You are on page 1of 18

Chapter 25

Differential Equations
of Mass Transfer
I n Chapter 9, the general differential equations for momentum transfer are derived by
the use of a differential control volume concept. By an analogous treatment, the general
differential equations for heat transfer are generated in Chapter 16. Once again, we
shall use this approach to develop the differential equations for mass transfer. By
making a mass balance over a differential control volume, we shall establish the
equation of continuity for a given species.
Additional differential equations will be obtained when we insert, into the
continuity equation, mass flux relationships developed in the previous chapter.
25.1 THE DIFFERENTIAL EQUATION FOR MASS TRANSFER
Consider the control volume, Dx Dy Dz, through which a mixture including component A is
flowing, as shown in Figure 25.1. The control-volume expression for the conservation of
mass is
ZZ ZZZ y
@
rðv : nÞdA þ r dV ¼ 0 (4-1)
c:s: @t c:v:
∆y
which may be stated in words as ∆z
8 9 8 9 ∆x
<net rate of mass= <net rate of accumulation=
efflux from þ of mass within control ¼ 0 x
: ; : ;
control volume volume
z
If we consider the conservation of a given species A,
this relation should also include a term that accounts for Figure 25.1 A differential control
volume.
the production or disappearance of A by chemical
reaction within the volume. The general relation for
a mass balance of species A for our control volume may
be stated as
8 9
8 9 8 9 > rate of chemical >
< net rate of mass = < net rate of accum = > < >
=
production of A
efflux of A from þ ulation of A within % ¼ 0 (25-1)
: ; : ; > > within the control >
>
control volume control volume : ;
volume
433
434 Chapter 25 Differential Equations of Mass Transfer
The individual terms will be evaluated for constituent A, and a discussion of their
meanings will be given below.
The net rate of mass efflux from the control volume may be evaluated by considering the
mass transferred across control surfaces. For example, the mass of A transferred across the
area Dy Dz at x will be rA v Ax Dy Dzjx , or in terms of the flux vector, nA ¼ rA vA , it would be
nA; x Dy Dzjx . The net rate of mass efflux of constituent A will be
in the x direction: nA,x Dy DzjxþDx % nA,x Dy Dzjx
in the y direction: nA,y Dx DzjyþDy % nA,y Dx Dzjy
and
in the z direction: nA,z Dx DyjzþDz % nA,z Dx Dyjz
The rate of accumulation of A in the control volume is
@rA
Dx Dy Dz
@t
If A is produced within the control volume by a chemical reaction at a rate rA, where rA has
the units (mass of A produced)/(volume)(time), the rate of production of A is
rA Dx Dy Dz
This production term is analogous to the energy generation term that appeared in the
differential equation for energy transfer, as discussed in Chapter 16.
Substituting each term in equation (25-1), we obtain
nA,x Dy DzjxþDx % nA,x Dy Dzjx þ nA,y Dx DzjyþDy
% nA,y Dx Dzjy þ nA,z Dx DyjzþDz % nA,z Dx Dyjz (25-2)
@rA
þ Dx Dy Dz % rA Dx Dy Dz ¼ 0
@t
Dividing through by the volume, Dx Dy Dz, and canceling terms, we have
nA,x jxþDx
%n j
A, x n A ,y j yþDy % n A, y jy n j
A , z %n j
A , z @r
x
þ þ zþDz z
þ A % rA ¼ 0 (25-3)
Dx Dy Dz @t
Evaluated in the limit as Dx, Dy, and Dz approach zero, this yields
@ @ @ @r
nA,x þ nA,y þ nA,z þ A %rA ¼ 0 (25-4)
@x @y @z @t
Equation (25-4) is the equation of continuity for component A. As nA,x ; nA,y ; and nA,z
are the rectangular components of the mass flux vector, nA, equation (25-4) may be
written
@rA
=: nA þ %rA ¼ 0 (25-5)
@t
A similar equation of continuity may be developed for a second constituent B in the
same manner. The differential equations are
@ @ @ @r
nB; x þ nB; y þ nB; z þ B %rB ¼ 0 (25-6)
@x @y @z @t
25.1 The Differential Equation for Mass Transfer 435
and
@rB
=: nB þ %rB ¼ 0 (25-7)
@t
where rB is the rate at which B will be produced within the control volume by a chemical
reaction. Adding equations (25-5) and (25-7), we obtain
@(rA þ rB )
=: (nA þ nB ) þ %ðrA þ rB Þ ¼ 0 (25-8)
@t
For a binary mixture of A and B, we have
nA þ nB ¼ rA vA þ rB vB ¼ rv
rA þ rB ¼ r
and
r A ¼ %r B
by the law of conservation of mass. Substituting these relations into (25-8), we obtain
@r
=: rv þ ¼0 (25-9)
@t
This is the equation of continuity for the mixture. Equation (25-9) is identical to the
equation of continuity (9-2) for a homogeneous fluid.
