Mass Transfer and Diffusion: Ass Transfer Is The Net Movement of A Component in A
Mass Transfer and Diffusion: Ass Transfer Is The Net Movement of A Component in A
Mass Transfer and Diffusion: Ass Transfer Is The Net Movement of A Component in A
66
3.1 Steady-State, Ordinary Molecular Diffusion 67
plane or stationary coordinate system. When a net flux the direction selected as positive. When the molecular and
occurs, it carries all species present. Thus, the molar flux of eddy-diffusion fluxes are in one direction and N is in the
an individual species is the sum of all three mechanisms. If opposite direction, even though a concentration difference
Ni is the molar flux of species i with mole fraction xi, and N or gradient of i exists, the net mass-transfer flux, Ni, of i can
is the total molar flux, with both fluxes in moles per unit time be zero.
per unit area in a direction perpendicular to a stationary In this chapter, the subject of mass transfer and diffusion
plane across which mass transfer occurs, then is divided into seven areas: (1) steady-state diffusion in
Ni = xi N + molecular diffusion flux of i stagnant media, (2) estimation of diffusion coefficients,
+ eddy diffusion flux of i (3-1) (3) unsteady-state diffusion in stagnant media, (4) mass
transfer in laminar flow, (5) mass transfer in turbulent flow,
where xiN is the bulk-flow flux. Each term in (3-1) is positive (6) mass transfer at fluid–fluid interfaces, and (7) mass
or negative depending on the direction of the flux relative to transfer across fluid–fluid interfaces.
lower to the upper region. After a long time, a dynamic equi- where, for convenience, the z subscript on J has been
librium will be achieved and the concentration of dye will be dropped, c = total molar concentration or molar density
uniform throughout the solution. Based on these observa- (c = 1/v = /M), and xA mole fraction of species A.
tions, it is clear that: Equation (3-4) can also be written in the following equiv-
alent mass form, where jA is the mass flux of A by ordinary
1. Mass transfer by ordinary molecular diffusion occurs
molecular diffusion relative to the mass-average velocity of
because of a concentration, difference or gradient; that
the mixture in the positive z-direction, is the mass density,
is, a species diffuses in the direction of decreasing
and wA is the mass fraction of A:
concentration.
2. The mass-transfer rate is proportional to the area normal dw A
jA = −DAB (3-5)
to the direction of mass transfer and not to the volume dz
of the mixture. Thus, the rate can be expressed as a flux.
Velocities in Mass Transfer
3. Net mass transfer stops when concentrations are
uniform. It is useful to formulate expressions for velocities of chemi-
cal species in the mixture. If these velocities are based on
the molar flux, N, and the molar diffusion flux, J, the molar
Fick’s Law of Diffusion
average velocity of the mixture, v M , relative to stationary
The above observations were quantified by Fick in 1855, who coordinates is given for a binary mixture as
proposed an extension of Fourier’s 1822 heat-conduction
N NA + NB
theory. Fourier’s first law of heat conduction is vM = = (3-6)
c c
dT Similarly, the velocity of species i, defined in terms of Ni, is
qz = −k (3-2)
dz relative to stationary coordinates:
where qz is the heat flux by conduction in the positive z- Ni
direction, k is the thermal conductivity of the medium, and vi = (3-7)
ci
dT/dz is the temperature gradient, which is negative in the
direction of heat conduction. Fick’s first law of molecular Combining (3-6) and (3-7) with xi = ci /c gives
diffusion also features a proportionality between a flux and a v M = xA vA + xB vB (3-8)
gradient. For a binary mixture of A and B,
Alternatively, species diffusion velocities, vi D , defined in
dcA terms of Ji, are relative to the molar-average velocity and are
JAz = −DAB (3-3a)
dz defined as the difference between the species velocity and
and the molar-average velocity for the mixture:
dcB
JBz = −DBA (3-3b) Ji
dz vi D = = vi − v M (3-9)
ci
where, in (3-3a), JAz is the molar flux of A by ordinary mol-
ecular diffusion relative to the molar-average velocity of the When solving mass-transfer problems involving net
mixture in the positive z direction, DAB is the mutual diffu- movement of the mixture, it is not convenient to use fluxes
sion coefficient of A in B, discussed in the next section, cA is and flow rates based on vM as the frame of reference. Rather,
the molar concentration of A, and dcA/dz is the concentra- it is preferred to use mass-transfer fluxes referred to station-
tion gradient of A, which is negative in the direction of ordi- ary coordinates with the observer fixed in space. Thus, from
nary molecular diffusion. Similar definitions apply to (3-3b). (3-9), the total species velocity is
The molar fluxes of A and B are in opposite directions. If the vi = v M + vi D (3-10)
gas, liquid, or solid mixture through which diffusion occurs
is isotropic, then values of k and DAB are independent of di- Combining (3-7) and (3-10),
rection. Nonisotropic (anisotropic) materials include fibrous
and laminated solids as well as single, noncubic crystals. Ni = ci v M + ci vi D (3-11)
The diffusion coefficient is also referred to as the diffusivity Combining (3-11) with (3-4), (3-6), and (3-7),
and the mass diffusivity (to distinguish it from thermal and
momentum diffusivities). nA dxA
Many alternative forms of (3-3a) and (3-3b) are used, NA = = xA N − cDAB (3-12)
A dz
depending on the choice of driving force or potential in the
gradient. For example, we can express (3-3a) as and
dxA nB dxB
JA = −cDAB (3-4) NB = = xB N − cDBA (3-13)
dz A dz
3.1 Steady-State, Ordinary Molecular Diffusion 69
Mole fraction, x
Mole fraction, x
xB xB
where in (3-12) and (3-13), ni is the molar flow rate in moles Thus, in the steady state, the mole fractions are linear in dis-
per unit time, A is the mass-transfer area, the first terms on the tance, as shown in Figure 3.1a. Furthermore, because c is
right-hand sides are the fluxes resulting from bulk flow, and constant through the film, where
the second terms on the right-hand sides are the ordinary mol-
ecular diffusion fluxes. Two limiting cases are important: c = cA + cB (3-20)
Thus,
Equimolar Counterdiffusion
dcA = −dcB (3-22)
In equimolar counterdiffusion (EMD), the molar fluxes of A
and B in (3-12) and (3-13) are equal but opposite in direc- From (3-3a), (3-3b), (3-15), and (3-22),
tion; thus,
DAB DBA
= (3-23)
N = NA + NB = 0 (3-14) dz dz
Thus, from (3-12) and (3-13), the diffusion fluxes are also Therefore, DAB = DBA.
equal but opposite in direction: This equality of diffusion coefficients is always true in a
binary system of constant molar density.
JA = −JB (3-15)
N2 over the length of the tube, The factor (1 − xA ) accounts for the bulk-flow effect. For a
mixture dilute in A, the bulk-flow effect is negligible or
cDN2 , H2
n N2 = [(xN2 ) 1 − (xN2 ) 2 ]A small. In mixtures more concentrated in A, the bulk-flow
z2 − z1
(4.09 × 10−5 )(0.784)(0.80 − 0.25)
effect can be appreciable. For example, in an equimolar
= (7.85 × 10−3 ) mixture of A and B, (1 − xA ) = 0.5 and the molar mass-
15
= 9.23 × 10−9 mol/s in the positive z-direction transfer flux of A is twice the ordinary molecular-diffusion
flux.
n H2 = 9.23 × 10−9 mol/s in the negative z-direction
For the stagnant component, B, (3-13) becomes
(b) For equimolar counterdiffusion, the molar-average velocity of dxB
the mixture, vM, is 0. Therefore, from (3-9), species velocities are 0 = xB NA − cDBA (3-28)
dz
equal to species diffusion velocities. Thus,
or
JN n N2
vN2 = ( vN2 ) D = 2 =
xB NA = cDBA
dxB
(3-29)
cN2 AcxN2
dz
9.23 × 10−9
= Thus, the bulk-flow flux of B is equal but opposite to its
[(7.85 × 10−3 )(4.09 × 10−5 )xN2 ]
0.0287 diffusion flux.
= in the positive z-direction
xN2 At quasi-steady-state conditions, that is, with no accumu-
lation, and with constant molar density, (3-27) becomes in
Similarly, integral form:
0.0287 z
vH2 = in the negative z-direction cDAB xA dxA
xH2 dz = − (3-30)
z1 NA xA1 1 − xA
Thus, species velocities depend on species mole fractions, as
which upon integration yields
follows:
vN2 , cm/s vH2 , cm/s cDAB 1 − xA
z, cm xN2 xH2 NA = ln (3-31)
z − z1 1 − xA1
0 (end 1) 0.800 0.200 0.0351 −0.1435
5 0.617 0.383 0.0465 −0.0749 Rearrangement to give the mole-fraction variation as a func-
10 0.433 0.567 0.0663 −0.0506
tion of z yields
15 (end 2) 0.250 0.750 0.1148 −0.0383
NA (z − z 1 )
Note that species velocities vary across the length of the connect- xA = 1 − (1 − xA1 ) exp (3-32)
ing tube, but at any location, z, vM = 0. For example, at z = 10 cm, cDAB
from (3-8),
Thus, as shown in Figure 3.1b, the mole fractions are non-
v M = (0.433)(0.0663) + (0.567)(−0.0506) = 0 linear in distance.
An alternative and more useful form of (3-31) can be
derived from the definition of the log mean. When z = z 2 ,
(3-31) becomes
Unimolecular Diffusion
In unimolecular diffusion (UMD), mass transfer of compo- cDAB 1 − xA2
NA = ln (3-33)
nent A occurs through stagnant (nonmoving) component B. z2 − z1 1 − xA1
Thus,
The log mean (LM) of (1 − xA ) at the two ends of the stag-
NB = 0 (3-24) nant layer is
and
(1 − xA2 ) − (1 − xA1 )
N = NA (3-25) (1 − xA ) LM =
ln[(1 − xA2 )/(1 − xA1 )]
xA1 − xA2 (3-34)
Therefore, from (3-12), =
ln[(1 − xA2 )/(1 − xA1 )]
dxA
NA = xA NA − cDAB (3-26) Combining (3-33) with (3-34) gives
dz
which can be rearranged to a Fick’s-law form, cDAB (xA1 − xA2 ) cDAB (−xA )
NA = =
z 2 − z 1 (1 − xA ) LM (1 − xA ) LM z
(3-35)
cDAB dxA cDAB dxA cDAB (−xA )
NA = − =− (3-27) =
(1 − xA ) dz xB dz (xB ) LM z
3.1 Steady-State, Ordinary Molecular Diffusion 71
As shown in Figure 3.2, an open beaker, 6 cm in height, is filled xA = 1 − 0.869 exp(0.281 z) (1)
with liquid benzene at 25◦ C to within 0.5 cm of the top. A gentle Using (1), the following results are obtained:
breeze of dry air at 25◦ C and 1 atm is blown by a fan across the
mouth of the beaker so that evaporated benzene is carried away by z, cm xA xB
convection after it transfers through a stagnant air layer in the 0.0 0.1310 0.8690
beaker. The vapor pressure of benzene at 25◦ C is 0.131 atm. The 0.1 0.1060 0.8940
mutual diffusion coefficient for benzene in air at 25◦ C and 1 atm is 0.2 0.0808 0.9192
0.0905 cm2/s. Compute: 0.3 0.0546 0.9454
(a) The initial rate of evaporation of benzene as a molar flux in 0.4 0.0276 0.9724
mol/cm2-s 0.5 0.0000 1.0000
(b) The initial mole-fraction profiles in the stagnant air layer These profiles are only slightly curved.
(c) The initial fractions of the mass-transfer fluxes due to molecu- (c) From (3-27) and (3-29), we can compute the bulk flow terms,
lar diffusion xANA and xBNA, from which the molecular diffusion terms are
(d) The initial diffusion velocities, and the species velocities (rela- obtained.
tive to stationary coordinates) in the stagnant layer
(e) The time in hours for the benzene level in the beaker to drop xiN Ji
2 cm from the initial level, if the specific gravity of liquid ben- Bulk-Flow Flux, Molecular-Diffusion
zene is 0.874. Neglect the accumulation of benzene and air in mol/cm2-s × 106 Flux, mol/cm2-s × 106
the stagnant layer as it increases in height
z, cm A B A B
xA = 0 vid Ji
Mass Molecular-Diffusion Species Velocity,
transfer 0.5 cm
xA = PAs /P
z Velocity, cm/s cm/s
Interface
z, cm A B A B
6 cm
Liquid 0.0 0.1687 −0.0254 0.1941 0
Benzene 0.1 0.2145 −0.0254 0.2171 0
0.2 0.2893 −0.0254 0.3147 0
0.3 0.4403 −0.0254 0.4657 0
Beaker
0.4 0.8959 −0.0254 0.9213 0
0.5 ∞ −0.0254 ∞ 0
Figure 3.2 Evaporation of benzene from a beaker—Example 3.2.
72 Chapter 3 Mass Transfer and Diffusion
Note that vB is zero everywhere, because its molecular-diffusion Table 3.1 Diffusion Volumes from Fuller,
velocity is negated by the molar-mean velocity. Ensley, and Giddings [J. Phys. Chem, 73,
(e) The mass-transfer flux for benzene evaporation can be equated 3679–3685 (1969)] for Estimating Binary Gas
to the rate of decrease in the moles of liquid benzene per unit cross Diffusivity by the Method of Fuller et al. [3]
section of the beaker. Letting z = distance down from the mouth of Atomic Diffusion Volumes Atomic
the beaker and using (3-35) with z = z, and Structural Diffusion-Volume Increments
DABP/(DABP)LP
1.5
Air—carbon dioxide 317.2 0.177 1.4
1.3
Air—ethanol 313 0.145 0.6
Air—helium 317.2 0.765 0.9 1.2
EXAMPLE 3.3
EXAMPLE 3.4
Estimate the diffusion coefficient for the system oxygen (A)/
benzene (B) at 38◦ C and 2 atm using the method of Fuller et al. Estimate the diffusion coefficient for a 25/75 molar mixture of argon
and xenon at 200 atm and 378 K. At this temperature and 1 atm, the
SOLUTION diffusion coefficient is 0.180 cm2/s. Critical constants are
RT where the units are cm2/s for DAB; cP (centipoises) for the
( DAB ) ∞ = (3-38)
6B RA NA solvent viscosity, B ; K for T; and cm3/mol for vA, the liquid
molar volume of the solute at its normal boiling point. The
where RA is the radius of the solute molecule and NA is parameter B is an association factor for the solvent, which
Avagadro’s number. Although (3-38) is very limited in its is 2.6 for water, 1.9 for methanol, 1.5 for ethanol, and 1.0 for
application to liquid mixtures, it has long served as a starting unassociated solvents such as hydrocarbons. Note that the
point for more widely applicable empirical correlations for effects of temperature and viscosity are identical to the pre-
the diffusivity of solute (A) in solvent (B), where both A and diction of the Stokes–Einstein equation, while the effect of
B are of the same approximate molecular size. Unfortu- the radius of the solute molecule is replaced by vA, which
nately, unlike the situation in binary gas mixtures, DAB = can be estimated by summing the atomic contributions in
DBA in binary liquid mixtures can vary greatly with compo- Table 3.3, which also lists values of vA for dissolved light
sition as shown in Example 3.7. Because the Stokes– gases. Some representative experimental values of diffusivity
Einstein equation does not provide a basis for extending in dilute binary liquid solutions are given in Table 3.4.
Table 3.3 Molecular Volumes of Dissolved Light Gases and Atomic Contributions for Other Molecules at the
Normal Boiling Point
Source: G. Le Bas, The Molecular Volumes of Liquid Chemical Compounds, David McKay, New York (1915).
3.2 Diffusion Coefficients 75
Table 3.4 Experimental Binary Liquid Diffusivities for Solutes, equation with experimental values for nonaqueous solutions.
A, at Low Concentrations in Solvents, B For a dilute solution of one normal paraffin (C5 to C32) in
another (C5 to C16),
Diffusivity,
Solvent, Solute, Temperature, DAB,
B A K cm2 /s × 105 T 1.47 B
( DAB ) ∞ = 13.3 × 10−8 (3-40)
Water Acetic acid 293 1.19 vA0.71
EXAMPLE 3.6
Estimate the diffusivity of formic acid (A) in benzene (B) at 25◦ C However, because formic acid is an organic acid, A is doubled
and infinite dilution, using the appropriate correlation of Hayduk to 187.4.
and Minhas [7]. The experimental value is 2.28 × 10−5 cm2/s. From (3-42),
SOLUTION 2981.29 (205.30.5 /187.40.42 )
( DAB ) ∞ = 1.55 × 10−8
0.60.92 960.23
Equation (3-42) applies, with T = 298 K
= 2.15 × 10−5 cm2 /s
A = 93.7 cm3 -g1/4 /s1/2 -mol B = 205.3 cm3 -g1/4 /s1/2 -mol
B = 0.6 cP at 25◦ C vB = 96 cm3 /mol at 80◦ C which is within 6% of the the experimental value.
