steady flow with a desired rate is

    established through the capillary and the ambient water

    is quiescent. The water and 2EH are mutually saturated

    prior to each experiment. The system is allowed to run

    for about 5 mins before measurements are taken. A peri-

    odic flow situation is then reached in which the drops

    form, grow, and detach from the outlet of the capillary.

    Since, in this dripping regime, primary and satellite drops

    of uniform size are continuously created, this techniqueprovides a reliable and repeatable illustration of the

    dynamics of drop formation. Reproducibility of results

    for the drop shape and volume has been found to be

    within 5% by making measurements under the same

    conditions, but at different times.

    4. Results and discussion

    This section presents the results of an investigation of

    all major effects governing the dynamics of drop forma-

    tion, which are represented by Re, Ca, G, , R/R, and

    u/u

    . Calculations are performed by systematically vary-

    ing one parameter while keeping the others fixed. The

    quantitative results to be reported have thus been made

    over wide ranges of the governing parameters to provide

    insight into the dynamics and to classify drastically

    different formation processes. In the nature of its im-

    portance and complexity of drop breakup and satellite

    generation, studies directed towards parameter correla-tions and/or maps to characterize satellite generation are

    underway in our laboratory. Whereas it is straightfor-

    ward to vary any one of the dimensionless governing

    parameters while holding all others fixed in numerical

    modeling, it is not possible to do so in laboratory experi-

    ments. Therefore, a typical case of 2EH drop formation

    in quiescent water is considered for comparative and

    verification purposes. The experimental investigation for

    drop formation has been performed by varying the vol-

    umetric flow rate, Q, to cover certain ranges of the

    Reynolds number and the capillary number: that is, the

    Reynolds number varies as 10(Re(70, the capillary

    number varies as 710(Ca(410, and the

    gravitational Bond number and viscosity ratio are fixed

    atG"0.32 and"8.9, respectively. Besides the surfaceprofile of forming drops, two measurements are used to

    characterize the features of drop evolution and breakup:

    the maximum length which a drop can reach prior to its

    breakup, , measured along the tube axis from the tip of

    the tube to the apex of the drop (see Fig. 1); and the

    volume of a primary (or breakoff ) drop, . As shown in

    Section 4.1, the excellent results obtained by comparisons

    of numerical simulations and experimental measure-ments give confidence as to the accuracy of the numerical

    model and its predictions.

    4.1. Typical case

    Fig. 2 shows a time sequence of interface profiles ob-

    tained by the numerical model (presented by the curved

    images on the right side of each interface profile) and

    experiments (the black images on the left side) for a 2EH

    drop forming in quiescent, ambient water from a tube of

    radius R"0.16 cm, at a constant liquid flow rate of

    Q"5 ml/min. As the volume of the drop increases

    1764 X. Zhang/Chemical Engineering Science 54 (1999) 17591774

  • 8/13/2019 1-s2.0-S0009250999000275-main

    7/16

    successively with the continuous addition of the drop

    liquid, the drop rises on the tip of the tube. The geometry

    of the drop gradually transforms from nearly spherical to

    pear shaped during its growth period. As the drop vol-

    ume becomes sufficiently large that the buoyancy force

    on the drop exceeds that of the surface tension (at

    t'1.21 s), the drop necks rapidly and a large portion

    starts to detach irreversibly from the tube. During thenecking sequence, the drop elongates rapidly and con-

    tracts in the middle to form a liquid thread. The liquid

    thread, which connects the detaching drop and the re-

    mainder of the liquid on the tube, is stretched along with

    rapid rise of the detaching drop and eventually breaks at

    its upper end at t"1.2712 s, resulting in a free primary

    drop. Immediately after the thread breaks at the upper

    end, the newly freed end of the thread recoils due to the

    unbalanced force of the surface tension. Meanwhile, the

    cone-shaped liquid remaining on the tube tends to re-

    trieve its apex and become spherical because of the capil-

    lary force. A large curvature, therefore, develops at the

    joining point of the liquid cone and the thread, leading to

    a large capillary pressure at and high velocities of the

    liquid out of that point (Zhang, 1999). Then, in a similar

    fashion to the initial breakup, the thread breaks again,

    this time at its lower end at t"1.275 s. As a result of the

    thread rupturing at both ends sequentially, a satellite

    droplet is created, lying between the primary drop and

    the cone-shaped liquid remaining on the tube. In fact, the

    satellite droplet is found to be very small, typically hav-

    ing a volume less than 1% that of the primary drop.

    Comparison of the numerical prediction with the experi-mental measurements shows excellent agreement. The

    numerical model accurately predicts the evolution of the

    interface profile of the drop, particularly, the double

    breakage of a liquid thread and generation of a satellite

    droplet subsequent to the thread breakup.

