ICH 2101 Mass Transfer 1 Modules
ICH 2101 Mass Transfer 1 Modules
ICH 2101 Mass Transfer 1 Modules
Course Objectives:
Learning outcomes
Recognise and explain mass transfer concepts and terminology for fluids in laminar and
turbulent flow.
Solve the problems associated to with mass transfer between liquid and gas phases.
Analyse and solve adsorption problems using solute movement analysis for linear
systems.
Solve the problems associated to with mass transfer between solid and fluid phases.
1
Detailed Course description:
Equilibrium, Diffusion between phases, Local Two-Phase Mass Transfer, Local Overall Mass-
transfer Coefficients, Materials Balances, Steady-State cocurrent Processes, Steady-State
Countercurrent Processes, Stages, Cascades, Cross-Flow Cascades, Countercurrent Cascades.
(Quiz)
(Quiz/test 1); Topic to be demonstrated using Aspen plus software/in a unit operation
demonstrations or industrial visits.
(Quiz).
Sparged vessels, mechanically agitated vessels, Tray towers, Bubble-cap trays, sieve trays.
Design of tray towers, weirs, downspouts, perforated plates, Packed towers packings.
(Presentations/test 1)
2
Mode of Delivery: Lectures/Hand outs (Interactive lectures/discussion), workshops,
assignments (personal assessment and it will be from these assignments that I will get to know
which areas u didn’t understand), tutorials (this will be more of interactive problem solving
session i.e we expect this to be more interactive and specific than formal class lectures and this
will be picked from assignments and will be conducted in form of group work) and tests.
Student Assessment
References
Heat and mass transfer, second edition by P K Nag (The McGraw-Hill companies)
Kochi Asano, “Mass transfer, from fundamental to modern industrial application”.
Hans Dieter Baehr, Karl Stephan, (2006),“Heat and Mass Transfer” 2nd Edition.
Theodore L. Bergman, Adrienne S. Lavine, Frank P. Incropera, David P. DeWitt ,
“Fundamentals of Heat and Mass Transfer” 7th Edition.
Binay K. Dutta, (2007), “Principles of Mass Transfer and Separation Processes”.
Christie John Geankoplis, “Transport Processes and Separation Process Principles (Includes
Unit Operations)” 4th Edition.
3
1.0 INTRODUCTION TO MASS TRANSPORT
Mass transfer is a transport of components under a chemical potential gradient. The component
moves to the direction of reducing concentration gradient. The transport occurs from a region
of higher concentration to lower concentration. Equilibrium is reached when the gradient is
zero. The transport or migration of one constituent from a region of higher concentration to
that of a lower concentration is known as mass transfer.
The transfer of mass within a fluid mixture or across a phase boundary is a process that plays a
major role in many industrial processes. Examples of such processes are:
2. Unit Operations
Unit operation is defined as a unified study of a particular separation technique used for a
chemical engineering application. Distillation is an example in which the vapor pressure
differences between two components are used to separate and purify these components
Common Mass Transfer–Based Separation Processes and the Phases Being Contacted
4
Wet solid Gas Drying
3. Product Development and Product Engineering
Transport phenomena are increasingly being exploited in product development. Example
applications include the design of drug capsules that should provide a constant release rate and
the design of polymer wrapping in food packaging to reduce oxygen diffusion.
4. Biomedical Applications
The focus of mass transfer analysis in the field of biomedical engineering is to bring together
fundamentals of transport models and life sciences principles. Key areas where mass transport
phenomena can be utilized include the following:
6. Catalytic Converter
A common example of mass transfer accompanied by reaction is found in an automobile. A
catalytic converter consists of a set of flow channels coated with an active layer of catalyst
such as platinum (Pt). This is an example of a flow system accompanied by mass transfer and
chemical reaction to reduce the release of pollutants such as CO and NOx.
7. Reacting Systems
Heterogeneous reactions are commonly encountered in chemical processing. These systems
include two or more phases, with reaction taking place mainly in one of the phases while the
reactants are usually present in the other phase
5
reactant
Many of air day-by-day experiences also involve mass transfer, for example:
(i) A lump of sugar added to a cup of coffee eventually dissolves and then eventually
diffuses to make the concentration uniform.
(ii) Water evaporates from ponds to increase the humidity of passing-air-stream
(iii) Perfumes presents a pleasant fragrance which is imparted throughout the
surrounding atmosphere.
The mechanism of mass transfer involves both molecular diffusion and convection.
Mass transfer operation plays an important role in many industrial processes. A group of
operations for separating the components of mixtures is based on the transfer of material from
one homogeneous phase to another. These methods are covered by the term mass transfer
operations which include techniques like gas absorption and stripping, liquid-liquid extraction,
leaching, distillation, humidification, drying, crystallization and a number of other separation
techniques.
6
situation, both the fluids (absorbent and absorbing fluid) are reprocessed and/or reused for the
operation.
Adsorption and desorption –Adsorption is the transfer of mass from either a gas or a
liquid to the surface of a solid. Desorption involves the transfer of mass from the solid surface
(adsorbents) to the gas or liquid medium (adsorbates). A few operations of practical
applications are; elimination of toxic gases and deodorization of air, recovery of solvents,
removal of ions from solution, as in demineralization of water and many other applications.
Extraction- It refers to a separation of the constituents of a liquid solution by contact
with another insoluble liquid. The liquid which is added to the solution to bring about the
extraction is known as the solvent. The solution which is to be extracted is called the feed. The
solvent-richen product of the operation is called the extract and the residual liquid from which
the solute is separated is called the raffinate. The separation of aromatics from kerosene based
fuel oils, the production of fuels in the nuclear industry and the separation of penicillin from
fermentation mixtures are examples of techniques of extraction.
Leaching-is a treatment of a finely divided solid with a liquid. Some examples of
leaching operations are oilseed extraction, extraction of sugar beets with hot water and
extraction of medicinal compounds from plant roots, leaves and stems.
Distillation-It is an operation whereby a liquid mixture of miscible and volatile
substances is separated into individual components or into group of components by partial
vaporization. The separation of a mixture of methanol and water into its components, and
separation of petroleum crude into gasoline, kerosene, fuel oil and lubricating stock are
examples of distillation process.
Humidification and dehumidification- In humidification, the enrichment of vapour
content in a gas stream takes place by passing the gas over a liquid. The transfer of water from
the liquid phase to the gas phase of a mixture of air and water vapour is a widespread
application of humidification. Dehumidification involves the transfer of water vapour from the
gas phase to the liquid phase. Removal of water vapour from air by condensation of a cold
surface and condensation of carbon tetrachloride out of a stream of nitrogen are examples of
dehumidification.
Drying and Evaporation- Drying refers to removal of relatively small amount of water
or other liquid from a solid material whereas evaporation refers to the removal of relatively
large amount of water from solutions. In evaporation the water is removed as vapour at its
7
boiling point. Drying involves the removal of water at temperature below the boiling point by
circulating air or some other carrier gas over the material.
8
1.2 Properties of Mixtures
Mass transfer always involves mixtures. Consequently, we must account for the variation of
physical properties which normally exist in a given system. When a system contains three or
more components, as many industrial fluid streams do, the problem becomes unwidely very
quickly. The conventional engineering approach to problems of multicomponent system is to
attempt to reduce them to representative binary (i.e., two component) systems.
In order to understand the future discussions, let us first consider definitions and relations,
which are often used to explain the role of components within a mixture.
1.2.1 Concentration of Species:
9
nA pA
CA = =
V RT
where pA is the partial pressure of species A in the mixture. V is the volume of gas, T is the
absolute temperature, and R is the universal gas constant.
The total molar concentration or molar density of the mixture is given by
C = ∑ Ci
i
Velocities
In a multicomponent system the various species will normally move at different velocities; and
evaluation of velocity of mixture requires the averaging of the velocities of each species
present.
