You are on page 1of 15

LECTURE 1: Introduction and Basic Concepts

Syllabus:

Transfer of Momentum, Heat Transfer (Conduction, Convection, Radiation),


Mass Transfer, Transport Phenomena Approach.

What is Transport Phenomena?

Transport Phenomena is the study of the transfer of momentum, heat and mass. These
three "modes" of transport are usually taught/grouped together because:
they have similar molecular origins
they yield similar governing equations/principles
they often occur simultaneously
they require similar mathematical/conceptual tools

Transport Phenomena and Unit Operations

The pioneering work Principles of Chemical Engineering was published in 1923 under the
authorship of Walker, Lewis, and McAdams [1]. This book was the first to emphasize the
concept of unit operations as a fundamental approach to physical separations such as distillation,
evaporation, drying, etc.
This was the era when the profession of chemical engineering matured into a separate area, no
longer the province of the industrial chemist. The study of unit operations such as distillation is
predicated on the idea that similarities in equipment and fundamentals exist regardless of the
process. In other words, the principles of distillation apply equally to the separation of liquid
oxygen from liquid nitrogen as well as to the thousands of other distillations routinely carried out
in industries around the world. The study of transport phenomena is undertaken because this
topic is the basis for most of the unit operations. In simple terms, transport phenomena comprise
three topics: heat transfer, mass transfer, and momentum transfer (fluid flow). In many of the unit
operations (such as distillation), all three transport phenomena (i.e., fluid flow, heat transfer, and
mass transfer) occur, often simultaneously. The concepts presented in transport phenomena
underlie the calculation procedures that are used in the design of unit operations.

Equilibrium and Rate Processes

Many problems can conveniently be divided into two classifications: equilibrium and nonequilibrium. Under conditions of non-equilibrium, one or more variables change with time. The
rates of these changes are of much interest, naturally. A typical chemical engineer is involved
with four types of rate processes: rate of heat transfer, rate of mass transfer, rate of momentum
transfer, and rate of reaction. The first three of these are the subject of this course. The fourth,
rate of reaction, will not be covered in any detail, except for the inclusion of the appropriate
terms in the general equations and in a few elementary examples.

Transfer of momentum

In Fig. 1 a pair of large parallel plates is shown, each one with area A, separated by a distance Y.
In the space between them is a fluid-either a gas or a liquid. This system is initially at rest, but at
time t = 0 the lower plate is set in motion in the positive x direction at a constant velocity V. As
time proceeds, the fluid gains momentum, and ultimately the linear steady-state velocity profile
shown in the figure is established. We require that the flow be laminar ("laminar" flow is the
orderly type of flow that one usually observes when syrup is poured, in contrast to "turbulent"
flow, which is the irregular, chaotic flow one sees in a high-speed mixer). When the final state of
steady motion has been attained, a constant force F is required to maintain the motion of the
lower plate. Common sense suggests that this force may be expressed as follows:
F
V

A
Y

(1)

That is, the force should be proportional to the area and to the velocity, and inversely
proportional to the distance between the plates. The constant of proportionality is a property
of the fluid, defined to be the viscosity.
We now switch to the notation that will be used throughout the book. First we replace F/A by the
symbol yx, which is the force in the x direction on a unit area perpendicular to the y direction. It
is understood that this is the force exerted by the fluid of lesser y on the fluid of greater y.
Furthermore, we replace V/Y by v x y . Then, in terms of these symbols, Eq. (1) becomes
yx

v x
y

(2)

This equation, which states that the shearing force per unit area is proportional to the negative of
the velocity gradient, is often called Newton's law of viscosity. (Actually we should not refer Eq.
(2) as a "law," since Newton suggested it as an empiricism