    The dynamic response of drop formation is especially

    notable during the period of drop breakup. Fig. 3 shows

    a summary of breakup features in terms of the maximum

    lengths that a drop can attain at the instant when the

    primary drop is about to detach (or the limiting lengths),

    , and the volume of the primary drop,, as functions of

    the flow rate,Q. The drop lengths and volumes presented

    in Fig. 3 are obtained for 2EH drops forming in water

    from a tube of the inner radius, R"0.16 cm. Three

    interface profiles of drops at the instant they are about to

    break are included in Fig. 3 for flow rates ofQ"5, 10,

    and 15 ml/min. The corresponding experimental results

    of /R, /R, and photographic profiles of drops are

    also shown in Fig. 3 by points and black images on the

    left, respectively. Again, excellent agreement between the

    numerical simulation and experimental measurements

    has been demonstrated in Fig. 3. As the flow rate in-

    creases, the pressure at the tube exit and the mass (orvolume) of liquid accumulating in the cone directly above

    the tube rise rapidly, due in part to the large axial

    Fig. 3. Dimensionless limiting length and breakoff volume of 2EH

    drops forming into quiescent water as a function of the flow rate from

    a tube ofR"0.16 cm.

    momentum of the entering liquid and also to the resist-

    ance that the thread exerts on the large amount of liquid

    exiting the tube during the necking and drop detachment

    steps. As a consequence of the pressure and mass buil-

    dup, a larger elongation of the liquid cone occurs as

    Q increases. As shown in the inserted profiles in Fig. 3,the increased deformation of the liquid cone retards, or

    even abolishes, development of the liquid thread at the

    larger Q, a determining factor that changes the features of

    thread breakup and satellite generation. The behavior of

    thread development and breakup is so interesting and

    important in drop formation, particularly in the creation

    of satellite droplets, these features will be discussed fur-

    ther in a separated part. Fig. 3 also shows that the

    breakoff volume of a drop increases with the flow rate,Q.

    The rate of increase in the breakoff volume of a drop

    decreases and tends to level off as Q becomes sufficiently

    large. Indeed, as the flow rate continues to increase, over

    a critical value, transition to jetting occurs and then the

    volume of a primary drop decreases (Richards et al.,

    1995). Moreover, in spite of the fact that the elongated

    liquid cone and enlarged detaching drop may contribute

    to the increase in the limiting length of a drop, the results

    in Fig. 3 indicate only a slight increase in the limiting

    length as Q increases. The primary reason for the weak

    dependence of the limiting length on the flow rate is the

    oblate deformation of the detaching drop caused by the

    viscous effects of the ambient fluid. Figure 3 depicts

    clearly that the additional deformation or flattening ofthe top of a detaching drop, brought about by the ambi-

    ent fluid, increases as Q increases because of the increased

    X. Zhang/Chemical Engineering Science 54 (1999) 17591774 1765

  • 8/13/2019 1-s2.0-S0009250999000275-main

    8/16

    Fig. 4. Variation of interaction time with the flow rate for 2EH drops

    forming in quiescent water from a tube ofR"0.16 cm. The times for

    the three images in inserts ofQ"5 ml/min are t"1.271, 1.2748, and

    1.277 s, respectively. The times for the three images in inserts ofQ"10

    ml/min are t"0.643, 0.646, and 0.65 s, respectively.

    Fig. 5. Dimensionless limiting length and breakoff volume of drops as

    a function of the Reynolds number at Ca"0.01, G"1, "1, andu/u

    "0.

    forming rate of the drop; therefore, it tends to counteract

    the elongation of the liquid cone.

    As a result of the double breakage of a liquid thread,

    a satellite droplet is predicted by the numerical model

    and is observed in experiments to be created under cer-

    tain operating conditions (see Fig. 2). Detailed examina-

    tion of the thread breakage and satellite creation reveals

    the importance of deformation of the liquid cone that issessile on the tube and joins the lower end of the thread.

    A quantitative measurement is defined to describe the

    breaking feature of the liquid thread as the time interval

    that elapses between the breakup of the two ends of the

    thread, commonly referred to as interaction time. Fig.

    4 shows the variation in the time interval, t, with theflow rate obtained by the numerical model (curve) and

    experiments (points) for 2EH drops forming into water

    from a tube of R"0.16 cm. Two series of interface

    profiles of breaking drops, one obtained from the model

    (the right side) and the other from experiments (the left

    side), are inserted in Fig. 4 for flow rates ofQ"5 and 10

    ml/min. These inserted profiles demonstrate typical situ-

    ations of drop formation with and without satellite

    generation. Evidently, as the flow rate increases, the in-

    creased axial momentum of the entering liquid causes the

    liquid cone to sustain a large axial elongation. It is more

    difficult for the surface tension force to restore the cone-

    shaped liquid mass to near-spherical form when compet-

    ing with increasing inertial force and then to develop

    a localized, large curvature at the point where the cone

    joins with the thread. Therefore, more time is required for

    the secondary rupture of the thread at the joining point.As shown in Fig. 4, the interaction time, t, increaseswith elevated flow rate. When the flow rate is sufficiently

    high, the secondary breakup of the thread can no longer

    occur before the thread recoils and coalesces with the

    liquid cone. In that case, no satellite droplet is created

    and the interaction time becomes infinity, as indicated in

    Fig. 4. Apparently, whether or not double breakage of

    a thread occurs and a satellite droplet is created are

    determined substantially by the relative importance of

    the inertial and capillary forces, which can be measured

    by a Weber number,e"ReCa. For the present study

    of the 2EH drops forming into water at different flow

    rates, the results from numerical model and experi-ments show that the threshold of the flow rate is

    8.5$0.5 ml/min, which corresponds to a critical Weber

    number of 0.031.

