Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

26.

4 Simultaneous Momentum, Heat, and Mass Transfer 479

26.4 SIMULTANEOUS MOMENTUM, HEAT, AND MASS TRANSFER


In previous sections, we have considered steady-state mass transfer independent of the other
transport phenomena. Many physical situations involve the simultaneous transfer of mass
and either energy or momentum, and in a few cases, the simultaneous transfer of mass,
energy, and momentum. The drying of a wet surface by a hot, dry gas is an excellent example
in which all three transport phenomena are involved. Energy is transferred to the cooler
surface by convection and radiation; mass and its associated enthalpy are transferred back
into the moving gas stream. The simultaneous transport processes are more complex,
requiring the simultaneous treatment of each transport phenomenon involved.
In this section, we consider two examples involving the simultaneous transfer of mass
and a second transport phenomenon.

Simultaneous Heat and Mass Transfer


Generally, a diffusion process is accompanied by the transport of energy, even within an
isothermal system. As each diffusing species carries its own individual enthalpy, a heat flux
at a given plane is described by
n
qD
¼ å Ni H i (26-62)
A i¼1

where qD /A is the heat flux due to the diffusion of mass past the given plane and H i is the
partial molar enthalpy of species i in the mixture. When a temperature difference exists,
energy will also be transported by one of the three heat-transfer mechanisms. For example,
the equation for total energy transport by conduction and molecular diffusion becomes
n
q
¼ k=T þ å Ni H i (26-63)
A i¼1

If the heat transfer is by convection, the first energy-transport term in equation (26-63)
would be replaced by the product of the convective heat-transfer coefficient and a DT
driving force.
A process important in many engineering processes as well as in day-to-day
Condensate Boundary of
liquid film
events involves the condensation of a vapor upon a cold surface. Examples of this
gas film
process include the ‘‘sweating’’ on cold water pipes and the condensation of moist
vapor on a cold window pane. Figure 26.12 illustrates the process that involves a
film of condensed liquid flowing down a cold surface and a film of gas through
which the condensate is transferred by molecular diffusion. This process involves
T = T (z)
T1
the simultaneous transfer of heat and mass.
The following conditions will be stipulated for this particular steady-state
physical situation. Pure component A will condense from a binary gas mixture. By
T2
psychrometry, the composition, yA, and the temperature, T1, are known at the plane
yA = yA(z) yA1
T3 z1. The temperature of the condensing surface, T3, is also known. By heat-transfer
considerations, the convective heat-transfer coefficients for the condensate liquid
yA2 film and the gas film can be calculated from equations given in Chapter 20. For
example, in the gas phase, when the carrier gas is air and the vapor content of the
diffusing species is relatively low, the heat-transfer coefficient for natural con-
vection can be estimated by equation (20-5)

z3 z2 z1
0:670RaL1/4
Figure 26.12 Vapor condensation NuL ¼ 0:68 þ (20-5)
on a cold surface. ½1 þ (0:492/Pr)9/16 4/9
480 Chapter 26 Steady-State Molecular Diffusion

Using the general differential equation for mass transfer, equation (25-11), we see that
the differential equation that describes the mass transfer in the gas phase is

d
NA, z ¼ 0 (26-64)
dz
Equation (26-64) stipulates that the mass flux in the z direction is constant over the
diffusion path. To complete the description of the process, the proper form of Fick’s law
must be chosen. If component A is diffusing through a stagnant gas, the flux is defined by
equation (26-3)