If I is the velocity of species i with respect to stationary fixed coordinates, then mass-average
velocity for a multicomponent mixture defined in terms of mass concentration is,
∑ ρi νi ∑ ρiν i
i i
ν= =
∑ ρi ρ
i
10
∑yi = 1
i
O2 7%
CO 10%
CO 2 15%
N2 68%
Determine
O2 = 0.07 mol
CO = 0.10 mol
CO 2 = 0.15 mol
N2 = 0.68 mol
O2 = 2 * 16 = 32 g/mol
CO = 12 + 16 = 28 g/mol
CO 2 = 12 + 2 * 16 = 44 g/mol
11
N2 = 2 * 14 = 28 g/mol
O2 = 0.07 * 32 = 2.24 g
CO = 0.10 * 28 = 2.80 g
CO 2 = 0.15 * 44 = 6.60 g
N2 = 0.68 * 28 = 19.04 g
= 30.68 g
2 .24
O 2= ∗100=7. 30 %
30. 68
2 . 80
CO= ∗100=9 .13 %
30 .68
6 .60
CO 2 = ∗100=21 .51 %
30 . 68
19 .04
N 2= ∗100=62 .06 %
30 .68
30 . 68
M= =30 . 68 g /mol
1
PV = nRT
n P
=
V RT
12
n
=molar density = ρ m
V
5
PM 1. 5∗10 ∗30 .68
Density = ρ m M= = kg /m 3
RT 8314∗273
= 2.03 kg/m 3
7
= ∗( 1.5∗10 5 )
100
= 0.07 * 1.5 * 10 5
= 0.105 * 10 5 Pa
Diffusion flux
Just as momentum and energy (heat) transfer have two mechanisms for transport-molecular and
convective, so does mass transfer. However, there are convective fluxes in mass transfer, even on a
molecular level. The reason for this is that in mass transfer, whenever there is a driving force, there is
always a net movement of the mass of a particular species which results in a bulk motion of molecules.
Of course, there can also be convective mass transport due to macroscopic fluid motion. In this chapter
the focus is on molecular mass transfer.
The mass (or molar) flux of a given species is a vector quantity denoting the amount of the particular
species, in either mass or molar units, that passes per given increment of time through a unit area
normal to the vector. The flux of species defined with reference to fixed spatial coordinates, N A is
N A =C A ν A ---------------------- (1)
This could be written in terms of diffusion velocity of A, (i.e., A - ) and average velocity of
mixture, , as
13
N A = C A (ν A − ν)+ C A ν --------------- (2)
By definition
∑ Ci νi
i
ν = ν∗ =
C
C
N A =C A (ν A − ν) +
C
A
∑C i ν i
i
= C A (ν A − ν ) + y A ∑ C i ν i
i
N A = C A (ν A − ν ) + y A (C A ν A + C B ν B )
= C A ( ν A − ν ) + y A (N A + N B )
N A = C A (ν A − ν ) + y A N ----------- (3)
The first term on the right-hand side of this equation is diffusional molar flux of A, and the second
term is flux due to bulk motion.
Fick’s law
An empirical relation for the diffusional molar flux, first postulated by Fick and, accordingly, often
referred to as Fick’s first law, defines the diffusion of component A in an isothermal, isobaric system.
For diffusion in only the Z direction, the Fick’s rate equation is
dCA
JA = − D A B
dZ
where D AB is diffusivity or diffusion coefficient for component A diffusing through component B, and
dCA/dZ is the concentration gradient in the Z-direction.
A more general flux relation, which is not restricted to isothermal, isobaric system could be written as
14
d yA
JA = − C DA B
dZ ----------------- (4)
d yA
N A = − C DA B + yAN
dZ --------------- (5)
NA=J A + yAN
or J A =N A − y A N ----------------------- (6)
Similarly,
J B =N B − y B N -------------------- (7)
J A + J B = N A + N B −( y A + y B) N ---------- (8)
By definition N = N A + N B and y A + y B = 1.
Therefore equation (8) becomes,
JA+JB=0
J A = -J B
d yA d yB
C D AB =−C D BA --------------- (9)
dz dZ
From yA+yB=1
Differentiate with respect to z
d y A −d y B
=
dz dZ
15
This leads to the conclusion that diffusivity of A in B is equal to diffusivity of B in A.
Diffusivity
Fick’s law proportionality, DAB, is known as mass diffusivity (simply as diffusivity) or as the
diffusion coefficient. DAB has the dimension of L 2 / t, identical to the fundamental dimensions
of the other transport properties: Kinematic viscosity, = / in momentum transfer, and
thermal diffusivity, = k / C in heat transfer.
Diffusivity is normally reported in cm2 / sec; the SI unit being m2/sec.
Diffusivity depends on pressure, temperature, and composition of the system.
In table below, some values of DAB are given for a few gas, liquid, and solid systems.
Diffusivities of gases at low density are almost composition independent, increase with the
temperature and vary inversely with pressure. Liquid and solid diffusivities are strongly
concentration dependent and increase with temperature.
In the absence of experimental data, semi-theoretical expressions have been developed which
give approximation, sometimes as valid as experimental values, due to the difficulties
encountered in experimental measurements.
Diffusivity in Gases:
1
D AB ∝
p (for moderate ranges of pressures, upto 25 atm).
And temperature dependency is according to
16
3
2
D AB ∝ T
1
D 1−mixture = ' ' '
y2 y y
+ 3 + .. .. . .. .. . .+ n
D1−2 D1−3 D1−n
Where D 1-mixture is the diffusivity for component 1 in the gas mixture; D 1-n is the diffusivity for
'
the binary pair, component 1 diffusing through component n; and y n is the mole fraction of
component n in the gas mixture evaluated on a component –1 – free basis, that is
'
y2
y2 =
y 2 + y 3 + .. .. ... y n
9. Determine the diffusivity of Co 2 (1), O 2 (2) and N 2 (3) in a gas mixture having the composition:
The gas mixture is at 273 K and 1.2 * 10 5 Pa. The binary diffusivity values are given as: (at 273 K)
D 12 P = 1.874 m 2 Pa/sec
D 13 P = 1.945 m 2 Pa/sec
D 23 P = 1.834 m 2 Pa/sec
Calculations:
Diffusivity of Co 2 in mixture
17
1
D 1 m=
y y
2′ 3′
+
D 12 D 13
y2 0. 15
y ′= = =0. 21
where
2 y 2+ y 3 0 . 15+0 .565
y3 0 .565
y ′= = =0 .79
3 y 2+ y 3 0 . 15+0 . 565
1
D 1 m P=
0 .21 0 .79
+
Therefore 1. 874 1. 945
= 1.93 m 2.Pa/sec
1 . 93 −5 2
D 1 m= =1. 61∗10 m / sec
1 .2∗10 5
1
D 2 m=
y y
1′ 3′
+
D 21 D 23
y1 0 . 285
y ′= = =0. 335
Where
1 y 1 + y 3 0 . 285+0 .565
18
and
y3 0. 565
y ′= = =0 .665
3 y 1 + y 3 0 . 285+0 .565
and
D 21 P = D 12 P = 1.874 m 2.Pa/sec
Therefore
1
D 2 m P=
0 .335 0 .665
+
1. 874 1. 834
= 1.847 m 2.Pa/sec
1 .847
D 2m = 5
= 1 .539∗10 −5 m 2 / sec
1 .2∗10
By Similar calculations Diffusivity of N 2 in the mixture can be calculated, and is found to be, D 3m =
1.588 * 10 –5 m 2/sec.
Diffusivity in liquids:
Diffusivity in liquid are exemplified by the values given in table … Most of these values are
nearer to 10-5 cm2 / sec, and about ten thousand times shower than those in dilute gases. This
characteristic of liquid diffusion often limits the overall rate of processes accruing in liquids
(such as reaction between two components in liquids).
In chemistry, diffusivity limits the rate of acid-base reactions; in the chemical industry,
19
diffusion is responsible for the rates of liquid-liquid extraction. Diffusion in liquids is
important because it is slow.
