Drying Theory

download Drying Theory

of 6

Transcript of Drying Theory

  • 8/20/2019 Drying Theory

    1/12

    J . A m .

    Cerum.

    Soc.

    73 [ l ]

    -14 1990)

    journal

    Theory of Drying

    George W. Scherer*

    Central Research

    and

    Development ,

    E.

    I.du Pont de Nemours & Co.,

    Wilmington, Delaware 19880-0356

    This review exam ines the stage s of dry-

    ing, with the emph asis on the constant

    rate period

    CRP),

    when the pores are

    full of liquid. It is during the CRP that

    most of the shrinkage occurs and the

    drying stresses rise to a maximum . We

    examine the forces that produce shrink-

    age and the mechanisms responsible or

    transpo rt of liqu id. By analyzing the in -

    terplay of fluid flow and s hrinkag e of the

    solid ne twork, it is possible to calculate

    the pressure distribution in the liquid in

    the pores. The tension in the liquid is

    found to be greatest near the drying sur-

    face, resulting in greater compressive

    stresses on the network in that region.

    This produces differential shrinkage of

    the solid, which is the cause of cracking

    during drying. The p robab ility of fracture

    is related to the size of the body, the rate

    of evaporation, and the strength of the

    network. A variety of strategies for avoid-

    ing fracture during drying are discussed.

    [Key words: drying, sh rinkage, cracking,

    models, gels.]

    1

    Introduction

    REMOVALof liquid is particularly trouble-

    some in sol-gel processing, because gels

    tend to warp and crack during drying, and

    avoid ing fracture requires inconveniently

    slow dryin g rates. Howe ver, liquid trans-

    port processes are also of imp ortance in

    other ce ramic-formingoperations, includ-

    A. H.

    Heuer-contributing editor

    Manuscript

    No.

    198096. Rece ived September 25,

    Mem ber, American Ceramic Society.

    1989; approved October

    17,

    1989.

    ing slip casting, tape casting, binder burn-

    out, liquid-p hase sintering, and dryin g of

    clays. Indeed, the princip les of flow in po-

    rous media are of such general interest

    that they have been frequently redisco-

    vered over the past 60 years, and rele-

    vant literature is found in fields including

    soil science, food science, and polymer

    materials science, as well as ce ramics . In

    most cases, liquid flows through a porous

    body in response to a gradient in pres-

    sure; at the same time, the pressure

    causes de formation of the solid network,

    and d ilatation of the p ores through wh ich

    the liquid moves. In this review we ana-

    lyze the interaction between low of the li-

    qu id and dilatation of the solid in order to

    predict the stresses and strains that de-

    velop during drying. Special attention is

    given to the problems encountered in dry-

    ing gels, bu t the analysis is quite general.

    The driving forces for shrinkage of the

    solid and the mechanisms for transport of

    the liquid are discussed n Section II.The

    stages of dryin g are outlined in Section 111,

    and a mo del for calculation of drying

    stresses is developed in Section IV. The

    cause of c racking during d rying an d vari-

    ous strategiesfor avoiding fracture are de -

    scribed in Section

    V.

    These topics are

    discussed in greater detail in a forthcom-

    ing book.'

    I I

    Deformation and

    Flow

    1) Driving Forces for Shrinkage

    The first stage of drying is illustrated in

    Fig. 1(B): for every unit volume of liquid

    that evaporates, the volume of the body

    decreases by one unit volume, so the li-

    quidlvapor interface (meniscus) remains

    at the surface of the body. In gels, this

    stage continues while the bo dy shrinks to

    as little as one-tenth of its orig ina l volume.

    The forces that pro duc e the shrinkage of

    the solid network are discussed below.

    A) CapillaryPressure: If evaporation

    George W. Scherer has bee n a mem ber

    of the Central Research Departm ent of

    E. I. du Pont de Nemours & Co. since

    1985 . His work at Du Pont has dealt prrn-

    cipally with sol-gel proces sing, and es-

    pecially with drying. In collaboration with

    Jeff Brinker of Sandia N ational Labs, he

    has written a book entitled Sol-Gel

    Science

    that will be published by Aca-

    dem ic Press in February. From 197 4 to

    1985 , Dr. Scherer was at Cornin g Glass

    Works, where his research included op -

    tical fiber fab rication, viscous sintering,

    and viscoelastic stress analysis. The lat-

    ter work was the su bject of his first bo ok,

    Relaxation in Glass and Composites

    (Wiley, 198 6). He receive d his

    B.S.

    and

    M.S.

    egrees in 1972 and his Ph.D. in

    materials science in 1974 , all from

    MIT,

    where his thesis work was on crystal

    growth in glass.

    3

  • 8/20/2019 Drying Theory

    2/12

    Vol. 73,No. 1

    Journal

    of

    the American Ceramic Society Scherer

    of

    liquid from the pores were to expose

    the solid phase, a solidlliquid interface

    would be replaced by a more energetic

    solidhapor interface. To prevent such an

    increase in the energy of the system, liquid

    tends to spread from the interior of the

    body to cover that interface. Since the vol-

    *The stress in the liquid, P. is positive when the

    liquid is in tension. The pressure, P i , follows the op-

    posite sign convention

    (Pi =

    - P ) ,

    so

    tension is

    negative pressure.

    Fig.

    1.

    Schematic illustration

    of

    drying process: black network represents solid phase and shad-

    ed area is liquid filling pores. A) Before evaporation begins, the meniscus is flat.

    B)

    Capillary

    tension develops in liquid as it "stretches" to prevent exposure

    of

    the solid phase, and network

    is drawn back into liquid. The network is initially so compliant that

    little

    stress

    is

    needed

    to

    keep

    it submerged, so the tension in the liquid is

    low

    and the radius of the meniscus is large.As the

    network stiffens, the tension rises and, at the critical point (end

    of

    the constant rate period), the

    radius of the meniscus drops

    to

    equal the pore radius. (C) During the falling rate period, the li-

    quid recedes into the gel.

    ume of liquid has been reduced by evapo-

    ration, the meniscus must become curved

    as indicated in Fig. 2. The tension 0 n

    the liquid

    is

    related to the radius of curva-

    ture

    (r)

    of the meniscus by*

    where y ~ vs the liquidlvapor interfacial

    energy (or surface tension). When the cen-

    ter of curvature s

    ir

    the vapor phase, the

    radius of curvature is negative and the

    li-

    quid is in tension

    PX).

    The maximum capillary tension PR)n

    the liquid occurs when the radius of the

    meniscus is small enough to fit into. the

    pore; for liquid in a cylindrical pore of ra-

    dius

    a ,

    he minimum radius of the menis-

    cus is

    where

    8

    s the contact angle.

    If 8 s go",

    then the liquid does not wet the solid and

    the liquidhapor interface is flat F

    w ,

    P =

    0).

    f

    8=

    0

    the solid surface

    is

    covered

    with a liquid film. Of course, the pores in

    a real body are not cylindrical, but it can

    be shownW that the maximum tension is

    related to the surface-to-volume ratio of

    the pore space,

    SplVp:

    where

    ysv

    and y s ~ re the solidlvapor

    and solidliiquid interfacial energies,

    respectively. The specific surface area of

    a porous body (interfacial area per gram

    of solid phase),S , s related to the surface-

    to-volume ratio by3

    where

    e is

    the relative density,

    e =e ,

    eb

    is

    the bulk density of the solid network

    (not counting the mass

    of

    the liquid), and

    es

    is

    the density of the solid skeleton (the

    skeletal density). The quantity

    VplSp is

    also known as the hydraulic radius.As we

    shall see, during most of the drying pro-

    cess the capillary tension is smaller than

    this maximum value.

    (6) Osmotic Pressure: Osmotic pres-

    sure

    (n)

    is

    produced by a concentration

    gradient, as in the case of pure water

    diffusing through a semipermeable mem-

    brane to dilute a salt-rich

    solution

    on the

    other side. As indicated

    in

    Fig.

    3

    pressure

    n

    must be exerted on the solution (or a

    tension of

    - n

    must be exerted on the

    pure liquid)

    to

    prevent :he water from en-

    tering the solution. The pressure s a meas-

    ure of the difference in chemical potential

    between the pure liquid and that in the so-

    lution. An analogous situation can arise if

    the pores of the drying body contain a so-

    lution of electrolyte: evaporation of solvent

    increases the salt concentration near the

    drying surface,so liquid diffuses from the

    interior

    to

    reduce the concentration gra-

    dient; the decrease in the volume of liquid

    in the interior causes tension in the liauid

  • 8/20/2019 Drying Theory

    3/12

    January 1990

    Theory of Drying 5

    that remains there. If the pores are large,

    the diffusive flux is matched by counter-

    flow of liquid toward the interior, and no

    stress develops. However, if the pores are

    small enough to inhibit flow, diffusion away

    from the interior can p roduce tension in

    the liquid in that region; then the balanc-

    ing compression in the solid phase (which

    in principle could approach

    fl

    can pro-

    duce shrinkage. In such a situation, the

    solid network plays the role of a semi-

    permeab le mem brane, p ermitting trans-

    port in only one direct ion. This

    phenom enon could b e of imp ortance in

    clays and gels. Alkoxide-derived gels

    generally contain a solution of liquids that

    differ in volatility (viz., alcohol and w ater),

    so

    evaporation creates a com position gra-

    dient, and osmotic flow may result.

