Anda di halaman 1dari 19

Extractive and azeotropic distillation

Distillation uses a very simple principle that consists of 3 steps

1. Generation of a two phase system


2. Mass transfer across the interface
3. Separation of the two phases

This principle is very effective and is therefore applied to alternative separation processes eg
absorption, adsorption and extraction were the two phase system is created by addition of an
external substance (entrainer or solvent) that has to be removed later on in an additional
separation process. In distillation, the two phase system (vapou-liquid) is generated by a supply
of heat which can be easily removed later on.

AZEOTROPIC SYSTEMS

Departures from Raoults law commonly manifest themselves in the formation of azeotropes;
particularly for mixtures of close-boiling species of different chemical types whose liquid
sulutions are nonideal. Azeotropic-forming mixtures exhibit either maximum- or minimum-
boiling points at some composition, corresponding, respectively, to negative and positive
deviations from Raoults law. Vapor and liquid compositions are identical for an azeotrope; thus,
all Kvalues are 1, AB = 1, and no separation can take place.

Therefore, an Azeotrope is a mixture of liquids that has a constant boiling point because the
vapour has the same composition as the liquid mixture. The boiling point of an azeotropic
mixture may be higher or lower than that of any of its components. The components of the
solution cannot be separated by simple distillation. If only one liquid phase exists, it is a
homogeneous azeotrope; if more than one liquid phase is present, the azeotrope is
heterogeneous. By the Gibbs phase rule, at constant pressure in a two-component system, the
vapor can coexist with no more than two liquid phases; in a ternary mixture, up to three liquid
phases can coexist with the vapor, and so on, Figures 1-3 show three types of azeotropes. The
most common by far is the minimum-boiling homogeneous azeotrope, e.g., isopropyl ether
isopropyl alcohol, shown in Figure 1a. At a temperature of 70C, the maximum total pressure is
greater than the vapor pressure of either component, as shown in Figure 1a, because activity

1
coefficients are greater than 1. The yx diagram in Figure 1b shows that for a pressure of 1 atm,
the azeotropic mixture is at 78 mol% ether. Figure 1c is a Tx diagram at 1 atm, where the
azeotrope is seen to boil at 66C. In Figure 1a, for 70C, the azeotrope occurs at 123 kPa (923
torr), for 72 mol% ether. Thus, the azeotropic composition and temperature shift with pressure.
In distillation, minimum-boiling azeotropic mixtures are approached in the overhead product.

Figure 1: Minimum-boiling-point azeotrope, isopropyl etherisopropyl alcohol system: (a)


partial and total pressures at 70_C; (b) vaporliquid equilibria at 101 kPa; (c) phase diagram at
101 kPa. [Adapted from O.A. Hougen, K.M.Watson, and R.A. Ragatz, Chemical Process
Principles. Part II, 2nd ed., John Wiley & Sons, New York (1959).]

2
For the maximum-boiling homogeneous azeotropic acetonechloroform system in Figure 2a, the
minimum total pressure at 60C is below the vapor pressures of the pure components because
activity coefficients are less than 1. The azeotrope is approached in the bottoms product in a
distillation operation. Phase compositions at 1 atm are shown in Figures 1b and c.

Figure 2: Maximum-boiling-point azeotrope, acetonechloroform system:(a) partial and total


pressures at 60C; (b) vaporliquid equilibria at 101 kPa; (c) phase diagram at 101 kPa pressure.
[Adapted from O.A. Hougen, K.M. Watson, and R.A. Ragatz, Chemical Process Principles. Part
II, 2nd ed., John Wiley & Sons, New York (1959).]

3
Heterogeneous azeotropes are minimum-boiling because activity coefficients must be
significantly greater than 1 to form two liquid phases. The region ab in Figure 3a for the water
n-butanol system is a two-phase region, where total and partial pressures remain constant as the
amounts of the phases change, but phase compositions do not. The yx diagram in Figure 3b
shows a horizontal line over the immiscible region, and the phase diagram of Figure 3c shows a
minimum constant temperature. To avoid azeotrope limitations, it is sometimes possible to shift
the equilibrium by changing the pressure sufficiently to break the azeotrope, or move it away
from the region where the required separation is to be made. For example, ethyl alcohol and
water form a homogeneous minimum-boiling azeotrope of 95.6 wt% alcohol at 78.15C and
101.3 kPa. However, at vacuums of less than 9.3 kPa, no azeotrope is formed.