Certain molecules diffuse as molecules, while others which are designated as electrolytes
ionize in solutions and diffuse as ions. For example, sodium chloride (NaCl), diffuses in water
as ions Na + and Cl-. Though each ions has a different mobility, the electrical neutrality of the
solution indicates the ions must diffuse at the same rate; accordingly it is possible to speak of a
diffusion coefficient for molecular electrolytes such as NaCl. However, if several ions are
present, the diffusion rates of the individual cations and anions must be considered, and
molecular diffusion coefficients have no meaning.
Diffusivity varies inversely with viscosity when the ratio of solute to solvent ratio exceeds five.
In extremely high viscosity materials, diffusion becomes independent of viscosity.
Diffusivity in solids:
Typical values for diffusivity in solids are shown in table. One outstanding characteristic of
these values is their small size, usually thousands of time less than those in a liquid, which are
inturn 10,000 times less than those in a gas.
Diffusion plays a major role in catalysis and is important to the chemical engineer. For
metallurgists, diffusion of atoms within the solids is of more importance.
This part is an application to the general differential equation of mass transfer. The objective is
to solve the differential equation of mass transfer under steady state conditions at different
conditions (chemical reaction, one dimensional or more etc.).
Two approaches will be used to simplify the differential equations of mass transfer:
1. Fick’s equation and the general differential equation for mass transfer can be simplified
by eliminating the terms that do not apply to the physical situation.
2. A material balance can be performed on a differential volume element of the control
volume for mass transfer.
20
One dimensional mass transfer independent of chemical reaction
The diffusion coefficient or mass diffusivity for a gas may be experimentally measured in an Arnold
diffusion cell. This cell is illustrated schematically in Figure below.
The narrow tube, which is partially filled with pure liquid A, is maintained at a constant temperature
and pressure. Gas B, which flows across the open end of the tube, has a negligible solubility in liquid A
and is also chemically inert to A.
Component A vaporizes and diffuses into the gas phase; the rate of vaporization may be physically
measured and may also be mathematically expressed in terms of the molar mass flux.
∂c A
∇ .⃗
NA+ −R A=0
∂t
21
∂ ∂ ∂ ∂c
N A , x+ N A ,y+ N A , z + A −R A =0
∂x ∂y ∂z ∂t
d
∴ N =0
dz A , z
It means that the molar flux of A is constant over the entire diffusion path from 𝑧1 to 𝑧2.
The molar flux is defined by the equation:
d yA
N A =−c D AB + y A( N A+ N B)
dZ
∴ N B=0
d yA
N A =−c D AB + yAN A
dZ
−c D AB d y A
N A=
1− y A dz
To get the flux 𝑁𝐴 the above equation has to be integrated between 𝑧1 and 𝑧2 by using the boundary
conditions:
At z=z 1 , y A = y A 1
At z=z 2 , y A = y A 2
z2 yA 2
−d y A
N A ∫ dz=c D AB ∫
z1 yA 1
1− y A
c D AB (1− y A 2 )
N A= ln
z 2−z 1 (1− y A 1 )
………………. (1)
22
The above equation (equation 1) is commonly referred to as equation for steady-state diffusion of one
gas through a second non-diffusing gas or stagnant gas.
d yA
For, N A =−c D AB dZ
+ y A ( N A+ N B)
−D AB d p A p A
N A= + ( −N B + N B )
RT dZ Pt
−D AB d p A
Therefore, N A =
RT dZ
z2 P A2
DAB
N A ∫ dz=− ∫ dp A
z1 RT P A1
D AB
RTz ( A1
N A= P −PA2 )
23
Graphically: Equimolal counterdiffusion
In mass transfer operations, two insoluble phases are brought into contact in order to permit
transfer of constituent substances between them. Transfer across each phase is dependent on
24
the concentration gradient existing within them; this is an indication of the departure from
equilibrium which exists between the two phases.
2.1 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.
Illustration;
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.
[N
1 ammonia-air water
H3
]
[NH3]
ammonia-air water
2
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
25
into the gas phase condenses into the solution. Dynamic equilibrium is 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.
Summarily, the following principles are common to all systems involving the distribution of
substances between two insoluble phases;
i. At a fixed set of conditions, such as temperature and pressure, there exists a set of
equilibrium relations exists, which may be shown in the form of an equilibrium
distribution curve.
ii. When the system is in equilibrium, there is no net mass transfer between the phases.
iii. 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.
Considering a solute is transferred from the gas phase into a liquid phase. The interphase
transfer involves three transfer steps:
i. The transfer of mass from the bulk conditions of one phase to the interfacial surface
ii. The transfer across the interface into the second phase
iii. The transfer to the bulk conditions of the second phase
26
Gas film Liquid film
NA
Gas-Liquid
interface
Figure 2 Gas absorption with solute A transferred from gas phase to liquid phase
i. The rate of mass transfer between the two phases is controlled by the rates of diffusion
through the phases on each side of the interface.
ii. No resistance is offered to the transfer of the diffusing component across the interface.
Consider the mass transfer of a solute A from the bulk of a gas phase to the bulk of a liquid
phase.
27
yAG interface
Concentration of diffusing solute A
xAi
Gas Liquid
yAi
xAL
Distance
Figure 3 Concentration gradients between two contacting phases where solute is transferred
from gas to liquid.
This can be shown graphically in terms of distance through the phases as shown in Figure
above. The concentration of A in the main body of the gas is y AG mole fraction and it falls to
y Ai at the interface. In the liquid, the concentration falls from x Ai at the interface to x AL in the
bulk liquid. There is no resistance to solute transfer across the interface separating the phases.
Only diffusional resistances are residing in the fluids. The equilibrium concentrations y Ai and
x Ai are obtained from the system’s equilibrium distribution curve. In the figure the
concentration rise at the interface from y Ai to x Ai is not a barrier to diffusion in the direction
gas to liquid. They are equilibrium concentrations.
Note; The interfacial partial pressure, p Ai can be less than, equal to or greater than the value
of c Ai according to the equilibrium conditions of the temperature and pressure of the system.
28
2.3 LOCAL TWO-PHASE MASS TRANSFER
N A, z =k G ( y A ,G −y A,i ).........................i
N A, z =k L (x A,i−x A ,L )..........................ii
Where kG & kL is the convective mass transfer coefficient in the gas phase and liquid
respectively.
The mole fraction difference y A ,G− y A ,i is the driving force necessary to transfer component A
from the bulk gas conditions to the interface separating the two phases. The concentration
difference, x A ,i−x A , L , is the driving force necessary to continue the transfer of A into the liquid
phase.
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 G ( y AG − y Ai )=−k L ( x Ai −x AL )
k y − y Ai
− L = AG
k G x AL −x Ai
The application of equation above for the evaluation of the interfacial compositions for a
specific set of bulk compositions as represented by point O is illustrated.
29
yAG O
-kL/kG
A composition in gas phase
Equilibrium
curve
yAi M
C
xAL xAi
A composition in liquid phase
30
2.4 OVERALL MASS-TRANSFER COEFFICIENTS
compositions,
y AG and x AL .
¿
In Figure 4, one observes the bulk liquid composition
x A ,L is in equilibrium with
y A is as
x
good measure of A , L as
x A , L itself and it has units consistent with y A ,G .
Equilibrium
yAG P curve
Slope=-kL/kG
Composition in gas phase
Slope
=m”
yAi Slope M
=m’
yA* C
xAL xAi xA*
Composition in liquid phase
Figure 5 Overall concentration differences
The entire two-phase mass-transfer effect can then be measured in terms of an overall-mass
transfer coefficient
KG .
¿
N A=K G ( y A ,G−y A ) ……………..1
31
K L which involves the resistance to diffusion in both phases and is in terms of the liquid
phase concentration driving force, is defined by
¿
N A=K L ( x A −x A ,L ) ……………….2
Figure 4 illustrates the driving forces associated with each phase and the overall driving forces.