Fig.1. The buildup to the steady, laminar velocity profile for a fluid contained between two
plates. The flow is called "laminar" because the adjacent layers of fluid ("laminae") slide past
one another in an orderly fashion.
the simplest proposal that could be made for relating the stress and the velocity gradient.
However, it has been found that the resistance to flow of all gases and all liquids with molecular
weight of less than about 5000 is described by Eq. (2), and such fluids are referred to as
Newtonian fluids. Polymeric liquids, suspensions, pastes, slurries, and other complex fluids are
not described by Eq. (2) and are referred to as non-Newtonian fluids.
Equation (2) may be interpreted in another fashion. In the neighborhood of the moving solid
surface at y = 0 the fluid acquires a certain amount of x-momentum. This fluid, in turn, imparts
momentum to the adjacent layer of liquid, causing it to remain in motion in the x direction.
Hence x-momentum is being transmitted through the fluid in the positive y direction. Therefore
yx may also be interpreted as the flux of x-momentum in the positive y direction, where the term
"flux" means "flow per unit area." This interpretation is consistent with the molecular picture of
momentum transport and the kinetic theories of gases and liquids. It also is in harmony with the
analogous treatment given later for heat and mass transport.
The above idea may be paraphrased by saying that momentum goes "downhill" from a region of
high velocity to a region of low velocity-just as a sled goes downhill from a region of high
elevation to a region of low elevation, or the way heat flows from a region of high temperature to
a region of low temperature. The velocity gradient can therefore be thought of as a "driving
force" for momentum transport.
In what follows we shall sometimes refer to Newton's law in Eq. (2) in terms of forces (which
emphasizes the mechanical nature of the subject) and sometimes in terms of momentum
transport (which emphasizes the analogies with heat and mass transport).
This dual viewpoint should prove helpful in physical interpretations.
Often the symbol represents the viscosity divided by the density (mass per unit volume) of the
fluid, thus:

(3)

This quantity is called the kinematic viscosity. The appropriate SI units for and are as
follows: [] = Pas , [] = m2s-1.

Heat Transfer

Internal energy may be viewed as the sum of the kinetic and potential energies of the molecules.
The portion of the internal energy of a system associated with the kinetic energy of the molecules
is called sensible energy or sensible heat. The internal energy is also associated with the
intermolecular forcers between the molecules of a system. These are the forcers that bind the
molecules to each other; they are strongest in solids and weakest in gases. If sufficient energy is
added to the molecules of a solid or liquid, they will overcome these molecular forcers and
simply break away, turning the system into a gas. This is a phase change process and because of
this added energy, a system in the gas phase is at a higher internal energy level than it is in the
solid or the liquid phase. The internal energy associated with the phase of a system is called
latent energy or latent heat.

The sensible and latent forms of internal energy can be transferred from one medium to another
as a result of a temperature difference, and are referred to as heat or thermal energy. Thus a heat
transfer is the exchange of the sensible and latent forms of internal energy between two mediums
as a result of a temperature difference.
[W]. The rate of heat
The amount of heat transferred per unit time is called heat transfer rate, Q
-2
transfer per unit area is called heat flux, q Q A [W m ]
The First Law of Thermodynamics manifests the principle of the conservation of energy:
Change
Total energy
Total energy

leaving in the total energy


entering
the system
the system

of the system

(4)

In heat transfer analysis, we are usually interested only in the form of energy that can be
transferred as a result of a temperature difference, that is, heat or thermal energy. In such cases it
is convenient to write a heat balance and to treat the conversion of nuclear, chemical,
mechanical, and electrical energies into thermal energy as heat generation. The energy balance in
that case can be expressed as
Qin Qout E gen Ethermal energy
(5)
When a stationary closed system involves heat transfer only and no work interactions across its
boundary, the energy balance relation reduces to
Q mcv T

(6)
Under steady conditions and in the absence of any work interactions, the conservation of energy
relation for a control volume with one inlet and one exit with negligible changes in kinetic and
potential energies can be expressed as
m
c p T
Q

(7)

-1

[kgs ] is the mass flow rate.


where m
Heat can be transferred in three different modes: conduction, convection, and radiation.
Conduction is the transfer of heat from the more energetic particles of a substance to the adjacent
less energetic ones as a result of interactions between the particles, and is expressed by Fouriers
law of heat conduction as
dT
Q cond k
,
dx

or, in general case

q kT

(8)
(8a)

Convection is the mode of heat transfer between a solid surface and the adjacent liquid or gas
that is in motion, and involves the combined effects of conduction and fluid motion. The rate of
convection heat transfer is expressed by Newtons law of cooling as

Q
convection hAs Ts T

(9)

where h [Wm-2K-1] is the convection heat transfer coefficient.


