Sei sulla pagina 1di 11

------------------ -

LIQUID-LIQUID EXTRACTION PROCESSES


James R. Fair and Jimmy L. Humphrey
The University of Texas at Austin
Austin, Texas
ABSTRACT
Liquid-liquid extraction is the separation of
one or more components of a liquid solution by con-
tact with a second immiscible liquid called the
solvent. If the components in the original liquid
solution distribute themselves differently between
the two liquid phases, separation will result.
This is the principle upon which separation by
liquid-liquid extraction is based, and there are a
number of important applications of this concept in
industrial processes. This paper will review the
basic concepts and applications as well as present
future directions for the liquid-liquid extraction
process.
INTRODUCTION
The concept of separating one or more components
from a liquid mixture by means of selective solvent
extraction has been known and practiced for dec-
ades, even centuries. In contrast to distillation,
however, the technology for analysis and design of
extraction processes has been slow to develop. Ex-
traction was not included with the original chemi-
cal engineering unit operations and did not find a
place in the first edition of Chemical Engineers'
Handbook (1934). In the 1936 edition of Elements
of Chemical Engineering, by Badger and McCabe,
knowledge of extraction was summarized as follows:
[Extraction] involves operations that are in
wide use not only in chemical engineering but
also in other arts ....The theory is quite
inadequate, and few important quantitative
studies have ever been made. Consequently,
the apparatus has developed along lines dic-
tated by convenience and experience, rather
than by a theoretical analysis of the
problem....
When the book Absorption and Extraction was written
by Sherwood in 1937, only one of eight chapters
dealt with extraction. Asimilar proportion of
coverage was found in the second edition of Chemi-
cal Engineers' Handbook (1941). It is clear that
the more quantitative aspects of extraction tech-
nology have been developed within the last forty
years or so. It seems clear also that chemical
engineers prefer not to use extraction instead of,
say, distillation because of their greater famil-
iarity with the latter operation and because meth-
ods for scaleup and design of extraction processes
are much less reliable than they are for other
separation processes such as distillation.
Despite what might be termed neglect of extraction
by those who develop chemical engineering technol-
ogy, the process does hold promise to lower energy
requirements for separating some liquid mixtures.
One can visualize a solvent contacting operation
carried out at ambient temperature and pressure, in
which a minor amount of a high-boiling material is
selectively removed from a large quantity of more
volatile material. In distillation it would be
necessary to supply considerable latent heat of
vaporization to carry out such a separation. For
the extraction case it would only be necessary to
vaporize the minor amount of extracted material.
Situations of this type are often encountered in
process design. Amore quantitative discussion of
them will be given later in the presentation.
The purpose of this paper is to provide the reader
with a general review of the state of the art of
the technology for the analysis and design of ex-
traction processes. Limitations in space will not
permit a detailed examination of any of the parts
of that technology. It is hoped that the reader
will obtain some appreciation of the place of
liquid-liquid extraction in the total scheme of
separation methods useful for the separation of
liquid mixtures at commercial scales of operation.
Although extraction can be applied to solid-liquid
processes, as implied by the title of the paper,
coverage here will be limited to liquid-liquid
systems.
CURRENT APPLICATIONS
One should not conclude at this point that extrac-
tion has remained as a laboratory curiosity. There
are some very important and large-scale processes
846
ESL-IE-83-04-130