The ratio of the resistances in an individual phase to the total resistance may be determined by
1/k G
Resistance ∈t h e gas p h ase
Total resistance∈bot h p h ases = 1 /K G
And
1/k L
Resistance∈the liquid phase
Total resistance∈both phases = 1 /K L
A relation between these overall coefficients and the individual phase coefficients can be
obtained when the equilibrium relation is linear as expressed by
y A ,i =mx A ,i
As always encountered at low concentrations where Henry’s law is obeyed; relating gas- and
liquid-phase concentrations
¿
y A,G=mx A
¿
y A=mx A, L
And
y A ,i =mx A ,i
Re-arranging 1;
¿ ¿
1 ( y A , G− y A ) ( y A , G− y A ,i ) ( y A , i− y A )
= = +
KG NA NA NA
'
1 ( y A , G− y A ,i ) m (x A ,i −x L )
= +
KG NA NA ,
Substituting equations (i) and (ii) into the above relation relates
K G to the individual phase
coefficients by
32
1 1 m'
= +
K G k G k L ……………………….3
¿
In a similar fashion,
x A is a measure of
y A ,G and can be used to define another overall
coefficient
KL :
¿
N A=K L ( x A −x A ,L )
¿ ¿
1 ( x A−x A , L ) ( x A −x A , i ) ( x A , i−x A , L )
= = +
KL NA NA NA
¿
1 ( x A−x A , L ) ( y A ,G − y A , i ) ( x A , i −x A , L )
= = +
KL NA ''
m NA NA
1 1 1
= '' +
K L m kG kL
…………………….4
Stipulate that the relative magnitudes of the individual phase resistances depend on the
solubility of the gas, as indicated by the magnitude of the proportionality constant. For a
system involving a soluble gas, such as ammonia in water, m’ is very small. From equation 3,
we may conclude that the gas-phase resistance is essentially equal to the overall resistance in
such a system. When this is true, the major resistance to mass transfer lies in the gas phase, and
such a system is said to be gas-phase controlled.
Systems involving gases of low solubility, such as carbon dioxide in water, have such a large
value of m” that equation 4 stipulates that the gas-phase resistance may be neglected, and the
overall coefficient, KL is essentially equal to the individual liquid phase coefficient, kL. This
type of system is designated liquid-phase controlled.
33
2.5 MATERIALS BALANCES
The material balance is important, as it provides expressions for evaluating the bulk
compositions of the two contacting phases at any plane in the separating equipments.
Consider any steady-state mass-transfer operation that involves the countercurrent contact of
two insoluble phases as Figure 6. The two insoluble phases will be identified as phase G and
phase L.
L2,x2,X2 G2,y2,Y2
z=z2
Lz,xz,Xz z Gz,yz,Yz
L1,x1,X1 G1,y1,Y1
Figure 6; Steady-state countercurrent process
At the bottom of the mass-transfer tower, the flow rates and concentrations are defined as
follows:
G1 Is the total moles of phase G entering the tower per hour per cross-sectional area of the
tower
L1 is the total moles of phase L leaving the tower per hour per cross-sectional area of the tower
y 1 Is the mole fraction of component A in G , expressed as moles of A per total moles in
1
phase G
x 1 Is the mole fraction of component A in L , expressed as moles of A per total moles in
1
phase L
34
z
Similarly at the top of the tower, or plane 2 , the total moles of each phase will be
G2 and L2
G1 y 1 +L z x z =G z y z +L1 x 1 …………………..1
Simpler relations in terms of solute-free concentration units
Y is the moles of A in G per mole of A-free G;
yA
Y A=
1− y A
X is the moles of A in L per mole of A-free L;
xA
X A=
1−x A
The solute-free concentration units are LS and GS
G1=Gs +G1 y 1
y1
G1 y 1=G s =G s Y 1
1− y 1
The overall balance on component A may be written using the solute-free terms as
Gs Y 1 + Ls X 2 =G s Y 2 +Ls X 1
Or
G s (Y 1−Y 2 )=Ls ( X 1− X 2 ) ………………..2
(Y 1 −Y 2 ) Ls
=
( X 1 −X 2 ) G s
A mass balance of component A around plane z1 and an arbitrary plane z=z in solute-free
terms is
Gs Y 1 + Ls X z =Gs Y z + Ls X 1
Or
35
Gs (Y 1−Y z )=L s ( X 1 −X z ) ……………….3
Re arranging
(Y 1 −Y z ) Ls
=
( X 1 −X z ) G s
Two straight lines having the same slope and a point in common lie on the same straight line.
Note; Equations (1) and (3), although both equations describe the mass balance for component
A, only equation (3) is an equation of a straight line. When written in the solute-free units, X
and Y, the operating line is straight because the mole-ratio concentrations are based on both
the constant quantities, Ls and Gs. When written in mole-fraction units, x and y, the total
moles in a phase, L or G, change as the solute is transferred into or out of the phase; this
produces a curved operating line on the x–y coordinates.
Figure 7 illustrates the location of the operating line relative to the equilibrium line when the
transfer of the solute is from phase G to phase L as in the case of absorption.
YA
Operating
YA1
Slope=Ls line
Equilibrium
/Gs curve
YAi vs XAi
YA2
XA2 XA1 XA
Figure 7 Steady-state countercurrent process, transfer from phase G to L.
The bulk equilibrium, located on the operating line, must be greater than the equilibrium
concentration in order to provide the driving forces, YAG-YAi needed for transfer from G phase
to L phase.
Figure 8 illustrates the location of the operating line relative to the equilibrium line when the
transfer of the solute is from phase L to phase G as in desorption or stripping.
36
YA Equilibrium
curve
YA2 YAi vs XAi
Operating
line
Slope=Ls/
YA1 Gs
XA1 XA2 XA
Figure 8 s-s countercurrent process transfer from phase to phase G
The closer the operating line is to the equilibrium curve, the smaller will be the driving force
for overcoming any mass transfer resistance. At the point of tangency, the diffusional driving
force is zero; thus, mass transfer between the two phases cannot occur. This then represents a
limiting condition, the minimum LS/GS ratio for mass transfer.
YA
YA1 P1 P2 P3 Equilibrium
curve
( L s/
1L /
)3
G s
(Ls /
Gs
Gs ) (
2
s )
YA2
XA2 XA
Figure 9 Operating-line locations.
37
2.7 STAGE
A device in which two insoluble phases are brought into intimate contact, where mass transfer
occurs between the two phases tending to bring them to equilibrium, and where the phases are
then mechanically separated. A process carried out in this manner is a single-stage process.
An equilibrium stage is one where the time of contact between phases is sufficient for the
effluents indeed to be in equilibrium. However, in principle this cannot be attained.
2.8 CASCADE
A group of stages interconnected so that the various streams flow from one to the other is
called a cascade. Its purpose is to increase the extent of mass transfer over and above that
which is possible with a single stage.
38
3.0 DISTILLATION
Distillation is method of separation of components from a liquid mixture which depends on the
differences in boiling points of the individual components and the distributions of the components
between a liquid and gas phase in the mixture. The liquid mixture may have different boiling point
characteristics depending on the concentrations of the components present in it. Therefore,
distillation processes depends on the vapor pressure characteristics of liquid mixtures. The vapor
pressure is created by supplying heat as separating agent. In the distillation, the new phases differ
from the original by their heat content. During most of the century, distillation was by far the most
widely used method for separating liquid mixtures of chemical components (Seader and Henley,
1998). This is a very energy intensive technique, especially when the relative volatility of the
components is low. It is mostly carried out in multi-tray columns. Packed column with efficient
structured packing has also led to increased use in distillation.
Vapor Pressure: The vaporization process changes liquid to gaseous state. The opposite process of
this vaporization is called condensation. At equilibrium, the rates of these two processes are same.
The pressure exerted by the vapor at this equilibrium state is termed as the vapor pressure of the
liquid. It depends on the temperature and the quantity of the liquid and vapor. From the following
Clausius-Clapeyron Equation or by using Antoine Equation, the vapor pressure can be calculated.