    (C) Disjoining Pressure: D isjoining

    forces are short-range forces resulting

    from the presence of a solidlliquid inter-

    face. The most important examples are

    dou ble-layer repulsion between ch arged

    surfaces and interactions caused by struc-

    ture created in the liquid by dispersion

    forces. Liquid molecules, especially

    water,43 tend to adop t a special structure

    in the vicinity of a solid surface. The inter-

    action with the surface is so strong that ad-

    sorbed layers

    =I

    nm thick resist freezing.6

    As

    evaporation occurs and solid surfaces

    are brought together, repulsive forces aris-

    ing from electrostatic repulsion, hydration

    forces, and solvent structure resist contrac-

    tion of the gel. The pore liqu id will diffuse

    or flow from the swollen interior of the gel

    toward the exterior to allow the solid sur-

    faces

    to

    move farther apart. The disjoin-

    ing forces thus pro duce an osmotic flow,

    where transport is driven by a gradient in

    chemical potential in the liquid phase.

    Since these forces become important

    when the separation between Surfaces is

    small, they are most likely to be impor-

    tant near the end of dryin g of gels, whe n

    the pore diameter may approach 2

    nm.

    Macey7 argues that electrostatic rep ul-

    sion between particles of clay produces

    tension in the liqu id that draws flow from

    the interior of a drying body. Even for

    clays, in which these phenomena are most

    evident, it has be en argued8 that osmotic

    forces must be less important than capil-

    lary pressure, because moisture gradients

    persist in clays for long periods when

    evaporation is prevented. In addition, it

    has been shown that the final shrinkage

    of kaolinite clay during drying

    is

    directly

    related to the surface tension of the pore

    liquid.9 The swelling pressure of clays in

    water is < I 0 MPa,lO which is com parable

    to the capillary pressure in pores with radii

    >14

    nrn (according to Eq.

    (l),

    assuming

    yLv=

    0.072

    Jlm2 for water). In the case of

    gels, the pores are generally smaller than

    that, so capillary forces are e xpected to

    dominate.

    0) Mo isture Stress: Moisture stress

    or moisture potential y) s the partial

    specific Gibbs free energy of liquid in a

    porous medium, and is given byir

    5)

    =

    where eL and V,,, are the density and mo -

    lar volume of the liquid, Rs is the ideal

    gas constant, T is the temperature, p v is

    the vapor pressure of the liquid in the sys-

    tem, and po is the vapor pressure over a

    flat surface of the pure liquid. In soil

    science12

    it

    is conventional to define the

    moisture potential in terms of the equilibri-

    um height to which it would draw a

    colum n of water, so a factor of g (the

    gravitational acceleration) would be includ-

    ed in the denominator on the right side of

    Eq.

    (5).

    The m oisture potential is quite in-

    clusive, because the vapor pressure is

    depressed by factors including capillary

    pressure, osmotic pressure, hydration

    ture potential subsumes all of the driving

    forces discussed above, and can be o b-

    tained by m easuring the vapor p ressure

    of the liquid in the system. For that rea-

    propriate potential driving shrinkage of On

    the

    solid

    phase

    shrinkage.

    gels during drying. The difficulty in im-

    plementing that suggestion is that capil-

    lary pressure gradients produce flow,

    while concentration gradients (that p ro-

    duce osmotic pressure) cause diffusion,

    so it is necessary to app ly portions of the

    total potential to different transport

    processes. In soil science, it is customary

    to assume that fluid flow is driven by the

    gradient in moisture potential, but it is done

    with the understanding that factors other

    than cap illary pressure and gravitation are

    negligible.12

    e x )

    k)

    forces, and adsorption forces. Thus, mois- A) (B)

    Fig. 2

    To

    prevent exposure of the solid

    Phase Ah the liquid must adopt a curved 11

    son, Zarzycki13 recom men ds it as the ap-

    quidhapor interface (B). Compressive forces

    Fig.

    3. Water diffuses into the salt solution to equilibrate the concentration on either side

    of

    the impermeable membrane; pressure il would have to be exerted

    on

    the solution

    to

    prevent

    the influx of water.

  • 8/20/2019 Drying Theory

    4/12

    6

    Journal

    of

    the American Ceramic Society Scherer

    VOl. 73, No.

    1

    2)

    Transport

    Processes

    A) Darcy's Law: Fluid flow throug h

    porous media obeys Darcy's la~,14~15

    which states that the flux of liquid, J , is

    proportional o the grad ient in pressure in

    the liquid,

    VP

    The flux is in units of volume per area of

    the porous body (not the area occupied

    by the liquid) per time,

    PL is

    the force per

    unit area of the liquid, q~ is the viscosity

    of the liquid, and D is called the p ermea-

    bility and has units of area. Positive flux

    moves in the direction of increasingly

    negative pressure (i.e., he flow is toward

    regions of greater tension in the liquid).

    Equation

    (6)

    s an empirical equation der-

    ived from observation of flow of water

    through soi1,16 but it is analogous to

    Poiseuille's aw for flow of liquid through

    a straight circular pipe. This analogy has

    given rise to man y models for the perme -

    ability

    of

    porous media based on

    representationsof the pores by arrays of

    tubes, many of which are discussed in the

    excellent texts by Scheideggerl4 and D ul-

    lien; l5 van Brake117 offers a critic al review

    of over

    300

    such models. The most popu -

    lar model, because of its simplicity and ac-

    curacy, is the Carman-Kozeny equation,

    which gives the permeability in terms of

    the relative density and specific surface

    area:

    (7)

    The factor of 5 is an empirical correction

    for the noncircular cross section and non-

    linear path of a ctual pores . This equation

    is reasonably successful for many types

    of granular m aterials, but it often fails, and

    should be applied with caution.

    The p roportionality of the flux to the

    pressure gradient is obeyed by many

    materials, including those with po res

    smaller than 10 nm, as in porous

    Vycort718 and alkoxide -derived gels.19

    Even in unsaturated bodies (i.e., where the

    pores contain both liquid and gas),

    Da rcy's law is obeyed1420 as long as the

    liquid

    phase is funicular (i.e., interconnect-

    ed); if the liquid is pendular (i.e., isolated

    in pockets), t can only b e transported by

    diffusion of the vapor. The permeability of

    unsa turated materials is a strong function

    of liquid content and shows considerable

    hysteresis as the liquid content is raised

    and lowered.

    In gels the pores are

    so

    small that a

    large portion of the liquid may be in struc-

    tured layers within =1 nm of a solid sur-

    face, so the effective viscosity may be

    greater than in the bulk liquid. The re-

    duced mobility in such lavers can be

    unfortunately, attempts at direct measure-

    ment of the viscosity near solid sur-

    faces243 have been shown26 to give

    incorrect results. Spe ctrosco pic methods

    indicate an increase in viscosity by a fac-

    tor of -3, so the effect of solvent structure

    on the flux in gels can be substantial.

    (B) Diffusion: According to Fick's

    law, the diffusive flux

    Jo)

    s proportional

    to the concentration gradient (VC):27

    where D, is the chem ical diffusion coeffi-

    cient, C is the concentration, and p is the

    chem ical potential. As noted above, diffu-

    sion can contribute to the shrinkage of gels

    in special cases (e.g., when the gel is im-

    merse d in a salt solution) and may be im -

    portant during evaporative drying, if a

    concentration gradient develops in the

    pore s by p referential evaporation of one

    component of the pore liquid.

    In some cases, a gradient in concen-

    tration of the solid phase can p roduce

    os-

    motic transport (as in the sw elling of som e

    organic polymers28 or clay29), but it is not

    clear whether transport occurs by diffusion

    or flow. One can com pare the fluxes giv-

    en by Eqs. (6) and

    (8)

    by converting the

    chemical potential gradient to pressure-

    volume work, then relating the diffusion

    coefficient to the viscosity by use of the

    Stokes-Einstein equation.1 The conclusion

    is that flow is faster than diffusion when-

    ever the pore diameter is more han a few

    times the diameter of the liquid molecule.