Proceedings from the Fifth Industrial Energy Technology Conference Volume II, Houston, TX, April 17-20, 1983
that utilize extraction, and perhaps the best known
of these is the one that produces virtually all of
the United States production of high-purity ben-
zene. In this process an aromatic mixture contain-
ing benzene, toluene, and xylene (called the "BTX"
fraction) is separated from close-boiling paraffins
and naphthenes in the product from reforming opera-
tions. The refining of lubricating oils is another
tonnage extraction process. Useful higher-valued
fuels are being extracted from heavy residual oils
from refinery cracking operations. Extraction is
widely used in the food and pharmaceutical indus-
tries; a modern example of such use is the removal
of caffeine from coffee beans by the use of super-
critical carbon dioxide solvent. There are many
more examples of extraction applications that could
be mentioned, but the above should provide some
perspective for the reader.
DEF INITIONS
The simplest extraction system comprises three com-
ponents: the solute, or the material to be ex-
tracted; the solvent, which must not be completely
miscible with the other liquids; and the "carrier,"
or nonsolute portion of the feed mixture to be sep-
arated. For the case of countercurrent extraction,
the flows of these materials is shown in Fig. 1.
It should be noted that distinction must be made
between the light and the heavy phase, be-
tween the disperse phase and the continuous phase,
and between the raffinate phase and the extract
phase. The terminal streams from an extractor are
the extract and the raffinate. Note also that the
location of the principal interface depends upon
which phase is dispersed.
AtYPical extraction system is shown in Fig. 2. As
contrasted with the simple systems of Fig. 1, the
feed stream is shown entering the extraction column
toward the center, and the column is provided with
reflux in the form of extract product that has been
separated from the solvent. While extraction ar-
rangements can differ, depending upon the system to
be separated as well as the type of equipment to be
used, the arrangement in Fig. 2 is presented to
enable the direct comparison with, say, a reboiled
absorption system. There is indeed a direct anal-
ogy between extraction and absorption. The strip-
per, which is a distillation column, is an impor-
tant part of the extraction system, and it can
consume large amounts of energy if the proportion
of solute in the feed is high and if a significant
amount of extract reflux is needed.
PHASE EQUILIBRIUM
The simplest ternary system is shown in Fig. 3;
this is called Type I, for one immiscible pair.
For such a system the carrier and the solvent are
essentially immiscible, while the carrier-solute
and solvent-solute pairs are miscible. The diagram
shows a single-phase region and a two-phase region;
for extraction to be feasible, compositions must be
such as to fall within the two-phase envelope.
Phase equilibrium relationships are indicated in
Fig. 3; the tie lines connect equilibrium phase
compositions and thus provide a basis for selec-
tivity:
(1 )
which will be recognized as the separation factor
equivalent to relative volatility in distillation.
:hus. SCA K
C
K
A,
where the equilibrium ratio K
1 S
(2)
Afinal point regarding the Type I system: the
Qlait tOint shown in Fig. 3 is the intersection of
the ra flnate phase and extract phase boundary
curves, and no separation can be made at that
point. The analogy is with the azeotrope composi-
tion in distillation.
FEED
EXTRACT LIGHT SOLVENT
AtC
B+C(+A)
A "CARRIER"
B SOLVENT
C SOLUTI;
(Distributed Component 1
A (+B+C) B
SOLVENT RAFFINATE
HEAVY SOLVENT
RAFFINATE SOLVENT
A(+C+Bl B
A "CARRIER"
B, SOLVENT
C ' SOLUTE
(Distributed Component l;
B+C (+Al A+C
EXTRACT FEED
Fig. 1. Extraction notation
Fig. 4 shows another type of ternary
system, one where there are immiscibilities between
solvent and solute, and between solvent and carrier
(thus, Type II). The tie lines are indicated, and
there is no plait point. With this type of system
it is possible to obtain an extract that is essenl
tially free of carrier, which is not possible
the Type I system shown in Fig. 3. There are also
a few Type III systems, in which immiscibilities
exist among all three pairs, but such systems are
relatively rare in extraction system design. For
all systems, temperature influences the locations
of the phase envelopes, and a normally immiscible
system can become completely miscible if the tem-
perature is raised suff,iciently.
Reliable liquid-liquid equilibrium data are crucial
to the rational and economic design of extraction
processes. Such data can be measured with less
difficulty than can vapor-liquid equilibria; the
R47
ESL-IE-83-04-130