Radiation is the energy emitted by matter in the form of electromagnetic waves (or photons) as a
result of the changes in the electronic configurations of the atoms and molecules. The maximum
rate of radiation that can be emitted from a surface at a thermodynamic temperature Ts is given
by the Stefan-Boltzmann law as
4

Q
(10)
emitt ,max AsTs
where = 5.6710-8 Wm-2K-4 is the Stefan-Boltzmann constant.
When a surface of emissivity and surface area As at a temperature Ts is completely enclosed by
a much larger (or black) surface at temperature Tsurr separated by a gas (such as air) that does not
intervene with radiation, the net rate of radiation heat transfer between these two surfaces is
given by
4
Q rad As Ts4 Tsurr
(11)
Ii this (limited) case, the emissivity () and the surface area of the surrounding surface do not
have any effect on the net radiation heat transfer.
The rate at which a surface absorbs radiation is determined from

Q
absorbed Qincident

(12)
Where is absorptivity of the surface.
Radiation is usually considered to be a surface phenomenon for solids that are opaque to thermal
radiation such as metals, wood, and rocks since the radiation emitted by interior region of such
material can never reach the surface, and the radiation incident on such bodies is usually
absorbed within a few microns from the surface.
The heat transfer problems encountered in practice can be considered in two groups: (1) rating
and (2) sizing problems. The rating problems deal with the determination of the heat transfer rate
for an existing system at a specified temperature difference. The sizing problems deal with the
determination of the size of a system in order to translate heat at a specified rate for a specified
temperature difference.

Mass Transfer

When a system contains two or more components whose concentrations vary from point to point,
there is a natural tendency for mass to be transferred, minimizing the concentration differences
within the system and moving it towards equilibrium.
The transport of one component from a region of higher concentration to that of a lower
concentration is called mass transfer. Distinction should be made between mass transfer and the
bulk fluid flow that occurs on a macroscopic level as a fluid is transported from one location to
another.
Many of our daily experiences involve mass-transfer phenomena. The invigorating aroma of a
cup of freshly brewed coffee and the sensuous scent of a delicate perfume both reach our nostrils
from the source by diffusion through air. A lump of sugar added to the cup of coffee eventually
dissolves and then diffuses uniformly throughout the beverage. Laundry hanging under the sun
during a breezy day dries fast because the moisture evaporates and diffuses easily into the
relatively dry moving air.

Mass transfer 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 covered by the term mass-transfer operationsinclude such techniques as distillation, gas absorption, humidification, liquid extraction,
adsorption, membrane separations, and others.
The driving force for transfer in these operations is a concentration gradient, much as a
temperature gradient provides the driving force for heat transfer. (N.B. Although it should be
remembered that a rigorous analysis employs the gradient of chemical potential as the driving
force for mass transfer.)

Returning to the lump of sugar added to the cup of coffee, it is evident that the time required for
the sugar to distribute uniformly depends upon whether the liquid is quiescent or whether it is
mechanically agitated by a spoon. In general, the mechanism of mass transfer depends upon the
dynamics of the system in which it occurs. Mass can be transferred by random molecular motion
in quiescent fluids, or it can be transferred from a surface into a moving fluid, aided by the
dynamic characteristics of the flow. These two distinct modes of transport, molecular mass
transfer and convective mass transfer, are analogous to conduction heat transfer and convective
heat transfer. Each of these modes of mass transfer will be described and analyzed. The two
mechanisms often act simultaneously. Frequently, when this happens, one mechanism can
dominate quantitatively so that approximate solutions involving only the dominant mode can be
used.
The description of diffusion involves a mathematical model based on a fundamental hypothesis
or law. Interestingly, there are two common choices for such a law. The more fundamental,
Ficks law of diffusion, uses a diffusion coefficient. The second, which has no formal name,
involves a mass transfer coefficient.
Now we want to illustrate the two basic ways in which diffusion can be described. To do this, we
first imagine two large bulbs connected by a long thin capillary (Fig.2). The bulbs are at constant
temperature and pressure and are of equal volumes. However, one bulb contains carbon dioxide,
and other is filled with nitrogen. To find out how fast these two gases will mix, we measure the
concentration of carbon dioxide in the bulb that initially contains nitrogen. We make these
measurements when only a trace of carbon dioxide has been transferred, and we find that the
concentration of carbon dioxide varies linearly with time. From this, we know the amount
transferred per unit time.