    Howe ver, this conclusion a pplies only

    to

    situations such as flow within a clay bod y

    (where tension in the liquid is produ ced

    by disjo ining orces), where there is a gra-

    dient in concentration of solid phase. Flow

    cannot reduce a concentration gradient in

    the liquid phase. For example, if a gel is

    immersed n a salt solution, flow of the

    so-

    lution into the pores does not affect the

    difference in salt concentration between

    the bath and the original pore liquid; that

    can be achieved only by diffusion. Simi-

    larly,

    if

    evaporation creates a concentra-

    tion gradient in the p ore liquid, flow from

    the interior of the gel cannot eliminate it;

    only interdiffusion within the po res can do

    I l l

    Stages of

    Drying

    The stages of drying were clearly

    discussed in the classic work of

    Shewmd30-32

    60

    years ago. Several texts

    prov ide qualitative desc riptions of the

    phenom enology and detailed discussion

    of the technology of drying,%-% he scien-

    tific aspects are discussed in several very

    good reviews (e.g., Refs. 8 and 36) and

    in the series of b ooks called Advance s in

    Dr~ing.37~38

    so.

    demonstrated. using nuclear m agnetic

    ( 1 )

    Constant

    Rate

    Period

    re so na nc gl or optical s p e c t r o ~ c o p y , ~ ~ ~ ~ ~

    he first stage of dryin g is called the

    constant rate period (CRP), because the

    rate of evaporatlon per unit area

    of

    the dry-

    ing surface is independent of time 7,8 The

    Corning

    Glass

    Works Corning

    NY

  • 8/20/2019 Drying Theory

    5/12

    January 1990

    Theory

    of

    Drying

    7

    evaporation rate is close to that from an

    open dish of liquid, as indicated, for ex-

    ample, by Dwivedi’s data for drying of alu-

    mina ge1.39 The rate of evaporation, V , , s

    proportional o the difference between

    pv

    and the ambient vapor pressure, pA:

    (9)

    where k s a factor that depends on the

    temperature, draft, and geometry of the

    system. The vapor pressure of the liquid

    is related to the capillary tension (P) by

    Pv Po exp(

    --PVm

    %IT

    From Eqs. ( l) , (9), and (10) we see that

    evaporation will continue as long as

    The fact that the evaporation rate is simi-

    lar to that of bulk liquid indicates that the

    vapor pressure reduction is insignificant

    during the CRP. However, in some gels

    the pores are so small that a significant

    reduction in pv could occur; moveover,

    the composition of the liquid in the pores

    could change with time if the initial liquid

    is a solution. The latter factors have been

    proposed

    to

    explain the absence of a CRP

    for an alkoxide-derived silica gel.4W

    It seems reasonable to conclude that

    the surface of the body must be covered

    with a film of liquid during the CRP, be-

    cause the proportion of the surface co-

    vered by menisci shrinks faster than the

    total area,

    so

    the rate would decrease as

    the body shrank if evaporation occurred

    only from the menisci. However, Suzuki

    and MaedaQ proved that the evaporation

    rate can remain constant even when dry

    patches form on the surface of the body.

    There is a stagnant (or slowly flowing)

    boundary layer of vapor over the drying

    surface, and

    if

    the breadth of the dry

    patches is small compared

    to

    the thick-

    ness of the layer, diffusion parallel

    to

    the

    surface homogenizes the boundary layer

    at the equilibrium concentration of vapor.

    This would certainly be expected in gels,

    where the expanse of dry solid phase be-

    tween menisci would be on the order

    of

    nanometers. Therefore, transport of vapor

    across the boundary layer obeys Eq. (9),

    and the rate of evaporation per unit area

    of surface is constant, whether or not there

    are small dry patches.

    Evaporation causes cooling of a body

    of liquid, but the reduced temperature

    leads to a lower rate of evaporation, and

    this feedback process equilibrates when

    the drying surface reaches the wet bulb

    temperature

    (T,).

    As indicated by Eq.

    (9),

    V ,

    ncreases as PA decreases,

    so

    T,

    decreases with the ambient humidity. The

    exterior surface of a drying body is at the

    wet bulb temperature during the CRP.3’

    The surface temperature rises only after

    the rate of evaporation decreases (in the

    falling rate period discussed in Section Ill

    (2)).

    For alkoxide-derived gels the vapor

    pressure must be kept high to avoid rap-

    id drying, so the temperature of the sam-

    ple remains near ambient.

    The tension in the liquid is supported by

    the solid phase, which therefore goes into

    compression. f the network is compliant,

    as it is in alkoxide-derived gels, the com-

    pressive forces cause it to contract into the

    liquid and the meniscus remains at the ex-

    terior surface, as indicated in Fig. 1 B). In

    a gel, it does not take much force to sub-

    merge the solid phase,

    so

    the capillary

    tension is low and the radius of the menis-

    cus is much larger than the pore radius.

    As drying proceeds, the network becomes

    increasingly stiff, because new bonds are

    forming and the porosity is decreasing; the

    meniscus deepens and the tension in the

    liquid rises correspondingly. Once the ra-

    dius of the meniscus becomes equal to

    the radius of the pores in the gel, the li-

    quid exerts the maximum possible force.

    That marks the end of the CRP: beyond

    that point the tension in the liquid cannot

    overcome further stiffening of the network,

    so

    the meniscus recedes into the pores,

    leaving air-filled pores near the outside of

    the gel (Fig. 1 C)). Thus, during the CRP,

    the shrinkage of the gel is equal

    to

    the vol-

    ume of liquid evaporated; the meniscus

    remains at the exterior surface, but r

    decreases continuously. This behavior is

    illustrated by the data of Kawaguchi etd . 4 3

    for alkoxide-derived gels; equivalent

    results have been reported for particulate

    gels made from fumed silica.44

    The end of the CRP is called the critical

    point (or leatherhard point, n clay technol-

    ogy), and it is at this point that shrinkage

    virtually stops.At the critical point, the ra-

    dius of curvature of the meniscus is small

    enough to enter the pores, so the capil-

    lary tension is found from Eqs. (3)and (4):

    For an alkoxide-derived gel with S-300 to

    800

    m*/g,

    eb-0.4

    to 1.6 glcm3, Q-0.2 to

    0.6, and yLv cos

    (0)-0.02

    to 0.07 J/m*,

    this is an enormous pressure: P p 3 to

    200

    MPa The amount of shrinkage that

    precedes the critical point depends on the

    magnitude of the maximum capillary

    stress,

    PR.

    Since PR increases with the in-

    terfacial energy

    (yLV)

    and with decreasing

    pore size, it is not surprising to find that

    the porosity of a dried body is greater (be-

    cause less shrinkage has occurred) when

    surfactants are added to the liquid. For ex-

    ample, Kingery and Franc19 found a line-

    ar proportionality between

    YLV

    and dried

    density for clay bodies mixed with surfac-

    tants. It is important to recognize, howev-

    er, that the pressure depends on the

    contact angle, and the surfactant could in-

    crease

    0

    while reducing

    y ~ v .

    he impor-

    tance of contact angle is nicely illustrated

    by the work of Mitsyuk et a1.45 They pre-

    pared aqueous silica gels from sodium sili-

    cate, then soaked them in various alcohols

  • 8/20/2019 Drying Theory

    6/12

    8 Jour

    Fig . 4. After

    the critical point, the li-

    quidlvapor meniscus retreats into the pores

    of

    the body. In the first falling rate period, liquid

    is

    in the funicular state, so transport by fluid

    flow

    is

    possible. There

    is also

    some diffusion

    in the vapor phase.

    Fig.

    5.

    During the second falling rate peri-

    od evaporation occurs inside the body,

    at

    the

    boundary between the funicular (continuous

    liquid) and pendular (isolated pockets of liquid)

    regions. Transpqrt in the pendular region oc-

    curs by diffusion.of vapor.

    wal of the American Ceramic Society Scherer

    Vol. 73, No.

    1

    (methanol, ethanol, 1 propanot, 1 butanol)

    to replace the pore liquid. When the gels

    were dried, the final porosity was found

    to be linearly related to the heat of wet-

    ting. The heat of wetting is related*

    to

    the

    quantity ysv-ysL=yLv cos

    (8); in

    this

    case, yLv is nearly the same for all the al-

    cohols, so the variation in capillary stress

    is caused by

    8.

    (2)

    First Falling

    Rate

    Period

    When shrinkage stops, further evapo-

    ration drives the meniscus into the body,

    as illustrated n Fig. 1(C); as air enters the

    pores, the surface may begin

    to

    lose its

    translucency.@ In the first falling rate peri-

    od (FRPI), the rate of evaporation

    decreases and the temperature of the sur-

    face rises above the wet bulb temperature.

    Most of the evaporation is still occurring

    at the exterior surface, so the surface re-

    mains below the ambient temperature,

    and the rate of evaporation is sensitive to

    the ambient temperature and vapor pres-

    sure.83 The liquid in the pores near the

    surface remains in the funicular condition,

    so

    there are contiguous pathways along

    which flow can cccur (Fig.

    4).

    At the same

    time, some liquid evaporates within the un-

    saturated pores and the vapor is transport-

    ed by diffusion. Analysis of this situation

    involves coupled equations for flow

    of

    heat

    and liquid and diffusion of vapor, with

    transport coefficients that are generally

    dependent on temperature and con-

    ,centration. There are several good

    revie~s36~47.48f the many theories that

    have been proposed o descibe the FRP1.