    4.2. Effect of the Reynolds number

    Fig. 5 demonstrates computational predictions of the

    variation of the dimensionless limiting length, /R, and

    the volume of a breakoff drop, /R, as a function of

    the Reynolds number, while holding other governing

    parameters fixed as Ca"0.01, G"1, and "1. Fornumerical results shown in Sections 4.24.5, a quiescent

    ambient fluid is considered with a outer container of the

    fixed size ofR/R"3. An obviously smaller value ofR

    /R

    has been used in numerical calculations in comparison

    with that in experiments in order to attain a high com-

    putational efficiency. In fact, as shown in Section 4.6, the

    wall effects of an ambient fluid container can be neglected

    as R/R*2.5. Moreover, the boundary condition of

    rigid free slip is used in calculations shown in Sections

    4.24.5 for the container wall to further minimize the wall

    effects. The effects of axial flow of the ambient fluid are

    shown in Section 4.7. Four profiles of interface shapes of

    1766 X. Zhang/Chemical Engineering Science 54 (1999) 17591774

  • 8/13/2019 1-s2.0-S0009250999000275-main

    9/16

    Fig. 6. Evolution with time of interface profiles of detaching drops at

    Reynolds numbers of (a) Re"1 and (b) Re"10, with Ca"0.01,

    G"1,"1, and u/u

    "0.

    drops at the instant they are about to detach are inserted

    in Fig. 5 for Re"0.1, 1, 10, and 100. As expected, the

    limiting length, , of a detaching drop increases mono-

    tonically as the Reynolds number increases, due prim-

    arily to the increase in the inertial force to elongate the

    liquid cone remaining on the tube (see Fig. 5). Moreover,

    the larger axial momentum of the liquid leaving the tube

    forces more liquid to squeeze into the drop and causes

    the volume of the detaching primary drop to increase as

    Reincreases. At a certain criticalRe, transition to jetting

    occurs, at which time the volume of the primary drop

    attains its maximum, as shown in Fig. 5.

    As the Reynolds number increases, the dynamic effects

    of the ambient fluid becomes pronounced on a detaching

    drop because of the increased speed of the drop rising in

    the medium. The increased tangential viscous stress on

    a detaching drop from the ambient fluid causes the drop

    to change its shape from prolate (at Re)1) to oblate(at Re&10) and, then, to dimpled ellipsoidal

    (at Re*100), as shown in inserted drop profiles in Fig. 5.

    Fig. 6 demonstrates breakup features of drops at

    Reynolds numbers ofRe"1 and 10, while holding other

    governing parameters fixed as Ca"0.01, G"1, "1,and u

    /u

    "0. At low Reynolds numbers (Fig. 6a), the

    liquid thread ruptures at both ends sequentially because

    of the stretching under the weight of the detaching pri-

    mary drop and development of large local mean curva-

    tures at the joining points of the thread with the twobodies of liquid. Peaks of capillary pressure are then

    induced inside the thread at its two ends and force the

    liquid out of these breaking points, accelerating thread

    rupture (see also Zhang, 1999). Consequent to the double

    breakage of the thread, a satellite drop is generated. In

    contrast, at the large Reynolds number (Fig. 6b), the

    liquid thread is no longer obvious as a result of the

    increasing elongation of the liquid cone on the tube

    during the drop necking sequence. Following breakoff of

    the primary drop, the tip of the liquid cone rebounds and

    forms a bulbous head under unbalanced capillary forces.

    Eventually, as shown in Fig. 6b, the bulbous head merges

    into the liquid cone and a new drop emerges. No satellite

    droplet is generated at this large Reynolds number.

    4.3. Effect of the capillary number

    The effects of the capillary number on drop breakup, in

    terms of the dimensionless limiting length, /R, and

    detaching volume of the drop, /R, are shown in Fig.

    7 with other parameters fixed at Re"1, G"1, "1,and u

    /u

    "0. Four drop profiles at the instant of detach-

    ment are inserted for Ca"0.001, 0.01, 0.05, and 0.1.Similar to the cases of drop forming into air (Zhang,

    1999), the limiting length, , and the volume of the

    primary drop,, are shown to increase monotonically as

    Ca increases.

    Noteworthy, as the capillary number becomes suffi-

    ciently large (such asCa"0.1 in Fig. 7), the capillary (or

    surface tension) force cannot restore the detaching pri-

    mary drop to spherical shape and, therefore, a large

    curvature can no longer be developed at the joining point

    of the detaching drop and the liquid thread. Moreover,

    the increased viscous force relative to the capillary force

    holds a relatively long liquid thread prior to its rupture.