The equation of continuity for the mixture and for a given species can be written in
terms of the substantial derivative. As shown in Chapter 9, the continuity equation for the
mixture can be rearranged and written
Dr
þr=: v ¼ 0 (9-5)
Dt
Through similar mathematical manipulations, the equation of continuity for species A in
terms of the substantial derivative may be derived. This equation is
rDvA
þ=: jA % rA ¼ 0 (25-10)
Dt
We could follow the same development in terms of molar units. If RA represents the rate
of molar production of A per unit volume, and RB represents the rate of molar production of B
per unit volume, the molar-equivalent equations are
for component A
@cA
=: NA þ % RA ¼ 0 (25-11)
@t
for component B
@cB
=: NB þ % RB ¼ 0 (25-12)
@t
and for the mixture
@ðcA þ cB Þ
=: ðNA þ NB Þ þ % ðRA þ RB Þ ¼ 0 (25-13)
@t
436 Chapter 25 Differential Equations of Mass Transfer
For the binary mixture of A and B, we have
NA þ NB ¼ cA v A þ cB v B ¼ cV
and
cA þ cB ¼ c
However, only when the stoichiometry of the reaction is
A fi B
which stipulates that one molecule of B is produced for each mole of A disappearing, can we
stipulate that RA ¼ %RB . In general, the equation of continuity for the mixture in molar
units is
@c
=: cV þ % ðRA þ RB Þ ¼ 0 (25-14)
@t
25.2 SPECIAL FORMS OF THE DIFFERENTIAL MASS-TRANSFER EQUATION
Special forms of the equation of continuity applicable to commonly encountered situations
follow. In order to use the equations for evaluating the concentration profiles, we replace the
fluxes, nA and NA , by the appropriate expressions developed in Chapter 24. These expres-
sions are
NA ¼ %cDAB =yA þ yA (NA þ NB ) (24-21)
or its equivalent
NA ¼ %cDAB =yA þ cA V
and
nA ¼ %rDAB =vA þ vA (nA þ nB ) (24-22)
or its equivalent
nA ¼ %rDAB =vA þ rA v
Substituting equation (24-22) into equation (25-5), we obtain
@rA
%=: rDAB =vA þ =: rA v þ %rA ¼ 0 (25-15)
@t
and substituting equation (24-21) into equation (25-11), we obtain
@cA
%=: cDAB =yA þ =: cA V þ %RA ¼ 0 (25-16)
@t
Either equation (25-15) or (25-16) may be used to describe concentration profiles within
a diffusing system. Both equations are completely general; however, they are relatively
unwieldy. These equations can be simplified by making restrictive assumptions.
Important forms of the equation of continuity, with their qualifying assumptions, include:
(i) If the density, r, and the diffusion coefficient, DAB can be assumed constant,
equation (25-15) becomes
0
@rA
%DAB =2 rA þ rA =: v þ v: =rA þ %rA ¼ 0
! @t
25.2 Special Forms of the Differential Mass-Transfer Equation 437
Dividing each term by the molecular weight of A and rearranging, we obtain
@cA
v: =cA þ ¼ DAB =2 cA þ RA (25-17)
@t
(ii) If there is no production term, RA ¼ 0, and if the density and diffusion coefficient
are assumed constant, equation (25-17) reduces to
@cA
þv: =cA ¼ DAB =2 cA (25-18)
@t
We recognize that ð@cA =@tÞ þ v: =cA is the substantial derivative of cA; rewriting the
left-hand side of equation (25-18), we obtain
DcA
¼ DAB =2 cA (25-19)
Dt
which is analogous to equation (16-14) from heat transfer
DT k 2
¼ = T (16-14)
Dt rcP
or
DT
¼ a=2 T
Dt
where a is the thermal diffusivity. The similarity between these two equations is the basis for
the analogies drawn between heat and mass transfer.
(iii) In a situation in which there is no fluid motion, v ¼ 0, no production term, RA ¼ 0,
and no variation in the diffusivity or density, equation (25-18) reduces to
@cA
¼ DAB =2 cA (25-20)
@t
Equation (25-20) is commonly referred to as Fick’s second ‘‘law’’ of diffusion. The
assumption of no fluid motion restricts its applicability to diffusion in solids, or stationary
liquids, and for binary systems of gases or liquids, where NA is equal in magnitude, but
acting in the opposite direction to N; that is, the case of equimolar counterdiffusion.
Equation (25-20) is analogous to Fourier’s second ‘‘law’’ of heat conduction
@T
¼ a=2 T (16-18)
@t
(iv) Equations (25-17), (25-18), and (25-20) may be simplified further when the
process to be defined is a steady-state process; that is, @cA =@t ¼ 0. For constant density
and a constant-diffusion coefficient, the equation becomes
v: =cA ¼ DAB =2 cA þ RA (25-21)
For constant density, constant diffusivity, and no chemical production, RA ¼ 0, we
obtain
v: =cA ¼ DAB r2 cA (25-22)
If additionally, v ¼ 0 the equation reduces to
=2 cA ¼ 0 (25-23)
Equation (25-23) is the Laplace equation in terms of molar concentration.
438 Chapter 25 Differential Equations of Mass Transfer
Each of the equations (25-15) through (25-23) has been written in vector form, thus
each applies to any orthogonal coordinate system. By writing the Laplacian operator, =2, in
the appropriate form, the transformation of the equation to the desired coordinate system is
accomplished. Fick’s second ‘‘law’’ of diffusion written in rectangular coordinates is
! "
@cA @ 2 cA @ 2 cA @ 2 cA
¼ DAB þ 2 þ 2 (25-24)
@t @x2 @y @z
in cylindrical coordinates is
! 2 "
@cA @ cA 1 @cA 1 @ 2 cA @ 2 cA
¼ DAB þ þ þ 2 (25-25)
@t @r 2 r @r r 2 @u2 @z
and in spherical coordinates is
! # $ # $ "
@cA 1 @ @cA 1 @ @cA 1 @ 2 cA
¼ DAB 2 r2 þ 2 sin u þ 2 2 (25-26)
@t r @r @r r sin u @u @u r sin u @f2
The general differential equation for mass transfer of component A, or the equation of
continuity of A, written in rectangular coordinates is
! "
@cA @NA;x @NA;y @NA;z
þ þ þ ¼ RA (25-27)
@t @x @y @z
in cylindrical coordinates is
! "
@cA 1@ 1 @NA;u @NA;z
þ (rNA;r ) þ þ ¼ RA (25-28)
@t r @r r @u @z
and in spherical coordinates is
! "
@c A 1 @ 1 u 1 @NA;f
þ 2 (r 2 NA;r ) þ (NA;u sin u) þ ¼ RA (25-29)
@t r @r r sin u @u r sin u @f
25.3 COMMONLY ENCOUNTERED BOUNDARY CONDITIONS
A mass-transfer process is fully described by the differential equations of mass transfer only
if the initial boundary and initial conditions are specified. Typically, initial and boundary
conditions are used to specify limits of integration or to determine integration constants
associated with the mathematical solution of the differential equations for mass transfer.
The initial and boundary conditions used for mass transfer are very similar to those used in
Section 16.3 for energy transfer. The reader may wish to refer to that section for further
discussion of initial and boundary conditions.