    'The most complete and rigorous treatment

    is by Whitaker.49,50

    Shaw51r52 performed an elegant series

    of experiments showing that the drying

    front (i.e., he liquidlvapor interface) s frac-

    tally rough on the scale of the pores, but

    stable on a much larger scale. It is the

    pressure gradient in the unsaturated re-

    gion that is responsible or the stability of

    the drying front: the capillary pressure is

    SO

    low in advanced regions of the front

    that the radius of the meniscus istoo large

    to pass through the pores. Since the ir-

    regularity in the drying front is on the scale

    of the pores, it is very small compared to

    tlhe dimensions of the body. Even in a par-

    ticulate gel with 60-nm pores.53 if a par-

    tially dried gel is broken in half, the drying

    front is visible as a sharp line between the

    translucent saturated region and the

    opaque dry region. No doubt this line

    would be rough if observed in the

    SEM,

    but it is quite smooth on a macroscopic

    scale.

    3) Second Falling Rate Period

    As the meniscus recedes into the body,

    the exterior does not become completely

    dry right away, because liquid continues

    to

    flow

    to

    the outside; as long as the flux

    of liquid is comparableto the evaporation

    ratel the funicular condition is preserved.

    However, as the distance from the exteri-

    or

    to the drying front increases, he capil-

    lary pressure gradient decreases and

    therefore so does the flux. Eventually (if the

    body is thick enough) it becomesso slow

    that the liquid near the outside of the body

    is isolated n pockets (i.e., enters the pen-

    dular condition), so flow to the surface

    stops and liquid s removed rom the body

    only by diffusion of its vapor. At this stage,

    drying is said to enter the second falling

    rate period (FRP2), where evaporation

    oc-

    curs inside the body (see Fig. 5 .31 The

    temperature of the surface approaches he

    ambient temperature and the rate of

    evaporation becomes ess sensitive to ex-

    ternal conditions (temperature, humidity,

    draft rate, etc.). As indicated n Fig.5, the

    drying front is drained by flow of funicular

    liquid which evaporates at the boundary

    of the funicularlpendular regions. In the

    pendular region, vapor is in equilibrium

    with isolated pockets of liquid and ad-

    sorbed films, and the principal transport

    process is expected to be diffusion of

    vapor.

    As

    the saturated region recedes into the

    body, the body expands slightly as the to-

    tal

    stress on the network is re lie~ed.32~43~~

    At the same time, differential strain builds

    up because the solid network is being

    compressed more in the saturated region

    than near the drying surface. This can

    cause warping in a plate dried from one

    side, as faster contraction of the wet side

    makes the plate convex toward the dry-

    ing side.% The fact that the warping is per-

    manent (i.e., does not spring back when

    drying is complete) indicates that the un-

    saturated region retains some viscosity or

    plasticity during FRP2. As the saturated re-

    gion becomes thinner, its contraction is

    more effectively prevented by the larger

    unsaturated region, and this raises the ten-

    sion in the network in the saturated region.

    This phenomenon probably accounts for

    the observation by Simpkins et

    a/.

    that

    cracks in drying gels often originated near

    the nondrying surface.

    Whitaker499 developed an analysis of

    heat and mass transfer during drying of

    rigid materials that offers the most com-

    plete description

    of

    the falling rate periods.

    He uses transport coefficients that are

    lo-

    cal averages for regions large compared

    to the pore size, but small compared to

    the sample. This is analogous

    o

    the aver-

    aging implicit in Darcy's law, where the

    permeability,

    D,

    "smears out" the ge-

    ometrical details of the microstructure.

    U s e

    of Whitaker's model requires knowledge

    of a large number of physical properties

    (permeability, thermal conductivity, diff u-

    sivity of vapor), and the analysis must be

    performed numerically.

    A

    successful test

    of the model was performed by Wei et

    a/.,55-56who studied the drying of porous

    sandstone.

    IV.

    Drying

    Stress

    If evaporation of liquid from a porous

    body exposed the solid network, a solid/

    vapor interface would appear where a

    solid/liquid nterface had been. This would

  • 8/20/2019 Drying Theory

    7/12

    January 1990

    Theory of Drying

    9

    raise the energy of the system, because

    ysv>vsL,

    so liquid tends to flow from the

    interior to prevent exposure of the solid.

    As it stretches toward the exterior, the

    li-

    quid goes into tension, and this has two

    consequences:

    1)

    liquid tends

    to

    flow

    from the interior along the pressure gra-

    dient, according to Darcy’s law; (2) the

    tension is balanced by compressive stress

    in the network that causes shrinkage. The

    lower the permeability, the more difficult

    it

    is to draw liquid from the inside of the

    body, and therefore the greater the pres-

    sure gradient that develops.

    As

    the pres-

    sure gradient increases, so does the

    variation in free strain rate, with the sur-

    face tending to contract faster than the in-

    terior. It is the differential strain (i.e., the

    spatial variation in strain (for an elastic ma-

    terial) or strain rate (for a viscous materi-

    al)) that produces stress. This is analogous

    to

    the development of thermal stresses in

    response to a temperature gradient, an

    observation that has been exploited by a

    number of authors.57-60 Just as calculation

    of thermal stresses requires knowledge of

    the temperature distribution, prediction of

    drying stresses depends on calculation of

    the pressure distribution, which we now

    explore.

    ( 1 ) Pressure Distribution

    If we consider an isolated region of a

    porous body, the rate

    of

    change of the vol-

    ume of liquid in that region depends on

    the divergence of the flux (i.e., the differ-

    ence between the flux entering and the

    flux leaving). During the CRP, when the

    pores are full of liquid, he change in liquid

    content must be equal to the change in

    pore volume,S which is. related

    to

    the

    volumetric strain rate,

    E .

    Setting these

    changes equal, we obtain the equation for

    continuity (conservation of matter):el

    We need

    to

    express in terms of the ten-

    sion in the liquid using a constitutive equa-

    tion for the network. Various authors have

    done this by using empirical (nonlinear

    elastic) equation~7~12r by assuming elas-

    tic behavior with the solid and liquid

    phases compressible57P or incompress-

    ible,63 a or allowing the network to be

    purely V~S CO US ,~ ~r viscoelastic.66-70 For

    the sake of discussion, we will employ the

    simpler elastic analysis. When the network

    is

    assumed

    to

    be elastic, Eq.

    13)

    has the

    mathematical orm of the diffusion equa-

    tion. For the CRP it

    is

    appropriate to in-

    troduce the boundary condition that the

    flux at the exterior surface is constant:

    $In his case, a pore is a space not occupied by

    the solid phase, which may be occupied by liquid

    andlor gas Duringthe CRP, there are no gas pock-

    ets, so a pore is full

    of

    liquid

    where

    V

    is the constant evaporation rate.

    For

    an elastic network Eq. 13) becomes

    In this equatior,

    L

    is the half-thickness of

    the drying plate, u =z /L is the coordinate

    normal to the drying surface, the dimen-

    sionless time is defined as

    O = flr

    and

    where

    Kp

    and

    GP

    are the bulk and shear

    moduli of the solid network (i.e., the

    properties that would be measured with

    the liquid drained away). By solving Eq.

    15)

    we obtain the pressure distribution

    in the drying body; the stresses and

    strains follow from the constitutive equa-

    tions.1 61

    2) Stress Distribution

    Philip12 discusses at length the

    methods for solving the nonlinear version

    of

    Eq.

    15)

    that results when the permea-

    bility and elastic properties vary with the

    porosity (and therefore with position n the

    body). In the simple case where the

    properties are constant and the shrinkage

    during drying is negligible, an analytical

    solution is readily obtained;63 typical

    results are shown in Fig. 6. The tension

    P in the liquid rises until at the critical point

    (time

    0)

    it

    reaches the maximum value

    at the exterior surface,

    P(L,OR)= PR,

    as

    shown in Fig. 6(A). If the evaporation rate

    is not too fast, the distribution through the

    plate becomes roughly parabolic at a

    much earlier stage (when O=eR/3 in Fig.

    6(C)) and, since the stress depends on

    the shape of the pressure distribution,a,

    is approximately constant (Figs. 6(B) and

    (D)) during the time interval OR/3

  • 8/20/2019 Drying Theory

    8/12

    10

    Journal of the American Ceramic Society Scherer

    Vol.

    73,

    No. 1

    distribution s parabolic, in which case Eq.

    (18) becomes61

    The stress increases in prop ortion

    to

    the

    thickness of the plate and the rate of

    evaporation, and in inverse proportion to

    the perm eability; that is, the stress

    s

    in-

    creased by those factors that steepen the

    pressure gradient. The reason that gels

    are so much more difficult to dry than

    conventional ceramics is that the perme-

    ability of gels is low, as a result of their

    very small pore size. Comparing the

    stress at the surface of a drying plate,

    cylinder, and sphe re, it is found71 that the

    tension decreases in the ratio plate/

    cylinderlsphere =; /a/ ; The lower stress

    reflects the shallower pressure gradients

    in the cylinder and sphere, where the

    li-

    quid flowing from the interior passes

    through a volum e that increases as P and

    r 3

    respectively. Since these results are

    derived from Eq.