Fig. 2. A simple diffusion experiment. Two bulbs initially containing different gases are
connected with a long thin capillary. The change of concentration in each bulb is a measure of
diffusion and can be analyzed in two different ways.
There are two basic ways in which diffusion (mass transfer) can be described.
Firstly, imagine two large bulbs connected by a long capillary (Fig.2). The bulbs are at constant
temperature and are of equal volumes. However, one bulb contains carbon dioxide, and the other
is filled with nitrogen. To find how fast these two gases will mix, we measure the concentration
of carbon dioxide in the bulb that initially contains nitrogen. We make these measurements when
only a trace of carbon dioxide has been transferred, and we find that the concentration of carbon
dioxide varies linearly with time. From this, we know the amount transferred per unit time. We
want to analyze this amount transferred to determine physical properties that will be applicable
not only to this experiment but also in other experiments. To do this, we first define the flux:

carbon dioxide flux

amount of gas removed


time area capillary

(13)

In other words, if we double the cross-sectional area, we expect the amount transported in
double. Defining the flux in this way is a first step in removing the influences of our particular
apparatus and making our results more general. We next assume that the flux is proportional to
the gas concentration:
carbon dioxide

carbon dioxide flux k concentration

difference

(14)

The proportionality constant k is called a mass transfer coefficient. Its introduction signals one of
the two basic models of diffusion.
Alternatively, we can recognize that increasing the capillarys length will decrease the flux, and
we can assume that

carbon dioxide flux D

carbon dioxide concentration difference

capillary length

(15)

The new proportionality constant D is the diffusion coefficient. Its introduction implies the other
model for diffusion, the model often called Ficks law.
These assumptions may seem arbitrary, but they are similar to those made in many other
branches of science. For example, they are similar to those used in developing Ohms law, which
states that

voltage, or
1

area times flux


potential
resistance

of electrons
difference

current, or

(16)

Thus, the mass transfer coefficient k is analogous to the reciprocal of the resistance. An
alternative form of Ohms law is
current density

potential difference
1

or flux of

length
resistivity

electrons

(17)

The diffusion coefficient D is analogous to the reciprocal of resistivity.


Neither the equation using the mass transfer coefficient k nor that using the diffusion coefficient
D is always successful. This is because of the assumptions made in their development. For
example, the flux may not be proportional to the concentration difference if the capillary is very
thin or if the two gases react. In the same way, Ohms law is not always valid at very high
voltages. But these cases are exceptions; both diffusion equations work well in most practical
situations, just as Ohms law does.
The parallels with Ohms law also provide a clue about how the choice between diffusion
models is made. The mass transfer coefficient in Eq. (14) and the resistance in Eq. (16) are
simpler, best used for practical situations and rough measurements. The diffusion coefficient in
Eq. (15) and the resistivity in Eq. (17) are more fundamental, involving physical properties like
those found in handbooks.
How these differences guide the choice between the two models?
The choice between the two models represents a compromise between ambition and
experimental resources. Obviously, we would like to express our results in the most general and
fundamental ways possible. This suggests working with diffusion coefficients. However, in many
cases our experimental measurements will dictate a more approximate and phenomenological
approach. Such approximations often imply mass transfer coefficients, but they usually still
permit us to reach our research goals.
This choice and the resulting approximations are best illustrated by two examples. In the first,
we consider hydrogen diffusion in metals. This diffusion substantially reduces a metals ductility,
so much so that parts made from the embrittled metal frequently fracture. To study this
embrittlement, we might expose the metal to hydrogen under a variety of conditions and measure
the degree of embrittlement versus these conditions. Such empiricism would be a reasonable first
approximation, but it
would
quickly flood us with
uncorrelated
information that would
be difficult
to use effectively.