Proceedings from the Fifth Industrial Energy Technology Conference Volume II, Houston, TX, April 17-20, 1983
RAFFINATE
.--,-+MAKEUP SOLVENT
N Condenser
EXTRACT
PRODUCT
EXTRACTION
COLUMN
FEED STRIPPER
EXTRACT REFLUX
Fig. 2. Extraction system, with extract reflux
Type I
System
A ~ 8 immiscible
"A Rich"
"8 Rich"
Raffinale- r ~
Ex I raeI Phase
Phase
/ (E Phase)
I
(R Phase) I.
A L - - - - - - - - - ~
(Carrier)
Fig. 3. Phase diagram, Type I system
Tie Line
Type IT
System
A f 8 } parlically
C f 8 immiscible
miscible
l----''--__L...)''---=-__~ 8
A
(Carrier) (Solvent)
Fig. 4. Phase diagram, Type II system
phases are brought to equilibrium in a suitable
container and then allowed to separate completely
before they are sampled for analysis.
Many equilibria have been published, and should be
consulted not only for possible use in scaleup for
also for calibrating equipment for laboratory mea-
surements. References 1 to Beither provide direct
data or furnish guidance to the literature where
the data can be found. Treyba1 [9] has 1i sted
representative values of the equilibrium ratio
(Eq. 2) for 278 systems. In general, the data
reported have not been subjected to a thermodynamic
consistency analysis, and allowance for errors
should be made.
STAGE CALCULATIONS
It is convenient to model, extraction processes on
an equilibrium stage basis, even if the equipment
operates in a countercurrent mode (as in packed
columns). For single-stage extractions, a mixer-
settler arrangement is used, with the stirred ves-
sel designed to provide a close approach to equi-
librium. For multiple-stage extractions, both
crosscurrent and countercurrent arrangements may be
used, as indicated in Fig. 5. The countercurrent
system is more usual and is more efficient in its
use of solvent. The stages are arranged as in dis-
tillation and it is often convenient to use distil-
lation-type equipment such as tray columns and
packed towers.
There is a minimum solvent rate that corresponds to
the minimum reflux ratio in distillation. At this
rate, an infinite number of stages would be re-
quired to make a given separation. Fig. 6 illus-
trates the minimum rate for a simple Type I system,
and also illustrates the graphical technique for
stepping off equilibrium stages. For a given feed
F and solvent R values of the extract E and raffi-
nate R can be obtained graphically, as shown. When
extract and feed are in equilibrium (the line con-
necting them coincides with a tie line), then the
resulting solvent rate is the minimum rate. If the
solvent-to-feed ratio is then increased (segment Mf
increases in proportion to segment ~ , stages can
be stepped off, using the difference point as the
pivot, as shown in Fig. 7.
Since feed and extract pass each other on the bot-
tom stage, equilibrium may limit the purity of the
extract (on a solvent-free basis). To increase
this purity, extract reflux may be used, a shown in
Fig. 8. The solvent stripper must be used in any
case, in order that the solute may go on to its
intended use and the solvent return to the extrac-
tor. Returning a portion of the extract is equiva-
lent to refluxing in distillation, and it permits a
much purer solute to be extracted. Fig. 8 shows
also how the analogy to heat in boilup is provided
by the solvent in extraction.
Another graphical approach to stage determination
embodies plotting the carrier-solute on a solvent-
free basis. Fig. 9 shows the equilibrium relation-
ships for typical Type I and Type II systems. Note
the analogy of the extract and raffinate phases to
the vapor and liquid phases in distillation. A
typical stage count for the Type I system is also
shown in the figure. The solvent-free plot is es-
pecially useful for Type II systems, and has a com-
plete analogy to the y-x McCabe-Thiele diagram for
distillation. The effect of raffinate and extract
reflux is seen readily, and the center area feed
location is easily handled.
For systems of more than three components, analyti-
cal approaches are required. It is possible, but
not practical, to handle a four-component system on
a three-dimensional plot. Rigorous-type models
848
ESL-IE-83-04-130

Proceedings from the Fifth Industrial Energy Technology Conference Volume II, Houston, TX, April 17-20, 1983
Composite
extract
FinoI
roffinote
SI S2 S,3


Solvent EN E2 Extract
Net flow to left=R-S =F-E=6
Fig. 5. Crosscurrent and countercurrent
extraction
F is Given c
Iin. FEll
S is
coincides with
F+S"R+E"M
tie line
S MF
-'=-"S"M , . II) sfages
6..R-S F-E
Fig. 6. Minimum solvent rate, infinite stages
c
4 Thea. Sloge.
,,"N'
WI
E, F
B
B
Fig. 7. Higher-than-minimum solvent rate, four
stages
S+F+Ro= RN+E
S
RN-S =F+(Ro-E
1
)
til = F + tl2
Roffinote
Stripping
E
E R
SOLVENT
1 i
STRIPPER
Enriching
Extract
R
o
s
E,
"'----------'
Fig. 8. Use of extract reflux
C Solute
A Carrier
Fig. 9. Solvent-free equilibrium diagram, typical
stages
have been developed for handling multicomponent I
systems and require computers for their solution,
In general, these models are proprietary, but a I
discussion of their approach has been published
Scheibel [101. Itappears that multicomponent mqd-
els are not used very often, partly because of the
lack of reliable multicomponent phase equilibria
and partly because the systems indicating their se
are poorly defined. It is often satisfactory to
use a pseudo-ternary system, with pseudo components
representing the properties of the design extracV
and raffinate streams.
SOLVENT SELECTION
The optimum solvent for a given separation is de-I
termined from a consideration of several
It is important to note that the "best" solvent f(lr
the laboratory or pilot plant development may not,
be feasible for the commercial plant. Some general
criteria for solvent selection, as outlined by
Treybal [11, 12J, are these:
1. Selectivity. Ahigh value of the separa-
tion factor enables fewer stages to be I
used.
2. Equilibrium ratio. Ahigh value of SCA
permits lower solvent/feed ratios.
3. Density. Ahigh density difference be-
tween extract and raffinate phases permits
higher capacities in equipment.
4. Insolubility of solvent. If the solvent
is too soluble in the raffinate, signifi-
cant solvent losses can occur.
5. Recoverability. It is desirable to make a
clean separation of extractant and solvent
in the stripper, without excessive energy
requirements.
H49
ESL-IE-83-04-130