    (13),

    hey are valid only

    as long as the pores rem ain filled with li-

    quid. At s ome point the netw ork will stop

    shrinking and the meniscus will retreat

    into the gel; then Eq. (19) will apply on ly

    within the saturated po res nsid e he ge1.63

    3) Diffusion

    If

    the po res conta in a solution of liqu ids

    with intrinsic diffusion coefficientsD , and

    D2,

    then diffusion contributes to the trans-

    port and the diffusion term m ust be a d-

    ded

    to

    the flow term. Then Eq.

    (13)

    becomes72

    h =

    v.

    P

    + I

    -@)V.

    i

    Note that d iffusion has no influence f the

    intrinsic diffusion coefficients of the two

    liquids are equal, because the diffusive

    volume fluxes are then equal and oppo-

    site (i.e., diffusion produces no volume

    flow). It has been shown72 that drying

    stresses can be reduced considerably

    when the diffusion term is significant. The

    reason is that a substantial flux can be

    produced by

    a

    shallow co ncen tration gra-

    dient (since interdiffusion of liquids is

    rapid ), so diffusion can e xtract liquid from

    the interior of the bod y almost as fast as

    it evaporates from the surface. Conse-

    quently, the pressure distribution is flat-

    ter, the differential strain is reduced, and

    the drying stresses are smaller when

    diffusion occurs.

    V. Fracture

    ( 1)

    Models of Fracture During Drying

    There is no generally accepted ex pla-

    nation for the phenom enon

    of

    cracking

    during drying. Any suitable he ory should

    a'ccount for the com mon observations

    that cracking is more likely

    if

    the body is

    thick or the drying rate is high, and that

    cracks generally appear at the critical

    point (i.e., when shrinkage stops and the

    vapor/liquid interface moves into the

    body

    of

    the gel). The tendency for slow-

    ly dried bodies

    to

    crack at the critical point

    has been noted for clay,35 particulate

    gels, l73 and alkoxide -deriv edgels.39174

    We no w examine two mod els of fracture

    during drying, a macroscopic model

    (described in Section IV) that attributes

    cracking

    to

    stresses produced b y a pres-

    sure gradien t in the liquid ph ase, and a

    microscopic mod el that explains crack-

    ing as a result of the distribution of p ore

    sizes.

    The stress that causes fracture is not

    the m acrosc opic stress,

    a

    that acts on

    the network. R ather, it is the stress con-

    centrated at the tip of a flaw of length c

    which is proportional to75

    Fracture occurs w hen o,>K\,, where KI,

    is a material property called the critical

    stress intensity.76 It is reasonable to as-

    sume that th e flaw size distribution is in-

    depen dent of the size and drying rate of

    the gel,

    so

    the tendency

    to

    fracture will

    increase with the stress,

    ax.

    Although Eq.

    (19) accounts qualitatively for the ob-

    served dependence of cracking on L and

    V . , t

    does not offer any explanation for

    the common observation that slowly

    dried ge ls crack at the critical point. The

    stress is p redic ted to rise continu ally until

    that moment, but there is no sudden ump

    predicted for a at time f3R that would en-

    hance the likelihood of cracking .

    The microscopic model for fracture is

    based on the idea77178 llustra ted in Fig .

    7. After the critical point, liquid s removed

    first from the largest pores; then the ten-

    sion in the neighboring small pores is

    claimed to deform the pore wall and

    cause cracking. This mechanism app ears

    to

    account quite clearly for the occur-

    rence of cracking at the critical point.

    How ever, he flaws produ ced in this way

    have lengths on the order of the space

    between pores, which is typically

    1

    to

    5

    nm in alkoxide-derived gels, and such

    flaws should be subcritical (i.e., non-

    propagating). This difficulty could be

    avoided b y suppo sing that the flaws per-

    colate through the structure until they

    achieve the critical length. A more im por-

    tant prob lem with this mechanism is that

    it does not explain the impo rtance of dry-

    ing rate or body size. The local stresses

    result from the local hetero geneity of the

    microstructure,

    so

    fracture should be in-

    evitable when the pore size distribution

    is wide.

    Another version of this mode l wou ld at-

    tribute the flaws to the irregu larity of the

    drying front, illustrated schematically in

    Fig. 8. The width

    of

    the drying front,

    w ,

    is 2 or 3 orders of m agnitude larger than

    the po re size, but the drying front is quite

    smooth on the scale of the thickness of

    the sample. The crack might be expect-

    ed to have a length similar

    to

    w ,

    so

    the

    stress intensity would be p roportional to

    ac= a x e (21)

  • 8/20/2019 Drying Theory

    9/12

    January

    1990

    PRwlQ, However, Shawsz has shown

    that

    w a p p ) - 1 ~ (VE)-1/2

    (22)

    which means that the drying front be-

    comes smoother (w decreases) as the

    drying rate increases. Thus

    (23)

    c

    =

    pRwl/2a VE)-1/4

    which means that the stress intensity

    decreases as the evaporation rate

    increases, in contradiction

    to

    the ex-

    perimental evidence. Further, no depen-

    dence of stress on sample size is

    expected according to this mechanism.

    On the other hand, if these flaws are act-

    ed upon by the stress predicted by the

    macroscopic mechanism, then the stress

    intensity is

    (241

    which increases almost in proportion

    to

    the drying rate. Thus, the flaws generat-

    ed by the irregular drying front, together

    with the macroscopic stress, may explain

    the appearance of cracking at the criti-

    cal point. The macroscopic nature of the

    stress also explains the observation that

    a drying body will often break into only

    two or three pieces;

    if

    the stresses were

    local, failure should always result in a very

    large number of fragments.

    2)

    Avoiding

    Fracture

    Since the capillary pressure sets the

    limit on the drying stress

    (o,OR);

    (F) normalized

    stress distribution at same times as in

    (E).

    From Ref.

    63.

    soaking f&

    24

    h in

    4N

    HCI or

    2N

    NH3.

  • 8/20/2019 Drying Theory

    10/12

      2

    Journal

    of

    the American Ceramic Society Scherer

    Vol.

    73,

    No. 1

    Constant rate period

    Fig.

    7.

    Illustration of microscopic model:

    during the constant rate period, meniscus has

    same radiusof curvature

    for

    poresof all sizes;

    after

    the

    critical point, the largest pores are

    emptied first.The capillary tension compress-

    ing the smaller pores causes local

    stresses

    that

    crack the network. After Ref. 77.

    Fig. 8. Drying front is fractally rough bound-

    ary between saturated (i.e., iquid-filled) and un-

    saturated regions. Flaw of length

    c is

    subjected

    to stress over width of drying front.

    and all of these features help to reduce

    cracking. The coarser structure may be

    a result of the higher pH produced by

    hydrolysis of formamide.84 Unfortunate-

    ly, the additive is difficult to remove upon

    heating, so bloating and cracking result.

    The original claims85 of rapid processing

    (-48

    h) for centimeter-thick pieces of gel

    processed with formamide have not been

    repeated nor reproduced, but promising

    results have been reported for dimethyl-

    formamide (DMF).86,87 That additive

    yields gels with larger pores, and they are

    even larger after aging at elevated tem-

    peratures (=150°C). Gels made with DMF

    do not crack at drying rates that destroy

    gels made with formamide, or without

    any DCCA. Interestingly, the dried gel

    cracks when exposed to vapors of water

    (yLv=0.072 J/m2) or formamide

    (0.058

    Jlmz), but not vapors of methanol

    (0.023

    Jlm2) or DMF

    (0.036

    J/m2), so the lower

    surface tension of the additive may be im-

    portant. To the extent that these additives

    are effective, their success can be at-

    tributed to coarsening of the microstruc-

    ture (which increases

    D

    and decreases

    PR) and strengthening of the network.

    They may also provide a medium through

    which the more volatile components

    (water and alcohol) can diffuse, thereby

    allowing diffusion to reduce the pressure

    differential within the body.72

    Since shrinkage and cracking are

    produced by capillary forces, KistleP

    reasoned that those problems could be

    avoided by removing the liquid from the

    pores above the critical temperature

    (T,)

    and critical pressure

    (Pc)

    of the liquid.

    Under such conditions there is no distinc-

    tion between the liquid and vapor phases:

    the densities become equal, there is no

    liquidhapor interface, and no capillary

    pressure. In the process of supercritical

    (or hypercritical) drying, a

    sol

    or wet gel

    is placed into an autoclave and heated

    along a path such as the one indicated

    iin Fig.

    9.

    The pressure and temperature

    ,are increased in such a way that the

    phase boundary is not crossed; once the

    critical point is passed, he solvent is vent-

    led

    at a constant temperature

    (>T,).

    The

    resulting gel, called an aerogel, has a

    volume similar

    to

    that of the original

    sol.