    Therefore, in contrast to drop formation at small capil-

    lary numbers, interface disturbances may cause the

    thread to rupture at its middle portion prior to the

    capillary-driven breakup of the thread at its upper end

    joining the detaching primary drop, as shown in Fig. 8

    for a series of interface profiles of a breaking drop under

    conditions ofCa"0.1 and all other conditions the same

    as those in Fig. 7. As a result of the sequential rupture of

    the thread at its top portion, a small satellite droplet is

    created. The rest of the thread then recoils rapidly and

    merges into the liquid cone. A similar breakage feature ofa drop of a highly viscous liquid has been observed in

    experiments with glycerol drops (Zhang and Basaran,

    X. Zhang/Chemical Engineering Science 54 (1999) 17591774 1767

  • 8/13/2019 1-s2.0-S0009250999000275-main

    10/16

    Fig. 7. Dimensionless limiting length and breakoff volume of drops as

    a function of the capillary number at Re"1, G"1, "1, andu/u

    "0.

    Fig. 8. Evolution with time of interface profiles of detaching drops at

    Re"1, Ca"0.1,G"1,"1, and u/u

    "0.

    1995). At intermediate range of the capillary number, the

    effects of the viscous force is no longer sufficiently great

    to cause a long liquid thread, and the surface tension

    force is still small relative to the inertial force, leading to

    drop formation at a small Weber number. Similar to the

    situation of large Reynolds numbers, the development of

    a thread is suppressed by the large elongation of the

    liquid cone on the tube, and obviously, the double break-

    age of the thread and satellite generation will no longeroccur (cf. Fig. 6b). At a very small capillary number, the

    capillary force becomes dominant and causes the double

    Fig. 9. Dimensionless limiting length and breakoff volume of drops as

    a function of the gravitational Bond number at Re"1, Ca"0.01,

    "1, and u/u

    "0.

    breakage of a thread at its two ends and subsequent

    satellite generation (cf. Fig 6a).

    4.4. Effect of the gravitational Bond number

    The gravitational Bond number, G, has been varied in

    calculations by changing the magnitude of the density

    difference between the two phases,. Fig. 9 shows thevariation of the dimensionless volume of a primary drop,

    /R, and the limiting length of the drop at the instant of

    its detachment, /R, as functions of the gravitational

    Bond number with the other parameters fixed at Re"1,

    Ca"0.01,"1, and u/u

    "0. In Fig. 9, instantaneous

    interface profiles of detaching drops are inserted for

    G"0.1, 1, 3, and 10. Fig. 9 shows monotonical decrease

    in the volume of the detaching drop and the limiting

    length with increasing values of G. In particular, the

    volume of the primary drop is shown to be a strong

    function of the magnitude of the gravitational force, orG,

    as G(2. The rate of decrease in the breakoff volume

    with increasing G falls and tends to level off asG becomes

    sufficiently large and the volume of the primary drop is

    sufficiently small, indicating the difficulty of reducing the

    volume of breakoff drops. Because of the large driving

    force presented to pull the liquid out of the tube at a large

    value of G, the drop is stretched immediately after its

    emergence from the tube and breaks rapidly. The large

    pulling effect of gravity significantly reduces the volumeof liquid remaining on the tube during drop breakoff, and

    the development of a liquid thread is no longer as evident

    1768 X. Zhang/Chemical Engineering Science 54 (1999) 17591774

  • 8/13/2019 1-s2.0-S0009250999000275-main

    11/16

    Fig. 10. Dimensionless limiting length and breakoff volume of drops asa function of the viscosity ratio at Re"1, Ca"0.01, G"1, and

    u/u

    "0.

    as with smallGvalues. Obviously, at such large values of

    G, the double breakage of the thread and satellite genera-

    tion cannot be expected, as predicted by the numerical

    model. As a result of increasing velocity of drop de-

    tachment at large G, the increasing viscous force of the

    ambient fluid causes the detaching drop to deform to

    a dimpled ellipsoidal shape; a similar phenomenon

    occurred as Re is increased.

    4.5. Effect of the viscosity ratio of two phases

    Despite the fact that liquid viscosities of drop and

    continuous phases have little influence on the volume of

    a breakoff drop (Kumar and Kuloor, 1970), the viscous

    force has been found to affect drop necking and breakup

    greatly, particularly the development, extension, and

    breakup of a liquid thread and generation of satellite

    droplets, as briefly discussed in Section 4.3. To address

    the viscous effects on drop formation, the viscosity ratio

    of the drop liquid to the ambient fluid, , is used and hasbeen changed by varying the viscosity of the drop liquid

    alone in present numerical simulations. Fig. 10 shows the

    variation of dimensionless breakoff volume, /R, and

    limiting length, /R, as a function of with other gov-

    erning parameters fixed atRe"1,Ca"0.01,G"1, and

    u/u

    "0. Instantaneous interface shapes when drops are

    about to break are also displayed in Fig. 10 for"0.01,1, 5, and 10. As the viscosity ratio (or the viscosity of the

    drop liquid) increases, the dimensionless limiting length,

    /R, increases substantially due to the significant in-

    crease in lengths of the liquid thread and cone, as shown

    in the inserted drop profiles in Fig. 10. Viscosity of the

    drop liquid plays a great role in stabilizing the drop

    interface, which makes possible greater thread elongation

    and extension by damping, and even eliminating, oscilla-

    tions of the interface. As a result of the increased exten-

    sion of the thread, a longer elapsed time is required for

    drop breakup and, thus, more liquid is retained in the

    outlet of the tube, resulting in a longer liquid cone on thetube. In spite of its weak dependence on the viscosity

    ratio, , the volume of breakoff drops varies with ina peculiar wayit exhibits a maximum in the neighbor-

    hood of"1, as shown in Fig. 10, for the ranges ofRe,Ca, andG studied. This feature of nonmonotonical vari-

    ation of the volume of breakoff drops with the liquid

    viscosity is similar to that observed in early experiments

    (Kumar and Kuloor, 1970).