3.2 Diffusion Coefficients 77
Diffusivity, DAB,
Protein MW Configuration Temperature, °C cm2/s × 105
Bovine serum albumin 67,500 globular 25 0.0681
–Globulin, human 153,100 globular 20 0.0400
Soybean protein 361,800 globular 20 0.0291
Urease 482,700 globular 25 0.0401
Fibrinogen, human 339,700 fibrous 20 0.0198
Lipoxidase 97,440 fibrous 20 0.0559
crystalline, or amorphous; and the type of solid material, mechanisms or combinations thereof may take place:
whether it be metallic, ceramic, polymeric, biological, or
1. Ordinary molecular diffusion through pores, which
cellular. Crystalline materials may be further classified ac-
present tortuous paths and hinder the movement of
cording to the type of bonding, as molecular, covalent, ionic,
large molecules when their diameter is more than 10%
or metallic, with most inorganic solids being ionic. How-
of the pore diameter
ever, ceramic materials can be ionic, covalent, or most often
a combination of the two. Molecular solids have relatively 2. Knudsen diffusion, which involves collisions of diffus-
weak forces of attraction among the atoms or molecules. In ing gaseous molecules with the pore walls when the
covalent solids, such as quartz silica, two atoms share two or pore diameter and pressure are such that the molecular
more electrons equally. In ionic solids, such as inorganic mean free path is large compared to the pore diameter
salts, one atom loses one or more of its electrons by transfer 3. Surface diffusion involving the jumping of molecules,
to one or more other atoms, thus forming ions. In metals, adsorbed on the pore walls, from one adsorption site to
positively charged ions are bonded through a field of elec- another based on a surface concentration-driving force
trons that are free to move. Unlike diffusion coefficients in 4. Bulk flow through or into the pores
gases and low-molecular-weight liquids, which each cover a
range of only one or two orders of magnitude, diffusion co- When treating diffusion of solutes in porous materials
efficients in solids cover a range of many orders of magni- where diffusion is considered to occur only in the fluid in the
tude. Despite the great complexity of diffusion in solids, pores, it is common to refer to an effective diffusivity, Deff,
Fick’s first law can be used to describe diffusion if a mea- which is based on (1) the total cross-sectional area of the
sured diffusivity is available. However, when the diffusing porous solid rather than the cross-sectional area of the pore
solute is a gas, its solubility in the solid must also be known. and (2) on a straight path, rather than the pore path, which
If the gas dissociates upon dissolution in the solid, the may be tortuous. If pore diffusion occurs only by ordinary
concentration of the dissociated species must be used in molecular diffusion, Fick’s law (3-3) can be used with an
Fick’s law. In this section, many of the mechanisms of diffu- effective diffusivity. The effective diffusivity for a binary
sion in solids are mentioned, but because they are exceed- mixture can be expressed in terms of the ordinary diffusion
ingly complex to quantify, the mechanisms are considered coefficient, DAB, by
only qualitatively. Examples of diffusion in solids are con- DAB
sidered, together with measured diffusion coefficients that Deff = (3-49)
can be used with Fick’s first law. Emphasis is on diffusion of
gas and liquid solutes through or into the solid, but move- where is the fractional porosity (typically 0.5) of the solid
ment of the atoms, molecules, or ions of the solid through it- and is the pore-path tortuosity (typically 2 to 3), which is
self is also considered. the ratio of the pore length to the length if the pore were
straight in the direction of diffusion. The effective diffusivity
is either determined experimentally, without knowledge of
Porous Solids
the porosity or tortuosity, or predicted from (3-49) based on
When solids are porous, predictions of the diffusivity of measurement of the porosity and tortuosity and use of the
gaseous and liquid solute species in the pores can be made. predictive methods for ordinary molecular diffusivity. As an
These methods are considered only briefly here, with details example of the former, Boucher, Brier, and Osburn [13] mea-
deferred to Chapters 14, 15, and 16, where applications are sured effective diffusivities for the leaching of processed soy-
made to membrane separations, adsorption, and leaching. This bean oil (viscosity = 20.1 cP at 120◦F) from 1/16-in.-thick
type of diffusion is also of great importance in the analysis and porous clay plates with liquid tetrachloroethylene solvent.
design of reactors using porous solid catalysts. It is sufficient The rate of extraction was controlled by the rate of diffusion
to mention here that any of the following four mass-transfer of the soybean oil in the clay plates. The measured value of
80 Chapter 3 Mass Transfer and Diffusion
Deff was 1.0 × 10−6 cm2/s. As might be expected from the Table 3.10 Diffusivities of Solutes in Crystalline Metals
effects of porosity and tortuosity, the effective value is about and Salts
one order of magnitude less than the expected ordinary mol-
Metal/Salt Solute T, °C D, cm2/s
ecular diffusivity, D, of oil in the solvent.
Ag Au 760 3.6 × 10−10
Crystalline Solids Sb 20 3.5 × 10−21
Sb 760 1.4 × 10−9
Diffusion through nonporous crystalline solids depends
Al Fe 359 6.2 × 10−14
markedly on the crystal lattice structure and the diffusing
Zn 500 2 × 10−9
entity. As discussed in Chapter 17 on crystallization, only
Ag 50 1.2 × 10−9
seven different lattice structures are possible. For the cubic
lattice (simple, body-centered, and face-centered), the dif- Cu Al 20 1.3 × 10−30
fusivity is the same in all directions (isotropic). In the six Al 850 2.2 × 10−9
other lattice structures (including hexagonal and tetragonal), Au 750 2.1 × 10−11
the diffusivity can be different in different directions Fe H2 10 1.66 × 10−9
(anisotropic). Many metals, including Ag, Al, Au, Cu, Ni, H2 100 1.24 × 10−7
Pb, and Pt, crystallize into the face-centered cubic lattice C 800 1.5 × 10−8
structure. Others, including Be, Mg, Ti, and Zn, form Ni H2 85 1.16 × 10−8
anisotropic, hexagonal structures. The mechanisms of diffu- H2 165 1.05 × 10−7
sion in crystalline solids include: CO 950 4 × 10−8
1. Direct exchange of lattice position by two atoms or W U 1727 1.3 × 10−11
ions, probably by a ring rotation involving three or AgCl Ag +
150 2.5 × 10−14
more atoms or ions Ag+ 350 7.1 × 10−8
2. Migration by small solutes through interlattice spaces Cl− 350 3.2 × 10−16
called interstitial sites KBr H2 600 5.5 × 10−4
Br2 600 2.64 × 10−4
3. Migration to a vacant site in the lattice
4. Migration along lattice imperfections (dislocations),
or gain boundaries (crystal interfaces)
Diffusion coefficients associated with the first three EXAMPLE 3.9
mechanisms can vary widely and are almost always at least
one order of magnitude smaller than diffusion coefficients in Gaseous hydrogen at 200 psia and 300◦ C is stored in a small,
low-viscosity liquids. As might be expected, diffusion by the 10-cm-diameter, steel pressure vessel having a wall thickness of
fourth mechanism can be faster than by the other three 0.125 in. The solubility of hydrogen in steel, which is proportional
to the square root of the hydrogen partial pressure in the gas, is
mechanisms. Typical experimental diffusivity values, taken
equal to 3.8 × 10−6 mol/cm3 at 14.7 psia and 300◦ C. The diffusiv-
mainly from Barrer [14], are given in Table 3.10. The diffu- ity of hydrogen in steel at 300◦ C is 5 × 10−6 cm2/s. If the inner sur-
sivities cover gaseous, ionic, and metallic solutes. The val- face of the vessel wall remains saturated at the existing hydrogen
ues cover an enormous 26-fold range. Temperature effects partial pressure and the hydrogen partial pressure at the outer sur-
can be extremely large. face is zero, estimate the time, in hours, for the pressure in the ves-
sel to decrease to 100 psia because of hydrogen loss by dissolving
Metals in and diffusing through the metal wall.
Differentiating (2) with respect to time, For both hydrogen and helium, diffusivities increase rapidly
dn A n Ao dpA with increasing temperature. At ambient temperature the
= (3)
dt pAo dt diffusivities are three orders of magnitude lower than in liq-
Combining (1) and (3), uids. At elevated temperatures the diffusivities approach
dpA DA A(3.8 × 10−6 ) pA0.5 pAo
those observed in liquids. Solubilities vary only slowly with
=− (4) temperature. Hydrogen is orders of magnitude less soluble
dt n Ao z(14.7) 0.5
in glass than helium. For hydrogen, the diffusivity is some-
Integrating and solving for t,
what lower than in metals. Diffusivities for oxygen are also
2n Ao z(14.7) 0.5 0.5
included in Table 3.11 from studies by Williams [17] and
t= pAo − pA0.5
3.8 × 10 DA ApAo
−6
Sucov [18]. At 1000◦ C, the two values differ widely be-
Assuming the ideal-gas law, cause, as discussed by Kingery, Bowen, and Uhlmann [19],
(200/14.7)[(3.14 × 103 )/6)] in the former case, transport occurs by molecular diffusion;
n Ao = = 0.1515 mol
82.05(300 + 273) while in the latter case, transport is by slower network diffu-
The mean-spherical shell area for mass transfer, A, is sion as oxygen jumps from one position in the silicate net-
3.14
work to another. The activation energy for the latter is much
A= [(10) 2 + (10.635) 2 ] = 336 cm2 larger than for the former (71,000 cal/mol versus 27,000
2
cal/mol). The choice of glass can be very critical in high-
The time for the pressure to drop to 100 psia is
vacuum operations because of the wide range of diffusivity.
2(0.1515)(0.125 × 2.54)(14.7) 0.5
t= (2000.5 − 1000.5 )
3.8 × 10−6 (5 × 10−6 )(336)(200)
Ceramics
= 1.2 × 106 s = 332 h
Diffusion rates of light gases and elements in crystalline
Silica and Glass ceramics are very important because diffusion must precede
chemical reactions and causes changes in the microstructure.
Another area of great interest is the diffusion of light gases Therefore, diffusion in ceramics has been the subject of
through various forms of silica, whose two elements, Si and numerous studies, many of which are summarized in
O, make up about 60% of the earth’s crust. Solid silica can Figure 3.4, taken from Kingery et al. [19], where diffusivity
exist in three principal crystalline forms (quartz, tridymite, is plotted as a function of the inverse of temperature in the
and cristobalite) and in various stable amorphous forms, high-temperature range. In this form, the slopes of the
including vitreous silica (a noncrystalline silicate glass or curves are proportional to the activation energy for diffu-
fused quartz). Table 3.11 includes diffusivities, D, and solu- sion, E, where
bilities as Henry’s law constants, H, at 1 atm for helium and
E
hydrogen in fused quartz as calculated from correlations of D = Do exp − (3-51)
experimental data by Swets, Lee, and Frank [15] and Lee RT
[16], respectively. The product of the diffusivity and the sol- An insert at the middle-right region of Figure 3.4 relates the
ubility is called the permeability, PM. Thus, slopes of the curves to activation energy. The diffusivity
PM = D H (3-50) curves cover a ninefold range from 10−6 to 10−15 cm2/s,
with the largest values corresponding to the diffusion of
Unlike metals, where hydrogen usually diffuses as the potassium in
-Al2 O3 and one of the smallest values for car-
atom, hydrogen apparently diffuses as a molecule in glass. bon in graphite. In general, the lower the diffusivity, the
higher is the activation energy. As discussed in detail by
Table 3.11 Diffusivities and Solubilities of Gases in Amorphous Kingery et al. [19], diffusion in crystalline oxides depends
Silica at 1 atm not only on temperature but also on whether the oxide is stoi-
chiometric or not (e.g., FeO and Fe0.95O) and on impurities.
Gas Temp, C Diffusivity, cm2/s Solubility mol/cm3-atm Diffusion through vacant sites of nonstoichiometric oxides
He 24 2.39 × 10−8 1.04 × 10−7 is often classified as metal-deficient or oxygen-deficient.
300 2.26 × 10−6 1.82 × 10−7 Impurities can hinder diffusion by filling vacant lattice or
500 9.99 × 10−6 9.9 × 10−8 interstitial sites.
1,000 5.42 × 10−5 1.34 × 10−7
H2 300 6.11 × 10−8 3.2 × 10−14 Polymers
500 6.49 × 10−7 2.48 × 10−13
Thin, dense, nonporous polymer membranes are widely
1,000 9.26 × 10−6 2.49 × 10−12
used to separate gas and liquid mixtures. As discussed in
O2 1,000 6.25 × 10−9
detail in Chapter 14, diffusion of gas and liquid species
(molecular)
through polymers is highly dependent on the type of poly-
1,000 9.43 × 10−15
mer, whether it be crystalline or amorphous and, if the latter,
(network)
glassy or rubbery. Commercial crystalline polymers are
82 Chapter 3 Mass Transfer and Diffusion
Temperature, °C
–5
1716 1393 1145 977 828 727
10
K in β – Al2O3
10 –6
Co in CoO
(air)
10 –7 Fe in Fe0.95
O in Ca0.14 Zr0.86O1.86
10 –8 O in Y2O3
Mg in MgO Cr in Cr2O3
(single)
Diffusion coefficient, cm2/s
10 –9 Y in Y2O3
O in Cu2O
N
(Po2 = 14 kPa)
in
U
N
O in CoO O in Cu2O
10 –10 (Po2 = 20 kPa) (Po2 = 20 kPa)
763
382
191
Al in O in Ni0.68Fe2.32O4
Al2O3 (single) 133
10 –11 90.8
57.8
O in Al2O3
(poly) kJ/mol
Ni in NiO
10 –12 (air)
O in Al2O3
(single)
10 –13
C in
graphite O in UO2.00
O in TiO2
10 –14 (Po2 = 101 kPa)
Ca in CaO
O in MgO O fused SiO2 Figure 3.4 Diffusion coefficients for single-
O in Cr2O3 (Po2 = 101 kPa)
U in UO2.00 and polycrystalline ceramics.
10 –15 [From W.D. Kingery, H.K. Bowen, and D.R.
0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 Uhlmann, Introduction to Ceramics, 2nd ed., Wiley
1/T × 1000/T, K –1 Interscience, New York (1976) with permission.]
about 20% amorphous. It is mainly through the amorphous where Ki, the equilibrium partition coefficient, is equal to the
regions that diffusion occurs. As with the transport of ratio of the concentration in the polymer to the concentration,
gases through metals, transport of gaseous species through ci, in the liquid adjacent to the polymer surface. The product
polymer membranes is usually characterized by the solution- KiDi is the liquid permeability.
diffusion mechanism of (3-50). Fick’s first law, in the fol- Values of diffusivity for light gases in four polymers, given
lowing integrated forms, is then applied to compute the mass in Table 14.6, range from 1.3 × 10−9 to 1.6 × 10−6 cm2 /s,
transfer flux. which is orders of magnitude less than for diffusion of the
same species in a gas.