    This process makes it possible to

    produce monolithic gels as large as the

    volume of the autoclave. Table I contains

    values of T, and

    Pc

    for some relevant li-

    quids. Two groups succeeded at about

    the same time in making large monolithic

    gels by supercritical drying. In one case89

    the aerogel itself was the objective: the

    LOW

    density of the silica gel was required

    for a Cherenkov radiation detector.90 The

    other group91’92 wanted to make

    rnonolithic gels

    to

    be sintered nto dense

    glasses or ceramics, and found that large

    crack-free bodies could be made within

    vvide ranges of concentration of reac-

    tants. Although supercritical drying gives

    very good results or silica, the high tem-

    peratures and pressures make the

    process expensive and dangerous. A

    convenient alternative s to exchange the

    pore liquid for a substance with a much

    lower critical point. As shown in Table I,

    carbon dioxide has T, = 31“C and

    Pc=7.4 MPa,

    so

    the process can be

    performed near ambient temperatures.

    Supercritical drying following COP ex-

    change has become a standard tech-

    nique for preparing biological samples for

    TEM

    examination.93 t was apparently first

    applied for producing monolithic silica

    gels by Woignier,94 and was indepen-

    dently developed by Tewari

    ef

    al.95 for

    making large windows. For some materi-

    als, supercritical treatment in alcohol

    causes dissolution,

    so

    a milder process

    is essential. Brinker et al.96 used CO2 ex-

    change

    to

    make aerogels of lithium

    borate compositions that would dissolve

    in alcohol. This would seemto be an ideal

    way of making aerogels, but it does have

    some disadvantages. Long times can be

    required to achieve complete solvent ex-

    change, especiaily because

    C 0 2

    is not

    miscible with water (Kistler88 notes that li-

    quidlliquid interfaces ormed by rmmisci-

    ble liquids could produce capillary

    compression of the gel). It may be neces-

    sary to exchange first with a mutual

    sol-

    vent such as amyl acetate,96 then

    to

    flush

    for hours with liquid C02. Moreover, be-

    cause of the low density of the dried

    body, sintering of monolithic crystalline

    aerogels to full density is impractical.

    Another way

    of

    avoiding the presence

    of the liquidlvapor interface is

    to

    freeze

    the pore liquid and sublime the resulting

    solid under vacuum. This process

    of

    freeze-drying s widely used in the prepa-

    ration of foods,47 but does not permit the

    preparation of monolithic gels. The re a

    son is that the growing crystals reject the

    gel network, pushing t out of the way until

    it is stretched to the breaking point. It is

    this phenomenon that allows gels to be

    used as hosts for crystal growth:97198 the

    gel is so effectively excluded that the crys-

    tals nucleated in the pore liquid are not

    contaminated with the gel phase; the

    crystals can grow up to a size of a few

    millimeters before the strain is

    so

    great

    that macroscopic ractures appear in the

    gel. If a silica sol is frozen, flakes of silica

    gel (sometimes called lepidoidal silica) are

    produced;99 f freezing is done unidirec-

    tionally, fibers of gel are obtained.’OO*”J’

    Attempts to freeze-dry gels typically result

    in flakes (e.g., Ref. 102) or in translucent

    bodies with large pores that are the “fos-

    sils” of the crystals.

    VI. Conclusions

    Although the principles of drying have

    been recognized for decades, the means

    of calculating drying stresses and strains

    have been developed relatively recently.

    The stresses result from a gradient in the

    pressure in the liquid in the pores of the

    drying body. The stress increases with

    the drying rate and the size of the body,

    and is inversely related to the permeabil-

  • 8/20/2019 Drying Theory

    11/12

    January 1990

    ity of the structure. It is the latter factor that

    makes gels so much harder to dry than

    conventional ceramics: their small pores

    result in very low permeability. Fracture

    may result from the action

    of

    drying

    stresses on preexisting flaws, but in many

    cases seems to result from flaws gener-

    ated by the irregularity of the drying front

    as it enters the body at the critical point.

    Unfortunately, many of the physical

    properties needed to predict failure (e.g.,

    permeability and critical stress intensity

    of wet bodies) have not yet been

    measured.

    A

    variety of strategies have been de-

    vised

    to

    avoid fracture during drying.

    These include strengtheningof the solid

    network by aging or chemical additives,

    increasing permeability by increasing

    pore size, and reducing capiliary pres-

    sure by increasing pore size, reducing in-

    terfacial energies, or drying under

    supercritical conditions. Each of these ap-

    proaches nvolves some tradeoff (for ex-

    ample, in processing time or sintering

    temperature),

    so

    the best method must

    be chosen by regarding the process as

    a whole.

    Theory of Drying

    and Pore Structure. Academic Press, New York,

    1979.

    16H. Darcy. Les Fontaines Publiques de la Ville de

    Diion. Libraire des Corps lmperiaux des Ponts et

    Chaussees et des Mines, Paris, France, 1856.

    17J. van Brakel, "Pore Space Models for Trans-

    port Phenomena in Porous Media-Review and

    Evaluation with Special Emphasis on Capillary Liquid

    Transport," Powder Techno/., 11, 205-36 (1975)

    lap. Debye and R. L. Cleland, "Flow of Liquid

    Hydrocarbons in Porous VYCOR,"

    J.

    Appl. Phys.,

    30 [6] 843-49 (1959).

    19G. W. Scherer and R. M. Swiatek, "Measurement

    of Permeability: II. Silica Gels"; to be published in

    Non-Cryst. Solids.

    20K. K. Watson, "An Instantaneous Profile Method

    for Determining he Hydraulic Conductivity of Unsalu-

    rated Porous Materials," Water Resources Res.,

    2

    [4] 709-15 (1966).

    21K. J. Packer, "The Dynamics of Water in Heter-

    ogeneous Systems." Pbilos. Trans.

    R.

    Soc. London,

    Ser.

    8

    278. 59-87 (1977).

    22J. Warnock, D. D. Awschalom, and M. W. Shafer,

    "Orientational Behavior of Molecular Liquids in Res-

    tricted Geometries,"

    P h y s

    Rev. 6, 34 [I ] 475-78

    (1986).

    23M. W. Shafer,

    G.

    D. Awschalom. J. Warnock, and

    G Ruben, "The Chemistry of and Physics with

    Porous Sol-Gel Glasses,"

    J .

    Appl. Pbys., 61

    [12]

    5438-46 (1987).

    24D. Y. C. Chan and

    R.

    G. Horn, "The Drainage

    of Thin Liquid Films between Solid Surfaces,"

    J.

    Cbem. Pbys..

    83 [l o] 5311-24 (1985).

    25J. N. Israelachvili, "Measurement of the Viscosi-

    ty of Liquids in Very Thin Films,"

    J .

    Colloid lnterface

    26J. van Alsten. S. Granick. and J.

    N.

    Israelachvili,

    "Concerning the Measurement of Fluid Viscosity be-

    tween Curved Surfaces,"

    J. Colloid

    Interface Sci.,

    27J. Crank, Mathematics of Diffusion. Clarendon

    Press, Oxford, U.K..

    1975.

    2*T, Tanaka and

    D.

    .

    Fillmore, "Kinetics of Swell-

    ing of Gels,"

    J.

    Cbem. Phys.,70

    [3] 1214-18 (1979).

    29H. van Olphen, Introduction

    to

    Clay Colloid

    Chemistry. Interscience, London,

    1977.

    30T. K. Sherwood, "The Drying of Solids-I,"

    lnd.

    Eng. Cbem.,

    21 [I] 12-16 (1929).

    3 T K. Sherwood, "The Drying of Solids-11," lnd.

    Eng.

    Chem

    , 21 [lo] 976-80 (1929).

    32T. K. Sherwood, "The Drying of Solids-Il l,'' lnd.

    Eng.

    Cbem..

    22 (21 132-36 (1930).

    33R.W. Ford, Ceramics Drying. Pergamon Press,

    New York, 1986.

    W . B. Keey, Drying, Principles and Practice. Per-

    gamon Press, New York,

    1972

    "F. H. Clews, Heavy Clay Technology. Academ-

    ic Press, New York.

    1969.

    36M. Fortes and

    M.

    R .

    Okos, "Drying Theories:

    Their Bases and Limitations as Applied to Foods and

    Sci., 110 [ I ] 263-71 (1986).

    125 [2] 739-40 (1988).

    13

    References

    C . J. Brinker and G. W. Scherer, Sol-Gel Science.

    Academic Press, New York.

    ZL. R. White, "Capillary Rise in Powders,"

    J . Col-

    loid hterface Sci., 90 [2j 536-38 (1982)

    3G.

    W. Scherer, "Correction

    of

    'Drying Gels:

    I.

    The-

    ory,'"

    J .

    Non-Cryst. Solids, 92,

    375-82 (1987)

    4K.

    J.

    Packer, "The Dynamics

    of

    Water in Heter-

    ogeneous Systems," Pbilos. Trans.

    R.

    SOC.London,

    Ser. 8 ,

    278, 59-87 (1977).

    5J

    N.

    Israelachvili, "Measurements of Hydralion

    Forces between Macroscopic Surfaces," Chem.

    Scr.. 25, 7-14 (1985).

    6M. Brun,

    A.

    Lallemand,J. F. Quinson, and C. Ey-

    raud, "A New Method for the Simultaneous Deter-

    mination of the Size and the Shape of Pores: The

    Thermoporometry," Thermocbim. Acta,

    21, 59-88

    7H. H. Macey, "Clay-Water Relationships and the

    Internal Mechanism

    of

    Drying," Trans Br. Ceram.

