You are on page 1of 30

Equation Section 4© May 22, 2008, Cerro, Higgins & WhitakerEquation Section 5

Chapter 5
Gas-Liquid Systems
In the previous chapter, we began our study of macroscopic mass and mole balances for
multicomponent systems. There we encountered a variety of measures of concentration and we summarize
these measures as

{
ρ A = mass of species A
per unit volume } (5-1a)

A= N
ρ = ∑ρ
A =1
A , total mass density (5-1b)

ωA = ρ A ρ , mass fraction (5-1c)

cA = ρ A MWA , molar concentration (5-1d)

A= N
c = ∑c
A=1
A , total molar concentration (5-1e)

y A or x A = c A c , mole fraction (5-1f)


In the analysis of gas-phase systems it is often important to relate the concentration to the pressure and
temperature. This is done by means of an equation of state, often known as a p-V-T relation. In this
chapter we will make use of the ideal gas relations; however, many processes operate under conditions
such that the ideal gas laws do not apply and one must make use of more general p-V-T relations. Non-
ideal gas phase behavior will be studied in a subsequent course in thermodynamics.

5.1 Ideal Gas-Phase Behavior


For an N-component ideal gas mixture we have the following relations
pA V = n A RT , A = 1, 2, ...N (5-2)

Here p A is the partial pressure of species A and R is the gas constant. Values of the gas constant in
different units are given in Table 5-1. If we sum Eq. 5-2 over all N species we obtain
pV = nRT (5-3)
where
A= N
p = ∑p
A=1
A (5-4)

A= N
n = ∑n
A=1
A (5-5)

Equations 5-2 through5-5 are sometimes referred to as Dalton's Laws.

108
Gas-Liquid Systems 109

Table 5-1. Numerical values of the gas constant, R


Numerical value Units
8.314 m3 Pa/ mol K
8.314 J/mol K
0.08314 liter bar/mol K
82.06 atm-cm3 /mol K
1.987 cal/mol K

In Figure 5-1 we have illustrated a constant pressure, isothermal mixing process. In the compartment

VB
VA

species A species B constant


pressure

a) before mixing

mixture of A and B constant


pressure

b) after mixing

Figure 5-1. Constant pressure, isothermal mixing process

containing species A illustrated in Figure 5-1a, we can use Eq. 5-2 to obtain

pVA = n A RT (5-6)

while in the compartment containing species B we have

pVB = nB RT (5-7)

Upon removal of the partition and mixing, we have the situation illustrated in Figure 5-1b. For that
condition, we can use Eq. 5-3 to obtain
110 Chapter 5

pV = (n A + nB )RT (5-8)

where the volume is given by

V = VA + VB + ∆Vmix (5-9)

By definition, an ideal gas mixture obeys what is known as Amagat's Law, i.e.

∆Vmix = 0 , Amagat's Law (5-10)

Amagat’s law can also be expressed as


A= N

∑V
A=1
A = V , Amagat's Law (5-11)

In the gas phase the mole fraction is generally denoted by y A while x A is reserved for mole fractions in the
liquid phase. For ideal gas mixtures it is easy to show, using Eqs. 5-1f, 5-2 and 5-3, that the mole fraction
is given by
yA = pA p (5-12)

This is an extremely convenient representation of the mole fraction in terms of the partial pressure;
however, one must always remember that it is strictly valid only for an ideal gas1.

EXAMPLE 5.1 Flow of an ideal gas in a pipeline


A large pipeline is used to bring natural gas from Oklahoma to Nebraska as illustrated in
Figure 5.1. Natural gas, consisting of methane with small amounts of ethane, propane, and other

Figure 5.1. Transport of natural gas from Oklahoma to Nebraska

low molecular mass hydrocarbons, can be assumed to behave as an ideal gas at ambient
temperature. At the pumping station in Glenpool, OK, the pressure in the 20-inch pipeline is 2900
psia and the temperature is To = 90 F . At the receiving point in Lincoln, NE, the pressure is down
to 2100 psia and the temperature is T1 = 45 F . The mass average velocity of the gas at Glenpool
is 50 ft/s. Assuming ideal gas behavior and the gas being 100% methane, compute the following:
a) Mass averaged velocity at the end of the pipeline
b) Mass and molar flow rates at both ends of the pipeline.
This problem has been presented in terms of a variety of units, and it is often convenient to
express all variables in terms of SI units as follows:

1
Non-ideal gases are discussed briefly in Section 5.3.
Gas-Liquid Systems 111

6890 Pa
( 2900 psia ) × = 19.98 MPa
psia
6890 Pa
( 2100 psia ) × = 14.47 MPa
psia
5C
( 90 F − 32 F ) × = 32.2 C = 305.4 K
9F
5C
( 45 F − 32 F ) × = 7.2 C = 280.4 K
9F
0.3048 m
( 50 ft/s ) × = 15.24 m/s
ft
0.0254 m
Do = D1 = ( 20 in ) × = 0.508 m
in
π ( 0.508 m )
2
π D2
A o = A1 = = = 0.2027 m 2
4 4
Since the natural gas is assumed to be pure methane, we can obtain the molecular mass from
Table 2B in Appendix B where we find MWCH 4 = 16.043 g/mol . The volume per mole of
methane can be determined using the ideal gas law given by Eq. 5-2 along with the value of the
gas constant found in Table 5-1. The volume per mole at the entrance is determined as

m3 Pa
V
8.314 ( 305.4 K )
RT mol K
Vˆo = o = = = 1.2 × 10−4 m3 / mol (1)
n p 19.98 106 Pa

while the volume per mole at the exit is given by


V RT
Vˆ1 = 1 = = 1.53 × 10−4 m3 / mol (2)
n p

The gas densities at the entrance and exit of the pipeline are given by

MWCH 4 16.043 g/mol


(ρCH 4 ) o = = = 133, 700 g/m3 = 133.7 kg/m3 (3)
Vˆo 1.2 × 10−4 m3 /mol

MWCH 4
(ρCH 4 )1 = = 104.8 kg/m3 (4)
Vˆ 1

To perform a mass balance for the pipeline, we begin with species mass balance given by

∫ρ ∫ (ρ v ) ⋅ n dA ∫ r dV
d
A dV + A A = A (5)
dt
V A V

Since there are no chemical reactions and no accumulation, the first and last terms in this result are
zero. The control volume is constructed in the obvious manner, thus there is an entrance in
Glenpool, OK and an exit in Lincoln, NE. Since the mass average velocity and the diameter of the
pipeline are given, we express the mass balance as

(ρCH 4 ) o v o Ao = (ρCH 4 )1 v1 A1 (6)


112 Chapter 5

The only unknown in this result is the velocity of the gas at the exit of the pipeline. Solving for
v1 we obtain

