applications of the differential balances to momentum transfer

Upload: suterian

Post on 02-Jun-2018

219 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    1/102

    A pp lica t io ns of t h e Dif ferent ia l B alan ces t o

    M o m e n t u m T r a n s f e r

    Let us now see w hat mu st be don e in order to arrive at detailed d escriptions for

    the configurations and motions of adjoining multiple phases within specified

    boundaries consistent with the principles we have developed in the preceding

    two chapters.

    In writing this chapter, we have assumed that you have already been in

    troduced to the more standard problems of fluid mechanics involving the flow

    of a single phase [42, Chap. 3; 99] and that you wish to learn more about

    analyzing multiphase flows when interfacial effects are important.

    3 . 1 P h i l o s o p h y

    3 .1 .1 S tru c ture o f Pr ob lem

    It is relatively easy to outline the issues that must be addressed in order to

    describe the configuration and motions of several adjoining phases.

    i)

    Within each phase,

    we must satisfy the differential mass balance, the

    differential momentum balance, and an appropriate model for bulk stress-

    deformation behav ior such as th e N ewto nian fluid [42, p. 42].

    ii) On each dividing surface, we must satisfy the jump mass balance and the

    jump momentum balance. Note that, as the result of the discussion in Sect.

    2.2.2, we will express the surface stress tensor simply in terms of an interfacial

    tension or energy.

    iii) At each common line,there are mass and mom entum balances to be obeyed

    (Sects. 1.3.8 and 2.1.9).

    iv)

    Everywhere,

    velocity must be finite.

    v)

    Within each phase,

    the stress tensor is bounded, since every second order

    tensor in three space is bounded [77, p. 140].

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    2/102

    160 3 Applications of the Differential Balances to Momentum Transfer

    vi)

    On each dividing surface,

    the surface stress tensor is bounded, by the same

    argument .

    vii) Within each phase in the neighborhood of a phase interface, it is important

    to correct for the intermolecular forces from the adjoining phase (see Sect. 2.2).

    In the case of a single interface, this may be a direct correction, although

    usually it will be done in terms of a surface tension or energy. In the case of

    thi n films, we comm only will introd uce a surface tension or energy a ttri bu tab le

    to unbounded adjoining phases with a further correction to account for the

    fact that at least one of the adjoining phases is not unbounded.

    viii) At each dividing surface, the tangential components of velocity are con

    tinuous. This is often referred to as the no-slip boundary condition. It is a

    consequence of assuming nearly local equilibrium at the phase interface (see

    Sect. 4.10.3). Lacking contrary experimental evidence, this condition is ap

    plied even when there is a phase change, the normal component of velocity

    is not continuous at the interface, and the adjacent phases axe clearly not

    in local equilibrium. A careful examination of the no-slip condition has been

    given by G oldste in [81, p. 676]. It is know n to fail, when o ne of th e pha ses is

    a rarefied gas.

    commo n lines on relatively rigid solids

    Prob lem s involving comm on

    lines on relatively rigid solids require special consideration, because we will

    normally decide to ignore the deformation of the solid within the immediate

    neighb orhoo d of th e comm on line. We can sum ma rize our discussions in Sects.

    2.1.10 and 2.1.13 with these additional comments.

    ix ) A common line can not move. As viewed on a macroscale, it appears

    to move relative to the solid as a succession of stationary common lines are

    formed on a microscale, driven by a negative disjoining pressure in the receding

    film (see Sect. 3.3.5). The amount of receding phase left stranded by the

    formation of this succession of common lines is too small to be easily detected,

    as in the experiments of Dussan V and Davis [16, see also Sect. 1.2.9].

    x)

    The m ass balance at a comm on line on a rigid solid is satisfied identically,

    since all velocities measured relative to the rigid solid go to zero at the common

    line.

    xi) The mom entum balance at a comm on line on a rigid solid reduces to

    Young's equation (Sect. 2.1.10)

    7^^^) cos 0 0 + 7^^^^

    - 1 ^ ^ ^

    = 0 (3.1.1-1)

    where 7^'^^ denotes the interfacial tension or energy acting in the

    i-j

    interface

    of Fig. 2.1.10-1 and OQ is the contact angle measured through phase A at the

    true common line (rather than the static or dynamic contact angle measured

    by experimentalists at some distance from the common line, perhaps with 10

    X magnification).

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    3/102

    3.1 Philosophy 161

    The preceding discussion assumes that one is willing to solve a complete

    problem. In the majority of applications, you will prefer to use a data corre

    lation to specify the static or dynamic contact angle at the apparent common

    line, the line at which the experimentalist measures the contact angle with

    perhaps 10x magnification. If th e ap par ent comm on line is in motion, it mu st

    be allowed to slip along the solid surface. Various empiricisms have been sug

    gested to account for this pheno me non [83, 85-8 7, 91 , 115, 183-185].

    3 . 1 . 2 A p p r o x i m a t i o n s

    Although it is relatively easy to outline as we have done in the preceding

    section the relationships and conditions that must be satisfied in describing

    any multiphase fiow, this is not suHicient, if we are to describe real physical

    phenomena. One must decide precisely what problem is to be solved and

    what approach is to be taken in seeking a solution. We think in terms of the

    approximations that must be introduced. Approximations are always made,

    when we describe real physical problems in mathematical terms.

    We can distinguish at least four classes of approximations.

    a) idealization of physical problem. Usually the physical problem with which

    we are concerned is too difficult to describe in detail. One answer is to replace

    it with a problem that has most of the important features of our original

    problem but that is sufficiently simple for us to analyze. Physical boundaries

    are replaced by geometric surfaces. Viscous effects may be neglected in a gas

    phase with respect to those in a liquid phase. The motion of boundaries is

    simplified. The disturbance caused by a tracer particle floating in an interface

    may be ignored. Bulk stress-deformation behavior may be described by the

    incompressible Newtonian fluid.

    b) limiting case. Even after such an idealization of our original physical prob

    lem, it may still be too difficult. We may wish to consider a limiting case in

    which one or more of the parameters appearing in our boundary-value prob

    lem approach zero. This is equivalent to asking for a solution correct to the

    zeroth perturbation in these parameters.

    c) integral average. In many cases, our requirements do not demand detailed

    solutions. Perhaps some type of integral average is of primary interest.

    d) mathematical approximation. Often we will introdu ce a m ath em atic al ap

    proximation in seeking a solution to a boundary-value problem. This is inher

    ent with numerical solutions.

    While we should attempt to hold approximations to a minimum, even

    greater effort should be devoted to recognizing those that we are forced to

    introduce. For it is in this way that we shall better understand the limitations

    of our theories in describing experimental data.

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    4/102

    162 3 Ap plication s of th e Differential Bala nces to M om en tum Transfer

    3 .2 On ly In t e r f a c i a l Te n s io n

    3 .2 .1 C las se s o f Pr ob lem s

    In this section, we will talk about phase interfaces in the absence of deforma

    tion. In all cases, there is a frame of reference with respect to which there is

    no motion anywhere within the system being considered. Usually this will be

    the laboratory frame of reference, but we will see that this is not necessary.

    Because of this lack of motion with respect to some preferred frame of

    reference, the interface and the adjoining phases axe static, but they should

    not necessarily be taken to be at equilibrium. Long after motion can be said

    to be absent from a system, its chemical composition may be slowly shifting

    as the various species present distribute themselves between the interface and

    the adjoining phases. You might expect this point to be made in Chap. 5,

    where we explicitly account for mass transfer. But in fact it is the rare ex

    periment in which surfactants, perhaps present only in the form of unwanted

    contaminants, have been eliminated.

    We can distinguish at least two classes of problems.

    In the simplest class, we neglect the effect of all external forces. Pressure is

    independent of posit ion within each phase and the jump momentum balance

    requires

    [ P ^ ] =

    2H^i

    (3.2.1-1)

    which implies that the mean curvature H is independent of position on the

    interface. These problems are discussed further in Exercise 3.2.1-1 and illus

    trated in the sections that follow.

    A second and often more difficult class of problems includes those in which

    the effect of one or more external forces can not be neglected. In Sect. 3.2.2,

    we consider the spinning drop measurement of interfacial tension in which the

    drop shape is controlled by the Coriolis force and the effect of gravity is ne

    glected. In Exercise 3.2.5-1, we discuss a rotating interface in which the effect

    both of gravity and of the Coriolis force play important roles in determining

    its shape.

    Exercise 3.2.1-1. Axially symm etric interface in absence of external forces [186]

    In Exercise

    A.5.3-11,

    we find that for an axially symmetric surface in cylindrical

    coordinates

    z = h{r) (3.2.1-2)

    the mean curvature H may be described by

    H=Yr[rh

    +

    h'+{h'f][l

    +

    {h'f]

    ^ ^ ^ '^'

    (3.2.1-3)

    -3 /2

    2'-dr- \^[i + (ft/)2]V2

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    5/102

    3.2 On ly Interfacial Tension 163

    where

    h' = ^

    (3.2.1-4)

    d r

    We confine our attention here to a static interface in the absence of gravity, in which

    case i f is a con stant . W e assume th at the surface coordina te system has been chosen

    such tha t

    H >0.

    Integrating (3.2.1-3), we find

    ^ ^ -,

    =Hr + Br-^

    (3.2.1-5)

    [l +

    ih'ff'

    Notice that a solution of (3.2.1-5) can exist only in an interval of

    r

    for which

    | i f r + S r - ^ | < l (3 .2 .1-6 )

    There aie three possibilit ies.

    i) If B > 0, prove that solutions exist only if

    B < ^ (3.2.1-7)

    Prove that there is a solution in the interval (r i ,T-2)

    0 < r - i < ^ < r 2 < - i : (3.2.1-8)

    where

    ri

    an d

    r2

    assumes all values in the indicated ranges as

    B

    varies from zero to

    i t s upper bound and

    as r -> n :

    h' ^ oo

    (3.2.1-9)

    as r ^ 1-2 : ft' ^ oo (3 .2.1 -10 )

    An example is i l lustra ted in Fig. 3.2.1-1.