Proceedings from the Fifth Industrial Energy Technology Conference Volume II, Houston, TX, April 17-20, 1983
--
--
6. Interfacial tension. low interfacial ten-
sion aids dispersion but hinders settling
and phase separation.
7. Toxicity, flammability. These are impor-
tant occupational health and safety consi-
derations.
8. Cost. An excellent solvent, based on
laboratory tests, may not be commercially
available or may represent a very large
initial cost for charging the system.
Also, losses occur in operating systems
and must be replaced.
There are approaches to solvent selection, at least
for preliminary screening, that are based on chemi-
cal structure and interactions between the chemical
species involved. Such approaches have been dis-
cussed recently by Robbins [13].
EXTRACTION DEVICES
There are a great many different devices that are
used commercially, many of them being proprietary
and requiring the involvement of the proprietor in
the scaleup procedure. Abasic type of device is
the mixer-settler system, useful if only a few
theoretical stages are required but tending to be
expensive if throughputs are high. Adiagram of a
mixer-settler system is show in Fig. 10. As shown,
a considerable amount of piping and pumping is re-
quired for the stages, but by suitable mixer design
100% stage efficiency can be approached. Compact
mixer-settler systems, with simple overflow-
underflow arrangements, can be designed to minimize
space and piping/pumping requirements. The mixer-
settler concept suggests a general caution in ex-
tractor design: intense agitation to provide high
rates of mass transfer and close approaches to 100%
stage efficiency can lead to liquid-liquid disper-
sions that are difficult to settle into the dis-
tinct phases. Thus, some balance between intensity
of dispersion and time of settling must be reached.
Many extractors are of the tower type, and examples
of this type are shown in Fig. 11. The simplest of
these, and the least efficient, is the spray ex-
tractor.
The spray extractor comprises a vertical vessel
with the only internal device being a distributor
for the phase to be dispersed. As shown, the sol-
vent is the heavy phase and is being dispersed
through a perforated pipe distributor. The extract
phase drops fall through the continuous phase (raf-
finate) and the solute diffuses from the continuous
phase to the dispersed phase. The spray extractor
is inexpensive, but suffers a low efficiency for
two reasons: there is considerable backmixing in
the continuous phase, thus lowering the available
concentration driving force for diffusion, and the
lack of re-formation of drops penalizes the overall
rate of mass transfer. (Experiments show that a
majority of the total mass transfer occurs during
drop formation; thus a device that causes coales-
cence/formation several times has a mass transfer
rate advantage.) Measurements show that spray ex-
tractors do not normally produce more than two or
three theoretical stages.
FINAL EXTRACT FRESH
SOLVENT
EXTRACT' HEAVY PHASE
RAFFINATE LIGHT PHASE
C COALESCER
S'SETTLER (deconter)
Fig. 10. Three-stage mixer-settler system
Perforated Plate
Phase
Heavy Heavy
Phase Phase
Interface
Light Light
Phase Phose
Heavy Phase
Interface
Coalesced
dispersed
phase
Phase
Unagitated Extractor Columns
Rotating
Scheibel Reciprocoting

Conlactor Plate
111
Mechanically Agitated Extractor Columns
Fig. 11. Representative column-type extractors
H'50
ESL-IE-83-04-130