A o (ρCH 4 )o 133.7 kg/m3


v1 = vo = (15.24 m/s ) = 19.44 m/s (7)
A1 (ρCH 4 )1 104.8 kg/m3

The mass flow rate is a constant given by

( )
m o = m 1 = (ρCH 4 ) o v o Ao = 133.7 kg/m3 (15.24 m/s ) 0.2027 m 2 ( ) = 413kg/s (8)

from which we determine the constant molar flow to be

(m CH 4 )o 413kg/s
M o = M 1 = = = 25.74 × 103 mol/s (9)
MWCH 4 16.043g/mol

In order to use Eq. 5-3 to estimate the density of a pure gas, we multiply by the molecular mass and
arrange the result in the form

n MW p MW
ρ = = (5-13)
V RT

For an ideal gas mixture, one uses the definitions of the total mass density of a mixture (Eq. 5-1b) and total
pressure (Eq. 5-4) with Eq. 5-2 to obtain
A= N A= N A= N

∑ ∑ ∑
n A MWA p A MWA p MW
ρ = ρA = = = (5-14)
V RT RT
A =1 A=1 A=1

Here we have used MW to represent the mean molecular mass defined by


A= N A= N
MW =
1
p ∑p
A =1
A MWA = ∑y A=1
A MWA (5-15)

These results are applicable when molecule-molecule interaction is negligible and this occurs for many
gases under ambient conditions. At low temperatures and high pressures, gases depart from ideal gas
behavior, and under those conditions one should use more accurate equations of state, such as those
presented in a standard course on thermodynamics2.
EXAMPLE 5.2 Molecular mass of air
The air we breathe has a composition that depends on position. Air pollution sources abound
and these sources add minute amounts of chemicals to the atmosphere. Combustion of fuels in
cars and power plants are a source for sulfur dioxide, oxides of nitrogen, and carbon monoxide.
Chemical industries add pollutants such as ammonia, chlorine, and even hydrogen cyanide to the
atmosphere. In some locations, there are minute amounts of other gases such as argon, helium,
and radon. The standard dry air, for the purpose of combustion computations, is assumed to be a
mixture of 79% by volume of nitrogen and 21% by volume of oxygen. In this example, we want
to determine the average molecular mass of standard air.
For an ideal gas, a volume percentage is also a percentage of the number of moles of a species
with respect to the total number of moles. Thus, the volume percentages of nitrogen and oxygen
can be simply translated to molar fractions:

Sandler, S.I. 2006, Chemical, Biochemical, and Engineering Thermodynamics, 4th edition, John Wiley and Sons,
2.

New York.
Gas-Liquid Systems 113

79% by volume of nitrogen → yN 2 = 0.79


21% by volume of oxygen → yO2 = 0.21

We can use Eq. 5-15 to compute the average molecular mass of standard air according to

MW = MWair = y N 2 MWN 2 + yO2 MWO2 = 28.85 g/mol

5.2 Liquid Properties and Liquid Mixtures


When performing material balances for liquid systems, one must have access to reliable liquid
properties. Unlike gases, the densities of liquids are weak functions of pressure and temperature, i.e., large
changes in pressure and temperature result in small changes in the density. The changes in liquid density
due to changes in pressure are determined by the coefficient of isothermal compressibility which is defined
by
coefficient of isothermal  1  ∂ρ 
κ =   =   (5-16)
 compresibility  ρ  ∂ p T

Changes in liquid density due to changes in temperature are determined by the coefficient of thermal
expansion which is defined by

coefficient of thermal  1  ∂ρ 
β =   = −   (5-17)
 expansion  ρ  ∂T p

The coefficient of thermal expansion, β , is defined with a negative sign since the density of most liquids
decreases with increasing temperature. Using a negative sign in the definition of the thermal expansion
coefficient makes β positive for most liquids. There is an interesting counterexample, however, and it is
the density of liquid water at low temperatures. For liquid water, β < 0 for the range of temperature
between 4 C and the freezing point of water, 0 C. If it were not for this feature, water in lakes and rivers
would freeze from the bottom during the winter, and this would destroy most aquatic life.
The density of ideal liquid mixtures is computed using Amagat’s law. Assuming that the total volume
of a mixture is equal to the sum of the volume of the components of the mixture, we obtain
A= N A= N

∑ρ ∑V
mA
V = o
= A (5-18)
A =1 A A=1

Here V is the volume of the mixture, while mA and ρoA represent the masses and densities of the pure
components. Equation 5-18 can be used to compute the density of the mixture according to

m m 1
ρ = = A= N
= A= N
(5-19)
V

A =1
mA
ρoA ∑
A =1
ωA
ρoA

Non-ideal behavior of liquid mixtures is a very complex topic. At low to moderate pressures, this version
of Amagat’s law for liquids is a satisfactory approximation. When some of the components of a liquid
mixture are above their boiling points, or if the components are polar, the use of Eqs. 5-18 or 5-19 may give
significant errors3.

3.
Reid, R. C., Prausnitz, J. M., and Sherwood, T. K., 1977, The Properties of Gases and Liquids, Sixth Edition, New
York, McGraw-Hill Books.
114 Chapter 5

5.3 Vapor Pressure of Liquids


If we study the p-V-T characteristics of a real gas using the experimental system shown in Figure 5-2,
we find the type of results illustrated in Figure 5-3. In the system illustrated in Figure 5-2, a single
component is contained in a cylinder immersed in a constant temperature bath. We can increase or
decrease the pressure inside the cylinder by simply moving the piston. When the specific volume, that is
the volume per mole, is sufficiently large, the distance between molecules is large enough (on the average)
so that molecular interaction becomes unimportant. For example, at the temperature T3 , and a large value
of V / n , we observe ideal gas behavior. This is illustrated by the fact that at a fixed temperature we have

pV n = constant (5-20)

However, as the pressure is increased (and the volume decreased) in the system illustrated in Figure 5-3, a
point is reached where liquid appears and the pressure remains constant as the volume continues to
decrease. This pressure is referred to as the vapor pressure and we will identify it as pvap . Obviously the

pressure = p

volume = V

moles = n

Constant Temperature Reservoir

Figure 5-2. Experimental study of p-V-T behavior

vapor pressure is a function of the temperature and knowledge of this temperature dependence is crucial for
the solution of many engineering problems.
Gas-Liquid Systems 115

Figure 5-3. p-V-T behavior of methane

In a course on thermodynamics you will learn that the Clausius-Clapeyron equation gives a reasonably
accurate representation of the vapor pressure as a function of temperature. The Clausius-Clapeyron
equation can be expressed as

 ∆H vap  1 1 
p A,vap = p oA,vap exp  −  −  (5-21)
 R T To  

in which p A,vap represents the vapor pressure at the temperature T. We have used p oA,vap to represent the
vapor pressure at the reference temperature To , while ∆H vap represents the molar heat of vaporization. A
more accurate, but empirical, expression for the vapor pressure is given by Antoine’s equation

B
log10 ( p A,vap ) = A − (5-22)
(θ + T )
in which p A,vap is determined in mm Hg and T is specified in C. The coefficients A, B, and θ are given in
Table 2B of the Appendix B for a variety of compounds. Note that Eq. 5-22 is dimensionally incorrect and
must be used with great care as we indicated in our discussion of units in Sec. 2.3

