You are on page 1of 16

Air Pollution Control

Absorption

Absorption ; Gas absorption ; Gas washing

Gas absorption is a unit operation by which a gas mixture is contacted with a liquid for
the purposes of preferentially dissolving one or more components from the gas in the liquid
and to provide a solution of these in the liquid. For example:
• the gas from by-product coke ovens is washed with water to remove ammonia and
again with oil to remove benzene and toluene vapours;
• offensive (objectionable) hydrogen sulphide is removed from such a gas or from
natural hydrocarbon gases by washing with various alkaline solutions in which it is
absorbed;
• valuable solvent carried by a gas stream may be recovered for reuse by washing the
gas with an appropriate solvent for the vapours.
Such operations require mass transfer of a substance from the gas stream to the liquid one.
When mass transfer occurs in the opposite direction, i.e. from the liquid to the gas, the
operation is called desorption, or stripping. For example, the benzene and toluene are
removed from the absorption oil mentioned above by contacting the liquid solution with
steam, whereupon the vapours enter the gas stream and are carried away, and the absorption
oil may be used again. The principle of both absorption and desorption are basically the same,
and we may study both operations at the same time.
Ordinarily, these operations are used only for solute recovery and solute removal.
Separation of solutes from each other requires the fractional techniques of distillation.
The rate at which a gaseous constituent will dissolve in an absorbent liquid depends
upon the departure from equilibrium which is maintained, and therefore it is necessary to
consider the equilibrium characteristics of gas-liquid systems. A very brief discussion willl be
given here.

Equilibrium
Suppose a fixed closed container partly filled with water and above the water with a
gaseous mixture of ammonia and air. The whole system is maintained at constant temperature
and pressure. Since ammonia is very soluble in water, some ammonia molecules will instantly
transfer from the gas into the liquid, crossing the interfacial surface separating the two phases.
A portion of ammonia molecules escapes back into the gas, at a rate proportional to their
concentration (number) in the liquid. As more ammonia enters the liquid, its concentration is
increasing within the liquid, consequently the rate at which ammonia returns to the gas
increases also, until eventually the rate at which it enters the liquid exactly will be equal to
that at which it leaves. At the same time, through the mechanism of diffusion, the
concentrations throughout each phase become uniform. A dynamic equilibrium now exists,
and while ammonia molecules continue to transfer back and forth one phase to the other, the
net transfer falls zero. The concentrations within each phase no longer change. To the
observer who can not see the individual molecules the diffusion has apparently stopped.
If now we inject additional ammonia into the container, increasing the concentration
of ammonia in the gas phase, a new set of equilibrium will eventually be established, with
higher concentrations in each phase than were earlier. In this manner we can obtain the
complete relationship between the equilibrium concentrations in both phases. If the ammonia
is designated as substance A, the equilibrium concentrations in the gas and liquid, y*A and xA
mole fractions, respectively, give the equilibrium-distribution curve of the type shown in
Figure 1.
Figure 1.

1/16
Dr. Parti Mihály; 463-2635
Department of Fluid Mechanics
26/10/2011
Air Pollution Control
Absorption

This curve does not depend on the amounts of water and air that we start with and is
influenced only by the conditions, such as temperature and pressure, imposed upon the three-
component system. It is important to note that at equilibrium the concentrations of the two
phases are not equal (remark: but chemical potential of the ammonia is the same in both
phases, consequently it is equality of chemical potentials, not concentrations, which cause the
net transfer of solute to stop).
The curve of Figure 1 does not show all the equilibrium concentrations existing within
the system. For example, the water will partially vaporize into the gas phase, the components
of the air will also dissolve to a small extent in the liquid, and equilibrium concentrations for
these substances will also be established. We do not consider these equilibria at present, since
they have minor importance to the present discussion. Obviously also, other concentration
units than mole fractions may be used to describe the equilibria.
Generally speaking, whenever a substance is distributed between two insoluble phases,
a dynamic equilibrium of this type can be established. Of course, the various equilibria are
peculiar to the particular system investigated. For example, replacement of the water in the
example considered above with another liquid such as benzene, or with a solid adsorbent such
as activated carbon, or replacement of the ammonia with another solute such as e.g. sulphur
dioxide will each result in new curves not at all related to the first.
Two-component system: If a quantity of single gas and a relatively non-volatile
liquid are brought to equilibrium in the manner described earlier, the resulting concentration
of dissolved gas in the liquid is said to be the gas solubility (the solubility of the gas) at the
prevailing temperature and pressure. At fixed temperature, the solubility concentration will
increase with partial pressure of the component in the manner, for example, of curve A in
Figure 2 which shows the solubility of ammonia in water at 30 C°.
Figure 2.
Different liquids and gases yield separate solubility curves, which must ordinarily be
determined experimentally for each system If the equilibrium partial pressure of a gas at a
given liquid concentration is high, as in the case of curve B (sulphur dioxide for 10 C°), the
gas is said to be relatively insoluble in the liquid, while if it is low, as for curve C (hydrogen
chloride for 10 C°), the solubility is said to be high.
The solubility of any gas is influenced by the temperature and by the total pressure. In
most cases, the solubility of a gas decreases with increasing temperature. As an example,
curve A for ammonia in water at 10 C° lies below the corresponding curve for 30 C°. At the
boiling point of the solvent, provided that its vapour pressure is less than that of the gas or
vapour solute, the gas solubility will be zero.
Multicomponent system: If a mixture of gases is brought into contact with a liquid,
under certain conditions the equilibrium concentrations of each gas will be independent of the
others, provided, however, that the equilibrium is described in terms of the partial pressures in
the gas mixture. If all but one of the components of the gas are substantially insoluble, their
concentrations in the liquid will be so small that they can not influence the solubility of the
relatively soluble component. For example, curve A on Figure 2 will also describe the
solubility of ammonia in water when the ammonia is diluted with air, since air is so insoluble
in water, provided that the ordinate of the plot is considered as the partial pressure of
ammonia in the gas mixture.
Ideal and non-ideal liquid solutions: When the liquid phase may be considered
ideal, we may compute the equilibrium partial pressure of a gas from the solution without
experimental results. In reality there are no ideal solutions, and actual mixtures only approach
ideal solutions as a limit. Ideality would require that the molecules of the constituent be
similar in size, structure, and chemical nature. Perhaps the isomers of organic compounds
2/16
Dr. Parti Mihály; 463-2635
Department of Fluid Mechanics
26/10/2011
Air Pollution Control
Absorption