Clapeyron Equation
( ) ( )( )
Pv
ln v =
P1
λ 1 1
−
R T1 T
(3.1)
v
Where P and P1 are the vapour pressures in pascal at absolute temperature T (0 K) and T in K. λ
v
39
Typical representative values of the constants A, B and C are given in the following Table 3. 1 (Ghosal et al.,
1993)
Table 3. 1 Typical representative values of the constants A, B and C
For binary mixture phase diagram only two-component mixture, (e.g. A (more volatile) and B (less volatile))
are considered. There are two types of phase diagram: constant pressure and constant temperature.
The Figure 10 shows a constant pressure phase diagram for an ideal solution (one that obeys Raoult's Law).
At constant pressure, depending on relative concentrations of each component in the liquid, many boiling
point temperatures are possible for mixture of liquids (solutions) as shown in phase diagram (Figure 10). For
mixture, the temperature is called bubble point temperature when the liquid starts to boil and dew point when
the vapor starts to condense. Boiling of a liquid mixture takes place over a range of boiling points. Likewise,
condensation of a vapor mixture takes place over a range of condensation points. The upper curve in the
boiling point diagram is called the dew-point curve (DPC) while the lower one is called the bubble-point
curve (BPC). At each temperature, the vapor and the liquid are in equilibrium. The constant pressure phase
diagram is more commonly used in the analysis of vapor-liquid equilibrium.
40
Figure 10 Phase diagram of binary system at constant pressure
The constant temperature phase diagram is shown in Figure 11. The constant temperature phase diagram is
useful in the analysis of solution behaviour. The more volatile liquid will have a higher vapor pressure (i.e.
pA at xA = 1.0)
41
More on phase diagrams to illustrate fractional distillation of ideal mixtures
If you boil a liquid mixture C1, you will get a vapor with composition C2, which you can condense to
give a liquid of that same composition (the pale blue lines).
If you reboil that liquid C2, it will give a vapor with composition C3. Again you can condense that to
give a liquid of the same new composition (the red lines).
Reboiling the liquid C3 will give a vapor still richer in the more volatile component B (the green lines).
You can see that if you were to do this once or twice more, you would be able to collect a liquid which
was virtually pure B.
The secret of getting the more volatile component from a mixture of liquids is obviously to do a
succession of boiling-condensing-reboiling operations. It is not quite so obvious how you get a sample
of pure A out of this.
42
Figure X: Fractional distillation setup. An Erlenmeyer flask is used as a receiving flask. Here the
distillation head and fractionating column are combined in one piece. from Wikipedia
The fractionating column is packed with glass beads (or something similar) to give the maximum
possible surface area for vapor to condense on. Some fractionating columns have spikes of glass sticking
out from the sides which serve the same purpose. If you sketch this, make sure that you do not
completely seal the apparatus. There has to be a vent in the system otherwise the pressure build-up when
you heat it will blow the apparatus apart. In some cases, where you are collecting a liquid with a very
low boiling point, you may need to surround the collecting flask with a beaker of cold water or ice. The
mixture is heated at such a rate that the thermometer is at the temperature of the boiling point of the
more volatile component. Notice that the thermometer bulb is placed exactly at the outlet from the
fractionating column.
Suppose you boil a mixture with composition C1. The vapor over the top of the boiling liquid will be
richer in the more volatile component, and will have the composition C2.
43
That vapor now starts to travel up the fractionating column. Eventually it will reach a height in the
column where the temperature is low enough that it will condense to give a liquid. The composition of
that liquid will, of course, still be C 2. So what happens to that liquid now? It will start to trickle down
the column where it will meet new hot vapor rising. That will cause the already condensed vapor to
reboil.
Some of the liquid of composition C2 will boil to give a vapor of composition C3. Let's concentrate first
on that new vapor and think about the unvaporized part of the liquid afterwards.
The vapor
This new vapor will again move further up the fractionating column until it gets to a temperature where
it can condense. Then the whole process repeats itself. Each time the vapor condenses to a liquid, this
44
liquid will start to trickle back down the column where it will be reboiled by up-coming hot vapor. Each
time this happens the new vapor will be richer in the more volatile component. The aim is to balance the
temperature of the column so that by the time vapor reaches the top after huge numbers of condensing
and reboiling operations, it consists only of the more volatile component - in this case, B.
Whether or not this is possible depends on the difference between the boiling points of the two liquids.
The closer they are together, the longer the column has to be.
The liquid
So what about the liquid left behind at each reboiling? Obviously, if the vapor is richer in the more
volatile component, the liquid left behind must be getting richer in the other one. As the condensed
liquid trickles down the column constantly being reboiled by up-coming vapor, each reboiling makes it
richer and richer in the less volatile component - in this case, A. By the time the liquid drips back into
the flask, it will be very rich in A indeed. So, over time, as B passes out of the top of the column into the
condenser, the liquid in the flask will become richer in A. If you are very, very careful over temperature
control, eventually you will have separated the mixture into B in the collecting flask and A in the
original flask. Finally, what is the point of the packing in the column?
To make the boiling-condensing-reboiling process as effective as possible, it has to happen over and
over again. By having a lot of surface area inside the column, you aim to have the maximum possible
contact between the liquid trickling down and the hot vapor rising. If you didn't have the packing, the
liquid would all be on the sides of the condenser, while most of the vapor would be going up the middle
and never come into contact with it.
There is no difference whatsoever in the theory involved. All that is different is what the fractionating
column looks like. The diagram shows a simplified cross-section through a small part of a typical
column.
45
The column contains a number of trays that the liquid collects on as the vapor condenses. The up-
coming hot vapor is forced through the liquid on the trays by passing through a number of bubble caps.
This produces the maximum possible contact between the vapor and liquid. This all makes the boiling-
condensing-reboiling process as efficient as possible. The overflow pipes are simply a controlled way of
letting liquid trickle down the column.
If you have a mixture of lots of liquids to separate (such as in petroleum fractionation), it is possible to
tap off the liquids from some of the trays rather than just collecting what comes out of the top of the
column. That leads to simpler mixtures such as gasoline, kerosene and so on.
Example 1
Use the phase diagram below to explain how you can obtain a pure sample of B from a mixture M by
successively boiling and condensing the liquid mixture.
46
Why is it important to carefully control how strongly the original mixture is heated during the
separation?
Explain briefly how the separation occurs, making use of the phase diagram above if you think it helps.
Solution
If you boil the mixture M, it will boil at a temperature T1. The vapor above the liquid at this temperature
will be richer in the more volatile substance B. If you condense that vapor, it will give a liquid of the
composition M1. If you reboil that, it will boil at a temperature T2. The vapor over that liquid will have
a composition M2, still richer in B. If you go on doing that, reboiling and recondensing, then the vapor
becomes richer and richer in B until it eventually becomes pure B. When you finally get to that point
and condense the vapor, then you will have pure B liquid.
You have to be sure that only the vapor of the more volatile of the two liquids passes into the condenser.
That means that the thermometer has to read exactly the boiling point of the more volatile liquid. If it is
47
below that, then nothing is going to pass out into the condenser. If it is above that, then your distillate
will still contain some of the less volatile component.
B is the more volatile liquid; A is the less volatile one. The vapor over the boiling liquid in the flask will
be richer in B than the original liquid is. That vapor will pass up the column until the temperature falls
enough for it to condense to give a liquid richer in B than the one in the flask (equivalent to M1 in the
diagram). This will start to trickle down the column. Hot vapor coming up from the flask will reboil the
condensed liquid, giving a vapor which will be even richer in B (M2 on the diagram). This will
condense to a liquid, trickle down the column and then be reboiled. This continuous process will go on
until the vapor is entirely B. The column is heated so that this is finally complete right at the top of the
column. Meanwhile, the liquids trickling down the column get richer and richer in A as the B is
removed and carried up the column. Eventually, the liquid in the flask will end up as pure A.
Relative volatility
Relative volatility is a measure of the differences in volatility between two components, and hence their
48
boiling points. It indicates how easy or difficult a particular separation will be. The relative volatility of
component ‘A’ with respect to component ‘B’ in a binary mixture is defined as
y A/ xA
α AB=
yB / xB (3.3)
Where, yA = mole fraction of component ‘A’ in the vapor, x A = mole fraction of component ‘A’ in the liquid.