Figure 3: Minimum-boiling-point (two liquid phases) water/n-butanol system: (a) partial and
total pressures at 100C; (b) vaporliquid equilibria at 101 kPa; (c) phase diagram at 101 kPa
pressure. [Adapted from O.A. Hougen, K.M. Watson, and R.A. Ragatz, Chemical Process
Principles. Part II, 2nd ed., John Wiley & Sons, New York (1959).]

4
Extractive Distillation

Extractive distillation is a partial vaporization process in the presence of a non-volatile


separating agent with a high boiling point, which is generally called solvent or separating agent,
and which is added to the azeotropic mixture to alter the relative volatility of the key component
with no additional formation of azeotropes (Perry, 1992; Black et al., 1972).

The principle behind extractive distillation is the introduction of a selective solvent that interacts

differently with each of the components of the original mixture and which generally shows a
strong affinity with one of the key components (Lee and Gendry, 1997; Doherty and Malone,
2001). In general terms, the solvent is fed at the upper zone of the column, above the feed
stream, and retains a significant concentration throughout the liquid phase. Then it is withdrawn
as bottom product and is forwarded to a vacuum-operated regeneration column.

Extractive distillation is used to separate azeotropes and close-boiling mixtures. If the feed is a
minimum-boiling azeotrope, a solvent, with a lower volatility than the key components of the
feed mixture, is added to a tray just a few trays below the top of the column so that (1) the
solvent is present in the down-flowing liquid, and (2) little solvent is stripped and lost to the
overhead vapor. If the feed is a maximum boiling azeotrope, the solvent enters the column with
the feed. The components in the feed must have different solvent affinities so that the solvent
causes an increase in a of the key components, to the extent that separation becomes feasible and
economical. The solvent should not form an azeotrope with any components in the feed. Usually,
a molar ratio of solvent to feed on the order of 1 is required. The bottoms are processed to
recover the solvent for recycle and complete thefeed separation. Extractive distillation was
introduced by Dunn et al. (1945) in connection with the commercial separation of toluene from a
paraffinhydrocarbon mixture, using phenol as solvent.

Table 1 lists industrial applications of extractive distillation.

5
Consider the case of the acetonemethanol system. Consider the case of the acetonemethanol
system. At 1 atm, acetone (nbp = 56.2C) and methanol (nbp = 64.7C) form a minimum-boiling
azeotrope of 80 mol% acetone at a temperature of 55.7C. The UNIFAC program was used to
predict vaporliquid equilibria for this system at 1 atm, the azeotrope was estimated to occur at
55.2C with 77.1 mol% acetone, reasonably close to measured values. At infinite dilution with
respect to methanol, A,M for acetone (A) with respect to methanol (M), is predicted to be 0.74 by
UNIFAC, with a liquid-phase activity coefficient for methanol of 1.88. At infinite dilution with

6
respect to acetone, A,M is 2.48; by coincidence, the liquid-phase activity coefficient for acetone
is also 1.88.

Water is a possible solvent for the system because at 1 atm: (1) it does not form a binary or
ternary azeotrope with acetone and/or methanol, and (2) it boils (100C) at a higher temperature.
The resulting residue-curve map with arrows directed from the azeotrope to pure water,
computed by ASPEN PLUS using UNIFAC, is shown in Figure 4, where it is seen that no
distillation boundaries exist. As discussed by Doherty and Caldarola (1985), this is an ideal
situation for the selection of an extractive distillation process. Their residue-curve map for this
type of system (designated 100) is included as an insert in Figure 4.

Ternary mixtures of acetone, methanol, and water at 1 atm give the following separation factors,
estimated from the UNIFAC equation, when appreciable solvent is present.