    Along with the increased elongation and extension of

    a liquid thread at large viscosity ratio, the creation of

    satellite droplets subsequent to the thread break is

    Fig. 11. Evolution with time of interface profiles of detaching drops at

    viscosity ratios of (a)"0.01 and (b) "5, with Re"1, Ca"0.01,G"1, and u

    /u

    "0.

    X. Zhang/Chemical Engineering Science 54 (1999) 17591774 1769

  • 8/13/2019 1-s2.0-S0009250999000275-main

    12/16

    expected to be interesting and complicated. Fig. 11 shows

    two series of interface profiles of drops. Each takes a close

    look at thread breakup and creation of satellite droplets,

    with viscosity ratios of "0.01 and 5 and all otherconditions the same as those in Fig. 10. Apparently, the

    stabilizing effects of the larger viscosity of the drop liquid

    (or the viscosity ratio) leads to the production of a longer,

    thicker liquid thread, as shown in Fig. 11b. Immediatelyafter its first breakup at the upper end, the thread recoils

    under the unbalanced capillary force to form a round

    knob at its top; this leads to the formation of a main,

    large satellite droplet following the secondary breakup of

    the thread. As a result of the large viscous force at the

    large viscosity ratio, a noticeably large interaction time

    (or the time interval between the two breakups), t, isobserved, namely that t"14 milliseconds for "5(Fig. 11b) as compared witht"1.6 milliseconds at thesmall viscosity ratio of"0.01 (Fig. 11a). Therefore, asshown in Fig. 11b, waves have sufficient time to develop

    and propagate on the thread surface. Along with the

    thread recoil and the secondary breakage, these waves

    grow and eventually break the thread when, at certain

    points, the amplitude of the waves is compatible with the

    radius of the thread. This leads to the creation of a sub-

    satellite droplet that has a much smaller size and lies

    between the main satellite and the cone-shaped liquid

    remaining on the tube. In contrast, the development and

    extension of a liquid thread are no longer so evident at

    the small viscosity ratio (Fig. 11a). As a result, a consider-

    ably smaller satellite droplet is created subsequent to

    double breakage of the thread. Thus, we expect the cre-ation of satellite droplets to cease or to be unmeasurable

    as the viscosity ratio becomes extremely small, for

    example, in the case of bubble generation in liquids (cf.

    Oguz and Prosperetti, 1993).

    4.6. Wall effects of the outer container

    When drops are formed in a bounded medium, the

    dynamics of their deformation and breakup is affected by

    the wall of the ambient fluid container. Wall effects have

    always been present, to a greater or lesser extent, as

    a result of disturbances of the ambient fluid by forming

    drops. Fig. 12 depicts computational results of the wall

    effects on drop formation, which are described by the

    variation of dimensionless breakoff volume, /R, and

    limiting length, /R, of a drop as a function of the ratio

    of the inner radius of the container to that of the capillary

    tube,R/R. The results in Fig. 12 are obtained by holding

    other governing parameters fixed, with Re"1,

    Ca"0.01, G"1, "1, u/u

    "0, and the boundary

    condition of rigid no slip for the container wall. Four

    interface profiles of detaching drops are inserted in

    Fig. 12 for R/R"1.35, 1.4, 2, and 4. Quantitativemeasurements of /R and

    /R in Fig. 12 indicate

    clearly that the wall effects become significant and the

    Fig. 12. Dimensionless limiting length and breakoff volume of drops as

    a function of the dimensionless inner radius of the ambient fluid

    container at Re"1,Ca"0.01,G"1,"1, and u/u

    "0.

    container radius, R, becomes a determining factor gov-

    erning the deformation and breakup of a forming drop at

    small values ofR/R. The presence of the wall causes an

    increased deformation of a detaching drop, with elonga-tion occurring in the axial direction to yield an approx-

    imately prolate ellipsoid shape (see Fig. 12). At a value of

    R/R sufficiently small that the drop size is compatible

    with the outer boundary, the drop fills most of the con-

    tainer cross section and a slug flow regime results.

    Apparently, the increased viscous effects of the ambient

    fluid delay drop breakup considerably, resulting in a sig-

    nificant increase in the breakoff volume of a drop asR/R

    decreases. As a result of increased volume of the detach-

    ing drop and its axial elongation, the limiting length of

    the drop increases dramatically when the slug flow oc-

    curs atR/R(1.5, as shown in Fig. 12. It is expected that

    as the distance between the wall and a drop, or R/R,

    increases, the wall effects will decrease. Beyond a certain

    value of R/R so that the drop no longer sees the

    container wall, the wall effects become negligible. As

    illustrated in Fig. 12, when the radius ratio R/Rexceeds

    a value of about 2.5, the variation of breakoff volume and

    limiting length of the drop with R/R becomes virtually

    unmeasurable, indicating negligible effects of the con-

    tainer wall.