The initial condition in mass transfer processes is the concentration of the diffusing
species at the start of the time interval of interest expressed in either mass or molar
concentration units. The concentration may be simply equal to a constant, for example
at t ¼ 0; cA ¼ cAo ðmolar unitsÞ
at t ¼ 0; rA ¼ rAo ðmass unitsÞ
or may be more complex if the initial concentration distribution within the control volume
for diffusion is specified. Initial conditions are associated only with unsteady-state or
pseudo-steady-state processes.
25.3 Commonly Encountered Boundary Conditions 439
Four types of boundary conditions are commonly encountered in mass transfer.
(1) The concentration of the transferring species A at a boundary surface is specified.
Surface concentration can assume a variety of units, for example, molar concentration cAs,
mass concentration rAs , gas mole fraction yAs, liquid mole fraction xAs, etc. When the
boundary surface is defined by a pure component in one phase and a mixture in the second
phase, then the concentration of transferring species A in the mixture at the interface is
usually at thermodynamic saturation conditions. Specifically, for a gas mixture in contact
with a pure volatile liquid A or pure volatile solid A, the partial pressure of species A in the
gas at the surface is saturation vapor pressure, PA, so that pAs ¼ PA . For a liquid mixture in
contact with a pure solid A, the concentration of species A in the liquid at the surface is the
solubility limit of A in the liquid, cA& so that cAs ¼ cA& .
For a contacting gas and liquid where transferring species A is present in both phases,
there are two ways to specify the concentration at the gas–liquid interface. First, if both of
the species in the liquid phase are volatile, then the boundary condition at the gas–liquid
surface is defined for an ideal liquid mixture by Raoult’s law
pAs ¼ xA PA
where xA is the mole fraction in the liquid, PA is the vapor pressure of species A evaluated at
the temperature of the liquid, and PAs is the partial pressure of species A in the gas. The
partial pressure of species A at the interface is related to surface mole fraction yAs by
Dalton’s law
pAs
yAs ¼
P
or to surface concentration cAs by the Ideal Gas law
pAs
cAs ¼
RT
Second, for solutions where species A is only weakly soluble in the liquid, Henry’s law
may be used to relate the mole fraction of A in the liquid to the partial pressure of A in
the gas
p A ¼ H ' xA
where coefficient H is known as Henry’s constant. Values of H in pressure units for
selected gaseous solutes dissolved in aqueous solution are listed in Table 25.1. A similar
equation may also be used to determine the boundary conditions at a gas–solid
interface
cA;solid ¼ S ' pA
Table 25.1 Henry’s constant for various gases in aqueous solutions (H in bars)
T (K) NH3 Cl2 H2S SO2 CO2 CH4 O2 H2
273 21 265 260 165 710 22,800 25,500 58,000
280 23 365 335 210 960 27,800 30,500 61,500
290 26 480 450 315 1300 35,200 37,600 66,500
300 30 615 570 440 1730 42,800 45,700 71,600
310 755 700 600 2175 50,000 52,500 76,000
320 860 835 800 2650 56,300 56,800 78,600
440 Chapter 25 Differential Equations of Mass Transfer
Table 25.2 Solubility constants for selected gas–solid combinations (1 bar ¼ 105 Pa)
S ¼ cA;solid /PA
Gas Solid T (K) (kg mol/m3 bar)
O2 Natural rubber 298 3:12 ( 10%3
N2 Natural rubber 298 1:56 ( 10%3
CO2 Natural rubber 298 40:15 ( 10%3
He Silicon 293 0:45 ( 10%3
H2 Ni 358 9:01 ( 10%3
where cA; solid is the molar concentration of A within the solid at the interface in units of
kg mol/m3 and pA is the partial pressure of gas phase species A over the solid in units of Pa.
The partition coefficient S, also known as the solubility constant, has units of kg mol/m3 : Pa.
Values of S for several gas–solid pairs reported by Barrer1 are listed in Table 25.2.
(2) A reacting surface boundary is specified. There are three common situations, all
dealing with heterogeneous surface reactions. First, the flux of one species may be related
to the flux of another species by chemical reaction stoichiometry. For example, consider the
generic chemical reaction at the boundary surface A þ 2B ! 3C, where reactants A and B
diffuse to the surface, and product C diffuses away from the surface. The fluxes for A and B
move in the opposite direction to the flux for C. Consequently, the flux NA is related to the
flux of the other species by NB ¼ þ2 NA or NC ¼ %3 NA. Second, a finite rate of chemical
reaction might exist at the surface, which in turn sets the flux at the surface. For example, if
component A is consumed by a first-order on a surface at z = 0, and the positive z direction
is opposite to the direction of flux of A along z, then
%
%
¼ %kc cAs
NA %z¼0
where ks is a surface reaction rate constant with units of m/s. Third, the reaction may be so
rapid that cAs ¼ 0 if species A is the limiting reagent in the chemical reaction.
(3) The flux of the transferring species is zero at a boundary or at a centerline of
symmetry. This situation can arise at an impermeable boundary, or at the centerline of
symmetry of the control volume, where the net flux is equal to zero. In either case, for a
one-dimensional flux along z
% @c % @cA %%
% A %
¼ %DAB ¼0 ¼0
NA %z¼0 % %
or
@z z¼0 @z z¼0
where the impermeable boundary or the centerline of symmetry is located at z = 0.
(4) The convective mass transfer flux at the boundary surface is specified. When a
fluid flows over the boundary, the flux can be defined by convection. For example, at some
surface located at z = 0, the convective mass transfer flux across the fluid boundary layer is
%
%
¼ kc ðcAs % cA1 Þ
NA %z¼0
where cA1 is the bulk concentration of A the flowing fluid, cAs is the surface concentration of
A at z = 0 and kc is the convection mass-transfer coefficient defined in Section 24.3.
1
R. M. Barrer, Diffusion In and Through Solids, Macmillan Press, New York, 1941.
25.4 Steps for Modeling Processes Involving Molecular Diffusion 441
25.4 STEPS FOR MODELING PROCESSES INVOLVING
MOLECULAR DIFFUSION
Processes involving molecular diffusion can be modeled by the appropriate simplifications
to Fick’s equation and the general differential equation for mass transfer. In general, most
molecular diffusion problems involve working through the following five steps:
Step 1: Draw a picture of the physical system. Label the important features, including
the system boundaries. Decide where the source and the sink of mass transfer are
located.