Gas species:
Diffusivities of liquids in rubbery polymers have been
studied extensively as a means of determining viscoelastic
Hi Di PMi
Ni = ( pi − pi2 ) = ( pi − pi2 ) (3-52) parameters. In Table 3.12, taken from Ferry [20], diffusivi-
z2 − z1 1 z2 − z1 1 ties are given for different solutes in seven different rubber
polymers at near-ambient conditions. The values cover a
where pi is the partial pressure of the gas species at a poly- sixfold range, with the largest diffusivity being that for
mer surface. n-hexadecane in polydimethylsiloxane. The smallest diffu-
Liquid species: sivities correspond to the case where the temperature is
approaching the glass-transition temperature, where the
K i Di polymer becomes glassy in structure. This more rigid struc-
Ni = (ci − ci2 ) (3-53)
z2 − z1 1 ture hinders diffusion. In general, as would be expected,
3.2 Diffusion Coefficients 83
Diffusivity,
Polymer Solute Temperature, K cm2/s
Polyisobutylene n-Butane 298 1.19 × 10−9
i-Butane 298 5.3 × 10−10
n-Pentane 298 1.08 × 10−9
n-Hexadecane 298 6.08 × 10−10
Hevea rubber n-Butane 303 2.3 × 10−7
i-Butane 303 1.52 × 10−7
n-Pentane 303 2.3 × 10−7
n-Hexadecane 298 7.66 × 10−8
Polymethylacrylate Ethyl alcohol 323 2.18 × 10−10
Polyvinylacetate n-Propyl alcohol 313 1.11 × 10−12
n-Propyl chloride 313 1.34 × 10−12
Ethyl chloride 343 2.01 × 10−9
Ethyl bromide 343 1.11 × 10−9
Polydimethylsiloxane n-Hexadecane 298 1.6 × 10−6
1,4-Polybutadiene n-Hexadecane 298 2.21 × 10−7
Styrene-butadiene rubber n-Hexadecane 298 2.66 × 10−8
The structure of wood, which often consists of (1) highly Typical results are given by Sherwood [25] and Stamm [24].
elongated hexagonal or rectangular cells, called tracheids in For example, for beech with a swollen specific gravity of
softwood (coniferous species, e.g., spruce, pine, and fir) and 0.4, the diffusivity increases from a value of about
fibers in hardwood (deciduous or broad-leaf species, e.g., 1 × 10−6 cm2 /s at 10◦ C to 10 × 10−6 cm2 /s at 60°C.
oak, birch, and walnut); (2) radial arrays of rectangular-like
cells, called rays, which are narrow and short in softwoods
3.3 ONE-DIMENSIONAL, STEADY-STATE
but wide and long in hardwoods; and (3) enlarged cells with
large pore spaces and thin walls, called sap channels because
AND UNSTEADY-STATE, MOLECULAR
they conduct fluids up the tree. The sap channels are less DIFFUSION THROUGH STATIONARY MEDIA
than 3 vol% of softwood, but as much as 55 vol% of For conductive heat transfer in stationary media, Fourier’s
hardwood. law is applied to derive equations for the rate of heat transfer
Because the structure of wood is directional, many of its for steady-state and unsteady-state conditions in shapes such
properties are anisotropic. For example, stiffness and as slabs, cylinders, and spheres. Many of the results are plot-
strength are 2 to 20 times greater in the axial direction of the ted in generalized charts. Analogous equations can be de-
tracheids or fibers than in the radial and tangential directions rived for mass transfer, using simplifying assumptions.
of the trunk from which the wood is cut. This anisotropy ex- In one dimension, the molar rate of mass transfer of A in
tends to permeability and diffusivity of wood penetrants, a binary mixture with B is given by a modification of (3-12),
such as moisture and preservatives. According to Stamm which includes bulk flow and diffusion:
[24], the permeability of wood to liquids in the axial direc-
dxA
tion can be up to 10 times greater than in the transverse n A = xA (n A + n B ) − cDAB A (3-54)
direction. dz
Movement of liquids and gases through wood and wood If A is a dissolved solute undergoing mass transfer, but B is
products takes time during drying and treatment with preser- stationary, n B = 0. It is common to assume that c is a constant
vatives, fire retardants, and other chemicals. This movement and xA is small. The bulk-flow term is then eliminated and
takes place by capillarity, pressure permeability, and (3-54) accounts for diffusion only, becoming Fick’s first law:
diffusion. Nevertheless, wood is not highly permeable be-
dxA
cause the cell voids are largely discrete and lack direct n A = −cDAB A (3-55)
interconnections. Instead, communication among cells is dz
through circular openings spanned by thin membranes with Alternatively, (3-55) can be written in terms of concentration
submicrometer-sized pores, called pits, and to a smaller ex- gradient:
tent, across the cell walls. Rays give wood some permeabil-
dcA
ity in the radial direction. Sap channels do not contribute to n A = −DAB A (3-56)
permeability. All three mechanisms of movement of gases dz
and liquids in wood are considered by Stamm [24]. Only dif- This equation is analogous to Fourier’s law for the rate of
fusion is discussed here. heat conduction, Q:
The simplest form of diffusion is that of a water-soluble
solute through wood saturated with water, such that no di- dT
Q = −k A (3-57)
mensional changes occur. For the diffusion of urea, glycer- dz
ine, and lactic acid into hardwood, Stamm [24] lists diffu- Steady State
sivities in the axial direction that are about 50% of ordinary
liquid diffusivities. In the radial direction, diffusivities are For steady-state, one-dimensional diffusion, with constant
about 10% of the values in the axial direction. For example, DAB, (3-56) can be integrated for various geometries, the
at 26.7◦ C the diffusivity of zinc sulfate in water is most common results being analogous to heat conduction.
5 × 10−6 cm2 /s. If loblolly pine sapwood is impregnated
1. Plane wall with a thickness, z 2 − z 1 :
with zinc sulfate in the radial direction, the diffusivity is
found to be 0.18 × 10−6 cm2 /s [24]. cA1 − cA2
The diffusion of water in wood is more complex. Mois- n A = DAB A (3-58)
z2 − z1
ture content determines the degree of swelling or shrinkage.
Water is held in the wood in different ways: It may be 2. Hollow cylinder of inner radius r1 and outer radius r2,
physically adsorbed on cell walls in monomolecular layers, with diffusion in the radial direction outward:
condensed in preexisting or transient cell capillaries, or DAB (cA1 − cA2 )
absorbed in cell walls to form a solid solution. n A = 2L (3-59)
ln(r2 /r1 )
Because of the practical importance of lumber drying
rates, most diffusion coefficients are measured under drying or
conditions in the radial direction across the fibers. Results cA1 − cA2
n A = DAB ALM (3-60)
depend on temperature and swollen-volume specific gravity. r2 − r1
3.3 One-Dimensional, Steady-State and Unsteady-State, Molecular Diffusion through Stationary Media 85
When r1 /r2 < 2, the arithmetic mean area is no more For one-dimensional diffusion in the radial direction only,
than 4% greater than the log mean area. When r1 /r2 < 1.33, for cylindrical and spherical coordinates, Fick’s second law
the arithmetic mean area is no more than 4% greater than the becomes, respectively,
geometric mean area.
∂cA DAB ∂ ∂cA
= r (3-70)
∂t r ∂r ∂r
Unsteady State
and
Equation (3-56) is applied to unsteady-state molecular diffu-
∂cA DAB ∂ 2 ∂cA
sion by considering the accumulation or depletion of a = 2 r (3-71)
∂t r ∂r ∂r
species with time in a unit volume through which the species
is diffusing. Consider the one-dimensional diffusion of Equations (3-68) to (3-71) are analogous to Fourier’s sec-
species A in B through a differential control volume with dif- ond law of heat conduction where cA is replaced by temper-
fusion in the z-direction only, as shown in Figure 3.5. ature, T, and diffusivity, DAB, is replaced by thermal diffu-
Assume constant total concentration, c = cA + cB , constant sivity, = k/C P . The analogous three equations for heat
diffusivity, and negligible bulk flow. The molar flow rate of conduction for constant, isotropic properties are, respec-
species A by diffusion at the plane z = z is given by (3-56): tively:
2
∂cA ∂T ∂ T ∂2T ∂2T
n Az = −DAB A (3-63) = + + (3-72)
∂z z ∂t ∂x2 ∂ y2 ∂z 2
At the plane, z = z + z, the diffusion rate is ∂T ∂ ∂T
= r (3-73)
∂cA ∂t r ∂r ∂r
n Az+z = −DAB A (3-64)
∂z z+z ∂T ∂ ∂T
= 2 r2 (3-74)
The accumulation of species A in the control volume is ∂t r ∂r ∂r
∂cA
A z (3-65) Analytical solutions to these partial differential equations in
∂t either Fick’s law or Fourier’s law form are available for a
Since rate in − rate out = accumulation, variety of boundary conditions. Many of these solutions are
∂cA ∂cA ∂cA derived and discussed by Carslaw and Jaeger [26] and Crank
−DAB A + DAB A =A z [27]. Only a few of the more useful solutions are presented
∂z z ∂z z+z ∂t
here.
(3-66)
∂ cA ∂ cA ∂ cA
Consider the semi-infinite medium shown in Figure 3.6,
nAz = – DAB A
∂z z
A
∂t
dz nAz +∆z = – DAB A
∂z z+∆z which extends in the z-direction from z = 0 to z = ∞. The x
and y coordinates extend from −∞ to +∞, but are not of
interest because diffusion takes place only in the z-direction.
Thus, (3-68) applies to the region z ≥ 0. At time t ≤ 0, the
concentration is cAo for z ≥ 0. At t = 0, the surface of the
z z+∆z semi-infinite medium at z = 0 is instantaneously brought to
Figure 3.5 Unsteady-state diffusion through a differential the concentration cAs > cAo and held there for t > 0. There-
volume A dz. fore, diffusion into the medium occurs. However, because
86 Chapter 3 Mass Transfer and Diffusion
transfer can be greatly increased by agitation to induce turbulent Center of slab Surface of slab
motion. For solids, it is best to reduce the diffusion path to as small 1.0 0
a dimension as possible by reducing the size of the solid. DAB t /a2 1.0
cAs – cAo = 1 – E
Consider a rectangular, parallelepiped medium of finite
cAs – cAo = E
thickness 2a in the z-direction, and either infinitely long
cAs – cA
cA – cAo
dimensions in the y- and x-directions or finite lengths of 2b 0.4 0.6
0.2
and 2c, respectively, in those directions. Assume that in Fig-
ure 3.7a the edges parallel to the z-direction are sealed, so
diffusion occurs only in the z-direction and initially the con- 0.2 0.1 0.8
b
y
z
b
a
x a c c
r
c x
c a
(a) Slab. Edges at x = + c and – c and (b) Cylinder. Two circular ends at x = + c
at y = +b and – b are sealed. and – c are sealed.
r
a
We can also determine the total number of moles transferred This equation is plotted in Figure 3.9. It is important to note
across either unsealed face by integrating (3-82) with respect that concentrations are in mass of solute per mass of dry
to time. Thus, solid or mass of solute/volume. This assumes that during dif-
t fusion the solid does not shrink or expand so that the mass of
8(cAs − cAo ) Aa
NA = n A |z=a dt = dry solid per unit volume of wet solid will remain constant.
o 2 Then, we can substitute a concentration in terms of mass or
∞
1 DAB (2n + 1) 2 2 t moles of solute per mass of dry solid, i.e., the moisture con-
× 1 − exp −
n=0
(2n + 1) 2 4a 2 tent on the dry basis.
When the edges of the slab in Figure 3.7a are not sealed,
(3-83) the method of Newman [31] can be used with (3-69) to
In addition, the average concentration of the solute in the determine concentration changes within the slab. In this
medium, cAavg , as a function of time, can be obtained in the method, E or Eavg is given in terms of the E values from the
case of a slab from: solution of (3-68) for each of the coordinate directions by
a
cAs − cAavg (1 − ) dz E = Ex E y Ez (3-86)
= o (3-84)
cAs − cAo a
Corresponding solutions for infinitely long, circular cylin-
Substitution of (3-80) into (3-84) followed by integration ders and spheres are available in Carslaw and Jaeger [26]
gives and are plotted in Figures 3.9, 3.10, and 3.11, respectively.
For a short cylinder, where the ends are not sealed, E or Eave
cAs − cAavg is given by the method of Newman as
E avgslab = (1 − ave ) slab =
cAs − cAo
E = Er E x (3-87)
8
∞
1 DAB (2n + 1) 2 2 t
= 2 exp −
n=0 (2n + 1) 2 4a 2 Some materials such as crystals and wood, have thermal
(3-85) conductivities and diffusivities that vary markedly with
direction. For these anisotropic materials, Fick’s second law
in the form of (3-69) does not hold. Although the general
1.0 anisotropic case is exceedingly complex, as shown in the
0.8
Ea , 2c
following example, the mathematical treatment is relatively
0.6 Eb ,
Ec (
slab 2a
simple when the principal axes of diffusivity coincide with
)
0.4 the coordinate system.
2b
0.3 E
r (c
yl
0.2 in
de
r)
2c
Axis of cylinder Surface of cylinder
0.10
Es
2a
(s
0.08 1.0 0
ph
2
er
0.06 D A Bt/a
Eavg = c s – c avg
e)
Ao
0.4
cA – cA
0.04
0.8 0.2
As
0.03
2a 0.6 0.4
0.2
cAs – cA
0.010
cA – cAo
0.008
0.4 0.6
0.006
0.004
0.003 0.1
0.2 0.8
0.002 0.04
0.01
0 1.0
0 0.2 0.4 0.6 0.8 1.0
0.001 r
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
a
DABt/a2, DABt/b2, DABt/c2
Figure 3.10 Concentration profiles for unsteady-state diffusion
Figure 3.9 Average concentrations for unsteady-state diffusion. in a cylinder.
[Adapted from R.E. Treybal, Mass-Transfer Operations, 3rd ed., McGraw- [Adapted from H.S. Carslaw and J.C. Jaeger, Conduction of Heat in
Hill, New York (1980).] Solids, 2nd ed., Oxford University Press, London (1959).]
3.3 One-Dimensional, Steady-State and Unsteady-State, Molecular Diffusion through Stationary Media 89
Center of sphere Surface of sphere Since this is the same form as (3-69) and since the boundary condi-
1.0 0 tions do not involve diffusivities, we can apply Newman’s method,
0.4 D ABt/a
2 using Figure 3.9, where concentration, cA, is replaced by weight-
percent moisture on a dry basis.
0.8 0.2 From (3-86) and (3-85),
0.2
cAave − cAs 5−2
E aveslab = E avgx E avg y E avgz = = = 0.167
cAo − cAs 20 − 2
cAs – cAo = E
cAs – cAo = 1 – E
0.6 0.4
cAs – cA
Let D = 1 × 10−5 cm2/s.
cA – cAo
0.4 0.6
z1 Direction (axial):
0.1 1/2 1/2
D 20 1 × 10−5
a1 = a = = 7.07 cm
Dz 2 2 × 10−5
0.2 0.8
Dt 1 × 10−5 t
0.04
0.01
= = 2.0 × 10−7 t, s
a12 7.072
0 1.0
0 0.2 0.4 0.6 0.8 1.0 y1 Direction:
r
1/2 1/2
a
D 10 1 × 10−5
Figure 3.11 Concentration profiles for unsteady-state diffusion in b1 = b = = 7.906 cm
Dy 2 4 × 10−6
a sphere.
Dt 1 × 10−5 t
[Adapted from H.S. Carslaw and J.C. Jaeger, Conduction of Heat in
2
= = 1.6 × 10−7 t, s
Solids, 2nd ed., Oxford University Press, London (1959).] b1 7.9062
x1-Direction:
1/2 1/2
D 5 1 × 10−5
EXAMPLE 3.12 c1 = c = = 3.953 cm
Dx 2 4 × 10−6
A piece of lumber, measuring 5 × 10 × 20 cm, initially contains Dt 1 × 10−5 t
20 wt% moisture. At time 0, all six faces are brought to an equilib- 2
= = 6.4 × 10−7 t, s
c1 3.9532
rium moisture content of 2 wt%. Diffusivities for moisture at 25◦ C
are 2 × 10−5 cm2/s in the axial (z) direction along the fibers and Use Figure 3.9 iteratively with assumed values of time in seconds
4 × 10−6 cm2/s in the two directions perpendicular to the fibers. to obtain values of Eavg for each of the three coordinates until
Calculate the time in hours for the average moisture content to drop (3-86) equals 0.167.
to 5 wt% at 25◦ C. At that time, determine the moisture content at
the center of the piece of lumber. All moisture contents are on a dry t, h t, s Eavgz Eavgy Eavg x Eavg
1 1 1
basis.
100 360,000 0.70 0.73 0.46 0.235
120 432,000 0.67 0.70 0.41 0.193
SOLUTION 135 486,000 0.65 0.68 0.37 0.164
In this case, the solid is anisotropic, with Dx = D y = 4 × 10−6 cm2/s Therefore, it takes approximately 136 h.
and Dz = 2 × 10−5 cm2/s, where dimensions 2c, 2b, and 2a in the For 136 h = 490,000 s, the Fourier numbers for mass transfer
x, y, and z directions are 5, 10, and 20 cm, respectively. Fick’s sec- are
ond law for an isotropic medium, (3-69), must be rewritten for this
anisotropic material as Dt (1 × 10−5 )(490,000)
2
= = 0.0980
a1 7.072
2
∂cA ∂ cA ∂ 2 cA ∂ 2 cA Dt (1 × 10−5 )(490,000)
= Dx + + Dz 2 (1) = = 0.0784
∂t ∂x 2 ∂y 2 ∂z 2
b1 7.9062
Dt (1 × 10−5 )(490,000)
as discussed by Carslaw and Jaeger [26].
2
= = 0.3136
To transform (1) into the form of (3-69), let c1 3.9532
From Figure 3.8, at the center of the slab,
D D D
x1 = x y1 = y z1 = z (2)
Dx Dx Dz
E center = E z1 E y1 E x1 = (0.945)(0.956)(0.605) = 0.547
cA − cAcenter 2 − cAcenter
where D is chosen arbitrarily. With these changes in variables, = s = = 0.547
(1) becomes cAs − cAo 2 − 20
Solving,
∂cA ∂ 2 cA ∂ 2 cA ∂ 2 cA
=D + + (3) cA at the center = 11.8 wt% moisture
∂t ∂ x12
∂ y12
∂z 12
90 Chapter 3 Mass Transfer and Diffusion
3.4 MOLECULAR DIFFUSION Thus, the maximum liquid velocity, which occurs at z = 0, is
IN LAMINAR FLOW g2
(u y ) max = (3-90)
Many mass-transfer operations involve diffusion in fluids in 2
laminar flow. The fluid may be a film flowing slowly down a The bulk-average velocity in the liquid film is
vertical or inclined surface, a laminar boundary layer that
forms as the fluid flows slowly past a thin plate, or the fluid u y dz g2
ū y = 0 = (3-91)
may flow through a small tube or slowly through a large pipe 3
or duct. Mass transfer may occur between a gas and a liquid Thus, the film thickness for fully developed flow is indepen-
film, between a solid surface and a fluid, or between a fluid dent of location y and is
and a membrane surface.
3ū y 1/2 3 1/3
= = (3-92)
g 2g
Falling Liquid Film
where = liquid film flow rate per unit width of film, W.
Consider a thin liquid film, of a mixture of volatile A and
For film flow, the Reynolds number, which is the ratio of
nonvolatile B, falling in laminar flow at steady state down
the inertial force to the viscous force, is
one side of a vertical surface and exposed to pure gas, A, on
the other side of the film, as shown in Figure 3.12. The sur- 4r H ū y 4ū y 4
NRe = = = (3-93)
face is infinitely wide in the x-direction (normal to the page).