Proceedings from the Fifth Industrial Energy Technology Conference Volume II, Houston, TX, April 17-20, 1983
The pulsed extractor became popular in the mid-
1950s, largely through experience with small-scale
units. The pulsing action is designed to create
frequent renewals of the interfacial surface,
thereby enhancing mass transfer rates. Perforated
plates in the column minimize departures from ef-
fective countercurrent flow of the phases. When
scaled up to large sizes, the pulse column was
found to suffer in efficiency because of the diffi-
culty of propagating the pulses. To correct for
this, the effect of pulsing was obtained by moving
the plates in a reciprocating fashion. Thus, the
pulse column is typified today by the Karr extrac-
tor [14], which contains a series of perforated
plates (without downcomers or upcomers) on one or
more shafts, with the assembly being given a re-
ciprocating movement.
Representative performance data for two reciprocat-
ing plate columns are given in Fig. 12, taken from
the paper by Karr and Lo [15] and based on the
o-oxylene/acetic acid/water system. The important
design variables appear to be length of stroke
("double amplitude") and reciprocating speed.
Height equivalent to a theoretical stage (HETS)
values indicate stage efficiencies of the order of
5-10%; however, it is possible to use a very low
tray spacing (one inch in the example shown) and
still maintain relatively high throughputs. Apos-
sible shortcoming of this type of device, other
than its cost, is the fact that the reciprocating
motion of the driver tends to give maintenance
problems.
The Scheibel column, marketed under the name York-
Scheibel Column, is designed to simulate a series
of mixer-settler extraction units, with self-
contained mesh-type coalescers at each contacting
stage. The dispersed phase holdup and mass trans-
fer efficiency'are controlled primarily by the
speed of the agitators. Typical data for a Schei-
bel column [16] are shown in Fig. 13 for the same
system represented in Fig. 12, the o-xylene/acetic
acid/water system. This system is considered a
"difficult" system for mass transfer because of its
relatively high interfacial tension; for "easy"
systems the stage efficiency can exceed 100%. The
reason the apparent limitation of equilibrium can
be exceeded is that a Scheibel stage is in reality
two stages--one for agitation and one for coales-
cence. The minimum HETS for the data of Fig. 13 is
about 13 cm. and is lower (i.e., efficiency higher)
when the acetic acid (solute) is transferred from
the aqueous phase to the hydrocarbon phase. Data
in Fig. 12 show the same effect of transfer direc-
tion on HETS. Flooding of the column would be
reached at about 15 rev/s for the total liquid
throughput (raffinate phase plus extract phase)
shown. Although moderately expensive, the Scheibel
column gives very high contacting efficiency.
The rotating disc contactor (ROC) was introduced in
the 1950s by the Shell companies [17] and has been
used extensively in the petroleum industry for ex-
tractions involving hydrocarbon systems. Rotors on
a central shaft create dispersion and movement of
the phases, while stators provide the countercur-
rent staging. Like the Scheibel unit, the ROC ef-
fectiveness can be controlled to some extent by
500...,.-------------..........,
a
u
CL>
...
o
CL>
-c
r
Vl 100
CL>
o 60
r U
c 40
c
CL> ell
-
30
00'
> 20
.:; (f)
0-
W 10
-C
0'
CL>
I
o 100 300 500
Re c ipracatin9 Speed
strakes/min
5
+--1----+---+--4----+---1---'
Oouble Plate Total
Cur4t D1alllltl!r Plw.u PkaSt Spacing
S)"lIlbOl ... 11,.) Extractant (tn.) (In.)- (gph/ft ) ,
C 3' "f1ter Waur 1 425
. 36 Water Xylene 1 442
0 3 "tter Water 1 424
424 ,
3 ....., Water 1/2
Fig. 12. Efficiency of the reciprocating plate
extractor [15 ]
System: o-xylene/acetic acid/water
100
80
CL>
.B
en
60
c
OJ
u
"-
OJ
0-
>,
40
u
c
OJ

U
4-
4-
OJ
OJ
en 20
+'
'"
trl
o
Rotating speed, rev/s
Fig. 13. Efficiency of the Scheibel column
extractor [16 ]
System: o-xylene/acetic acid/water
Water extractant
0
[), Xylene extractant
5 10 15
851
ESL-IE-83-04-130

Proceedings from the Fifth Industrial Energy Technology Conference Volume II, Houston, TX, April 17-20, 1983
varying the speed of rotation of the disk dis-
persers.
The sieve tray extractor resembles a sieve tray
distillation column. Downcomers (or upcomers) are
provided to move the continuous phase downward or
upward, depending on whether it is the heavy phase
or the light phase. Tray perforations provide for
drop formation at each stage, thus aiding the mass
transfer process. The sieve tray device is nonpro-
prietary, but because there is relatively little
published information on its performance at the
large scale, engineering firms with experience on
such performance tend to play the role of proprie-
tor. Adiagram of a sieve tray column section is
shown in Fig. 14.
The sieve tray extractor is amenable to mechanistic
modeling by chemical engineers, and approaches to
the modeling have been given by Skelland and Conger
[18], Treybal [12J, and Schulz and Pilhofer [19].
The approach to 100% stage efficiency is governed
by three mechanisms, occurring in sequence:
1. Formation of drops at the sieve tray per-
forations. As noted above, a significant
amount of the total mass transfer occurs
during this process. Factors governing
transfer rate include surface created
(drop size), interfacial tension, wetta-
bility of the tray material by the dis-
persed phase, rate of drop formation (flow
rate through the perforations).
2. Rise of drops through the crossflowing
continuous phase. Mass transfer resis-
tances inside and outside of the drops
must be considered, and the approach to
free rise velocity of the drops must be
taken into account. Distance of drop
travel, a function of tray spacing, influ-
ences the amount of mass transferred, as
shown in Fig. 15, taken from the paper by
Pilhofer [20).
3. Coalescence of drops under the tray above.
The contribution of this mechanism to the
total mass transfer process is usually
quite small.
In the preceding discussion of sieve tray mecha-
nisms, and in the diagram of Fig. 14, the light
phase is dispersed. If it is desirable to disperse
the heavy phase, the tray arrangement can be in-
verted, with continuous phase flowing through up-
comers and the dispersed phase forming drops at the
perforations, which then fall to the tray below,
often attaining a free-fall velocity.
Packed extractors are designed on bases that are
analogous to those for gas-liquid packed columns.
Packing materials are the same (e.g., Intalox sad-
dles, Pall rings, ordered packings of the gauze or
mesh type), and equivalent devices are used for
phase distribution and collection. Adiagram of a
packed extractor is shown in Fig. 16.
The maximum rates of phase flows are obtained from
flood correlations such as that of Nemunaitis et
al. [211. Transfer unit requirements are computed
CONTINUOUS
PHASE