Step 2: Make a ‘‘list of assumptions’’ based on your consideration of the physical
system. As appropriate, make a ‘‘list of nomenclature’’ and update the list as you
add more terms to the model development.
Step 3: Pick the coordinate system that best describes the geometry of the physical
system: rectilinear (x, y, z), cylindrical (r, z, u), or spherical (r, u, f). Then
formulate differential material balances to describe the mass transfer within a
volume element of the process based on the geometry of the physical system and
the assumptions proposed, making use of Fick’s law and the general differential
equation for mass transfer. Two approaches may be used to simplify the general
differential equation for mass transfer. In the first approach, simply reduce or
eliminate the terms that do not apply to the physical system. For example:
@cA
(a) If the process is steady state, then ¼ 0.
@t
(b) If no chemical reaction occurs uniformly within the control volume for
diffusion, then RA ¼ 0.
(c) If the molecular mass transfer process of species A is one-dimensional in
the z direction,
@NAz
= ' NA ¼
@z
by cylindrical geometry in the r and z directions,
@NAz 1 @(rNAr )
= ' NA ¼ þ
@z r @r
for radial symmetry in spherical coordinates,
1 @(r2 NAr )
= ' NA ¼
r2 @r
In the second approach, perform a ‘‘shell balance’’ for the component of interest
on a differential volume element of the process. Both of these approaches are
discussed and illustrated in Chapter 26. Next, Fick’s law is simplified by
establishing the relationship between the fluxes in the bulk-contribution
term. For example, recall the one-dimensional flux of a binary mixture of
components A and B
dyA
NAz ¼ %cDAB þ yA (NAz þ NBz )
dz
If NAz ¼ %NBz , then yA (NAz þ NBz ) ¼ 0. If yA (NAz þ NBz ) does not equal 0, then
NA is always equal to cA Vz and reduces to cA v z for low concentrations of A in the
442 Chapter 25 Differential Equations of Mass Transfer
Differential
equation for
Fick's equation mass transfer
Assumptions Assumptions
Fick's equation Simplified differential
equation for
differential from (NA)
mass transfer (NA)
Analytical Boundary More
integration conditions assumptions
Integral form Simplified differential
flux (NA) or equation for
transfer rate (wA) mass transfer (cA)
Analytical Boundary
integration conditions
Figure 25.2 Model
Integral form Differentiation Concentration development pathways for
flux (NA) or
profile (cA) processes involving
transfer rate (wA)
molecular diffusion.
mixture. If a differential equation for the concentration profile is desired, then the
simplified form of Fick’s law must be substituted into the simplified form of the
general differential equation for mass transfer. Figure 25.2 illustrates this
process.
Step 4: Recognize and specify the boundary conditions and initial conditions. For
example
(a) Known concentration of species A at a surface or interface at z ¼ 0 e.g.,
cA ¼ cAo . This concentration can be specified or known by equilibrium
relationships such as Henry’s law.
(b) Symmetry condition at a centerline of the control volume for diffusion,
or no net diffusive flux of species A at a surface or interface at z ¼ 0;
NAz jz¼0 ¼ 0 ¼ dcA =dz.
(c) Convective flux of species A at a surface or interface, e.g., NA ¼
kc (cAs % cA1 ).
(d) Known flux of species A at a surface or interface, e.g., at z ¼ 0,
NAz jz¼0 ¼ NAo .
(e) Known chemical reaction at a surface or interface. For the rapid disappearance
of species A at the surface or interface, e.g., at z ¼ 0; cAs ¼ 0. For a slower
chemical reaction at the surface or interface with finite cAs at z ¼ 0, e.g.,
NAz ¼ k0 cAs , where k0 is a first-order chemical reaction rate constant.
Step 5: Solve the differential equations resulting from the differential material balances
and the boundary/initial conditions described to get the concentration profile,
the flux, or other parameters of engineering interest. If appropriate, consider
asymptotic solutions or limiting cases to more difficult problems first.
The following examples illustrate how physical and chemical processes involving
molecular diffusion can be modeled by the appropriate simplifications of Fick’s equation
and the general differential equation for mass transfer. The examples cover many of
25.4 Steps for Modeling Processes Involving Molecular Diffusion 443
typically encountered boundary conditions in both rectilinear and cylindrical geometry. The
examples emphasize the first four steps of model development outlined, and the final model
equations are generally left in differential-equation form. Chapters 26 and 27 provide
analytical solution techniques for steady-state and unsteady-state diffusion processes.
We have taken extra time at the beginning of each example to describe the interesting
technology behind the process.
EXAMPLE 1 Microelectronic devices are fabricated by forming many layers of thin films onto a silicon wafer.
Each film has unique chemical and electrical properties. For example, a thin film of solid silicon
(Si) serves as a semiconductor. Silicon thin films are commonly formed by the chemical
vapor deposition, or CVD, of silane vapor (SiH4) onto the surface of the wafer. The chemical
reaction is
SiH4 ðgÞ ! SiðsÞ þ 2H2 ðgÞ
This surface reaction is usually carried out at very low pressure (100 Pa) and high temperature
(900 K). In many CVD reactors, the gas phase over the Si film is not mixed. Furthermore, at high
temperatures, the surface reaction is very rapid. Consequently, the molecular diffusion of the
SiH4 vapor to the surface often controls the rate of Si film formation. Consider the very simplified
CVD reactor shown in Figure 25.3. A mixture of silane and hydrogen gas flows into the reactor. A
diffuser provides a quiescent gas space over the growing Si film. Develop a differential model for
this process, including statements of assumptions and boundary conditions.
SiH4 vapor
To vacuum
+ H2 gas
Diffuser
z=0
SiH4 H2
Quiescent
gas Si thin
film z=δ
Figure 25.3 Chemical
Heated plate vapor deposition of silicon
SiH4(g) Si(s) + 2 H2 (g) hydride.