    Besides the change in shape of a detaching drop with

    R/R, the inserted interfacial profiles of breaking drops inFig. 12 indicate clearly that the shape of the liquid cone

    remaining on the tube varies as well. As R/Rdecreases,

    1770 X. Zhang/Chemical Engineering Science 54 (1999) 17591774

  • 8/13/2019 1-s2.0-S0009250999000275-main

    13/16

    Fig. 13. Evolution with time of interface profiles of detaching drops at

    Re"1, Ca"0.1,G"1,"1, R/R"1.35, and u

    /u

    "0.

    the increased viscous effects delay the drop breakup, in

    a manner similar to that of increasing (discussed in

    Section 4.5). Therefore, more liquid is accumulated in theliquid cone, causing the cone shape to change from con-

    cave to convex prior to drop breakup. Fig. 13 depicts this

    breakup feature of a drop at the slug flow regime with

    R/R"1.35 and all other conditions the same as those in

    Fig. 12. Unlike drops forming into an unbounded me-

    dium (cf. Fig. 6a), no measurable thread is developed

    between the breakoff drop and the liquid cone and hence

    the drop breaks sharply at its middle portion, as the

    result of the accumulation of a large volume of liquid in

    and convexity of the liquid cone at such a small value of

    R/R. Obviously, in such a situation, a satellite droplet

    can no longer be created subsequent to the drop breakup.Moreover, a comparison of the breakup features of drops

    within a tightly bounded medium (Fig. 13) with those in

    an infinite medium (Fig. 6a) indicates clearly that the

    breakup time of the drop, along with its breakoff volume

    and limiting length, is significantly increased by the wall

    effects as R/R decreases.

    4.7. Effect of the externalflow

    Flowing an ambient fluid concurrently with the dis-

    persed phase has been used as an effective and practicalmethod for reducing the volume (or size) of breakoff

    drops by accelerating drop detachment with an

    Fig. 14. Evolution with time of the dimensionless length of drops

    forming in an ambient fluid which flows at different characteristic

    velocities ofu/u

    "0, 50, and 200 at Re"1, Ca"0.1, G"1, "1,

    and R/R"2.

    additional viscous shear force (Clift et al., 1978). Besides,

    an external uniform flow is found to vary the interface

    profile and breaking length of a drop during drop forma-tion considerably at infinitesimal inertial effects (Zhang

    and Stone, 1997). Fig. 14 presents the entire history of the

    formation of three drops that are each subjected to an

    external flow of different velocities. The drop evolution is

    presented quantitatively in terms of the drop length, ,

    nondimensionalized by the inner radius of the tube, R.

    The results shown in Fig. 14 are calculated for the three

    situations of drop forming in an ambient fluid for which

    fully developed flows are applied concurrently at differ-

    ent average velocities of u/u

    "0, 50, and 200, respec-

    tively. Other controlling parameters are held constant:

    Re"1, Ca"0.01, G"1, "1, R/R"2, and R

    /R

    "1.1 (see Fig. 1). Evidently, the external flow of the

    ambient fluid plays an important role in drop evolution

    and breakup, as clearly depicted by the two series of

    instantaneous profiles of drops at u/u

    "0 (the lower

    series) and 200 (the upper series), respectively, in Fig. 14.

    When injected into a container with a specified flow rate

    (or average velocity), the ambient fluid interacts with

    a forming drop and brings about an additional hy-

    drodynamic force on it. Initially, at the growing stage, the

    drop of small volume grows and rises slowly at the tip of

    the tube. The entering ambient fluid, if possessing a suffi-ciently high velocity as shown in Fig. 14 with u

    /u

    "200,

    passes over the drop rapidly and expands, resulting in

    X. Zhang/Chemical Engineering Science 54 (1999) 17591774 1771

  • 8/13/2019 1-s2.0-S0009250999000275-main

    14/16

    a reduction in its velocity and, consequently, an increase

    in the ambient pressure (Batchelor, 1967). As a result of

    the large pressure on its top, the drop deforms and

    flattens at its tip; then the drop length, , is measured to

    be shorter than that in quiescent or slowly flowing me-

    dium, as shown in Fig. 14 at t(0.12 s. Along with the

    increase in the interface area of the drop at longer time,

    the viscous force on the drop interface induced by theexternal flow becomes important and stretches the drop

    into a spheroid in the axial direction, whereas the drop is

    seen to be pear shaped under the gravitational force

    when the external flow is vanishing, as shown in the

    inserted profiles in Fig. 14. Differing from body forces,

    which pull the drop liquid away from the tube, such as

    the gravitational force (cf. Fig. 8), the external flow acts

    on the drop interface as a viscous shear stress, which

    stretches and elongates the drop and, therefore, leads to

    a great increase in the drop length. At the extremely high

    velocity of the external flow, such as that ofu/u

    "200 in

    Fig. 14, a thread about three time longer than the tube

    radius is developed during necking and breakup of the

    drop. It is expected that a large satellite droplet, or

    multiple satellites, will be created subsequent to the

    thread breakup. This phenomenon is similar to that

    observed in the formation of conducting or insulating

    drops driven by an external electric field, which acts with

    electric charges and induces an electrostatic force on

    drop surfaces (Zhang and Basaran, 1996).