EXAMPLE 5.3: Vapor pressure of a single component


In this example we wish to estimate the vapor pressure of methanol at 25 C using the
Clausius-Clapeyron equation. The heat of vaporization of methanol is ∆H vap = 8426 cal/mole at
the normal boiling point of methanol, 337.8 K . The heat of vaporization is a function of
temperature and pressure. The data given for the heat of vaporization is for the temperature
T = 337.8 K = 64.6 C . At this temperature, the vapor pressure of methanol is equal to
atmospheric pressure. In order to estimate the vapor pressure at 25 C , we use the normal boiling
temperature as the reference temperature. Normally we would want to compute the value of the
116 Chapter 5

heat of vaporization at 25 C using a thermodynamic relationship and then use an average value
for ∆H vap in Eq.5-21. For the purpose of this example, we will estimate the vapor pressure at
25 C using the heat of vaporization at 64.6 C . All variables can be converted into SI units as
follows:
Temperature: 25 C + 273.16 = 298.16 K
To = 337.8 K
o
pM,vap = 1 atm = 101,300 Pa
∆H vap = (8426 cal/mol) ( 4.186 J/cal) = 35, 271 J/mol

Substitution of these results into Eq. 5-21 gives

 35271 J/mol  1 1 
pM,vap = (101,300 Pa ) exp  − 3  −   = 19,112 Pa
 8.314 m Pa/mol K  298.2 K 337.8 K  

Vapor pressures estimated using the Clausius-Clapeyron equation can exhibit substantial errors with
respect to experimental values of vapor pressure. This is caused by the fact that the assumptions made in
the development of this equation are not always valid. The semi-empirical equation known as Antoine’s
equation has the advantage that it is based on the correlation of experimental values of the vapor pressure.

EXAMPLE 5.4. Vapor pressure of single components using Antoine’s equation


In this example, we determine the vapor pressure of methanol at 25 C using Antoine’s equation,
Eq. 5-22, and compare the result with the vapor pressure computed in Example 5.3. The
numerical values of the coefficients in Antoine’s equation are obtained from Table II in the
appendix, and they are given by
A = 8.07246 , B = 1574.99 , θ = 238.86

Substitution of these values into Eq. 5-22 gives


1,574.99
log pM,vap = 8.07246 − = 2.10342
238.86 + 25

and the vapor pressure of methanol at 25 C is pM,vap = 126.9 mmHg = 16,912 Pa. The result
computed using the Clausius-Clapeyron equation was, pM,vap = 19,112 Pa, thus the two results
differ by 11%. The results of this example clearly indicate that it is misleading to represent
calculated values of the vapor pressure to five significant figures.

5.3.1 Mixtures
The behavior of vapor-liquid systems having more than one component can be quite complex;
however, some mixtures can be treated as ideal. In an ideal, vapor-liquid multi-component system, the
partial pressure of species A in the gas phase is given by

pA = x A p A,vap (5-23)

Here p A is the partial pressure of species A in the gas phase, x A is the mole fraction of species A in the
liquid phase, and p A,vap is the vapor pressure of species A at the temperature under consideration. It is
important to remember that Eq. 5-23 is an equilibrium relation; however, when the condition of local
thermodynamic equilibrium is valid Eq. 5-23 can be used to calculate values of p A for dynamic systems.
Gas-Liquid Systems 117

For an ideal gas, the gas-phase mole fraction can be expressed as

yA = pA p (5-24)

and this result can be used with Eq. 5-23 to obtain a relation between the gas and liquid-phase mole
fractions that is given by
yA = (
x A p A,vap p ) (5-25)

This result is sometimes referred to as Raoult’s law. For a two-component system we can use Eq. 5-25
along with the constraint on the mole fractions

x A + xB = 1, y A + yB = 1 (5-26)

to obtain the following expression for the mole fraction of species A in the gas phase:

α AB x A
yA = (5-27)
1 + x A (α AB − 1)

Here α AB is the relative volatility defined by

p A,vap
α AB = (5-28)
pB ,vap

For a dilute binary solution of species A, one can express Eq. 5-27 as
yA = α AB x A , for x A (α AB − 1) << 1 (5-29)

and this special form of Raoult’s law is often referred to as Henry’s law. For an N-component system, one
can express Henry’s law as
y A = K A xA , for x A << 1 (5-30)

Here K A is referred to as the Henry’s law constant even though it is not a constant since it depends on the
composition of the liquid, i.e.
K A = F ( xB , xC , ....xN ) (5-31)

This treatment of gas-liquid systems is extremely brief and devoid of the rigor that will be encountered in a
comprehensive discussion of phase equilibrium. However, we now have sufficient information to solve a
few simple mass balance problems that involve gas and liquid phases.

5.4 Equilibrium Stages


Mass transfer of a chemical species from one phase to another is an essential feature of the mixing and
purification processes that are omnipresent in the chemical and biological process industries. A
comprehensive analysis of mass transfer requires an understanding of the prerequisite subjects of fluid
mechanics, thermodynamics and heat transfer; however, there are some mass transfer processes that can be
approximated as equilibrium stages and these processes can be analyzed using the techniques presented in
this text. Most students are familiar with an equilibrium stage when it is carried out in a batch-wise
manner, since this is a common purification technique used in organic chemistry laboratories.
If an organic reaction produces a desired product that is soluble in an organic phase and an undesirable
product that is soluble in an aqueous phase, the product can be purified by liquid-liquid extraction as
illustrated in Figure 5-4. In the first step of step of this process, the mixture from a reactor is placed in a
separatory funnel. Water is added, the system is agitated, and the phases are allowed to equilibrate and
separate. The amount of the undesirable product in the organic phase is reduced by an amount related to
the volumes of the organic and aqueous phases and the distribution coefficient defined by
118 Chapter 5

solubility in the organic phase (moles of A/vol)


KA = distribution coefficient = (5-32)
solubility in the aqueous phase (moles of A/vol)

The analysis of the process illustrated in Figure 5-4 is relatively simple provided that the following
conditions are valid: (1) There are negligible changes in the volumes of the organic and aqueous phases

Figure 5-4. Batch-wise liquid-liquid extraction

because of the mass transfer process, and (2) the linear equilibrium relation given by Eq. 5-32 is valid. If
the batch process illustrated in Figure 5-4 is repeated n times, the concentration of species A in the organic
phase is given by
1
( c A ) final = ( c A )initial n (5-33)
 Vaq 
 1 + 
 n K A Vorg 

Here we have used Vaq to represent the volume of the aqueous phase and Vorg to represent the volume of
the organic phase. Equation 5-33 indicates that repeated batch-wise extractions can be used to reduce the
concentration of species A in the organic phase to arbitrarily small values.

EXAMPLE 5.5 Mass transfer in a gas-liquid system


Sulfur dioxide is “scrubbed” from a stream of air in the absorption column illustrated in
Figure 5.5a. The column is designed to treat 1200 ft3/min of the contaminated air stream, and the
mole fraction of sulfur dioxide in the incoming air stream is 0.02. The air stream leaving the
scrubber is assumed to contain a negligible amount of SO2 . By negligible, we mean that ( ySO2 )4
can be set equal to zero provided that small causes produce small effects4. This approximation is
consistent with the idea that the air stream leaving the scrubber is in equilibrium with the water
entering the scrubber. Under these circumstances, the scrubber is considered to be an “equilibrium
stage” and we will study equilibrium stages in more detail in Sec. 5.6.