    H

    F i g . 3 . 2 . 1 - 1 . Section of rotationally symmetric interface for B > 0

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    6/102

    164 3 Applications of the Differential Balances to Momentum Transfer

    ii) If S = 0, prove that the unique solution for (3.2.1-5) is a sphere of radius (in

    spherical coordinates)

    (3.2.1-11)

    iii) If B < 0, prove that there is a solution in the interval (ri,r2)

    0 < r-i < oo (3.2.1-12)

    4 < ^2 < oo (3.2.1-13)

    where ri and r2 increase through all values of their ranges as

    B

    decreases from 0 to

    oo and

    as r ^ r-i : h' - o o (3.2.1-14)

    as r ^ r2 :

    h' ^ oo

    (3.2.1-15)

    One case is shown in Fig. 3.2.1-2.

    r = 0 r i r2

    Fig. 3.2.1-2.

    Section of rotationally symmetric interface for S < 0

    3 .2 .2 Spin ning D ro p Interfac ia l Te ns io m eter [21]

    Althoug h not restricted to this l imit, only three experiments have been dem on

    strated as being suitable for measuring ultra low interfacial tensions (less than

    10 '^

    dyne / cm

    :

    th e spinning drop [187-191], th e microp end ant drop

    [192],

    and

    the sessile drop [193]).

    The spinning drop experiment sketched in Fig. 3.2.2-1 is perhaps the sim

    plest to construct and to operate. At steady state, the hghter drop phase {A),

    the heavier continuous phase

    {B),

    and the tube all rotate as rigid bodies with

    the same angular velocity Q.

    Although the analysis for the spinning drop presented by Princen et al.

    [188] an d by Ca yias et a l. [190] is correct, it is inconven ient t o use bec ause the ir

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    7/102

    3.2 Only Interfacial Tension 165

    ^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^

    y y ^ ^ ^ y y ^ ^ ^ y y y ^ ^ y y y ^ ^ y y y ^ ^ y y y ^ ^ ^ y y ^ ^ ^ y y ^ ^ -

    Fig. 3.2.2-1. At steady state, the drop phase A, the continuous phase B, and the

    tube all rotate as rigid bodies with the same angular velocity Q

    results axe given in terms of the radius of a sphere having the same volume

    as the drop. An experimentalist normally measures the maximum diameter

    (2rniax after correction for refraction) and the length (2/imax) of the drop.

    The revised solution that follows avoids this difficulty. It is more than a

    simple rearrangement of their result, since the radius of curvature of the drop

    at its axis of rotation (a in their notation) is treated in their development as

    thought it were known.

    Prom the differential momentum balance for each phase (i = A, B)

    (3.2.2-1)

    5()

    i .

    where V), is th e modified press ure of ph ase i on the axis of the cylinder. The

    (3.2.2-2)

    jump momentum balance requires

    where we have assumed that the Bond number

    AT,

    o

    (P'

    B) _ n(^ )

    )9rl

    7

    < 1

    (3.2.2-3)

    Here H is the mean curvature of the interface, 7 is the interfacial tension, and

    To

    is a char acte ristic len gth of th e sys tem t h a t will be defined shortly. If we

    describe the configuration of the interface by

    z = h{T)

    the mean curvature takes the form (Table

    B .

    1.2-7)

    -1 /2

    I

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    8/102

    166 3 Applications of the Differential Balances to Momentum Transfer

    It is convenient to introduce the change of variable [188]

    ta.n9 = - (3.2.2-6)

    d r

    in terms of which (3.2.2-5) becomes

    -2H

    = - - ^

    (r-

    sin6) (3.2.2-7)

    r dr

    Given (3.2.2-1) and (3.2.2-7), we can rearrange (3.2.2-2) as

    \ - ^ {r*sin6) = A- r*^ (3.2.2-8)

    where we have found it convenient to define

    * = IL

    ~ To

    (3.2.2-9)

    A=- (p^^ ) - P ^ ^ ) ) To (3.2 .2-10)

    and

    27

    I

    (3.2.2-11)

    Equation (3.2.2-8) can be integrated once consistent with the boundary con

    ditions

    (3.2.2-12)

    (3.2.2-13)

    (3.2.2-14)

    at r*

    and

    at r*

    to find

    sin^

    = 0 :

    = r*

    ' max

    =4

    61

    = 0

    : e=

    ^ * 3

    with the requirement

    A =

    2

    ' m a x

    r* 2

    ' max

    2

    n

    2

    (3.2.2-15)

    We can now visualize integrating (3.2.2-6) to learn

    h*m^= '

    1/2^^*

    (3.2.2-16)

    Jo ( l - s i n 2 6 l)^

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    9/102

    3.2 Only Interfacial Tension 167

    in which sin

    0

    is given by (3.2.2-14) and

    h*

    is defined by analogy with (3.2.2-9).

    With the change of variable

    g ^ - - (3.2.2-17)

    (3.2.2-16) can be written as

    rA/2

    K.^ =

    / ^

    - ^ dq

    (3.2.2-18)

    91

    [{q - qi) [q -

    ^2)

    [q - qs)]

    with

    qi = ^

    (3.2.2-19)

    ' m a x

    r* 2 1 / r * * V^^

    92 = ^ + 2

    [ - ^ +

    ^ m a. j (3.2.2-20)

    93 = ^ - 2 [f^ +^ m a. j (3.2.2-21)

    Let us consider two limiting cases. For a cylindrical drop

    sin6l = l (3.2.2-22)

    and it follows immediately from (3.2.2-8) and (3.2.2-15) that

    < a x = 2 ^ / ' (3.2.2-23)

    For a spherical drop whose dimensionless radius is

    R*

    r*

    s i n 6 l = - - (3.2.2-24)

    R*

    and we conclude from (3.2.2-8) and (3.2.2-15) that

    < a x = 0 (3.2.2-25)

    This suggests that we confine our attention to the range 0 < r^^x ^ 2^/^

    for which

    qi>qi> qa (3.2.2-26)

    With this restriction, (3.2.2-18) can be integrated to find [194, p. 78].

    2

    h* =-

    ' max , .1 /2

    [qi -

    93)

    qiF {o, k) - {qi - qs) E {^o, k)

    +

    (91 - 93) (t an(t>o)

    ya - k^

    sin^ 0o

    3.2.2-27)

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    10/102

    168 3 Applications of the Differential Balances to Mom entum Transfer

    in which

    E

    {(j)o,

    k)= f Vl-k'^sm'^^d^

    Jo

    Fic^o.k)^

    /

    Jo

    v l - ^ si

    sin^

    ^

    3.2.2-28)

    3.2.2-29)

    (3.2.2-30)

    (3.2.2-31)

    qi -qs

    Following a suggestion of Princen et al.

    [188],

    the volume of the drop

    0o = arcsin

    . M-2 , ,^^ /^

    A-2q2

    k

    2 _ q2-q3

    becomes

    V^

    V

    +

    /?* - 1

    3.2.2-32)

    We computed rmax/^max = r^^^JKi^ from (3.2.2-27) an d V from

    (3.2.2-32) as functions oir'^^^. In these com puta tions , the incom plete elliptic

    integrals (3.2.2-28) and (3.2.2-29) were expanded in series [195, p. 300]. The

    results are presented in Table 3.2.2-1 and Fig. 3.2.2-2 in a form convenient for

    ex pe rim en tal ists : r^g^^ ^^ ^ ^ * ^^ func tions of rmax/Z^max-

    ' max max

    0.4 0.6

    1.259

    1.250

    1.255

    0.25

    ^ m a x / ' ^ m a x

    Fig. 3.2.2-2.

    rJiax as a function of rmax/^max from (3.2.2-27)

    0.35

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    11/102

    3.2 Only Interfacial Tension 169

    Table 3.2.2-1.r^a^ andV*as function of rmax/Vax from (3.2.2-27)

    and (3.2.2-32)^

    1 '

    -msn-v

    h

    ^

    h

    1.0

    0 0

    0.3198 1.252 30.6741

    0.9997 0.1 0.0042 0.3122 1.253 31.6546

    0.9980 0.2 0.0336 0.3038 1.254 32.7883

    0.9932 0.3 0.1139 0.2945 1.255 34.1321

    0.9840 0.4 0.2725 0.2837 1.256 35.7818

    0.9687 0.5 0.5406 0.2708 1.257 37.9195

    0.9459 0.6 0.9571 0.2543 1.258 40.9617

    0.9140 0.7 1.5745 0.2297 1.259 46.2969

    0.8710 0.8 2.4714 0.2262

    1.2591

    47.1309

    0.8148 0.9 3.7782 0.2225 1.2592 48.0733

    0.7415 1 0 5.7457 0.2183 1.2593 49.1567

    0.6432 1.1 8.9812 0.2136 1.2594 50.4307

    0.4928

    1 2

    15.9687 0.2081

    1.2595

    51.9769

    0.3332 1.25 29.0379 0.2016 1.2596 53.9444

    0.3268 1.25 29.8101 0.1932

    1.2597

    56.6522

    In order to use the spinning drop technique to determine interfacial ten

    sion, one must measure both r-max and /imaxfo rgiven values of (p^^^ p^ ^^)

    and /?. Having rmax/Zimax, one determines r^^x from Table 3.2.2-1 or Fig.

    3.2.2-2. When rmax//imax< 0.25, (3.2.2-23) can be used with less than 0.4

    error in the interfacial tension. The interfacial tension can then be found by

    rearranging the definition for r^ax-

    ^ ^ 1 / r ^ A UB)_p(A)\^2 (3.2.2-33)

    2 \' max/

    Princen et al. [188] and Cayias et al. [190] expressed their results as func

    tions of a dimensionless shape parameter a, which here becomes

    a = (^j^ (3.2.2-34)

    While their results can not be readily rearranged in the form given here, the

    relation betweenaand rmax/Zimax calculated from (3.2.2-34) and Table 3.2.2-1

    agrees with that given by Princen et al. [188].

    There are several cautions to be observed in using this technique.

    ^This table,aminor revision of that presented by Slattery and Chen [21], is

    due to Chen

    [196],

    who gives the computer program as well as a discussion of the

    truncation errors.

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    12/102

    170 3 Applications of the Differential Balances to Momentum Transfer

    1) Prom (3.2.2-3) and (3.2.2-11),

    2 V 3 ( p ( B ) _ p ( A ) ) i / 3 g

    < 1 (3.2.2-35)

    This implies that the angular velocity

    O

    should be as large as possible. Unfor

    tunately, so long as (p^^^ p^^^) i 0, we can approach the rigid body motion

    assumed here only asymptotically as

    N-QO

    ^ 0 or as /2> OO. For any finite

    value of i7, the axis of the drop or bubble will be slightly displaced from the

    axis of the tube, resulting in convective mixing within the two phases

    [191].