In the absence of mass transfer of A into the liquid film, the where rH = hydraulic radius = (flow cross section)/(wetted
liquid velocity in the z-direction, uz, is zero. In the absence of perimeter) = (W )/W = and, by the equation of continu-
end effects, the equation of motion for the liquid film in fully ity, = ū y .
developed laminar flow in the downward y-direction is As reported by Grimley [32], for NRe < 8 to 25, depend-
ing on the surface tension and viscosity, the flow in the film
d 2u y
+ g = 0 (3-88) is laminar and the interface between the liquid film and the
dz 2 gas is flat. The value of 25 is obtained with water. For 8 to
25 < NRe < 1,200, the flow is still laminar, but ripples and
Usually, fully developed flow, where uy is independent of the waves may appear at the interface unless suppressed by the
distance y, is established quickly. If is the thickness of the addition of wetting agents to the liquid.
film and the boundary conditions are u y = 0 at z = (no- For a flat liquid–gas interface and a small rate of mass
slip condition at the solid surface) and duy /dz = 0 at z = 0 transfer of A into the liquid film, (3-88) to (3-93) hold and
(no drag at the liquid–gas interface), (3-88) is readily inte- the film velocity profile is given by (3-89). Now consider a
grated to give a parabolic velocity profile: mole balance on A for an incremental volume of liquid film
z 2 of constant density, as shown in Figure 3.12. Neglect bulk
g2
uy = 1− (3-89) flow in the z-direction and axial diffusion in the y-direction.
2 Then, at steady state, neglecting accumulation or depletion
of A in the incremental volume,
z=δ ∂cA
z z = 0, y = 0 −DAB (y)(x) + u y cA | y (z)(x)
∂z z
∂cA
y = −DAB (y)(x) + u y cA | y+y (z)(x)
cA i (in liquid) ∂z z+z
Liquid (3-94)
Bulk
Gas flow
Rearranging and simplifying (3-94),
u y cA | y+y − u y cA | y (∂cA /∂z) z+z − (∂cA /∂z) z
= DAB
y y z
Liquid (3-95)
uy {z} Diffusion
film
element
of A In the limit, as z → 0 and y → 0,
z + ∆z
y +∆y
∂cA ∂ 2 cA
z
uy = DAB 2 (3-96)
∂y ∂z
cA {z} Substituting (3-89) into (3-96),
z 2 ∂c
Figure 3.12 Mass transfer from a gas into a falling, laminar g2 A ∂ 2 cA
1− = DAB 2 (3-97)
liquid film. 2 ∂y ∂z
3.4 Molecular Diffusion in Laminar Flow 91
This equation was solved by Johnstone and Pigford [33] For mass transfer, a composition driving force replaces
and later by Olbrich and Wild [34], for the following bound- T . As discussed later in this chapter, because composition
ary conditions: can be expressed in a number of ways, different mass-
cA = cAi at z = 0 for y > 0 transfer coefficients are defined. If we select cA as the dri-
cA = cA0 at y = 0 for 0 < z < ving force for mass transfer, we can write
∂cA /∂z = 0 at z = for 0 < y < L n A = kc A cA (3-105)
where L = height of the vertical surface. The solution of which defines a mass-transfer coefficient, kc, in mol/
Olbrich and Wild is in the form of an infinite series, giving time-area-driving force, for a concentration driving force.
cA as a function of z and y. However, of more interest is the Unfortunately, no name is in general use for (3-105).
average concentration at y = L, which, by integration, is For the falling laminar film, we take cA = cAi − c̄A ,
which varies with vertical location, y, because even though
1
c̄Ay = u y cAy dz (3-98) cAi is independent of y, the average film concentration, c̄A ,
ū y 0
increases with y. To derive an expression for kc, we equate
For the condition y = L, the result is (3-105) to Fick’s first law at the gas–liquid interface:
cAi − c̄AL
= 0.7857e−5.1213
+ 0.09726e−39.661
∂cA
cAi − cA0 (3-99) kc A(cAi − c̄A ) = −DAB A (3-106)
+ 0.036093−106.25
+ · · · ∂z z=0
where Although this is the most widely used approach for defin-
2DAB L 8/3 8/3 ing a mass-transfer coefficient, in this case of a falling film it
= 2
= = (3-100) fails because (∂cA/∂z) at z = 0 is not defined. Therefore, for
3 ū y NRe NSc (/L) (/L) NPe M
this case we use another approach as follows. For an incre-
NSc = Schmidt number = mental height, we can write for film width W,
DAB
(3-101) n A = ū y W d c̄A = kc (cAi − c̄A )W dy
momentum diffusivity,/ (3-107)
=
mass diffusivity, DAB This defines a local value of kc, which varies with distance y
NPeM = NRe NSc = Peclet number for mass transfer because c̄A varies with y. An average value of kc, over a
4ū y (3-102) height L, can be defined by separating variables and inte-
= grating (3-107):
DAB c
L
The Schmidt number is analogous to the Prandtl number, k c dy ū y cAAL [d c̄A /(cAi − c̄A )]
used in heat transfer: kcavg = 0 = 0
L L
CP (/ ) momentum diffusivity ū y cAi − cA0 (3-108)
NPr = = = = ln
k (k/C P ) thermal diffusivity L cAi − c̄AL
The Peclet number for mass transfer is analogous to the
In general, the argument of the natural logarithm in
Peclet number for heat transfer:
4ū y C P (3-108) is obtained from the reciprocal of (3-99). For values
NPeH = NRe NPr = of
in (3-100) greater than 0.1, only the first term in (3-99)
k
is significant (error is less than 0.5%). In that case,
Both Peclet numbers are ratios of convective transport to
molecular transport. ū y e5.1213
kcavg = ln (3-109)
The total rate of absorption of A from the gas into the L 0.7857
liquid film for height L and width W is
Since ln ex = x,
n A = ū y W ( c̄AL − cA0 ) (3-103) ū y
kcavg = (0.241 + 5.1213
) (3-110)
Mass-Transfer Coefficients L
Mass-transfer problems involving fluids are most often In the limit, for large
, using (3-100) and (3-102), (3-110)
solved using mass-transfer coefficients, analogous to heat- becomes
transfer coefficients. For the latter, Newton’s law of cooling DAB
kcavg = 3.414 (3-111)
defines a heat-transfer coefficient, h:
Q = h A T (3-104) In a manner suggested by the Nusselt number,
where NNu = h/k for heat transfer, where = a characteristic
length, we define a Sherwood number for mass transfer,
Q = rate of heat transfer
which for a falling film of characteristic length is
A = area for heat transfer (normal to the direction of
heat transfer) kcavg
NShavg = (3-112)
T = temperature-driving force for heat transfer DAB
92 Chapter 3 Mass Transfer and Diffusion
From (3-111), NShavg = 3.414, which is the smallest value The error function is defined as
that the Sherwood number can have for a falling liquid film. z
2
e−t dt
2
The average mass-transfer flux of A is given by erf z = √ (3-118)
0
n Aavg
NAavg = = kcavg (cAi − c̄A ) mean (3-113) Using the Leibnitz rule with (3-116) to differentiate this in-
A
tegral function,
For values
< 0.001 in (3-100), when the liquid-film
∂cA 3ū y
flow regime is still laminar without ripples, the time of con- = −(cAi − cA0 ) (3-119)
tact of the gas with the liquid is short and mass transfer is ∂z z=0 2 DAB y
confined to the vicinity of the gas–liquid interface. Thus, the Substituting (3-119) into (3-117) and introducing the Peclet
film acts as if it were infinite in thickness. In this limiting number for mass transfer from (3-102), we obtain an expres-
case, the downward velocity of the liquid film in the region sion for the local mass-transfer coefficient as a function of
of mass transfer is just u ymax , and (3-96) becomes distance down from the top of the wall:
∂cA ∂ 2 cA
u ymax = DAB 2 (3-114) 2
3DAB NPeM 3DAB
∂y ∂z kc = = (3-120)
8y 2y
Since from (3-90) and (3-91), u ymax = 3ū y /2, (3-114) can be
rewritten as The average value of kc over the height of the film, L, is ob-
tained by integrating (3-120) with respect to y, giving
∂cA 2DAB ∂ 2 cA
= (3-115) 6DAB 2
3DAB
∂y 3ū y ∂z 2 kcavg = = NPeM (3-121)
L 2L
where the boundary conditions are
Combining (3-121) with (3-112) and (3-102),
cA = cA0 for z > 0 and y > 0
cA = cAi for z = 0 and y > 0 3 4
NShavg = NPe M = (3-122)
cA = cA0 for large z and y > 0 2L
Equation (3-115) and the boundary conditions are equivalent where, by (3-108), the proper mean to use with kcavg is the log
to the case of the semi-infinite medium, as developed above. mean. Thus,
Thus, by analogy to (3-68), (3-75), and (3-76) the solution is (cAi − c̄A ) mean = (cAi − c̄A ) LM
cAi − cA z (cAi − cA0 ) − (cAi − cAL ) (3-123)
E =1− = = erf (3-116) =
cAi − cA0 2 2DAB y/3ū y ln[(cAi − cA0 )/(cAi − c̄AL )]
Assuming that the driving force for mass transfer in the film When ripples are present, values of kcavg and NShavg can be
is cAi − cA0 , we can use Fick’s first law at the gas–liquid considerably larger than predicted by these equations.
interface to define a mass-transfer coefficient: In the above development, asymptotic, closed-form solu-
tions are obtained with relative ease for large and small
∂cA values of
, defined by (3-100). These limits, in terms of the
NA = −DAB = kc (cAi − cA0 ) (3-117)
∂z z=0 average Sherwood number, are shown in Figure 3.13. The
100
Sherwood number
Sh
ort
res
ide
10 Eq nce-t Gen
era
. (3 im l so
-12 e s lutio
2) olu n
tio
n
1
0.001 0.01 0.1 1 10
η=
8/3 Figure 3.13 Limiting and general solutions for
( δ /L)NPe M mass transfer to a falling, laminar liquid film.
3.4 Molecular Diffusion in Laminar Flow 93
Thus, the exiting liquid film is saturated with CO2, which implies
EXAMPLE 3.13 equilibrium at the gas–liquid interface. From (3-103),
Water (B) at 25◦ C, in contact with pure CO2 (A) at 1 atm, flows as
n A = 0.0486(1.15 × 10−4 )(3.4 × 10−2 ) = 1.9 × 10−7 kmol/s
a film down a vertical wall 1 m wide and 3 m high at a Reynolds
number of 25. Using the following properties, estimate the rate of
adsorption of CO2 into water in kmol/s: Boundary-Layer Flow on a Flat Plate
−5
DAB = 1.96 × 10 cm /s; 2
= 1.0 g/cm ; 3
Consider the flow of a fluid (B) over a thin, flat plate parallel
L = 0.89 cP = 0.00089 kg/m-s with the direction of flow of the fluid upstream of the plate,
Solubility of CO2 in water at 1 atm and 25◦ C = 3.4 × as shown in Figure 3.14. A number of possibilities for mass
10−5 mol/cm3 . transfer of another species, A, into B exist: (1) The plate
might consist of material A, which is slightly soluble in B.
SOLUTION (2) Component A might be held in the pores of an inert solid
plate, from which it evaporates or dissolves into B. (3) The
From (3-93),
plate might be an inert, dense polymeric membrane, through
NRe 25(0.89)(0.001) kg which species A can pass into fluid B. Let the fluid velocity
= = = 0.00556
4 4 m−s profile upstream of the plate be uniform at a free-system ve-
locity of uo. As the fluid passes over the plate, the velocity ux
From (3-101),
in the direction of flow is reduced to zero at the wall, which
(0.89)(0.001) establishes a velocity profile due to drag. At a certain dis-
NSc = = = 454
DAB (1.0)(1,000)(1.96 × 10−5 )(10−4 ) tance z, normal to and out from the solid surface, the fluid ve-
locity is 99% of uo. This distance, which increases with
From (3-92),
increasing distance x from the leading edge of the plate, is
arbitrarily defined as the velocity boundary-layer thickness,
3(0.89)(0.001)(0.00556) 1/3
= = 1.15 × 10−4 m . Essentially all flow retardation occurs in the boundary
1.02 (1,000) 2 (9.807)
layer, as first suggested by Prandtl [38]. The buildup of this
From (3-90) and (3-91), ū y = (2/3)u ymax . Therefore, layer, the velocity profile in the layer, and the drag force can
be determined for laminar flow by solving the equations
2 (1.0)(1,000)(9.807)(1.15 × 10−4 ) 2
ū y = = 0.0486 m/s of continuity and motion (Navier–Stokes equations) for the
3 2(0.89)(0.001) x-direction. For a Newtonian fluid of constant density and
From (3-100), viscosity, in the absence of pressure gradients in the x- and
8/3
= = 6.13 Free
(25)(454)[(1.15 × 10−4 )/3] stream
uo
Therefore, (3-111) applies, giving uo
uo
y- (normal to the x–z plane) directions, these equations for If mass transfer begins at the leading edge of the plate and if
the region of the boundary layer are the concentration in the fluid at the solid–fluid interface is
constant, the additional boundary conditions are
∂u x ∂u z
+ =0 (3-124) cA = cAo at x = 0 for z > 0,
∂x ∂z
cA = cAi at z = 0 for x > 0,
∂u x ∂u x ∂ 2u x
ux + uz = (3-125) and cA = cAo at z = ∞ for x > 0
∂x ∂z ∂z 2
If the rate of mass transfer is low, the velocity profiles are
The boundary conditions are undisturbed. The solution to the analogous problem in heat
u x = u o at x = 0 for z > 0 u x = 0 at z = 0 for x > 0 transfer was first obtained by Pohlhausen [42] for NPr > 0.5,
u x = u o at z = ∞ for x > 0 u z = 0 at z = 0 for x > 0 as described in detail by Schlichting [40]. The results for
mass transfer are
The solution of (3-124) and (3-125) in the absence of heat NShx 0.332
and mass transfer, subject to these boundary conditions, was 1/3
= 0.5 (3-132)
first obtained by Blasius [39] and is described in detail by NRex NSc NRex
Schlichting [40]. The result in terms of a local friction factor, where
fx, a local shear stress at the wall, w x , and a local drag coef- xkcx
NShx = (3-133)
ficient at the wall, C Dx , is DAB
C Dx fx w 0.322 and the driving force for mass transfer is cAi − cAo .
= = x2 = 0.5 (3-126) The concentration boundary layer, where essentially all
2 2 uo NRex
of the resistance to mass transfer resides, is defined by
where
cAi − cA
xu o = 0.99 (3-134)
NRex = (3-127) cAi − cAo
and the ratio of the concentration boundary-layer thickness,
Thus, the drag is greatest at the leading edge of the plate, c , to the velocity boundary thickness, , is
where the Reynolds number is smallest. Average values of c / = 1/NSc
1/3
(3-135)
the drag coefficient are obtained by integrating (3-126) from
x = 0 to L, giving Thus, for a liquid boundary layer, where NSc > 1, the concen-
tration boundary layer builds up more slowly than the veloc-
C Davg f avg 0.664 ity boundary layer. For a gas boundary layer, where N SC ≈ 1,
= = (3-128)
2 2 ( NReL ) 0.5 the two boundary layers build up at about the same rate. By
The thickness of the velocity boundary layer increases with analogy to (3-130), the concentration profile is given by
3
distance along the plate: cAi − cA z z
= 1.5 − 0.5 (3-136)
4.96 cAi − cAo c c
= 0.5 (3-129)
x NRex Equation (3-132) gives the local Sherwood number. If
this expression is integrated over the length of the plate, L,
A reasonably accurate expression for the velocity profile
the average Sherwood number is found to be
was obtained by Pohlhausen [41], who assumed the empiri-
1/2 1/3
cal form ux = C1z + C2z3. NShavg = 0.664 NReL NSc (3-137)
If the boundary conditions,
where
Lkcavg
u x = 0 at z = 0 u x = u o at z = ∂u x /∂z = 0 at z = NShavg = (3-138)
DAB
are applied to evaluate C1 and C2, the result is
ux z z 3 EXAMPLE 3.14
= 1.5 − 0.5 (3-130) Air at 100◦ C, 1 atm, and a free-stream velocity of 5 m/s flows over
uo
a 3-m-long, thin, flat plate of naphthalene, causing it to sublime.
This solution is valid only for a laminar boundary layer,
(a) Determine the length over which a laminar boundary layer
which by experiment persists to NRex = 5 × 105 . persists.
When mass transfer of A into the boundary layer occurs,
the following species continuity equation applies at constant (b) For that length, determine the rate of mass transfer of naphtha-
lene into air.
diffusivity:
2 (c) At the point of transition of the boundary layer to turbulent
∂cA ∂cA ∂ cA flow, determine the thicknesses of the velocity and concentra-
ux + uz = DAB (3-131)
∂x ∂z ∂x2 tion boundary layers.
3.4 Molecular Diffusion in Laminar Flow 95
Assume the following values for physical properties: build up as shown at planes b, c, and d. In this region, the
Vapor pressure of napthalene = 10 torr
central core outside the boundary layer has a flat velocity
Viscosity of air = 0.0215 cP profile where the flow is accelerated over the entrance ve-
Molar density of air = 0.0327 kmol/m3 locity. Finally, at plane e, the boundary layer fills the tube.