cDAB dyA
N A, z ¼ (26-3)
1  yA dz
As a temperature profile exists within the film, and the diffusion coefficient and total gas
concentration vary with temperature, this variation with z must often be considered.
Needless to say, this complicates the problem and requires additional information before
equation (26-3) can be integrated.
When the temperature profile is known or can be approximated, the variation in
the diffusion coefficient can be treated. For example, if the temperature profile is of the
form
 n
T z
¼ (26-65)
T1 z1
the relation between the diffusion coefficient and the length parameter may be deter-
mined by using equation (24-41) as follows:
 3/2  3n/2
T z
DAB ¼ DAB jT1 ¼ DAB jT1 (26-66)
T1 z1
The variation in the total concentration due to the temperature variation can be evaluated
by
P P
c¼ ¼
RT RT1 (z/z1 )n
The flux equation now becomes
 n/2
PDAB jT1 z dyA
NA, z ¼ (26-67)
RT1 (1  yA ) z1 dz
This is the same approach used in Example 15.2, which discussed heat transfer by con-
duction when the thermal conductivity was a variable.
Over a small temperature range, an average diffusion coefficient and the total molar
concentration may be used. With this assumption, equation (26-3) simplifies to

(cDAB )avg dyA


NA; z ¼  (26-68)
(1  yA ) dz
Integrating this equation between the boundary conditions
at z ¼ z1 y A ¼ y A1
and

at z ¼ z2 yA ¼ yA 2
26.4 Simultaneous Momentum, Heat, and Mass Transfer 481

we obtain the relation


(cDAB )avg (yA1  yA2 )
N A, z ¼ (26-69)
(z2  z1 )yB, lm
The temperature, T2, is needed for evaluating (cDAB )avg , the temperature difference
between the liquid surface and the adjacent vapor, and the vapor pressure of species A at
the liquid surface. This temperature may be evaluated from heat-transfer considerations.
The total energy flux through the liquid surface also passes through the liquid film. This
can be expressed by
qz
¼ hliquid (T2  T3 ) ¼ hc (T1  T2 ) þ NA, z MA (H1  H2 ) (26-70)
A
where hliquid is the convective heat-transfer coefficient in the liquid film, hc is the natural
convective heat-transfer coefficient in the gas film, MA is the molecular weight of A, and
H1 and H2 are the enthalpies of the vapor at plane 1 and the liquid at plane 2, respectively,
for species A per unit mass. It is important to realize that there are two contributions to the
energy flux entering the liquid surface from the gas film, convective heat transfer and the
energy carried by the condensing species.
To solve equation (26-70), a trial-and-error solution is required. If a value for the
temperature of the liquid surface is assumed, T2, hc, and ðcDAB Þavg may be calculated. The
equilibrium composition, yA2 , can be determined from thermodynamic relations. For
example, if Raoult’s law holds
p A2 ¼ x A P A
where xA for a pure liquid is 1.0, and the partial pressure of A above the liquid surface is equal
to the vapor pressure PA. By Dalton’s law, the mole fraction of A in the gas immediately
above the liquid is
pA 2 PA
y A2 ¼ or
P P
where P is the total pressure of the system and PA is the vapor pressure of A at the assumed
temperature T2. Knowing (cDAB )avg and yA2 , we can evaluate NAz by equation (26-69).
The liquid-film heat-transfer coefficients can be evaluated, using equations presented in
Chapter 20. A value is now known for each term in equation (26-70). When the left- and
right-hand sides of the equation are equal, the correct temperature of the liquid surface has
been assumed. If the initially assumed temperature does not yield an equality, additional
values must be assumed until equation (26-70) is satisfied.
There are several industrial-unit operations in which heat and mass transfer between gas
and liquid phases occur simultaneously. Distillation, humidification or dehumidification of
air, and water cooling are such operations. In early space exploration, the cooling of the
reentry vehicles by sublimation of ablative material is another example where simultaneous
transfer played an important engineering role.
In the following example, consideration of the simultaneous transfer of mass and heat is
required to predict the flux relations described by Fick’s law.

