for coalescence

Upload: abhijeet-h-thaker

Post on 07-Jul-2018

221 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/19/2019 For coalescence

    1/8

    Chemfcol Engineering Scknce, Vol. 41, No. I, pp. 65-72, 1986.

    ocQ9-2509/86 s3.00 + 0.00

    Printed in Great Britain.

    0 1986. Pergamon Press Ltd.

    BREAKAGE OF VISCOUS AND NON NEWTONIAN DROPS

    IN STIRRED DISPERSIONS

    J. S. LAGISET TY, P. K. DAS and R. KUMAR

    Department of Chem ical En gineering, Indian Institute of Science, Bangalore 560 012, India

    and

    K. S. GANDHI

    Department of Chem ical Engineering, Indian Institute of Technology. Kanpur 20801 6, India

    (Received 15 February 1985)

    Abstract-A model of breakage of drops in a stirred vessel has been proposed to account for the effect of

    rheology of the dispersed phase. The deformation of the drop is represented by a Voigt element. A realistic

    description of the role of interracial tension is incorporated by treating it as a restoring force w hich passes

    through a maxim um as the drop deforms and eventually reaching a zero value at the break point. It is

    considered that the drop will break when the strain of the drop has reached a value equal

    to

    its diameter. An

    expression for maxim um stable drop diameter, d,.

    is derived from th e model and found to be applicable

    over a wide range of variables, as well as to data already existing in literature. The model could be naturally

    extended to predict observed values of d,, when the dispersed ph ase is a power law fluid or a Bingham

    plastic.

    INTRODUCTION

    The analysis of rate processes in liquid-liquid disper-

    sions requires knowledge of the sizes of the drops

    existing in the vessel To estimate the drop sizes, it is

    essential to know the break-up mecha nism of a drop in

    a turbulent dispersion. Hinze (1955) suggested a model

    for predicting maximum stable drop diameter by

    comparing the restoring elastic stress in

    the

    drop due

    to interfacial tension with the inertial stress across the

    drop diame ter. He proposed that break-up of a drop

    occurs w hen the ratio of the inertial stress to the elastic

    stress, i.e.

    PCu* (4 40,

    exceeds a critical value.

    Assum ing that turbulence is isotropic a nd that the

    diameter

    d

    1 (Kolmogoroff length), the mean square

    velocity fluctuation across distance, d, is given by

    UZ(4) cc (& ‘3.

    (1)

    If the Reynolds number is sufficiently large, then for a

    fully baffled vessel, the power dissipation per unit

    mass, E, can be expre ssed a s

    E a N3D2 .

    (2)

    Substituting eqs (1) and (2) in Hinze’s proposed

    criterion, Shinnar (1961) derived the following equa-

    tion for the maxim um stable drop diameter,

    d ,,,:

    d

    -225 = constant (We)-o.6

    D

    (3)

    where We is the Weber number.

    Equation (3) has been used by a numbe r of investi-

    gators. Coulaloglou and Tavlarides (1976) have dis-

    cussed exhaustively all the correlations available to

    predict

    d,

    in a turbulent dispersion.

    Equation (3) is based on the assump tion that there is

    hardly any difference in densities and viscosities of the

    two phases. The viscosity of the dispersed phase has,

    however, been found to have a significant influence on

    the drop size (Arai et al. , 1977; Ko nno et al ., 1982). The

    viscous stress resists the flow inside the drop, leading

    to

    its breakage, and som etimes this stress has the same

    order of magnitude as the inertial stress due to pressure

    fluctuation. There exists very limited information (Arai

    et al . , 1977; Konno et al. 1982) on the prediction of

    maxim um stable drop size taking into account the

    effect of dispersed phase viscosity. Konno et a l . (1977)

    were the first to propose a model for predicting the

    maxim um stable drop diameter incorporating the

    effect of dispersed phase viscosity. They considered

    that the deformation of a drop under external stress

    can be described by the Voigt model which simul-

    taneously takes both interfacial tension (restoring

    force) and viscous dissipation (due to resistance to flow

    inside the drop) into account. In deriving the model

    equations, they assumed that the pressure fluctuation

    across a distance due to turbulent flow is periodic and

    the break-up of a drop occurs when the deformation

    strain (0) reaches a critical value, 13~~. They have

    obtained a semi-empirical correlation for the maxi-

    mum stable drop diameter in terms of two dimension-

    less groups, Weber numbe r and viscous number.