approach this condition mostly. Practically however, many solutions are so closely ideal that
for engineering applications they may be considered ideal. When the gas mixture with an
ideal liquid solution also follows the ideal-gas law, the partial pressure p*A of a solute gas A
in equilibrium with the liquid is equal to the product of its vapour pressure PA at the same
temperature and its mole fraction in the solution xA. This is Raoult’s law:
p*A = PA x A
The nature of the solvent liquid does not enter into consideration as it establishes an ideal
solution, and it follows that the solubility of a particular gas in ideal solution in any solvent is
always the same.
For not ideal liquid solutions Raoult’s law will give highly incorrect results. Line D in
Figure 2, for example, is the calculated partial pressure for ammonia in equilibrium with
water solutions at 10 C° and it clearly does not represent the experimental data. On the other
hand, the straight line E is seen to represent the 10 C° ammonia-water data very well up to
mole fractions of 0.06 in the liquid. The equation of such a line is the Henry’s law which
relates the partial pressure of the solute A in gas phase and the mole fraction of the dissolved
solute A in liquid phase solvent B:
p*A = H A x A
where HA is the Henry’s law coefficient expressed in kPa (or bar) per mole-fraction solute in
liquid. For dilute concentrations of many gases and over a fairly wide concentration range for
some gases it may be applicable with different values of H for each gas and also applicable
for the gases shown in the figure. This equation is quite useful when it can be applied, but it
should be checked experimentally to determine the accuracy with which it can be used. If
Henry’s law holds, the solubility is defined by stating the value of the constant H along with
the temperature and the solute partial pressure for which it is employed. Failure to follow
Henry’s law over wide concentration ranges may be the results of chemical interaction with
the liquid or electrolytic dissociation, as is the case with ammonia-water, or non-ideality in
the gas phase.
For quite a number of gases, Henry’s law holds very well when the partial pressure of
the solute is less than 100 kPa (1 atm). For partial pressures of the solute gas greater than 100
kPa, H is seldom independent of the partial pressure of the solute gas, and a given value of H
can be used over a narrow range of partial pressures. The less soluble gases, such as nitrogen
or oxygen in water, can be expected to follow the Henry’s law up to equilibrium partial
pressure of 1 bar, and gases of the vapour type (which are below their critical temperature)
will generally follow the law up to pressures of approximately 50 percent of the saturation
value at the prevailing temperature, provided no chemical action occurs in the liquid. There is
a strong nonlinear variation of Henry’s law coefficient with temperature. Investigation is
recommended before considering temperature extrapolations of Henry’s-law data. In any case
H must be determined experimentally. Additional data and information on the applicability of
Henry’s-law constants can be found in literature.
High values of H indicate very poor solubility of the pollutant in the liquid or, in other
words, low absorption capacity of the liquid for the pollutant. On the other hand, a very low
value for H indicates very effective solution of the pollutant in the liquid, that is, the
absorption capacity of the liquid for the pollutant is very high. To absorb a certain amount of
pollutant, a much larger amount of liquid is required with a high value of H than the case of a
system with a small value of H (see Figure 2.).
The use of Henry’s law is illustrated by the following examples.
Example 1. It is desired to calculate how much ammonia can be dissolved in water from a gas
mixture when the total pressure is 101325 Pa (1 atm), the mole fraction of the ammonia in the
gas phase is 0.03, and the temperature is 20 ºC. The value of the Henry’s coefficient is 3·105
3/16
Dr. Parti Mihály; 463-2635
Department of Fluid Mechanics
26/10/2011
Air Pollution Control
Absorption

Pa (3 bar). To apply the Henry’s law the partial pressure of the ammonia should be calculated
which by using the Dalton law is as follows:
p *A = p*NH 3 = py *A = py *NH 3 = 101325 ⋅ 0.03 = 3040 Pa
and according to Henry’s law, the mole fraction of ammonia in the liquid phase:
p*NH 3 3040
x NH 3 = = = 1.013 ⋅ 10 − 2
H NH 3 300000
Example 2. It is desired to calculate how much sulphur dioxide can be dissolved in water
from a gas mixture when the total pressure is 101325 Pa (1 atm), the mole fraction of the
sulphur in the gas phase is 0.03, and the temperature is 20 ºC. The value of the Henry’s
coefficient is 3·106 Pa (30 bar). The partial pressure of the sulphur dioxide is the same as
earlier (since the mole fraction in gas phase and total pressure are the same):
p*SO 2 = py *SO 2 = 101325 ⋅ 0.03 = 3040 Pa

and according to Henry’s law, the mole fraction of sulphur dioxide in the liquid phase:
p*SO 2 3040
x SO 2 = = = 1.013 ⋅ 10 − 3
H SO 2 3000000
To evaluate the results we calculate how much ammonia and sulphur dioxide can be solved in
1 kg of solution (the mass fraction of the pollutant in the liquid). The relation between the two
types of the concentrations is shown by the next equation:
mk M k nk M k xk
ωk = = =
m k + m w M k n k + M w n w M k x k + M w (1 − x k )
since
x k + x ww = 1
Substituting the mole fractions calculated above the solubility of ammonia and sulphur
dioxide is as follows:
ω NH3 = 0.009576
ωSO 2 = 0.002473
or 9576 as well as 2473 ppm (parts per million weight). To make a comparison we calculate
how much oxygen can be solved in water when the partial pressure of the oxygen in the gas
mixture is about 21000 Pa (as in the atmospheric air) and the value of the Henry’s coefficient
is about 39·108 Pa (39000 bar). The following result is received:
y O2 = 0.21
x O2 = 5.38 ⋅ 10 −6
ωO2 = 0.00000957
which is 9.57 ppm.