    Fig. 15 shows variations of breakoff volume and limit-

    ing length of drops as a function of the average velocity of

    an external flow, u, normalized by the average velocity ofthe drop liquid ejecting from the tube, u

    under condi-

    tions that are otherwise the same as those used in Fig. 14.

    Three typical interface profiles of drops at the instant

    when they are about to detach are inserted for external

    flows ofu/u

    "1, 100, and 200. As expected, the external

    flow, acting as an additional force assisting to break up

    drops, reduces the breakoff volume of the drops. As the

    velocity of the external flow (or its dimensionless term

    u/u

    ) increases, the volume of breakoff drops decreases

    and, correspondingly, the dripping rate increases mono-

    tonically. It is noteworthy that since the volume of

    breakoff drops decreases as the velocity of an external

    flow increases, the increase in the limiting length, /R,

    withu/u

    (shown in Fig. 15) is obviously attributed to an

    increase in the thread length. The concurrent increase in

    drop elongation and decrease in the volume of breakoff

    drops greatly affect the dynamics of thread rupture and

    the subsequent generation of satellite droplets. Fig. 16

    depicts the feature of a drop breaking into a flowing

    medium with u/u

    "200 and all other conditions the

    same as those in Fig. 14. Because of the large shear stress

    on the drop interface generated by the external flow, the

    drop produces a long, thick thread prior to its breakup.Immediately following the detachment of a primary

    drop, the newly freed tip of the thread tends to retract

    Fig. 15. Dimensionless limiting length and breakoff volume of drops as

    a function of the dimensionless velocity of the flowing media atRe"1,

    Ca"0.01,G"1,"1, and R/R"2.

    Fig. 16. Evolution with time of interface profiles of detaching drops at

    Re"1, Ca"0.1, G"1,"1, R/R"2, and u

    /u

    "200.

    downward under the unbalanced capillary force. How-

    ever, owing to the great opposition from the external

    flow, the tip of the distorted thread reaches a minimum

    height and then starts to ascend att"0.154 s. As it does

    so, a bulbous head develops at the tip of the thread

    and continues to expand. A secondary neck then be-gins to develop immediately below the head at

    t"0.17 s. Eventually, a second breakup occurs to create

    1772 X. Zhang/Chemical Engineering Science 54 (1999) 17591774

  • 8/13/2019 1-s2.0-S0009250999000275-main

    15/16

    a secondary drop with the volume almost 10% of that of

    the primary drop. This is in contrast to the situation with

    quiescent or slowly flowing ambient fluid where the sec-

    ondary (or satellite) drop has a volume less than 1% of

    that of the primary drop (cf. Fig. 6(a)). The large differ-

    ence in the thread breakup is also indicated by difference

    in the interaction time. Namely, in a quiescent ambient

    fluid, the time interval that elapses between the twobreakups is about 3.6 ms as shown in Fig. 6a, while with

    an external flow ofu/u

    "200, it takes more than 17 ms

    (see Fig. 16). Hence, for the latter case, the secondary

    breakup can be regarded as a breakup entirely separate

    from the first one; evidently the liquid jetting occurs to

    generate drops of nonuniform sizes.

    5. Concluding remarks

    Through both theoretical and experimental means, we

    have investigated the effects of inertial, viscous, gravi-

    tational, and surface tension forces on drop formation.

    The emphasis has been on determining the importance of

    the dynamic effects of an ambient viscous fluid on drop

    evolution and breakup. The VOF/CSF numerical algo-

    rithm, on which our numerical model is based, allows

    calculations to pass the breaking point during drop

    formation continuously without numerical modifications

    to overcome the singular nature of the interface rupture.

    This feature makes it possible to characterize numerically

    the double breakage of a liquid thread and the sub-

    sequent generation of satellite droplets under differentconditions. Reassuringly, the numerical predictions show

    excellent agreement in comparison with experimental

    measurements for a typical liquidliquid system of 2EH

    drops forming in water.

    According to the results obtained in the present study,

    the maximum (or limiting) length that a drop attains

    prior to its breakup and the volume of a drop that breaks

    off from the tube increase significantly with increased

    Reynolds and capillary numbers and with decreased

    gravitational Bond number of the drop liquid. During

    necking and breakup of a drop, a liquid thread, which

    connects to a detaching drop and a cone-shaped liquid

    mass sessile on the tube, develops and stretches. Under

    certain conditions, the thread breaks at both ends se-

    quentially, resulting in satellite droplets. When drops

    form into a quiescent ambient fluid, whether the satellite

    droplet is created subsequent to the thread breakup is

    substantially dependent on the relative importance of the

    inertial and surface tension forces. Dynamic effects of an

    ambient fluid have been shown to play an important role

    in drop formation. In particular, the increasing wall ef-

    fects of an ambient fluid container cause a significant

    increase in the volume of breakoff drops and a largereduction in the length of the liquid threada factor that

    may alter the possibility of creation of satellite droplets.