The silane in the feed gas serves as the source for mass transfer, whereas the Si film serves as
the sink for silane mass transfer. In contrast, the formation of H2 at the Si film surface serves as the
source for H2 mass transfer, whereas the feed gas serves as the sink for H2 mass transfer. The physical
system possesses rectilinear geometry, and the major assumptions for model development are listed
here.
(1) The reaction occurs only at the surface of growing Si thin film. Consequently, there is no
homogeneous reaction of silane within the diffusion zone, so that RA ¼ 0. In this context, the
surface reaction is the sink for silane mass transfer. (2) The gas space in the ‘‘diffusion zone’’ is
not externally mixed, so that molecular diffusion dominates. (3) The feed gas provides silane in
high excess relative to that consumed by reaction, so the silane concentration in the gas space at
the diffusion-zone boundary is constant. (4) The flux of silane is one-dimensional along z, as
the source and sink for silane mass transfer are aligned at the boundaries along the z direction.
(5) The thickness of the Si film is very thin relative to d, the diffusion path length along the
z direction. Therefore, d is essentially constant. (6) The mass transfer process within the diffusion
zone is at steady state.
The assumptions are used to reduce the general forms of the differential equation for mass
transfer and Fick’s equation. The general differential equation for mass transfer in terms of rectilinear
coordinates is
# $
@NAx @NAy @NAz @cA
% þ þ þ RA ¼
@x @y @z @t
444 Chapter 25 Differential Equations of Mass Transfer
For steady-state one-dimensional flux along the z direction with no homogeneous chemical reaction
ðRA ¼ 0Þ, the general differential equation for mass transfer reduces to
dNAz
¼0
dz
which shows that the flux is constant along the z direction. As the diffusion flux is with respect to only
one dimension, the partial derivative becomes an ordinary derivative. Fick’s equation for the one-
dimensional flux silane through a binary mixture in the gas phase is
dyAz
NAz ¼ %cDAB þ yA ðNAz þ NBz Þ
dz
where species A represents silane vapor (SiH4) reactant and species B represents the hydro-
gen gas (H2) product. The flux of the gaseous reactant is opposite in direction to the flux of the
gaseous product. From the reaction stoichiometry and Figure 25.3, NAz is related to NBz as
follows:
NAz %1 mol SiH4 reacted 1
¼ ¼%
NBz þ2 mol H2 formed 2
Therefore, NBz ¼ %2 NAz and Fick’s equation further reduces to
dyA cDAB dyA
NAz ¼ %cDAB þ yA ðNAz % 2NAz Þ ¼ %
dz 1 þ yA dz
It is interesting to note that increasing yA decreases the flux. Two boundary conditions must be
specified. At the surface of the Si film, the reaction is so rapid that the concentration of silane vapor is
zero. Furthermore, the concentration of silane in the feed gas is constant.
At the Si film surface, z ¼ d; yA ¼ yAs , and yBs ¼ 1 (yA þ yB ¼ 1 for binary mixture).
At the diffusion screen, z ¼ 0; yA ¼ yAo ; and yB ¼ yBo ¼ 1 % yAo .
The differential model is now specified. Although the analytical solution was not asked for in
the problem statement, it is easy to obtain. We first recognize that for this particular system, NAz is a
constant along z. If NAz is a constant, then Fick’s equation can be integrated by separation of
dependent variable yA from independent variable z, with integration limits defined by the boundary
conditions
Zd ZyAs
%cDAB dyA
NAz dz ¼
1 þ yA dz
0 yAo
If the system temperature T and total system pressure P are constant, then the total molar
concentration of the gas, c ¼ P=RT, is also constant. Likewise, the binary gas phase diffusion
coefficient of silane vapor in hydrogen gas, DAB, is also constant. The final integrated equation is
# $
cD AB 1 þ yAo
NAz ¼ ln
d 1 þ yAs
If yAs is specified, then NAz can be determined. With the silane flux NAz known, parameters of
engineering interest, such as the Si film formation rate, can be easily determined. These questions are
considered in a problem exercise at the end of Chapter 26.
EXAMPLE 2 The formation of a tungsten thin film on unmasked surfaces of a silicon wafer is an important step in
the fabrication of solid-state microelectronic devices. The tungsten metal serves as conductor for
current flow between devices on the wafer. In one typical process, the tungsten thin film is formed by
25.4 Steps for Modeling Processes Involving Molecular Diffusion 445
the chemical vapor deposition of tungsten hexafluoride (WF6) onto the surface of the wafer in the
presence of hydrogen (H2) gas and an inert helium carrier gas
3H2 ðgÞ þ WF6 ðgÞ ! WðsÞ þ 6HFðgÞ
as shown in Figure 25.4. What is the differential form of Fick’s law for the flux of WF6 gas onto the
surface?
Tungsten thin film formation
WF6(g) H2(g) HF(g)
z=0
Tungsten
z=δ
thin film Figure 25.4 Chemical vapor
Silicon substrate deposition of tungsten
hexafluoride.
The assumptions for analysis are similar to those made in example 1. The flux of WF6
(species A) is one dimensional in the z direction. As there are four components in the gas phase
mixture, Fick’s equation is
dyA
NAz ¼ %cDA%mixture þ yA (NAz þ NBz þ NCz þ NDz )
dz
where DA%mixture is the diffusion coefficient of WF6 in the mixture of H2 gas (species B), HF gas
(species C), and inert He gas (species D). The flux of the gaseous reactants (WF6, H2) is opposite in
direction to the flux of the gaseous product (HF). The reaction stoichiometry at the surface of the
tungsten film relate the fluxes of all diffusing species to WF6
NAz %1 mol WF6 reacted 1
¼ ¼ or NBz ¼ þ3NAz
NBz %3 mol H2 reacted 3
NAz %1 mol WF6 reacted 1
¼ ¼% or NCz ¼ %6NAz
NCz þ6 mol HF formed 6
The net flux of He, NDz , is zero because it has no sink for mass transfer. Accordingly, Fick’s equation
for WF6 reduces to
dyA
NAz ¼ %cDA%mixture þ yA ðNAz þ 3NAz % 6NAz þ 0Þ
dz
or
%cDA%mixture dyA
NAz ¼
1 þ 2yA dz
The flux of WF6 to the surface is hindered by the flux of the HF product gas from the surface since the
denominator term (1 þ 2yA ) is greater than one. The differential form of Fick’s equation cannot be
integrated analytically unless a mean value for the mixture-based diffusion coefficient is taken.