In general, relative volatility of a mixture changes with the mixture composition. For binary mixture, x B = 1-
xA. So Equation (3.3) can be rearranged, simplifying and expressed by dropping subscript A' for more
volatile component as:
α ave x
y= (3.4)
1+ ( α ave −1 ) x
The Equation (3.4) is a non-linear relationships between x and y. This Equation can be used to determine the
Some optimal relative volatility that are used for distillation process design
49
(normal bpt °C) (normal bpt °C) relative
volatility
benzene (80.1) Toluene (110.6) 2.34
Toluene (110.6) p-Xylene (138.3) 2.31
benzene (80.1) p-Xylene (138.3) 4.82
m-Xylene (139.1) p-Xylene (138.3) 1.02
pentane (36.0) Hexane (68.7) 2.59
Hexane (68.7) Heptane (98.5) 2.45
Hexane (68.7) p-Xylene (138.3) 7.0
Ethanol (78.4) iso-Propanol (82.3) 1.17
iso-Propanol (82.3) n-Propanol (97.3) 1.78
Ethanol (78.4) n-Propanol (97.3) 2.1
Methanol (64.6) Ethanol (78.4) 1.56
It is useful for graphical design in determining the number of theoretical stages required for a distillation
column. A typical equilibrium curve for a binary mixture on x-y plot is shown in Figure 12. It can be plotted
by the Equations (3.4) or (3.5) as discussed earlier section. It contains less information than the phase
diagram (i.e. temperature is not included), but it is most commonly used. The VLE plot expresses the bubble-
point and the dew-point of a binary mixture at constant pressure. The curved line in Figure 12 is called the
equilibrium line and it describes the compositions of the liquid and vapor in equilibrium at some fixed
pressure. The equilibrium line can be obtained from the Equation (3.4) once the relative volatility is known.
50
Figure 12 Equilibrium diagram(yx) for a benzene-toluene mixture at 1 atmosphere
This particular VLE plot (Figure 12) shows a binary ideal mixture that has a uniform vapor-liquid equilibrium
that is relatively easy to separate. On the other hand, the VLE plots shown in Figure 13 represented for non-
ideal systems. These non-ideal VLE systems will present more difficult separation.
51
The most challenging VLE curves are generated by azeotropic systems. An azeotrope is a liquid mixture
which when vaporized, produces the same composition as the liquid. Two types of azeotropes are known:
minimum-boiling and maximum-boiling.
Ethanol-water system (at 1 atm, 89.4 mole %C-OH, 78.2 oC) Carbon-disulfide - acetone (61.0 mole% CS2,
39.25 oC, 1 atm) and Benzene - water (29.6 mole% H 2O, 69.25 oC, 1 atm) are minimum-boiling azeotropes.
Hydrochloric acid - water (11.1 mole% HCl, 110 oC, 1 atm); Acetone - chloroform (65.5 mole%
chloroform, 64.5 oC, 1 atm) are the examples of Maximum-boiling azeotropes. The Figure 14 shows two
different azeotropic systems, one with a minimum boiling point ( Figure 14a) and one with a maximum boiling
point (Figure 14b). The points of intersections of the equilibrium curves with the diagonal lines are called
azeotropic points. An azeotrope cannot be separated by conventional distillation. However, vacuum
distillation may be used as the lower pressures can shift the azeotropic point.
Figure 14 VLE curves for azeotropic systems: (a) for maximum boiling point, (b) for minimum
boiling point
Although most distillations are carried out at atmospheric or near atmospheric pressure, it is common to
distill at other pressures. High pressure distillation (typically 3 - 20 atm) usually occurs in thermally
integrated processes. In those cases the equilibrium curve becomes narrower at higher pressures as shown in
Figure 15. Separability becomes less at higher pressures.
52
Figure 15 Variation of equilibrium curve with pressure
53
Suppose you are going to distil a mixture of ethanol and water with composition C 1 as shown on the
next diagram. It will boil at a temperature given by the liquid curve and produce a vapor with
composition C2.
When that vapor condenses it will, of course, still have the composition C 2. If you reboil that, it will
produce a new vapor with composition C3.
54
You can see that if you carried on with this boiling-condensing-reboiling sequence, you would
eventually end up with a vapor with a composition of 95.6% ethanol. If you condense that you
obviously get a liquid with 95.6% ethanol.
What happens if you reboil that liquid? The liquid curve and the vapor curve meet at that point. The
vapor produced will have that same composition of 95.6% ethanol. If you condense it again, it will still
have that same composition. You have hit a barrier. It is impossible to get pure ethanol by distilling any
mixture of ethanol and water containing less than 95.6% of ethanol. This particular mixture of ethanol
and water boils as if it were a pure liquid. It has a constant boiling point, and the vapor composition is
exactly the same as the liquid. It is known as a constant boiling mixture or an azeotropic mixture or
an azeotrope.
The implications of this for fractional distillation of dilute solutions of ethanol are obvious. The liquid
collected by condensing the vapor from the top of the fractionating column cannot be pure ethanol. The
best you can produce by simple fractional distillation is 95.6% ethanol. What you can get (although it
isn't very useful!) from the mixture is pure water. As ethanol rich vapor is given off from the liquid
boiling in the distillation flask, it will eventually lose all the ethanol to leave just water.
To Summarize
Distilling a mixture of ethanol containing less than 95.6% of ethanol by mass lets you collect:
a distillate containing 95.6% of ethanol in the collecting flask (provided you are careful with the
temperature control, and the fractionating column is long enough);
55
pure water in the boiling flask.
What if you distil a mixture containing more than 95.6% ethanol? Work it out for yourself using the
phase diagram, and starting with a composition to the right of the azeotropic mixture. You should find
that you get:
a distillate containing 95.6% of ethanol in the collecting flask (provided you are careful with the
temperature control and the fractionating column is long enough);
pure ethanol in the boiling flask.
A negative deviation from Raoult's Law
Nitric acid and water form mixtures in which particles break away to form the vapor with much more
difficulty than in either of the pure liquids. You can see this from the vapor pressure / composition curve
discussed further up the page. That means that mixtures of nitric acid and water can have boiling points
higher than either of the pure liquids because it needs extra heat to break the stronger attractions in the
mixture.
In the case of mixtures of nitric acid and water, there is a maximum boiling point of 120.5°C when the
mixture contains 68% by mass of nitric acid. That compares with the boiling point of pure nitric acid at
86°C, and water at 100°C. Notice the much bigger difference this time due to the presence of the new
ionic interactions (see above). The phase diagram looks like this:
Distilling dilute nitric acid. Start with a dilute solution of nitric acid with a composition of C 1 and trace
through what happens.
56
The vapor produced is richer in water than the original acid. If you condense the vapor and reboil it, the
new vapor is even richer in water. Fractional distillation of dilute nitric acid will enable you to collect
pure water from the top of the fractionating column. As the acid loses water, it becomes more
concentrated. Its concentration gradually increases until it gets to 68% by mass of nitric acid. At that
point, the vapor produced has exactly the same concentration as the liquid, because the two curves meet.
You produce a constant boiling mixture (or azeotropic mixture or azeotrope) and if you distil dilute
nitric acid, that's what you will eventually be left with in the distillation flask. You cannot produce pure
nitric acid from the dilute acid by distilling it.
You cannot produce pure nitric acid from the dilute acid (<68%) by distilling it.
Distilling nitric acid more concentrated than 68% by mass
This time you are starting with a concentration C2 to the right of the azeotropic mixture.
57
The vapor formed is richer in nitric acid. If you condense and reboil this, you will get a still richer
vapor. If you continue to do this all the way up the fractionating column, you can get pure nitric acid out
of the top. As far as the liquid in the distillation flask is concerned, it is gradually losing nitric acid. Its
concentration drifts down towards the azeotropic composition. Once it reaches that, there cannot be any
further change, because it then boils to give a vapor with the same composition as the liquid. Distilling a
nitric acid / water mixture containing more than 68% by mass of nitric acid gives you pure nitric acid
from the top of the fractionating column and the azeotropic mixture left in the distillation flask.