-DISPERSED
l
0
0
0
0

0
0
0
0
0
0
0
0
0
0
0
0
0
0
0")
PHASE
0 0 0 0 0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
!
0 0 0 0 0
Fig. 14. Flows in a sieve tray extraction column
o 10 20

2 3
2.8 4.7 6.0
28.4 23.0 23.6
30 40
Height H/cm
ON hole diameter; D
p
drop diameter;
W hole velocity
N
Fig. 15. Dependence of point efficiency of single
drops on the height of rise for formation
at high hole velocities, with jetting [20]
System: toluene/acetone/water
H52
ESL-IE-83-04-130

Proceedings from the Fifth Industrial Energy Technology Conference Volume II, Houston, TX, April 17-20, 1983
PHASE OUT
interface
HEAVY PHASE IN
heavy phose distributor
pocking holddown grid
pock ing support!
phose redist ri butor
pocking halddown grid
light phose disperser
LIGHT PHASE IN
packing support
HEAVY
)
PHASE OUT
Fig. 16. Diagram of a packed extractor
in standard fashion, and empirical methods are used
to determine transfer unit heights. Examples of
experimentally determined heights of transfer units
are shown in Fig. 17, from the paper by Nemunaitis
et al. The term HOR refers to the height of an
overall transfer unit, based on concentrations in
the raffinate phase. Use of such information is
for the determination of required height of pack-
ing:
where the number of transfer units NOR is calcu-
lated on a basis equivalent to that for equilibrium
stages.
Some guidance in extractor selection may be ob-
tained from Table 1, taken from the paper by Todd
[221. Not all of the currently available extrac-
tion devices are included in the table. Fig. 18,
from the same source, shows approximate areas of
application.
Table 2 shows approximate capacity and efficiency
data for several types of column extractor, drawn
in part from the book by Laddha and Degaleesan
[231. The efficiency criterion is
(4)
combined Phase)( stages per)
(
flow rate unit height '
where
qo _
3
m
u
--2
o
AT -
hr. m
3
_ qc
m
u
--2
c
- AT
hr. m
Note that E
L
is not a fraction or a percentage.
An earlier review of extraction equipment has been
given by Morrelo and Poffenberger [24]. A recent
and excellent treatment of commercially available
extractors is that of Lo [25]. For sieve tray col-
umn design an early article, still useful from a
practical standpoint, is that of Mayfield and
Church [26].
ENERGY CONSIDERATIONS
At the outset of this presentation it was noted
that for some mixture separation problems it is
possible that extraction can have energy
tion advantages over distillation. For extraction,
the major cost of energy is for the solvent strip,-
per. The energy required for dispersing the phases
(mechanical or pressure) is low in comparison.
Thus, as for absorption or straight distillation,
the energy analysis deals primarily with a distil-
lation step.
Comparisons of energy requirements for several
separation processes have been provided by Null
[271. While it is clearly difficult to make gen+
eralizations, Null has presented guidelines that
are of definite interest to process engineers.
comparisons between distillation and extraction are
shown in Figs. 19 and 20. The conditions for
figures represent extreme conditions.
For Fig. 19, a completely nonvolatile solvent is
assumed. This solvent does not contaminate the
raffinate, and requires only simple flash steps nor
separation from the extract. As an example use cif
the figure, if 60% of the feed would be taken over-
head in distillation (DfF = 0.6) and the required
heating medium temperature for distillation is I
149 C (300 F), then any indicated distillation re-
flux ratios (RD) greater than 2.0 would suggest I
consideration of extraction as a viable alternate.
!
Fig. 20 covers the opposite extreme where two
vent strippers would be needed, one for the solvent-
extract separation and one for a
separation: If each of these strippers required
reflux ratlo RE of 2.0, and for the same DfF = 0.6
in distillation, then extraction would merit consi-
deration if the distillation reflux ratio was
greater than 4.0. For Fig. 20, the temperature of
the heating medium for solvent stripping is assumi.d
to be the same as the temperature of the heating
medium for distillation, an extreme situation, and
not a likely one.
Any comparisons of distillation and extraction must
take into account the state of the art on equipment
scaleup and design. One can be much more confident
in distillation than in extraction, because of the
853
ESL-IE-83-04-130