    Although their semi-empirical expression predicts

    their experimental re sults reasonably well, the model is

    open to criticism on several grounds. The Voigt mode l

    has a maxim um equilibrium deformation and it is

    reversible. As such, it is hard to v isualize the breaking

    of a drop through a classical Voigt model. Arai ef al.

    (1977) overcame this problem through the artifice of

    associating drop breakage with a maxim um deforma-

    tion, Qr,

    which is left as an arbitrary parameter.

    Further, the regular and periodic pattern assigned by

    them to the turbulent flow around the drop does no t

    CES 41:1-E

    65

  • 8/19/2019 For coalescence

    2/8

    66

    J. S. LAGISEITY et al.

    appear realistic, especially on the length scale of drop

    size being dealt with. Apart from this, their mode l does

    not give rise to the low viscosity limit of the maximum

    stable drop size naturally and had to be introduced in

    an ad hoc mann er. Finally, their model does not yield a

    final expression which can be use d directly for evaluat-

    ing the two constants contained in the model. Instead,

    the final expression is assumed and the two constants

    involved in it are evaluated from their own experimen-

    tal data. Such a procedure does not allow extension of

    their model to other rheologically complex fluids.

    The present work aims at proposing a new phenom-

    enological model which accounts for the effect of

    viscosity of the dispersed phase in a more ra tional way.

    It also tests the validity of the model against exper-

    imental results.

    EXPERIMENTAL

    The experimental apparatus consisted of a glass

    vessel of 14.5 cm i-d. and 20 cm height. The impeller

    used was a six-bladed disk turbine, placed centrally in

    the mixing vessel through a stainless-steel shaft, which

    in turn was connected to a motor, the speed of which

    could be regulated. A set of four, equally spaced

    stainless-steel balIles, arranged vertically at the wa ll of

    the vessel, was used. The width o f each baffle was equal

    to one-tenth of the diameter of the vessel. A schem atic

    diagram of the equipment is shown in Fig. 1.

    The continuous phase was taken in the stirred vessel

    and the stirrer speed ra ised to the desired value. The

    dispersed phase w as then added to the vessel. The

    dispersed phase volume fraction was kept at less than

    0.02 to minimize coalescence. The equipment was run

    for 1 h to achieve steady state. Samples were then

    D.14.5 cm

    Fig. 1. Diagram of the stirredvessel.

    drawn from the vessel and the particle sizes measu red

    with an optical microscope. At least 150 drop di-

    ameters were measu red to obtain the value of d,. To

    avoid coalescence of drops during sampling and

    subsequent size measu rements, polyvinyl alcohol

    (PVA) was added immediately before drawing the

    sample. This was done when the aqueous phase was

    continuous_ When the aqueous phase was dispersed,

    PVA was added to it at the beginning itself.

    The liquids used together with their properties and

    experimental conditions are listed in Tables 1 and 2,

    respectively. Interfacial tensions were determined by

    the pendant drop m ethod. Viscosities w ere measu red

    with a coaxial cylinder viscometer.