Evaluation of the results


The results show that the solubility of ammonia and sulphur dioxide is much larger
than the solubility of oxygen although the concentration of the first two components is much
less than the oxygen concentration in the gas mixture. t is a general situation for most of the
pollutants to be removed from waste gases or flue gases before emitted into the atmosphere.
Similar results may be calculated for nitrogen. Therefore the natural components of the air
(oxygen and nitrogen) have much less solubility than the pollutants and this gives means for
solving the air pollution control. Physical absorption was assumed for the calculation but
4/16
Dr. Parti Mihály; 463-2635
Department of Fluid Mechanics
26/10/2011
Air Pollution Control
Absorption

chemical absorption may be also used when the reagent concentration in the absorbent
determines the absorptive capacity which may be much larger than for physical absorption
(see later).
Chemical Equilibrium Conditions: In a chemical absorption process the pollutant A
is transferred from the gas phase to the liquid phase where a chemical reaction takes place. It
means that the pollutant A and reactant B produce a new compound as follows:
A + ν BB ⇔ ν PP
where νk is stoichiometric coefficient for the components and the arrow indicates forward and
backward (a reversible) reactions. The rate of the chemical reaction is:
r = k 1 c A c Bν B − k 2 c νPP
where k1 and k2 is the rate constant for the forward and the backward reactions and ck is the
mole concentration of the components. The chemical equilibrium constant is defined as
follows:
k cνP
K = 1 = Pν
k 2 c A c BB
For the chemical absorption process it is of great advantage when the chemical conversion is
product oriented, that is: the forward reaction is much faster than the backward. This will be
the case for high values of k1 and K.

In this case the reaction is very fast and proceeds in the liquid phase very close to the
interface.
For practical applications this type of chemical absorption is the most advantageous, since for
environmental protection the concentrations of the pollutants are generally very low and
physical absorption will never be able to reduce the pollutant concentration very close to zero
(or to zero), but it is often desirable. Such conditions can be only met by chemical absorption.
Hence, chemical absorption is of great importance in the field of air pollution control.

Selection of solvent for absorption


If the principal purpose of the absorption operation is to produce a specific solution, as
in the manufacture of hydrochloric acid, for example, the solvent is specified by the nature of
the product (it is water in hydrochloric acid production). If the principal purpose is to remove
some constituent from the gas mixture (as for solving environmental problems that is to
remove some pollutant from the air), some choice is frequently possible. When choice is
possible, liquids with high solubilities for the solute are preferred as a high solubility reduces
the amount of solvent to be circulated. Water is, of course, the cheapest and most plentiful
solvent, but the following properties are important for considerations:
Gas solubility: The gas solubility should be high, thus increasing the rate of
absorption and decreasing the quantity of solvent required. Generally solvents of a chemical
nature similar to that of the solute to be absorbed will provide good solubility. Thus
hydrocarbon oils, and not water, are used to remove benzene from coke-oven gas. For cases
where the solutions formed are ideal, the solubility of the gas is the same in terms of mole
fractions for all solvents. But it is greater in terms of weight fractions for solvents of low
molecular weight, and smaller weights of such solvents, as measured in kilograms, need to be
used. Chemical reaction of solvent with the solute will frequently result in very high gas
solubility, but if the solvent is to be recovered for reuse, the reaction must be reversible. For
example, hydrogen sulphide may be removed from gas mixture using (mono- or di-)
ethanolamine solutions since the sulphide is readily absorbed at low temperatures and easily
stripped at high temperatures. Caustic soda (NaOH) absorbs hydrogen sulphide or calcium
5/16
Dr. Parti Mihály; 463-2635
Department of Fluid Mechanics
26/10/2011
Air Pollution Control
Absorption

hydroxide (Ca(OH)2) absorbs sulphur dioxide excellently but will not release it in a stripping
operation.
Volatility: The solvent should have a low vapour pressure since the gas leaving the
absorption unit is ordinarily saturated with the solvent and therefore much may be lost by
using solvent of high volatility. Solvent loss may be costly and may present environmental
contamination problems. If necessary, a second, less volatile liquid may be used to recover the
evaporated portion of the first. This is sometimes done, for example, in the case of
hydrocarbon absorbers, where a relatively volatile solvent oil is used in the principal portion
of the absorber because of the superior solubility characteristics and the volatilised solvent is
recovered from the gas by a non-volatile oil. Similarly, hydrogen sulphide may be absorbed
by water solution of sodium phenolate, but the desulphurized gas is further washed with water
to recover the evaporated phenol.
Corrosiveness: The materials of construction required for the equipment should not
be unusual (rarely used) or expensive.
Cost: The solvent should be inexpensive, so that losses are not costly, and should be
readily available.
Viscosity: Low viscosity is preferred for reasons of rapid absorption rates, improving
flooding characteristics in absorption towers, low pressure drops on pumping, and good heat-
transfer characteristics (if necessary).
Miscellaneous: The solvent if possible should be non-toxic, non-flammable and
chemically stable and should have a low freezing point (the equipment are generally operated
outside).
To sum up: the solvent should have high gas solubility, should be relatively non-
volatile, inexpensive, non-corrosive, chemically stable, non-viscous, non-foaming, and
preferably non-flammable. Generally low-cost solvents should be chosen over more
expensive ones of higher solubility or lower volatility. Water is generally used for gases fairly
soluble in water, oils for light hydrocarbons, and special chemical solvents for acid gases such
as carbon dioxide, sulphur dioxide and hydrogen sulphide. Sometimes a reversible chemical
reaction will result in a very high solubility and a minimum solvent rate. Data on actual
systems are desirable when chemical reactions are involved. Examples of gas-liquid
combinations for air pollution control, widely used in industry, are given in Table 1.
Table 1.