Diffusivity of napthalene in air = 0.94 × 10−5 m2 /s From here the velocity profile is fixed and the flow is said to
be fully developed. The distance from the plane a to plane e
SOLUTION is the entry region.
For fully developed laminar flow in a straight, circular
(a) NRex = 5 × 105 for transition. From (3-127), tube, by experiment, the Reynolds number, NRe = D ū x /,
NRex [(0.0215)(0.001)](5 × 105 ) where ū x is the flow-average velocity in the axial direction,
x=L= = = 2.27 m x, and D is the inside diameter of the tube, must be less than
uo (5)[(0.0327)(29)]
2,100. For this condition, the equation of motion in the axial
at which transition to turbulent flow begins.
direction for horizontal flow and constant properties is
10(0.0327)
(b) cAo = 0 cAi = = 4.3 × 10−4 kmol/m3 ∂ ∂u x dP
760 r − =0 (3-139)
r ∂r ∂r dx
From (3-101),
[(0.0215)(0.001)]
where the boundary conditions are
NSc = = = 2.41 r = 0 (axis of the tube), ∂u x /∂r = 0
DAB [(0.0327)(29)](0.94 × 10−5 )
and r = rw (tube wall), u x = 0
From (3-137),
Equation (3-139) was integrated by Hagen in 1839 and
NShavg = 0.664(5 × 105 ) 1/2 (2.41) 1/3 = 630
Poiseuille in 1841. The resulting equation for the velocity
From (3-138), profile, expressed in terms of the flow-average velocity, is
630(0.94 × 10−5 ) 2
kcavg = = 2.61 × 10−3 m/s r
2.27 u x = 2ū x 1 − (3-140)
rw
For a width of 1 m,
A = 2.27 m2
or, in terms of the maximum velocity at the tube axis,
2
n A = kcavg A(cAi − cAo ) = 2.61 × 10−3 (2.27)(4.3 × 10−4 ) r
u x = u xmax 1 − (3-141)
= 2.55 × 10−6 kmol/s rw
(c) From (3-129), at x = L = 2.27 m, From the form of (3-141), the velocity profile is parabolic in
3.46(2.27)
nature.
= = 0.0111 m The shear stress, pressure drop, and Fanning friction fac-
(5 × 105 ) 0.5
From (3-135), tor are obtained from solutions to (3-139):
c =
0.0111
= 0.0083 m
∂u x 4ū x
(2.41) 1/3 w = − = (3-142)
∂r r=rw rw
dP 32ū x 2 f ū 2x
Fully Developed Flow in a Straight, Circular Tube − = = (3-143)
dx D2 D
Figure 3.15 shows the formation and buildup of a laminar
velocity boundary layer when a fluid flows from a vessel into with
a straight, circular tube. At the entrance, plane a, the veloc- 16
f = (3-144)
ity profile is flat. A velocity boundary layer then begins to NRe
The entry length to achieve fully developed flow is de- The Graetz solution of (3-147) for the temperature profile
fined as the axial distance, Le, from the entrance to the point or the concentration profile is in the form of an infinite
at which the centerline velocity is 99% of the fully devel- series, and can be obtained from (3-146) by the method of
oped flow value. From the analysis of Langhaar [43] for the separation of variables using the method of Frobenius. A
entry region, detailed solution is given by Sellars, Tribus, and Klein [45].
Le From the concentration profile, expressions for the mass-
= 0.0575NRe (3-145) transfer coefficient and the Sherwood number are obtained.
D
When x is large, the concentration profile is fully developed
Thus, at the upper limit of laminar flow, NRe = 2,100, Le/D = and the local Sherwood number, NShx , approaches a limiting
121, a rather large ratio. For NRe = 100, the ratio is only 5.75. value of 3.656. At the other extreme, when x is small such
In the entry region, Langhaar’s analysis shows the friction fac- that the concentration boundary layer is very thin and con-
tor is considerably higher than the fully developed flow value fined to a region where the fully developed velocity profile is
given by (3-144). At x = 0, f is infinity, but then decreases ex- linear, the local Sherwood number is obtained from the
ponentially with x, approaching the fully developed flow value classic Leveque [46] solution, presented by Knudsen and
at Le. For example, for NRe = 1,000, (3-144) gives f = 0.016, Katz [47]:
with Le/D = 57.5. In the region from x = 0 to x/D = 5.35, the
average friction factor from Langhaar is 0.0487, which is k cx D NPeM 1/3
NShx = = 1.077 (3-148)
about three times higher than the fully developed value. DAB (x/D)
In 1885, Graetz [44] obtained a theoretical solution to
the problem of convective heat transfer between the wall of where
a circular tube, held at a constant temperature, and a fluid
flowing through the tube in fully developed laminar flow. D ū x
Assuming constant properties and negligible conduction in NPeM = (3-149)
DAB
the axial direction, the energy equation, after substituting
(3-140) for ux, is The limiting solutions, together with the general Graetz
2
r ∂T k 1 ∂ ∂T solution, are shown in Figure 3.16, where it is seen that
2ū x 1 − = r (3-146) NShx = 3.656 is valid for NPeM /(x/D) < 4 and (3-148) is
rw ∂x C P r ∂r ∂r
valid for NPeM /(x/D) > 100. The two limiting solutions can
The boundary conditions are be patched together if one point of the general solution is
available where the two solutions intersect.
x = 0 (where heat transfer begins), T = T0 , for all r Over a length of tube where mass transfer occurs, an av-
x > 0, r = rw , T = Ti x > 0, r = 0, ∂ T /∂r = 0 erage Sherwood number can be derived by integrating the
general expression for the local Sherwood number. An em-
The analogous species continuity equation for mass pirical representation for that average, proposed by Hausen
transfer, neglecting bulk flow in the radial direction and dif- [48], is
fusion in the axial direction, is
2
r ∂cA 1 ∂ ∂cA 0.0668[NPeM /(x/D)]
2ū x 1 − = DAB r (3-147) NShavg = 3.66 + (3-150)
rw ∂x r ∂r ∂r 1 + 0.04[NPeM /(x/D)]2/3
with analogous boundary conditions. which is based on a log-mean concentration driving force.
100
Sherwood number
10
lut ion
General so
lution q ue so
Leve
Solution for
fully developed
concentration profile
1 Figure 3.16 Limiting and general solutions
1 10 100 1000
for mass transfer to a fluid in laminar flow in a
NPeM /(x /D)
straight, circular tube.
3.5 Mass Transfer in Turbulent Flow 97
Fluid Mechanics
2FD Drag force
Drag Coefficient CD =
Au 2 Projected area × Velocity head
P D Pipe wall shear stress
Fanning Friction Factor f =
L 2ū 2 Velocity head
ū 2 Inertial force
Froude Number NFr =
gL Gravitational force
L ū L ū LG Inertial force
Reynolds Number NRe = = =
Viscous force
ū 2 L Inertial force
Weber Number NWe =
Surface-tension force
Heat Transfer
j-Factor for Heat Transfer j H = NStH ( NPr ) 2/3 jM
hL Convective heat transfer
Nusselt Number NNu = NSh
k Conductive heat transfer
L ūC P Bulk transfer of heat
Peclet Number for Heat Transfer NPeH = NRe NPr = NPeM
k Conductive heat transfer
CP v Momentum diffusivity
Prandtl Number NPr = = NSc
k Thermal diffusivity
NNu h Heat transfer
Stanton Number for Heat Transfer NStH = = NStM
NRe NPr CP G Thermal capacity
Mass Transfer
j-Factor for Mass Transfer j M = NStM ( NSc ) 2/3 jH
NSc k Thermal diffusivity
Lewis Number NLe = = =
NPr C P DAB DAB Mass diffusivity
L ū Bulk transfer of mass
Peclet Number for Mass Transfer NPeM = NRe NSc = NPeH
DAB Molecular diffusion
Momentum diffusivity
Schmidt Number NSc = = NPr
DAB DAB Mass diffusivity
kc L Convective mass transfer
Sherwood Number NSh = NNu
DAB Molecular diffusion
NSh kc Mass transfer
Stanton Number for Mass Transfer NStM = = NStH
NRe NSc ū Mass capacity
L = characteristic length G = mass velocity = u
Subscripts: M = mass transfer H = heat transfer
mean fluid velocity in the direction of flow and on position By analogy, the same mixing length is valid for turbulent-
in the fluid with respect to the solid boundaries. flow heat transfer and mass transfer. To use this analogy,
In 1925, in an attempt to quantify turbulent transport, (3-151) to (3-153) are rewritten in diffusivity form:
Prandtl [52] developed an expression for t in terms of an zx du x
eddy mixing length, l, which is a function of position. The = −( + M ) (3-154)
dz
eddy mixing length is a measure of the average distance that
qz dT
an eddy travels before it loses its identity and mingles with = −( + H ) (3-155)
other eddies. The mixing length is analogous to the mean CP dz
free path of gas molecules, which is the average distance a dcA
NAz = −( DAB + D ) (3-156)
molecule travels before it collides with another molecule. dz
3.5 Mass Transfer in Turbulent Flow 99
where M , H , are D are momentum, heat, and mass eddy NPr = NSc = 1. Thus, the Reynolds analogy has limited
diffusivities, respectively; v is the momentum diffusivity practical value and is rarely applied in practice. Reynolds
(kinematic viscosity), / ; and is the thermal diffusivity, postulated the existence of the analogy in 1874 [53] and
k/C P . As a first approximation, the three eddy diffusivities derived it in 1883 [50].
may be assumed equal. This assumption is reasonably valid
for H and D , but experimental data indicate that Chilton–Colburn Analogy
M / H = M / D is sometimes less than 1.0 and as low as
A widely used extension of the Reynolds analogy to Prandtl
0.5 for turbulence in a free jet.
and Schmidt numbers other than 1 was presented by Colburn
[54] for heat transfer and by Chilton and Colburn [55] for
Reynolds Analogy
mass transfer. They showed that the Reynolds analogy for
If (3-154) to (3-156) are applied at a solid boundary, they can turbulent flow could be corrected for differences in velocity,
be used to determine transport fluxes based on transport temperature, and concentration distributions by incorporat-
coefficients, with driving forces from the wall, i, at z = 0, to ing NPr and NSc into (3-162) to define the following three
the bulk fluid, designated with an overbar, –: Chilton–Colburn j-factors, included in Table 3.13.
zx d( u x /ū x ) f f
jM ≡ = jH ≡
h
( NPr ) 2/3
= −( + M ) = ū x (3-157) 2 GC P
ū x dz z=0 2 (3-165)
kc
d(C P T ) = jD ≡ ( NSc ) 2/3
qz = −( + H ) = h(Ti − T̄ ) (3-158) G
dz z=0
Equation (3-165) is the Chilton–Colburn analogy or the
dcA
NAz = −( DAB + D ) = kc (cAi − c̄A ) (3-159) Colburn analogy for estimating average transport coeffi-
dz z=0
cients for turbulent flow. When NPr = NSc = 1, (3-165)
We define dimensionless velocity, temperature, and solute reduces to (3-162).
concentration by In general, j-factors are uniquely determined by the geo-
metric configuration and the Reynolds number. Based on the
ux Ti − T cA − cA
= = = i (3-160) analysis, over many years, of experimental data on momen-
ū x Ti − T̄ cAi − c̄A tum, heat, and mass transfer, the following representative
If (3-160) is substituted into (3-157) to (3-159), correlations have been developed for turbulent transport to
or from smooth surfaces. Other correlations are presented in
∂ f ū x h
= = other chapters. In general, these correlations are reasonably
∂z z=0 2( + M ) C P ( + H ) accurate for NPr and NSc in the range of 0.5 to 10, but should
(3-161)
kc be used with caution outside this range.
=
( DAB + D ) 1. Flow through a straight, circular tube of inside
This equation defines the analogies among momentum, heat, diameter D:
and mass transfer. Assuming that the three eddy diffusivities j M = j H = j D = 0.023( NRe ) −0.2 (3-166)
are equal and that the molecular diffusivities are either for 10,000 < NRe = DG/ < 1,000,000
everywhere negligible or equal,
2. Average transport coefficients for flow across a flat
f h kc
= = (3-162) plate of length L:
2 C P ū x ū x
j M = j H = j D = 0.037( NRe ) −0.2 (3-167)
Equation (3-162) defines the Stanton number for heat
for 5 × 105 < NRe = Lu o / < 5 × 108
transfer,
h h 3. Average transport coefficients for flow normal to a
NStH = = (3-163) long, circular cylinder of diameter D, where the drag
C P ū x GC P
coefficient includes both form drag and skin friction,
where G = mass velocity = ū x , and the Stanton number but only the skin friction contribution applies to the
for mass transfer, analogy:
kc kc
NStM = = (3-164) ( j M ) skin friction = j H = j D = 0.193( NRe ) −0.382
ū x G
for 4,000 < NRe < 40,000 (3-168)
both of which are included in Table 3.13. ( j M ) skin friction = j H = j D = 0.0266( NRe ) −0.195
Equation (3-162) is referred to as the Reynolds analogy.
It can be used to estimate values of heat and mass transfer for 40,000 < NRe < 250,000 (3-169)
coefficients from experimental measurements of the DG
with NRe =
Fanning friction factor for turbulent flow, but only when
100 Chapter 3 Mass Transfer and Diffusion
1.000
Pa
cke
db
ed
0.100
j -factor
Sp
her
e
0.010
Cy
lin
der
Tube Flat p
flow late
0.001
1 10 100 1,000 10,000 100,000 1,000,000 10,000,000 Figure 3.17 Chilton–Colburn
Reynolds number j-factor correlations.
4. Average transport coefficients for flow past a single Other improvements were made by van Driest [64], who
sphere of diameter D: used a modified form of the Prandtl mixing length, Reichardt
( j M ) skin friction = j H = j D = 0.37( NRe ) −0.4 [65], who eliminated the zone concept by allowing the eddy
(3-170) diffusivities to decrease continuously from a maximum to
DG
for 20 < NRe = < 100,000 zero at the wall, and Friend and Metzner [57], who modified
the approach of Reichardt to obtain improved accuracy at
5. Average transport coefficients for flow through beds very high Prandtl and Schmidt numbers (to 3,000). Their
packed with spherical particles of uniform size DP: results for turbulent flow through a straight, circular tube are
f /2
j H = j D = 1.17( NRe ) −0.415 NStH = √ (3-172)
DP G (3-171) 1.20 + 11.8 f /2( NPr − 1)( NPr ) −1/3
for 10 < NRe = < 2,500 f /2
NStM = √ (3-173)
1.20 + 11.8 f /2( NSc − 1)( NSc ) −1/3
The above correlations are plotted in Figure 3.17, where the
curves do not coincide because of the differing definitions of Over a wide range of Reynolds number (10,000–
the Reynolds number. However, the curves are not widely 10,000,000), the Fanning friction factor is estimated from
separated. When using the correlations in the presence of the explicit empirical correlation of Drew, Koo, and
appreciable temperature and/or composition differences, McAdams [66],
Chilton and Colburn recommend that NPr and NSc be evalu-
f = 0.00140 + 0.125( NRe ) −0.32 (3-174)
ated at the average conditions from the surface to the bulk
stream. which is in excellent agreement with the experimental data
of Nikuradse [67] and is preferred over (3-165) with (3-166),
which is valid only to NRe = 1,000,000. For two- and three-
Other Analogies
dimensional turbulent-flow problems, some success has
New turbulence theories have led to improvements and ex- been achieved with the (kinetic energy of turbulence)–
tensions of the Reynolds analogy, resulting in expressions (rate of dissipation) model of Launder and Spalding [68],
for the Fanning friction factor and the Stanton numbers for which is widely used in computational fluid dynamics
heat and mass transfer that are less empirical than the (CFD) computer programs.
Chilton–Colburn analogy. The first major improvement was
by Prandtl [56] in 1910, who divided the flow into two re- Theoretical Analogy of Churchill and Zajic
gions: (1) a thin laminar-flow sublayer of thickness next to An alternative to (3-151) to (3-153) or the equivalent diffu-
the wall boundary, where only molecular transport occurs; sivity forms of (3-154) to (3-156) for the development of
and (2) a turbulent region dominated by eddy transport, with transport equations for turbulent flow is to start with the
M = H = D . time-averaged equations of Newton, Fourier, and Fick. For
Further important theoretical improvements to the example, let us derive a form of Newton’s law of viscosity
Reynolds analogy were made by von Karman, Martinelli, for molecular and turbulent transport of momentum in paral-
and Deissler, as discussed in detail by Knudsen and Katz lel. In a turbulent-flow field in the axial x-direction, instanta-
[47]. The first two investigators inserted a buffer zone be- neous velocity components, u x and u z , are
tween the laminar sublayer and turbulent core. Deissler
gradually reduced the eddy diffusivities as the wall was u x = ū x + u x
approached. u z = u z
3.5 Mass Transfer in Turbulent Flow 101
where the “overbarred” component is the time-averaged Equation (3-180) is a highly accurate quantitative represen-
(mean) local velocity and the primed component is the local tation of turbulent flow because it is based on experimental
fluctuating component that denotes instantaneous deviation data and numerical simulations described by Churchill and
from the local mean value. The mean velocity in the perpen- Zajic [70] and in considerable detail by Churchill [71]. From
dicular z-direction is zero. The mean local velocity in the x- (3-142) and (3-143), the shear stress at the wall, w , is
direction over a long period of time is given by related to the Fanning friction factor by
1 1
ū x = u x d = ( ū x + u x ) d (3-175) f =
2 w
(3-181)
0 0 ū 2x
The time-averaged fluctuating components u x and u z equal
zero. where ū x is the flow-average velocity in the axial direction.