    DEVELOPMENT OF THE MODEL

    When a viscous droplet becomes deformed in a

    turbulent flow field, the viscous stress due to the

    internal flow will act simultaneously with the interfa-

    cial tension force to resist the deformation of the drop

    against the external inertial stress arising from the

    turbulent p ressure fluctuations. If the inertial stress is

    sufficiently large and sustained over a long enough

    time interval, the drop w ill deform and a thin column

    of liquid will appear so mew here in its bulk. The

    column will break due to instability producing two

    daughter droplets. This has been assumed to occur

    when the magnitude of deformation is of the order of a

    drop diameter. As the interfacial tension restoring

    force is absent both at zero deformation and at the

    breakage point, this restoring force should pass

    through a maxim um as deformation proceeds. In this

    connection, it is interesting to note that based on

    calculations for low Reynolds numbers, Rallison

    (1984) states that for globular drops, surface tension

    acts as a restoring force but that surface tension may

    even promote breakage once the drop becom es

    elongated. As the interfacial tension and viscous stress

    simultaneously oppose the inertial stress, the Voigt

    model offers a suitable description of drop deforma-

    tion. How ever, the elastic stress due to interfacial

    tension needs to be represented in a more realistic

    fashion. The Voigt m odel corresponding to the process

    of deformation is presented in Fig. 2.

    In accordance with the Voigt. model, the applied

    stress (T) must be equal to the opposing elastic stress

    due to interfacial tension (rJ plus the viscous stress 7,).

    The governing equation therefore has the following

    form

    7 = 7, + 7”.

    (4)

    Table 1. Experimental conditions

    Impeller speed

    Reynolds number of the

    continuous phase

    Dispersed phase volume

    fraction

    Temperature

    3.33-iOrev/s

    1.5 x l@-5 x 104

    0.02

    26°C

  • 8/19/2019 For coalescence

    3/8

    Breakage of drops in a stirred suspension

    Table 2. Properties of the $ontinuous and dispersed phases for the systems studied

    67

    Continuous phase

    Dispersed phase

    Description

    c1

    P

    Description

    K

    T

    (Poise)

    (Mn3)

    13/(cn=2-“)

    (dynes/cm) (g/&3)

    Water

    0.01

    1.0

    Polystyrene

    0.43-

    1 20

    0X8-0.92

    in styrene

    37.50

    lO-30% by

    wt

    Kerosene 0.021 0.78 lOOm1 of 14.5 2/3 50 1

    CMC in water

    (2.5 %I +

    60ml of 2%

    PVA

    CaCO, aqueous

    0.137 1 45.2 1.47

    suspension

    (59.5 % Caco3 +

    2.00 y0 polyvinyl

    alcohol)

    -

    i

    Fig. 2. Voigt model for the deformation of a drop.

    Constitutive equation for T,

    When the drop suffers a small but finite strain, the

    interfacial tension force generates a restoring stress

    proportional to the strain. However, as the drop

    undergoes further deformation and approaches the

    breaking point, its structure will be more like that of a

    dumb bell: two “yet to be born” daughter droplets

    connected by a thin filament of liquid. The filament will

    break because of its inherent instability. Thus near the

    break point,- surface tension has no restoring effect.

    Hence a more realistic description of the role of surface

    tension must be to assign to it a retractive force that

    increases for small deformations of the drop but

    decreases afterwards and e ventually reaches zero at the

    break point. Assuming that the total deformation

    undergone by the drop up to the point of breakage is of

    the order of ma gnitude of

    d,

    the following simple

    functional relationship between 5, and 0, has been

    assumed.

    TS= $ es (1 -e,>

    e, c 1

    =

    0

    e 3 I.

    (5)

    The condition that 7, is zero for 0, > 1 explicates the

    essential feature of the model, i.e. the drop has reached

    its break point when the dimensionless strain has

    reached the value of unity and surface tension can no

    longer bring it back to its original state. Thus the Voigt

    model proposed in the present work cannot retract to

    its original state, once 0, & 1, even when the applied

    stress is withdrawn.

    Constitu tive equation for TV

    For the fluids considered in this work, the constitut-

    ive equation for viscous stress can be written in a

    generalized form as

    7v = TV + K (de,/dtY.

    (6)

    Equation (6) simplifies to the case of Newtonian,

    Bingham plastic and power law models when ap-

    propriate simplifications are made.

    Model equation

    For a Voigt element, the total strain 8 = 8, = 0,.

