You are on page 1of 4

INTERPHASE MASS TRANSFER THEORY

The process of mass transfer from the bulk of one phase to the interphase
surface and then from the interphase to the bulk of another phase is called
interphase mass transfer.
In cases involving ideal gas and liquid phases, the Raoults law can be applied
for relating the equilibrium concentrations in the two-phases.
y A P=x A PA (1)
where PA is the vapour pressure of pure solute A at the equilibrium condition
and P is the equilibrium pressure. If the liquid phase does not behave ideally, the
following modified form of Raoults law can be applied:
y A P=x A A P A (2)
where A is the activity coefficient of solute A in solution. For dilute solution
Henrys law can be used to express the equilibrium relations which is expressed
by
p A = y A P=H x A (3)
where p A is the equilibrium partial pressure of solute A in the vapour phase
and H is the Henrys law constant.
Two-film Resistance Theory
Consider the mass transfer of a solute A from the bulk of a gas phase to the bulk
of a liquid phase. This can be shown graphically in terms of distance through the
phases as shown in Figure 3.6.
The concentration of A in the main body of the gas is y AG mole fraction and it
falls to yAi at the interface. In the liquid, the concentration falls from x Ai at the
interface to xAL in the bulk liquid.
According to Lewis and Whitman (1924), there is no resistance to solute transfer
across the interface separating the phases. Only diffusional resistances are
residing in the fluids.
The equilibrium concentrations yAi and xAi are obtained from the systems
equilibrium distribution curve. This concept has been called the two-resistance
theory.
In the figure the concentration rise at the interface from y Ai to xAi is not a
barrier to diffusion in the direction gas to liquid. They are equilibrium
concentrations. For steady state mass transfer, the rate at which A reaches the
interface from the gas must be equal to the rate at which it diffuses to the bulk
liquid, so that no accumulation or depletion of A at the interface occurs.
Therefore the mass transfer flux of A in terms of mass transfer film coefficient
for each phase can be written as:

N A =k y ( y AG y Ai ) =k x ( x Aix AL )( 4) After rearranging it can be written as:

( y AG y Ai ) k x
(x A ix AL )

ky

(5)

Experimentally the mass transfer film coefficients Ky and Kx are difficult to


measure except for cases where the concentration difference across one phase is
small and can be neglected. Under these circumstances, the overall mass transfer
coefficients Ky and Kx are measured on the basis of the gas phase or the liquid
phase. The entire two-phase mass transfer effect can then be measured in terms
of gas phase molar fraction driving force as:
N A =K y ( y AG y A )(6)
where, Ky is based on the overall driving force for the gas phase, in mole/m 2.s and
y A is the value of concentration in the gas phase that would be in the
equilibrium with x AL . Similarly, the entire two-phase mass transfer effect can
then be measured in terms of liquid phase molar fraction driving force as:

N A =K x ( x A x AL )(7)
where Kx is based on the overall driving force for the liquid phase, in mole/m 2.s

and x A is the value of concentration in the liquid phase that would be in the
y AG . A relation between the overall coefficients and the
equilibrium with
individual mass transfer film coefficients can be obtained when the equilibrium
relation is linear as y Ai=m x Ai .
The linear equilibrium condition can be obtained at the low concentrations, where
Henrys law is applicable. Here the proportionality constant m is defined as m=
H/P. Utilizing the relationship, y Ai=m x Ai , gas and liquid phase concentrations
can be related by
y A =m x AL (8) y AG=m x A (9)
Rearranging Equation (6), one can get


1 ( y AG y A )
=
(10)
Ky
NA

From geometry, y AG y A

can be written as

y
( AG y Ai )+( y Ai y A )(11)
y AG y A =
Substituting Equation (11) in Equation (10)
y
y
x
m( Aix AL )
(12)
NA
( AG y Ai )
+
NA

( AG y Ai) ( y Ai y A )
+
=
NA
NA

1 ( y AG y A )
=
=
Ky
NA
The substitution of equation(6) in (12) relates overall gas phase mass transfer
coefficient ( K y ) to the individual film coefficients by
1
1 m
= + (13)
K y k y kx

Similarly the relation of overall liquid phase mass transfer coefficient ( K x ) to


the individual film coefficients can be derived as follows:
y
( AG y Ai) ( x Aix AL )
+
( 14)
mN A
NA

1 ( x A x AL )
=
=
Kx
NA
or

1
1
1
=
+ (15)
K x mk y k x

The following relationships between the mass transfer resistances from the
Equations (13) and (15)
1/k y
Resistance gas phase
=
(16)
Total resistanceboth phases 1/ K y

1/ k x
Resistance liquid phase
=
(17)
Total resistanceboth phases 1/ K x
If solute A is very soluble in the liquid, m is very small. Then the term m/k x in
Equation (13) becomes minor and consequently the major resistance is
represented by 1/ky. In this case, it is said that the rate of mass transfer is gas
phase controlled. In the extreme it becomes:

1
1
(18)
K y ky
The total resistance equals the gas film resistance. The absorption of a very
soluble gas, such as ammonia in water is an example of this kind. Conversely
when solute A is relatively insoluble in the liquid, m is very large. Consequently
the first term of Equation (15) becomes minor and the major resistance to the
mass transfer resides within the liquid. The system becomes liquid film
controlling. Finally this becomes:
1 1
(19)
K x kx
The total resistance equals the liquid film resistance. The absorption of a gas of
low solubility, such as carbon dioxide or oxygen in water is of this type of
system.

You might also like