Material balances
Remark: the concentration-difference driving forces are existing at one position in the equipment used
to contact the immiscible phases. In the case of a steady-state process, because of the transfer of the solute from
one phase to the other one, the concentration within each phase changes as the phase moves through the
equipment. These changes produce changes in the driving forces, which can be followed with the help of
material balances.
The basic expressions for material balances and their graphical interpretations will be
presented for countercurrent flow (a similar method may be used for cocurrent flow). Figure
3 shows a countercurrent tower which may be either a packed or spray tower, a plate tower
(filled with bubble-cap or any other trays), or may have any internal constructions to bring
about liquid-gas contact.
Figure 3.
Steady-state process is assumed. It means that concentrations at any position in the tower
remain constant with passing time. This requires continuous, invariable flow of all phases into
and out of the apparatus, a continuous and constant flow conditions within the apparatus,
constant concentrations of the feed streams, and unchanging conditions of temperatures and
pressures. In the figure the apparatus used is represented simply as a rectangular box and the
6/16
Dr. Parti Mihály; 463-2635
Department of Fluid Mechanics
26/10/2011
Air Pollution Control
Absorption

two insoluble phases are identified as gas phase G and liquid phase L. It is assumed that only
a single substance A diffuses from gas phase G to liquid phase L during their contact in the
apparatus. The other constituents of the phases, solvents for the diffusing solutes, are then
considered not to diffuse (these are the inert components).
The gas stream at any point in the tower consists of G total moles/hr, made up of
diffusing solute A of mole fraction yA, partial pressure pA, or mole ratio YA, and non-
diffusing, essentially insoluble gas GS moles/hr. The relationships among these are:
G GA G − GS
yA = A = =
G GS + G A G
GA yA pA
YA = = =
GS 1 − y A p − pA
G
G S = G (1 − y A ) =
1 + YA
Similarly the liquid stream consists of L total moles/hr, containing xA mole fraction soluble
gas, or mole ratio XA, and essentially non-volatile solvent LS moles/hr. The relationships
among them are:
xA
XA =
1 − xA
L
L S = L(1 − x A ) =
1 + XA
In the figure the subscripts 1 indicate that end of the tower where the gas phase enters
and 2 that end where the gas phase leaves. At the entrance to the tower (device, apparatus) in
which the two phases are contacted, phase G contains G1 total moles per unit time, consisting
of non-diffusing solvent gas GS moles per unit time and diffusing solute A, whose
concentration is y1 mole fraction. As the gas phase G moves through the equipment, A
diffuses to the liquid phase L and consequently the total quantity of G decreases to G2 moles
per unit time at the exit, although the rate of flow of non-diffusing solvent GS remains is the
same as at the entrance. The concentration of A has decreased to y2 mole fraction. Similarly
the liquid phase L at the entrance contains L2 total moles per unit time, of which LS moles per
unit time is non-diffusing solvent, and an A concentration of x2 mole fraction. Owing to the
accumulation of A to a concentration x1, liquid phase L increases in amount to L1 moles per
unit time at the exit, although the solvent content LS remained constant. Envelope I, the closed
irregular line drawn about the lower part of the equipment, will help to establish a material
balance for substance A, since an accounting for substance A must be made wherever the
envelope is crossed by an arrow representing a flowing stream. The A content of the entering
G phase is G1y1, that of the L phase is Lx. Similarly the A content of the leaving streams is
Gy and L1x1, respectively. Thus the material balance (for y and x subscript A is omitted):
G 1 y 1 + Lx = Gy + L1 x1
which is rearranged for the gas phase mole fraction of y as:
L G y − L 1x1
y = x+ 1 1
G G
The last equation represents a line with varying slope on x-y coordinates, of local slope L/G,
passing through the point whose coordinates are (x1,y1). This line represents the relation
between the liquid and gas concentration at any level in the lower part of the tower.
A similar material balance can be written on the upper part of the equipment, giving
the following rearranged equation:
7/16
Dr. Parti Mihály; 463-2635
Department of Fluid Mechanics
26/10/2011
Air Pollution Control
Absorption