You can produce pure nitric acid from the concetrated acid (>68%) by distilling it
McCabe and Thiele (1925) developed a graphical method to determine the theoretical number of stages
required to effect the separation of a binary mixture. This method uses the equilibrium curve diagram to
determine the number of theoretical stages (trays) required to achieve a desired degree of separation. It
assumes constant molar overflow and this implies that:
(i) Molal heats of vaporization of the components are roughly the same.
(ii) Heat effects are negligible.
The information required for the systematic calculation are the VLE data, feed condition
(temperature, composition), distillate and bottom compositions; and the reflux ratio, which is
58
defined as the ratio of reflux liquid over the distillate product. For example, a column is to be
designed for the separation of a binary mixture as shown in Figure 16.
The feed has a concentration of xF (mole fraction) of the more volatile component (MVC), and a
distillate having a concentration of xD of the MVC and a bottoms having a concentration of xB is desired.
In its essence, the method involves the plotting on the equilibrium diagram three straight lines:
i. The rectifying section operating line (ROL)
ii. The feed line (also known as the q-line)
iii. The stripping section operating line (SOL)
An important parameter in the analysis of continuous distillation is the Reflux Ratio, R defined as the
quantity of liquid returned to the distillation column over the quantity of liquid withdrawn as product
from the column, i.e. R = L/D. The reflux ratio R is important because the concentration of the more
volatile component in the distillate (in mole fraction x D) can be changed by changing the value of R. The
steps to be followed to determine the number of theoretical stages by McCabe-Thiele Method:
i. Determination of the Rectifying section operating line (ROL).
ii. Determination the feed condition (q).
iii. Determination of the feed section operating line (q-line).
59
iv. Determination of required reflux ratio (R).
v. Determination of the stripping section operating line (SOL).
vi. Determination of number of theoretical stage.
Consider the rectifying section as shown in the Figure 17. Material balance can be written around the
envelope shown in Figure 17:
Overall or total balance;
V n+1=L n +D (3.6)
Component balance for MVC
V n+1 y n+1=Ln x n +Dx D (3.7)
From Equations (3.6) and (3.7), it can be written as
60
Figure 17 Outline graph of rectifying section
Introducing reflux ratio defined as: R = L/D, the Equation (3.10) can be expressed as:
R 1
y n+1= xn+ x
R+1 R+1 D (3.11)
The Equation (3.11) is the rectifying section operating line (ROL) having slope R/(R+1) and intercept,
xD/(R+1) as shown in Figure 18. If xn = xD, then yn+1 = xD, the operating line passed through the point (x D,
xD) on the 45o diagonal line. When the reflux ratio R changed, the ROL will change. Generally the
61
rectifying operating line is expressed without subscript of n or n+1. Without subscript the ROL is
expresses as:
R 1
y= x+ x
R+1 R+1 D (3.12)
62
The feed enters the distillation column may consists of liquid, vapor or a mixture of both. Some portions
of the feed go as the liquid and vapor stream to the rectifying and stripping sections. The moles of liquid
flow in the stripping section that result from the introduction of each mole of feed, denoted as ‘q’. The
limitations of the q-value as per feed conditions are shown in Table 3.2.
Table 3.2 Limitations of q-value as per feed conditions
Calculation of q-value
When feed is partially vaporized?
Other than saturated liquid (q=1) and saturated vapor (q=0), the feed condition is uncertain. In that
case one must calculate the value of q. The q-value can be obtained from enthalpy balance around the
feed plate. By enthalpy balance one can obtain the q-value from the following form of Equation:
q=
( H V −H F
H V −H L ) (3.13)
Where HF, HV and HL are enthalpies of feed, vapor and liquid respectively which can be obtained
from enthalpy-concentration diagram for the mixture
q can be alternatively defined as the heat required to convert 1 mole of feed from its entering
condition to a saturated vapor; divided by the molal latent heat of vaporization. Based on this
63
definition, one can calculate the q-value from the following Equations for the case whereby q > 1
(cold liquid feed) and q < 0 (superheated vapor feed) as:
For cold liquid feed;
C p , L (T bpt −T f )+λ
q=
λ (3.14)
For superheated vapor feed;
C p , V (T dpt−T f )
q=
λ (3.15)
Where Tbpt is the bubble point, λ is the latent heat of vaporization and Tdpt is the dew point of the
feed respectively.
Consider the section of the distillation column (Figure 16) at the tray (called feed tray) where the feed is
introduced. In the feed tray the feed is introduced at F moles/hr with liquid of q fraction of feed and
vapor of (1-q) fraction of feed as shown in Figure 19 Overall material balance around the feed tray:
'
L =L+qF and V =V ' +(1−q)F (3.16)
Component balances for the more volatile component in the rectifying and stripping sections
are:
For rectifying section;
Vy=Lx +Dx D (3.17)
For striping section;
V ' y =L' x−Bx B (3.18)
At the feed point where the two operating lines (Equations (3.17) and (3.18) intersect can be
written as:
(V −V ' ) y=( L−L' ) x+ Dx D +Bx B (3.19)
64
Figure 19Feed tray with fraction of liquid and vapor of feed
y=− ( ) ( )
q
1−q
x+
1
x
1−q F (3.21)
For a given feed condition, xF and q are fixed, therefore the q-line is a straight line with slope -q / (1-q)
and intercept xF/(1-q). If x = xF , then from Equation (3.21) y = xF. At this condition the q-line passes
through the point (xF, xF) on the 45o diagonal. Different values of q will result in different slope of the q-
line. Different q-lines for different feed conditions are shown in Figure 20.
65
Figure 20 Different q-lines for different feed conditions
The stripping section operating line (SOL) can be obtained from the ROL and q-line without doing any
material balance. The SOL can be drawn by connecting point x B on the diagonal to the point of
intersection between the ROL and q-line. The SOL will change if q-line is changed at fixed ROL. The
change of SOL with different q-lines for a given ROL at constant R and xD is shown in Figure 21.
66
Figure 21 Stripping section operating line with different q-lines
The stripping section operating line can be derived from the material balance around the stripping
section of the distillation column. The stripping section of a distillation column is shown in Figure 22.
The reboiled vapor is in equilibrium with bottoms liquid which is leaving the column.
67
Consider the constant molal overflow in the column. Thus
' ' ' ' ' '
Lm=Lm+1 =. .. .. ..=L =cons tan t and V m=V m+1=. .. . .. .=V =cons tant
Overall material balance gives
' '
L =V +B
MVC balance gives;
' '
L x m=V y m+1 +Bx B
L' B
y= ' x− ' x B
V V
' '
Substituting V =L −B
( ) ( )
'
L B
y= '
x− ' xB
L −B L −B
This is called the stripping operating line (SOL) which is a straight line with slope
(L'/L'-B) and intercept (BxB/L'-B). When x = xB , y = xB, the SOL passes through (xB, xB ) on the 45o
diagonal line.
Suppose a column is to be designed for the separation of a binary mixture where the feed has a
concentration of xF (mole fraction) of the MVC and a distillate having a concentration of x D of the more
volatile component whereas the bottoms having a desired concentration of x B. Once the three lines
(ROL, SOL and q-line) are drawn, the number of theoretical stages required for a given separation is
then the number of triangles that can be drawn between these operating lines and the equilibrium curve.
The last triangle on the diagram represents the reboiler. A typical representation is given in Figure 23.
68
Figure 23 A typical representation of identifying number of theoretical stages
Reflux Ratio, R
The separation efficiency by distillation depends on the reflux ratio. For a given separation (i.e. constant
xD and xB) from a given feed condition (xF and q), higher reflux ratio (R) results in lesser number of
required theoretical trays (N) and vice versa. So there is an inverse relationship between the reflux ratio
and the number of theoretical stages. At a specified distillate concentration, x D, when R changes, the
slope and intercept of the ROL changes (Equation (3.12)). From the Equation (3.12), when R increases
(with xD constant), the slope of ROL becomes steeper, i.e. (R/R+1) and the intercept (x D/R+1) decreases.
