4.1 Transferencia Convectiva Entre Fases

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 4

Chapter 29

Convective Mass Transfer


Between Phases

In Chapter 28, convective mass transfer within a single phase was considered; in this
case, mass is exchanged between a boundary surface and a moving fluid and the flux is
related to an individual mass-transfer convective coefficient. Many mass-transfer
operations, however, involve the transfer of material between two contacting phases
where the flux may be related to an overall mass-transfer convective coefficient. These
phases may be a gas stream contacting a liquid stream or two liquid streams if they are
immiscible. In this chapter, we shall consider the mechanism of steady-state mass
transfer between the phases and the interrelations between the individual convective
coefficients for each phase and the overall convective coefficient.
Chapter 30 presents empirical equations for the individual mass-transfer convective
coefficients involved in the interphase transfer. These equations have been established
from experimental investigations. Chapter 31 presents methods of applying interphase
concepts to the design of mass-transfer equipment.

29.1 EQUILIBRIUM
The transport of mass within a phase, by either molecular or convective transport
mechanisms, has been shown to be directly dependent upon the concentration gradient
responsible for the mass transfer. When equilibrium within the system is established, the
concentration gradient and, in turn, the net diffusion rate of the diffusing species
becomes zero. Transfer between two phases also requires a departure from equilibrium
that might exist between the average or bulk concentrations within each phase. As the
deviations from equilibrium provides the concentration driving force within a phase, it is
necessary to consider interphase equilibrium in order to describe mass transfer between
the phases.
Initially, let us consider the equilibrium characteristics of a particular system and then
we will generalize the results for other systems. Consider a two-phase system involving a
gas contacting a liquid; for example, let the initial system composition include air and
ammonia in the gas phase and only water in the liquid phase. When first brought into contact,
some of the ammonia will be transferred into the water phase in which it is soluble and some
of the water will be vaporized into the gas phase. If the gas–liquid mixture is contained
within an isothermal, isobaric container, a dynamic equilibrium between the two phases will
eventually be established. A portion of the molecules entering the liquid phase returns to the
gas phase at a rate dependent upon the concentration of the ammonia in the liquid phase and
the vapor pressure exerted by the ammonia in the aqueous solution. Similarly, a portion of
the water vaporizing into the gas phase condenses into the solution. Dynamic equilibrium is

551
552 Chapter 29 Convective Mass Transfer Between Phases

0.250 0.050

0.200 0.040

PA(atm NH3)

PA(atm NH3)
0.150 0.030

0.100 0.020

0.050 0.010

0.000 0.000
0.00 0.05 0.10 0.15 0.00 0.50 1.00 1.50
XA (mole fraction dissolved NH3 in water) Concentration of dissolved NH3, CAL(kg.mol/m3)

(a) (b)

Figure 29.1 Ammonia solubility in water vs. partial pressure of ammonia at 308C.

indicated by a constant concentration of ammonia in the liquid phase and a constant


concentration or partial pressure of ammonia in the gas phase.
This equilibrium condition can be altered by adding more ammonia to the isothermal,
isobaric container. After a period of time, a new dynamic equilibrium will be established
with a different concentration of ammonia in the liquid and a different partial pressure of
ammonia in the gas. Obviously, one could continue to add more ammonia to the system;
each time a new equilibrium will be reached. Figure 29.1 (a) and (b) illustrates the equ-
ilibrium distribution of ammonia in the gas and liquid phases at 308C.
Figure 29.1(a) presents the concentrations in terms of the partial pressure of the solute
in the gas phase and the mole fraction of the dissolved solute in the liquid phase. Figure
29.1(b) presents the equilibrium distribution as the concentration of ammonia approaches
zero, in terms of the partial pressure in the gas phase and the molar concentration in the
liquid phase; in this dilute concentration range, the equilibrium distribution is linear and the
two concentrations are related by Henry’s law, equation (29-4). There are many graphical
forms of equilibrium data due to the many ways of expressing concentrations in each of the
phases. We will find use for many types of equilibrium plots in Chapter 31.
Equations relating the equilibrium concentrations in the two phases have been
developed and are presented in physical chemistry and thermodynamic textbooks. For
the case of nonideal gas and liquid phases, the relations are generally complex. However, in
cases involving ideal gas and liquid phases, some fairly simple yet useful relations are
known. For example, when the liquid phase is ideal, Raoult’s law applies
pA ¼ x A PA (29-1)
where pA is the equilibrium partial pressure of component A in the vapor phase above the
liquid phase, xA is the mole fraction of A in the liquid phase, and PA is the vapor pressure of
pure A at the equilibrium temperature. When the gas phase is ideal, Dalton’s law is obeyed
pA ¼ y A P (29-2)
where yA is the mole fraction of A in the gas phase and P is the total pressure of the
system. When both phases are ideal, the two equations may be combined to obtain a
relation between the concentration terms, xA and yA, at constant pressure and temperature,
the combined Raoult–Dalton equilibrium law stipulates
yA P ¼ xA PA (29-3)
Another equilibrium relation for gas and liquid phases where dilute solutions are involved
is Henry’s law. This law is expressed by
pA ¼ HcA (29-4)
where H is the Henry’s law constant and cA is the equilibrium composition of A in the
dilute liquid phase. Table 25.1 lists Henry’s constant for selected aqueous solutions.
29.1 Equilibrium 553