    SOC.,

    41 [4] 73-121 (1942).

    8F.

    Moore, "The Mechanism of Moisture Move-

    ment in Clays with Particular Reference to Drying.

    A Concise Review,"

    Trans.

    6r. Ceram. Soc., 60,

    9W. D. Kingery and J. Francl, "Fundamental Study

    of Clay: XIII. Drying Behavior and Plastic Properties,"

    J. Am. Ceram. SOC.,37

    [12] 596-602 (1954).

    l0H. van Olphen. An Introduction o Clay Colloid

    Chemistry, 2d ed. Wiley, New York,

    1977.

    1lR. Q. Packard, "Moisture Stress in Unfired Cer-

    amic Clay Bodies,"

    J.

    Am. Ceram. Soc.. 50 [5]

    223-29 (1967).

    12J. R. Philip, "Theory

    of

    Infiltration," Adv.

    Hydrosci., 5,

    215-96 (1969).

    '33. Zarzycki, "Monolithic Xero- and Aerogels for

    Gel-Glass Processes"; pp .

    27-42

    in Ultrastructure

    Processing of Ceramics, Glasses, and Composites.

    Edited by L. L.Hench and D. R. Ulrich. Wiley, New

    York,

    1984.

    14A. E. Scheidegger, The Physics of Flow Through

    Porous Media, 3d ed. University of Toronto Press,

    Toronto, Canada, 1974.

    15F. A. L. Dullien, Porous Media-Fluid Transport

    (1977).

    517-39 (1961).

    I

    I

    Temperature Tc

    Fig. 9. Schematic phase diagram, indicat-

    ing temperature/pressure path followed dur-

    ing supercritical drying. At the critical point

    (T,,P,)

    the densities

    of

    the liquid and vapor

    phases are the same and the surface tension

    is zero.

    Table I: Critical Points of Selected Solvents

    Substance Formula T ("C) P, ( M W

    31.1 7.36

    19.7

    2.97

    Carbon dioxide CO2

    Freon

    11

    6 CF3CF3

    Methanol CH30H 240 7.93

    Ethanol C~H SO H 243 6.36

    Water H20 374

    22.0

    'Data collected by Tewari, Hunt, and Lofltus (Ref.

    95)

  • 8/20/2019 Drying Theory

    12/12

    14

    Journal of the American Ceramic Society Scherer

    Vol. 73, No. 1

    Grains , pp 119-54 in Advances in Drying, Vol. 1 .

    Edited by A. S. Mujumdar Hemisphere Publishing

    Corp., New York, 1980.

    37Advances in Drying, Vol. 1. Edited by A. S.

    Mulumdar. H emisphere Publishing Corp., N ew York.

    1980

    JeAdvances in Drying, Vol. 2. Edited by A.

    S.

    Mujumdar Hemisphere Publishing Corp ., New York,

    1983.

    3%

    K.

    Dwivedi, Drying Behavior of Alumina

    Gels.

    J.

    Mater. SC Lee .,

    5 ,

    373-76 (1986).

    40M

    J.

    R .

    Wilson, Drying Kinetics of Pure Silica

    Xerogels ; M S Thesis. University

    of

    Florida, Gaines-

    ville, FL, 1989

    "L. L. Hench and M. J. R. Wilson, Processmg

    of Gel-Silica Monoliths for Optic s: Drying Beha vior

    of Small Pore Gels ;

    to

    be published in Proceedings

    of the 5th International Workshop on Glasses and

    Ceramics from Geis, Rio de Janiero, 1989,J. Non-

    Cryst. Sohds.

    42M. Suzuki and

    S.

    Maeda, On the Mechanism

    of Drying of Granular Beds-Mass Transfer from Dis-

    continuousSource,

    J.

    Chem.

    Eng

    Jpn., 1

    [ I ]

    26-31

    1 968).

    43T. Kawag uchi, J. lura, N. Taneda . H. Hishikura,

    and Y. Kokubu, Structural Changes of Monolithic

    Silica Gel during the Gel-to-GlassTransition,

    J.

    Non-

    Cryst. Solids,

    82,

    50-56 (1986).

    44P. G. Simpkins, D. W. Johnson, Jr., and D. A.

    Flem ing, Drying Behavior of Colloidal Silica Gels,

    J . Am. Ceram. SOC.,

    72

    [l o] 1816-21 (1989).

    458. M. Mitsyuk, Z. Z. Vysotskii, and M. V. Poly-

    akov, The Part Played by the Polarity of the Inter-

    micellar Liquid and the Intensity of

    Its

    Interac tion with

    the Surface of Silicic Acid Hydrogel Particles in the

    Formation of S ilica Gel Textures, Dokl.

    Akad. Nauk

    SSSR, 155 16) 416-18 (1964).

    46A. W. Adamson, Physical Chemistry of Surfaces,

    p. 541 Interscience, New York, 1967.

    475 Bruin and K. Ch. A. M. Luybe n, Drying of

    Food Materials: A Review of Recent Developments ;

    pp. 155-21 5 in Advan ces in D rying, Vol. 1. Edited

    by A.

    S.

    Mujumdar. Hemisphere Publishing Corp.,

    New York. 1980.

    aJ.

    van Brakel, Mass Transfer in Convective Dry-

    ing ; pp 217-67 in Advances in Drying,

    Vol

    1. Edit-

    ed by A

    s.

    Mujumdar. Hemisphere Publishing

    Corp.,

    New

    York, 1980.

    49s Whitaker. Simultaneous Heat, Mass, and

    Mom entum Transfer in Porous M edia: A Theory of

    Drying, Adv.

    Heat

    Transfer, 13, 119-203 (1977).

    5 .

    Whitaker. Heat and MassTransfer n Granu-

    lar Porous Media ; pp. 23-61 in Advances in Dry-

    ing, Voi. 1 Edited by A. s. Mujumdar. Hemisphere

    Publishing Corp., New York, 1980.

    51T. M. Shaw, Mov eme nt of a Drying Front in a

    Porous Material ; pp

    215-23

    in Materials Research

    Society Sympos ia Proceedings, Vol. 73, Better Cer-

    amics Through Chemistry 11 Edited by C. J. Brinker,

    D . E. Clark, and D. R. Ulrich. M aterials Research So-

    ciety, Pittsburgh, PA, 1986.

    5ZT. M. Shaw, Dry ing as an Immiscible Displac e-

    ment Process with Fluid Counterflow, Phys. Rev.

    Len.,

    59

    [15] 1671-74 (1987).

    53G. W Sch erer, unreported work invo lving gels

    of the type described by G. W Scherer and J. C.

    Luong,

    J.

    Non-Cryst. Solids, 63, 163-72 (1984).

    SG. W. Scherer, Drying Gels

    Ill.

    Warping Plate,

    J.

    Non-Cryst. Solids, 91 83-1 00 1987).

    55C .

    Wei.

    H.

    T. Davis, E. A. Davis, and J . Gor-

    don, Hea t and Mass Transfer

    in

    Water-Laden Sand-

    stone: Convective Heating. AlChE J.,

    31

    [8]

    56C. K. Wei. H T. Davis, E. A Davis, and J. Gor-

    don, Hea t and Mass Transfer

    in

    Water-Laden Sand-

    1338-48 (1 985).

    stone: Microwave Heating, AfChE J . , 31 15) 842-48

    57J. Geertsma, A Remark on the Analogy be-

    tween Therm oelasticity and the E lasticity of Saturat-

    ed Porous Media,

    J .

    Mech.

    Phys.

    Solids,

    6 ,

    13-16

    58P. F. Less e, Osm otic Stress in Wood-Part

    I:

    The Analogy betw een Thermal and Swelling Stress,

    Wood Sci. Techno/.. 6 206-14 (1972); Osmotic

    Stress in Wood-Part II: On the Com putation of Dry-

    ing Stresses in Wood , /bid.,

    ,

    272-83 (1972).

    59A.

    R Cooper, Quantitative Theory

    of

    Cracking

    and Warping during the Drying of Clay Bodies ; pp.

    261-76 in Ceramics Processing Before Firing. Ed it-

    ed by G.

    Y

    Onoda, Jr., and L. L Hench. Wiley, New

    York, 1978.

    6 % W. Scherer. Drying Gels: II. Film and Flat

    Plate,

    J. Non-Cryst.

    Solids.

    89,

    217-38 (1987).

    61G. W. Scherer, Drying Gels:

    VIII.

    Revison and

    Review, J . Non-Cryst. Solids,

    109,

    171-82 (1989).

    62M. A Biot, Gen eral Theory of Three -

    Dimensional Consolidation, J . Appl. Phys., 12,

    6 %

    W. Scherer, Drying Gels:

    V.

    Rigid Gels, J .

    MT. Tanaka, L. 0. Hocker, and G. B Benedek,

    Spectrum of Light Scattered from a Viscoelastic

    Gel,

    J.

