You are on page 1of 85

5 Mass Transfer

Mass transfer is the movement of a chemical species from a region of higher to lower
concentration. We can see the effect of mass transfer when colored solid crystals
dissolve in a clear liquid, air pollutants disperse, and materials dry. Mass transfer is
achieved by concentration differences that give rise to concentration gradients
(changes of concentration over a distance).

Concentration gradients are not the only driving forces for mass transfer. Mass
transfer can also be driven by differences in pressure (as in centrifugations), external
forces (electrolytic cells, Figure 5.1) or temperature (thermal diffusion). To
understand mass transfer, therefore, we must examine the interplay between the
different driving forces and fluxes.

FIGURE 5.1. The operation of a mercury cell is a good example of mass transfer
driven by external forces rather than concentration gradients.

TABLE 5.1. Flux-Gradient Interactions

Mass transfer, is governed by Fick's law of diffusion. The basic equation governing
mass transfer is Fick's first law

(5.1)

where DAB is the diffusivity of component A through B, and JA is the molar flux in
moles per unit time per unit area. Because mass flux is a vector,

(5.2)

1/85
where i, j, and k are unit vectors in the x, y, and z direction, and JA is the mass flux
vector.

In defining JA, it is stipulated that this flux must be referred to a plane across which
there is no net volume transport (i.e., the plane moves with respect to the fixed
apparatus although the fluid is stagnant; Figure 5.2). While this definition of JA is
physically correct, it is not useful in design and process work. What we need, instead,
is a flux defined relative to the apparatus itself, rather than to a moving plane. We call
this new flux N, which we shall derive.

The velocity of the moving plane can be given by

(5.3)

where Uy is the velocity of the moving plane in distance per unit time, Ni is the mass
flux of the ith component in moles per unit area per unit time, and v I = is the partial
molal volume of the ith component in volume per mole.

For a binary system with components A and B

(5.4)

(5.5)

and

(5.6)

FIGURE 5.2. In this depiction of the motion of a diffusion boundary, flux JA


consists of the molecules that move across the moving boundary of the mass
transferred. NA combines convective mass transfer and molecular diffusion.

2/85
A mass balance on an overall basis is simple because outflow minus inflow equals
accumulation or depletion. To consider a mass balance of individual species, we m
treat each species separately with its own mass balance. A mass balance for a sit~
species is called the equation of continuity of species.

To derive such a specific mass balance equation, we must consider the fluxes mass in
and out of a space element (Figure 5.3). We also have to consider chemical reaction
because it can change one species to another.

The equation of continuity for a given species A in rectangular coordinates is

(5.7)

where RA is the rate of chemical reaction in moles per cubic centimeter-second. For
cylindrical coordinates, the mass balance is

(5.8)

Equations 5.7 and 5.8, and their counterparts in spherical coordinates, represent the
most important equations in chemical and petroleum processing, for they tie together
chemical reaction and mass transfer.

Diffusivity, the transfer coefficient or property that relates flux and gradient, varies
according to the combination of phases involved. Table 5.2 shows typical diffusivity
values. The diffusivities of components in gases are orders of magnitude greater than
their values in liquids or solids. These values point out that diffusion within a liquid or
solid is much slower than the diffusion of a molecule at the interface of a liquid or
solid with a gas.

Not shown in Table 5.2 is the effect of the diffusing species' concentrations the
diffusion coefficient. Concentration has a large effect on the diffusion coefficient in
liquid and solid systems but not in gas systems at a given pressure regardless of the
amounts of A or B (DAB is the same value for gaseous A_B mixes whether the
concentrations are 10-90, 30-70, 50-50, 60-40.)

3/85
FIGURE 5.3. A mass balance on a volume in space accounts for mass flux into
and out of the space in three coordinates plus chemical reaction.

TABLE 5.2 Typical Diffusivity Values for Various Systems

The process of mass transfer is analogous to that of heat transfer as we can see by
comparing eq 4.6, the equation of energy

and eq 5.7, the equation of continuity of species

Both equations have accumulation terms ( Cp T/ t and CA/ t ), generation terms (


.
q and RA), and flux terms (conduction and the NA term). Temperature and

concentration are analogous parameters.

4/85
Earlier we discussed heat transfer by considering steady-state heat conduction across
a slab. Similarly, we begin considering mass transfer with steady-state molecular
diffusion, illustrated in Example 5.1.

MASS TRANSFER AT A PHASE BOUNDARY

The theoretical and basic equations for mass transfer can be solved for simple
geometries and laminar flow where the flow field is specified. These cases can be
mathematically complicated. If we consider turbulent flow and complex geometries,
we cannot handle the situation analytically and classically simply or in some cases not
at all. This situation clearly parallels that of heat transfer. Hence, we can define a mass
transfer coefficient k and extend the analogy to heat transfer:

MASS TRANSFER FOR DIFFERENT GEOMETRIES

As mentioned earlier, the transfer of heat and mass are analogous for similar
geometries, boundary conditions, and flow For example (5),

(5.18)

where jD, the mass transfer flux, is equal to jH, the heat transfer flux,

(5.19)

and

(5.20)

This identity holds for flow over flat plates, cylinders, through packed beds, and in
pipes and conduits (where Re 10,000).

Correlations exist for various geometries. The Reynolds number is correlated with the
coefficients for flow through tubes (Figure 5.5).

5/85
FIGURE 5.5. Heat and mass transfer are analogous for certain geometries. For
flow through tubes, the Reynolds number Re can be correlated with both the
heat transfer coefficient and the mass transfer coefficient, using the same data.
(Adapted from reference 20.)

EXAMPLE 5.2. SUBLIMATION RATE FROM A CYLINDER

Air passes through a naphthalene tube that has an inside diameter of 0.0254 m and a
length of 1.83 m. The velocity is 15.24 m/s and the air is at 10 C and atmospheric
pressure. What is the percent saturation of air with naphthalene, and what is the rate of
naphthalene sublimation at 15.24 m/s given air properties:

6/85
To deal with the mass transferred, we need a balance analogous to Q= H in heat
transfer. Our balance equates the mass taken up by the air and the mass transferred.
We also use differential changes for the length of the tube (dx) and the change of
density of the napthalene vapor (d A). The ( A sat - A bulk) is the driving force for mass
transfer. Hence, the mass taken up by the air equals the mass transferred, so

Hence, saturation is 35%.

At 10 C, saturation of naphthalene in air is 1.52 X 10-4 kg/m3 because

A sat = (2.79 N/m2)(128)(1.249 kg/m3)/(1.01 X l05 N/m2)(29)

7/85
Then,

total evaporation rate =(/4)(0.0254 m) 2 (15.24 m/s)(0.35)(1.52 x 10-4 kg/m3)

total evaporation rate = 4.2 X 10-7 kg/s

This last example has some points that deserve further discussion. The first point is
the use of the mass transfer coefficient k based on density, which was not previously
mentioned. Saturation calculations, such as for humidity, are based on densities. Also,
using k means that PBM, a term that requires the partial pressures of air at various
points for calculations, did not have to be used.

A second point is that when calculating Sc, the diffusivity value was that for
naphthalene in air, DBA, whereas the other values such as for density and viscosity
were those for air. The use of these values appears to be inconsistent. However, the
diffusivities of naphthalene in air and air in naphthalene are identical. An explanation
for this can be derived from the kinetic theory of gases, which yields identical
expressions for DAB and DBA. (In brief, the diffusivities are derived from a collection
of the same terms involving temperature, pressure, and other parameters multiplying
either [(1/MA) + (1/MB)]0.5 for DAB or [(1/MB) + (1/MA)]0.5 for DAB; MA and MB are the
molecular weights of components A and B, respectively. Hence, DAB = DBA in a binary
system.) The relationship DAB equals DBA holds for binary mixtures, but it does not
hold for multicomponent systems.

EXAMPLE 5.3. EVAPORATION RATE FROM A DROPLET

A spherical drop of water (0.05 cm in diameter) is falling at a velocity of 215 cm/s


through dry, still air at 1 atmosphere. Estimate the instantaneous rate of evaporation
from the drop if the drop surface is at 70 F and the air is at 140 F.

To solve this problem, we shall assume ideal gas behavior, insolubility of air in water,
equilibrium at the interface, and pseudosteady-state conditions.

For a small evaporation rate, we can write a mass transfer coefficient equivalent to eq.
5.6,

8/85
where WA is the molar rate of exchange of A, and XA0 and XA are the water mole
fractions at the drop surface and in the air beyond the boundary layer around the drop.

9/85
This molar flow rate amounts to a decrease of 1.23 x 10-3 cm/s in the drop diameter.
Hence, a drop would fall a considerable distance before evaporating.

If we were to do Example 5.3 on a rigorous basis, not making the assumptions we did,
we would combine heat and mass transfer, and also the change in drop diameter as the
water evaporates. This change would require a differential equation. However, the
simpler approach yielded acceptable results.

Equilibrium Stage Operations, The Ideal Stage, And Phase Equilibria

Diffusional mass transfer describes many physical processes. However, it is difficult


in many instances to use this technique to design large-scale equipment. For such
cases, we use an approach based on an ideal stage (Figure 5.12). Its contents are so
well mixed that streams leaving it are in equilibrium. In an ideal stage, the mass
transfer takes place because the entering streams, which are not at equilibrium, reach
equilibrium. The differences in composition between the entering and equilibrium
values constitute the driving forces.

Mass transfer operations such as distillation, absorption, extraction, leaching, and


crystallization can be analyzed by the ideal stage concept. Solutions for these mass
transfer processes or the design of devices to carry out such operations require
calculations of (1) a material or mass balance, (2) an enthalpy or energy balance, and
(3) equilibrium data. The first two items are the conservation of mass and
conservation of energy. The third item represents the appropriate relationship between
the phases present, for example, vapor-liquid equilibrium data for distillation. Table
5.3 shows the type of data required for a given process (see also Figures 5.13-5.16).

Figures 5.14 and 5.15 both represent the vapor-liquid equilibrium between benzene
and toluene. Their interchangeability can be explained by the use of the Gibbs phase
rule

where P is the number of phases, V represents the variables or degrees of freedom,


and C is the number of components. There are two phases and two components in the
benzene-toluene system. Hence, there are two degrees of freedom. One of these
degrees is specified by pressure, which leaves just one variable. If the benzene mole
fraction in either liquid or vapor is specified, then we automatically fix the

10/85
temperature and mole fraction in the other phase. The reader should verify this
exercise by studying Figures 5.14 and 5.15.

MASS BALANCES FOR AN IDEAL STAGE

The concept of an ideal stage can be explained by examining a continuous distillation


column (Figure 5.17). Calculations for a mass balance across a continuous distillation
are simpler than for a batch distillation. Once the continuous column reaches

FIGURE 5.12. Calculations for mass transfer operations, such as distillation and
absorption, use the concept of the process as a series of ideal stages. An ideal or
equilibrium stage is so well-mixed that the exiting streams, D and C in this
illustration, are in equilibrium. Non-equilibrium streams, A and B, provide the
driving force for mass transfer within the stage.