Proceedings from the Fifth Industrial Energy Technology Conference Volume II, Houston, TX, April 17-20, 1983
Continuous Phase Velocity, m/hr
15 30 45 60 75 90
3,-----r----+--t---....----+-__+,
0.75
uo- 7.5 mi.
o.l.-._-+-_-+-__+-_-+-_-t_----J
o 50 100 150 200 250 300
Continuous Phase Velocity. ft/hr
Continuous Phose Velocity. m/hr
15 30 45 60 75 90
8...---+----+-----l----+--t---+,
ZT=1.5m 2.0
\ 1.5
uO"17m/s H ,m
x_ OR
x-r-:-ti---i===i1.0
2
Uo'7
5m
/s 0.5
O.l.--+--+-----t---+--__+---'
o 50 100 150 200 250 300
Continuous Phase Velocity.ft/hr
Fig. 17. Heights of transfer units for kerosene/
MEK/water system, 25-mm metal Pall rings,
0.46-m column, 1.5-m packed height (21)
9f-
::::::::
----.:
8
::::::: Centrifugal
:::::::.
--
Extractors
7-
Pulse Columns
ROC Columns
--
--
--
--
--
6-

Stages
-
-
4 IITITIlI
3 Steve tray
'- columns
2 Mixer-Settlers
I I-//////////[/j
Spray
EASY HARD
Difficulty of Dispersion
low phose phose denity
difference difference
low interfacial----7high interfacial
tension tension
Fig. 18, Areas of application of extraction
devices [22]
c
20
o
+-
u
a
"-
+-
><"0
Ww
.cO
"-
u>
10
.- a
.cLL
3
Vl
w'-
>
o
.D
a
o
o
0::
o 0.2 0.4 0.6 0.8
DfF in DistiIIation Process
required heating medium temperature
for distillation column
reqUired heating medium temperature
for extraction solvent stripper =600 F
(316 C)
R
O
= required reflux ratio for distillation
Fig. 19. Extraction vs. distillation selection,
nonvolatile solvent [27]
c
20
0
+-
u
a
"-
+-
><"0
WW
.cO
"-
10
u >
.- a
.cLL
3
Vl
w
>
0
.D
a
0
0
0::
0 0.2 0.4 0.6 0.8
DfF in Distillation Process
Heating medium temperatures for distillation
and solvent stripping are equal;
reflux ratio for salvent-extract separation (R
E
) =
reflux ratio for solvent-raffinate separation;
distillatian reflux ratio = R
O
'
Fig. 20. Extraction vs. distillation selection,
difficult solvent separation [27]
ESL-IE-83-04-130

Proceedings from the Fifth Industrial Energy Technology Conference Volume II, Houston, TX, April 17-20, 1983
TABLE 1. RATINGS OF SEVERAL COMMERCIAL EXTRACTORS
Abi
Operating
and
to Handle