Fig. 3. Hydrogen diffusion into a metal. This process can be described with either a mass transfer
coefficient k or a diffusion coefficient D. The description with a diffusion coefficient correctly
predicts the variation of concentration with position and time, and so is superior.
As an improvement, we can undertake two sets of experiments. First, we can saturate metal
samples with hydrogen and determine their degrees of embrittlement. Thus we know metal
properties versus hydrogen concentration. Second, we can measure hydrogen uptake versus time,
as suggested in Fig. 3, and correlate our measurements as mass transfer coefficients. Thus we
know average hydrogen concentration versus time. To our dismay, the mass transfer coefficients
in this case will be difficult to interpret. They are anything but constant. At zero time, they
approach infinity; at large time, they approach zero. At all times, they vary with the hydrogen
concentration in the gas surrounding the metal. They are inconvenient way to summarize our
results. Moreover, the mass transfer coefficients give only the average hydrogen concentration in
the metal. They ignore the fact that the hydrogen concentration very near metals surface will
reach saturation but the concentration deep within the bar will remain zero. As a result, the metal
near the surface may be very brittle but that within may be essentially unchanged.
We can include these details in the diffusion model that assumes that

hydrogen

flux D

hydrogen concentration at z 0 hydrogen concentration at z l


(18)
thickness at z l thickness at z 0

or, symbolically,
j1 D

c1 z 0 c1 z l
l 0

(19)
Where the subscript 1 symbolizes the diffusing species. In this equations, the distance l is that
over which diffusion occurs. In the case described in Fig.2, the length of the capillary was
appropriately this distance; but in this case, it seems uncertain what distance should be taken into
consideration. If we assume that it is very small,
j1 D lim
l 0

c1 z z c1 z z l
z z l z z

d c1
dz

(20)
We can use this relation and the techniques developed later in this course to correlate our
experiments with only one parameter, the diffusion coefficient D. We then can correctly predict
the hydrogen uptake versus time and the hydrogen concentration in the gas. Moreover, we get the
hydrogen concentration at all positions and times within the metal.

Thus the model based on the diffusion coefficient gives results of more fundamental value than
the model based on mass transfer coefficients. In mathematical terms, the diffusion model is said
to have distributed parameters, for the dependent variable (the concentration) is allowed to vary
with all independent variables (like position and time). In contrast, the mass transfer model is
said to have lumped parameters (like the average hydrogen concentration in metal).

Fig.4. Rates of drug dissolution. In this case, describing the system with a mass transfer
coefficient k is best because it easily correlates the solutions concentration versus time.
Describing the system with a diffusion coefficient D gives a similar correlation but introduces an
unnecessary parameter, the film thickness l.
These results would appear to imply that the diffusion model is superior to the mass transfer
model and so should always be used. However, in many interesting cases the models are
equivalent. To illustrate this, imagine that we are studying the dissolution of a solid drug
suspended in water, as schematically suggested by Fig. 4. The dissolution of this drug is known
to be controlled by the diffusion of the dissolved drug away from the solid surface of the
undissolved material. We measure the drug concentration versus time as shown, and we want to
correlate these results in terms of as few parameters as possible.
One way to correlate the dissolution results is to use a mass transfer coefficient. To do this, we
write a mass transfer balance on the solution:
accumulation

total rate of
of drug in
dissolution

solution

d
V c1 V dc1 A j1 A k c1( sat ) c1
dt
dt

(21)
where V is the volume of solution, A is the total area of the drug particles, c1(sat) is the drug
concentration at saturation and at the solids surface, and c1 is the concentration in the bulk
solution. Integrating this equation allows quantitatively fitting our results with one parameter, the
mass transfer coefficient k. This quantity is independent of drug solubility, drug area, and