TABLE 5.3. Types of Equilibrium Data Needed To Calculate the Mass Transfer
in Ideal Stage Operations

11/85
FIGURE 5.13. Equilibrium calculations for an ideal stage in mass transfer
equipment, such as an absorption column, may require solubility data. These
data are for the solubility of ammonia in water

FIGURE 5.14. Boiling point data, such as this diagram for benzene and toluene,
is helpful when calculating the number of ideal stages needed for a distillation
column.

its steady state, all of the streams' concentrations remain constant with time. Batch
columns, however, have concentrations that change with time as the volatile materials
first distill, followed by less volatile components. Illustrations of distillation columns,
packing materials, and packed columns are given (Figures 5.17-5.21).

In setting up the mass balances for the column shown in Figure 5.17, we will only
consider the streams cut by given boundary. On an overall basis, we use the AAAA
boundary. The overall mass balance for total streams is

12/85
F=D+B (5.30)
Where F, D, and B are the molecules per hour of the feed, overhead (distillate)
product, and the bottoms product stream, respectively. The mass balance on the most
volatile stream would be

FXF = DXD + BXB (5.31)

FIGURE 5.15. If ideal stage calculations are performed for a two-component


system, the equilibrium data needed may be a vapor-liquid mole fraction
diagram, such as this one for benzene and toluene.

FIGURE 5.16: Reading a ternary diagram. Separation processes sometimes


make use of an extracting agent in which the solute is soluble and the solvent is
not. The three components form a ternary system. This diagram is for a ternary

13/85
system of water MIK (methyl isobutyl ketone), and acetone. Each corner of the
triangle represents 100% of a component. Each side opposite a point represents
0% of the component. The area under the curved line ACEDB is formed of a
mixture of two saturated and insoluble liquid phases, the composition of each
defined by the points at the end of tie lines crossing the area. In such a ternary
mixture, the closer points A and B come to the points of the triangle, the more
insoluble are the liquids A and B in each other The driving force behind the
extraction at each ideal stage is the difference in the solubility of acetone in A-
rich and B-rich phases related by a tie line (see Example 5.5). (Adapted from
reference 47.)

FIGURE 5.17. A distillation column can be examined from several mass balance
points of view.A balance for the entire column would be defined by boundary
AAAA, the enriching section above the feed point by boundary BBBB, and the
reflux section by boundary CCCC. F is the feed, D is the distillate, and B is the
bottoms product, all in moles per hour X is the mole fraction of component in the
liquid, and Y is the fraction in the vapoc (Adapted from reference 48.)

where X represents the fraction of component A in the liquid feed, distillate, and
bottoms streams.

Solving eqs 5.30 and 5.31 for the ratios of distillate and bottoms to feed results in

D/F = (XF XB)/(XD - XB) (5.32)

14/85
B/F = (XD - XF)/(XD - XB) (5.33)
Next we will look at the enriching or rectifying section of the column (the sections
above the feed in Figure 5.17) with boundaries BBBB and CCCC in which n is the
stage number, Y is the mole fraction of a in the vapor stream, V is the molar flow rate
of the vapor, and L is the molar flow rate of the liquid. The boundary BBBB gives

(5.34)

and the boundary CCCC yields

(5.35)

Combining the two equations yields

(5.36)

Finally, a boundary (not shown in Figure 5.17) around the entire portion of the
column and reflux unit that cuts Ln, Vn+1, and D, shows that Vn+1 equals Ln + D.
Substitution in eq 5.36 gives

(5.37)

15/85
FIGURE 5.18. A stage can take various mechanical forms in process equipment.
Process units, such as a distillation column in a petrochemical plant, can be more
than 100 ft tall with hundreds of stages (trays). (a) In a bubble tray, each stage is
composed of a perforated sheet of metal with a weir cutting across one side and a
bubble cap covering each perforation. The cap (b) has slits through which rising
vapor passes, bubbling through the downflowing liquid.

16/85
FIGURE 5.19. Contact between vapor and gas in separation processes, such as
extraction or distillation columns, can be enhanced by filling the columns with
packing material. Column packing comes in many shapes, sizes, and materials.

A similar approach for the stripping section (the stages below the feed) gives eq 5.38
and 5.39. To understand the mass balance process, it is best to verify these equations
by working them out in a manner similar to that used for eq 5.36 and 5.37.

(5.38)

(5.39)

FIGURE 5.20. In a disk-and doughnut column, vapor is forced to contact the


downflowing liquid by being diverted to the edge by a disk.

17/85
FIGURE 5.21. In a packed separation column, liquid is fed into the top of the
column and flows down over the packing, while gas flows upward.

ENERGY BALANCE FOR AN IDEAL STAGE

To balance the energy across a stage, we use a base enthalpy of zero for a liquid (Ln)
at Tn and the types of data specified in Table 5.4, given stage n at temperature Tn.
The energy balance is then

a+bcd+ef=0 (5.40)
In eq 5.40, a and d are much larger than the other terms, and the remaining terms are
about the same order of magnitude. Hence we can neglect terms b, e, c, and f, leaving

a=d (5.41)
TABLE 5.4. Energy Values Needed for an Ideal Stage Balance

18/85
(5.42)

(5.43)

(5.44)

THE McCABE-THIELE METHOD

When eqs 5.41-5.44 hold, the conditions are known as constant molal vaporization
and constant molal overflow. These conditions mean that the liquid and vapor flow
rates in the enriching and stripping sections are, respectively, equal. Therefore, eqs
5.36-5.39 are straight lines.

This method, known as the McCabe-Thiele method, is used to solve the relationship
graphically between material balances and equilibrium for constant molal overflow
and to calculate the number of stages in mass transfer equipment, such as distillation
columns. A typical McCabe-Thiele solution is shown in Figure 5.22, in which each
step represents a stage.

The starting point in a McCabe-Thiele diagram is an equilibrium diagram of the type


shown in Figure 5.15. The first step is to depict the appropriate material balances on
the graph.

The rectifying operating line (the upper line in Figure 5.22) can be placed. It must
pass through the point XD = YD because at the top of the column (Figure 5.17), YA =
XD = XA, and it intercepts the y axis at XD/( RD + 1 ) where RD is the reflux ratio, La/D.
The stripping line (the lower line in Figure 5.22) must intercept both the feed line and
the rectifying line. The feed line slope depends on the quality of the feed, whether it is
a saturated liquid, super-heated vapor, etc. (Figure 5.23). The equation of the feed line
is

19/85
(5.45)

where f is the feed mole condition. Values of f for various conditions are shown in
Table 5.5.

Once the rectifying and stripping operating lines are correctly placed, we can
determine the number of stages (the steps in Figure 5.22) required for separation. To
create the first step, we start from the point XD = YD on the line X = Y, and move
horizontally left to the equilibrium line, then vertically down to the rectifying
operating line. (This sequence actually involves using the material balance of the
rectifying operating line with the equilibrium data.) The step procedure continues
until the feed point is reached. At this juncture, the material balance switches to the
stripping line. Steps are continued until the XB intercept is reached.

FIGURE 5.22. The McCabe-Thiele method graphically determines the number


of stages (the steps drawn in this diagram) needed for a separation by
determining the relationship between a material balance and vapor-liquid
equilibrium for one component, as shown here. The solution starts with an
equilibrium diagram, such as Figure 5.15, which is the top limit for the steps.
The feed composition XF is at the junction of two mass balance lines, one for the
rectifying section above the feed point of the column and the other for the
stripping section below the feed point. The rectifying line intersects the point X D
= YD (the mole fraction of A in the liquid and in the vapor of the distillate D) and

20/85
the y axis at a point determined by the reflux ratio Ro and the composition of the
distillate (XD/(RD + I)). The lower line, limiting the stages, is the stripping line,
which must intersect the rectifying and feed lines and

XB = YB for the bottoms B product. (Adapted from reference 46.)

FIGURE 5.23. The equation that defines the feed line

(eq 5.45, Y = [-(I - f )/f ]x + XF/f ) is determined by the condition of the feed. Feed
may be cold, saturated liquid, saturated vapor or superheated vapor (see Table
5.5 for values of f ). (Adapted from reference 51.)

21/85
EXAMPLE 5.4. DESIGNING A FRACTIONATING COLUMN WITH
DIFFERENT FEED CONDITIONS

We need to design a continuous fractionating distillation column to separate 3.78 kg/s


of 40% benzene and 60% toluene into an overhead product containing 97 mass
percent benzene and a bottom product containing 98 mass percent toluene. We must
use a reflux ratio of 3.5 mol to 1 mol of product. The latent molal heat of both ben -
zene and toluene is about 357.1 kJ/kg. We will consider the following cases for the
design: (1) liquid feed at its boiling point, (2) liquid feed at 20 C ( C liquid = 1.75
kJ/kg-C), and (3) feed is two-thirds vapor.

First, we determine the necessary material balances, mole fractions, etc.

Now the McCabe-Thiele diagram can be drawn. First, we plot the equilibrium line
and X = Y line together with the vertical XB, XF, and XD (Figure 5.24). Next we locate
the rectifying operating line (passes through XD = YD and the y axis intercept of XD/
(RD + 1) or 0.974/(3.5 + 1) = 0.216).

The f value for case 1 (liquid feed at the boiling point) is zero, giving a vertical feed
line. The intercept of the feed line and the rectifying operating line gives one intercept
for the stripping operating line (the other intercept is XB = YB).

For case 2, f is given by

22/85
TABLE 5.5 Feed Values for Varying Feed Conditions

FIGURE 5.24. Example 5.4 uses a McCabe-Thiele diagram to determine the


number of stages needed to separate benzene and toluene, given different feed
scenarios. This diagram solves for the feed at its boiling point.

The McCabe-Thiele diagram for case 2 (feed line slope of 3.70) is shown in Figure
5.25.

Finally, for case 3, the f value is 2/3 (because 2/3 of the feed is vapor), thus giving a
slope of -1/2 for the feed line. The resultant solution is given in Figure 5.26. There are
eleven ideal stages plus a reboiler with feed entering on the seventh plate from the top
for cases 1 and 3. Case 2 gives 9 ideal stages plus reboiler with feed on the fifth stage
from the top.

23/85
Minimum and Maximum Stages In An Ideal Stage Separation

The two limiting conditions in distillation are the minimum number of plates (total
reflux) and the minimum reflux (infinite number of plates). The minimum number of
plates or stages can be determined (Figure 5.27) by using the line x = y as the
operating line for both rectifying and stripping. Minimum reflux can be determined
graphically (Figure 5.28). When one or both operating lines touch the equilibrium
line, an infinite number of stages will result because we cannot infinitely subdivide
matter. Understanding how to find the stages needed for minimum reflux is important
because the actual reflux ratio used in distillation columns is 1.3 to 1.7 times the
minimum (a figure of 1.5 is frequently used).

FIGURE 5.25. In Example 5.4, the McCabe-Thiele diagram for feed at 20 C.

24/85
FIGURE 5.26. In Example 5.4, the McCabe-Thiele diagram for feed that is two-
thirds vapor.

Relative Volatility-Using Enthalpy- Concentration Diagrams

A useful quantity in describing distillation is the relative volatility AB defined as

(5.46)

This equation for relative volatility indicates the ease or difficulty of achieving
separation of A and B (large or small a values indicate easy separation while a value
of 1 means A and B are inseparable). It also provides a means of generating
equilibrium data because is a fairly constant quantity at a given total pressure.