EXAMPLE 3 An emerging area of biotechnology called ‘‘tissue engineering’’ develops new processes to grow
organized living tissues of human or animal origin. A typical configuration is the engineered tissue
bundle. Engineered tissue bundles have several potential biomedical applications, including the
production of replacement body tissue (skin, bone marrow, etc.) for transplantation into the human
body, or in the future, may serve as artificial organs for direct implantation into the human body.
446 Chapter 25 Differential Equations of Mass Transfer
Living tissues require oxygen to stay alive. The mass transport of oxygen (O2) to the tissue is an
important design consideration. One potential system is schematically illustrated in Figure 25.5. Thin
tubes arranged on a triangular pitch pass longitudinally through the tissue bundle. The tubes serve as
a ‘‘scaffold’’ for supporting the living tissue matrix and supply oxygen and nutrients to the tissue at
the same time. Let us focus on a single O2 delivery tube with tissue surrounding it, as illustrated in
Figure 25.5. Pure oxygen (O2) gas flows through the tube. The tube wall is extremely permeable
to O2, and the O2 partial pressure through the porous tube wall can be taken as the O2 partial
100% O2 gas
Tube Tissue bundle
(100% O2 gas)
Nutrient
Tissue
Tissue
r = R1, cA = cAs
Cross section
Tube
NAr r = R2, dcA/dr = 0
Simplified cross section
Figure 25.5 Oxygen transport within an engineered tissue bundle.
pressure inside the tube. Oxygen is only sparingly soluble in the tissue, which is mostly water. The
concentration of dissolved O2 at r ¼ R1 , is
pA
cAs ¼
H
where H is the Henry’s law constant for the dissolution of O2 in living tissue at the process
temperature, and pA is the partial pressure of O2 in the tube. The dissolved O2 diffuses through the
tissue and is metabolically consumed. The metabolic consumption of dissolved O2 is described by a
kinetic rate equation of the form
RA; max cA
RA ¼ %
K A þ cA
A key parameter in the design of the engineered tissue bundle is the spacing between the tubes. If the
tube spacing is too wide, the dissolved O2 concentration will go to near zero and starve the tissue.
Therefore, it is important to know the radial concentration profile, cA ðrÞ of dissolved O2. Develop a
differential model to predict cA ðrÞ.
The physical system possesses cylindrical geometry, and the following assumptions for model
development are listed here. (1) The source for O2 mass transfer is the pure O2 gas inside the tube, and
the sink for mass transfer is the metabolic consumption of dissolved oxygen by the tissue. If the O2
partial pressure pA is maintained constant inside the tube along longitudinal coordinate z, then the
flux of oxygen through the tissue is one dimensional along the radial (r) direction. (2) Tissue remains
viable and maintains constant physical properties. (3) The O2 transfer process is at steady state. (4)
The tissue is stationary, and the dissolved O2 concentration is dilute. (5) At r ¼ R1 , the tube material
25.4 Steps for Modeling Processes Involving Molecular Diffusion 447
is thin and highly permeable to O2 so that the dissolved O2 concentration in the tissue is in
equilibrium with the O2 partial pressure in the tube. (6) At r ¼ R2 , there is no net flux of O2.
The general differential equation for mass transfer in cylindrical coordinates is
# $
1 @ 1 @N Au @N Az @cA
% ðrNAr Þ þ þ þ RA ¼
r @r r @u @z @t
For steady-state one-dimensional flux along the r direction, the general equation for mass transfer
reduces to
1 @
% ðrNAr Þ þ RA ¼ 0
r @r
For a one-dimensional system, the partial derivatives can be replaced with ordinary derivatives.
Alternatively, we can perform a material balance for dissolved O2 on the differential element of
volume 2pLr Dr shown in Figure 25.5 and get the same result. Specifically, for steady-state one-
dimensional flux along the r direction with a homogeneous reaction RA within the differential volume
element, we have
% %
2pLr NAr %r¼r %2pLr NAr %r¼rþDr þ RA : 2pLr : Dr ¼ 0
Diving through by 2pL Dr, and rearranging, we get
% % !
rNAr %r¼rþDr % rNAr %r¼r
% þ RA r ¼ 0
Dr
Finally, taking the limits as Dr ! 0 yields
1 d
% ðrNAr Þ þ RA ¼ 0
r dr
For one-dimensional flux of dissolved O2 through the stagnant tissue in cylindrical coordinates
along the r direction, Fick’s equation reduces to
dcA cA dcA
NAr ¼ %DAB þ ðNAr Þ ffi % DAB
dr c dr
because O2 is only sparingly soluble in the tissue so that cA * c, where c is the total molar
concentration of the tissue, which approximates the molar concentration of water. In cylindrical
geometry, NAr is not constant along diffusion path r, because (a) cross-sectional area for flux is
increasing along r and (b) the RA term is present. As a result, the flux equation cannot be integrated, as
was the case in example 1. It is now necessary to combine Fick’s equation and the differential
equation for mass transfer in order to get the concentration profile
# $
1 d dcA
% %rDAB þ RA ¼ 0
r dr dr
or
! "
d2 c A 1 dc A RA; max cA
DAB þ % ¼0
dr 2 r dr K A þ cA
The concentration profile cA ðrÞ is now expressed as a second-order differential equation. Therefore,
two boundary conditions on cA ðrÞ must be specified:
dcA
r ¼ R1 ; ¼ 0 ðnet flux NA ¼ 0 at r ¼ R1 Þ
dr
pA
r ¼ R2 ; cA ¼ cAs ¼
H
The analytical solution for cA ðrÞ and its extension to predicting the overall rate of oxygen
consumption in the tissue bundle has been left as a problem exercise in Chapter 26.
448 Chapter 25 Differential Equations of Mass Transfer
25.5 CLOSURE
The general differential equation for mass transfer was developed to describe the mass
balances associated with a diffusing component in a mixture. Special forms of the general
differential equation for mass transfer that apply to specific situations were presented.