solution volume, but it does vary with physical properties like stirring rate and solution viscosity.
Correlating the effects of these properties turns out to be straightforward.
The alternative to mass transfer is diffusion theory, for which the mass balance is
d
V c1 V dc1 A j1 A D c1( sat ) c1
(22)
dt
dt
l
in which l is an unknown parameter, equal to the average distance across which diffusion occurs.
This unknown, called a film or unstirred layer thickness, is a function not only of flow and
viscosity but also of the diffusion coefficient itself.
Equations (21) and (22) are equivalent, and they share the same successes and shortcomings. In
the former, we must determine the mass transfer coefficient experimentally; in the latter, we
determine instead the thickness l. The choice between the mass transfer and diffusion models is
often a question of taste rather than precision. The diffusion model is more fundamental and is
appropriate when concentrations are measured or needed versus both position and time. The
mass transfer model is simpler and more approximate and is especially useful when only average
concentrations are involved.
The additional examples given below should help us decide which model is appropriate for our
purposes. (This choice is often difficult, a junction where many have trouble).
Example 1: Ammonia scrubbing. Ammonia, the major material for fertilizer, is made by reacting
nitrogen and hydrogen under pressure. The product gas can be washed with water to dissolve the
ammonia and separate it from unreacted gases. How can you correlate the dissolution rate of
ammonia during washing?
Concept of solution: The easiest way is to use mass transfer coefficients. If you use diffusion
coefficients, you must somehow specify the distance across which diffusion occurs. This distance
is unknown unless the detailed flows of gases and the water are known; they rarely are.
Example 2: Reaction in porous catalysts. Many industrial reactions use catalysts containing
small amount of noble metals dispersed in a porous inert material like silica. The reaction on
such a catalyst are sometimes slower in large pellets than in small ones. This is because the
reagents take longer to diffuse into the pellet than they do to react. How should you model this
effect?
Concept of solution: You should use the diffusion coefficients to describe the simultaneous
diffusion and reaction in the pores in the catalyst. You should not use mass transfer coefficients
because you cannot easily include effect of reaction.
Example 3: Corrosion of marble. Industrial pollutants in urban areas like Venice, Athens etc.
cause significant corrosion of marble statues. You want to study how these pollutants penetrate
marble. Which diffusion model should you use?
Concept of solution: The model using diffusion coefficients is the only one that will allow you to
predict concentration versus position in the marble. The model using mass transfer coefficients
will only correlate how much pollutant enters the statue, not what happens to the pollutant.
Example 4: Protein size in solution. You are studying a variety of proteins that you hope to purify
and use as food supplements. You want to characterize the size of the proteins in solution. How
can you use diffusion to do this?

Concept of solution: Your aim is determining the molecular size of the protein molecules. You
are not interested in the protein mass transfer except as a route to these molecular properties. As
a result, you should measure the proteins diffusion coefficient, not its mass transfer coefficient.
The proteins diffusion coefficient will turn out to be proportional to its radius in solution.
Example 5: Antibiotic production. Many drugs are made by fermentations in which
microorganisms are grown in a huge stirred what of a dilute nutrient solution or beer. Many of
these fermentations are aerobic, so the nutrient solution requires aeration. How should you model
oxygen uptake in the type of solution?
Concept of solution: Practical models use mass transfer coefficients. The complexities of the
problem, including changes in air bubble size, flow effects of the non-Newtonian solution, and
foam caused by biological surfactants, all inhibit more careful study.
Example 6: Facilitated transport across membranes. Some membranes contain a mobile carrier,
a reactive species that reacts with diffusing solutes, facilitating their transport across the
membrane. Such membranes are used to concentrate copper ions from industrial waste and to
remove carbon dioxide from coal gas. Similar membranes are believed to exist in the human
intestine and liver. Diffusion across these membranes does not vary linearly with the
concentration difference across them. The diffusion can be highly selective, but it is often easily
poisoned. Should this diffusion be described with mass transfer coefficients or with diffusion
coefficients?
Concept of solution: This system includes not only diffusion but also chemical reaction.
Diffusion and reaction couple in a nonlinear way to give the unusual behavior observed.
Understanding such behavior will certainly require the more fundamental model of diffusion
coefficients.
Example 7: Flavor retention. When food products are spray-dried, they lose a lot of flavor.
However, they lose less than would be expected on the basis of the relative vapor pressures of
water and the flavor compounds. The reason apparently is that the drying food often forms a tight
gel-like skin across which diffusion of the flavor compounds is inhibited. What diffusion model
should you use to study this effect?
Concept of solution: Because spray drying is a complex, industrial-scale process, it is usually
modeled using mass transfer coefficients. However, in this case you are interested in the
inhibition of diffusion. Such inhibition will involve the sizes of pores in the food and of
molecules of the flavor compounds. Thus you should use the more basic diffusion model, which
includes these molecular factors.
Example 8: The smell of marijuana. Recently, a large shipment of marijuana was seized in the
Minneapolis-St. Paul airport. The police said their dog smelled it. The owners claimed that it was
too well wrapped in plastic to smell and that the police had conducted an illegal search without a
search warrant. How could you tell who was right?
Concept of solution: In this case, you are concerned with the diffusion of odor across the thin
plastic film. The diffusion rate is well described by either mass transfer or diffusion coefficients,
However, the diffusion model explicitly isolates the effect of the solubility of the smell in the
film, which dominates the transport. This solubility is the dominant variable. In this case, the
search was illegal.