The local instantaneous rate of momentum transfer by Combining (3-179) with (3-181) and performing the required
turbulence in the z-direction of x-direction turbulent mo- integrations, both numerically and analytically, lead to the
mentum per unit area at constant density is following implicit equation for the Fanning friction factor as
a function of the Reynolds number, NRe = 2a ū x /:
u z ( ū x + u x ) (3-176) 1/2 1/2 2
2 2
The time-average of this turbulent momentum transfer is 1/2
equal to the turbulent component of the shear stress, zxt , 2 f f
= 3.2 − 227 + 2500
f NRe NRe
zxt = u ( ū x + u x ) d 2 2
0 z (3-182)
(3-177) NRe
= u ( ū x ) d + u z (u x ) d 1
0 z 0 + ln 21/2
0.436 2
Because the time-average of the first term is zero, (3-177) f
reduces to
zxt = (u z u x ) (3-178) This equation is in excellent agreement with experimental
data over a Reynolds number range of 4,000–3,000,000 and
which is referred to as a Reynolds stress. Combining (3-178) can probably be used to a Reynolds number of 100,000,000.
with the molecular component of momentum transfer gives Table 3.14 presents a comparison of the Churchill–Zajic
the turbulent-flow form of Newton’s law of viscosity, equation, (3-182), with (3-174) of Drew et al. and (3-166)
du x of Chilton and Colburn. Equation (3-174) gives satis-
zx = − + (u z u x ) (3-179) factory agreement for Reynolds numbers from 10,000 to
dz
10,000,000, while (3-166) is useful only from 100,000 to
If (3-179) is compared to (3-151), it is seen that an alterna- 1,000,000.
tive approach to turbulence is to develop a correlating equa- Churchill and Zajic [70] show that if the equation for the
tion for the Reynolds stress, (u z u x ), first introduced by conservation of energy is time averaged, a turbulent-flow
Churchill and Chan [73], rather than an expression for a tur- form of Fourier’s law of conduction can be obtained with the
bulent viscosity t . This stress, which is a complex function fluctuation term (u z T ). Similar time averaging leads to a
of position and rate of flow, has been correlated quite accu- turbulent-flow form of Fick’s law of diffusion with (u z cA ).
rately for fully developed turbulent flow in a straight, circu- To extend (3-180) and (3-182) to obtain an expression for
lar tube by Heng, Chan, and Churchill [69]. In generalized the Nusselt number for turbulent-flow convective heat trans-
form, with a the radius of the tube and y = (a − z) the dis- fer in a straight, circular tube, Churchill and Zajic employ an
tance from the inside wall to the center of the tube, their analogy that is free of empircism, but not exact. The result
equation is
+ 3 −8/7
y −1
(u z u x ) = 0.7
++
+ exp
10 0.436y + Table 3.14 Comparison of Fanning Friction Factors for Fully
−7/8 Developed Turbulent Flow in a Smooth, Straight Circular Tube
1 6.95y + −8/7
− 1+
0.436a + a+ f, Drew et al. f, Chilton–Colburn f, Churchill–Zajic
NRe (3-174) (3-166) (3-182)
(3-180)
10,000 0.007960 0.007291 0.008087
where
100,000 0.004540 0.004600 0.004559
(u z u x ) ++ = − u z u x / 1,000,000 0.002903 0.002902 0.002998
a + = a( w ) 1/2 / 10,000,000 0.002119 0.001831 0.002119
100,000,000 0.001744 0.001155 0.001573
y + = y( w ) 1/2 /
102 Chapter 3 Mass Transfer and Diffusion
for Prandtl numbers greater than 1 is Chilton–Colburn correlation, which is widely used, is within
10% of the more theoretically based Churchill–Zajic equa-
1
NNu = 2/3 (3-183) tion for Reynolds numbers up to 1,000,000. However, beyond
NPrt 1 NPrt 1 that, serious deviations occur (25% at NRe = 10,000,000 and
+ 1−
NPr NNu1 NPr NNu∞ almost 50% at NRe = 100,000,000). Deviations of the
Friend–Metzner correlation from the Churchill–Zajic equa-
where, from Yu, Ozoe, and Churchill [72], tion vary from about 15% to 30% over the entire range of
Reynolds number in Table 3.15. At all Reynolds numbers,
0.015
NPrt = turbulent Prandtl number = 0.85 + (3-184) the Churchill–Zajic equation predicts higher Nusselt num-
NPr bers and, therefore, higher heat-transfer coefficients.
which replaces (u z T ), as introduced by Churchill [74], At a Prandtl number of 1,000, which is typical of high-
viscosity liquids, the Friend–Metzner correlation is in fairly
NNu1 = Nusselt number for ( NPr = NPrt )
close agreement with the Churchill–Zajic equation, predict-
f ing values from 6 to 13% higher. The Chilton–Colburn cor-
NRe
2 relation is seriously in error over the entire range of
= −5/4 (3-185)
2 Reynolds number, predicting values ranging from 74 to 27%
1 + 145 of those from the Churchill–Zajic equation as the Reynolds
f
number increases. It is clear that the Chilton–Colburn corre-
NNu∞ = Nusselt number for ( NPr = ∞) lation should not be used at high Prandtl numbers for heat
1/2 transfer or (by analogy) at high Schmidt numbers for mass
NPr 1/3 f
= 0.07343 NRe (3-186) transfer.
NPrt 2
The Churchill–Zajic equation for predicting the Nusselt
The accuracy of (3-183) is due to (3-185) and (3-186), which number provides an effective power dependence on the
are known from theoretical considerations. Although (3-184) Reynolds number as the Reynolds number increases. This is
is somewhat uncertain, its effect is negligible. in contrast to the typically cited constant exponent of 0.8, as
A comparison of the Churchill et al. correlation of in the Chilton–Colburn correlation. For the Churchill–Zajic
(3-183) with the Nusselt forms of (3-172) of Friend and equation, at a Prandtl number of 1, the exponent increases
Metzner and (3-166) of Chilton and Colburn, where from with Reynolds number from 0.79 to 0.88; at a Prandtl num-
Table 3.13, NNu = NSt NRe NPr , is given in Table 3.15 for a ber of 1,000, the exponent increases from 0.87 to 0.93.
wide range of Reynolds number and Prandtl numbers of Extension of the Churchill–Zajic equation to low Prandtl
1 and 1,000. numbers, typical of molten metals, and to other geometries,
In Table 3.15, at a Prandtl number of 1, which is typical of such as parallel plates, is discussed by Churchill [71], who
low-viscosity liquids and close to that of most gases, the also considers the important effect of boundary conditions
Table 3.15 Comparison of Nusselt Numbers for Fully Developed Turbulent Flow in a
Smooth, Straight Circular Tube
(e.g., constant wall temperature and uniform heat flux) at Churchill–Zajic equation:
low-to-moderate Prandtl numbers. Using mass-transfer analogs,
For calculation of convective mass-transfer coefficients,
kc , for turbulent flow of gases and liquids in straight, (3-184) gives NSct = 0.850
smooth, circular tubes, it is recommended that the (3-185) gives NSh1 = 94
Churchill–Zajic equation be employed by applying the (3-186) gives NSh∞ = 1686
analogy between heat and mass transfer. Thus, as illustrated
(3-183) gives NSh = 1680
in the following example, in (3-183) to (3-186), from
Table 3.13, the Sherwood number, NSh , is substituted for the From Table 3.13,
Nusselt number, NNu ; and the Schmidt number, NSc , is sub-
stituted for the Prandtl number, NPr . NSh 1680
NStM = = = 0.0000324,
NRe NSc (35800)(1450)
pA pA
Bulk liquid
Liquid
film
cAi
cAi
Gas Gas Well-mixed
bulk region
cAb
at cAb
Interfacial
cAb region
gaseous A. Because the gas is pure A at total pressure the bulk liquid. If the diffusivity of SO2 in water is
P = pA , there is no resistance to mass transfer in the gas 1.7 × 10−5 cm2 /s, determine the mass-transfer coefficient, kc , and
phase. At the gas–liquid interface, phase equilibrium is as- the film thickness, neglecting the bulk-flow effect.
sumed so the concentration of A, cAi , is related to the partial
pressure of A, pA , by some form of Henry’s law, for exam- SOLUTION
ple, cAi = HA pA . In the thin, stagnant liquid film of thick-
ness , molecular diffusion only occurs with a driving force 0.027(1,000) mol
NSO2 = 2
= 7.5 × 10−7 2
of cAi − cAb . Since the film is assumed to be very thin, all of (3,600)(100) cm -s
the diffusing A passes through the film and into the bulk liq- For dilute conditions, the concentration of water is
uid. If, in addition, bulk flow of A is neglected, the concen- 1
tration gradient is linear as in Figure 3.18a. Accordingly, c= = 5.55 × 10−2 mol/cm3
18.02
Fick’s first law, (3-3a), for the diffusion flux integrates to
From (3-188),
DAB cDAB
JA = (cAi − cAb ) = (xAi − xAb ) (3-187) DAB NA
kc = =
c(xAi − xAb )
If the liquid phase is dilute in A, the bulk-flow effect can be 7.5 × 10−7
neglected and (3-187) applies to the total flux: = = 6.14 × 10−3 cm/s
5.55 × 10−2 (0.0025 − 0.0003)
DAB cDAB
NA = (cAi − cAb ) = (xAi − xAb ) (3-188) Therefore,
DAB 1.7 × 10−5
= = = 0.0028 cm
If the bulk-flow effect is not negligible, then, from (3-31), kc 6.14 × 10−3
cDAB 1 − xAb cDAB which is very small and typical of turbulent-flow mass-transfer
NA = ln = (xAi − xAb ) processes.
1 − xAi (1 − xA ) LM
(3-189)
where Penetration Theory
xAi − xAb A more realistic physical model of mass transfer from a
(1 − xA ) LM = = (xB ) LM fluid–fluid interface into a bulk liquid stream is provided by
ln[(1 − xAb )/(1 − xAi )]
the penetration theory of Higbie [59], shown schematically
(3-190) in Figure 3.18b. The stagnant-film concept is replaced by
Boussinesq eddies that, during a cycle, (1) move from the
In practice, the ratios DAB / in (3-188) and DAB /
bulk to the interface; (2) stay at the interface for a short,
(1 − xA ) LM in (3-189) are replaced by mass transfer coeffi-
fixed period of time during which they remain static so that
cients kc and kc , respectively, because the film thickness, ,
molecular diffusion takes place in a direction normal to the
which depends on the flow conditions, is not known and the
interface; and (3) leave the interface to mix with the bulk
subscript, c, refers to a concentration driving force.
stream. When an eddy moves to the interface, it replaces an-
The film theory, which is easy to understand and apply, is
other static eddy. Thus, the eddies are alternately static and
often criticized because it appears to predict that the rate of
moving. Turbulence extends to the interface.
mass transfer is directly proportional to the molecular diffu-
In the penetration theory, unsteady-state diffusion takes
sivity. This dependency is at odds with experimental data,
place at the interface during the time the eddy is static. This
which indicate a dependency of D n , where n ranges from
process is governed by Fick’s second law, (3-68), with
about 0.5 to 0.75. However, if DAB / is replaced with kc ,
boundary conditions
which is then estimated from the Chilton–Colburn analogy,
Eq. (3-165), we obtain kc proportional to DAB 2/3
, which is in cA = cAb at t = 0 for 0 ≤ z ≤ ∞;
better agreement with experimental data. In effect, de- cA = cAi at z = 0 for t > 0; and
pends on DAB (or NSc ). Regardless of whether the criticism cA = cAb at z = ∞ for t > 0
of the film theory is valid, the theory has been and continues
to be widely used in the design of mass-transfer separation These are the same boundary conditions as in unsteady-state
equipment. diffusion in a semi-infinite medium. Thus, the solution can
be written by a rearrangement of (3-75):
cAi − cA z
EXAMPLE 3.17 = erf √ (3-191)
cAi − cAb 2 DAB tc
Sulfur dioxide is absorbed from air into water in a packed absorp-
tion tower. At a certain location in the tower, the mass-transfer flux where tc = “contact time” of the static eddy at the interface
is 0.0270 kmol SO2 /m2 -h and the liquid-phase mole fractions are during one cycle. The corresponding average mass-transfer
0.0025 and 0.0003, respectively, at the two-phase interface and in flux of A in the absence of bulk flow is given by the
3.6 Models for Mass Transfer at a Fluid–Fluid Interface 105
Fraction of
F{t} exit stream
φ {t} older than t1
Total Figure 3.19 Residence-time distrib-
0 area = 1 ution plots: (a) typical F curve;
0 (b) typical age distribution.
0 t 0 t1
[Adapted from O. Levenspiel, Chemical
t t
Reaction Engineering, 2nd ed., John Wiley
(a) (b) and Sons, New York (1972).]
106 Chapter 3 Mass Transfer and Diffusion
The instantaneous mass-transfer rate for an eddy with an From (3-196), the residence-time distribution is given by
age t is given by (3-192) for the penetration theory in flux
F{t} = 1 − e−t/0.45 , (2)
form as
where t is in seconds. Equations (1) and (2) are plotted in
DAB Figure 3.20. These curves are much different from the curves
NAt = (cAi − cAb ) (3-198)
t of Figure 3.19.
1
1 = 2.22 s–1
t
Area = t φ {t} 1 e–t / t
F{t} t
1 – e–t / t
Area = 1
0 t 0 t
0 t = 0.45 s 0 t = 0.45 s Figure 3.20 Age distribution curves for
(a) (b) Example 3.19: (a) F curve; (b) {t} curve.
3.7 Two-Film Theory and Overall Mass-Transfer Coefficients 107
at different rates. Equations (3-202) and (3-203) become, the Marangoni effect, is discussed in some detail by Bird,
respectively, Stewart, and Lightfoot [28], who cite additional references.
The effect can occur at both vapor–liquid and liquid–liquid
NAavg = kc (cAi − cAb ) = (cAi − cAb )(s DAB ) 1/2
interfaces, with the latter receiving the most attention. By
∞
s adding surfactants, which tend to concentrate at the inter-
× 1+2 exp −2n (3-204)
n=1
DAB face, the Marangoni effect may be reduced because of stabi-
lization of the interface, even to the extent that an interfacial
DAB mass-transfer resistance may result, causing the overall
NAavg = kc (cAi − cAb ) = (cAi − cAb )
mass-transfer coefficient to be reduced. In this book, unless
otherwise indicated, the Marangoni effect will be ignored
∞ (3-205)
1 and phase equilibrium will always be assumed at the phase
× 1 + 2
DAB interface.
n=1 1 + n 2 2
s2
In the limit, for a high rate of surface renewal, s2 /DAB ,
Gas–Liquid Case
(3-204) reduces to the surface-renewal theory, (3-200). For Consider the steady-state mass transfer of A from a gas
low rates of renewal, (3-205) reduces to the film theory, phase, across an interface, into a liquid phase. It could be
n
(3-188). At conditions in between, kc is proportional to DAB , postulated, as shown in Figure 3.21a, that a thin gas film ex-
where n is in the range of 0.5–1.0. The application of the ists on one side of the interface and a thin liquid film exists
film-penetration theory is difficult because of lack of data for on the other side, with the controlling factors being molecu-
and s, but the predicted effect of molecular diffusivity lar diffusion through each film. However, this postulation is
brackets experimental data. not necessary, because instead of writing
( DAB ) G ( DAB ) L
NA = (cAb − cAi ) G = (cAi − cAb ) L
3.7 TWO-FILM THEORY AND OVERALL G L
MASS-TRANSFER COEFFICIENTS (3-206)
Separation processes that involve contacting two fluid we can express the rate of mass transfer in terms of mass-
phases require consideration of mass-transfer resistances in transfer coefficients determined from any suitable theory,
both phases. In 1923, Whitman [63] suggested an extension with the concentration gradients visualized more realisti-
of the film theory to two fluid films in series. Each film pre- cally as in Figure 3.21b. In addition, we can use any number
sents a resistance to mass transfer, but concentrations in the of different mass-transfer coefficients, depending on the se-
two fluids at the interface are assumed to be in phase equi- lection of the driving force for mass transfer. For the gas
librium. That is, there is no additional interfacial resistance phase, under dilute or equimolar counter diffusion (EMD)
to mass transfer. This concept has found extensive applica- conditions, we write the mass-transfer rate in terms of partial
tion in modeling of steady-state, gas–liquid, and liquid– pressures:
liquid separation processes.