4
See page 4 in Birkhoff, G. 1960, Hydrodynamics: A Study in Logic, Fact, and Similitude, Princeton University Press,
Princeton, New Jersey.
Gas-Liquid Systems 119

Figure 5.5a. Sulfur dioxide scrubber

The volumetric flow rate of the pure water entering the scrubber is 20 ft3/min. Given these
conditions, we want to know the mole fraction of SO2 in the water stream leaving the system and
the molar fraction of water in the air leaving the system. The scrubber operates at atmospheric
pressure and the temperature is constant at 20 C . In determining the mole fraction of SO2 in the
water, we will ignore the reaction with water to form sulfurous acid, and it will be left as an
exercise for the student to explore the impact of this assumption. In addition to this simplification,
we will assume that the transfer of air from the gas stream to the liquid stream is negligible. This
last assumption is justified by the small solubility of both nitrogen and oxygen in water at room
temperature.
To construct a control volume for this system, we follow the rules described previously and
make cuts where information is given and required. One such cut is illustrated in Figure 5.5b
where we have shown a cut where information is given, i.e., at the incoming stream of water.
Information is also given at the incoming contaminated air stream, and we are required to
determine the both amount of water in the outgoing air stream and the amount of sulfur dioxide in
the outgoing water stream. In conclusion, cuts must be made at all four streams that are entering
and leaving the system. When these cuts are joined where c A v ⋅ n = 0 , we obtain the control
volume illustrated in Figure 5.5c where we have constructed the surface of the control volume in a
manner that makes it visible. In reality, the surface of the control volume, in regions other than at
120 Chapter 5

Figure 5.5b. Cut for sulfur dioxide scrubber

the entrances and exits, is located at the solid-air interface where the molar flux of all species is
zero. For convenience, we transform given flow rates into SI units:

Q1 =
(1200 ft 3
)(
/ min 0.0283m3 /ft 3 ) = 0.566
m3
, Q3 = 0.00943
m3
60s/min s s

For steady-state conditions and no chemical reaction, we make use of the molecular species
balance in the form

∫ c v ⋅ n dA
A
A = 0 (1)

Here we have replaced the species velocity v A with the mass average velocity v on the basis of
the ideas discussed in Secs. 4.3 through 4.5. Evaluation of the four terms associated with
Eq. 1 for the sulfur dioxide leads to

Sulfur dioxide: (cSO2 )1 Q1 = (cSO2 ) 2 Q2 (2)

It will be left as an exercise (see Problem 5-9) for the student to develop arguments supporting the
idea that air can be treated as a single species for this particular process. Under these
circumstances, use of Eq. 1 for air leads to

Air: (cair )1 Q1 = (cair ) 4 Q4 (3)

This result requires the assumption that the absorption of nitrogen, oxygen, etc., in the water is
Gas-Liquid Systems 121

Figure 5.5c. Control volume for sulfur dioxide scrubber

negligible. The molar concentration of SO2 can be expressed in terms of the mole fraction
according to
cSO2 = ySO2 c , in the gas phase (4a)

cSO2 = xSO2 c , in the liquid phase (4b)

where c is the total molar concentration. Use of Eqs. 4 in Eq. 2 leads to

Sulfur dioxide: (ySO2 )1 c1 Q1 = (xSO2 )2 c2 Q2 (5)

and a similar representation for Eq. 3 provides

Air: c1 (yair )1 Q1 = c4 (yair ) 4 Q4 (6)

A little thought will indicate that this latter result can be expressed as

c1 1 − (ySO2 )1  Q1 = c4 1 − (yH 2O )4  Q4 (7)

Finally, the balance of water leads to

Water: (
c 3 xH 2O )3 Q3 = c2 ( xH O )2 Q2 + c4 ( yH O )4 Q4
2 2
(8)

Since the temperature and pressure are assumed to be constant, the total molar concentration in the
gas (air) streams can be treated as a constant

101300 Pa
c1 = c4 = p RT = = 0.0416 kmol/m3 (9)
8.314 Pa m3 /mol K 293.16 K
122 Chapter 5

Finally, we must determine the vapor pressure of water at 20 C using Antoine’s equation. From
Appendix II, Antoine’s constants for water are:
A = 7.94915, B = 1657.46, C = 227.03

Substitution of these results into Antoine’s equation5 gives the vapor pressure of water at 20 C ,
and we can use Raoult’s law to compute the mole fraction of water in stream #4 as

pHo 2O
pHo 2O = 17.36 mm Hg , ( yH 2 O ) 4 = = 0.0228 (10)
p

Here we have assumed that the water vapor in stream #4 is in equilibrium with the pure liquid
water entering the scrubber at 20 C . This type of simplification will be explored carefully in
subsequent courses on heat and mass transfer.
On the basis of Eqs. 2 to 10 we can complete the degree-of-freedom table, given by Table 5.5,
in order to conclude that we have a solvable problem.

Table 5.5. Degrees-of-Freedom

Stream Variables
compositions N x M = 12
flow rates M=4

Generic Degrees of Freedom (A) 16

Number of Independent Balance Equations


mass/mole balance equations N=3 Eqs. 5, 7, and 8

Number of Constraints for Compositions 4

Generic Specifications and Constraints (B) 7

Specified Stream Variables


compositions 0
flow rates 2

Constraints for Compositions 4

Auxiliary Constraints 3 Eqs. 9 and 10

Particular Specifications and Constraints (C) 9

Degrees of Freedom (A - B - C) 0

To develop a solution to this problem, we note that Eq. 5 takes the form

c1 ( ySO2 )1 Q1 = 4.71 10−4 kmol/s = c2 ( xSO2 )2 Q2 (11)

while the mole balance of air, Eq. 7, reduces to.

5. See Sec. 2.4 for a discussion of the convenience unit mm Hg as a measure of pressure.
Gas-Liquid Systems 123

c1 1 − (ySO2 )1  Q1 = 0.0231 kmol/s = c4 1 − (yH 2O )4  Q4 (12)

Since we know the total molar concentration and the mole fraction of water in the exit air stream,
we can compute the flow rate as

kmol m3
0.0231 Kmol/s = 0.0416 [1 − 0.0228] Q4 ; Q4 = 0.5682 (13)
m3 s

The mole balance for water given by Eq. 8 reduces to

(
c3 xH 2O )3 Q3 (
− c4 yH 2O )4 Q4 = 0.5225 kmol/s = c2 xH 2O ( )2 Q2 (14)

We add now Eqs. 11 and 14, using the composition constraint, to get

0.523 kmol/s = c2 xH 2O ( )2 Q2 (
+ c2 xSO2 )2 Q2 = c2 Q2 (15)

It is easy now to compute the mole fractions of water and sulfur dioxide in stream #2 using
Eqs. 11, 14, and 15. This leads to
(xH 2 O )2 = 0.9991 , (xSO2 ) 2 = 0.0009 (16)

In the previous example, we studied a process for which the use of molar quantities was appropriate. In
other situations, mass fractions and mass flow rates are more appropriate and in some cases it is convenient
to use both. The following example is an illustration of the latter situation.