L G y − L2x2
y= x+ 2 2
G G
which also represents a line with varying slope on x-y coordinates, of local slope L/G, passing
through the point whose coordinates are (x2,y2). It is easy to prove that the second parts of the
last two equations are equal, by writing a material balance on the whole equipment which can
be rearranged as follows:
G 1 y 1 − L1x1 = G 2 y 2 − L 2 x 2
It follows that the lines represented by the last two equations coincide. This line is illustrated
on Figure 4 and is called operating line of the equipment. This line represents the relation
between the liquid and gas concentration at any level in the tower.
Figure 4.
The equilibrium-solubility data (the equilibrium curve) for the solute gas in the solvent liquid
may also be plotted on the same diagram as shown in the figure.
Since the solvent gas and solvent liquid are essentially unchanged in quantity as they
pass through the tower, it is convenient to express the material balance in terms of these. A
solute balance about the lower part of the tower (envelope 1) is:
G S Y1 + L S X = G S Y + L S X1
and after rearranging:
L G Y − L S X1
Y = S X+ S 1
GS GS
This is the equation of a straight line (the operating line) on X,Y coordinates, of slope LS/GS,
which passes through the point whose coordinates are (X1,Y1). Substitution of X2 and Y2 for
X and Y shows the line passes through the point whose coordinates are (X2, Y2) also. This
line represents the relation between the liquid and gas concentration at any level in the tower
as at point P on Figure 5.
Figure 5.
The equilibrium data are also plotted in terms of these concentration units on the same
diagram as shown in the figure (as curve MN). Each point on this curve represents the gas
concentration in equilibrium with the corresponding liquid at its local concentration and
temperature. For an absorber (for mass transfer from gas to liquid) the operating line always
lies above the equilibrium curve, while for stripper (for mass transfer from liquid to gas) the
operating line is always below.
In mass transfer theory it is proved that the overall driving force at any position of the
tower is represented by a vertical or a horizontal line between the operating line and the
equilibrium curve on Figure 4 or Figure 5. The driving force obviously changes in magnitude
from place to place or from one end of the equipment to the other. If the operating line should
touch the equilibrium curve anywhere, so that the phases contacted are in equilibrium, the
driving force and hence the rate of mass transfer would become zero, and the time required
for a finite material transfer would be infinite. It is clearly seen that when the operating line is
closer to the equilibrium curve the driving force at any position will be smaller, consequently,
larger equipment is required to fulfil the same mass transfer.

Minimum liquid-gas ratio for absorption


In the design of absorbers, the quantity of gas to be treated G1 or GS, the terminal
concentrations Y1 and Y2 (the inlet and the outlet concentrations) and the composition of the
entering liquid X2 are ordinarily fixed by process requirements, but the quantity of liquid to be
used for the solution of the problem is subject to choice. The operating line must pass through
point D and must end at the ordinate Y1 as shown on Figure 6.
8/16
Dr. Parti Mihály; 463-2635
Department of Fluid Mechanics
26/10/2011
Air Pollution Control
Absorption

Figure 6.
If such a liquid is used to give the operating line DE, the exit liquid concentration will be X1.
If less liquid is used, the exit-liquid concentration will be clearly greater, as point F shows,
but the driving forces for the mass transfer are less, the absorption becomes more difficult.
The contact time between gas and liquid must be greater, and the absorber must be
correspondingly taller. The minimum liquid quantity which may be used corresponds to the
operating line DM, which has the greatest slope for any line touching or crossing the
equilibrium curve and is the tangent to the curve at P. At P the diffusional driving force is
zero, the required contact time for the concentration change desired is infinite, and an
infinitely tall tower results. This then represents the limiting (the minimum) liquid-gas ratio.
The equilibrium curve is frequently concave upward, as is shown in Figure 6 also, and the
minimum liquid-gas ratio then corresponds to an exit liquid concentration in equilibrium with
the entering gas.
The actual liquid-gas ratio normally will be greater than the minimum one by as much
as 25 to 100 percent and may be determined by economic considerations as well as by
judgement and experience. For example, in some packed-tower applications involving very
soluble gases or vacuum operation, the minimum quantity of solvent needed to dissolve the
solute may be insufficient to keep the packing surface thoroughly wet, leading to poor
distribution of the liquid stream.
These principles also apply to strippers, where an operating line which anywhere
touches the equilibrium curve represents a maximum ratio of liquid to gas and a maximum
exit-gas concentration.

Equipment for absorption


The purpose of the equipment used for absorption is to provide intimate contact of the
two phases in order to permit inter-phase diffusion of the constituents. The mass transfer rate
is directly dependent on the interfacial surface exposed between the phases, and the nature
and degree of dispersion of one of the phases in the other one are therefore of prime
importance. The equipment may be classified according to whether its principal action is to
disperse the gas or the liquid, although sometimes both phases become dispersed. The gas is
dispersed in agitated vessel (rarely used for absorption) and in tray towers. The liquid is
dispersed in Venturi scrubbers, in wetted-wall towers, in spray towers, in baffle towers (or
shower trays) and in packed towers. Only a brief discussion of tray towers and packed towers
will be given here.
Tray (or Plate) towers Tray towers are vertical cylinders in which the liquid and gas
are contacted in stepwise fashion on trays or plates. A schematic representation of a bubble-
cap tray tower is shown on Figure 7.
Figure 7.
The liquid enters at the top and flows downward by gravity. On the way, it flows
across each tray horizontally and through a downcomer to the next tray below. Figure 8
shows different tray arrangements and the direction of liquid flow on the trays.
Figure 8.
The gas passes upward through the openings in the tray, and then bubbles through the
liquid to form froth, and disengaging from the froth passes on to the next tray above. The
overall effect is a multiple countercurrent contact of gas and liquid, although each tray is
characterized by a cross flow of the two phases. Each tray of the tower is a stage, since on the
tray the fluids are brought into intimate contact, interphase mass transfer occurs, and the
fluids are separated. The number of theoretical trays (or stages) in a column (or tower) is
dependent only upon the difficulty of the separation to be carried out and is determined solely
9/16
Dr. Parti Mihály; 463-2635
Department of Fluid Mechanics
26/10/2011
Air Pollution Control
Absorption