7
The presence of appreciable water increases the liquid phase activity coefficient of acetone and
decreases that of methanol; thus, over the entire concentration range of acetone and methanol,
the a of acetone to methanol is at least 2.0. This makes it possible, with extractive distillation, to
obtain a distillate of acetone and a bottoms of methanol and water. The a values of acetone to
water and methanol to water average 4.5 and 2.0, thus making it relatively easy to prevent water
from reaching the distillate, and, in subsequent operations, to separate methanol from water by
distillation.

In selecting a solvent for extractive distillation, a number of factors are considered, which
include availability, cost, corrosivity, vapor pressure, thermal stability, heat of vaporization,
reactivity, toxicity, infinite-dilution activity coefficients in the solvent of the components to be
separated, and ease of solvent recovery for recycle. In addition, the solvent should not form
azeotropes. Initial screening is based on the measurement or prediction of infinite-dilution
activity coefficients.

Berg (1969) discusses selection of separation agents for both extractive and azeotropic
distillation. He points out that all successful solvents for extractive distillation are highly
hydrogen-bonded liquids, such as (1) water, amino alcohols, amides, and phenols that form
three-dimensional networks of strong hydrogen bonds; and (2) alcohols, acids,phenols, and
amines that are composed of molecules containing both active hydrogen atoms and donor atoms
(oxygen, nitrogen, and fluorine). Solvents most commonly used in extractive distillation of
ethanol include glycols (Perry, 1992; Meirelles et al., 1992), glycerol (Lee and Pahl, 1985;
Uyazn et al., 2006), gasoline (Chianese and Zinnamosca, 1990) and, in the case of saline
extractive distillation, acetate and inorganic salts: CaCl2, AlCl3, KNO3, (CuNO3)23H2O,

8
Al(NO)39H2O, K2CO3 (Barba et al., 1985; Furter, 1992; Ligero and Ravagnani, 2003; Llano and
Aguilar, 2003; Pinto et al., 2003; Schmit and Vogelpohl, 1983). It is difficult or impossible to
find a solvent to separate components having the same functional groups.

HOMOGENEOUS AZEOTROPIC DISTILLATION

An azeotrope can be separated by extractive distillation, using a solvent that is higher boiling
than the feed components and does not form any azeotropes. Alternatively, the separation can be
made by homogeneous azeotropic distillation, using an entrainer not subject to such restrictions.
Like extractive distillation, a sequence of two or three columns is used. Alternatively, the
sequence is a hybrid system that includes operations other than distillation, such as solvent
extraction.

The conditions that a potential entrainer must satisfy have been studied by Doherty and
Caldarola (1985); Stichlmair, Fair, and Bravo (1989). If it is assumed that a distillation boundary,
if any, of a residue curve map is straight or cannot be crossed, the conditions of Doherty and
Caldarola apply. These are based on the rule that for entrainer E, the two components, A and B,
to be separated, or any product azeotrope, must lie in the same distillation region of the residue-
curve map. Thus, a distillation boundary cannot be connected to the AB azeotrope.
Furthermore, A or B, but not both, must be a saddle.

The maps suitable for a sequence that includes homogeneous azeotropic distillation together with
ordinary distillation are classified into the five groups illustrated in Figure 5a, b, c, d, and e,
where each group includes applicable residue-curve maps and the sequence of separation
columns used to separate A from B and recycle the entrainer. For all groups, the residue-curve
map is drawn, with the lowest-boiling component, L, at the top vertex; the intermediate-boiling
component, I, at the bottom-left vertex; and the highest-boiling component, H, at the bottom-
right vertex. Component A is the lower-boiling binary component and B the higher. For the first
three groups, A and B form a minimum-boiling azeotrope; for the other two groups, they form a
maximum-boiling azeotrope.

9
Figure 5: Residue-curve maps and distillation sequences for homogeneous azeotropic distillation.
(a) Group 1: A and B form a minimum-boiling azeotrope, I = E, E forms no azeotropes.
(continued )

10
Figure 5: (Continued ) (b) Group 2: A and B form a minimum-boiling azeotrope, L = E, E forms
a maximum-boiling azeotrope with A. (c) Group 3: A and B form a minimum-boiling azeotrope,
I = E, E forms a maximum-boiling azeotrope with A.