    In contrast, an external flow of the ambient liquid leads

    to a great increase in the thread length and radius but to

    a decrease in the volume of breakoff drops. Moreover,

    as a result of viscous damping of the interface distur-

    bances, a long, thin thread is developed when the ratio of

    the viscosity of the drop liquid to that of the ambient

    fluid becomes sufficiently large, resulting in the genera-

    tion of multiple satellite droplets subsequent to dropbreakup.

    Acknowledgments

    This work was supported by the Division of Chemical

    Sciences, Office of Basic Energy Sciences, U.S. Depart-

    ment of Energy under contract DE-AC05-96OR22464

    with Lockheed Martin Energy Research Corp. The

    author performed this study when he was working at

    Oak Ridge National Laboratory, Tennessee. The author

    thanks Dr. J. R. Richards of Dupont for providing the

    details of his calculations.

    References

    Batchelor, G.K. (1967). An introduction to fluid dynamics, New York:

    Cambridge at the University Press.

    Berg, J.C. (1993). ettability. New York: Marcel Dekker.

    Bird, R.B., Stewart, W.E., & Lightfoot, E.N. (1960). ransport phe-

    nomena. New York: Wiley.Bogy, D.B. (1979). Drop formation in a circular liquid jet. Annual

    Review of Fluid Mechanics, 11, 207228.

    Brackbill, J.U., Kothe, D.B., & Zemach, C. (1992). A continuum method

    for modeling surface tension.Journal of Computational Physics 100,

    335354.

    Clift, R., Grace, J.R., & Weber, M.E. (1978). Bubbles, drops, and par-

    ticles. New York: Academic Press.

    Cram, L.E. (1984). A numerical model of drop formation. In J. Noye,

    & C. Fletcher (Eds.), Computational techniques and applications:

    CAC-83. Amsterdam: Elsevier.

    Eggers, J., & Dupont, T.F. (1994). Drop formation in a one-dimensional

    approximation of the NavierStokes equation. Journal of Fluid

    Mechanics262, 205221.

    Harris, M.T., & Byers, C.H. (1989). An advanced technique for inter-facial tension measurements in liquidliquid systems. ORNL/TM-

    10734.

    Heertjes, M.P., de Nie, L.H., & de Vries, H.J. (1971). Drop formation

    in liquidliquid systems I Prediction of drop volumes at

    moderate speed of formation. Chemical Engineering Science 26,

    441449.

    Hirt, C.W., & Nichols, B.D. (1981). Volume of fluid (VOF) method for

    the dynamics of free boundaries.Journal of Computional Physics39,

    201225.

    Kothe, D.B., & Mjolsness, R.C. (1992). Ripple: A new model for

    incompressible flows with free surfaces. AIAA Journal 30,

    26942700.

    Kumar, R., & Kuloor, N.R. (1970). The formation of bubbles and drops.

    Advances in Chemical Engineering 8, 255368.

    Michael, D.H. (1981). Meniscus stability.Annual Review of Fluid Mech-

    anics13, 189215.

    X. Zhang/Chemical Engineering Science 54 (1999) 17591774 1773

  • 8/13/2019 1-s2.0-S0009250999000275-main

    16/16

    Oguz, H.N., & Prosperetti, A. (1993). Dynamics of bubble growth and

    detachment from a needle. Journal of Fluid Mechanics 257,111145.

    Peregrine, D.H., Shoker, G., & Symon, A. (1990). The bifurcation of

    liquid bridge.Journal of Fluid Mechanics 212, 2539.

    Richards, J.R. (1994). Fluid mechanics of liquid-liquid system. Ph.D.

    dissertation, University of Delaware.

    Richards, J.R., Beris, A.N., & Lenhoff, A.M. (1995). Drop formation in

    liquidliquid systems before and after jetting. Physics of Fluids, 7,

    26172630.Scheele, G.F., & Meister, B.J. (1968). Drop formation at low velocities

    in liquidliquid systems: Part I. Prediction of drop volume.

    A.I.Ch.E. Journal 14, 919.

    Schulkes, R.M.S.M. (1993). Nonlinear dynamics of liquid columns:

    A comparative study. Physics of Fluids A, 5, 21212130.

    Schulkes, R.M.S.M. (1994). The evolution and bifurcation of a pendant

    drop. Journal of Fluid Mechanics 278, 83100.

    Shi, X.D., Brenner, M.P., & Nagel, S.R. (1994). A cascade of structure in

    a drop falling from a faucet. Science, 265, 219222.

    Zhang, D., & Stone, H.A. (1997). Drop formation in viscous flows at

    a vertical capillary tube. Physics of Fluids 9, 22342242.

    Zhang, X. (1999) Dynamics of growth and breakup of viscous

    pendant drops into air. Journal of Colloid and Interface Science,

    In press.Zhang, X., & Basaran, O.A. (1995). An experimental study of dynamics

    of drop formation. Physics of Fluids 7, 11841203.

    Zhang, X., & Basaran, O.A. (1996). Dynamics of drop formation from

    a capillary in the presence of an electric field. Journal of Fluid

    Mechanics326, 239263.

    1774 X. Zhang/Chemical Engineering Science 54 (1999) 17591774