from material balances and equilibrium considerations. The stage or tray efficiency, and
therefore the number of real trays, is determined by the mechanical design used and the
conditions of operation. The diameter of the tower, on the other hand, depends upon the
quantities of liquid and gas flowing through the tower per unit time. Once the number of
theoretical trays required has been determined the principal problem in the design of the tower
is to choose dimensions and arrangements which will represent the best compromise among
several opposing tendencies, since it is generally found that conditions leading to a high tray
efficiency will ultimately lead to operational difficulties. The most frequently used types of
trays are bubble-cap trays, sieve trays and valve trays. Figure 9 shows two typical bubble-cap
designs.
Figure 9.
Plate columns may be economically preferable for large-scale operations and are
needed when liquid rates are so low that packing would be inadequately wetted, when the gas
velocity is so low (owing to a very high liquid-to-gas ratio) that axial dispersion or “pumping”
of the gas back down the (packed) column can occur, or when intermediate cooling is desired.
Also, plate towers may have a better turndown ratio and less subject to fouling by solids than
are packed towers.
Packed towers Packed towers are vertical cylinders which have been filled with
packing or devices of large surface. A schematic representation of a packed tower is shown on
Figure 10.
Figure 10.
The liquid is distributed over, and trickles down through the packed bed, thus exposing a
large surface to contact the two phases. The adequate initial liquid distribution is very
important. Dry packing is of course completely ineffective for mass transfer between the two
phases, and various devices are used for more efficient liquid distribution. Two major types of
packing are used, random or regular. Random packing is that which is simply dumped into the
tower and allowed to fall at random. Regular packing is carefully stacked in the tower. Both
types of packing have advantages and disadvantages from economic standpoint (investment
and operational costs). Figure 11 shows several types of packing.
Figure 11.
Packed columns usually are chosen for very corrosive materials, for liquids that foam badly,
for either small- or large-diameter towers involving very low allowable pressure drops, and
for small-scale operations requiring diameters of less than 0.6 m (2 ft). The type of packing is
selected on the basis of resistance to corrosion, mechanical strength, capacity for handling the
required flows, mass-transfer efficiency and cost.
Details on the operating characteristics of plate and packed towers are given in special
literature.

Absorption plant
The general layout of an absorption plant is shown in Figure 12.
Figure 12.
An absorption plant consists of the absorption unit, the regeneration unit and all necessary
auxiliary equipment and machinery. The gas mixture with the gaseous pollutant, namely the
solute or the absorptive, enters the absorption unit at the bottom. Within the absorption unit
the gas mixture is brought into contact with the liquid (solvent or absorbent). During the
contact time the pollutant is at least partly transferred from the gas to the liquid. The purified
gas mixture leaves the absorption unit at the top. The liquid normally enters at the top of the
absorber and leaves it at the bottom. The liquid leaving the absorption unit and carrying the

10/16
Dr. Parti Mihály; 463-2635
Department of Fluid Mechanics
26/10/2011
Air Pollution Control
Absorption

pollutant is directed to some regeneration unit where the absorbent is treated in such a way
that it may be recycled to the absorption unit and the pollutant is disposed appropriately.
Regeneration may be achieved by various processes such as vaporisation, rectification,
steam stripping, desorption, extraction etc. If necessary, regenerative medium is introduced to
extract the pollutant.
The solubility of the different gases is decreasing with increasing temperature and
decreasing pressure (as mentioned earlier). It follows that the absorption process should be
carried out at a relatively low temperature and high pressure. However in air pollution control
many times the gas mixture to be treated has a higher temperature and lower pressure.
Therefore, the waste gas has to pass through a pre-cooler in order to reduce its temperature as
low as possible. Another important property which should be observed is the low (sometimes
very low) concentration of the pollutant in the gas mixture. In many cases the pollutant
concentration is so low that it hardly can be detected in the gas mixture so a very efficient
absorption technique is required for the removal of such a tracer component.
Taking into account the considerations mentioned above the absorption unit is only
one part of the absorption plant which may consists of a generation unit, heat exchangers, as
well as fans and pumps for transporting the gas and the liquid, respectively. In many cases,
the absorption unit is the smallest part of the plant. Earlier the consideration was focused on
the absorption process and the absorption unit which is the heart and the most important part
of the whole absorption plant. But it should be emphasized that the absorption unit (or the
absorber) should be considered in combination with the auxiliary equipment for the
optimisation of the process. In the following an important application will be shown.

Sulphur dioxide removal from stack gases.


Sulphur dioxide is an important chemical component of waste gases from power plants
when some fossil fuel (coal, crude oil, natural gas) containing sulphur is used as energy
source. Before the eighties in Europe (also in the USA) many countries used coal which had
high sulphur content (3-4 % m/m). Burning the coal in the boiler the sulphur is also burnt and
sulphur dioxide is produced:
S + O 2 = SO 2
If the sulphur dioxide is emitted into the atmosphere it causes the acid rain. Strict
international regulation has been applied to decrease the emission of the sulphur dioxide
below the emission limit since 1985 (signed by most European countries as well as by the
USA, Canada and the UN). At the beginning the requirement was fulfilled by changing the
fuel (using coal or crude oil with less sulphur content or natural gas which practically has no
sulphur content) but when the regulation was tightening up the desulphurization of the waste
gas was introduced. The desulphurization of the waste gas can be carried out with several
processes classified into three groups as wet processes, dry processes and half-dry (or half-
wet) processes but mainly a wet process is used all over the world. In this process the
absorbent (the solvent) contains calcium hydroxide which reacts with the sulphur dioxide in
the liquid phase creating gypsum. Figure 13 schematically describes the desulphurization of
stack gases.
Figure 13.
The volumetric flow rate of these waste gases is of the order of several million m3/h. The
polluted gas containing the sulphur dioxide passes through a dust separator (because of the
strict regulation generally the electrostatic precipitators or sometimes bag filters are used).
The waste gas leaving the dust separator moves to the absorber that may be operated in one or
(rarely) two stages (two absorbers connected into series).