    In addition to negating rigid body motion, these convection patterns have the

    potential of creating interfacial tension gradients in the interface leading to a

    loss of stability, if the re is a surfac tant pre sen t.

    If

    fl

    is too large, the drop or bubble phase will contact the ends of the

    t ube ,

    introducing end effects not taken into account in this analysis.

    It may be possible to find a compromise by studying 7 as a function of Q

    [197].

    2) Let us assume that the phases are equilibrated before they are introduced

    into the apparatus . As fl begins to increase towards its final value, the drop

    lengthens and the area of the drop surface increases. If there is no surfactant

    present, the system reaches a steady state rapidly. If there is a surfactant

    present, the system approaches a steady state more slowly: as surfactant is

    adsorbed in the interface, the interfacial tension falls, the interfacial area of

    the drop increases allowing more surfactant to come into the interface,

    The problem is obvious, when the phases are not pre-equilibrated [198,

    199]. But it could also be a serious problem, if the primary source or surfac

    tant is a dilute solution forming the drop phase. If sufficient surfactant were

    removed from the drop phase by adsorption in the expanded interface, the

    equilibrium compositions of the two phases could be changed.

    These may have been the conditions studied by Kovitz

    [200].

    The drop

    phases were dilute solutions of surfactant in octane and the continuous phases

    were aqueous solutions of NaCl. After pre-equilibration, the drop phase re

    mained the primary source of surfactant. He observed that days or weeks

    might be required to achieve an equilibrium interfacial tension with the spin

    ning drop experiment compared with minutes or hours in static experiments

    such as the sessile drop and pendant drop. Frequently an equilibrium in

    terfacial tension could never be measured with the spinning drop interfacial

    tensiometer, either because the drop disappeared entirely or because the drop

    phase evolved into a shape that was clearly not amenable to data reduction.

    (This may have been the result of buoyancy-induced convection becoming

    more important as the single-phase region was approached on the phase di

    agram

    [191].

    Lee and Tadros [201] observed interfacial instabilities leading

    to disintegration of the interface and the formation of droplets that they

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    13/102

    3.2 Only Interfacial Tension 171

    attributed to interfacial turbulence.) When an equilibrium interfacial tension

    was measured, it often differed by an order of magnitude from the values found

    with static experiments [meniscal breakoff interfacial tensiometer (Sect. 3.2.3),

    du Noiiy ring (Exercise 3.2.3-2), sessile drop (Sect. 3.2.5)]. In contrast, for sin

    gle component systems he found excellent agreement between measurements

    made with the spinning drop and those made with static experiments.

    3) Vib rations shou ld be carefully redu ced, since the y prom ote the developm ent

    of instabilities

    [191].

    4) The spinning tube should be horizontal, in order to prevent migration of

    the bubble or drop to one of the ends.

    In spite of all of these potential problems, the spinning drop technique has

    been used successfully in the measurement of interfacial (and surface) tensions

    having a wide range of values.

    Wiistneck and Warnheim [202] report excellent agreement between mea

    surements made with a du Noiiy ring (Exercise 3.2.3-2) and those made with

    a spinning drop for an aqueous sodium dodecylsulphate solution against air

    (their Fig. 2). Their surfactant was confined to the continuous phase, elimi

    nating any significant change in the equilibrium phase compositions.

    In the experiments of Ryden and Albertsson [189] and Torza

    [203], signif

    icant amounts of surfactant were not present.

    3 .2 .3 Menisca l Breakoff Interfac ia l Tens iometer

    Figure 3.2.3-1 shows two geometries that may prove to be the basis for a useful

    technique for measuring interfacial tension. A circular knife-edge is positioned

    so as to just touch the A-B pha se interface. In Fig. 3.2. 3-l(a) , the general level

    of th e A-B interface is slowly lowered un til it bre aks away from th e kn ife-edge;

    in Fig. 3.2.3-l(b), theA-B interfac e is slowly raise d unt il breakoff

    occurs.

    From

    the measured value

    oi H =

    -ffmax at

    breakoff,

    the interfacial tension can be

    calculated.

    Let us describe the configuration of th e interface in cylindrical co ordina tes

    by

    r = c{z) (3.2.3-1)

    From either the r- or z-component of the jump momentum balance at the

    A-B ph ase interface, we have (Table B.1.2-8)

    d^c / dc

    1

    c

    dz^

    Vd : i + ' d z y

    d c V '

    (3.2.3-2)

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    14/102

    172 3 Applications of the Differential Balances to Momentum Transfer

    (a)

    gravity

    z = 0

    (b)

    z = -H

    z = 0

    gravity

    Fig. 3.2.3-1.

    Meniscal breakoff interfacial tensiometer. (a) When lower phase

    pref

    erentially wets the knife-edge, the A-B interface is slowly lowered un til breakoff

    occurs, (b) When upper phase preferentially wets the knife-edge, the

    A-B

    interface

    is slowly raised until breakoff occurs

    Here

    po

    is th e pressure within phase

    A

    as the interface is approached in a

    region sufficiently iai away from the knife-edge that the curvature of the inter

    face is zero;po is the pressure with in pha se B as the interface is approached

    in the same region. Equation (3.2.3-2) requires that sufficiently far away from

    the knife-edge

    at z = 0

    Pi^^ Pi^^

    (3.2.3-3)

    Equation (3.2.3-2) can be more conveniently written in terms of the dimen-

    sionless variables

    -1/2

    (3.2.3-4)

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    15/102

    -1/2

    as

    {piA)-p{B))g

    d^c*

    f dc

    3.2 Only Interfacial Tension 173

    (3.2.3-5)

    z =

    dz* dz*

    - 1

    1 +

    d^

    dz*

    -3/2

    (3.2.3-6)

    If our objective is to describe the elevated meniscus shown in Fig.

    3.2.3-1

    (a),

    we must solve (3.2.3-6) consistent with the boundary conditions

    as z* ^ 0 : c*> oo

    at

    z* = H* : c* = R*

    (3.2.3-7)

    (3.2.3-8)

    In order to describe the depressed meniscus pictured in Fig.

    3.2.3-1

    (b),

    we

    must find a solution for (3.2.3-6) consistent with (3.2.3-7) and

    at

    z* -H*

    R* (3.2.3-9)

    Notice that the elevated meniscus problem takes the same form as that for

    the depressed meniscus, when

    z

    is replaced by

    z

    and

    [p^^ p^^^)

    by

    [p^^^ p^^^)- For this reason, we will confine our atte nti on for th e m om ent

    to the elevated meniscus.

    Bo th Pa dd ay a nd P it t [204] and K ovitz [205] have considered this p roblem .

    Selected members of the family of meniscus profiles as calculated by Kovitz

    [205] axe shown in Fig. 3.2.3-2; they are identified by the value of a parameter

    Q introduced in his numerical solution. For a selected value of R* (a given

    knife-edge diameter and chemical system), the interface can assume a contin

    uous sequence of configurations corresponding to increasing the elevation

    H*

    from 0 to some maximum value -ffmax) beyond which static solutions do not

    exist and the interface ruptures.

    A sequence of such static configurations for an air-water interface is pre

    sented in Fig. 3.2.3-3. As an example. Fig. 3.2.3-4 shows that the profile pic

    tured in Fig. 3.2.3-3(f) compares well with Kovitz's numerical computations.

    Figure 3.2.3-2 indicates the possibility of two different configurations pass

    ing through a given knife-edge denoted by a point (i?*,

    h*).

    One of these two

    configurations always touches the envelope i^^ax before contacting the knife-

    edge.

    Kovitz [205] has never observed such a configuration in his experiments

    and believes it to be unstable.

    Kovitz [205] provides analytical expressions for various portions of the

    envelope curve H^^ = H^^^ (R*) in Fig. 3.2.3-2. For large R*,

    H*^.^

    = 2-IR*-'-IR*-^ +

    0{R*-^

    (3.2.3-10)

    which may be used with less than 1% error for

    R* >

    1.9. For 0.2

    < R*

    i^max?

    no

    solutions exist

    and the

    interface

    ruptures

    D

    n = l

    where

    C i = 0.18080

    Ci = 0.29211

    C2 = 0.84317

    C5 = - 0 . 0 6 2 3 3

    C3 = - 0 . 6 9 4 9 4

    C s

    =

    0.00532

    For sufBciently small

    R*,

    R*

    =

    ^e-a+ (1

    _

    coscj,)-^

    i^iax = 4 e - ( ^ - ) e - c ^ ( l ^ ) t a n ^

    \

    1 COS(p J

    in which

    a = 0 . 5 7 7 2 1 5 7 -

    (3.2.3-11)

    (3.2.3-12)

    (3.2.3-13)

    (3.2.3-14)

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    17/102

    3.2 On ly Interfac ial Ten sion 175

    F i g . 3 . 2 . 3 - 3 . P ho to gr ap hs by Kov itz [205] of an air -w at er interface for i?* = 1.12.

    (a)

    H*

    = 0.24. Th e knife-edge is below the ai r-w ate r interface, which in tersects

    the rod with a f inite contact angle, (b)

    H^

    = 0.24. Th e air -w ate r interface interse cts

    the knife-edge. The contact angle boundary condition is removed, (c)

    H*

    = 1.20.

    (d)

    H* =

    1.25. (e) if* = 1.30. (f) if* = 1.36. See th e com pari son w ith the or y in

    Fig. 3.2.3-4. Breakoff occurred at if* = 1.38

    i s t h e E u le r c ons t a n t a n d c o t

    (j)

    = d c * /dz * . E q ua t io n ( 3 .2 .3 -13 ) c o r r e c t s K o v i t z

    [205] Eq . (46) . E qu a t io n (3 .2 .3-13) m us t be so lved s im ul t an eo us ly for va r iou s

    values of

    0 for

    which only one metastable drop configuration exists (G = 3).