    Noting this and substituting eqs (5) and (6), respect-

    ively, in eq. (4), we obtain

    r-=,--ae l-e8)+K

    d

    c

    1.

    (7)

    The flow inside the drop during breakage will be quite

    complex. In the present model, however, it has been

    assumed to be a simple shear flow.

    Description of z

    In eq. (4), T represents the dynamic pressure dif-

    ference across the drop diam eter. In a turbulent flow

    field the pressure difference normally associated with

    the arrival of eddies is a quantity which fluctuates with

    time and distance over which it operates. However,

    exact knowledge of the fluctuations is not yet available.

    Therefore, it is assum ed that the deformation takes

    place under the influence of a mean stress which

    remains constant over the average life time of an eddy.

  • 8/19/2019 For coalescence

    4/8

    68

    J. S. Llsorszrrv et al.

    Expr ess ions for aver age va lu es of 7 and T

    Let the average values of 7 be + and the average life

    time of an eddy beF The expressions for these average

    values given by Hinze (1955) a nd Coulaloglou and

    Tavlarides (1977) have been used in the present

    investigation. They are

    and

    7 ~PcU2W (8)

    =&*

    (9)

    It is necessary to express u2 (d) in terms of the

    parameters associated with the stirred vessel. The

    expressions for the rate of energy dissipation per unit

    mass, E, and u2 (d) given by Coulaloglou and

    Tavlarides (1977) are

    E = 0.407 IV= D 2

    (10)

    and

    u2 (d ) = 1.88 s2’3 d2’3.

    (11)

    Therefore + can be expressed as:

    + -_ C pc N2 D4/= d2j3 (12)

    where C is a constant dependent only on the geometry

    of the tank and agitator. The average life time is given

    by

    (13)

    The drop experiences no stress prior to its coming

    under the influence of an eddy a nd from then onwards,

    a constant stress equal to + for a tim e interval of T

    Thus

    r = O t t o

    =p

    o< t gT

    Substituting these into eq. (7), we obtain

    C PcN2 D413 ‘ 213 _

    de n

    Tg +1 -e)e+K -&

    .

    (14)

    (>

    Equation (14) is more conveniently represented in

    dimensionless form as

    (dlJ/dt = C We (d/D)‘ / ’

    - (q,d/o) - (0 - 02) (15)

    where rl is the nondimen sional time given by

    q = (o /dK)“” t .

    (16)

    Max i mu m s tab l e d r op s i ze

    The drop is exposed to a stress of i only for a time

    period T After this time, the eddy wou ld have dis-

    sipated its energy and the external stress on the drop

    becomes zero once again. If at the end of the time

    interval T the value of B -Z 1, then the surface tension

    spring w ould still have a finite retractive stress and the

    drop would return to its original state. Thus for the

    drop to break it is not on ly ne cessary that 8 = 1 is

    reached but also that it should be attained at some

    point during the time intervalT. The time required to

    reach 0 = 1 can be obtained by integrating eq. (15); an

    initial condition is needed to do so. The dashpo t

    cannot show any instantaneous deformation on the

    application of the stress. Hence

    8=0 at q=O_

    (17)

    By separating the variables, eq. (15) can be written as

    18)

    where

    a = C We (d/D )s’3 - (T,, d/u) - l/4.

    (19)

    The solution of eq. (18) w ith the initial condition

    depends on the constant a in eq. (19). There are two

    possible solutions corresponding to a < 0 or a > 0. If

    a -Z 0, it can be shown that w hen t -B co, 8 --r [l/2

    - fi]_ Thus, Bean never reach a value of unity, and

    breakage is not possible in the finite time interval ofT_

    But tI can reach unity in a finite time period when a

    > 0, and therefore to predict the maxim um drop

    diameter, only values of a greater than zero need to be

    considered.