The assumption of phase equilibrium at the phase inter- NA = k p ( pAb − pAi ) (3-207)
face, while widely used, may not be valid when gradients of where kp is a gas-phase mass-transfer coefficient based on a
interfacial tension are established during mass transfer be- partial-pressure driving force.
tween two fluids. These gradients give rise to interfacial tur- For the liquid phase, we use molar concentrations:
bulence resulting, most often, in considerably increased
mass-transfer coefficients. This phenomenon, referred to as NA = kc (cAi − cAb ) (3-208)
cAi cAi
cAb cAb
Transport Transport
Figure 3.21 Concentration
gradients for two-resistance
theory: (a) film theory; (b) more
(a) (b) realistic gradients.
108 Chapter 3 Mass Transfer and Diffusion
At the phase interface, cAi and pAi are assumed to be in Alternatively, (3-207) to (3-209) can be combined to
phase equilibrium. Applying a version of Henry’s law differ- define an overall mass-transfer coefficient, KG, based on the
ent from that in Table 2.3,1 gas phase. The result is
cAi = HA pAi (3-209) pAb − cAb /HA
NA = (3-216)
(1/k p ) + (1/HA kc )
Equations (3-207) to (3-209) are a commonly used combina-
tion for vapor–liquid mass transfer. Computations of mass- In this case, it is customary to define: (1) a fictitious gas-
transfer rates are generally made from a knowledge of bulk phase partial pressure pA∗ = cAb /HA , which is the partial
concentrations, which in this case are pAb and cAb . To obtain pressure that would be in equilibrium with the bulk liquid;
an expression for NA in terms of an overall driving force for and (2) an overall mass-transfer coefficient for the gas phase,
mass transfer, (3-207) to (3-209) are combined in the fol- KG, based on a partial-pressure driving force. Thus, (3-216)
lowing manner to eliminate the interfacial concentrations, can be rewritten as
cAi and pAi . Solve (3-207) for pAi : ( pAb − pA∗ )
NA = K G ( pAb − pA∗ ) = (3-217)
NA (1/k p ) + (1/HA kc )
pAi = pAb − (3-210)
kp where
Solve (3-208) for cAi : 1 1 1
= + (3-218)
NA KG kp HA kc
cAi = cAb + (3-211)
kc In this, the resistances are 1/kp and 1/(HAkc). When
1/kp >> 1/HAkc,
Combine (3-211) with (3-209) to eliminate cAi and combine
the result with (3-210) to eliminate pAi to give NA = k p ( pAb − pA∗ ) (3-219)
pAb HA − cAb Since the resistance in the liquid phase is then negligible, the
NA = (3-212)
( HA /k p ) + (1/kc ) liquid-phase driving force is cAi − cAb ≈ 0 and cAi ≈ cAb .
The choice between using (3-213) or (3-217) is arbitrary,
It is customary to define: (1) a fictitious liquid-phase but is usually made on the basis of which phase has the
concentration cA∗ = pAb HA , which is the concentration that largest mass-transfer resistance; if the liquid, use (3-213); if
would be in equilibrium with the partial pressure in the bulk the gas, use (3-217). Another common combination for
gas; and (2) an overall mass-transfer coefficient, KL. Thus, vapor–liquid mass transfer uses mole fraction-driving forces,
(3-212) is rewritten as which define another set of mass-transfer coefficients:
(cA∗ − cAb ) NA = k y ( yAb − yAi ) = k x (xAi − xAb )
NA = K L (cA∗ − cAb ) = (3-213) (3-220)
( HA /k p ) + (1/kc )
In this case, phase equilibrium at the interface can be
where expressed in terms of the K-value for vapor–liquid equilib-
1 HA 1 rium. Thus,
= + (3-214)
KL kp kc K A = yAi /xAi (3-221)
in which KL is the overall mass-transfer coefficient based on Combining (3-220) and (3-221) to eliminate yAi and xAi ,
the liquid phase. The quantities HA/kp and 1/kc are measures yAb − xAb
NA = (3-222)
of the mass-transfer resistances of the gas phase and the (1/K A k y ) + (1/k x )
liquid phase, respectively. When 1/kc >> HA/kp, (3-214)
This time we define fictitious concentration quantities and
becomes
overall mass-transfer coefficients for mole-fraction driving
NA = kc (cA∗ − cAb ) (3-215) forces. Thus, xA∗ = yAb /K A and yA∗ = K A xAb . If the two
Since resistance in the gas phase is then negligible, the gas- values of KA are equal, we obtain
phase driving force is pAb − pAi ≈ 0 and pAb ≈ pAi . xA∗ − xAb
NA = K x (xA∗ − xAb ) = (3-223)
(1/K A k y ) + (1/k x )
1
Many different forms of Henry’s law are found in the literature. They and
include
yAb − yA∗
pA = HA xA , pA =
cA
, and yA = HA xA NA = K y ( yAb − yA∗ ) = (3-224)
HA (1/k y ) + (K A /k x )
When a Henry’s-law constant, HA, is given without citing the equation that
defines it, the defining equation can be determined from the units of the
where Kx and Ky are overall mass-transfer coefficients based
constant. For example, if the constant has the units of atm or atm/mole on mole-fraction driving forces with
fraction, Henry’s law is given by pA = HAxA. If the units are mol/L-mmHg, 1 1 1
cA
Henry’s law is pA = . = + (3-225)
HA Kx KAk y kx
3.7 Two-Film Theory and Overall Mass-Transfer Coefficients 109
Table 3.16 Relationships among Mass-Transfer Coefficients Case of Large Driving Forces for Mass Transfer
Equimolar Counterdiffusion (EMD): When large driving forces exist for mass transfer, phase
equilibria ratios such as HA, KA, and K DA may not be con-
Gases: NA = k y yA = kc cA = k p pA stant across the two phases. This occurs particularly when
P one or both phases are not dilute with respect to the diffusing
k y = kc = k p P if ideal gas
RT solute, A. In that case, expressions for the mass-transfer flux
Liquids: NA = k x xA = kc cA must be revised.
k x = kc c, where c = total molar For example, if mole-fraction driving forces are used, we
concentration (A + B)
write, from (3-220) and (3-224),
Unimolecular Diffusion (UMD): NA = k y ( yAb − yAi ) = K y ( yAb − yA∗ ) (3-234)
Gases: Same equations as for EMD with k replaced Thus,
by k =
k 1 yAb − yA∗
( yB ) LM = (3-235)
Ky k y ( yAb − yAi )
Liquids: Same equations as for EMD with k replaced by or
k =
k 1 ( yAb − yAi ) + ( yAi − yA∗ ) 1 1 yAi − yA∗
= = +
(xB ) LM Ky k y ( yAb − yAi ) ky k y yAb − yAi
When using concentration units for both phases, it is convenient (3-236)
to use: From (3-220),
k G (cG ) = kc (c) for the gas phase
kx ( yAb − yAi )
k L (c L ) = kc (c) for the liquid phase = (3-237)
ky (xAi − xAb )
110 Chapter 3 Mass Transfer and Diffusion
x
m
e
* is a fictitious x in
xA
op A 0.1700 0.00676
equilibrium with yAb.
Sl
yA *
yA
* is a fictitious y in
yA A Experimental values of the mass transfer coefficients are as
equilibrium with xAb.
xAb xAi * follows.
xA
Liquid phase: kc = 0.18 m/h
kmol
Gas phase: k p = 0.040
xA
h-m2 -kPa
Figure 3.22 Curved equilibrium line. Using mole-fraction driving forces, compute the mass-transfer
flux by:
(a) Assuming an average Henry’s-law constant and a negligible
bulk-flow effect.
Combining (3-234) and (3-237),
(b) Utilizing the actual curved equilibrium line and assuming a
1 1 1 yAi − yA∗ negligible bulk-flow effect.
= + (3-238)
Ky ky k x xAi − xAb (c) Utilizing the actual curved equilibrium line and taking into
account the bulk-flow effect.
In a similar manner,
In addition,
1 1 1 xA∗ − xAi (d) Determine the relative magnitude of the two resistances and
= + (3-239) the values of the mole fractions at the interface from the results
Kx kx ky yAb − yAi
of part (c).
A typical curved equilibrium line is shown in Figure 3.22
with representative values of yAb , yAi , yA∗ , xA∗ , xAi , and xAb
indicated. Because the line is curved, the vapor–liquid equi- SOLUTION
librium ratio, K A = yA /xA , is not constant across the two The equilibrium data are converted to mole fractions by assuming
phases. As shown, the slope of the curve and thus, KA, de- Dalton’s law, yA = pA/P, for the gas and using xA = cA/c for the
creases with increasing concentration of A. Denote two liquid. The concentration of the liquid is close to that of pure water
slopes of the equilibrium line by or 3.43 lbmol/ft3 or 55.0 kmol/m3. Thus, the mole fractions at equi-
librium are:
yAi − yA∗
mx = (3-240) ySO2 xSO2
xAi − xAb
0.0191 0.000563
and 0.0303 0.000846
0.0546 0.001408
yAb − yAi
my = (3-241) 0.0850 0.001971
xA∗ − xAi
These data are fitted with average and maximum absolute devia-
Substituting (3-240) and (3-241) into (3-238) and (3-239), tions of 0.91% and 1.16%, respectively, by the quadratic equation
respectively, gives
ySO2 = 29.74xSO2 + 6,733xSO
2
2
(1)
1 1 mx
= + (3-242) Thus, differentiating, the slope of the equilibrium curve is given by
Ky ky kx
dy
m= = 29.74 + 13,466xSO2 (2)
and dx
1 1 1 The given mass-transfer coefficients can be converted to kx and ky
= + (3-243)
Kx kx m y ky by (3-227) and (3-228):
kmol
k x = kc c = 0.18(55.0) = 9.9
h-m2
EXAMPLE 3.20 kmol
k y = k p P = 0.040(2)(101.3) = 8.1 .
h-m2
Sulfur dioxide (A) is absorbed into water in a packed column. At a
certain location, the bulk conditions are 50°C, 2 atm, yAb = 0.085, (a) From (1) for xAb = 0.001, yA∗ = 29.74(0.001) + 6,733(0.001) 2
and xAb = 0.001. Equilibrium data for SO2 between air and = 0.0365. From (1), for yAb = 0.085, we solve the quadratic
water at 50°C are equation to obtain xA∗ = 0.001975.
Summary 111
SUMMARY
1. Mass transfer is the net movement of a component in a mixture 5. Diffusivity values vary by orders of magnitude. Typical values
from one region to another region of different concentration, often are 0.10, 1 × 10−5 , and 1 × 10−9 cm2 /s for ordinary molecular
between two phases across an interface. Mass transfer occurs by diffusion of a solute in a gas, liquid, and solid, respectively.
molecular diffusion, eddy diffusion, and bulk flow. Molecular dif- 6. Fick’s second law for unsteady-state diffusion is readily ap-
fusion occurs by a number of different driving forces, including plied to semi-infinite and finite stagnant media, including certain
concentration (the most important), pressure, temperature, and ex- anisotropic materials.
ternal force fields.
7. Molecular diffusion under laminar-flow conditions can be deter-
2. Fick’s first law for steady-state conditions states that the mass- mined from Fick’s first and second laws, provided that velocity pro-
transfer flux by ordinary molecular diffusion is equal to the product files are available. Common cases include falling liquid-film flow,
of the diffusion coefficient (diffusivity) and the negative of the con- boundary-layer flow on a flat plate, and fully developed flow in a
centration gradient. straight, circular tube. Results are often expressed in terms of a mass-
3. Two limiting cases of mass transfer are equimolar counterdif- transfer coefficient embedded in a dimensionless group called the
fusion (EMD) and unimolecular diffusion (UMD). The former is Sherwood number. The mass-transfer flux is given by the product of
also a good approximation for dilute conditions. The latter must in- the mass-transfer coefficient and a concentration driving force.
clude the bulk-flow effect. 8. Mass transfer in turbulent flow is often predicted by analogy
4. When experimental data are not available, diffusivities in gas to heat transfer. Of particular importance is the Chilton–Colburn
and liquid mixtures can be estimated. Diffusivities in solids, in- analogy, which utilizes empirical j-factor correlations and the
cluding porous solids, crystalline solids, metals, glass, ceramics, dimensionless Stanton number for mass transfer. A more accurate
polymers, and cellular solids, are best measured. For some solids— equation by Churchill and Zajic should be used for flow in tubes,
for example, wood—diffusivity is an anisotropic property. particularly at high Schmidt and Reynolds numbers.
112 Chapter 3 Mass Transfer and Diffusion
9. A number of models have been developed for mass transfer 10. The two-film theory of Whitman (more properly referred to as
across a two-fluid interface and into a liquid. These include the film a two-resistance theory) is widely used to predict the mass-transfer
theory, penetration theory, surface-renewal theory, and the film- flux from one fluid phase, across an interface, and into another fluid
penetration theory. These theories predict mass-transfer coeffi- phase, assuming equilibrium at the interface. One resistance is often
cients that are proportional to the diffusivity raised to an exponent controlling. The theory defines an overall mass-transfer coefficient
that varies from 0.5 to 1.0. Most experimental data provide expo- that is determined from the separate coefficients for each of the two
nents ranging from 0.5 to 0.75. phases and the equilibrium relationship at the interface.
REFERENCES
1. TAYLOR, R., and R. KRISHNA, Multicomponent Mass Transfer, John 28. BIRD, R.B., W.E. STEWART, and E.N. LIGHTFOOT, Transport Phenom-
Wiley and Sons, New York (1993). ena, 2nd ed., John Wiley and Sons, New York (2002).
2. POLING, B.E., J.M. PRAUSNITZ, and J.P. O’CONNELL, The Properties of 29. CHURCHILL, R.V., Operational Mathematics, 2nd ed., McGraw-Hill,
Liquids and Gases, 5th ed., McGraw-Hill, New York (2001). New York (1958).
3. FULLER, E.N., P.D. SCHETTLER, and J.C. GIDDINGS, Ind. Eng. Chem., 30. ABRAMOWITZ, M., and I.A. STEGUN, Eds., Handbook of Mathematical
58 (5), 18–27 (1966). Functions, National Bureau of Standards, Applied Mathematics Series 55,
4. TAKAHASHI, S., J. Chem. Eng. Jpn., 7, 417–420 (1974). Washington, DC (1964).
5. SLATTERY, J.C., M.S. thesis, University of Wisconsin, Madison 31. NEWMAN, A.B., Trans. AIChE, 27, 310–333 (1931).
(1955). 32. GRIMLEY, S.S., Trans. Inst. Chem. Eng. (London), 23, 228–235
6. WILKE, C.R., and P. CHANG, AIChE J., 1, 264–270 (1955). (1948).
7. HAYDUK, W., and B.S. MINHAS, Can. J. Chem. Eng., 60, 295–299 33. JOHNSTONE, H.F., and R.L. PIGFORD, Trans. AIChE, 38, 25–51 (1942).
(1982). 34. OLBRICH, W.E., and J.D. WILD, Chem. Eng. Sci., 24, 25–32 (1969).
8. QUAYLE, O.R., Chem. Rev., 53, 439–589 (1953). 35. CHURCHILL, S.W., The Interpretation and Use of Rate Data: The Rate
9. VIGNES, A., Ind. Eng. Chem. Fundam., 5, 189–199 (1966). Concept, McGraw-Hill, New York (1974).
10. SORBER, H.A., Handbook of Biochemistry, Selected Data for Molecu- 36. CHURCHILL, S.W., and R. USAGI, AIChE J., 18, 1121–1128 (1972).
lar Biology, 2nd ed., Chemical Rubber Co., Cleveland, OH (1970). 37. EMMERT, R.E., and R.L. PIGFORD, Chem. Eng. Prog., 50, 87–93
11. GEANKOPLIS, C.J., Transport Processes and Separation Process (1954).
Principles, 4th ed., Prentice-Hall, Upper Saddle River, NJ (2003). 38. PRANDTL, L., Proc. 3rd Int. Math. Congress, Heidelberg (1904);
12. FRIEDMAN, L., and E.O. KRAEMER, J. Am. Chem. Soc., 52, 1298–1304, reprinted in NACA Tech. Memo 452 (1928).
1305–1310, 1311–1314 (1930). 39. BLASIUS, H., Z. Math Phys., 56, 1–37 (1908); reprinted in NACA Tech.
13. BOUCHER, D.F., J.C. BRIER, and J.O. OSBURN, Trans. AIChE, 38, Memo 1256.
967–993 (1942). 40. SCHLICHTING, H., Boundary Layer Theory, 4th ed., McGraw-Hill,
14. BARRER, R.M., Diffusion in and through Solids, Oxford University New York (1960).
Press, London (1951). 41. POHLHAUSEN, E., Z. Angew. Math Mech., 1, 252 (1921).
15. SWETS, D.E., R.W. LEE, and R.C. FRANK, J. Chem. Phys., 34, 17–22 42. POHLHAUSEN, E., Z. Angew. Math Mech., 1, 115–121 (1921).
(1961). 43. LANGHAAR, H.L., Trans. ASME, 64, A-55 (1942).