    H*-G:

    -1.05

    -0.3-5 0.10

    G

    I G

    =1.90

    :0.5

    2.35

    Fig. 3.2.4-4. Sequence of pendant drops for R* = 1.18 asH* increases [220]

    Limiting cases of the inner and outer envelopes in Fig. 3.2.4^2 are derived

    by Kovitz

    [220],

    by Hida and Miura

    [221],

    and by Hida and Nakanishi

    [222].

    There are two ways in which the pendant drop or emerging bubble can be

    used to measure interfacial tension.

    The preferred way of measuring interfacial tension with the pendant drop

    is to com pare the experim entally observed drop configuration w ith those pre

    dicted theoretically. Tabular numerical solutions of (3.2.4-3) consistent with

    (3.2.4-5) and (3.2.4-6) have been presented by a variety of workers. The most

    recent tables are those prepared by Hartland and Hartley [223] and by Padday

    [224;

    these tables are available on microfiches from Kodak Limited, Research

    Laboratories, Wealdstone, Harrow, Middlesex, England; see also 212]. Padday

    [225,

    pp . I l l and 152] summ arizes with comm ents the earlier tables p repared

    by Andreas et al .

    [226],

    by Niederhauser and Bartell [227, p. 114], by Ford-

    ham

    [228],

    by Mills

    [229],

    and by Stauffer

    [230].

    In the tables of Hartland and

    Hartley

    [223],

    B

    G

    (3.2.4-9)

    The tables of Padday and Pitt [204, 212] are based upon a rearrangement of

    (3.2.4-3)

    2+

    /3z**

    =

    2 * *

    C

    dz**^

    +

    d^

    dz*

    + 1 1 +

    dc^

    dz*

    - 3 / 2

    (3.2.4-10)

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    29/102

    3.2 Only Interfacial Tension 187

    where

    b

    is th e radiu s of curv ature at th e ape x of the drop and

    (3.2.4-11)

    b

    z

    0-

    b2(^piB) _ piA)> g

    (3.2.4-12)

    (3.2.4-13)

    In arriving at (3.2.4-10), we have noted that at

    z** =

    0

    -2Hb =

    c*

    2

    G

    2**

    d^c

    c

    +

    z**^ \ dz

    dc*

    + 1 1 +

    dc^

    dz*

    -3/2

    {piA)_p{B)^g

    - 1 / 2

    (3.2.4-14)

    From (3.2.4-13) and (3.2.4-14), we see th at the sha pe pa ram ete r /3 is negative

    for a pendant drop and simply related to G:

    0

    4

    " G 2

    (3.2.4-15)

    Perhaps the simplest technique is to slowly increase the volume of the drop

    until rupture occurs at the intersection of c* = R* with the outer envelope.

    Unfortunately, the volume of this drop is not identically equal to the volume

    of the drop which breaks away and which would be measured experimentally.

    A correc tion m ust b e employed P ad da y [225, p . 131]. Th is is referred t o as

    th e drop we ight t echn ique , since experim entally one would mea sure the

    weight of the drop which breaks away.

    The pendant drop and emerging bubble shown in Fig. 3.2.4-1 may be

    thought of as radius-limited, in the sense that the drop configurations must

    pass through the knife-edge tip of the capillary. Special cases of this problem

    are the contact-angle-limited pendant drop and emerging bubble shown in Fig.

    3.2.4-5. Equation (3.2.4-3) must again be solved consistent with boundary

    conditions (3.2.4-5) and (3.2.4-6), but boundary condition (3.2.4-7) is now

    replaced by a specification of the contact angle at z* = H*. Figure 3.2.4-2

    again applies. Given the contact angle between the meniscus and the solid

    wall from which the drop is suspended. Fig. 3.2.4-2 can be used to determine

    the configuration of the drop for a specified value

    H*

    G.

    Notice that for a

    given value of H* G, only one solution is possible. Furthermore, the drop

    phase must wet the wall. Multiple solutions are possible as H*G varies (as

    the volume of the drop varies).

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    30/102

    188 3 Applications of the Differential Balances to Momentum Transfer

    (a)

    z = H

    z H A

    gravity

    ^ +

    (b)

    gravity

    z = -H

    Fig. 3.2.4-5. (a) Contact-angle-limited pendant drop in which the density of the

    drop is greater than that of the continuous phase, (b) Contact-angle-limited emerg

    ing bubble in which the density of the bubble is less than that of the continuous

    phase

    As in the previous section, the radius of the knife-edge tip of the capil

    lary is presumed here to be sufHciently small that the outer solution (outside

    the immediate neighborhood of the common line where the effects of mutual

    forces such as London-van der Waals forces can be ignored; see Sect. 3.1.1) is

    independent of the static contact angle. It is only necessary that the bubble

    or drop wet the capillary.

    3 .2 .5 Sess i le Drop

    Th e captive bu bble a nd sessile drop are shown in Fig.

    3.2.5-1.

    They are directly

    analogous with the contact angle l imited pendant drop and emerging bubble

    sketched in Fig. 3.2.4-5. The only difference is that the densities of the two

    phases are interchanged. As with the pendant drop and emerging bubble,

    interfacial tension can be inferred by comparing their measured profiles with

    those expected experimentally.

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    31/102

    3.2 On ly Interfac ial Ten sion 189

    (a)

    z = H

    gravity

    (b)

    z = -H

    gravity

    F ig . 3 . 2 .5 - 1 . ( a ) Captive bubble in which densi ty of the drop is less than that of

    the continuous phase, (b) Sessile drop in which density of the drop is greater than

    that of the continuous phase

    Le t us p roceed as we d id in Sec t . 3 .2 .4 to desc r ibe the conf igura t ion of

    th e i n t e r f a c e i n c y l ind r i c a l c oo r d ina t e s . Ra th e r t h a n ( 3 .2 .4 - 3 ) , w e ha ve f r om

    e i t h e r t h e

    r-

    o r

    z-

    c o m p o n e n t o f t h e j u m p m o m e n t u m b a l a n c e a t t h e

    A-B

    pha se i n t e r f a c e

    -G -

    * 1

    z =

    c*

    where

    G =

    (.A) {1

    PH -PH

    ' - (P'

    1+

    dc^

    dz^

    -\

    - 3 / 2

    ,(^)-p(^))5^][7(p()-^(^))5]

    - 1 / 2

    7

    1 - 1 / 2

    7

    (p(B)-piA))g

    - 1 - 1 / 2

    (3.2.5-1)

    (3.2.5-2)

    (3.2.5-3)

    (3.2.5-4)

    I f w e w i sh t o de sc r ibe t he c on f igu r a t i on o f t he c a p t ive bubb le show n in F ig .

    3 .2 .5 - l ( a ) , w e m us t so lve ( 3 .2 .5 -1 ) c ons i s t e n t w i th

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    32/102

    190 3 Applications of the Differential Balances to Momentum Transfer

    at z* = 0 : c* = 0 (3.2.5-5)

    0 : ^ ^ 00 (3.2.5-6)

    s z

    at

    z* = H*

    c* = 0

    dc*

    dz*

    dc*

    dz*

    = cot 0('*'**) (3.2.5-7)

    in which 0(s*^*) is the static contact angle measured through phase

    A.

    In order to describe the sessile drop sketched in Fig. 3.2.5-1(b), we must

    determine a solution for (3.2.5-1) consistent with (3.2.5-5), (3.2.5-6) and

    at z*

    -H* :

    dc*

    dz*

    -cot6('*^*)

    (3.2.5-8)

    The sessile drop problem takes the same form as that for the captive bubble

    when z is replaced by z and

    [p^^

    p^ ^^) by

    [p^-^^ p^^^),

    which mea ns

    that it is not necessary to give separate discussions for these two problems.

    In the limit of the large flat captive bubble shown in Fig. 3.2.5-2, (3.2.5-1)

    simplifies to

    - G - z *

    d V

    d ^

    1+

    /dc*Y

    -1 - 3 / 2

    (3.2.5-9)

    If the bubble does not wet the solid surface upon which it rests, then there

    z = 0

    B

    Fig. 3.2.5-2. Large flat captive bubble that does not wet the solid

    is an axial position

    h

    such that

    dc*

    at z* =

    h*

    dz

    = 0

    Since the bubble is relatively flat at its apex, we will require

    dc*

    dz*

    d^c*

    as z

    0

    oo

    dz*2

    remains finite

    (3.2.5-10)

    (3.2.5-11)

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    33/102

    3.2 Only Interfacial Tension 191

    This permits us to conclude from (3.2.5-9) that

    G = 0 (3.2.5-12)

    Finally, we can integrate (3.2.5-9) consistent with (3.2.5-10) and (3.2.5-11) to

    find

    - 3 / 2

    r--/:-(sr

    (s)

    P^^^)9

    (3.2.5-13)

    (3.2.5-14)

    While this is a simple expression, it is relatively difficult to locate the position

    h

    of the maximum diameter.

    For greater accuracy, it is preferable to consider a smaller bubble and to

    compare the experimental and theoretical profiles at several axial positions.

    In these cases, we must rely upon the tabular numerical solutions of (3.2.5-1)

    consistent with (3.2.5-5) through (3.2.5-7) which are available. The most re

    cent tables are those prepare d by Ha rtla nd and H artley [223; see also footnote

    a of Sect. 3.2.4] and by Padday^ [224; these tables are available on microfiches

    from Kodak Limited, Research Laboratories, Wealdstone, Harrow, Middle

    sex, England; see also 212]. Padday [225, pp. 106 and 151] summarizes with

    comments the earlier tables prepared by Bashforth and Adams [231] and by

    Blaisdell

    [232].

    The contact angle limited captive bubble and sessile drop are

    stable in all configurations as are the analogous contact angle limited pendant

    drop and emerging bubble discussed in Sect. 3.2.4.

    ^These tables were prepared with (3.2.5-1) in the form of (3.2.4-10). Note now

    however that

    at

    z**

    = Q:

    -2Hb =

    = 2

    = G

    , d_c_

    dz**^

    H)'"'lh()

    - 3 / 2

    -1-1/2

    (p (S ) -p (A) )3 j

    which implies

    4

    P =

    G 2

    (3.2.5-15)

    (3.2.5-16)

    The shape factor /? is positive for the captive bubble-sessile drop but negative for

    the pendant drop-emerging bubble.

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    34/102

    192 3 Applications of the Differential Balances to Momentum Transfer

    Besides their use in determining interfacial tension, the contact angle-

    limited captive bub ble and sessile drop ca n also be used to measu re the con tact

    angle

    [150].