11
In Group 1, the intermediate boiler, I, is E, which forms no azeotropes with A and/or B. As
shown in Figure 5a, this case, like extractive distillation, involves no distillation boundary. Both
sequences assume that fresh feed F, of A and B, as fed to Column 1, is close to the azeotropic
composition. This feed may be distillate from a previous column used to produce the azeotrope
from the original A and B mixture. Either the direct sequence or the indirect sequence may be
used. In the former, Column 2 is fed by the bottoms from Column 1 and the entrainer is
recovered as distillate from Column 2 and recycled to Column 1. In the latter, Column 2 is fed by
the distillate from Column 1 and the entrainer is recovered as bottoms from Column 2 and
recycled to Column 1. Although both sequences show entrainer combined with fresh feed before
Column 1, fresh feed and recycled entrainer can be fed to different trays to enhance the
separation.

In Group 2, low boiler L is E, which forms a maximum boiling azeotrope with A. Entrainer E
may also form a minimum-boiling azeotrope with B, and/or a minimum-boiling (unstable node)
ternary azeotrope. Thus, in Figure 5b, any of the five residue-curve maps may apply. In all cases,
a distillation boundary exists, which is directed from the maximum-boiling azeotrope of AE to
pure B, the high boiler. A feasible indirect or direct sequence is restricted to the subtriangle
bounded by the vertices of pure components A, B, and the binary AE azeotrope. An example of
an indirect sequence is included in Figure 5b. Here, the AE azeotrope is recycled to Column 1
from the bottoms of Column 2. Alternatively, as in Figure 5c for Group 3, A and E may be
switched to make A the low boiler and E the intermediate boiler, which again forms a maximum-
boiling azeotrope with A. All sequences for Group 3 are confined to the same subtriangle as for
Group 2.

Groups 4 and 5, in Figures 5d and e, are similar to Groups 2 and 3. However, A and B now form
a maximum boiling azeotrope. In Group 4, the entrainer is the intermediate boiler, which forms a
minimum-boiling azeotrope with B. The entrainer may also form a maximum-boiling azeotrope
with A, and/or a maximum-boiling (stable node) ternary azeotrope. A feasible sequence is
restricted to the subtriangle formed by vertices A, B, and the BE azeotrope. In the sequence, the
distillate from Column 2, which is the minimum-boiling BE azeotrope, is mixed with fresh feed
to Column 1, which produces a distillate of pure A. The bottoms from Column 1 has a
composition such that when fed to Column 2, a bottoms of pure B can be produced. Although a

12
direct sequence is shown, the indirect sequence can also be used. Alternatively, as shown in
Figure 5e for Group 5, B and E may be switched to make E the high boiler. In the sequence
shown, as in that of Figure 5d, the bottoms from Column 1 is such that, when fed to Column 2, a
bottoms of pure B can be produced. The other conditions and sequences are the same as for
Group 4.

The distillation boundaries for the hypothetical ternary systems in Figure 5 are shown as straight
lines. When a distillation boundary is curved, it may be crossed, provided that both the distillate
and bottoms products lie on the same side of the boundary. It is often difficult to find an
entrainer for a sequence involving homogeneous azeotropic distillation and ordinary distillation.
However, azeotropic distillation can also be incorporated into a hybrid sequence involving
separation operations other than distillation. In that case, some of the restrictions for the entrainer
and resulting residue-curve map may not apply. For example, the separation of the close boiling
and minimum-azeotrope-forming system of benzene and cyclohexane using acetone as the
entrainer violates the restrictions for a distillation-only sequence because the ternary system
involves only two minimum-boiling binary azeotropes.

However, the separation can be made by the sequence shown in Figure 6, which involves: (1)
homogeneous azeotropic distillation with acetone entrainer to produce a bottoms product of
nearly pure benzene and a distillate close in composition to the minimum-boiling binary
azeotrope of acetone and cyclohexane; (2) solvent extraction of distillate with water to give a
raffinate of cyclohexane and an extract of acetone and water; and (3) ordinary distillation of
extract to recover acetone for recycle.