11/16
Dr. Parti Mihály; 463-2635
Department of Fluid Mechanics
26/10/2011
Air Pollution Control
Absorption

The sulphur dioxide is absorbed in the absorbent, where reacts with the calcium
hydroxide producing calcium sulphite. The problem is that the calcium sulphite is soluble in
water and can not be disposed since the rainwater will wash it into the soil and underground
water. It means that an oxidation process is required to get calcium sulphate (gypsum). The
steps are as follow:
• absorption in the water, producing sulphuretted (or sulphurous) acid:
SO 2 + H 2O = H 2SO 3 ,
• chemical reaction in the liquid phase with the calcium hydroxide, producing calcium
sulphite:
Ca(OH )2 + H 2SO 3 = CaSO3 + 2H 2O ,
• oxidation of the calcium sulphite producing calcium sulphate (which is practically
insoluble in the water and it is the reason that gypsum is used as building material):
CaSO3 + 0.5O 2 = CaSO 4 ,
which contains crystal water (not shown in the chemical reaction), removable only with
heating. The flue gas generally contains the oxygen necessary to the oxidation as some excess
air is used in the boiler, but more excess air or directly oxygen can be introduced into the
lower part of the absorber in order to enhance the efficiency of the oxidation.
The desulphurized waste gas from the absorber moves through a mist separator and to
a heat exchanger for reheating it to about 110 °C or higher and finally to stack.
Remark: a quick calculation shows that 1 kg of absorbent containing 0.1 kg calcium
hydroxide (10 % by mass) can absorb about 0.086 kg sulphur dioxide which is 86000 ppm (in
ideal conditions). It is about 35 times higher than received for physical absorption (it was
2473 ppm).

Advantages and disadvantages of absorption systems (packed and plate columns)


Advantages
• Relatively low pressure drop
• Capable of achieving relatively high mass-transfer coefficient and high mass transfer
rate
• Increasing the height and/or type of packing or try numbers capable of improving
mass transfer without purchasing a new piece of equipment
• Relatively low capital cost
• Relatively small space requirements
• Ability to collect particles as well as gases
• Collected substances may be recovered by distillation
• Fiberglass-reinforced plastic construction permits operation in highly corrosive
atmospheres.
Disadvantages
• Possibility of creating water (or liquid) disposal problem (remark: IPPC)
• Product collected wet
• Particulates deposition possibly causing plugging of the bed or the plates
• Fibreglass-reinforced plastic is sensitive to temperature
• Relatively high maintenance cost
• Must be protected from freezing (if operated in outside)
Comparison of packed and plate towers
Packed towers
• Lower pressure drop
12/16
Dr. Parti Mihály; 463-2635
Department of Fluid Mechanics
26/10/2011
Air Pollution Control
Absorption

• Simpler and cheaper construct


• Preferable for liquids with high foaming tendencies
Plate towers
• Less susceptible (sensitive) to plugging
• Less weight
• Less susceptible (sensitive) to channelling
• Temperature variation resulting in less damage

13/16
Dr. Parti Mihály; 463-2635
Department of Fluid Mechanics
26/10/2011
Air Pollution Control
Absorption

Introduction

Definitions Gas absorption is a unit operation in which soluble components of a gas


mixture are dissolved in a liquid. The inverse operation, called stripping or desorption, is
employed when it is desired to transfer volatile components from a liquid mixture into a gas.
Both absorption and desorption (stripping), in common with distillation, make use of special
equipment for bringing gas and liquid phases into intimate contact. This section is concerned
with the application of absorption and stripping processes as well as with several special
questions of their design and operation.

Equipment Absorption and stripping (as well as distillation) operations are usually
carried out in vertical, cylindrical columns or towers in which devices such as plates or
packing elements are placed. The gas and liquid normally flow countercurrently, and the
devices provide the contacting interfacial surface through which mass transfer takes place
(cocurrent flow is also possible from economic standpoint). Background material on this mass
transfer process is given in a different class.

General design procedure The designer is ordinary required to determine (1) the
best solvent; (2) the vessel diameter; (3) the height of the vessel and its internal members,
which is the height and type of pacing or the number of the contacting trays; (4) the optimum
solvent circulation through the absorber and stripper; (5) the temperatures of streams entering
and leaving the absorber and the quantity of heat to be removed to account for heat of solution
and other thermal effects; (6) the pressures at which the absorber and stripper will operate;
and (7) mechanical design of the absorption and stripping vessels (normally columns or
towers), including flow distributors, packing supports, and so on.
The problem presented to the designer of a gas-absorption unit usually specifies the
following quantities: (1) gas flow rate; (2) gas composition, at least with respect to the
component or components to be removed (to be absorbed); (3) operating pressure and
allowable pressure drop across the absorber; (4) minimum degree of recovery of one or more
solutes; and, possibly, (5) the solvent to be employed. Items 3, 4, and 5 may be subject to
economic considerations and therefore are sometimes left up to the designer. For determining
the number of variables that must be specified in order to fix a unique solution for the design
of an absorber one can use the same phase-rule approach as for distillation.
Recovery of the solvent, sometimes by chemical means but more often by distillation,
is almost always required, and the recovery system ordinarily is considered an integral part of
the absorption-system process design. A more efficient solvent-stripping operation normally
will result in a less costly absorber because of a smaller concentration of residual dissolved
solute in the regenerated solvent; however, this may increase the overall cost of solvent
recovery. A more detailed discussion of these and other economic considerations is not
presented here because of time shortage.

Selection of solubility data Solubility values determine the liquid rate necessary for
complete or economic solute recovery and so are essential to design. Equilibrium data
generally will be found in one of three forms: (1) solubility data expressed either as solubility
in weight or mole percent or as Henry’s-law coefficients (2) pure-component vapour
pressures, or (3) equilibrium distribution coefficients (K values as mentioned in literature).
Data for special systems may be found in different references.