    The contact angle limited captive bubble and sessile drop in Fig. 3.2.5-1

    may be thought of as special cases of the radius-limited captive bubble and

    sessile drop shown in Fig. 3.2.5-3. Equation (3.2.5-1) is again to be solved

    consistent with boundary conditions (3.2.5-8) and (3.2.5-9), but boundary

    condition (3.2.5-10) is now to be replace by a specification of the bubble

    radius

    ai z = H.

    The radius limited captive bubble is analogous to the radius

    limited pendant drop in the sense that not all configurations satisfying these

    bo un da ry conditions are stab le [145, 204].

    (a)

    (b)

    z = H

    gravity

    gravity

    z = -H

    Fig. 3.2.5-3. (a) Radius limited captive bubble in which density of bubble is less

    than that of continuous phase, (b) Radius limited sessile drop in which density of

    drop is greater than that of continuous phase

    Finally, note that the analysis presented here is for the outer solution

    (outside the immediate neighborhood of the common fine, where the effects of

    mutual forces such as London-van der Waals forces and electrostatic double-

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    35/102

    3.2 On ly Interfacial Tension 193

    laye r forces can be ignored ; see Sec t . 3 .1 .1) . The s ta t ic contac t ang le i s one

    t h a t a n e x p e r i m e n t a l i s t m i g h t m e a s u r e w i t h p e r h a p s 1 0 x m a g n i f i c a t io n .

    E x e r c i s e 3 . 2 . 5 - 1 .

    Rotating menisci

    Fi gu re 3.2.5-4 shows an interface formed in

    a vertical tube of radius

    R

    that rota tes with a constant angular veloci ty

    O.

    T he

    contac t ang le measured th rough phase

    B

    is

    0.

    z = 0

    F i g . 3 . 2 . 5 - 4 .

    Interface formed between phases

    A

    and B in a ver t ica l tube that

    rota tes with a constant angular veloci ty

    f2

    Assume that the configuration of the interface in cylindrical coordinates is de

    scribed by z =

    h{r).

    D e te r mine t ha t

    h

    can be found by satisfying

    1

    dr*^ ^ dr* ^ \dr* J ^ \dr* J

    - 3 /2

    ( ^ ( B ) _ ^ ( A )

    ^

    I h*h^-^^r^^^+C

    (3.2.5-17)

    subject to the boundary condit ions

    at r* = 0 :

    h* = 0,

    ^ = 0

    '

    dr*

    at r* = 1

    Here

    * _ r

    ' =R'

    dh*

    dr*

    ( s i n 0 )

    R'

    ( P ^ ^ ' - P ^ )

    (3.2.5-18)

    (3.2.5-19)

    (3.2.5-20)

    Here pg

    '

    is th e press ure as (2: = 0, r = 0) is app roac hed wit hin pha se

    A,

    and pg

    '

    is th e pressure as this sam e point is approach ed w ithin ph ase

    B.

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    36/102

    194 3 Ap plication s of th e Differential Balan ces to M om en tum Transfer

    This problem was first solved by Wasserman and Slattery [233]. Fur the r numer

    ica l resul ts have been presented by Pr incen and Aronson [234], a portion of which

    are shown in Fig. 3.2.5-5.

    2.4

    2.0

    1.6

    1.2

    .8

    .4

    /i^ 0

    .4

    .8

    1.2

    1.6

    9 n

    1 1 1 1

    Q _ ^ ^ . > - < ^

    2^^

    J\

    i Q - ^ / /

    3 J / /

    1 1 1 1

    1 .8 .6 .4 .2 0 .2 .4 .6 .8 1

    F i g . 3 . 2 . 5 - 5 .

    Computed profiles of rotating menisci from Princen and Aronson [234]

    for 6> = 0,

    [p^^^ - p^^A gE?h = 1-3389,

    and various values of

    Q

    {R/gf^'^

    3 .3 A p p l i c a t i o n s o f O u r E x t e n s i o n of C o n t i n u u m

    M e c h a n i c s t o t h e N a n o s c a l e

    I n S e c t . 2 . 2 , w e de ve lope d a n e x t e ns ion o f c on t inuum me c ha n ic s t o t he

    na n osc a l e ( no t t he mo le c u la r s c a l e ). By na nosc a l e , w e m e a n th e im m e d ia t e

    ne i gh bo r ho od o f a ph a se in t e r f a c e . O ur pa r t i c u l a r c onc e r n is w i th t hose p r o b

    lem s w he re i t i s no t suff ic ien t to cor re c t for in te rm ole cu la r fo rces f rom an

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    37/102

    3.3 Applications of Our Extension of Continuum Mechanics to the Nanoscale 195

    adjacent phase with an interfacial tension or energy as we did in the prior

    section.

    3 . 3 . 1 S u p e r c r i t i c a l A d s o r p t i o n [ 2 2 ]

    Far from a solid-fluid interface on a molecular scale, the density of a gas,

    either subcritical or supercritical, approaches the bulk density. Within a few

    molecular diameters of the interface, the gas is subjected to intense attractive

    intermolecular forces attributable to the solid. The magnitudes of these forces

    increase as the distance to the interface decreases. With a subcritical gas or

    vapor, condensation occurs. With a supercritical fluid, the density increases

    as the interface is approached, becoming similar to that of a liquid. We will

    confine our attention here to the case of a supercritical fluid. We explore the

    adsorption or densification of thr ee supercritical gases on Gra pho n: argon,

    krypton, and methane.

    Specovius and Findenegg [235] used a gravimetric method for the deter

    mination of surface excess isotherms of supercritical argon and supercritical

    methane on Graphon (a graphitized carbon black). Blumel et al. [236] did a

    similar study for supercritical krypton on Graphon.

    Five analyses of these data starting from a molecular point of view have

    been published.

    Egorov [237] pre sen ted a microscop ic sta tisti ca l me chan ical the ory of ad

    sorption of supercritical fluids. The theory is in excellent agreement with

    experimental observations of Specovius and Findenegg [235] and Blumel

    et al. [236] for argon and krypton, although the error increases at higher

    densit ies as the temperature approaches the cri t ical temperature.

    Ra ng araja n et al. [238] developed a mean-field mod el th at superim poses

    the fluid-solid potential on a fluid equation of state to predict adsorption

    on a flat wall. Their paper shows some predicted adsorption isotherms for

    krypton but none for argon or methane.

    Sokolowski [239] used the Percus-Y evick e qua tion for th e local den sity of

    a gas in contact with a flat solid surface to calculate the adsorption char

    acteristics of argon and of methane on graphite. The calculations used the

    Boltzmann-averaged potential

    [240],

    which is calculated from the pa rtic le-

    graphite basal-plane potential

    [241],

    for the gas-surface interaction and

    used the Lennard-Jones (12,6) potential for the gas-gas interactions.

    Fischer [242] used a mod el for high tem pe rat ur e and high pressure adsorp

    tion t h at was the sam e as the one introduced in previous pa pers [243, 244]

    for moderate pressures. He assumed that a hard sphere gas was in con

    tact with a structureless plane wall and that a Lennard-Jones potential

    described the wall-particle interaction.

    Aranovich and Donohue [245] used two adjustab le para m ete rs in com par

    ing the Ono and Kondo [49] theory to the data reported by Specovius and

    Findenegg

    [235].

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    38/102

    196 3 Applications of the Differential Balances to Momentum Transfer

    The first four of these analyses will be discussed in more detail later.

    There have been a number of studies of supercritical adsorption on acti

    vated carbo ns, both theore tical and experim ental [246-253]. W hile activate d

    carbons are of immense practical importance, they are not well-suited to test

    theories, because they are porous and the effects of intermolecular forces are

    more complex. For this reason, they will not be discussed here.

    In Sect. 2.2, we introduced a new extension of continuum mechanics to

    the nanoscale. In Sect. 2.2.2 we tested this theory in predicting the surface

    tension for the n-alkanes using one adjustable parameter. What follows is the

    first test of the theory using no adjustable parameters.

    Referring to Fig. 2.2.1 -l(a), let phase A be a static sup ercritical fluid (/ )

    and phase B an impermeable crystalline carbon solid (s). Our objective is to

    determine the density distribution within the static fluid and therefore the

    apparent adsorption of the fluid on the crystalline carbon solid.

    This analysis will be done in the context of view (iv) of Sect.

    2.2.1,

    but ,

    because we are interested in only the gas phase, we will have no need for the

    jump momentum balance. Start ing with the differential momentum balance

    (2.2.1-1),

    we will neglect any effects of gravity to write for the fluid

    _ V P ( / )

    +

    b(/'

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    39/102

    3.3 Applications of Our Extension of Continuum Mechanics to the Nanoscale 197

    The Lennard-Jones (6-12) potential (2.2.0-1) is commonly recommended

    for non-polar dilute gases. In this case,

    f( /. /) = - V (n(^)'

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    40/102

    198 3 Applications of the Differential Balances to Momentum Transfer

    Here (dP^-'^^/dc^-''))^ will be ev alua ted using a n em pirical eq uatio ns of stat e for

    argon

    [257],

    for krypton

    [258],

    an d for m et ha ne [259, 260]; AT is A vog adro 's

    number;

    c^^' = n'^^'/N

    is th e molar density in th e interfacial region. The

    solution of this differential equation with the boundary condition

    a s 2 > 00 :

    c(/) ^ c^f:^-^^) (3.3.1-9)

    gives a density distribution.

    The surface excess or the apparent adsorption

    r ( - )

    = r

    (c (/) - c(/' '' ^))

    Az

    (3.3.1-10)

    Since

    5

    is not a known para m eter , it m ust be defined. Prom (3.3.1-4), (3.3.1-6),

    and (3.3.1-7), there are two values of

    z

    at which ^'^f^'^'^'^)

    =

    0 and c'-^'>

    =

    ^(/,buik) Qj g is as z ^ oo; th e oth er we will define to be

    2;

    = 8. The graphi te

    is assumed to be impermeable to the fluid.