13
Figure 5: (Continued ) (d) Group 4: A and B form a maximum-boiling azeotrope, I = E, E forms
a minimum-boiling azeotrope with B. (e) Group 5: A and B form a maximum-boiling azeotrope,
H = E, E forms a minimum-boiling azeotrope with B.

14
Figure 6: Sequence for separating cyclohexane and benzene using homogeneous azeotropic
distillation with acetone entrainer. [From Perrys Chemical Engineers Handbook, 6th ed., R.H.
Perry and D.W. Green, Eds., McGraw-Hill, New York (1984) with permission.]

HETEROGENEOUS AZEOTROPIC DISTILLATION

Required for homogeneous azeotropic distillation is that A and B lie in the same distillation
region of the residue-curve map as entrainer E. This is so restrictive that it is difficult, if not
impossible, to find a feasible entrainer. The Group 1 map in Figure 5a requires that the entrainer
not form an azeotrope but yet be the intermediate-boiling component, while the other two
components form a minimum-boiling azeotrope. Such systems are rare, because most
intermediate boiling entrainers form an azeotrope with one or both of the other two components.
The other four groups in Figure 5 require formation of at least one maximum-boiling azeotrope.

15
However, such azeotropes are far less common than minimum-boiling azeotropes. Thus,
sequences based on homogeneous azeotropic distillation are rare and a better alternative is
needed. A better, alternative technique that finds wide use is heterogeneous azeotropic
distillation to separate close-boiling binaries and minimum-boiling binary azeotropes by
employing an entrainer that forms a binary and/or ternary heterogeneous (two-phase) azeotrope.
A heterogeneous azeotrope has two or more liquid phases. If it has two, the overall, two-liquid-
phase composition is equal to that of the vapor phase. Thus, all three phases have different
compositions. The overhead vapor from the column is close to the composition of the
heterogeneous azeotrope. When condensed, two liquid phases form in a decanter. After
separation, most or all of the entrainer-rich liquid phase is returned to the column as reflux, while
most or all of the other liquid phase is sent to the next column for further processing.

Because these two liquid phases usually lie in different distillation regions of the residue-curve
map, the restriction that dooms homogeneous azeotropic distillation is overcome. Thus, in
heterogeneous azeotropic distillation, the components to be separated need not lie in the same
distillation region. Heterogeneous azeotropic distillation has been practiced for a century, first by
batch and then by continuous processing. Two of the most widely used applications are (1) the
use of benzene or another entrainer to separate the minimum boiling azeotrope of ethanol and
water, and (2) use of ethyl acetate or another entrainer to separate the close-boiling mixture of
acetic acid and water. Other applications, cited by Widagdo and Seider (1996), include
dehydrations of isopropanol with isopropylether, sec-butyl-alcohol with disec-butylether,
chloroform with mesityl oxide, formic acid with toluene, and acetic acid with toluene. Also,
tanker-transported feedstocks such as benzene and styrene, which become water-saturated during
transport, are dehydrated. Consider separation of the ethanolwater azeotrope by heterogeneous
azeotropic distillation. The two most widely used entrainers are benzene and diethyl ether

Young (1902) discussed the use of benzene as an entrainer for a batch process, in perhaps the
first application of heterogeneous azeotropic distillation.

A residue-curve map, computed by Bekiaris, Meski, and Morari (1996) for the ethanol (E)water
(W)benzene (B) system at 1 atm, using the UNIQUAC equation for liquid-phase activity
coefficients (with parameters from ASPEN PLUS), is shown in Figure 7. Superimposed on the
map is a bold-dashed binodal curve for the two-liquid-phase region boundary. The normal