EXAMPLE 5 An ethanol/water vapor mixture is being distilled by contact with an ethanol/water liquid solution.
The ethanol is transferred from the liquid to the vapor phase and the water is transferred in the
opposite direction. The condensation of water vapor provides the energy for vaporization of ethanol.
Both components are diffusing through a gas film 0.1 mm thick. The temperature is 368 K and the
482 Chapter 26 Steady-State Molecular Diffusion

pressure is 1:013  105 Pa. At these con- Ethanol/water Gas


bulk vapor film
ditions, the pure component enthalpie of

Saturated liquid ethanol/water mixture


(δ)
vaporization of the ethanol and water are
840 and 2300 kJ/kg, respectively. Develop
N EtOH
the flux equation for ethanol vapor. Then vaporization
develop the flux equation assuming that
the components have equimolar heats of Adiabatic
vaporization. wall
We will assume a one-dimensional,
adiabatic molecular mass-transfer process NH2O
condensation
across a gas film of thickness d, as illustra-
ted in Figure 26.13. For one-dimensional,
steady-state mass transfer, the general dif-
ferential equation for mass transfer of
ethanol simplifies to
Figure 26.13 Adiabatic rectification of an ethanol/
dNEtOH, z
¼0 water mixture.
dz
For a binary gas-phase mixture, Fick’s
equation is
dyEtOH
NEtOH, z ¼ cDEtOHH2 O þ yEtOH (NEtOH, z þ NH2 O, z )
dz
We perform an energy balance to relate the flux of ethanol vapor to the flux of water vapor. If the
distillation operation is adiabatic, then all of the energy released when the water condenses must
equal the energy used to produce the alcohol vapor. The energy balance is

NEtOH, z DHy, EtOH ¼ NH2 O, z DHy, H2 O


or
     
kJ kg kJ kg
NEtOH, z 840 46 ¼ NH2 O, z 2300 18
kg kg mol kg kg mol

NEtOH, z ¼ 1:071 NH2 O, z


We recognize that flux is a vector quantity, and that according to Figure 26.13, the flux of ethanol is
opposite in direction to the flux of water. Substituting this relationship into Fick’s equation, we obtain
cDEtOHH2 O dyEtOH
NEtOH, z ¼ 
1 þ 0:071yEtOH dz
As the flux is constant along the z direction, this equation can be integrated directly to obtain
Z d Z yEtOH; 2 dyEtOH
NEtOH, z dz ¼ cDEtOHH2 O
0 yEtOH; 1 1 þ 0:071yEtOH
or
 
cDEtOHH2 O 1 þ 0:071yEtOH;1
N EtOH, z ¼ ln
0:071d 1 þ 0:071yEtOH;2
Now consider a simplified case where the molar heats of vaporization are essentially equal, i.e.,
DHy, EtOH ffi DHy, H2 O . Then from the adiabatic energy balance, it is easy to show that
NEtOH ¼ NH2 O . Then
dyEtOH
NEtOH, z ¼ cDEtOHH2 O þ yEtOH (NEtOH, z þ NH2 O, z )
dz
26.4 Simultaneous Momentum, Heat, and Mass Transfer 483

reduces to
dyEtOH
NEtOH, z ¼ cDEtOHH2 O
dz
which upon integration yields
cDEtOHH2 O
NEtOH, z ¼ (yEtOH, 1  yEtOH, 2 )
d
So we see that equimolar heats of vaporization result in an equimolar-counterdiffusion flux for an
adiabatic-distillation process.

Simultaneous Momentum and Mass Transfer


In several mass-transfer operations, mass is exchanged between two phases. An important
example that we have previously encountered is absorption, the selective dissolution of one
of the components of a gas mixture by a liquid. A wetted-wall column, as illustrated in
Figure 26.14, is commonly used to study the mechanism of this mass-transfer operation, as it
provides a well-defined area of contact between the two phases. In this operation, a thin
liquid film flows along the wall of the column while in contact with a gas mixture. The time
of contact between the two phases is relatively short during normal operation. As only a
small quantity of mass is absorbed, the properties of the liquid are assumed to be unaltered.
The velocity of the falling film will thus be virtually unaffected by the diffusion process.

x=0

W L

∆x W
∆x
∆y

∆y
x=L y

Figure 26. Absorption into a falling-liquid film.

You might also like