    Our main concern is to know the max imum size of a

    drop that can ex ist in the stirred vessel. All drops above

    such a size will break. As stated earliei, the drop has

    been assum ed to de form under the influence of T for a

    period of?? T he max imum stable drop size therefore is

    the max imum threshold drop size which reaches a

    deformation of unity when exposed to a constant stress

    F during the time interval F. Such a drop size can be

    calculated as follows.

    Equation (18) is integrated to find rl at 8 = 1. If rl at

    8 = 1, the dimensionless time required to reach defor-

    mation correspon ding to breakage, is more than the

    nondimen sional life time of the eddy, breakage wo uld

    not occur. Thus for breakage to occur, the following

    condition must be satisfied:

    q (0 = 1) <

    (u /d IQ1 ’“Z

    (20)

    A s

    the applied stress increases with diameter, q (0 = 1)

    decreases with increasing diameter. The maximum

    stable drop size is that for which

    q (0 = 1) = (a/d_, K)““T.

    (21)

    Thus eq. (18) can be integrated up to rl(0 = 1)

    = (a /d_K)““T to find the maxim um stable drop

    size. Using the expression for T in eq. (9), ~(0 = 1) can

    be written as

    r.t(e = 1) =

    (Re/We)“” (d/D)2’3-“n

    where

    Re =

    D” NoI2 - n

    ,

    2”-‘ (3+l /n r K

    (23)

    Forr,=Oandn= 1,1/3,1/2and2/3,eq.(18)canbe

    integrated analytically; the solutions are listed in

    Table 3. It should be noted that C = 8.0 has been used

    in these results for both Newtonian and non-

    Newtonian fluids. This will be discussed further in the

    Results and Discussion.

    It should be noted that the mode l developed in the

  • 8/19/2019 For coalescence

    5/8

    Breakage of drops in a stirred suspension

    Table 3. Solution of eq. (18) for non-Newtonian fluids

    n

    Solution of the model equation

    [32 We (d_/D)s’3 - I]“* tanm

    =

    (RejWe) (d-/D)- I i3

    >I

    Re/We)3/2 (d-/D)- 5’ 6

    u2

    1.6

    51.2

    We d_/D)“ ’ (32 We d_/D)513 -

    1) +

    [32 We d_/D)5’3 - l]3/2

    x

    tan-’

    c32 We(d_;D)“” _ 1,~,2 = cRe/~~’ axP-4’3

    1’3

    2.66(44We(d_/D)5’ 3 -1

    [(32We(d_/D)“ ” -l)* Wez(d_/D)‘ o’ “] + (32 We(d,,,f/ ” - 1)5f*

    x tan-’

    [32 We d_;D)= - l]“*

    = [Re/

    We]

    (d- /D) - ‘ j3

    present work involves a single parameter whereas

    the model proposed by Arai et al. (1977) requires the

    knowledge of two parameters to be evaluated from

    experimental data. A positive

    test that can be

    applied to

    the present mode l is that the constant C involved in eq.

    (12), which relates the turbulent pressure fluctuations

    to d, N and

    D,

    should not depend on the rheological

    properties of the dispersed phase. If this is satisfied,

    then clearly the present mod el offers a distinct ad-

    vantage over the model of Arai et al. (1977) which

    requires the evaluation of two parameters.

    RESULTS

    AND DISCUSSION

    Newtonian fluids and inviscid limit

    One of the interesting aspects of the model is that as

    the dispersed phase viscosity becomes very small, i.e.

    ,ud+ 0, the solution of the model equation redu ces to

    the well-known and widely tested equation

    given

    by

    Shinnar (1961). For example, looking at the equation

    for a Newtonian fluid (n = 1) presented in Table 3, the

    only solution that is possible as pd - 0 or

    Re -+ co

    is

    that We (d_/D)5’ 3

    -+ constant. This result is identical

    to eq. (3) reported by Shinnar (1961).