EXAMPLE 5.6. Use of air to dry wet solids


In Figure 5.6 we have illustrated a counter-current air drier. The solids entering the drier
contain 20% water on a mass basis and the mass flow rate of the wet solids entering the drier is
1000 lbm/hr. The dried solids contain 5% water on a mass basis. The air leaving the drier is
assumed to be in equilibrium6 with the entering wet solids and this leads to a partial pressure of
water vapor that is equivalent to 200 mm Hg. If the partial pressure of water vapor in the fresh air
entering the drier is equivalent to 10 mm Hg, what is the total molar flow rate of fresh air entering
the system? Assume that the drier operates at a pressure of one atmosphere.

dry air wet air

drier
dry solids, wet solids,
5% water 20% water

Figure 5.6a. Air drier

To construct a control volume for the analysis of this system, we need only make cuts where
information is given and required. These cuts are illustrated in Figure 5.6b and they lead to the
control volume shown in Figure 5.6c.

6 . This type of assumption is generally associated with what is known as an “equilibrium stage”.
124 Chapter 5

drier

Figure 5.6b. Cuts for the construction of a control volume

We begin this problem with the species macroscopic mole balance for a steady-state process
in the absence of chemical reaction and note that the species velocity, v A , can be replaced with
the mass average velocity, v, at entrances and exits to obtain

∫ c v ⋅ n dA
A
A = 0 (1)

It is important to remember that the molar concentration can be expressed as

cA = yA c , gas streams (2)

where y A is the mole fraction of species A and c is the total molar concentration. The form given

control
volume

dry air wet air


3 4

dry solids,
drier wet solids,
5% water 20% water
2 1

Figure 5.6c. Control volume for the analysis of the drier

by Eq. 2 is especially useful in the analysis of the air stream; however, for the wet solids stream it
is convenient to work in terms of mass rather than moles and make use of

ρA
cA = , wet solid streams (3)
MWA

If we let species A be water and apply Eq. 1 to the control volume shown in Figure 5.6c, we obtain
Gas-Liquid Systems 125

∫ (ρ
A1
H 2O )
MWH 2O v ⋅ n dA +
∫ (ρ
A2
H 2O )
MWH 2O v ⋅ n dA +




molar flow rate of water molar flow rate of water


entering with the solid leaving with the solid
Water: (4)

∫y
A3
H 2O c v ⋅ n dA +
A4
∫y H 2O c v ⋅ n dA = 0





molar flow rate of water molar flow rate of water


entering with the air leaving with the air

Here we have used Eqs. 2 and 3 in order to arrange the fluxes in forms that are convenient, but not
necessary, for this particular problem, and we can express those fluxes in terms of averaged
quantities to obtain

〈ρH 2O 〉1 Q1 〈ρH2 O 〉 2 Q2
Water: − + − (yH2 O )3 M 3 + (yH 2O ) 4 M 4 = 0 (5)
MWH 2O MWH2O

In this representation of the macroscopic mole balance for water, we have drawn upon the analysis
presented in Sec. 4.5. Specifically, we have imposed the following assumptions:
Gas streams: c v ⋅ n = constant (6a)

Wet solid streams: v ⋅ n = constant (6b)

in which “constant” means constant across the area of the entrances and exits. The three phases
contained in the wet solid streams are illustrated in Figure 5.6d for stream #1. The total density in
these streams consists of the density of the solid, the water, and the air, and this density can be
written explicitly as
〈ρ〉 = 〈ρsolid 〉 + 〈ρH 2O 〉 + 〈ρair 〉 (7)

The mass fraction of water in the wet solids is defined by

〈ρH 2O 〉
〈ωH 2O 〉 = (8)
〈ρ〉

and use of this representation in Eq. 5 leads to


〈ωH 2O 〉1 m 1 〈ωH 2O 〉 2 m 2
Water: − + − (yH2 O )3 M 3 + (yH 2O ) 4 M 4 = 0 (9)
MWH2 O MWH2 O

Here we have identified the mass flow rates of the wet solid streams according to

m 1 = 〈ρ〉1 Q1 , m 2 = 〈ρ〉 2 Q2 (10)

A little thought (see Problem 5-12) will indicate that a mass balance for the solid material leads to

Solid: 1 − 〈ωH O 〉1  m 1 = 1 − 〈ωH2 O 〉 2  m 2 (11)


 2 

and this result can be used in Eq. 7 to obtain

 (yH O ) 4   〈ωH 2O 〉 2 − 〈ωH2 O 〉1 


 
M 3 =  2  M 4 +   m 1 (12)
 (yH 2O )3   (yH2 O )3 1 − 〈ωH 2O 〉 2  MWH2 O 
126 Chapter 5

A molar balance for the air will allow us to eliminate M 4 from this result and the calculation of
M easily follows. This will be left as an exercise for the student (see Problem 5-13).
3

drier

water air
solid

Figure 5.6d. Wet solids entering the drier

5.5 Saturation, Dew Point and Bubble Point of Liquid Mixtures


When a pure component A in the liquid phase is in equilibrium with a gas, the vapor pressure of the
component in the gas phase is equal to the vapor pressure of the pure component.

pA = p A, vap (5-34)

This result is obtained by letting x A = 1 in Eq. 5-23 so that the liquid is pure component A. In general,
when air is the gas phase we will assume that the vapor pressure in the gas phase adjacent to the liquid is
saturated with the liquid component and that the concentration of the air in the liquid is negligible. When
the liquid phase is a mixture, Raoult’s law (Eq. 5-23) must be used to compute the composition of the gas
phase. If a liquid mixture is in equilibrium with its own vapors, then the overall pressure is equal to the
sum of the vapor pressures of the individual components.
A= N
p = ∑p
A=1
A, vap xA (5-35)

When a liquid mixture is heated, the vapor pressure of the components in the mixture increases and the sum
of the partial pressures, given by Eq. 5-35, increases accordingly. If the liquid mixture is in contact with air
at atmospheric pressure, the partial pressure of the components of the mixture is also given by Eq. 5-35.
When the sum of the partial pressures of the components of the mixture is equal to the atmospheric
pressure, the liquid mixture boils. The difference between a liquid mixture and a pure liquid is that the
boiling temperature of a mixture is not constant. For a mixture in equilibrium with its own vapors, the
bubble point of a mixture is the pressure at which the liquid starts to vaporize. Similarly, for a vapor
mixture the dew point of the mixture is the pressure at which the vapors start to condense. The use of the
terms has been extended when the liquid is in contact with air, i.e. it is customary to call bubble point at the
temperature at which the liquid mixture starts boiling and dew point the temperature at which the first
condensed liquid appears.
Gas-Liquid Systems 127

EXAMPLE 5.7. Bubble point of a water-alcohol mixture.