An equation similar to Henry’s law relation describes the partition of a solute between
two immiscible liquids. This equation, the ‘‘distribution-law’’ equation is
cA; liquid 1 ¼ KcA; liquid 2 (29-5)
where cA is the concentration of solute A in the specified liquid phase and K is the
partition or distribution coefficient.
A complete discussion of equilibrium relations must be left to physical chemistry and
thermodynamic textbooks. However, the following basic concepts common to all systems
involving the distribution of a component between two phases are descriptive of interphase
mass transfer:
1. At a fixed set of conditions, such as temperature and pressure, Gibbs’s phase rule
stipulates that a set of equilibrium relations exists, which may be shown in the
form of an equilibrium distribution curve.
2. When the system is in equilibrium, there is no net mass transfer between the
phases.
3. When a system is not in equilibrium, components or a component of the system
will be transported in such a manner as to cause the system composition to shift
toward equilibrium. If sufficient time is permitted, the system will eventually
reach equilibrium.
The following examples illustrate the application of equilibrium relations for determin-
ing equilibrium compositions.

EXAMPLE 1 An exhaust stream from a semiconductor fabrication unit contains 3 mol % acetone and 97 mol % air.
In order to eliminate any possible environmental pollution, this acetone-air stream is to be fed to a
mass-transfer column in which the acetone will be stripped by a countercurrent, falling 293 K water
stream. The tower is to be operated at a total pressure of 1:013  105 Pa. If the combined Raoult–
Dalton equilibrium relation may be used to determine the distribution of acetone between the air and
the aqueous phases, determine
(a) the mole fraction of acetone within the aqueous phase, which would be in equilibrium with
the 3 mol % acetone gas mixture.
(b) the mole fraction of acetone in the gas phase, which would be in equilibrium with 20 ppm
acetone in the aqueous phase.
At 293 K, the vapor pressure of acetone is 5:64  104 Pa.
(a) By Raoult–Dalton law when yA ¼ 0:03

YA P ¼ xA PA
(0:03)(1:013  105 Pa) ¼ xA (5:64  104 Pa)
or xA ¼ 0:0539 mole fraction acetone

(b) 20 ppm acetone in solution

20 g acetone
¼
999; 980 g water
20 g/(58 g/mol)
¼
999; 980 g(18 g/mol)
mol acetone
¼ 6:207  106
mol water
554 Chapter 29 Convective Mass Transfer Between Phases

For the dilute solution, th mole fraction of acetone will be


86:027  106 mol acetone
xA ¼ ¼ 6:207  106
1:0 mol water þ 6:027  106 mol acetone
By Raoult-Dalton law
yA P ¼ xA PA
yA (1:013  105 Pa) ¼ (6:207  106 )(5:64  104 Pa)
or yA ¼ 3:45  106 mole fraction acetone

EXAMPLE 2 The Henry’s law constant for oxygen dissolved in water is 4:06  109 Pa/(mol of O2 per total mol of
solution) at 293 K. Determine the solution concentration of oxygen in water that is exposed to dry air
at 1:013  105 Pa and 293 K.

Henry’s law can be expressed in terms of the mole fraction units by


pA ¼ H 0 xA

where H0 is 4:06  109 Pa/(mol of O2/total mol of solution).


From example 24.1, we recognize that dry air contains 21 mol% oxygen. By Dalton’s law
pA ¼ yA P ¼ (0:21)(1:013  105 Pa) ¼ 2:13  104 Pa:

The equilibrium mole fraction of the liquid at the interface is computed by Henry’s law
PA 2:13  104 Pa
xA ¼ ¼
H 4:06  109 Pa/(mol O2 /mol soln)
¼ 5:25  106 (mol O2 /mol soln)

For one cubic meter of very dilute solution, the moles of water in the solution will be approximately
 
1
nwater ¼ (1 m )(1  10 kg/m )
3 3 3
0:018 kg=mol
¼ 5:56  104 mol
The total moles in the solution is essentially the moles of water because the concentration of oxygen
is quite low. Accordingly, the moles of oxygen in one cubic meter of solution is
noxygen ¼ (5:25  106 mol O2 /mol soln)(5:56  104 mol soln)
¼ 0:292 mol of O2
The saturation concentration is
(0:292 mol/m3 )(0:032 kg/mol) ¼ 9:34  103 kg O2 /m2 (9:34 mg/L)

29.2 TWO-RESISTANCE THEORY


Many mass-transfer operations involve the transfer of material between two contacting
phases. For example, in gas absorption, as illustrated in Figure 29.2, a solute is transferred
from the gas phase into a liquid phase. The interphase transfer involves three transfer steps:
(1) the transfer of mass from the bulk conditions of one phase to the interfacial surface,
(2) the transfer across the interface into the second phase, and (3) the transfer to the bulk
conditions of the second phase.

You might also like