FIGURE 5.27. If the x = y line of an equilibrium diagram is used as the lower


limits of a McCabe-Thiele solution for both the rectifying and stripping sections

25/85
(the condition of total reflux), the number of stages will be minimum. (Adapted
from reference 46.)

FIGURE 5.28. If the reflux is minimum, the number of stages needed for
separation will be infinite, according to a McCabe-Thiele diagram.

The Ponchon-Savarit Method Of Nonconstant Molal Mass Transfer

The McCabe-Thiele method relies on the assumptions of constant molal vaporization


and constant molal overflow When we cannot assume constant molal overflow (i.e.,
when there are significant thermodynamic nonidealities in the system), the overall
enthalpy balance (eq 5.40) is important for determining the number of stages needed
in separation equipment. In such cases, a graphical solution is found by using
enthalpy-concentration diagrams that include equilibrium lines (Figure 5.29); this is
called the Ponchon-Savarit method. Descriptions of the techniques can be found in a
number of sources (6,8,14); the method is described briefly here and a full description
is provided in reference 6:

1. Start with an enthalpy concentration diagram (Figure 5.29), with saturated


vapor and liquid lines (the bold curves).

26/85
FIGURE 5.29. When constant molal overflow in a separation process cannot be
assumed (as in the McCabe-Thiele method) and the energy balance becomes
necessary, the number of stages needed can be determined by the Ponchon-
Savarit method. The method is complex and is not covered in detail in this text.
For further information, see references 6, 8, and 14.

2. Assume that the distillate and bottoms products (D and B) are at bubble
temperatures that fall on the saturated liquid line.

3. The feed F can be at any thermal condition; it is shown in Figure 5.29 as a


mixture of liquid and vapor, with enthalpy, HF, and concentration, xF.

4. The process is nonadiabatic; for an adiabatic condition on this figure, product


streams D and B must be corrected for the heat removed by the condenser (-q c)
and the heat added by the reboiler (qr):

In the diagram, points D' and B' represent corrected product streams. Line B'FD' is
the overall enthalpy line.

5. Either -qc (usually the case) or qr is taken to be independent; -qc is fixed by the
reflux ratio

for which Hy1 is the specific enthalpy of vapor leaving the first stage.

27/85
6. An enthalpy operating line can then be constructed for each plate: in Figure
5.29, the line LnVn+1 describes the transfer taking place between plates n and
n+1, where Ln is the liquid going from plate n to plate n+1, and Vn + 1 is the
vapor leaving plate n+1 to go to plate n.

7. In a stepwise procedure, the number of ideal plates can be determined.

Estimating The Stages Needed For A Multicomponent Separation With Relative


Volatilities

The design of a distillation column becomes complex when the system involves
separation of more than two components. Separation of more than two components
requires stage-by-stage calculations that carefully balance mass and energy, a
laborious, time-consuming task best handled by a computer. There are methods,
however, of estimating the number of stages needed for a given multicomponent
system.

The first step in the process is to use the relative volatilities of the components to
evaluate the system. Each volatility is considered relative to the least volatile
component. For example, consider a system of five components (Table 5.6).

TABLE 5.6. Relative Volatilities for a System with More Than Two Components

The components closest to the point of separation are B and C; most of B is in the
overhead stream and a little in bottoms and vice versa for C. This means that B is the
light key (Ik) and C is the heavy key (hk).

To separate a multicomponent mixture such as that as shown in Table 5.6, the


minimum steps or stages are given (15) by

(5.47)

28/85
where X represents the desired mole fractions of the heavy or light keys in the
distillate (D) and bottom (W) products. The ' is the volatility of the light key relative
to the heavy key (in the case cited above a' equals 3.6/1.7). Just as we assume for
binary mixtures, a minimum number of stages assumes total reflux.

Next, for minimum reflux (14),

(5.48)
The s are volatilities relative to the least volatile, the XDs are the mole fractions in
the overhead product, and is an empirical constant. The constant is obtained by
trial and error (14) from

(5.49)

where the f is the feed condition and the XFs are the mole fractions of the components
in the feed. Again minimum reflux, just as with binary mixtures implies an infinite
number of stages.

Once both Smin and Rmin are known, we can determine the number of stages needed for
any rellux ratio using a graphical correlation (Figure 5.30).

Differential Distillation For A Batch Distillation

The quantitative treatment of batch distillation is difficult and complex because the
distillate composition changes with time. One aspect of these distillations that can be
treated more simply is that of differential distillation. In differential distillation, the
liquid is vaporized and each segment of vapor is removed from contact with the liquid
as it is formed. Although the vapor can be in equilibrium with the liquid as it is
formed, the average vapor formed will not be in equilibrium with the liquid residue.

For an amount of liquid -dW to be vaporized,

(5.50)

(5.51)

(5.52)

29/85
FIGURE 5.30. When a system with more than two components is to be
separated, we can estimate the number of stages needed using a correlation
between Smin, the minimum number of stages needed for the separation, and Rmin,
the minimum reflux needed for separation. For further information, see eqs 5.47
through 5.49 and Table 5.6.

Finally,

(5.53)

Similarly for any two components in a differential distillation,

(5.54)

where A and B are the moles in the still at time t, and A0 and B0 are the moles at the
start of the distillation. The is the volatility of A relative to B.

Extractive Distillation-Adding A Third Component To Improve Volatilities

When substances are difficult to separate, we can add a third component to increase
system relative volatilities. The extractive agent has an affinity for one of the
components. The resulting mixture of the component and extractive agent constitutes
the bottoms in the distillation column. Some typical binary systems that rely on an

30/85
extractive agent are shown in Table 5.7, and a schematic of a typical extractive
distillation system is shown in Figure 5.31.

FIGURE 5.31. When a binary system is hard to separate, a third component, an


extractive agent, can be added to facilitate distillation. The agent, which usually
has an affinity for the bottoms product, can be added either to the feed or the top
of the distillation column as shown. The agent is recycled by distillation of the
bottoms in a column separate from the main column.

TABLE 5.7. Difficult-To-Separate Binary Systems and Their Extractive Agents

Graphical Solution Of Absorption In A Plate Column

31/85
Absorption is a separation process based on a vapor-liquid system. When absorption
is carried out in a column with plates or stages, the number of stages can be
determined by the McCabe-Thiele method as if for a distillation column that has only
a stripping section (with no rectification).

We can use Figure 5.32 for a mass balance. In absorption, we convert the mole
fractions and flow rates to a solute-free basis if the system is not dilute. This process
makes the mass balance lines linear. The primed quantities are created by the
conversion Y' = [Y/(1 - Y)] and X' = [X/(1 - X)]. For dilute cases, Y = Y' and X = X'.

Then,

(5.55)

where LM and GM are the molal mass velocities in pound-moles per hour-square foot
for solute-free liquid and gas streams.

If we do a mass balance on an overall basis, then

(5.56)

For a dilute gas

(5.57)

In solving absorption graphically, we use principles similar to those for the McCabe-
Thiele method. First, we locate the equilibrium line (Figure 5.33). Then we locate the
operating line (its slope is LM /GM or LM/GM and it passes through X0, Y, and Xn,
Yn+1 (Figure 5.33)). Stages are then stepped off. In Figure 5.33, four stages are
required to effect the separation.

In absorption, the condition of minimum liquid rate corresponds to minimum reflux.


Both rates require an infinite number of stages. This value is found by drawing the
operating line so that it touches the equilibrium line at Xn, Yn (Figure 5.34).

The McCabe-Thiele approach works for plate-type absorbers because thermal effects
are low (the component, found in the gas in small amounts, is being absorbed into a
large mass of liquid). Example 5.5 considers the design of a plate absorber column. In
this example, we use the concept of 1.5 times a minimum value of liquid flow rate as
an operating parameter.

32/85
FIGURE 5.32. Absorption is a separation process, based on a liquid-vapor
system, that can be treated like distillation with only a stripping process. The
countercurrent process has the liquid feed, L, fed into the top of the column, and
a gas, G, fed into the bottom. The McCabe-Thiele method can be used to
determine the number of plates needed.

Example 5.5. Graphical Estimation Of The Number Of Stages Needed For An


Extraction

We want to remove alcohol vapor (0.01 mole fraction) from a carbon dioxide gas
stream. Water for the absorption contains 0.0001 mole &action of alcohol. A total of
227 moles of gas are to be treated per hour. The equilibrium relationship for alcohol
and water is given by Y = 1.0682 X. For this case, how many theoretical plates would
be required for 98% absorption at a liquid rate of 1.5 times minimum?

At minimum liquid rate, the operating line would intersect the equilibrium line at

33/85
FIGURE 5.33. Application of the McCabe-Thiele method to an absorption
column begins with an equilibrium line for the solute. The operating line must
pass through X0 (mole fraction of the solute in the feed of the absorbing liquid)
and Y1 (the mole fraction solute in the exiting gas), and through X N (the mole
fraction in the exiting liquid) and YN+1 (the mole fraction of solute in the feed
gas).

34/85
FIGURE 5.34. In absorption, minimum liquid rate corresponds to minimum
reflux in distillation. To find the number of stages needed for a minimum liquid
rate, the stages must be drawn using an operating line that intersects X" and Y".

and

Hence, point C in Figure 5.35 is (0.009362, 0.01).

Point A is determined by X0 and Y1. The X0 value is given as 0.0001 in the problem
statement and Y1 by mass balance.

35/85
FIGURE 5.35. Given the desired liquid flow rate for an absorption process, we
can determine the number of plates needed in the process by using the McCabe-
Thiele method, as illustrated in Example 5.5, for which this diagram is the
solution. First, point C is determined by assuming a minimum liquid flow rate.
Then, point A is determined by mass balance. With the slope of line AC, the
numerical value of the minimum flow rate is found, leading to the desired flow
rate. Point B is determined last, and the stages are drawn. (Adapted from
reference 10.)

From L we can calculate the slope of a material or mass balance line as 357/224.78 or
1.588. Point A (0.0001, 0.000202) represents one end of the mass balance as well as
the composition at the top of the tower. The other end of the line is point B
(composition at the bottom of the tower). For this point, Y is 0.01. Hence, line AB
passes through (0.0001, 0.000202) with a slope of 1.588 and ends at Y = 0.01. Now
the number of stages or plates can be stepped off from either A or B to yield 9 stages.

The solution for the number of plates can also be obtained algebraically, which uses
the relationship shown in Figure 5.36.

36/85
CHAPTER 5 - UNIT OPERATIONS INTRODUCTION TO MASS
TRANSFER

Mass transfer operations include distillation, absorption, stripping, drying, and


extraction. These operations are widely used in various chemical and petrochemical
separation processes. During such an operation either one (or more) component(s) in
the vapor phase is transferred to the liquid phase and/or from the liquid phase to the
vapor phase. Examples include:

Distillation is used in every oil refinery and also in many chemical


manufacturing plants in the separation and purification of the desired products.
Distillation is the most important separation technique, in general. Distillation
columns are very visible in the skyline of any refinery and many chemical
plants.
Absorption is used for removal of gaseous hydrogen sulfide and/or carbon
dioxide and/or mercaptans from Natural Gas, Synthesis Gas, etc. (any process
gas stream, really) by dissolving them in a (reacting) liquid stream.
Stripping is the reverse of Absorption. Stripping of dissolved and
contaminating volatile organic components (VOC's) from (ground) water is a
good and actual example.
Drying (removal of water vapor) from Natural Gas by dehydrating liquids (i.e.
Triethyleneglycol TEG) is an actual process example.
Extraction is a liquid/liquid contacting operation, and is used in the petroleum
industry (to separate aromatics and aliphatic species) and in the
pharmaceutical industry (to recover penicillins).