The ROL therefore rotates around the point (xD, xD). The reflux ratio may be any value between a
minimum value and an infinite value. The limit is the minimum reflux ratio (result in infinite stages) and
the total reflux or infinite reflux ratio (result in minimum stages). With x D constant, as R decreases, the
slope (R/R+1) of ROL (Equation (3.12)) decreases, while its intercept (xD/R+1) increases and rotates
69
upwards around (xD, xD) as shown in Figure 24. The ROL moves closer to the equilibrium curve as R
decreases until point Q is reached. Point Q is the point of intersection between the q-line and the
equilibrium curve.
At this point of intersection the driving force for mass transfer is zero. This is also called as Pinch Point.
At this point separation is not possible. The R cannot be reduced beyond this point. The value of R at
this point is known as the minimum reflux ratio and is denoted by Rmin.
70
4.0 ABSORPTION
Gas absorption: It is a mass transfer operation in which one or more gas solutes is removed by
dissolution in a liquid. The inert gas in the gas mixture is called “carrier gas”. In the absorption process
of ammonia from air-ammonia mixture by water, air is carrier gas, ammonia is „solute” and water is
absorbent. An intimate contact between solute gas and absorbent liquid is achieved in suitable
absorption equipment, namely, tray tower, packed column, spray tower, venture scrubber, etc.
Desorption or stripping operation is the reverse of absorption. Absorption operation is of two types;
physical and chemical.
solute+carrier gas⃗
absorbent solute absorbed ∈absorbent +carrier gas
71
iii. Viscosity: For better absorption, a solvent of low viscosity is required. In mechanically agitated
absorber, greater amount of power is required for high viscous solvent and flooding is also
caused at lower liquid and gas flow rates.
iv. Corrosiveness: Non-corrosive or less corrosive solvent reduces equipment construction cost as
well as maintenance cost.
v. Cost: The solvent should be cheap so that losses will be insignificant and should be easily
available.
vi. Toxicity and Hazard: The solvent should be non-toxic, non-flammable, non-hazardous and
should be chemically stable.
Steam is generally used in desorption or stripping medium as stripped solute can be recovered very
easily by condensing steam leaving desorption tower.
Two common gas absorption equipments are packed tower and plate tower. Other absorption
equipments are, namely, spray column, agitated contactor, venture scrubber, etc. The gas and the liquid
phases come in contact in several discrete stages. Thus, a stage wise contact is there in a plate column.
But in packed tower, the up-flowing gas remains in contact with down-flowing liquid throughout the
packing, at every point of the tower. Therefore, packed tower is known as “continuous differential
contact equipment It is different from the stage-wise distillation column. In the stage distillation column
the equilibrium in each stage will vary not in a continuous fashion whereas in the packed column the
equilibrium is changed point wise in each axial location.
72
Equilibrium data, gas and liquid flow rates, solute concentration in two terminals, individual and overall
volumetric mass transfer coefficients should be known for the design of a packed absorption tower.
Packing Materials
Packing materials are utilized to provide large interfacial area of contact between two phases. These are
made from either of ceramics, metals or plastics. A number of packing materials with various size,
shape and performance are available. These are classified into three types, namely, dumped or random,
structured and grid.
(A) Dumped or random packing materials: Dumped or random packing materials are classified into
three categories as first generation (1907 to mid 1950); second generation (mid 1950 to mid 1970) and
third generation (mid 1970 to till date). The first generation random packing materials are of three
categories, such as, (a) Raschig rings; (b) Lessing rings and modified Raschig rings and (c) Berl saddles.
These are shown in Figure 9.
73
(a) Raschig rings; (b) Lessing rings and (c) Berl saddle
modified Raschig rings
(Cross-partition rings)
Figure 25 First generation dumped or random packing materials
The second generation random packing materials are mainly (a) Intalox saddle and modification; (b)
Pall ring and modification. Intalox saddle is the modified version of Berl saddle and offers less friction
resistance due to particular shape (two saddles will never nest). Pall rings are modified version of
Raschig rings.
The third generation random packing materials are numeral; (a) Intalox Metal Tower Packing (IMTP);
(b) Nutter ring; (c) Cascade Mini-Ring (CMR); (d) Jaeger Tripac; (e) Koch Flexisaddle; (f) Nor-Pac; (g)
Hiflow ring, etc. These are shown in Figure 27.
74
(a) IMTP (b) Nutter ring (c) CMR
(d) Jaeger Tripac
(B) Structured packing materials: These materials are used widely as packing materials in packed
tower due to low gas pressure drop and improved efficiency. Corrugated metal sheet structured
packing and Wire mesh structured packing materials are widely used in the industries. These
include Mellapak, Flexipak, Gempak, Montz and MaxPak. These are shown in Figure 28.
(C) Grid packing materials: This packing material is used for high gas or vapor capacities at low
75
pressure drop. Mellagrid series, Flexigrid series, Snap grid series are among these grids.
A substantial number of industrial operations in which the compositions of solutions and or mixtures are
changed involve interphase mass transfer. Typical examples of such operations could include
i. The transfer of a solute from the gas phase into a liquid phase as encountered in absorption,
dehumidification, and distillation
ii. The transfer of a solute from the liquid phase into a gas phase, as encountered in desorption or
stripping and humidification
iii. The transfer of a solute from one liquid phase into a second immiscible liquid phase (such as the
transfer from an aqueous phase to a hydrocarbon phase), as encountered in liquid–liquid
extraction
iv. The transfer of a solute from a solid into a fluid phase as encountered in drying and leaching
v. The transfer of a solute from a fluid onto the surface of a solid as encountered in adsorption and
ion exchange.
Mass-transfer operations are commonly encountered in towers or tanks that are designed to provide
intimate contact of the two phases. This equipment may be classified into one of the four general types
according to the method used to produce the interphase contact.
Many varieties and combinations of these types exist or are possible; we will restrict our discussion to
the major classifications.
Bubble towers consist of large open chambers through which the liquid phase flows and into which the
gas is dispersed into the liquid phase in the form of fine bubbles.
76
The small gas bubbles provide the desired contact area. Mass transfer takes place both during the bubble
formation and as the bubbles rise up through the liquid. The rising bubbles create mixing action within
the liquid phase, thus reducing the liquid-phase resistance to mass transfer.
Bubble towers are used with systems in which the liquid phase normally controls the rate of mass
transfer; for example, it is used for the absorption of relatively insoluble gases, as in the air oxidation of
water. Figure above illustrates the contact time and the direction of phase flow in a bubble tower. As one
would expect, the contact time, as well as the contact area, plays an important role in determining the
amount of mass transferred between the two phases. The basic mass-transfer mechanism involved in
bubble towers is also encountered in batch bubble tanks or ponds where the gas is dispersed at the
bottom of the tanks. Such equipment is commonly encountered in biological oxidation and in
wastewater-treatment operations.
Here gas is dispersed into bubbles or foams. In this case chemical reaction between the dissolved gas
and constituent of the liquid is required.
Applications
Carbonation of lime slurry
Chlorination of paper stock
Hydrogenation of vegetable oils
77
Aeration of fermentation broths e.g. production of penicillin, aeration of activated sludge for
biological oxidation
This is a vertical tower in which the liquid and gas are contacted in stepwise fashion on trays or plates.
The liquid enters at the top and flows downward by gravity. On the way, it flows across each tray and
through a downspout to the tray below. The gas passes upward through openings of one sort or another
in the tray, then bubbles through the liquid to form a froth, disengages from the froth and passes on to
the next tray above.
Illustration
78
The overall effect is a multiple countercurrent contact of gas and liquid, although each tray is
characterized by across flow of the two. Each tray of the tower is a stage since on the tray the fluids are
brought into intimate contact, interphase diffusion occurs and fluids are separated.
-End-
79