14/16
Dr. Parti Mihály; 463-2635
Department of Fluid Mechanics
26/10/2011
Air Pollution Control
Absorption

In order to define completely the solubility of a gas in a liquid, it generally is


necessary to state the temperature, the equilibrium partial pressure of the solute gas in the gas
phase, and the concentration of the solute gas in the liquid phase. Strictly speaking, the total
pressure on the system also should be stated, but for low total pressures, less than about 500
kPa (5 atm), the solubility for a particular partial pressure of solute gas normally will be
relatively independent of the total pressure of the system.

Pure-component vapour pressures can be used for predicting solubilities for systems in
which Raoult’s law is valid. For such systems
p A = PA x A
where PA is the pure-component vapour pressure of the solute, pA is its partial pressure in the
gas phase and xA is its mole-fraction in the liquid phase. Extreme care should be exercised
when attempting to use pure-component vapour pressures to predict gas-absorption behaviour.
Both liquid-phase and gas-phase nonidealities can cause significant deviations from the
behaviour predicted by Raoult’s law. Vapour-pressure data are available in literature.

Whenever data are available for a given system under similar conditions of
temperature, pressure, and composition, equilibrium distribution coefficients (K values)
provide a much more reliable tool for predicting vapour-liquid distributions. The distribution
coefficient gives a direct relation between the mole fraction of A component in the gas phase
and the liquid phase as follows:
y A = K AxA
Detailed discussion of equilibrium K values may be found in references.

Calculation of Liquid-to-Gas Ratio The minimum possible liquid rate is readily


calculated from the composition of the entering gas and the solubility of the solute in the exit
liquor, saturation being assumed. It may be necessary to estimate the temperature of the exit
liquid based on the heat of solution of the solute gas. Values of latent and specific heats and
values of heats of solution (at infinite dilution) are given in references.

Column Diameter and Pressure Drop Flooding determines the minimum possible
diameter of the absorber column, and the usual design is for 60 to 80 percent of the flooding
velocity. Maximum allowable pressure drop may be determined by the cost of energy for
compression of the feed gas. For systems having a significant tendency to foam, the
maximum allowable velocity will be lower than estimated flooding velocity, especially for
plate towers. The safe range of operating velocities should include the velocity one would
derive from economic considerations. Methods for predicting flooding velocities and pressure
drops are given are references.

Computation of Tower Height The required height of a gas-absorption or stripping


tower depends on (1) the phase equilibria involved, (2) the specified degree of removal of the
solute from the gas, and (3) the mass-transfer efficiency of the apparatus. These same
considerations apply both to plate towers and to packed towers. Items 1 and 2 dictate the
required number of theoretical stages (plate tower) or transfer units (packed tower). Item 3 is
derived from the tray efficiency and tray spacing (plate tower) or from the height of one
transfer unit (packed tower). Solute-removal specifications normally are derived from
economic considerations.
For plate towers, the approximate design methods may be used to estimate the number
of theoretical stages, and the tray efficiencies and tray spacing for the tower can be specified
15/16
Dr. Parti Mihály; 463-2635
Department of Fluid Mechanics
26/10/2011
Air Pollution Control
Absorption

on the basis of the information given later. Considerations involved in the rigorous design of
the theoretical stages for plate towers are treated in references. For packed towers, the
continuous differential nature of the contact between gas and liquid leads to a design
procedure involving the solution of differential equations, described in references.

Selection of Stripper-Operating Conditions Stripping involves the removal of one


or more volatile components from a liquid by contacting it with a gas such as steam, nitrogen,
or air. The operating conditions chosen for stripping normally results in a low solubility of the
solute (i.e. a high value of m), so that the ratio mG/L will be larger than unity. A value of 1.4
may be used for rule-of-thumb calculations involving pure physical desorption. For plate-
tower calculations the stripping factor S=KG/L, where K=y/x, usually is specified for each
tray.
When the solvent from an absorption operation must be regenerated for recycling back
to the absorber, one may employ a “pressure–swing concept”, a “temperature-swing concept”
or a combination of both in specifying stripping conditions. In pressure-swing operation the
temperature of the stripper is about the same as that of the absorber, but the stripping pressure
is much lower. In temperature-swing operation the pressures are about the same, but the
stripping temperature is much higher than the absorption temperature.
In pressure-swing operation a portion of the dissolved gas may be sprung from the
liquid by the use of a flash dram upstream of the stripping-tower feed point. If the flashing of
the feed liquid takes place inside the stripping tower, this effect must be accounted for in the
design of the upper section in order to avoid overloading and flooding near the top of the
tower.

Design of Absorber-Stripper Systems The solute-rich liquor leaving a gas absorber


normally is distilled or stripped to regenerate the solvent for recirculation back to the
absorber. It is apparent that the conditions selected for the absorption step (e.g., temperature,
pressure, L/G) will affect the design of the stripping tower, and, conversely, a selection of
stripping conditions will affect the absorber design. The choice of optimum operating
conditions for an absorber-stripper system therefore involves a combination of economic
factors and practical judgements as to the operability of the system within the context of the
overall process flow sheet.

Absorption is the process by which a gas component is dissolved in a liquid. In most


cases, absorption is the process of selective removal of a component from a gas mixture. The
gas mixture consists of insoluble and soluble components, the first one is called inert
component and the second one is the diffusive or solute component. The liquid consists of the
solvent and the absorbate selectively removed from the gas mixture. The liquid mixture
including the absorbate is called the absorbent. Absorption is one of the most advanced
techniques applied to the separation of gas mixtures. Absorption is widely used in industry,
and in the areas of air pollution control it has become one of the important processes for the
abatement of gaseous pollutants. In Table 1 pollutants are compiled that may be removed
from a gas mixture by means of absorption.

16/16
Dr. Parti Mihály; 463-2635
Department of Fluid Mechanics
26/10/2011

You might also like