These operations are usually performed in cylindrical columns. These columns come
in a wide range of sizes, their diameter ranges from 0.05 m (for a typical laboratory
scale column) to about 10 to 12 m (for the largest industrial columns) and their height
ranges from about 0.5 m to about 100 m. The required contacting needed for the
separation(s) is provided by filling these columns with packings and/or trays.
Packings (either structured or random) come in many types and sizes and so do trays.

37/85
Commonly used tray types include:

Bubble cap trays,


Sieve tray,
Valve trays, etc.

Distillation

A process in which a liquid or vapor mixture of two or more substances is


separated into its component fractions of desired purity, by the application and
removal of heat.

Distillation is based on the fact that the vapor of a boiling mixture will be richer in the
components that have lower boiling points. Therefore, when this vapor is cooled and
condensed, the condensate will contain more volatile components. At the same time,
the original mixture will contain more of the less volatile material. Distillation
columns are designed to achieve this separation efficiently.

Although many people have a fair idea what distillation means, the important
aspects that seem to be missed from the manufacturing point of view are that:

distillation is the most common separation technique

it consumes enormous amounts of energy, both in terms of cooling and heating


requirements

it can contribute to more than 50% of plant operating costs

The best way to reduce operating costs of existing units is to improve their efficiency
and operation via process optimization and control. To achieve this improvement, a
thorough understanding of distillation principles and how distillation systems are
designed is essential.

The purpose of this course is to expose you to the terminology used in distillation
practice and to give a very basic introduction to:

types of columns

basic distillation equipment and operation

column internals

38/85
reboilers

distillation principles

vapor liquid equilibria

distillation column design and

the factors that affect distillation column operation

Basic Distillation Equipment And Operation

Main Components of Distillation Columns

Distillation columns are made up of several components, each of which is used either
to transfer heat energy or enhance material transfer. A typical distillation contains
several major components:

a vertical shell where the separation of liquid components is carried out

column internals such as trays/plates and/or packings which are used to


enhance component separations

a reboiler to provide the necessary vaporization for the distillation process

a condenser to cool and condense the vapor leaving the top of the column

a reflux drum to hold the condensed vapor from the top of the column so that
liquid (reflux) can be recycled back to the column

The vertical shell houses the column internals and together with the condenser and
reboiler, constitutes a distillation column. A schematic of a typical distillation unit
with a single feed and two product streams is shown below:

39/85
Basic Operation and Terminology

The liquid mixture that is to be processed is known as the feed and this is introduced
usually somewhere near the middle of the column to a tray known as the feed tray.
The feed tray divides the column into a top (enriching or rectification) section and a
bottom (stripping) section. The feed flows down the column where it is collected at
the bottom in the reboiler.

Heat is supplied to the reboiler to generate vapor. The source of heat input can be any
suitable fluid, although in most chemical plants this is normally steam. In refineries,
the heating source may be the output streams of other columns. The vapor raised in
the reboiler is re-introduced into the unit at the bottom of the column. The liquid
removed from the reboiler is known as the bottoms product or simply, bottoms.

40/85
The vapor moves up the column, and as it exits the top of the unit, it is cooled by a
condenser. The condensed liquid is stored in a holding vessel known as the reflux
drum. Some of this liquid is recycled back to the top of the column and this is called
the reflux. The condensed liquid that is removed from the system is known as the
distillate or top product.

Thus, there are internal flows of vapor and liquid within the column as well as
external flows of feeds and product

Column Internals

Trays and Plates

Usually, trays are horizontal, flat, specially prefabricated metal sheets, which are
placed at a regular distance in a vertical cylindrical column.

Trays have two main parts:

1) The part where vapor (gas) and liquid are being contacted; the contacting area
2) The part where vapor and liquid are separated, after having been contacted;
the downcomer area.

Classification of trays is based on:

type of plate used in the contacting area


type and number of downcomers making up the downcomer area
direction and path of the liquid flowing across the contacting area of the tray
vapor (gas) flow direction through the (orifices in) the plate
presence of baffles, packing or other additions to the contacting area to
improve the separation performance of the tray

Common plate types, for use in the contacting area:

Bubble cap tray. Caps are mounted over risers fixed on the plate. These caps
come in a wide variety of sizes and shapes, round, square, rectangular (tunnel),
etc.

41/85
Sieve trays come with different hole shapes (round, square, triangular,
rectangular (slots), star), various hole sizes (from ~2 mm to ~25 mm) and
several punch patterns (triangular, square, rectangular).
Valve tray come in a variety of valve shapes (round, square, rectangular,
triangular), valve sizes, valve weights (light and heavy), orifice sizes and
either as fixed or floating valves.
Combinations of these types are applied: plates with both sieve openings and
valves, as well as plates with both light and heavy valves.

Trays usually have one or more downcomers. Dual flow trays, i.e. trays without
downcomers, are used in exceptional fouling services, only. (Examples: the Stone &
Webster Ripple tray and the Shell Turbogrid tray).

Type and number of downcomers used mainly depends on the amount of downcomer
area required to handle the liquid flow. Single pass trays are trays with one
downcomer delivering the liquid from the next higher tray, a single bubbling area
across which the liquid passes to contact the vapor and one downcomer for the liquid
to the next lower tray.

The terms "trays" and "plates" are used interchangeably. There are many types of tray
designs, but the most common ones are:

Bubble cap trays A bubble cap tray has riser


or chimney fitted over each hole, and a cap
that covers the riser. The cap is mounted so
that there is a space between riser and cap to
allow the passage of vapor. Vapor rises
through the chimney and is directed
downward by the cap, finally discharging
through slots in the cap, and finally bubbling
through the liquid on the tray.

42/85
Valve trays

In valve trays, perforations are covered by


liftable caps. Vapor flows lifts the caps, thus
self creating a flow area for the passage of
vapor. The lifting cap directs the vapor to
flow horizontally into the liquid, thus
providing better mixing than is possible in
sieve trays.
Sieve trays

Sieve trays are simply metal plates with


holes in them. Vapor passes straight upward
through the liquid on the plate. The
arrangement, number and size of the holes
are design parameters.

Because of their efficiency, wide operating range, ease of maintenance and cost
factors, sieve and valve trays have replaced the once highly thought of bubble cap
trays in many applications.

Liquid and Vapor Flows in a Tray Column

The next few figures show the direction of vapor and liquid flow across a tray, and
across a column.

43/85
Each tray has 2 conduits, one on each side, called downcomers. Liquid falls through
the downcomers by gravity from one tray to the one below it. The flow across each
plate is shown in the above diagram on the right.

A weir on the tray ensures that there is always some liquid (holdup) on the tray and is
designed such that the holdup is at a suitable height, e.g. such that the bubble caps are
covered by liquid.

Being lighter, vapor flows up the


column and is forced to pass through
the liquid, via the openings on each
tray. The area allowed for the passage
of vapor on each tray is called the
active tray area.

The picture on the left is a photograph of a


section of a pilot scale column equipped with
bubble capped trays. The tops of the 4 bubble
caps on the tray can just be seen. The down-
comer in this case is a pipe, and is shown on
the right. The frothing of the liquid on the
active tray area is due to both passage of vapor
from the tray below as well as boiling.

As the hotter vapor passes through the liquid on the tray above, it transfers heat to the
liquid. In doing so, some of the vapor condenses adding to the liquid on the tray. The
condensate, however, is richer in the less volatile components than is in the vapor.
Additionally, because of the heat input from the vapor, the liquid on the tray boils,
generating more vapor. This vapor, which moves up to the next tray in the column, is

44/85
richer in the more volatile components. This continuous contacting between vapor and
liquid occurs on each tray in the column and brings about the separation between low
boiling point components and those with higher boiling points.

Tray Designs

A tray essentially acts as a mini-column, each accomplishing a fraction of the


separation task. From this we can deduce that the more trays there are, the better the
degree of separation and that overall separation efficiency will depend significantly on
the design of the tray. Trays are designed to maximize vapor-liquid contact by
considering the

liquid distribution on the tray and

vapor distribution on the tray

This is because better vapor-liquid contact means better separation at each tray,
translating to better column performance. Fewer trays will be required to achieve the
same degree of separation. Attendant benefits include less energy usage and lower
construction costs.

There is a clear trend to improve separations by supplementing the use of trays by


additions of packings.

Packings

Packings are passive devices that are designed to increase the interfacial area for
vapor-liquid contact. The following pictures show 3 different types of packings.

These strangely shaped pieces are supposed to impart good vapor-liquid contact when
a particular type is placed together in numbers, without causing excessive pressure-

45/85
drop across a packed section. This is important because a high pressure drop would
mean that more energy is required to drive the vapor up the distillation column.

Packings versus Trays

A tray column that is facing throughput problems may be de-bottlenecked by


replacing a section of trays with packings. This is because:

packings provide extra inter-facial area for liquid-vapor contact

efficiency of separation is increased for the same column height

packed columns are shorter than trayed columns

Packed columns are called continuous-contact columns while trayed columns are
called staged-contact columns because of the manner in which vapor and liquid are
contacted.

Column Reboilers

There are a number of designs of reboilers. It is beyond the scope here to delve into
their design principles. However, they can be regarded as heat-exchangers that are
required to transfer enough energy to bring the liquid at the bottom of the column to
boiling point. The following are examples of typical reboiler types.

46/85
47/85
Distillation Principles

Separation of components from a liquid mixture via distillation depends on the


differences in boiling points of the individual components. Also, depending on the
concentrations of the components present, the liquid mixture will have different
boiling point characteristics. Therefore, distillation processes depends on the vapor
pressure characteristics of liquid mixtures.

Vapor Pressure and Boiling

The vapor pressure of a liquid at a particular temperature is the equilibrium pressure


exerted by molecules leaving and entering the liquid surface. Here are some important
points regarding vapor pressure:

energy input raises vapor pressure

vapor pressure is related to boiling

a liquid is said to boil when its vapor pressure equals the surrounding
pressure

the ease with which a liquid boils depends on its volatility

liquids with high vapor pressures (volatile liquids) will boil at lower
temperatures

the vapor pressure and hence the boiling point of a liquid mixture depends on
the relative amounts of the components in the mixture

distillation occurs because of the differences in the volatility of the


components in the liquid mixture

The Boiling Point Diagram

The boiling point diagram shows how the equilibrium compositions of the
components in a liquid mixture vary with temperature at a fixed pressure. Consider an
example of a liquid mixture containing 2 components (A and B) - a binary mixture.
This has the following boiling point diagram.

48/85
The boiling point of A is that at
which the mole fraction of A is 1.
The boiling point of B is that at
which the mole fraction of A is 0. In
this example, A is the more volatile
component and therefore has a lower
boiling point than B. The upper
curve in the diagram is called the
dew-point curve while the lower one
is called the bubble-point curve.

The dew-point is the temperature at which the saturated vapor starts to condense.

The bubble-point is the temperature at which the liquid starts to boil.

The region above the dew-point curve shows the equilibrium composition of the
superheated vapor while the region below the bubble-point curve shows the
equilibrium composition of the subcooled liquid.