    Chem. Phys., 59 [9] 5151-59 (1973).

    65G.

    W.

    Scherer, Drying Gels:

    I

    General Theory.

    J. Non-Cryst. solids, 87, 199-225 (1986).

    66M. A Biot, Theory of Stress-Strain Rela tions n

    Aniso tropic Viscoelasticity and Relaxation Phenome-

    na, J . Appf. Phys, 25 [ l l ] 1385-91 (1954).

    67M. A. Biot, Gene ralized Theory of Acoust ic

    Propagation in Porous Dissipative Med ia,

    J.

    Acoust.

    SOC.A m . , 34 [9] 1254-64 (1962).

    68M. A. Biot, Mec hanics of Defocmation and

    Acoustic Propagation in Porous Media, J . Appl.

    Phys., 33 [4] 1482-98 (1962).

    6 %

    W. Scherer, Drying Gels: VI. Viscoelastic

    Plate, J. Non-Crysf. Solids.

    99,

    324-58

    1

    988).

    7oP. J. Banks. Theo ry of Con stant-R ate Expre s-

    sion and Subsequent Relaxation ; pp.

    102-108

    in

    Drying 85. Edited by

    R.

    Toei and A.

    S.

    Mujumdar.

    Hemisphere Publishing Corp., New York, 1985.

    71G. W. Schere r, Drying Gels. IV. Cylinde r and

    Sphere, J. Non-Crysf. Sohds,

    91,

    101-21 (1987).

    72G. W. Schere r, Drying Gels: VII. Diffusion du r-

    ing Drying,

    J.

    Non-Cryst. Solids,

    107,

    135-48

    73R. Clasen, Prepa ration and Sintering of High-

    Density Green Bodies to High-PuritySilica Glasses,

    J .

    Non-Cryst. Solids,

    89, 335-44 1987).

    74P. Anders on an d L. C. Klein, Shrink age of Lithi-

    um Aluminosilicate GelB during Dryin g,

    J.

    Non-

    Cryst. Solids. 93, 15-22 (1987).

    75A. A. Griffith, The Phenom enon of Ru pture and

    Flow in Solids, fh i los . Trans.

    R.

    SOC London. Ser.

    A,

    221,

    163-98 (1920).

    7 6 8 . R . Lawn and T.

    R.

    Wilshaw, Fracture of Brittle

    Solids. Cambridge University Press, Cambridge,

    U.K., 1975

    77J. Zarzy cki, M. Prassas, and J. Phalippo u. Syn-

    thesis of Glasses from

    Gels:

    The Problem of

    Monolithic Gels,

    J.

    Mater. Sci.,

    17, 3371-79 (1982).

    78L.

    L.

    Hench; pp. 52-64 In Science of C eramic

    Chemical Processing. Edited by L. L. Hench and D.

    R.

    Ulrich. Wiley, N ew York, 1986.

    79G. W. Scherer.

    S.

    A. Pa rdenek, and

    R.

    M. Swia-

    tek, Visco elasticity in Silica Gel, J. Non-Crysf.

    8OJ. K. West,

    R.

    hlikles, and G. Latorre, Correla-

    tions between Processing Parameters, U ltrastructure,

    and Strength of Gel-Silica ; pp. 219-24 in Materi-

    als Research Society Sym posia Proce edings, Vol.

    121,Better CeramicsThrough Chemistry 111. Edited

    by C. J. Brinker. D. E. Clark, and D.

    R.

    Ulrich. Materi-

    als Research Society, Pittsburgh, PA, 1988.

    8lJ. Zarzyck i, Critical Stress Intensity Factors of

    'Wet Gels,

    J. Non Crysf.

    Solids.

    100,359-63 (1988).

    (1985).

    (1957).

    155-64 (1941).

    Non-Cryst. SOlIdS. 92,

    122-44 (1987).

    (1989).

    Solids 107,

    14-22 (1988).

    82T. Mizuno, H. Nagata, and S. Manabe. At-

    tempts to Avoid Cracks during Drying, J. Non-Cryst.

    ass. H. Wang and L. L. Hench, Processing and

    Prop erties of Sol-Gel Derive d 20 mol% Na20-80

    mol%

    S i02

    (2ON)

    Materials ; pp.

    71 -77

    in Materials

    Research Society Symposia Proce edings , Vol. 32,

    Better Ceramics Through Chem istry. Edited by C.

    J.

    Brinker. D.

    E.

    Clark, and D.

    R.

    Ulrich. North-Holland

    Publishing Co., New York, 1984.

    84G. Orcel and L. L. Hench. Physical-Chemical

    Variables in Processing Na20-B203 -Si02 Gel

    Monoliths ; pp. 79-84 in Ma terials Research Socie-

    ty Sympos ia Proceedings, Vol. 32, Better Ceramics

    Through Chemistry. Edited by C. J. Brinker, D. E.

    Clark, and D.

    R.

    Ulrich. North-Holland Publishing

    Co., Ne w York, 1984.

    85s.

    Wallace and L. L. Hench, The Processing

    and Characterization of DCCA Modified Gel-Derive d

    Silica ; pp. 47-52 in Materials Research Society

    Symposia Proceedings,

    Vol.

    32, Better Ceramics

    Through Chemistry. Edited by C.

    J.

    Brinker, D.

    E.

    Clark, and D. R. Ulrich. North-Holland Publishing

    Co.. New York, 1984.

    86T. Ada chi and S. Sakka, Prepa ration of

    Monolithic Silica Gel and Glass by the Sol-Gel

    Method Using N,N-Dimethylformamide,

    J

    Mater.

    Sci..

    22,

    4407-10 (1987).

    87T. Adachi and S. Sakka, The Role of N,N-

    Dimethylformamide, a DCCA , in the Formation of Sil-

    ica Gel Monoliths by Sol-Gel Method,

    J.

    Non-Crysf.

    Solids,

    99,

    118-28 (1988).

    88s

    S Kistler, Coherent Expanded Aerogels,

    J.

    Phys. Chem., 36, 52-64 (1932).

    9s

    Henning and L . Svensson, Production of Sil-

    ica Aerogel, Phys.

    Scr.,

    23, 697-702 (1981).

    9 %

    Poelz, Aero gel in High Energy Physics ;

    pp . 176-87 in Aerogels. Edited by J. Fricke.

    Springer-Verlag, New York.

    91M. Prassas, S ynthesis of Gels in the Si0 2- Na 20

    System and of Mon olithic Silica Gels: Study of Their

    Con version into Glass (in Fr.); Thesis. University of

    Science and Eng ineering of Lan guedoc, Montpe lli-

    er, France, 1981.

    92M. Prassas, J. Ph alippou , and J. Zarzycki. Syn-

    thesis of M onolithic Silica Gels by Hypercritical Sol-

    vent Evacuation,

    J.

    Mafer.

    Sci.,19,

    1656-65 (1984).

    93A. A. Bartlett and H. P. Burstyn, A Review of

    the Physics of Critical Point Drying ; pp. 305-16 in

    Scanning Electron Microscopy 1975, Part I.

    ITT

    Research Institute, Chicago, IL. 1975.

    94T. Woignier, Contribution to Preparation of

    Glasses by the Gel Rou te (in Fr.); Thesis. Universi-

    ty of Science and Engineering

    01

    Languedoc, Mont-

    pellier, France, 1984.

    95P. H.

    Tewari, A

    J.

    Hunt, and

    K.

    D. Lofftus,

    Ambient-Temp erature Supercritical Drying of Trans-

    parent Silica Aerogels, Mater. Lett.,

    3

    [9,10]363-67

    1

    985).

    96c.

    J. Brinker,

    K.

    J. Ward, K. D . Keefer,

    E.

    Holup-

    ka. P. J. Bray, and

    R.

    K. Pearson, Synthesis and

    Structure of Borate Based Aerogels ; pp. 56-67 in

    Aerogels. Edited by J. F ricke. Springer-Verlag, New

    York, 1986.

    97H. K. He nisch, Crysta l Grow th in Gels. Pennsyl-

    vania State U niversity Press, University Park, PA,

    1970.

    9aM. C. Robett and F. Lefau cheu x, Crys tal

    Growth in Gels: P rinciple and Applications,

    J. Crysf

    99R.

    K. Iler, The Chemistry of Silica, pp.

    21-23.

    Wiley, New York,

    1979.

    1WW. Mahler and M. Bechtold, Freeze-Formed

    Silica Fibres, Nafure(London),

    285,

    27-28 (1980).

    1olT. Maki a nd

    S.

    Sakka, Formation of Alumina

    Fibers by Unidirectional Freezing of Ge l, J .

    Non-

    Cryst.

    Solids,

    82,

    239-45 (1986).

    1OzE. De gn Egebe rg and J. Engell, Freeze D ry-

    ing of Silica Gels Prepared from Siliciumethoxid,

    Rev.

    Phys.

    Appl..

    24,

    C4-23-C4-28 (1989).

    Solids, 100,

    236-40 (1988).

    GroMh,

    90,

    358-67 (1988).