A mixture of ethanol, xEt = 0.5 , and water is slowly heated under well-stirred conditions in
an open beaker. Using Antoine’s equation to determine the vapor pressure of the components and
Raoult’s Law to estimate the partial pressures, we can estimate the bubble point of this mixture,
i.e. the temperature at which the first bubbles will start forming at the bottom of the beaker as well
as the composition of the first bubbles.
The vapor pressure of pure ethanol and water can be computed using Antoine’s equation. The
partial pressure of the components in the gas phase in equilibrium with the liquid mixture are
computed using Raoult’s Law given by Eq. 5-25. The bubble point will be determined as the
temperature at which the sum of the partial pressures of ethanol and water is equal to atmospheric
pressure.

p H 2O + p Et − p = x H 2O p H 2O,vap + x Et p Et,vap − 760 mmHg = 0 (1)

This problem, in principle, can be solved by substitution of Antoine’s equation for the vapor
pressures of the pure components in Eq. 1, and then solving for the temperature. A much simpler
route consists of guessing values of the temperature until we satisfy Eq. 1. This procedure is
easily done using a spreadsheet where one of the columns are values of the temperature and the
columns to the right are values of the partial pressure of the components and the residue of Eq. 1.
The values of Antoine’s coefficient for the components, from Appendix II are:
Water: A = 7.94915, B = 1657.46, C = 227.03

Ethanol: A = 8.1629, B = 1623.22, C = 227.03

and computed values of the vapor pressure are given in the following table:

Computation of dew point of Ethanol and water mixture


Temp p H 2O p Eth Residue
Degrees C mmHg mmHg mmHg
60 74.7483588 175.7127 −509.539
70 116.9527 270.8179 −372.229
80 177.72766 405.8736 −176.399
86 225.542371 511.0501 −23.4076
86.8 232.662087 526.6464 −0.6915
86.9 233.565109 528.6235 2.188576
87 234.471058 530.6067 5.077746
90 263.047594 593.0441 96.09172

We could continue the computation by inserting additional rows between the temperatures
T = 86.8 C and T = 86.9 C . However, for the purpose of this example we will accept the boiling
point of the mixture as T = 86.85 ± 0.05 C .

5.5.1 Humidity
In air-water mixtures the humidity is often used as the measure of concentration. The humidity is
defined by
mass of water
humidity = (5-36)
mass of dry air

It will be left as an exercise for the student to show that


128 Chapter 5

MWH O pH O
humidity = 2 2
(5-37)
(
MWair p − pH O
2
)
where p is the total pressure and pH O is the partial pressure of the water vapor. This result can be derived
2

from the definition given by Eq. 5-36 only if the air-water mixture is treated as an ideal gas. The percent
relative humidity is often used as a measure of concentration since our personal comfort may be closely
connected to this quantity. It is defined by

pH O
% relative humidity = 2
× 100 (5-38)
pHo O
2

where pHo O is the vapor pressure of water. When the percent relative humidity is 100% the air is
2

completely saturated and the addition of further water will result in condensation. Values of the vapor
pressure of water are listed in Table 5-2.

5.5.2 Modified mole fraction


In general, the most useful measures of concentration are the molar concentration c A and the species
density ρ A . Associated with these concentrations are the mole fraction defined by Eq. 5-1f and the mass
fraction defined by Eq. 5-1c. Sometimes it is convenient to use a modified mole fraction or mole ratio

Table 5-2. Vapor Pressure of Water as a Function of Temperature

T, C Vapor Pressure, mm Hg T, F Vapor Pressure, in. Hg


0 4.579 32 0.180
5 6.543 40 0.248
10 9.209 50 0.363
15 12.788 60 0.522
20 17.535 70 0.739
25 23.756 80 1.032
30 31.824 90 1.422
35 42.175 100 1.932
40 55.324 110 2.596
45 71.88 120 3.446
50 92.51 130 4.525
55 118.04 140 5.881
60 149.38 150 7.569
65 187.54 160 9.652
70 233.7 170 12.199
75 289.1 180 15.291
80 355.1 190 19.014
85 433.6 200 23.467
90 525.76 212 29.922
95 633.90 220 34.992
100 760.00 230 42.308
105 906.07 240 50.837
110 1074.56 250 60.725
115 1267.98 260 72.134
120 1489.14 270 85.225
125 1740.93 280 100.18
130 2026.16 290 117.19
135 2347.26 300 136.44
Gas-Liquid Systems 129

which is based on all the species except one. If we identify that one species as species N, we express the
modified mole fraction as
B = N −1
XA = cA ∑c B =1
B (5-39)

When it is convenient to work in terms of this modified mole fraction, one usually needs to be able to
convert from x A to X A and it will be left as an exercise for the student to show that this relation is given
by

XA = x A (1 − xN ) , A = 1, 2,...N (5-40)

In some types of analysis it is convenient to choose species N to be species A. Under those circumstances
we need to express Eq. 5-39 as
B= N
XD = cD ∑c B =1
B (5-41)

B≠ A

and Eq. 5-39 takes the form

XD = xD (1 − x A ) , D = 1, 2,...N (5-42)

Similar relations can be developed for the modified mass fraction Ω A and they will be left as exercises for
the student. It is important to note that the sum over all species of the modified mass or mole fraction is not
one.

EXAMPLE 5.8 Condensation of water in humid air


On a warm spring day in Baton Rouge, LA, the atmospheric pressure is 755 mm Hg, the
temperature is 80 F, and the humidity is 80%. A large industrial air conditioner treats 1000 kg/hr
of air (dry air basis) and lowers the air temperature to 15 C. How much liquid water, in kg/h, is
removed from the air by the air conditioning system? It is convenient to transform all the units to
a consistent set of units, such as the SI system of units. In this case the transformation is given by

 755 mm Hg   101,300 Pa 
p = 755 mm Hg =    = 100,634 Pa
 760 mm Hg/atm   atm 

T = (80 F − 32 F ) × ( 5C 9 F ) = 26.7 C = 299.8 K

From Table 5-2 we find the saturation vapor pressure of water at 80 F and at 15 C to be given by

Saturation pressure at 80 F: (1.032 in Hg) (25.4 mm/in) (101,300 Pa/760 mm Hg) = 3,494 Pa

Saturation Pressure at 15 C: (12.788 mm Hg) (101,300 Pa/760 mm Hg) = 1705 Pa

The vapor pressure of water in air at 15 C is assumed to be at saturation, i.e. at the bubble point of
a mixture of air and water. This is a standard assumption during condensation and it is based on
the assumption of local thermodynamic equilibrium. The vapor pressure of water at 80 F is only
80% of the value at saturation and this leads to
Vapor pressure of water at 80 F: 0.80 × ( 3494 Pa ) = 2795 Pa

We use these values to compute the humidity at the input and output conditions:
130 Chapter 5

M H 2O pH 2O
humidity (80 F) = =
(
M air p − pH 2O )
18.015 g H 2 O / mol × 2, 795 Pa g H 2O
= = 0.0179
28.84 g air/mol × (100,634 − 2,795) Pa g air

18.015 g H 2O × 1, 705 Pa g H 2O
humidity (15C) = = 0.0108
28.84 g air × (100,634 − 1, 705) Pa g air

The difference between input humidity and output humidity gives the amount of grams of water
condensed per gram of dry air.
g H 2O g H 2O g water condensed
0.0179 − 0.0108 = 0.0071
g air g air g dry air

Multiplying the grams of water condensed per gram of dry air by the load of the air conditioning
unit in kilograms of dry air per unit time, we obtain the amount of water condensed per unit time:
g water condensed kg dry air kg water condensed
0.0071 × 1000 = 7.1
g dry air hour hour

EXAMPLE 5.9. Humid air flow


Humid air exits a dryer at atmospheric pressure, 75 C, 25% relative humidity, and at a
volumetric flow rate of 100 m3/min. In this example we wish to determine:
a) Absolute humidity of the air in kg water/kg air.
b) Molar flow rates of water and dry air.