For example, when a subcooled liquid with mole fraction of A=0.4 (point A) is
heated, its concentration remains constant until it reaches the bubble-point (point B),
when it starts to boil. The vapor evolved during the boiling has the equilibrium
composition given by point C, approximately 0.8 mole fraction A. This is
approximately 50% richer in A than the original liquid.

This difference between liquid and vapor compositions is the basis for distillation
operations.

Relative Volatility

Relative volatility is a measure of the differences in volatility between 2 components,


and hence their boiling points. It indicates how easy or difficult a particular separation
will be. The relative volatility of component i with respect to component j is
defined as

49/85
yi = mole fraction of component i in the vapor

xi = mole fraction of component i in the liquid

Thus if the relative volatility between 2 components is very close to one, it is an


indication that they have very similar vapor pressure characteristics. This means that
they have very similar boiling points and therefore, it will be difficult to separate the
two components via distillation.

Vapor Liquid Equilibria

Distillation columns are designed based on the boiling point properties of the
components in the mixtures being separated. Thus the sizes, particularly the height, of
distillation columns are determined by the vapor liquid equilibrium (VLE) data for the
mixtures.

Vapor-Liquid-Equilibrium (VLE) Curves

Constant pressure VLE data are obtained from boiling point diagrams. VLE data of
binary mixtures is often presented as a plot, as shown in the figure on the right. The
VLE plot expresses the bubble-point and the dew-point of a binary mixture at
constant pressure. The curved line is called the equilibrium line and describes the
compositions of the liquid and vapor in equilibrium at some fixed pressure.

This particular VLE plot shows a binary


mixture that has a uniform vapor-liquid
equilibrium that is relatively easy to
separate.

50/85
The next two VLE plots below on the other hand, show non-ideal systems which will
present more difficult separations. We can tell from the shapes of the curves and this
will be explained further later on.

The most intriguing VLE curves are generated by azeotropic systems. An azeotrope is
a liquid mixture which when vaporized, produces the same composition as the liquid.
The two VLE plots below, show two different azeotropic systems, one with a
minimum boiling point and one with a maximum boiling point. In both plots, the
equilibrium curves cross the diagonal lines, and this are azeotropic points where the
azeotropes occur. In other words azeotropic systems give rise to VLE plots where the
equilibrium curves crosses the diagonals.

Note the shapes of the respective equilibrium lines in relation to the diagonal lines
that bisect the VLE plots.

51/85
Both plots are however, obtained from homogenous azeotropic systems. An azeotrope
that contains one liquid phase in contact with vapor is called a homogenous azeotrope.
A homogenous azeotrope cannot be separated by conventional distillation. However,
vacuum distillation may be used as the lower pressures can shift the azeotropic point.
Alternatively, an additional substance may add to shift the azeotropic point to a more
favorable position.

When this additional component appears in appreciable amounts at the top of


the column, the operation is called azeotropic distillation.

When the additional component appears mostly at the bottom of the column,
the operation is called extractive distillation

The VLE curve on the left is also generated by


an azeotropic system, in this case a
heterogeneous azeotrope. Heterogeneous
azeotropes can be identified by the flat portion
on the equilibrium diagram.

They may be separated in 2 distillation columns since these substances usually form
two liquid phases with widely differing compositions. The phases may be separated
using settling tanks under appropriate conditions.

Next, we will look at how VLE plots/data are used to design distillation columns.

Distillation Column Design

As mentioned, distillation columns are designed using VLE data for the mixtures to
be separated. The vapor-liquid equilibrium characteristics (indicated by the shape of
the equilibrium curve) of the mixture will determine the number of stages, and hence
the number of trays, required for the separation. This is illustrated clearly by applying
the McCabe-Thiele method to design a binary column.

52/85
McCabe-Thiele Design Method

The McCabe-Thiele approach is a graphical one, and uses the VLE plot to determine
the theoretical number of stages required to effect the separation of a binary mixture.
It assumes constant molar overflow and this implies that:

molal heats of vaporization of the components are roughly the same

heat effects (heats of solution, heat losses to and from column, etc.) are
negligible

for every mole of vapor condensed, 1 mole of liquid is vaporized

The design procedure is simple. Given the VLE diagram of the binary mixture,
operating lines are drawn first.

Operating lines define the mass balance relationships between the liquid and
vapor phases in the column.

There is one operating line for the bottom (stripping) section of the column,
and on for the top (rectification or enriching) section of the column.

Use of the constant molar overflow assumption also ensures the operating
lines are straight lines.

Operating Line for the Rectification Section

The operating line for the rectification section is constructed as follows. First the
desired top product composition is located on the VLE diagram, and a vertical line
produced until it intersects the diagonal line that splits the VLE plot in half. A line
with slope R/(R+1) is then drawn from this intersection point as shown in the diagram
below.

R is the ratio of reflux flow (L) to distillate flow (D) and is called the reflux ratio and
is a measure of how much of the material going up the top of the column is returned
back to the column as reflux.

53/85
Operating Line for the Stripping Section

The operating line for the stripping section is constructed in a similar manner.
However, the starting point is the desired bottom product composition. A vertical line
is drawn from this point to the diagonal line, and a line of slope L s/Vs is drawn as
illustrated in the diagram below.

Ls is the liquid rate down the stripping section of the column, while Vs is the vapor
rate up the stripping section of the column. Thus the slope of the operating line for the
stripping section is a ratio between the liquid and vapor flows in that part of the
column.

54/85
Equilibrium and Operating Lines

The McCabe-Thiele method assumes that the liquid on a tray and the vapor above it
are in equilibrium. How this is related to the VLE plot and the operating lines is
depicted graphically in the diagram on the right.

A magnified section of the operating line for the stripping section is shown in relation
to the corresponding nth stage in the column. L's are the liquid flows while V's are the
vapor flows. x and y denote liquid and vapor compositions and the subscripts denote
the origin of the flows or compositions. That is 'n-1' will mean from the stage below
stage 'n' while 'n+1' will mean from the stage above stage 'n'. The liquid in stage 'n'
and the vapor above it are in equilibrium, therefore, x n and yn lie on the equilibrium
line. Since the vapor is carried to the tray above without changing composition, this is
depicted as a horizontal line on the VLE plot. Its intersection with the operating line
will give the composition of the liquid on tray 'n+1' as the operating line defines the
material balance on the trays. The composition of the vapor above the 'n+1' tray is
obtained from the intersection of the vertical line from this point to the equilibrium
line.

55/85
Number of Stages and Trays

Doing the graphical construction repeatedly will give rise to a number of 'corner'
sections, and each section will be equivalent to a stage of the distillation. This is the
basis of sizing distillation columns using the McCabe-Thiele graphical design
methodology as shown in the following example.

Given the operating lines for both stripping and rectification sections, the graphical
construction described above was applied. This particular example shows that 7
theoretical stages are required to achieve the desired separation. The required number
of trays (as opposed to stages) is one less than the number of stages since the
graphical construction includes the contribution of the reboiler in carrying out the
separation.

The actual number of trays required is given by the formula:

(Number of theoretical trays) / (tray efficiency)

Typical values for tray efficiency range from 0.5 to 0.7 and depend on a number of
factors, such as the type of trays being used, and internal liquid and vapor flow
conditions. Sometimes, additional trays are added (up to 10%) to accommodate the
possibility that the column may be under-designed.

56/85
The Feed Line (q-line)

The diagram above also shows that the binary feed should be introduced at the 4 th
stage. However, if the feed composition is such that it does not coincide with the
intersection of the operating lines, this means that the feed is not a saturated liquid.
The condition of the feed can be deduced by the slope of the feed line or q-line. The
q-line is that drawn between the intersection of the operating lines, and where the feed
composition lies on the diagonal line.

Depending on the state of the feed, the feed lines will have different slopes. For
example,

q = 0 (saturated vapor)

q = 1 (saturated liquid)

0 < q < 1 (mix of liquid and


vapor)

q > 1 (subcooled liquid)

q < 0 (superheated vapor)

The q-lines for the various feed conditions are shown in the diagram on the left.

Using Operating Lines and the Feed Line in McCabe-Thiele Design

If we have information about the condition of the feed mixture, then we can construct
the q-line and use it in the McCabe-Thiele design. However, excluding the
equilibrium line, only two other pairs of lines can be used in the McCabe-Thiele
procedure. These are:
feed-line and rectification section operating line
feed-line and stripping section operating line
stripping and rectification operating lines

This is because these pairs of lines determine the third.

57/85
Overall Column Design

Determining the number of stages required for the desired degree of separation and
the location of the feed tray is merely the first steps in producing an overall
distillation column design. Other things that need to be considered are tray spacing;
column diameter; internal configurations; heating and cooling duties. All of these can
lead to conflicting design parameters. Thus, distillation column design is often an
iterative procedure. If the conflicts are not resolved at the design stage, then the
column will not perform well in practice. The next set of notes will discuss the factors
that can affect distillation column performance.

Effects of the Number of Trays or Stages

Here we will expand on the design of columns by looking briefly at the effects of

the number of trays, and

the position of the feed tray, and

on the performances of distillation columns.

Effects of the Number of Trays

It can be deduced from the previous section on distillation column design that the
number of trays will influence the degree of separation. This is illustrated by the
following example.

Consider as a base case, a 10 stage column. The feed is a binary mixture that has a
composition of 0.5 mole fraction in terms of the more volatile component, and
introduced at stage 5. The steady-state terminal compositions of about 0.65 at the top
(stage 1) and 0.1 at the bottom (stage 10) are shown below:

Composition Profile: 10 stages, feed at stage 5

58/85
Suppose we decrease the number of stages to 8, and keep the feed at the middle stage,
i.e. stage 4. The resulting composition profile is:

Composition Profile: 8 stages, feed at stage 4

We can see that the top composition has decreased while the bottom composition has
increased. That is, the separation is poorer.

Now, if we increase the number of stages to 12, and again introduce the feed at mid-
column, i.e. stage 6, the composition profile we get is:

Composition Profile: 12 stages, feed at stage 6

Again, the composition has changed. This time the distillate is much richer in the
more volatile component, while the bottoms have less, indicating better separation.

Thus, increasing the number of stages will improve separation.

59/85
Effect of Feed Tray Position

Here we look at how the position of the feed tray affects separation efficiency.
Suppose we have a 20 stage column, again separating a binary mixture that has a
composition of 0.5 mole fraction in terms of the more volatile component. The
terminal compositions obtained when the feed is introduced at stages 5, 10 and 15 (at
fixed reflux and reboil rates) are shown in the following plots.

Composition profile: 20 stages, feed at stage 5

Composition profile: 20 stages, feed at stage 10

60/85
Composition profile: 20 stages, feed at stage 15
As the feed stage is moved lower down the column, the top composition becomes less
rich in the more volatile component while the bottoms contains more of the more
volatile component. However, the change in top composition is not as marked as the
bottoms composition.

The preceding examples illustrate what can happen if the position of the feed tray is
shifted for this particular system. They should not be used to generalize to other
distillation systems, as the effects are not straightforward.

Factors Affecting Distillation Column Operation

The performance of a distillation column is determined by many factors, for example:

feed conditions
state of feed
composition of feed
trace elements that can severely affect the VLE of liquid mixtures

internal liquid and fluid flow conditions

state of trays (packings)

weather conditions

Some of these will be discussed below to give an idea of the complexity of the
distillation process.