Example 9: Scale-up of wet scrubbers. You want to use a wet scrubber to remove sulfur oxides
from the flue gas of a large power. A wet scrubber is essentially a large piece of pipe set on its
end and filled with inert ceramic material. You pump the flue gas up from the bottom of the pipe
and pour a lime slurry down from the top. In the scrubber, there are various reactions, such as:
CaO + SO2 CaSO3
The time reacts with the sulfur oxides to make in insoluble precipitate, which is discarded. You
have been studying a small unit and want to use these results to predict the behavior of a larger
unit. Such an increase in size is called a scale-up. Should you make these predictions using a
model based on diffusion or mass transfer coefficients?
Concept of solution: This situation is complex because of the chemical reactions and the
irregular flows within the scrubber. Your first try at correlating your data should be a simple
model based on mass transfer coefficients. Should these correlations prove unreliable, you may
be forced to use the more difficult diffusion model.
When facing the mass transfer operation problem, keep both models in mind. People involved in
basic research tend to be overcommitted to diffusion coefficients, whereas those with broader
objectives tend to emphasize mass transfer coefficients. Each group should recognize that the
other has a complimentary approach that may be helpful for the case in hand.
The purpose of the equipment used for mass-transfer operations is to provide intimate contact of
the immiscible phases in order to permit interphase diffusion of the constituents. The rate of
mass transfer is directly dependent upon the interfacial area exposed between the phases, and the
nature and degree of dispersion of one phase into the other are therefore of prime importance.
The operations which include humidification and dehumidification, gas absorption and
desorption, and distillation all have in common the requirement that a gas and a liquid phase be
brought into contact for the purpose of diffusional interchange between them. The equipment for
gas-liquid contact can be broadly classified according to whether its principal action is to
disperse the gas or the liquid, although in many devices both phases become dispersed.

Transport Phenomena Approach (Transport Phenomena-II)

If we introduce a general variable the conservative form for all fluid flow equations, including
equations for scalar quantities such as temperature and species concentration etc., can usually be
written in the following form:
(23)

In

words,

Eqn. (23) is the so-called transport equation for property . It clearly highlights the various
transport processes: the rate of change term and the convective term on the LHS and the diffusive
term ( = diffusion coefficient) and the source term respectively in the RHS.
The integration of Eqn. (23) over a three-dimensional control volume (CV) yields:

(24)

In

The
energy

words:

following
equation:

(25)
or

is

the

Here we are interested in a stationary volume element, fixed in space, through which a fluid is
flowing (control volume). Both kinetic energy and internal energy may be entering and leaving the
system by convective transport. Heat may enter and leave the system by heat conduction as well.
Heat conduction is fundamentally a molecular process. Work may be done on the moving fluid by the
stresses, and this, too, is a molecular process. This term includes the work done by pressure forces
and by viscous forces. In addition, work may be done on the system by virtue of the external forces,
such as gravity.

References:
1. Walker, W. H., W. K. Lewis, and W. H. McAdams: Principles of Chemical Engineering,McGrawHill, New York, 1923.
2. Y. A. engel, A. J. Ghajar . Heat and Mass Transfer. Fundamentals and Applications. 4 th Ed.,
McGraw-Hill.
3. E.L. Cussler. Diffusion Mass Transfer in Fluid Systems. 2nd Ed., Cambridge University Press.
4. R.B. Bird, W.E. Stewart, E.N. Lightfoot. Transport Phenomena. 2nd Ed., John Wiley & Sons.

You might also like