16
boiling points of E, W, and B are 78.4, 100, and 80.1C, respectively. The UNIQUAC equation
predicts that homogeneous minimum-boiling azeotropes AZ1 and AZ2 are formed by E and W at
78.2C and 10.0 mol% W, and by E and B at 67.7_C and 44.6 mol% E, respectively. A
heterogeneous minimum-boiling azeotrope AZ3 is predicted for Wand B at 69.3C, with a vapor
composition of 29.8 mol% W. The overall two-liquid-phase composition is the same as that of
the vapor, but each liquid phase is almost pure. The B-rich liquid phase is predicted to contain
0.55 mol% W, while the W-rich liquid phase contains only 0.061 mol% B. A ternary minimum-
boiling heterogeneous azeotrope AZ4 is predicted at 64.1_C, with a vapor composition of 27.5
mol% E, 53.1 mol% B, and 19.4 mol% W. The overall two-liquid-phase composition of the
ternary azeotrope equals that of the vapor, but a thin, dashed tie line through AZ4 shows that the
benzene-rich liquid phase contains 18.4 mol% E, 79.0 mol% B, and 2.6 mol% W, while the
water-rich liquid phase contains 43.9 mol% E, 6.3 mol% B, and 49.8 mol% W.

In Figure 7, the map is divided into three distillation regions by three thick, solid-line distillation
boundaries that extend from AZ4 to binary azeotropes at AZ1, 2, and 3. Each distillation region
contains one pure component. The ternary azeotrope is the lowest-boiling azeotrope, because it is
an unstable node. Because all three binary azeotropes boil below the boiling points of the three
pure components, the binary azeotropes are saddles and the pure components are stable nodes.
Accordingly, all residue curves begin at the ternary azeotrope and terminate at a pure-component
apex. Liquidliquid solubility (binodal curve) is shown as a thick, dashed, curved line. However,
this curve is not like the usual ternary solubility curve, because it is for isobaric, not isothermal,
conditions. Superimposed on the distillation boundary that separates Regions 2 and 3 are thick
dashes that represent vapor composition in equilibrium with two liquid phases.

Compositions of the two equilibrium-liquid phases for a particular vapor composition are
obtained from the two ends of the straight tie line that passes through the vapor composition
point and terminates at the liquid solubility curve. The only tie line shown in Figure 7 is a thin,
dashed line that passes through AZ4. Other tie lines, representing other temperatures, could be
added; however, in all distillations, a strenuous attempt is made to restrict the formation of two
liquid phases to the decanter because when two liquid phases form on a tray, the tray efficiency
decreases.

17
Figure 7: Residue-curve map for the ethanolwaterbenzene system at 1 atm.

Figure 7 clearly shows how a distillation boundary is crossed by the tie line through AZ4 to form
two liquid phases in the decanter. This phase split is utilized in a typical operation, where the
tower is treated as a column with no condenser, a main feed that enters a few trays below the top
of the column, and the reflux of benzene-rich liquid as a second feed. The composition of the
combined two feeds lies in Region 1. Thus, from the residue-curve directions, products of the
tower can be a bottoms of nearly pure ethanol and an overhead vapor approaching the AZ4
composition. When that vapor is condensed, phase splitting occurs to give a water-rich phase that
lies in Region 3 and an entrainer-rich phase in Region 2. If the water-rich phase is sent to a
reboiled stripper, the residue curves indicate that a nearly pure-water bottoms can be produced,
with the overhead vapor, rich in ethanol, recycled to the decanter. When the entrainer-rich phase
in Region 2 is added to the main feed in Region 1, the overall composition lies in Region 1. To
avoid formation of two liquid phases on the top trays of the azeotropic tower, the composition of
the vapor leaving the top tray must have an equilibrium liquid that lies outside of the two-phase-

18
liquid region in Figure 7. Shown in Figure 8, from Prokopakis and Seider (1983), are 18 vapor
compositions that form two liquid phases when condensed, but are in equilibrium with only one
liquid phase on the top tray, as restricted to the very small expanded window. That window is
achieved by adding to the entrainer-rich reflux a portion of the water-rich liquid or some
condensed vapor prior to separation in the decanter.

Figure 8: Overhead vapor compositions not in equilibrium with two liquid phases. [From J.
Prokopakis and W.D. Seider, AIChE J., 29, 4960 (1983) with permission.]

19

Anda mungkin juga menyukai