Feed Conditions

61/85
The state of the feed mixture and feed composition affects the operating lines and
hence the number of stages required for separation. It also affects the location of feed
tray. During operation, if the deviations from design specifications are excessive, then
the column may no longer be able handle the separation task. To overcome the
problems associated with the feed, some columns are designed to have multiple feed
points when the feed is expected to containing varying amounts of components.

Reflux Conditions

As the reflux ratio is increased, the gradient of operating line for the rectification
section moves towards a maximum value of 1. Physically, what this means is that
more and more liquid that is rich in the more volatile components are being recycled
back into the column. Separation then becomes better and thus fewer trays are needed
to achieve the same degree of separation. Minimum trays are required under total
reflux conditions, i.e. there is no withdrawal of distillate.

On the other hand, as reflux is decreased, the operating line for the rectification
section moves towards the equilibrium line. The pinch between operating and
equilibrium lines becomes more pronounced and more and more trays are required.
This is easy to verify using the McCabe-Thiele method.

The limiting condition occurs at minimum reflux ratio, when an infinite number of
trays will be required to effect separation. Most columns are designed to operate

62/85
between 1.2 to 1.5 times the minimum reflux ratio because this is approximately the
region of minimum operating costs (more reflux means higher reboiler duty).

Vapor Flow Conditions

Adverse vapor flow conditions can cause


a) foaming
b) entrainment
c) weeping/dumping
d) flooding

a) Foaming

Foaming refers to the expansion of liquid due to passage of vapor or gas.


Although it provides high interfacial liquid-vapor contact, excessive foaming
often leads to liquid buildup on trays. In some cases, foaming may be so bad
that the foam mixes with liquid on the tray above. Whether foaming will occur
depends primarily on physical properties of the liquid mixtures, but is
sometimes due to tray designs and condition. Whatever the cause, separation
efficiency is always reduced.

b) Entrainment

Entrainment refers to the liquid carried by vapor up to the tray above and is
again caused by high vapor flow rates. It is detrimental because tray efficiency
is reduced: lower volatile material is carried to a plate holding liquid of higher
volatility. It could also contaminate high purity distillate. Excessive
entrainment can lead to flooding.

c) Weeping/Dumping

This phenomenon is caused by low vapor flow. The pressure exerted by the
vapor is insufficient to hold up the liquid on the tray. Therefore, liquid starts to
leak through perforations. Excessive weeping will lead to dumping. That is the
liquid on all trays will crash (dump) through to the base of the column (via a
domino effect) and the column will have to be re-started. Weeping is indicated
by a sharp pressure drop in the column and reduced separation efficiency.

d) Flooding

63/85
Flooding is brought about by excessive vapor flow, causing liquid to be
entrained in the vapor up the column. The increased pressure from excessive
vapor also backs up the liquid in the downcomer, causing an increase in liquid
holdup on the plate above. Depending on the degree of flooding, the
maximum capacity of the column may be severely reduced. Flooding is
detected by sharp increases in column differential pressure and significant
decrease in separation efficiency.

Column Diameter

Most of the above factors that affect column operation is due to vapor flow
conditions: either excessive or too low. Vapor flow velocity is dependent on column
diameter. Weeping determines the minimum vapor flow required while flooding
determines the maximum vapor flow allowed, hence column capacity. Thus, if the
column diameter is not sized properly, the column will not perform well. Not only
will operational problems occur, the desired separation duties may not be achieved.

State of Trays and Packings

Remember that the actual number of trays required for a particular separation duty is
determined by the efficiency of the plate, and the packing, if packing is used. Thus,
any factors that cause a decrease in tray efficiency will also change the performance
of the column. Tray efficiencies are affected by fouling, wear and tear and corrosion,
and the rates at which these occur depends on the properties of the liquids being
processed. Thus appropriate materials should be specified for tray construction.

Weather Conditions

Most distillation columns are open to the atmosphere. Although many of the columns
are insulated, changing weather conditions can still affect column operation. Thus the
reboiler must be appropriately sized to ensure that enough vapors can be generated
during cold and windy spells and that it can be turned down sufficiently during hot
seasons. The same applies to condensers.

These are some of the more important factors that can cause poor distillation column
performance. Other factors include changing operating conditions and throughputs,

64/85
brought about by changes in upstream conditions and changes in the demand for the
products. All these factors, including the associated control system, should be
considered at the design stages because once a column is built and installed, nothing
much can be done to rectify the situation without incurring significant costs. The
control of distillation columns is a field in its own right, and is beyond the scope of
this course.

Tray Efficiency, in general

The separation performance of a tray is the basis of the performance of the column as
a whole. The primary function of, for instance, a distillation column is the separation
of a feed stream in (at least) one top product stream and one bottom product stream.
The quality of the separation performed by a column can be judged from the purity of
the top and bottom product streams. The specification of the impurity levels in the top
and bottom streams and the degree of recovery of pure products set the targets for a
successful operation of a distillation column.

Earlier developed mathematical models for the separation behavior of distillation


columns used the concept of equilibrium stages (theoretical trays) and incorporated
the tray efficiency, either as the Fenske Overall efficiency (Eo) or the Murphree vapor
efficiency (Emv). Newer computing models (like ChemSep and Aspen
Ratefrac) use non-equilibrium stages i.e. the rate based approach, which is a better
more general approach, esp. for multicomponent systems. This approach uses
information on flow rates, mass transfer coefficients, interfacial area and liquid hold
up, which can conveniently be combined in one variable prosaically called; the NTU
(Number of Transfer Units). The rate based approach is more computationally
intensive, however, and also requires much more input information. At the moment,
both approaches are in use, while the rate based approach appears to be gaining
ground on the traditional equilibrium stage method, because its reliability is
increasing.

In due course, typical tray efficiency values have become available from plant test
runs and various experimental programs (usually executed with well-defined binary
test systems). Perry's Chemical Engineers Handbook (table 18-4) gives representative
plate efficiencies, Emv's, which all vary between 60 - 120 %. These typical efficiency

65/85
values depend to a more or lesser extend on the test systems used, the tray type and
geometry tested, the liquid submergence, etc., etc. Picking the right value for a new
column design is still largely based on experience, but may be underpinned by one (or
more) of the following methods:

data interpolation (from previous tests/experience and/or from published


literature).
direct scale up from efficiencies measured in a
a. Oldershaw column;
b. pilot plant column;
c. existing commercial column.
empirical prediction methods (Drickamer & Bradford; O'Connell; Bakowski).
theoretical prediction methods (the procedure given in AIChE Bubble Tray
Design Manual and the Chan & Fair correlation).

When test data are available (or become available), the first two methods are
preferred, because of the 'proven' nature of the numbers involved and because the
uncertainty in these methods is the least.

Summarizing available prior experience, it has become evident that tray efficiency (or
NTU) can be influenced by:

The specific component under consideration (this holds especially for multi-
component systems in which the efficiency can be different for each
component, because of different diffusivities, diffusion interactions, different
stripping factors, etc.).
The vapor flow rate; usually increasing the flow rate increases the effective
mass transfer rate, while it decreases the contact time, at the same time. These
counteracting effects lead to a roughly constant efficiency value, for a tray in
its normal operating range. Upon approaching the lower operating limit a tray
starts weeping and looses efficiency. The tray also looses efficiency upon
approaching the upper operating limit, because of droplet entrainment.
The liquid flow rate; usually increasing the flow rate increases the tray liquid
hold up and contact time and hence tray efficiency.

66/85
The outlet weir height; higher weirs raise the liquid level, increase the
interfacial area and contact time giving improved efficiencies.
The operating pressure; tray efficiency increases with pressure. This apparent
pressure effect may be a reflection of the increase in liquid hold up (either
because of relative larger liquid flow rates or relatively less weir lengths).
The type of tray appears to be of secondary importance, as tray efficiencies of
valve and sieve tray are quite comparable, in their normal operating ranges.
The free area (fractional hole area); efficiency increases with a reduction in
free area.
The liquid flow path length; longer path lengths enhance efficiency, as liquid
mixing is less able to flatten out the axial liquid composition profile
developing, across the entire tray.
The presence of non-uniform flow distributions for the liquid and/or vapor
flow reduce tray efficiency.
The presence of dead water regions (stagnant zones) is detrimental to tray
efficiency; especially so on trays were these regions on successive trays "stack
up" (because they lie directly above each other) and allow part of the vapor
flow to effectively bypass contact with liquid across more trays.
The surface tension and surface tension gradients; a lower surface tension
increases the interfacial area and hence efficiency. Surface tension gradients
can either lower or enhance efficiency depending on the sign (direction) of the
gradient. In the liquid continuous flow regime, a positive gradient can enhance
efficiency (reduces bubble coalescence and increases interfacial area). A
negative gradient diminishes efficiency for the opposite reasons.
The liquid viscosity; there is no conclusive evidence for the effect of liquid
viscosity on tray efficiency, in distillation. In absorption, this situation can be
different for components for which the mass transfer becomes controlled by
liquid phase mass transfer.

There are still a number of Unknowns remaining, i.e. topics on which consensus has
not yet been reached. For instance; the best way to correlate the vapor phase mass
transfer coefficient; the mechanisms determining the amount of interfacial area
available for mass transfer; the 'stubborn' use of the inadequate Francis' weir equation
in the calculation of liquid hold up on a tray operating in the spray regime; the

67/85
estimation of the contribution of the liquid phase mass transfer resistance; and last but
not least the influence of the two phase flow regime.

It will be appreciated from the above, that the estimation of distillation tray efficiency
is wrought with uncertainties and inaccuracies. Hence, confidence in the validity of
available data is important.

Because of the considerable commercial value attached to these efficiency numbers,


the proper numbers have been (and are) slow in becoming available, publicly. Either
they are considered to be proprietary know how or are kept confidential (as company-
sponsored FRI does), because of the high costs involved in obtaining them. This has
slowed down the scientific development of fundamental theories of mass transfer on
trays (which could have been of benefit to the companies involved, as well).

Operating range of trays

The successful operation of a refinery and chemical plant depends on the operability
of its separation columns. The range of operation of the installed trays governs the
maximum and minimum gas and liquid loads a column can handle and consequently
the capacity and turndown capability of a plant.

In the illustrative operating diagram below, an operating line as well as an operating


point are indicated.

For both the vapor and the liquid flow rate, lower and upper limits exist. Several
different hydraulic mechanisms control these limits. The operating point of a column
should be chosen by carefully considering these limitations.