    Before comparing (3.3.1-10) with experimental data, let us consider the

    sources of experimental error. For graphitized carbon blacks, the uncertainty

    in the measured surface area is approximately 10% [235, 254], which trans

    lates to 10% uncertainty in the surface excess mass. Specovius and Findenegg

    [235] and Blumel et al. [236] found that the maximum relative error in their

    buoyancy correction to be 3% and 4% respectively, the values of which could

    be expected to increase as the critical point was approached. Other errors

    such as base-line drift were rep orted t o be < 1 % . Preclu ding any ran do m or

    systematic error, the uncertainty of the results should be less than 15%.

    In the comparisons of (3.3.1-10) with experimental data in Figs. 3.3.1-1

    through 3.3.1-3, we would like to emphasize that no adjustable parameters

    have been used. Equation (3.3.1-10) describes these data within the uncer

    tainty range for argon and krypton as well as for methane at 50C.

    It is clear that, for both the krypton and for the methane, (3.3.1-10) under

    estimates th e dat a, th e error becoming progressively larger as the tem pe ratu re

    decreases. There are at least two possible explanations.

    As noted earlier, Steele [254, p. 52] expressed some doubt about the as

    sumption of additivity of intermolecular forces in using his semi-empirical

    extension of the Lennard-Jones potential to the interactions between a crys

    talline solid and a gas. He proposed that this problem could be at least par

    tially avoided by using a more realistic description of the solid. For this reason,

    he described the solid as being discretely rather than continuously distributed

    in space. While this proposal has not been previously tested experimentally

    to our knowledge, we believe that the excellent agreement between the predic

    tions of our theory and the experimental observations for argon does support

    the use of pairwise additivity, at least when temperature is sufficiently above

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    41/102

    3.3 Applications of Our Extension of Continuum M echanics to the Nanoscale

    7 :

    199

    (a)

    (b)

    6

    5

    4

    ^ 3

    2

    1

    /

    /

    f

    ^

    ^

    ^ - ^ '

    ^ - ^ '

    /

    1 2 3 4 5

    T

    6

    25C

    Fig. 3.3.1-1. r (^mol/m^) for argon on Graphon predicted by (3.3.1-10) (the

    solid curves) (a) as a function of

    P

    (MPa) and (b) as a function of c (mol/L). The

    experimental observations are from Specovius and Findenegg

    [235].

    The dashed and

    dashed-double dot curves in (a) represent the computations of Sokolowski [239] for

    0C and the computations of Fischer [242] for 50C, respectively. The dashed-dot

    curve in (b) indicates the computations of Egorov [237] for 25C

    the cri t ical temperature Tc- Notice that for argon Tc = 150.8 K. For kry pto n,

    Tc =

    209.4 K, a nd we have excellent agreem ent between our prediction s and

    th e expe rime ntal observations at 100C, bu t errors increase as th e t em pe rat ur e

    decreases, the density increases, and the interference of neighboring of kryp

    ton increases. In summary, we believe that pairwise additivity can be used

    as suggested by Steele [254, p. 52], so long as the temperature is sufficiently

    above the cri t ical temperature.

    The poor agreement between our predictions and the experimental obser

    vations for methane is due in part to the same failure of pairwise additivity as

    th e critical tem pe ra tur e Tc = 190.4 K is appro ached . Bu t the errors are larger

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    42/102

    200 3 Applications of the Differential Balances to Momentum Transfer

    ,

    i

    *^ ,.

    10

    (a) r^

    (b) r'

    ^ ^ ^ _

    25 C

    Fig. 3.3.1-2. r (nmol/m^) for krypton on Graphon predicted by (3.3.1-10) (the

    solid curves) (a) as a function of

    P

    (MPa) and (b) as a function of c (mol/L). The

    experimental observations are from Blumel et al.

    [236].

    The dashed curve in (a)

    represents the computations of Rangarajan et al. [238] for0C;the dashed-dot curve

    for 50C; the dashed-double dot curve for 100C. The dashed curve in (b) indicates

    the computations of Egorov [237] for 25C

    than those seen with krypton, which has a higher Tc- We do see the errors

    decrease as the temperature is raised, but we don't have measurements at

    hight temperatures. However, we aje also concerned that the Lennard-Jones

    potential may not be appropriate, particularly in denser fluids, because the

    methane molecule is not spherical.

    As mentioned in the introduction, there are four analysis based upon sta

    tistical mechanics which have been compared with these same data: Egorov

    [237],

    Rangarajan et al .

    [238],

    Sokolowski

    [239],

    and Fischer

    [242].

    In all cases.

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    43/102

    3.3 Applications of Our Extension of Continuum Mechanics to the Nanoscale 201

    per

    Fig. 3.3.1-3.

    r ^

    (|i.mol/m^) as a function of

    P

    (MPa) for methane on Graphon

    predicted by (3.3.1-10) (the solid curves). The experimental observations are from

    Specovius and Findenegg

    [235].

    The dashed curve represents the computations of

    Sokolowski [239] for0C;the dashed-dot curve for 25C; the dashed-double do t curve

    for 50C

    we have used their results from their figures without repeating their compu

    ta t ions .

    Figure 3 .3.1-l(b) com pares the co mp utatio ns of Egorov [237] with (3.3.1-10)

    for the argon data of Specovius and Findenegg [235] at 25C; Figure 3.3.1-2(b)

    compares the two theories for the krypton data of Blumel et al.

    [236].

    The

    computations of Egorov [237] are excellent, although, as he notes, they begin

    to fail at higher densities. Unfortunately, he did not report computations for

    higher temperatures.

    Figure 3.3.1-2(a) compares the model of Rangarajan et al. [238] with

    (3.3.1-10) for the krypton data of Blumel et al.

    [236].

    Figures 3.3.1-l(a) and

    3.3.1-3 compare the Sokolowski [239] model with both (3.3.1-10) and the argon

    and methane data of Specovius and Findenegg

    [235].

    Fischer [242] presented

    a plot of his results compared with the data of Specovius and Findenegg [235]

    for argon at only one temperature 50C. Figure 3.3.1-l(a) compares his results

    with (3.3.1-10). In all cases, (3.3.1-10) was superior.

    Our theory does not do as well in describing the supercritical adsorption

    of krypton or of methane as the critical point is approached. We do not feel

    that this indicates a failure or limitation of the theory, but rather a limitation

    in th e way th a t t h e theo ry w as executed . Following Steele [241, 254, 255],

    we have assumed the Lennard-Jones potential can be used to describe the

    carbon-fluid as well as the fluid-fluid interactions, and we have assumed that

    long-range intermolecular forces are pairwise additive.

    We believe that the poorer agreement between our predictions and the ex

    perimental observations for krypton at lower temperatures is at tr ibutable to

    the assumption of pairwise additivity beginning to fail. The critical tempera-

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    44/102

    202 3 Applications of the Differential Balances to Momentum Transfer

    ture for argon

    Tc =

    150.8 K; for krypton,

    T^ =

    209.4 K. Since the assumption

    of pairwise additivity would begin to fail

    as T

    >

    Tc

    and the fluid became

    denser, we could expect to see these effects at higher temperatures with kryp

    ton than with argon.

    The poor agreement between our predictions and experimental observar

    tions for methane is certainly due in large measure to this same failure of the

    assumption of pairwise additivity. For methane, Tc = 190.4 K, an d t h e er

    ror in our predictions decreases as the temperature is increased. However, we

    are also concerned that the Lennard-Jones potential may not be appropriate,

    particularly in the denser fluids as Tc is approached, because the methane

    molecule is not spherical.

    Steele [254, p. 52] clearly recognized that pairwise additivity of intermolec-

    ular forces could not be justified in condensed m aterials . He propos ed th at this

    problem could be at least partially avoided by using a more realistic descrip

    tion of the solid, one in which the crystalline solid was distributed discretely

    rather than continuously in space. We believe that the excellent agreement

    between the predictions of our theory and the experimental observations for

    argon and krypton , particularly at higher tempe ratur es, supp orts his proposal.

    We wish to emphasize that the use of pairwise additivity is not an inherent

    limitatio n in the use of this theory. For exam ple, in describing the interactio ns

    between two condensed media, the limitations of pairwise additivity can be

    avoided by describing the two-point interactions using an effective, Lifshitz

    t ype ,

    Hamaier constant that is both screened and retarded

    [170].

    3 .3 .2 Stat ic Contact Angle [20]

    It has been recognized for a long time that the molecular interactions play

    an important role within the immediate neighborhood of a three-phase line of

    contac t or comm on line where the meniscal film is very thin. T he configuration

    of the fluid-fluid interface in the common line region and static contact angle

    depend on the correction for intermolecular forces (see Sect. 2.2).

    Rayleigh [261] recognized that, in two dimensions sufHciently far away

    from the common line, the fluid-fluid interface approaches a plane inclined

    at a angle 0(*^*) to the solid surface. As the common line is approached,

    the curvature of the interface is determined by a force balance to form a

    different angle 0o with the solid surface. More recently, the configuration of

    the interface near the common line has been investigated in the context of

    statistical mechanics [262, 263] and by molecular dynamics simulations

    [264].

    See Dussan V [115] for a further review.

    This problem has been discussed theoretically by many investigators [124,

    162, 163, 180, 181, 265-270]

    Figure 3.3.2-1 shows in two dimensions the three-phase line of contact or

    common line formed at equilibrium between phases A, C, and a solid B. The

    static contact angle 0(^*^*) might be measured by an experimentalist at some

    distance from the common line, using perhaps 10x ma gnification.

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    45/102

    3.3 Applications of Our Extension of Continuum Mechanics to the Nanoscale 203

    Fig. 3.3.2-1. A thin discontinuous film

    C

    forms a common line

    Our objective is to determine 0(stat)^ which we will distinguish from 0O)

    the contact angle at the common line introduced in the context of Young's

    equation (2.1.10-3). We will focus on the thin (precursor) film (Sect.

    1.2.9)

    formed in the neighborhood of the true common line, and we will employ the

    correction for intermolecular forces discussed in Sect. 2.2.5.

    We will make three assumptions.

    i) The solid is rigid, and its surface is smooth and planar. Because the effect

    of gravity will be ignored with respect to the correction for intermolecular

    forces, its orientation with respect to gravity is arbitrary.

    ii) The system is static, and equilibrium has been established. There are no

    interfacial tension gradients developed in the system.

    iii) In discussing the thin film we will adopt view (v) of Sect.

    2.2.1.

    The

    corresponding correction for intermolecular forces from adjoining phases will

    be taken from Sect. 2.2.4.