16. LEE, R. W., J. Chem, Phys., 38, 44–455 (1963). 44. GRAETZ, L., Ann. d. Physik, 25, 337–357 (1885).
17. WILLIAMS, E.L., J. Am. Ceram. Soc., 48, 190–194 (1965). 45. SELLARS, J.R., M. TRIBUS, and J.S. KLEIN, Trans. ASME, 78, 441–448
18. SUCOV, E.W., J. Am. Ceram. Soc., 46, 14–20 (1963). (1956).
19. KINGERY, W.D., H.K. BOWEN, and D.R. UHLMANN, Introduction to 46. LEVEQUE, J., Ann. Mines, [12], 13, 201, 305, 381 (1928).
Ceramics, 2nd ed., John Wiley and Sons, New York (1976). 47. KNUDSEN, J.G., and D.L. KATZ, Fluid Dynamics and Heat Transfer,
20. FERRY, J.D., Viscoelastic Properties of Polymers, John Wiley and McGraw-Hill, New York (1958).
Sons, New York (1980). 48. HAUSEN, H., Verfahrenstechnik Beih. z. Ver. Deut. Ing., 4, 91 (1943).
21. RHEE, C.K., and J.D. FERRY, J. Appl. Polym. Sci., 21, 467–476 (1977). 49. LINTON, W.H. Jr., and T.K. SHERWOOD, Chem. Eng. Prog., 46, 258–264
22. BRANDRUP, J., and E.H. IMMERGUT, Eds., Polymer Handbook, 3rd ed., (1950).
John Wiley and Sons, New York (1989). 50. REYNOLDS, O., Trans. Roy. Soc. (London), 174A, 935–982 (1883).
23. GIBSON, L.J., and M.F. ASHBY, Cellular Solids, Structure and Proper- 51. BOUSSINESQ, J., Mem. Pre. Par. Div. Sav., XXIII, Paris (1877).
ties, Pergamon Press, Elmsford, NY (1988).
52. PRANDTL, L., Z. Angew, Math Mech., 5, 136 (1925); reprinted in
24. STAMM, A.J., Wood and Cellulose Science, Ronald Press, New York NACA Tech. Memo 1231 (1949).
(1964).
53. REYNOLDS, O., Proc. Manchester Lit. Phil. Soc., 14, 7 (1874).
25. SHERWOOD, T.K., Ind. Eng. Chem., 21, 12–16 (1929).
54. COLBURN, A.P., Trans. AIChE, 29, 174–210 (1933).
26. CARSLAW, H.S., and J.C. JAEGER, Heat Conduction in Solids, 2nd ed.,
55. CHILTON, T.H., and A.P. COLBURN, Ind. Eng. Chem., 26, 1183–1187
Oxford University Press, London (1959).
(1934).
27. CRANK, J., The Mathematics of Diffusion, Oxford University Press,
56. PRANDTL, L., Physik. Z., 11, 1072 (1910).
London (1956).
Exercises 113
57. FRIEND, W.L., and A.B. METZNER, AIChE J., 4, 393–402 (1958). 68. LAUNDER, B.E., and D.B. SPALDING, Lectures in Mathematical Models
58. NERNST, W., Z. Phys. Chem., 47, 52 (1904). of Turbulence, Academic Press, New York (1972).
59. HIGBIE, R., Trans. AIChE, 31, 365–389 (1935). 69. HENG, L., C. CHAN, and S.W. CHURCHILL, Chem. Eng. J., 71, 163
(1998).
60. DANCKWERTS, P.V., Ind. Eng. Chem., 43, 1460–1467 (1951).
70. CHURCHILL, S.W., and S.C. ZAJIC, AIChE J., 48, 927–940 (2002).
61. LEVENSPIEL, O., Chemical Reaction Engineering, 3rd ed., John Wiley
and Sons, New York (1999). 71. CHURCHILL, S.W., “Turbulent Flow and Convection: The Prediction of
Turbulent Flow and Convection in a Round Tube,” in Advances in Heat
62. TOOR, H.L., and J.M. MARCHELLO, AIChE J., 4, 97–101 (1958). Transfer, J.P. Hartnett, and T.F. Irvine, Jr., Ser. Eds., Academic Press, New
63. WHITMAN, W.G., Chem. Met. Eng., 29, 146–148 (1923). York, 34, 255–361 (2001).
64. vAN DRIEST, E.R., J. Aero Sci., 1007–1011 and 1036 (1956). 72. YU, B., H. OZOE, and S.W. CHURCHILL, Chem. Eng. Sci., 56, 1781
65. REICHARDT, H., Fundamentals of Turbulent Heat Transfer, NACA (2001).
Report TM-1408 (1957). 73. CHURCHILL, S.W., and C. CHAN, Ind. Eng. Chem. Res., 34, 1332
66. DREW, T.B., E.C. KOO, and W.H. MCADAMS, Trans. Am. Inst. Chem. (1995).
Engrs., 28, 56 (1933). 74. CHURCHILL, S.W., AIChE J., 43, 1125 (1997).
67. NIKURADSE, J., VDI-Forschungsheft, p. 361 (1933).
EXERCISES
Section 3.1 level was originally 0.5 in. below the top of the tube. The diffusiv-
3.1 A beaker filled with an equimolar liquid mixture of ethyl ity of water in air at 25°C is 0.256 cm2/s.
alcohol and ethyl acetate evaporates at 0°C into still air at 101 kPa (a) How long will it take for the liquid level in the tube to drop 3 in.?
(1 atm) total pressure. Assuming Raoult’s law applies, what will be (b) Make a plot of the liquid level in the tube as a function of time
the composition of the liquid remaining when half the original ethyl for this period.
alcohol has evaporated, assuming that each component evaporates 3.5 Two bulbs are connected by a tube, 0.002 m in diameter and
independently of the other? Also assume that the liquid is always 0.20 m in length. Initially bulb 1 contains argon, and bulb 2 con-
well mixed. The following data are available: tains xenon. The pressure and temperature are maintained at 1 atm
and 105°C, at which the diffusivity is 0.180 cm2/s. At time t = 0,
Vapor Pressure, Diffusivity in Air diffusion is allowed to occur between the two bulbs. At a later time,
kPa at 0°C m2/s the argon mole fraction in the gas at end 1 of the tube is 0.75, and
Ethyl acetate (AC) 3.23 6.45 × 10−6 0.20 at the other end. Determine at the later time:
Ethyl alcohol (AL) 1.62 9.29 × 10−6 (a) The rates and directions of mass transfer of argon and xenon
(b) The transport velocity of each species
3.2 An open tank, 10 ft in diameter and containing benzene at (c) The molar average velocity of the mixture
25°C, is exposed to air in such a manner that the surface of the liq-
uid is covered with a stagnant air film estimated to be 0.2 in. thick. Section 3.2
If the total pressure is 1 atm and the air temperature is 25°C, what 3.6 The diffusivity of toluene in air was determined experimen-
loss of material in pounds per day occurs from this tank? The spe- tally by allowing liquid toluene to vaporize isothermally into air
cific gravity of benzene at 60°F is 0.877. The concentration of ben- from a partially filled vertical tube 3 mm in diameter. At a temper-
zene at the outside of the film is so low that it may be neglected. For ature of 39.4°C, it took 96 × 104 s for the level of the toluene to
benzene, the vapor pressure at 25°C is 100 torr, and the diffusivity drop from 1.9 cm below the top of the open tube to a level of 7.9 cm
in air is 0.08 cm2/s. below the top. The density of toluene is 0.852 g/cm3, and the vapor
3.3 An insulated glass tube and condenser are mounted on a pressure is 57.3 torr at 39.4°C. The barometer reading was 1 atm.
reboiler containing benzene and toluene. The condenser returns liq- Calculate the diffusivity and compare it with the value predicted
uid reflux so that it runs down the wall of the tube. At one point in from (3-36). Neglect the counterdiffusion of air.
the tube the temperature is 170°F, the vapor contains 30 mol% 3.7 An open tube, 1 mm in diameter and 6 in. long, has pure hy-
toluene, and the liquid reflux contains 40 mol% toluene. The effec- drogen blowing across one end and pure nitrogen blowing across
tive thickness of the stagnant vapor film is estimated to be 0.1 in. the other. The temperature is 75°C.
The molar latent heats of benzene and toluene are equal. Calculate
(a) For equimolar counterdiffusion, what will be the rate of trans-
the rate at which toluene and benzene are being interchanged by
fer of hydrogen into the nitrogen stream (mol/s)? Estimate the dif-
equimolar countercurrent diffusion at this point in the tube in
fusivity from (3-36).
lbmol/h-ft2.
(b) For part (a), plot the mole fraction of hydrogen against distance
Diffusivity of toluene in benzene = 0.2 ft2/h. from the end of the tube past which nitrogen is blown.
Pressure = 1 atm total pressure (in the tube). 3.8 Some HCl gas diffuses across a film of air 0.1 in. thick at
Vapor pressure of toluene at 170°F = 400 torr. 20°C. The partial pressure of HCl on one side of the film is 0.08
3.4 Air at 25°C with a dew-point temperature of 0°C flows past atm and it is zero on the other. Estimate the rate of diffusion, as mol
the open end of a vertical tube filled with liquid water maintained at HCl/s-cm2, if the total pressure is (a) 10 atm, (b) 1 atm, (c) 0.1 atm.
25°C. The tube has an inside diameter of 0.83 in., and the liquid The diffusivity of HCl in air at 20°C and 1 atm is 0.145 cm2/s.
114 Chapter 3 Mass Transfer and Diffusion
3.9 Estimate the diffusion coefficient for the gaseous binary sys- 3.16 Experimental liquid-phase activity-coefficient data are
tem nitrogen (A)/toluene (B) at 25°C and 3 atm using the method of given in Exercise 2.23 for the ethanol/benzene system at 45°C. Es-
Fuller et al. timate and plot diffusion coefficients for both ethanol and benzene
3.10 For the mixture of Example 3.3, estimate the diffusion coef- over the entire composition range.
ficient if the pressure is increased to 100 atm using the method of 3.17 Estimate the diffusion coefficient of NaOH in a 1-M aqueous
Takahashi. solution at 25°C.
3.11 Estimate the diffusivity of carbon tetrachloride at 25°C in a 3.18 Estimate the diffusion coefficient of NaCl in a 2-M aqueous
dilute solution of: (a) Methanol, (b) Ethanol, (c) Benzene, and solution at 18°C. Compare your estimate with the experimental
(d) n-Hexane by the method of Wilke–Chang and Hayduk–Minhas. value of 1.28 × 10−5 cm2 /s.
Compare the estimated values with the following experimental 3.19 Estimate the diffusivity of N2 in H2 in the pores of a catalyst
observations: at 300°C and 20 atm if the porosity is 0.45 and the tortuosity is 2.5.
Assume ordinary molecular diffusion in the pores.
Solvent Experimental DAB, cm2/s
3.20 Gaseous hydrogen at 150 psia and 80°F is stored in a small,
Methanol 1.69 × 10−5 cm2 /s at 15◦ C
spherical, steel pressure vessel having an inside diameter of 4 in. and
Ethanol 1.50 × 10−5 cm2 /s at 25◦ C
a wall thickness of 0.125 in. At these conditions, the solubility of hy-
Benzene 1.92 × 10−5 cm2 /s at 25◦ C
drogen in steel is 0.094 lbmol/ft3 and the diffusivity of hydrogen in
n-Hexane 3.70 × 10−5 cm2 /s at 25◦ C
steel is 3.0 × 10−9 cm2 /s. If the inner surface of the vessel remains
3.12 Estimate the liquid diffusivity of benzene (A) in formic acid saturated at the existing hydrogen pressure and the hydrogen partial
(B) at 25°C and infinite dilution. Compare the estimated value to pressure at the outer surface is assumed to be zero, estimate:
that of Example 3.6 for formic acid at infinite dilution in benzene. (a) The initial rate of mass transfer of hydrogen through the metal
3.13 Estimate the liquid diffusivity of acetic acid at 25°C in a wall
dilute solution of: (a) Benzene, (b) Acetone, (c) Ethyl acetate, and (b) The initial rate of pressure decrease inside the vessel
(d) Water by an appropriate method. Compare the estimated values (c) The time in hours for the pressure to decrease to 50 psia,
with the following experimental values: assuming the temperature stays constant at 80°F
3.21 A polyisoprene membrane of 0.8-m thickness is to be used
Solvent Experimental DAB, cm2/s
to separate a mixture of methane and H2. Using the data in
Benzene 2.09 × 10−5 cm2 /s at 25◦ C Table 14.9 and the following compositions, estimate the mass-
Acetone 2.92 × 10−5 cm2 /s at 25◦ C transfer flux of each of the two species.
Ethyl acetate 2.18 × 10−5 cm2 /s at 25◦ C
Water 1.19 × 10−5 cm2 /s at 20◦ C Partial Pressures, MPa
3.14 Water in an open dish exposed to dry air at 25°C is found to Membrane Side 1 Membrane Side 2
vaporize at a constant rate of 0.04 g/h-cm2. Assuming the water
surface to be at the wet-bulb temperature of 11.0°C, calculate the Methane 2.5 0.05
effective gas-film thickness (i.e., the thickness of a stagnant air film Hydrogen 2.0 0.20
that would offer the same resistance to vapor diffusion as is actually
encountered at the water surface). Section 3.3
3.15 Isopropyl alcohol is undergoing mass transfer at 35°C and 3.22 A 3-ft depth of stagnant water at 25°C lies on top of a
2 atm under dilute conditions through water, across a phase bound- 0.10-in. thickness of NaCl. At time < 0, the water is pure. At time =
ary, and then through nitrogen. Based on the date given below, 0, the salt begins to dissolve and diffuse into the water. If the con-
estimate for isopropyl alcohol: centration of salt in the water at the solid–liquid interface is main-
(a) The diffusivity in water using the Wilke–Chang equation tained at saturation (36 g NaCl/100 g H2O) and the diffusivity of
(b) The diffusivity in nitrogen using the Fuller et al. equation NaCl in water is 1.2 × 10−5 cm2 /s, independent of concentration,
(c) The product, DAB M , in water estimate, by assuming the water to act as a semi-infinite medium,
the time and the concentration profile of salt in the water when
(d) The product, DAB M , in air
(a) 10% of the salt has dissolved
where M is the molar density of the mixture.
(b) 50% of the salt has dissolved
Using the above results, compare:
(c) 90% of the salt has dissolved
(e) The diffusivities in parts (a) and (b)
(f) The diffusivity-molar density products in Parts (c) and (d) 3.23 A slab of dry wood of 4-in. thickness and sealed edges is
exposed to air of 40% relative humidity. Assuming that the two
Lastly: unsealed faces of the wood immediately jump to an equilibrium
(g) What conclusions can you come to about molecular diffusion moisture content of 10 lb H2O per 100 lb of dry wood, determine
in the liquid phase versus the gaseous phase? the time for the moisture to penetrate to the center of the slab (2 in.
Data: from either face). Assume a diffusivity of water in the wood as
8.3 × 10−6 cm2 /s.
Component Tc, °R Pc, psia Zc vL, cm3/mol 3.24 A wet, clay brick measuring 2 × 4 × 6 in. has an initial uni-
Nitrogen 227.3 492.9 0.289 — form moisture content of 12 wt%. At time = 0, the brick is exposed
Isopropyl alcohol 915 691 0.249 76.5 on all sides to air such that the surface moisture content is
Exercises 115
Temperature of inlet water, 25.15°C, Temperature of outlet water, The operation was countercurrent, with the gas entering at the bot-
25.35°C tom of the vertical tower and the acid passing down in a thin film
Partial pressure of water in inlet air, 6.27 torr, and in outlet air, on the inner wall. The change in acid strength was inappreciable,
20.1 torr and the vapor pressure of ammonia over the liquid may be as-
The value for the diffusivity of water vapor in air is 0.22 cm2/s at sumed to have been negligible because of the use of a strong
0°C and 1 atm. The mass velocity of air is taken relative to the pipe acid for absorption. Calculate the mass-transfer coefficient, kp,
wall. Calculate: from the data.
(a) Rate of mass transfer of water into the air 3.40 A new type of cooling-tower packing is being tested in a lab-
(b) KG for the wetted-wall column oratory column. At two points in the column, 0.7 ft apart, the fol-
lowing data have been taken. Calculate the overall volumetric
3.39 The following data were obtained by Chamber and mass-transfer coefficient Ky a that can be used to design a large,
Sherwood [Ind. Eng. Chem., 29, 1415 (1937)] on the absorption of packed-bed cooling tower, where a is the mass-transfer area, A, per
ammonia from an ammonia-air system by a strong acid in a wetted- unit volume, V, of tower.
wall column 0.575 in. in diameter and 32.5 in. long:
Inlet acid (2-N H2SO4) temperature, °F 76 Bottom Top
Outlet acid temperature, °F 81
Water temperature, °F 120 126
Inlet air temperature, °F 77
Water vapor pressure, psia 1.69 1.995
Outlet air temperature, °F 84 Mole fraction H2O in air 0.001609 0.0882
Total pressure, atm 1.00 Total pressure, psia 14.1 14.3
Partial pressure NH3 in inlet gas, atm 0.0807 Air rate, lbmol/h 0.401 0.401
Partial pressure NH3 in outlet gas, atm 0.0205 Column area, ft2 0.5 0.5
Air rate, lbmol/h 0.260 Water rate, lbmol/h (approximate) 20 20