The physical mechanisms behind the most common limitations are:

Entrainment flooding is caused by an excessive liquid flow rate generated by


droplets carried out of the gas-liquid dispersion on the tray and up to the next
tray by the gas stream. Preferentially, this occurs on trays operating in the
spray regime, when most of the liquid is present as droplets.
Bed expansion flooding (priming) sets in when the gas-liquid layer on the
tray extends up to the next higher tray, i.e. the dispersion height becomes equal

68/85
to the tray spacing. This condition depends on the liquid height on a tray and
its expansion, by the gas flowing through it. Actually, this condition is
equivalent to entrainment flooding. As the dispersion level approaches the
next higher tray, drop entrainment rises rapidly, causing accumulation of liquid
on the next tray and increasing the liquid flow, for the downcomers to cope
with.
Choking (blockage) of the downcomer entrance can set a limit to the liquid
flow rate, when the inflowing gas-liquid dispersion needs more space, when a
large amount of gas is set free in the downcomer. The separated gas rises
counter to the downflowing gas-containing liquid and can induce a (vapor)
blocking condition.
Downcomer overflowing (flooding) sets a limit to the liquid handling
capacity of a downcomer. This happens when the height of the gas-liquid layer
in a downcomer exceeds the height of the downcomer. This condition is
controlled by the height of liquid in the downcomer and by the expansion
caused by entrapped gas bubbles. The liquid height is given by the condition
that the hydrostatic liquid head should be sufficient to make the liquid flow
down against the tray pressure drop.
Weeping (raining) of liquid occurs when gas flowing through the perforations
in the tray floor is no longer able to counterbalance the hydrostatic head of
liquid on a tray and liquid starts leaking.
Sealing (dumping) of the downcomers requires, that at least some liquid flows
through them. This condition is fulfilled when the gas-liquid layer on a tray is
sufficiently expanded by the gas flow to make liquid flow over the outlet weir.
Blowing occurs when insufficient liquid flows through the downcomers to
prevent gas from bypassing up through them. This is detrimental to the
separation performance of a tray, because a significant part of the gas bypasses
the contacting area of the tray in this way. During the (re-) start up phase of a
column, this phenomenon needs special attention, as the penalty will be a
severe underperformance of the column (in both throughput and separation
performance).

Hydraulics and Two Phase Flow regimes on trays

69/85
The full spectrum of all possible two phase flow regimes can be encountered on trays,
depending on flow conditions, system properties and tray geometry. In general, four
two phase flow regimes are recognized to classify the flow patterns in the contacting
area:

Bubbling; gas bubbles in liquid, < ~ 50 % gas content, wide bubble size
distribution.

Foaming; gas bubbles in liquid, > ~ 65% gas content, fairly uniform bubble
sizes.

Churn turbulent / Compounded regime; ~ 20 to 60 % liquid content, at


bottom gas in liquid and at top droplets in gas.

Spray; droplets in gas, < ~ 25 % liquid content, wide drop size distribution.

The Bubbly Flow regime occurs at lower gas velocities and relatively thick liquid
layers (while bubbles or jets forming at the tray floor are fully submerged). The
bottom layer has the highest liquid content (may be 80-90 % liquid) and the formation
of bubbles or jets at the orifices can be seen clearly. The upper layer has higher gas
content and is in constant chaotic motion. The bubble population has a wide size
distribution; large bubbles may be a few centimeter (~0.02 - 0.05 m) in diameter,
while small bubbles measure a few millimeter (~0.002 - 0.005 m). Chaotic (turbulent)
motion causes large bubbles to break up, while coalescence of colliding bubbles
makes them grow. The wide bubble size distribution is the result of a dynamic
equilibrium between break up and coalescence. The chaotic motion is caused by the
rising bubbles and makes the surface level fluctuate constantly, as well.

The bubbly flow regime can be found:

in high pressure absorption or distillation; over whole operating range.

in medium pressure distillation; from the lower limit up to somewhere in the


middle of the operating range.

in sub atmospheric or vacuum pressure distillation: only at operation near the


lower limit.

70/85
The bubbly liquid flows over the outlet weir into the downcomer entrance. On its way,
the liquid is degassing, as bubbles keep escaping from the liquid phase. The gas
content of the liquid falling into the surface level in the downcomer will depend on
the rate of degassing and its original gas content (on exit from the contacting zone).
This degassing process causes an upflow of gas out of the downcomer. This upflow of
gas may interfere with the downflow of liquid. Particularly, when the dispersion
height becomes tall in comparison to the horizontal width of the downcomer (the
available distance for the throw of liquid becomes too small) and choking of the
entrance occurs.

The two phase dispersion sitting in the downcomer has two strata, usually. A clear
bottom layer with an upper bubbly layer. In the upper gas-liquid layer, bubbles come
in with the falling liquid and rise up and out of the dispersion, again. Thereby
generating the upflow of gas out of the downcomer.

The Foaming regime, also, occurs at lower gas velocities and relatively thick liquid
layers (while bubbles or jets forming at the tray floor are fully submerged). The
difference with the Bubbly Flow regime is caused by inhibition of bubble coalescence,
as break up of liquid films in between the bubbles is retarded by the presence of
surface tension gradients, very fine solids, a high liquid viscosity, etc. This leads to an
increased residence time of the gas in the dispersion, thus giving rise to a high gas
content. The degassing process is inhibited as well.

Commonly, the bottom layer is fairly clear. However, for fairly strongly foaming
systems, it can turn into a 'milky' emulsion containing many tiny bubbles (< ~ 50 %
gas content).

The mobile foam layer on top contains a fair amount of motion. Its structure is that of
a typical foam with fairly stable liquid lamella in between bubbles, of various sizes.
The gas content can be quite high, up to 90 - 95 %.

The foamy liquid falling over the outlet weir into the downcomer entrance takes along
a large amount of gas, as a result of the inhibited bubble coalescence and hampered
degassing. A much larger amount of gas now becomes available from the degassing
process in the downcomer. This gives rise to a larger upflow of gas out of the

71/85
downcomer, which can interfere even more strongly with the volumetrically enlarged
incoming foamy liquid. Thus leading to earlier inducement of choking phenomena
and hence premature flooding of the downcomer.

The two phase dispersion sitting in the downcomer again has two strata, but their
relative contributions differ. The upper bubbly (foamy) layer is enlarged, as an
enlarged gas flow has to be separated, while the separation takes more time, as well.
Very small bubbles may not separate at all and can be 'carried under', with the liquid
leaving the downcomer exit.

The Churn turbulent regime (or Compounded regime) might also be called a
'sandwiched' regime, as it is intermediate to both the bubbly regime and the spray
regime. In the structure of this two phase dispersion, these regimes can still be
identified: the bubbly regime in the bottom layer and the spray regime in the upper
layer. Often, an approximately uniform intermediate layer is present, as well. The
dispersion is very chaotic and full of fluctuations of many size and frequencies scales.
The 'bubble' size distribution is even wider than in the bubbly flow regime, as now
large clusters of bubbles or even 'slugs' (large voids) can be seen rising at high
velocity through the two phase layers.

The churn turbulent regime can be found:

in high pressure absorption or distillation; over the whole operating range.


in medium pressure distillation; from somewhat above the lower operating
limit up to somewhere before the upper limit.
In sub atmospheric or vacuum pressure distillation: somewhat above the lower
limit.

The liquid flow passing into the downcomer entrance is composed of a part
contributed by drops thrown over (out of the spray layer) and a part contributed by a
disintegrating bubbly liquid flowing over (out of the liquid continuous bottom layer).
Their relative contributions vary, depending on the operating conditions and the
specifics of the outlet weir geometry. The entering liquid flow is composed of a very
wide range of 'drop' sizes falling down and impacting on the surface level. Gas

72/85
bubbles are generated on impact, in addition to being taken along as the bubbly part in
the liquid.

Again, the dispersion in the downcomer has a two layered structure, with a bubbly
upper layer and a fairly clear liquid layer at the bottom.

The Spray regime occurs at higher gas velocities and relatively thin liquid layers
(while bubbles or jets forming at the tray floor are only partly submerged). The
bottom layer has the highest liquid content. The high velocity of the gas flowing out
the perforations in the tray floor atomizes the liquid and accelerates the droplets
upward. The drops can acquire substantial initial propagation velocities. Because the
vapor transferred its kinetic energy and momentum to the liquid phase, the result is a
fairly chaotic (randomized) movement of droplets, with a very wide distribution in
sizes, ranging from several cm (~ 0.02 m) for the largest drops, to a few hundred (~
0.0001 m) micron for the smallest droplets. A fraction of the very smallest droplets is
dragged along by the gas flow and carried away to the next higher tray, as droplet
entrainment (or carry over).

The upper layer is primarily a droplet propagating layer, whose height depends on the
initial propagation velocity acquired by the drops and the shearing interaction
between the moving drops and upflowing gas. Ultimately, the dispersion height is
constrained by the spacing between successive trays, which is somewhere between
0.3 to 0.9 m high. The liquid content of this upper layer typically ranges from ~1 -
~15 %.

The spray regime can be found:

in high pressure absorption or distillation; only at operation near the upper


limit.
in medium pressure distillation; from somewhere in the middle of the
operating range up to the upper limit.
in sub atmospheric or vacuum pressure distillation: starting somewhat above
the lower limit it occurs over the whole operating range.

73/85
The liquid passes into the downcomer entrance by splashing over the outlet weir.
Hence, a cloud of droplets rains down in and impacts on the liquid surface in the
downcomer. On impact some gas is entrained and bubbles forming from entrained gas
are dragged down for some distance, before they escape the impacting liquid flow and
rise up and separate. The upper layer in the downcomer will (again) be a bubbly
dispersion, while the bottom layer will be fairly clear liquid.

Tray pressure drop

Typical tray pressure drops lie in the range of 250 - 1500 N/m 2 (or 2.5 mbar - 15 mbar
or 25 - 150 mm Water Column, in whatever units one prefers).

Usually, the drop in pressure caused by gas flowing through a tray is small in
comparison to the system pressure. Except for vacuum columns, where it can become
quite substantial and the gas velocity in the perforations may become comparable to
the velocity of sound.

The tray pressure drop plays an important part in filling up the downcomers. To
compensate for the pressure drop, a liquid head builds up in the downcomers, to
enable the liquid to flow down against it. When the tray pressure drop becomes
excessive with respect to the height of the downcomers, flooding will be the result.

The tray pressure drop is composed of (at least) two (major) contributions:

1. A pressure drop caused by the gas flowing through the perforations in the tray
floor. This contribution depends on gas flow rate, fraction free area and the
pressure drop coefficient of the particular perforations (or valves) being used.
This pressure drop coefficient depends on relative hole thickness (i.e. the ratio
of tray thickness over hole diameter), hole shape and nearness of other holes
(ratio of hole pitch to hole diameter).
2. A pressure drop caused by the liquid present on the tray. This liquid hold up
effect primarily increases with an increase in outlet weir height, decreases with
an increase in gas flow rate and increases with an increase in liquid flow rate.
To a lesser extent, it depends on physical properties of the gas/liquid system.

74/85
The figure shows a typical tray pressure drop curve for a particular sieve tray. The
tray pressure drop is shown in comparison to the pressure drop caused by gas flowing
through a tray without liquid on it; the dry tray pressure drop. The gas velocity in the
holes is given here as the hole loadfactor, which essentially is the gas hole velocity
times the square root of the ratio of the gas over the liquid density.

The pressure drop of a valve tray differs from the sieve tray pressure drop mainly,
because the gas flowing through the valves in the tray floor experiences a different
flow resistance. A valve tray dry pressure drop curve exhibits two horizontally shifted
parabolas connected by a horizontal plateau. Operating at this plateau, the force
exerted on the valves by the gas pressure drop overcomes the weight of the valves.

Weeping, sealing & blowing: the lower operating limit(s)

For assessment of the lower operating limit, three ranges in gas flow rate should be
considered.