Commonly encountered boundary conditions for molecular diffusion processes were also
listed. From this theoretical framework, a five-step method for mathematically modeling
processes involving molecular diffusion was proposed. Three examples illustrated how the
differential form of Fick’s equation presented in Chapter 24, and the general differential
equation for mass transfer presented in this chapter, are reduced to simple differential
equations that describe the molecular diffusion aspects of a specific process. The approaches
presented in this chapter serve as the basis for problem solving in Chapters 26 and 27.
PROBLEMS
25.1 Derive equation (25-11) for component A in terms of process, list at least five reasonable assumptions for the
molar units, starting with the control-volume expression for the mass-transfer aspects of the water-evaporation process and
conservation of mass. simplify the general differential equation for mass transfer
25.2 Show that the (25-11) may be written in the form in terms of the flux NA.
b. What is the simplified differential form of Fick’s equation
@rA
þð= ' rA vÞ % DAB r2 rA ¼ rA for water vapor (species A)?
@t
25.5 A large deep lake, which initially had a uniform oxygen
25.3 The following sketch illustrates the gas diffusion in the
concentration of 1kg/m3, has its surface concentration suddenly
neighborhood of a catalytic surface. Hot gases of heavy hydro-
raised and maintained at 9 kg/m3 concentration level.
carbons diffuse to the catalytic surface where they are cracked
Reduce the general differential equation for mass transfer to
into lighter compounds by the reaction: H ! 2L, the light
write the specific differential equation for
products diffuse back into the gas stream.
a. the transfer of oxygen into the lake without the presence of a
H L chemical reaction;
z=0 b. the transfer of oxygen into the lake that occurs with the
simultaneous disappearance of oxygen by a first-order
z=δ biological reaction.
25.6 The moisture in hot, humid, stagnant air surrounding a
cold-water pipeline continually diffuses to the cold surface
a. Reduce the general differential equation for mass transfer to where it condenses. The condensed water forms a liquid film
write the specific differential equation that will describe this around the pipe, and then continuously drops off the pipe to the
steady-state transfer process if the catalyst is considered a ground below. At a distance of 10 cm from the surface of the pipe,
flat surface. List all of the assumptions you have made in the moisture content of the air is constant. Close to the pipe, the
simplifying the general differential equation. moisture content approaches the vapor pressure of water eval-
b. Determine the Fick’s law relationship in terms of only uated at the temperature of the pipe.
compound H and insert it into the differential equation a. Draw a picture of the physical system, select the coordinate
you obtained in part (a). system that best describes the transfer process and state at
25.4 A hemispherical droplet of liquid water, lying on a flat least five reasonable assumptions of the mass-transfer
surface, evaporates by molecular diffusion through still air aspects of the water condensation process.
surrounding the droplet. The droplet initially has a radius R. b. What is the simplified form of the general differential
As the liquid water slowly evaporates, the droplet shrinks slowly equation for mass transfer in terms of the flux of water
with time, but the flux of the water vapor is at a nominal steady vapor, NA?
state. The temperature of the droplet and the surrounding still air c. What is the simplified differential form of Fick’s equation
are kept constant. The air contains water vapor at an infinitely for water vapor, NA?
long distance from the droplet’s surface. d. What is the simplified form of the general differential
a. After drawing a picture of the physical process, select a equation for mass transfer in terms of the concentration
coordinate system that will best describe this diffusion of water vapor, cA?
Problems 449
25.7 A liquid flows over a thin, flat sheet of a slightly soluble to explain the metabolic consumption of the oxygen to produce
solid. Over the region in which diffusion is occurring, the liquid carbon dioxide.
velocity may be assumed to be parallel to the plate and to be Use the general differential equation for mass transfer of oxygen
given by v ¼ ay, where y is the vertical distance from the plate to write the specific differential equation that will describe the
and a is a constant. Show that the equation governing the mass diffusion of oxygen in the human tissue. What would be the form of
transfer, with certain simplifying assumptions, is Flicks relationship written in terms of only the diffusing oxygen?
# $ 25.10 A fluidized coal reactor has been proposed for a new
@ 2 cA @ 2 cA @cA
DAB þ 2 ¼ ay power plant. If the coal can be assumed to be spherical, reduce
@x2 @y @x
the general differential equation for mass transfer to obtain a
List the simplifying assumptions, and propose reasonable specific differential equation for describing the steady-state
boundary conditions. diffusion of oxygen to the surface of the coal particle.
25.8 Consider one of the cylindrical channels that run through Determine the Fick’s law relationship for the flux of oxygen
an isomerization catalyst as shown below. A catalyst coats the from the surrounding air environment if
inner walls of each channel. This catalyst promotes the isomer- a. only carbon monoxide, CO, is produced at the surface of the
ization of n-butane ðn % C4 H10 Þ species A to isobutene ði % carbon particle;
C4 H10 Þ species B. b. only carbon dioxide, CO2, is produced at the surface of the
carbon particle.
n % C4 H10 ðgÞ ! i % C4 H10 ðgÞ If the reaction at the surface of the carbon particle is instan-
tenous, give two boundary conditions that might be used in
solving the differential equation.
25.11 In the manufacture of semiconducting thin films, a thin
film of solid arsenic laid down on the surface of a silicon water by
the diffusion-limited chemical vapor deposition of arsine, AsH3.
2AsH3 ðgÞ ! 2AsðsÞ þ 3H2 ðgÞ
The gas head space, 5 cm above the surface of the wafer, is
stagnant. Arsenic atoms deposited on the surface then diffuse
into the solid silicon to ‘‘dope’’ the wafer and impart semicon-
The gas phase above the channels contains mixture of A and B ducting properties to the silicon, as shown in the figure below.-
maintained at a constant composition of 60 mol % n % C4 H10 Well mixed feed gas (constant composition).
(A) and 40 mol % i % C4 H10 (B). Gas phase species A diffuses Well-mixed feed gas (constant composition).
down a straight channel of diameter d ¼ 0:1 cm and length
L ¼ 2:0 cm. The base of each channel is sealed. This is rapid Diffuser screen
reaction so that the production rate of B is diffusion limited.