The vapor pressure of water at 75 C is found in Table 5-2 to be pwo = 289.1 mm Hg . The density
of mercury is found in Table I of the Appendix. We convert all parameters into SI units according
to
pw = ( 0.25 )
289.1mmHg
1000 mm/m
( )(
9.81m/s 2 13546 kg/m3 = 9605 Pa )
and we use Eq. 5-26 to compute the molar fraction of water in the air as

pw 9605 Pa
yw = = = 0.095
p 101,300 Pa

In order to determine the absolute humidity, we use Eq. 5-36

mass of water mass of water/volume


humidity = =
mass of dry air mass of dry air/volume

and this leads to an expression for the humidity given by

MWw yw MWw yw
humidity = =
MWa ya MWa (1 − yw )

(18.05 kg water/kmol )( 0.095)


= = 0.066 kg water / kg air
( 28.84 kg air/kmol ) (1 − 0.095)
Gas-Liquid Systems 131

Assuming that water vapor and air behave as ideal gases at atmospheric pressure, we use the ideal
gas law given by Eq. 5-3 to compute the total concentration of the mixture. The concentration of
the gas mixture is the total number of moles of air and water per unit volume of the mixture. This
can be expressed as

n p 101,300 Pa
c = = = = 35 mol/m3 (1)
V RT m3 Pa
8.314 × 348.16 K
mol K

This result gives the total number of moles of gas per unit volume of mixture. If one considers a
material volume of a flowing gas mixture, and the gases satisfy the ideal gas law, the
concentration of the flowing mixture in the material volume should be equal to the concentration
computed using Eq. 1. In order to determine the molar flow rates of water and dry air, we note
that

M w = cwQ = yw cQ = 0.095 × 35 mol/m3 × 100 m3 /min


(2a)
= 332.5 mol water/min

M a = ca Q = ya cQ = (1 − yw ) c Q = 3167.5 mol air /min (2b)

5.6 Staged Processes


In Sec. 5.3 we examined several systems that consisted of a single contacting process in which
equilibrium conditions were assumed to exist at the exit streams. Knowing when the condition of
equilibrium is a reasonable approximation requires a detailed study of the heat and mass transfer processes
taking place. These details will be studied in subsequent courses where it will be shown that the condition
of equilibrium is a reasonable approximation for many mass transfer processes.

*
5.7 Problems

Section 5.1
5-1. Show that the mole fraction in an ideal gas mixture can be expressed as y A = p A p .

5-2. Assuming ideal gas behavior, determine the average molecular mass of a mixture made of equal
amounts of mass of chlorine, argon, and ammonia.

Section 5.2
5-3. A liquid mixture of hydrocarbons has 40% by weight of cyclohexane, 40% of benzene, and 20%
toluene. Assuming that volumes are additive compute the following:
(a) partial densities of components in the mixture.
(b) overall density of the mixture
(c) concentration of components in moles/m3
(d) molar fractions of components in the mixture.

*
Problems marked with the symbol  will be difficult to solve without the use of computer software.
132 Chapter 5

Section 5.3
5-4. Determine the vapor pressure, in Pascal, of ethyl ether at 25 C and at 30 C. Estimate the heat of
vaporization of ethyl ether using these two vapor pressures and the Clausius-Clapeyron equation.

5-5. Determine the vapor pressure of methanol at 25 C and compare it to that of ethanol at the same
temperature. Consider the ethanol-methanol system to be an ideal solution in the liquid phase and an ideal
gas mixture in the vapor phase. Determine the mole fraction of methanol in the vapor phase when the
liquid phase mole fraction is 0.50. If the liquid phase is allowed to slowly evaporate, will it become richer
in methanol or ethanol? Here you are asked to provide an intuitive answer concerning the composition of
the liquid phase during the process of distillation. In Chapter 8 a precise analysis of the process will be
presented.

5-6. Determine the vapor-liquid equilibrium curve of a binary mixture of acetone and benzene. Plot the
mole fraction of acetone in the vapor phase versus the mole fraction of acetone in the liquid phase at one
atmosphere (760 mm Hg).

5-7. Use Eqs. 5-25 and 5-26 order to derive 5-27.

5-8. Demonstrate that Eq. 5-30 is valid for an ideal system containing three components, and think about
replacing the constraint x A << 1 with something more appropriate.

Section 5.4
5-9. Consider air to consist of nitrogen and oxygen and indicate under what circumstances the mole
balances for these two components can be added to obtain the special form

A
∫c air v ⋅ n dA = 0

that was used in Example 5.5. Here the molar concentration of air is defined by cair = cN 2 + cO2 .

5-10. In Example 5.5 we used a control volume located outside of the scrubber. This control volume is
illustrated in Figure 5.5c, and in this problem you are asked to construct a suitable control volume that lies
insider the scrubber.

5-11. In Example 5.5 we used a macroscopic mole balance to determine the mole fraction of SO2 in the
exit water stream. In actual fact the exit water stream contains two molecular species that involve sulfur
dioxide, i.e. SO2 and H 2SO3 . This occurs because of the reaction

SO 2 + H 2 O R H 2SO3

A more precise analysis of this problem would make use of a total macroscopic mole balance for SO2 and
H 2SO3 . Repeat the analysis given in Example 5.5 for the molar concentration of species C where

cC = c A + cB

Here c A represents the concentration of SO2 and cB represents the concentration of H 2SO3 . Be careful
to consider the reaction rate term in the species mole balance and comment on your treatment of the species
velocities, v A and v B .

5-12. Derive the form of the solid phase mass balance given by Eq. 8 in Example 5.6.

5-13. Complete the analysis in Example 5.6 in order to determine the total molar flow rate of fresh air
entering the drier.
Gas-Liquid Systems 133

5-14. A gas mixture leaves a solvent recovery unit as illustrated in Figure 5.14. The partial pressure of
benzene in this stream is 80 mm Hg and the total pressure is 750 mm Hg. The volumetric analysis of the
gas, on a benzene-free basis, is 15% CO 2 , 4% O 2 and the remainder is nitrogen. This gas is compressed
to 5 atm and cooled to 100 F. Calculate the percentage of benzene condensed in the process. Assume that
CO 2 , O 2 and N 2 are insoluble in benzene, thus the liquid phase is pure benzene.