In the normal operating range, the gas flow rate exceeds the critical weep point gas
flow rate needed to stop liquid from leaking through the perforations in the tray. All
liquid goes in cross flow from the inlet to the outlet downcomer. The separation
performance of a tray is at its best in this range.

In the weeping range, the gas flow rate is below the critical weep point gas flow rate
and above the critical seal point gas flow rate. The liquid changes from flowing
across the tray and over the outlet weir into flowing counter currently to the gas
through the perforations. The separation performance remains good (at the same level
as in its normal operating range) quite some way down in gas rate (and up in weep
rate).

75/85
In the dumping range, the gas flow rate is below the critical seal point (or dump
point) gas flow rate. All gas and liquid flow counter currently through the
perforations. A tray operating in its dumping range operates similar to a
downcomerless tray, i.e. a dual flow tray. The separation performance falls off
significantly upon operation below the seal point.

The hole loadfactor is essentially the gas hole velocity times the square root of the
ratio of gas density over liquid density.

The weep point gas flow rate (wp).

The weep point gas flow rate is the gas flow rate where the first leakage (raining) of
liquid occurs, because the gas flowing through the perforations is no longer able to
counterbalance the hydrostatic head of liquid on a tray. The perforations are no longer
exclusively used by the gas. The weep point gas flow rate (or weep point for short) is
fairly well defined and numerous studies have reported their values and several
empirical correlations are available in the open literature. As a rule of thumb for sieve
trays, the gas flow rate expressed as loadfactor (= capacity factor) based on hole area
usually lies between 0.30 - 0.40 m/s. Recommended can be the article by M.J. Lockett

76/85
and S. Banik, "Weeping from Sieve Trays", Ind. Eng. Chem. Process Des. Dev.,
25(1986)25, pp 561-569.

The seal point (or dump point) gas flow rate (sp).

The seal point (or dump point) gas flow rate is the gas flow rate where the gas-liquid
layer on a tray is just expanded enough to make some liquid flow over the outlet
weir into the downcomer. At a lower gas rate, the dispersion height is less than the
weir height and all liquid is leaking away through the perforations (countercurrent to
the gas). Although the seal point gas flow rate (or seal point for short) is fairly well
defined, only a few studies are available in the open literature.

Upward blowing of gas through the downcomers occurs when insufficient liquid
flows through them. This is detrimental to the separation performance of a tray,
because a significant part of the gas bypasses the contacting area of the tray in this
way. During the (re-) start up phase of a column, this phenomenon needs special
attention, as in this mode of operation, the penalty can be a severe underperformance
of the column (in both throughput and separation performance). A proper tray design
can take care of this potentially deleterious situation.

Flooding, priming & entrainment: the upper operating limit(s)

The common mechanisms governing the upper limit of operation of a tray are:

Entrainment flooding is caused by an excessive liquid flow rate generated by


droplets carried out of the gas-liquid dispersion by the gas stream and up to the next
higher tray. Preferentially, this occurs on trays operating in the spray regime, when
most of the liquid is already present as droplets.

Bed expansion flooding (priming) sets in when the gas-liquid layer on the tray
extends up to the next tray, i.e. the dispersion height becomes equal to the tray
spacing. This condition depends on the liquid height on a tray and the expansion by
the gas flowing through it. Actually, this condition is 'the twin brother' of entrainment
flooding. As the dispersion level approaches the next higher tray, droplet entrainment
rises rapidly as well, causing an increasing liquid level on the next tray, thereby
increasing the liquid flow through the downcomers.

77/85
Choking (blockage) at the downcomer entrance (or in the downcomer itself) can set a
limit to the liquid flow rate, as the inflowing gas-liquid dispersion needs more space,
in the case when a large amount of gas has to be set free in the downcomer. The
separated gas rises counter to the downflowing gas-containing liquid and can induce a
vapor blocking condition.

Downcomer overflowing (flooding) sets a limit to the liquid handling capacity of a


downcomer. This happens when the height of the gas-liquid layer in a downcomer
exceeds the height of the downcomer. This condition is controlled by the height of
liquid in the downcomer and by the expansion caused by entrapped gas bubbles. The
liquid height is given by the condition that the hydrostatic liquid head should be
sufficient to make the liquid flow down against the tray pressure drop.

A downcomer velocity limitation occurs when the liquid downflow velocity


becomes large enough to stop gas bubbles from rising out of the liquid level and
begins to drags them down (resulting in carry under of these gas bubbles). The
downcomer gets filled up with a 'frothy' mass, which takes up more space than pure
liquid.

The open literature on upper operating limits (or maximum capacities) of sieve and
valve trays is fairly large and several decades old already. Because of its importance,
the subject has been reviewed several times and their results have become text book
material as empirical design rules and design graphs, see: Henry Z. Kister,
Distillation Design, McGraw-Hill, Inc., 1992, and Section 18. Liquid-gas Systems by
J.R. Fair in Perry's Chemical Engineers Handbook, McGraw-Hill, Inc...

Rather unfortunately, a generic and unifying hydrodynamic theory of two phase flow
of gas and liquid on trays has not (yet) been developed. Consequently, a theory
predicting the upper operating limit(s) is still lacking and we will have to make the
best use we can of past experience, as embodied by these empirical design rules and
graphs as given.

Usually, the maximum column loadfactor for a conventional sieve tray lies in the
range of 0.06 to 0.09 m/s, depending on tray spacing (typically 0.6 m), liquid load,
weir height, etc. For conventional valve trays, the maximum capacity is about the

78/85
same or somewhat lower. Specially developed high capacity trays can handle
substantially more, however.

The column loadfactor is essentially the column gas velocity times the square root of
the ratio of gas density over liquid density.

79/85
Absorption and Stripping

Absorption is one of the oldest unit operations used in the gas processing industry.
Rich gas enters the bottom of the absorber and flows upward contacting the
countercurrent lean oil stream. The lean oil preferentially absorbs the heavier
components from the gas and is then termed "rich oil". The rich oil is sent to a stripper
(or still) where the absorbed components are removed by heating and/or stripping
with steam. The lean oil is recycled to the absorber to complete the process loop.

For a given gas, the fraction of each component in the gas that is absorbed by the oil
is a function of the equilibrium phase relationship of the components and lean oil, the
relative flow rates, and the contact stages. The phase relation is a function of pressure,
temperature, and the composition of the lean oil.

As components are absorbed, the temperature of the gas and oil phases will increase
due to heat of absorption. The heat released is proportional to the amount of gas
absorbed. In many cases, side coolers are used on the absorber to limit the
temperature rise and aid in absorption.

Lean oil will have a molecular weight in the 100 to 200 range. For ambient
temperature absorbers, a heavy lean oil of 180 to 200 molecular weight will normally
be used. For refrigerated absorbers, a lighter lean oil of 120 to 140 molecular weight
is used. Lower molecular weight lean oil will contain more moles per gallon resulting
in a lower circulation rate.

80/85
However, lower molecular weight lean oil will have higher vaporization losses.

The stripping column is operated at low pressure and high temperatures. In older
plants, "live" steam is injected into the bottom of the column to strip the NGL
components. The steam lowers the partial pressure of the light hydrocarbons which is
equivalent to lowering the effective operating pressure.

Refrigerated lean oil plants normally use direct fired heaters to vaporize a portion of
the rich oil in the stripper (still) to provide the necessary stripping vapor.

Absorber Calculations

Absorber and stripper calculations, like fractionation column calculations, can be


accomplished with tray-by-tray computer models. However, hand calculations can be
performed to estimate the absorption of components in a lean oil absorber.

The stripping operation is essentially the reverse of absorption and can be handled in a
similar fashion.

Many attempts have been made to define an "average" absorption factor method to
short-cut the time consuming rigorous calculation procedures. The sole restriction of
such a method is how well the average factor, as it is defined, will represent the
absorption that actually occurs. One of the simplest definitions of an average
absorption factor was by Kremser and Brown. They defined it as:

A=Lo/(kavgVn+1)

or Lo=A(Kavg)(Vn+1)

The use of an average absorption factor, as defined in the equation above, ignores the
change in gas volume from inlet to outlet.

Using an average absorption factor, the extraction of any component from a rich gas
can be described by:

Yn 1 Y1 An 1 A
n 1 Ea
Yn 1 Yo A 1

A graphical solution of the formula above is given below:

81/85
Also, the assumptions of average temperature and K-values can cause significant
errors in the preceding calculation method.

The above figure can also be used to determine the trays required for a given lean oil
rate or to calculate recoveries with a given oil rate and tray count. It shows that oil
rate declines with increasing number of trays and that beyond about eight theoretical
trays little increase in efficiency is achieved.

Since higher oil rates require more energy for heating, cooling, and pumping, the
optimum design is usually one that uses the minimum possible oil rate with a
reasonable size absorber.

The lowest molecular weight lean oil should be used. This will be fixed by oil vapor
pressure and absorber operating temperature.

Most problems in absorber operation center on oil quality and rates. Proper stripping
of the oil is necessary to minimize lean oil losses to the gas and to maximize
absorption capacity.

82/85
STRIPPER CALCULATIONS

In a calculation sense, a stripper is simply an upside-down absorber. For hand


calculations, a stripping factor is defined as

KV
ST
L

X m 1 X 1 STm 1 ST
m1 Ex
X m 1 X o ST 1

The same figure above can be used to perform stripper calculations in a similar
manner to absorber calculations.

SOUR WATER STRIPPERS

Sour water is a term used for water containing dissolved hydrogen sulfide. Facilities
for processing sour gas may have several sources of sour water. These include water
from inlet separators, water from compressor discharge scrubbers, quench water from
certain Claus unit tail-gas cleanup processes, and water from the regeneration of solid
bed product treaters or dehydrators. In some plants it is possible to dispose of this
water by using it for makeup to the gas treating solution.

However, most sour gas plants have an excess of water and the hydrogen sulfide must
be removed to a level of 1 to 2 ppmw before disposing of the water. Sour water
strippers are used for this purpose.

Sour water strippers commonly have 10 to 15 trays or 6 to 9 m of packing. The feed


enters at the top and heat is supplied either by a reboiler or by steam injection directly
below the bottom tray. Typical operating conditions are:
Pressure, kPa 69 - 103
Feed Temperature, C 93 - 110
Bottom Temperature, C 115 - 121
Reboil Heat, kW 223 - 446
Residual H2S, mg/kg 0.5 - 2.0
Overhead vapors from sour water strippers contain hydrogen sulfide, steam, trace
amounts of hydrocarbons and, in some plants, carbon dioxide. These vapors are
usually sent to the regenerator (still) condenser in plants using aqueous treating
solutions. Alternatively, the vapors may be sent directly to the sulfur recovery unit, or
incinerated if emission standards are not exceeded.

83/85
Foaming occurs in sour water strippers and the tower diameter should be based on
operation at 50 to 70 percent of the flooding loads for a non-foaming system.

The required number of theoretical trays and stripping vapor quantity can be
calculated as shown in the following example.

However, the results of such calculations must be used only as a guide to the relative
effects of changing vapor rates and trays. This is because tray efficiencies or packing
HETPs are not known accurately and the effects of other components in the sour
water change the apparent solubility of hydrogen sulfide. Ammonia, which is
common in refinery sour waters, can increase the hydrogen sulfide solubility by a
factor of 10 or more. A more detailed design procedure is then required for refining
sour water strippers than that given in the following example.

84/85
85/85

You might also like