    Referring to Fig. 3.3.2-1 and Table B.1.2-6, our initial objective will be to

    determine the configuration of the A-C interface

    Z3 = /l (Z i)

    assuming that the C B interface is

    Z3 = 0

    (3.3.2-1)

    (3.3.2-2)

    Both the z\ and Z3 comp onents of the jum p m om entum balance for the AC

    interface reduces to

    a t z3 = / i : p ( ^ ) - p ( ^ ) = 7

    dzi2

    1+

    ( ^ ) '

    - 3 / 2

    (3.3.2-3)

    Here

    p'^^^

    and

    p^^^

    are the pressures in these phases evaluated at the

    AC

    in

    terface. In view of ass um ptio n iii, we conclud e from th e differential m om en tum

    balance (2.2.1-1) that

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    46/102

    204 3 Applications of the Differential Balances to Momentum Transfer

    p ( ^ ) = p ( ^ ) +

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    47/102

    3.3 Applications of Our Extension of Continuum Mechanics to the Nanoscale 205

    atz^ = Q: h* = 1

    dh*

    at 2; = 0 : -p- r = ta nOQ

    dz*

    where

    &o

    is the conta ct angle at the t ru e com mon Une.

    Noting that

    d^h*

    _ dh* d^h* dz^

    dzf ~~d4dzfdh*

    _ l_d_ /dh*V dzt

    ~2dzf \d^J dh*

    _ l_d_ /dfe*^^

    ~

    2dh* \dzt^

    (3.3.2-12)

    (3.3.2-13)

    (3.3.2-14)

    we may write (3.3.2-11) as

    h*^

    ~ 2dh* \dz*) ^ V d ^ y

    -3 /2

    (3.3.2-15)

    Integrating this once consistent with (3.3.2-12) and (3.3.2-13),we have

    -1 /2

    ( l - F t a n ^ ^ o )

    -1 /2

    1 +

    dh*

    dz*

    T\h*

    - 1

    In the limits

    a s 2 ;* > 0 0 : / i *

    >

    CX3

    and

    dzi

    Eq. (3.3.2-16) further reduces to

    (1 + tan^ 0o ) ' ^ ' - (1 + tan^ (^t^*))

    1/2 ^

    B*

    or

    | c O S 0n |

    cos

    0(s ta t )

    = T-

    127r7

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    48/102

    206 3 Applications of the Differential Balances to Momentum Transfer

    | co s0o |

    COS

    0(s ta t )

    = T-

    127r7(5(sc')^

    (3.3.2-21)

    The angle

    OQ

    at the true common line should be determined using Young's

    equation (2.1.10-3)

    7(^^ ) cos 00 =

    7 ^^^

    -

    7^-^^^

    After some arrangement, we rearrange (3.3.2-21) as

    cos

    +

    3.3.2-22)

    3.3.2-23)

    The Hamaker constants

    A^^*^^

    and yl^^^^ were calculated by Hough an d

    W hi t e

    [162],

    and

    j(-^'^)

    were measured by Fox and Zisman

    [271].

    Girifalco

    an d G ood [272] (see also A dam son [273, p . 360]) prop ose d th a t 7(^C') could

    be estimated as

    ^ ( B C ) o o ^ f^^{AC)oo _ y ' ^ ( A B ) > y

    7

    {AB)oo _^_

    ^ ( ^ C ) o o

    2 ( ^ ( A B ) o o ^ ( A C ) o o - j l / 2

    3.3.2-24)

    It is important to recognize that the n-alkanes are volatile liquids and

    that the gas phase is dependent upon the specific n-alkane being considered.

    Fox and Zisman [271] observed that there is no difference between contact

    angles measured in saturated air and those measured in unsaturated air . The

    conclusion is that 7(^^) is independent of the particular n-alkane being

    considered. Prom (3.3.2-22) and (3.3.2-24), we conclude that

    ^{AC)oo^^^Q^^^{AB)oc

    iAB)oo _^ ^ ( A C ) o o

    2(^{AB)oo^{AC}oo\

    1/2-

    3.3.2-25)

    or

    cos 00

    -1 + 2

    ry{AB)00

    3.3.2-26)

    As the alkane number n is decreased, there is a limiting value at which spon

    taneous wetting occurs, and 0o = 0(^*' '*) = 0 and 7(^^) =

    r^(AC)oo_ ^^

    .j jg

    way, Zisman [274, p. 190] (see also Zisman [160, p. 20]) concluded that

    ^(AB)oo = i 8 .5n iN /m

    3.3.2-27)

    Working with different liquids on PTFE, Owens and Wendt [275] came to the

    same conclusion.

    In their analysis of supercritical adsorption of argon and krypton on im

    permeable carbon spheres, Fu et al. [22] chose 6 as the position at which

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    49/102

    3.3 Applications of Our Extension of Continuum Mechanics to the Nanoscale 207

    attractive and repulsive forces between the carbon and the gas, as described

    by a two-point Leonard-Jones potential, were balanced and

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    50/102

    208 3 Ap plicatio ns of th e DifTerential Balan ces to M om en tum Transfer

    a n d a pp l i e d a t t h e a p pa r e n t c o m m on l i ne ( a s s e e n w i th 10x m a gn i f i c a t i on ) .

    T a b le s 3 .3 .2 - 1 a nd 3 .3 .2 - 2 de mons t r a t e t ha t ( 3 . 3 .2 - 29 ) doe s no t r e p r e se n t t he

    e xpe r ime n ta l me a su r e me n t s o f 0 ( ^ *^ * ) w e l l .

    S l a t te r y [108 , p . 169] de r iv es Y ou ng 's eq ua t io n as (3 .3 .2-22) , an d app l ie s

    i t a t t h e t r ue c o m m on l i ne a s w e do he r e . E qu a t io n ( 3 .3 .2 - 23 ) , w h ic h w a s

    d e v e l o p e d u s i n g ( 3 . 3 .2 - 2 2 ), r e p r e s e n t s t h e e x p e r i m e n t a l d a t a i n t h e a b o v e

    ta b l e s mu c h be t t e r t h a n do e s ( 3 .3 .2 -29 ) . O ur c onc lu s ion is t h a t ( 3 .3 .2 - 22) i s

    t h e p r e f e r r e d f o r m of Y o ung ' s e q ua t ion .

    T a b l e 3 . 3 . 2 - 2 . Comparison of ca lcula t ion and exper imenta l da ta of

    static contact angle 0^^*^' ' for dispersive liquids on PDMS

    liquids

    d i iodomethane

    bromonapha lene

    methylnaphtha lene

    ie r t -bu ty l naphtha lene

    liq. paraffin

    Hexadecane

    ^{AC)oa

    m N / m

    50.8

    AAA

    39.8

    33.7

    32.4

    27.6

    ^Csc)

    A

    3.178

    2.497

    2.378

    3.384

    3.139

    3.044

    ^(BC') fc

    10-2 J

    5.57

    5.05

    4.80

    4.76

    5.15

    4.80

    ^ ( C C ) ' '

    10-2 J

    7.18

    5.93

    5.35

    5.23

    6.03

    5.22

    0(cal)'=

    Deg.

    71.9

    66.4

    61.3

    52.5

    50.2

    38.6

    Q(CB.\)d-

    Deg.

    66.8

    61.0

    50.5

    50.1

    44.5

    34.4

    @(exp)

    Deg.

    70.0^

    62.0^

    52.0^

    49.0

    40.0^

    36.09

    Calculated by (3.3.2-28).

    ' ' P rov ided by Drummond and Chan

    [279].

    '^ Calcula ted using a common form of Young's equat ion (3.3.2-29) .

    C alc ula ted by (3.3.2-23 ).

    ^ Measured by Chau dhury and W hi tes ides [276].

    ' ' Measured by Baler and Meyer [277].

    9 Measured by Bowers and Zisman

    [278].

    T hi s i s pe rh ap s th e fi rs t t e s t o f th e form of Yo un g 's equ a t io n (3 .3 .2-22)

    de r ive d by S l a t t e r y [ 108 , p . 169 ] f or t h e t r ue c om m on l i ne . C om m on ly Y o ung ' s

    e q u a t i o n i s a p p l i e d a t t h e a p p a r e n t c o m m o n l in e w i t h 0 o ( a s o b s e r v e d a t t h e

    t r u e c om m on l ine ) r e p l a c e d by 0(^*^*) ( a s obse r v e d a t t h e a ppa r e n t c o m m on

    l ine a s s e e n by a n e xp e r im e n ta l i s t w i th , s a y , 10 x m a gn i f i c a t i on ) .

    3 . 3 . 3 A R e v i e w o f C o a l e s c e n c e ( w i t h J . D . C h e n )

    T he r a t e a t w h ic h d r ops o r bubb le s , su spe nde d in a l i qu id , c oa l e sc e i s impor

    t a n t t o t h e p r e p a r a t i o n a n d s t a b i l i t y of e m u l s ions , of f oa ms , a n d of d i spe r s ions ;

    t o l i qu id - l i q u id e x t r a c t i on s ; t o t h e f o r m a t ion o f a n o il ba nk du r in g th e d i s

    pl ac em en t of oi l f rom a rese rvo ir rock ; to m in er a l f lota tion.

    O n a sma l l e r s c a l e , w he n tw o d r ops ( o r bubb le s ) a r e f o r c e d to a pp r oa c h one

    a n o t he r i n a l i qu id pha s e o r w he n a d r op i s d r ive n th r o ug h a li qu id pha s e t o a

    so l id sur face , a th in l iqu id f i lm forms be tween the two in te r faces and begins to

    d r a in . A s t he t h i c kne s s o f t he d r a in ing f i lm be c ome s su f f i c i e n t ly sma l l ( a bou t

    100 n m ) , th e ef fec ts o f th e cor rec t ion for in t e rm ole cu la r fo rces (Sec t . 2 .2 ) an d of

    e l e c t r o s t a t i c doub le - l a ye r f o r c e s be c ome s ign i f i c a n t . D e pe nd ing upon the s ign

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    51/102

    3.3 Applications of Our Extension of Continuum Mechanics to the Nanoscale 209

    and the magnitude of the disjoining pressure attr ibutable to the correction

    for intermolecular forces and the repulsive force of electrostatic double-layer

    forces, there may be a critical thickness at which the film becomes unstable,

    ruptures and coalescence occurs. (See Exercise 2.2.4-1 for the definition of the

    disjoining pressure.)