    Using the data of drop diam eters against dispersed

    phase viscosity for different speeds, the value of

    constant C in eq. (12) has been determined. The value

    of the constant worked out to be 8.0. Figures 3 and 4

    show the experimental drop diameter, d,, against

    dispersed phase viscosity, pd. and impeller speed. The

    results predicted by the model are shown as lines

    whereas the points are experimental results. The

    theoretical predictions are in good agreement with the

    observations. From this, it can be seen that C is

    dependent on ly on the geometry of

    the

    vessel and

    agitator, and not on cl,, . Further, it is worth noting that

    the effect of dispersed phase viscosity on d, becomes

    significant only after about 20 cP.

    The predictions of the present model and the

    numerical value of constant C can be further tested

    using the experimental observations of other workers.

    69

    Fig. 3. Com parisonof experimental esultswith the present

    model (effectof Jo)_

    When the dispersed phase viscosity becomes small, i.e.

    p, -V 0, the model predicts that

    (d-/D) = 0.125 (We)-“* 6.

    Sprow (1967) ha s given the following empirical re-

    lationship when the dispersed ph ase viscosity is small:

    (d-/D) = a1 (We)-“.6.

    The value of a 1 ies between 0.126 and 0.15, w hich is in

    agreement with the value found in the present work.

    The present value of 0.125 is somew hat better because

    it fits a large amo unt of other experimental data. For

    instance, the experimental results of Arai et al. (1977)

    also agree with the present model. This is shown in

    Fig. 5.

    The present model can be easily modified to in-

    corporate the effect of the dispersed ph ase hold-up.

    The value of the mean square velocity, u* (d), changes

    with the change in the dispersed ph ase hold-up. The

  • 8/19/2019 For coalescence

    6/8

    J. S. LAGISETN et al.0

    0

    0.

    E 0.

    1

    0.

    0

    I

    I

    L

    I 300

    Loo m 6

    Impcllsr speed N (r.p.m 1

    pd c ‘3 CP

    o- z 20 dyneslcm

    D = 4.5 cm

    Fig. 4. Effect of impeller speed on the maximu m stable drop

    diameter.

    I

    1.5

    d

    I

    B

    /

    i L-0

    d

    I J

    / /

    /’

    d’

    0.5

    /‘)/’

    _ A;./’

    ___--b-

    _--,_A=

    Experimental

    Theoretical

    l3.P.M

    D

    xl0

    0

    400

    0 -.- 600

    a- = 20 dyneslcm j D= 6. 5 c m

    Fig. 5. Comparison of experimental results (Arai et al. with

    the present model.

    value of m

    changes with 4 as (Coulaloglou and

    Tavlarides, 1977)

    m+=+ = (1

    +a2+)-2.0[uz(d) ]+o.

    (24)

    Using the modified expression for u2 (d) , the equation

    for maximum drop size for Newtonian fluids in the

    limit of small viscosity becomes

    d

    z = 0.125 (1 + r~~+)“ ~ (We)-“.6.

    D

    c-w

    The data available in the literature is in terms of dx2.

    Sprow (1967) and Coulaloglou and Tavlarides (1976)

    indicate that da2 is linearly related to d,. For

    continuous system s, Coulaloglou and Tavlarides

    (1976) indicate that d, = 1.5 dS2. Therefore

    d

    2 = 0.083 (1 + a2 )‘.” (We)-o.6.

    D

    (26)

    It is found that for a2 = 4.0, the model predictions

    given by eq. (26) fit the experimental data of

    Coulaloglou and Tavlarides (1976) on continuous

    systems reasonably well. This is shown in Fig. 6.

    It can be concluded that the value of 8.0 assigned to

    constant C is related only to the geometry of the vessel

    and agitator, and arises from the nature of the

    turbulent fluctuations. It is not affected by the prop-

    erties of the dispersed phase, which have been varied

    over a wide range. The m odel itself can then be seen to

    account for the variation of the viscosity of the

    dispersed phase. This can be tested further by examin-

    ing non-Newtonian fluids.

    Non -New ton i a n i d s

    Figure 7 presents the experimental results of maxi-

    mum stable drop diameter with change in impeller

    speed for a power law fluid (2.5 0A CMC in water

    + PVA, n = 2/3). The grade of CM C used was the

    same as that employed by Kumar and Saradhy (1972).