Figure 5.14. Recovery-condenser system

5-15. Small amounts of an inorganic salt contained in an organic fluid stream can be removed by contacting
the stream with pure water as illustrated in Figure 5.15. The process requires that the organic and aqueous
streams be contacted in a mixer that provides a large surface area for mass transfer, and then separated in a
settler. If the mixer is efficient, the two phases will be in equilibrium as they leave the settler and you are
to assume that this is the case for this problem. You are given the following information:
a) Organic stream flow rate: 1000 lbm/min
b) Specific gravity of the organic fluid: ρorg / ρH2 O = 0.87
c) Salt concentration in the organic stream entering the mixer: (c A )org = 0.0005 moles/liter
d) Equilibrium relation for the inorganic salt: (c A ) aq = K eq (c A )org where K eq = 60

Here (c A ) aq represents the salt concentration in the aqueous phase that is in equilibrium with the salt
concentration in the organic phase, (c A )org . In this problem you are asked to determine the mass flow rates
of the water stream that will reduce the salt concentration in the organic stream to 0.1, 0.01 and 0.001 times
the original salt concentration. The aqueous and organic phases are to be considered completely
immiscible, i.e., only salt is transferred between the two phases. In addition, the amount of material
transferred is so small that the volumetric flow rates of the two streams can be considered constant.
134 Chapter 5

Figure 5.15. Liquid-liquid extraction

5-16. In this problem, we examine the process of recovering fission materials from spent nuclear fuel rods.
This is usually referred to as reprocessing of the fuel to recover plutonium, Pu, and the active isotope of
uranium, U235. Reprocessing can be done by separation of the soluble isotope nitrates from a solution in
nitric acid by a solvent such as a 30% solution of tributyl phosphate (TBP) in dodecane in which the
nitrates are preferentially soluble. Industrial reprocessing of nuclear fuels is done by countercurrent
operation of many liquid-liquid separation stages. These separation stages consist of well-mixed contacting
tanks where UO 2 (NO3 ) 2 is exchanged between two immiscible liquid phases, and separation tanks where
the organic and aqueous phases are separated. A schematic of a separation stage is shown in Figure 5.16a.

Figure 5.16a. Liquid-liquid separation stage for reprocessing

In this process an aqueous solution of uranil nitrate, UO 2 (NO3 ) 2 is one of the feed streams to the
separation stage, and the mass flow rate of the aqueous feed phase is, m 1 = 400 kg/hr . The second feed
stream is an organic solution of TBP in dodecane, which we assume to be a single component. The organic
and inorganic phases are assumed to be completely immiscible, thus only the uranil nitrate is transferred
from one stream to the other. The process specifications are indicated in Figure 5.16b, and for this problem
it is the mass flow rates that we wish to determine.
Gas-Liquid Systems 135

Figure 5.16b. Specified stream variables

5-17. The concept of an equilibrium stage is a very useful tool for the design of multi-component
separations, and a typical equilibrium stage for a distillation column is shown in Figure 5.17. A liquid
stream, S1, flowing downward encounters a vapor stream, S2, flowing upward. We assume that the vapor

Figure 5.17. Sketch of an equilibrium stage process

and liquid streams exchange mass inside the equilibrium stage until they are in equilibrium with each other.
Equilibrium is determined by a ratio of the molar fractions of each component in the liquid and vapor
streams according to
y A,4
KA = A = 1, 2, .., N
x A,3

The streams leaving the stage, S3 (liquid), and S4 (vapor) are in equilibrium with each other and therefore
satisfy the above relation. The ratio of the molar flow rates of the output streams is a function of the energy
balance within the stage. In this problem we assume that the ratio of the liquid output molar flow rate to
the vapor output molar flow rate, M 3 / M 4 , is given. Assuming that the compositions, i.e. the mole
fractions of the components and the molar flow rates of the input streams, S1 and S2, are known, and the
equilibrium constant for one of the components is given, develop the mass balances for a two component
vapor-liquid equilibrium stage.

5-18. A single stage, binary distillation process is illustrated in Figure 5.18. The total molar flow rate
entering the unit is M 1 and the mole fraction of species A in this liquid stream is ( x A )1 . Heat is supplied in
order to generate a vapor stream, and the ratio, M / M = β , depends on the rate at which heat is supplied.
2 3
At the vapor-liquid interface, we can assume local thermodynamic equilibrium in order to express the
vapor-phase mole fraction in terms of the liquid-phase mole fraction according to
136 Chapter 5

α AB x A
yA = , at the vapor-liquid interface
1 + x A (α AB − 1)

Here α AB represents the relative volatility. If the distillation process is slow enough, one can assume that
the vapor and the liquid leaving the distillation unit are in equilibrium; however, at this point in our studies
we do not know what is meant by slow enough. In order to proceed with an approximate solution to this
problem, we replace the equilibrium relation with a process equilibrium relation given by
α AB ( x A )3
( y A )2 = , process equilibrium relation
1 + ( x A )3 (α AB − 1)

Given a detailed study of mass transfer in a subsequent course, one can make a judgment concerning the
conditions that are required in order that this process equilibrium relation be satisfactory. For the present,
you are asked to use the above relation to derive an implicit expression for ( y A ) 2 in terms of ( x A )1 and
examine three special cases: α AB → 0 , β → 0 , α AB = 1 .

Figure 5.18. Single stage binary distillation

5-19. A saturated solution of calcium hydroxide enters a boiler as shown in Figure 5.19 and a fraction, ϕ, of
the water entering the boiler is vaporized. Under these circumstances a portion of the calcium hydroxide
precipitates and you would like to know the mass fraction of this suspended solid calcium hydroxide in the
liquid stream leaving the boiler. The solubility can be expressed as

solubility = S = g of Ca(OH) 2 g of H 2 O

Assume that no calcium hydroxide leaves in the vapor stream, that none accumulates in the boiler, and that
the temperature of the liquid entering and leaving the boiler is a constant. Develop a general solution for
the mass fraction of the suspended solid in the liquid stream leaving the boiler in terms of ϕ and S. For ϕ =
0.50, 0.21, and 0.075, determine the mass fraction of suspended solid when S = 2.5 × 10−3 .

Figure 5.19. Precipitation of calcium hydroxide in a boiler


Gas-Liquid Systems 137

Section 5.5
5-20. An equi-molar mixture of ethanol and ethyl ether is kept in a closed container at 103 KPa and 95 C.
The temperature of the container is slowly reduced to the dew point of the mixture. Determine:
(a) What is the dew point temperature of the mixture?
(b) What is the pressure of the container at the dew point temperature of the mixture?
(c) What is the composition of the first drop of liquid at the dew point?

5-21. A liquid mixture of n-hexane (mole fraction equal to 0.32) and n-heptane is heated until it begins
boiling. Find the bubble point at p = 760 mm Hg. What are the molar fractions of the vapor when the
mixture starts boiling?

5-22. A vapor mixture of benzene and toluene is slowly cooled inside a constant volume vessel. Initially
the pressure inside the vessel is 300 mm Hg and the temperature is 70 C. As the vessel is cooled, the
pressure inside the vessel decreases. Assume the vapor behaves like an ideal gas and take the dew point of
the mixture to be 60 C. What is the mole fraction of benzene in the initial vapor?

5-23. Consider a day when the percent relative humidity is 70%, the temperature is 80 F and the barometric
pressure is 1 atm. What is the humidity, mole fraction of water in the air, and dew point of the air?

5-24. A mole of air is sampled from the atmosphere when the atmospheric pressure is 765 mm Hg, the
temperature is 25 C, and relative humidity is 75%. The sample of air is placed inside a closed container
and heated to 135 C and then compressed to 2 atm. What are the relative humidity, the humidity, and the
mole fraction of water in the air?

5-25. A humidifier is used to introduce moisture into air supplied to an office building during winter days.
Outside air at atmospheric pressure and 5 C is introduced into the heating system at a rate of 100 m3/min,
on a dry air basis. The relative humidity of the outside air is 95%, and the heating system delivers warm air
into the building at 20 C. How much water must be introduced into the warm air, in kg/min, the keep the
relative humidity inside the building at 75%?

Section 5.6
5-26.

5-27.

You might also like