    In examining prior studies, it is helpful to consider separately those per

    tinent to the early stage of thinning in which the effects of any disjoining

    pressure are negligible and those describing the latter stage of thinning in

    which the disjoining pressure may be controlling. Since the literature is ex

    tensive, we will also separate studies of coalescence on solid planes from those

    at fluid-fluid interfaces. (Comprehensive reviews of thin films are given by

    Kitchener

    [280],

    Sheludko

    [281],

    Clunie et al.

    [282],

    Buscall and Ottewill

    [283],

    and Ivanov

    [284].

    Reviews concerned with the drainage and stability of thin

    liquid films are given by Liem and Woods

    [285],

    Ivanov and Jain

    [286],

    Jain

    et al.

    [287],

    Ivanov

    [288],

    Ivanov and Dimitrov

    [289],

    and Maldarelli and Jain

    [290].)

    In i t ia l s tage of th inning: a t a so l id p lane For th e mo m ent, let us

    confine our attention to the initial stage of thinning in which the effects of the

    correction for intermolecular forces and of electrostatic forces can be neglected.

    When a drop or bubble approaches a solid plane, the thin liquid film

    formed is dimpled as the result of the radial pressure distribution developed

    [291-297].

    It is only in the limit as the approach velocity tends to zero that

    the dimple disappears.

    Prankel and Mysels [298] have developed approximate expressions for the

    thickness of the dimpled film at the center and at the rim or barrier ring as

    functions of time. Their expression for the thinning rate at the rim is nearly

    equal to that predicted by the simple analysis of Reynolds [299] for two plane

    parallel discs and it is in reasonable ag reemen t w ith some of th e me asure m ents

    of Platikanov [294, 300, 301]. Their predicted thinning rate at the center of

    the film is lower than that seen experimentally [294, 300, 301]. However, it

    should be kept in mind that their result contains an adjustable parameter (an

    initial time) that has no physical significance.

    Hartland [295] developed a more detailed model for the evolution of the

    thinning fi lm assuming that the shape of the drop beyond the rim did not

    change with time and that the configuration of the fiuid-fluid interface at the

    center was a spherical cap. He proposed that the initial film profile be taken

    directly from experimental data.

    H ar tla nd an d Ro binso n [302] assum ed th a t the fluid-fluid interface co nsists

    of two parabolas, th e radius of curvatu re at the apex varying with t ime in th e

    central parabola and a constant in the peripheral parabola. A priori knowledge

    was required of the radial position outside the dimple rim at which the film

    pressure equaled the hydrostatic pressure.

    For the case of a small spherical drop or bubble approaching a solid plane,

    the development of Lin and Slattery [300] is an improvement over those of

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    52/102

    210 3 Applications of the Differential Balances to Momentum Transfer

    Hartland [295] and of Hartland and Robinson

    [302].

    Since the initial and

    bo un da ry con ditions are more comp lete, less a priori expe rime ntal information

    is required. The results are in reasonable agreement with Platikanov [294]'s

    data for the early stage of thinning in which any effects of the correction for

    intermolecular forces and of electrostatic forces can generally be neglected.

    (Since the theory is for small spherical drops, it cannot be compared with the

    data of Hartland [295] for large drops.)

    Initial stage of thinning: at a fluidfluid interface Let us continue

    to confine our attention to the initial stage of thinning in which the effect

    of the correction for intermoleculax forces and of electrostatic forces can be

    neglected.

    As a drop (or bubble) is forced to approach a fluid-fluid interface, the

    minimum film thickness is initially at the center. As thinning proceeds, the

    minimum film thickness moves to the rim or barrier ring [303-305] and a

    dimpled film is formed. Allan et al. [303] found that the dimpling developed

    wh en th e film was 0.3 to 1.2 jim thick. T he rim ra diu s is a function of tim e

    [303-305].

    Princen [306] extended the Prankel and Mysels [298] theory to estimate

    the thinning rate both at the center and at the rim as a small drop (or bubble)

    approaches a fluid-fluid interface. His prediction for the thinning rate at the

    rim is again nearly eq ual to th at given by the simple analysis of Reyn olds

    [299].

    But, just as in the Prankel and Mysels [298] theory, there is an adjustable

    parameter (an initial time) that has no physical significance.

    Ha rtlan d [307] developed a more de tailed analysis to predict film thickness

    as a function b ot h of tim e and of radial position. He assum ed th at bo th fluid-

    fluid interfaces were equidistant from a spherical equilibrium surface at all

    t imes and that the fi lm shape immediately outside the rim was independent

    of time. The initial film profile had to be given experimentally.

    Lin and Slattery [23] developed a more complete hydrodynamic theory for

    th e thinn ing of a liquid film between a small, nearly spherical drop an d a fluid-

    fluid interface that included an a priori estimate of the initial proflle. Their

    theory is in reasonable agreement with data of Woods and Burrill

    [308],

    B urrill

    and Woods

    [309],

    and Liem and Woods [310] for the early stage of thinning in

    which any effects of London-van der Waals forces and of electrostatic forces

    generally can be neglected. (Because their theory assumes that the drop is

    small and nearly spherical, it could not be comp ared w ith Ha rtla nd [311, 312,

    313] data for larger drops.)

    Barber and Hartland [314] included the effects of the interfacial viscosities

    in describing the thinning of a liquid film bounded by partially mobile parallel

    planes. The impact of their argument is diminished by the manner in which

    they combine the effects of the interfacial viscosities with that of an interfacial

    tension gradient.

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    53/102

    3.3 Applications of Our Extension of Continuum Mechanics to the Nanoscale 211

    Latter s tage of th inning: a t a so l id p lane Exp erim entally we find

    that, when a small bubble is pressed against a solid plane, the dimpled hq-

    uid film formed may either rupture at the rim [293, 296] or become a flat,

    (app arently ) stable film [291, 293, 315-321].

    When the thickness of the draining film becomes sufficiently small (about

    1000

    A),

    the effects of the London-van der Waals forces and of electrostatic

    double-layer forces become significant. Depending upon the properties of the

    system and the thickness of the film on the solid, the London-van der Waals

    forces can contribute either a positive or negative component to the disjoining

    pressure [174, 321 , 322]. An electros tatic dou ble layer also ma y con tribu te

    either a positive or negative component to the disjoining pressure. A positive

    disjoining pressure will slow the rate of thinning at the rim and stabilize the

    thinning film. A negative disjoining pressure will enhance the rate of thinning

    at the rim and destabilize the film.

    When London-van der Waals forces were dominant, with t ime the fi lm

    either ruptures or becomes flat. Platikanov [294] and Blake [321] showed ex

    perimentally that a positive London-van der Waals disjoining pressure can be

    responsible for the formation of a flat, equilibrium wetting film on a solid sur

    face. Schulze [323] found that, for an aqueous film of dodecylamine between a

    bub ble and a qua rtz surface a t a high KCl concen tration (0.01 to 0.1 N) , the

    eff'ects of the electrostatic double-layer forces are negligible compared with

    those of the London-van der Waals forces. If the electrostatic double-layer

    disjoining pressure is positive, the wetting film thickness decreases with in

    creasing electrolyte concentration. Read and Kitchener [316] showed that for

    a dilute K Cl aqueou s film (< 10~^ N) on a silica surface thicker th an 300 A,

    the electrostatic double-layer forces dominate. As the KCl concentration in

    creases from 2 X 10~^ to 10~^ N, they found th at th e w etting film thickness

    decreases from 1200 to 800 A due to th e decrease in the e lectrostatic double-

    layer repulsion. Similarly, Schulze and Cichos [320] found that the thickness of

    the aqueous wetting film between a bubble and a quartz plate decreases with

    increasing KCl concentration in the aqueous film. The wetting fllm thickness

    varied from almost 500 A in 10 ^ N K Cl solution to abo ut 100 A in 0.1 N KCl

    solution. The y concluded t ha t, for film thickness greater t ha n 200 A, the con

    tribution of London-van der Waals forces to the disjoining pressure could be

    neglected in aqueous electrolyte solutions of less th an 0.01 N conc entration .

    A positive contribution to the disjoining pressure by electrostatic double-

    layer forces can be changed to a negative one by the adsorption in one of the

    interfaces of an ion having the opposite charge. For example, a stable aqueous

    wetting film may be formed in the absence of an ionic surfactant, but the film

    may rupture in its presence [323-325].

    When the contributions to the disjoining pressure of both the London-

    van der Waals forces and the electrostatic double-layer forces are negative,

    the dimpled film always ruptures, and the rupture thickness decreases with

    increasing elec trolyte c onc entr atio n [296, 318, 323].

  • 8/10/2019 Applications of the Differential Balances to Momentum Transfer

    54/102

    212 3 Applications of the Differential Balances to Momentum Transfer

    W he n these two com pone nts to th e disjoining pressure have different signs,

    th e dimpled film will either ru ptu re or become flat, depe nding u po n which con

    tribution is dominant. Schulze and Cichos [324] explained that the existence

    of a 0.1 N KC l aque ous film on a q ua rtz surface is du e to th e ex istence of a

    positive London-van der Waals disjoining pressure. They also explained the

    ru pt ur e cause d by t he ad dit ion of AICI3 is due t o th e a dso rptio n of Al *

    ions in the quartz surface, resulting in a negative electrostatic double-layer

    contribution to the disjoining pressure.

    Because of mathematical difficulties, especially when using the dimpled

    film model, only London-van der Waals forces have been taken into account

    in most analyses of the stability of thin liquid films on solid planes. Only Jain

    and Ruckenstein [326] and Chen [327] have included the effect of electrostatic

    double-layer forces in their disjoining pressure. Chen [327] has discussed the

    interaction of positive and negative components to the disjoining pressure on

    the dynamics of a dimpled film. The numerical results are consistent with the

    experimental observations mentioned above.

    Jain and Ruckenstein [326, 328] studied the stability of an unbounded,

    stagnant, plane, liquid film. Jain and Ivanov [329] considered a ring-shaped

    film. Their results suggest that the critical thickness decreases as the rim

    radius increases and as the width of the ring decreases. Blake and Kitchener

    [318] found experimentally that films of smaller diameter were more stable t