The quiescent gas space in the channel consists of only species A AsH3 (g) H2 (g)
and B.
a. State three relevant assumptions for the mass transfer As, thin film
process. Si wafer
b. Based on your assumptions, simplify the general differential NA
equation for the mass transfer of species A, leaving the
equation in terms of the flux NA. The process temperature is 1050+ C. The diffusion coefficient of
c. Using equations for the flux of A in your determined aresenic in silicon is 5 ( 10%13 cm/s at this temperature and the
equation, express the general differential equation in terms maximum solubility of aresenic in silicon is 2 ( 1021
of the concentration cA. atoms/cm3 . The density of solid silicon is 5 ( 1022 atoms=
d. Specify relevant boundary conditions for the gas phase cm3 . As the diffusion coefficient is so small, the aresenic atoms
concentration cA. do not ‘‘penetrate’’ very far into the silicon solid, usually less
than a few microns. Consequently, a relatively thin silicon water
25.9 An early mass-transfer study of oxygen transport in
can be considered as a ‘‘semi-infinite’’ medium for diffusion.
human tissue won a Nobel prize for August Krough. By con-
sidering a tissue cylinder surrounding each blood vessel, he a. State at least five reasonable assumptions for the mass
proposed the diffusion of oxygen away from the blood vessel into transfer of aresenic in this doping process.
the annular tissue was accompanied by a zero-order reaction, that b. What is the simplified form of the general differential
is, RA ¼ %m, where m is a constant. This reaction was necessary equation for the mass transfer of the aresenic concentration
450 Chapter 25 Differential Equations of Mass Transfer
within the silicon? Purpose reasonable boundary and initial without disintegrating. A water-soluble drug (solute A) is uni-
conditions. formly dissolved within the gel, has an initial concentration,
25.12 A device has been proposed that will serve as a ‘‘blood cAo of 50 mg/cm3 . The drug loaded within the spherical gel
oxygenator’’ for a heart–lung bypass machine. In this process, capsule is the sink for mass transfer. Consider a limiting case
blood (which is mostly water, species B) containing no dissolved where the drug is immediately consumed or swept away once it
oxygen (O2 species A) enters the top of the chamber and then reaches the surface, i.e., @ R, cA ¼ 0.
falls vertically down as a liquid film of uniform thickness, along a a. In analyzing the process, choose a coordinate system and
surface designed to appropriately wet blood. Contacting the simplify the general differential equation for the mass
liquid surface is a 100% O2 gas phase. transfer of the drug in terms of the flux.
b. What reasonable assumptions were used in your simplifying
Inlet blood containing of the general differential equation.
no dissolved oxygen, 40°C c. Simplify Fick’s equation for the drug species and obtain a
z=0
differential equation in terms of concentration, cA.
Falling
liquid 25.14 Consider a single, porous, spherical, inert mineral par-
film ticle. The pores inside the particle are filled with liquid water
Inert (species B). We are interested in analyzing the molecular diffu-
solid surface 100% O2 gas sion of the contaminant benzene C6H6 species A within the
Vm water-filled pores of the particle. The average diameter of
the pores is 150 nm and the void fraction (porosity) is 0.40.
The benzene solute does not adsorb onto the inter surfaces of the
z=L pores. Benzene is very sparingly soluble in water, and has a
molecular diameter of 0.15 nm. The process is isothermal at 298
x=d x=0 K. The concentration of dissolved benzene in the water sur-
exiting oxygenated rounding the particle, cAo , is constant with time. The critical
blood, 40 °C volume (Vc) of benzene is 259 cm3 =gmol. The effective diffu-
Oxygen is soluble in blood, with the equilibrium solubility sion coefficient of benzene inside the porous particle was to be
described by Henry’s law, cA& ¼ pA =H, where pA is the partial calculated in Problem 24.25.
pressure of oxygen (atm), H is the Henry’s law constant, and cA & is Starting with the general differential equation for mass trans-
the solubility concentration limit of oxygen dissolved in blood fer of benzene, develop a differential model to describe the
(mmol/L) at pA. concentration profile of benzene, species A within the single,
In analyzing the mass transport of dissolved oxygen into the porous, spherical, inert mineral particle. State reasonable
falling film, you may assume the following: (1) the process is assumptions and boundary/initial conditions for the process.
dilute with respect to dissolved oxygen in the fluid; (2) the falling 25.15 A large tank truck overturns and spills a herbicide over a
liquid film has a flat velocity profile with velocity v max; (3) the field. The fluid remains on the soil 30 min before evaporating into
gas space always contains 100% oxygen; (5) the width of the the atmosphere. Simplify the general differential equation for the
liquid film, W, is much larger than the length of the liquid film, L. mass transfer of the herbicide to write the following:
a. Simplify the general differential equation for O2 transfer. If a. the steady-state differential equation that will describe the
your analysis suggests more than one dimension for flux, evaporation of the herbicide into the air;
provide a simplified flux equation for each coordinate of b. the differential equation that will describe the diffusion of
interest. the herbicide into the soil.
b. Provide one simplified differential equation in terms of the
25.16 Consider the drug treatment system shown below.
fluxes and another simplified differential equation in terms
A hemisphere cluster of unhealthy cells is surrounded by a
of the oxygen concentration cA.
larger hemisphere of stagnant dead dead tissue (species B),
c. List boundary conditions associated with the oxygen mass which is in turn surrounded by a flowing fluid. The bulk, well-
transfer process. mixed fluid contains a drug compound (species A) of constant
25.13 One way to deliver a timed dosage within the human but dilute concentration cAo . Drug A is also soluble in the
body is to ingest a capsule and allow it to settle in the gastro- unhealthy tissue but does not preferentially partition into it
intestinal system. Once inside the body, the capsule slowly relative to the fluid. The drug (species A) enters the dead tissue
releases the drug to the body by a diffusion-limited process. and homes in on the unhealthy cells. At the unhealthy cell
A suitable drug carrier is a spherical bead of a nontoxic gela- boundary ðr ¼ R1 Þ the flux of A to the unhealthy cells is
tinous material that can pass through the gastrointestinal system diffusion limited. All metabolites of drug A produced by

You might also like