    These authors u sed a test attributed to Philipoff to

    ensure the absence of viscoelasticity in CMC solutions

    up to a concentration of 4% by weight. The other

    compon ent of the solution use d in the present work

    was PVA, with a molecular weight of about 70,000. At

    this molecular w eight and low concentrations, PVA

    will also not impart an elastic character to the sol-

    utions. Therefore the solution em ployed in this work

    can be safely considered to be inelastic. The drop

    diameter decreases with impeller speed in a nonlinear

    fashion. As m entioned earlier, according to the model,

    +

    0 0.15

    *

    0.Y)

    . 0 05

    0 0.025

    a2 z 4.0

    Fig. 6. Comparison of

    Coulaloglou and Tavlarides’ exper-

    imental values (effect of 4) with those predicted by the present

    model.

  • 8/19/2019 For coalescence

    7/8

    Breakage of drops in a stirred suspension

    71

    Impeller speed ti (r.p.m 1

    Fig. 7. Comparison of experimental results with those pre-

    dicted by the present model for a power law fluid.

    constant C must be independent of the rheology of the

    dispersed phase. The solid line represents the results

    calculated from the model using constant C evaluated

    with the experimental results of Newtonian fluids, i.e.

    calculated from the second equation of Table 3. It was

    found that the model predicts the results reasonably

    well. This lends further credence to the model.

    Gene ra l i zed rep resen t a t i on

    The model predictions can be represented in gener-

    alized graphs b y plotting

    (d- /D)

    against the Reynolds

    number,

    Re

    defined on the basis of the viscosity of the

    dispersed phase) with the Weber numb er, We, as a

    parameter. Figure 8 shows plots of

    d- /D

    against

    Re,

    with We as a parameter for the values of n equal to I,

    2j3 and 1 3. The plots show that at low Reynolds

    numb ers, i.e. at high va lues of p,, all the curves merge,

    indicating that interfacial tension does not play a

    significant role in the breakage of drops. This result is

    expected, since when the drop is sufficiently viscous,

    the applied force essentially overcomes the mean

    viscous resistance, compared to which the restoring

    surface tension force is insignificant. On the other

    hand, at high Reynolds numb ers, the restoring force

    due to interfacial tension is predominan t compared to

    the viscous forces and the applied force essentially has

    to overcome the restoring interfacial tension forces.

    That is why at high Reynolds numb ers different lines

    are obtained for different W eber numb ers.

    B i n am p l a s t i c s

    For Bingh am plastic fluids the theoretical equations

    can be easily developed by replacing

    7

    in the differential

    equations for purely viscous fluids with 7 - 70)_ I t is

    easily shown that this corresponds to replacing

    [32 We (d _j D )s’3 - l]

    by

    [32 We (d _jD )5/3 - 4 7,, d-/a) -

    l]

    in equations of Table 3. This implies the usage of the

    same value of C determined from New tonian fluids.

    Figure 9 presents the experimental results of the

    maximum stable drop diam eter with change in the

    impeller speeds for a Bingham plastic fluid aqueous

    CaCOS suspension with PVA, c = 45.2 dynes/cm T,,

    = 75 dynes/cm’, n = 1, p,, = 13.7 cP). It is found that

    the drop diameter decreases with increase in the

    impeller speed in a nonlinear fashion. The solid line

    represents the theoretical results calculated from the

    first equation of Table 3 with the modification de-

    scribed above. The dashed line represents the drop

    diameters that would have been observed if the fluid

    had been Newtonian, i.e.

    7. =

    0. The experimental

    results are higher than those for a Newtonian fluid

    since the effective applied stress is now lower. The

    model reflects this and predicts the d , values quite

    well, thus accruing more evidence in its favour.

    Fig. 8. Effect of the Reynolds numb er on d /D with the Weber number as a parameter.

  • 8/19/2019 For coalescence

    8/8