Contactor
Capital
Cost
Maintenance
Cost Efficiency
Total
Capacity
Flexi-
bi1ity Vol. Eff.
Space
Vert. Floor
That
Emulsify
Spray
Baffle plate
Packed
5
4
4
5
5
5
1
2
2
2
4
2
2
2
2
1
3
2
0
1
1
5
5
5
3
3
3
ROC 3 4 4 3 5 4 3 5 3
Pulsed plate
Mixer sett1er
3
2
3
2
4
3
3
4
4
3
4
3
3
5
5
1
1
0
Centrifugal 1 2 5 3 5 5 5 5 5
5 =desirable; 1 =undesirable
Source: Todd, O. B., Chern. Eng. 69 (14) 156 (July 9, 1962)
TABLE 2. APPROXIMATE EFFICIENCY AND CAPACITY CHARACTERISTICS OF EXTRACTION COLUMNS
Contactor
U
o
+ U
c
(m/hr)
HETS
(m)
Stages/
meter
ii:tJ
(hr:"1)
Spray
Sieve
15-75
3-60
3.0 -6.0
0.3 -1.8
0.3-0.15
0.5-3.30
3-7
1- i 20
Packed 6-45 0.9 -3.0 0.3-1.00 1-27
Karr 18-70 0.2 -0.6 1.6-6.00 17
ROC 18-40 0.15-0.6 1.6-6.60 22 -180
Scheibel 15-30 0.3 -0.6 1.6-3.30 29-pO
much larger amount of development work done in dis-
tillation. There has been no extraction equivalent
of Fractionation Research, Inc., in the area of
large-scale equipment performance testing. How-
ever, for those cases clearly indicating a superi-
ority of extraction as a separation process, rea-
sonable allowances for unknown scaleup factors may
over'come the apparent economic penaIt i es of pro-
ceeding with distillation as the selected method.
CONCLUSIONS
The technology of extraction process design and
development has advanced materially during the past
few decades. Agreat deal of work has been done on
liquid-liquid equilibria, but reliable and general
predictive methods are still not available; direct
experimentation is still needed for all but the
relatively simple systems. Methods are available
for computing theoretical stages or transfer units,
once the equilibria are in hand. The design of the
extraction devices, however, remains mostly in the
proprietary art, and practice is often dependent on
the proprietors for scaleup and design information.
This is a situation that needs correcting.
It is likely that for many applications, a simple
and nonproprietary extraction device such as a
crossflow sieve tray column or a packed column
would suffice. To aid in the development of reli-
able models for such devices, larger-scale perfbrm-
ance data are needed. Such data for distillation
columns have been provided by Fractionation Re-'
search, Inc.; a similar undertaking for extraction
would be welcome.
Finally, there appear to be many situations
extraction would be less energy intensive than 9is-
tillation. Improvement in the state of extraction
technology would enable exploitation of such situa-
tions.
ACKNOWLEDGMENT
This work was supported by the Center for Energy
Studies, The University of at Austin. .
REFERENCES
1. O'Ans-Lax. Taschenbuch fur Chemiker und
Physiker, 3rd. Ed., Vol. 1. Berlin:
Springer-Verlag, 1967.
2. Francis, A. W. LiqUid-Liiuid Equilibriums.
New York: Interscience, 963.
855
ESL-IE-83-04-130

Proceedings from the Fifth Industrial Energy Technology Conference Volume II, Houston, TX, April 17-20, 1983
3. Francis, A. W. Handbook for in
Solvent Extraction. New York: ordon and
Breach, 1972.
4. Himmelblau, D. M., Brady, B. L., McKetta,
J. J. Survey of Solubility Diagrams for
Ternar and uaternar Li uid Sstems. Spe-
Cla Publ. No.3, Bureau of Englneering Re-
search, The University of Texas, Austin,
Texas, 1959.
5. Landolt-Bornstein. Zahlenwerte und Funktionen
aus Naturwissenschaft aus Physik, Chemie,
Geophysik, und Technik, 6th Ed.,
Vol. II (2 ,2c). Ber in: Springer-Verlag,
1950+.
6. Lewis, J. B. Extraction: ACritical
Review. . . London: H. M. Stationery Office,
1953.
7. Prausnitz, J., Anderson, T., Grens, E.,
Eckert, C., Hsieh, R., O'Connell, J. Computer
Calculations for Multicomponent
and Liquid-Liquid Esuilibria. Eng ewood
Cliffs, N.J.: Prentlce-Ha", 1980.
8. Wisniak, J., Tamir, A. Liquid-Liquid Equi-
librium and Extraction: ALiterature Source
Book. Amsterdam: Elsevier, 1980. (Part A).
9. Treybal, R. E., in Chemical Engineers'
Handbook (R. H. Perry and C. H. Chilton,
eds.), 5th Ed., pp. 15-7 to 15-12. New York:
McGraw-Hill, 1973.
10. Scheibel, E. G. Petrol. Refiner 38 (9) 227
(1959).
11. Treybal, R. E. LiaUid Extraction, 2nd Ed.
New York: McGraw- ill, 1963.
12. Treybal, R. E. Mass Transfer 0Gerations, 3rd
Ed. New York: McGraw-Hi", 19 O.
13. Robbins, L. A. "Liquid-Liquid Extraction,"
Section 1.9 in Handbook of Separation Tech-
niques for Chemical Engineers, P. A. Schweit-
zer, Ed. New York: McGraw-Hill, 1979.
14. Karr, A. E. AIChEJ. 5 (4) 446 (1959).
15. Karr, A. E., Lo, T. C. Chem. Eng. Progr. 72
(11) 68 (1976).
16. Scheibel, E. G., Karr, A. E. Ind. Eng. Chem.
42, 1048 (1950).
17. Strand, C. P., Olney, R. B., Ackerman, G. H.
AIChE J. 8, 252 (1962).
18. Skelland, A. H. P., Conger, W. L. Ind. Eng.
Chem., Proc. Des. Devel. 12 (4) 448 (1973).
19. Schulz, L., Pilhofer, T. Intern. Chem. Eng.
22 (1) 61 (1982).
20. Pllhofer, T. Chem. Eng. Commun. II, 241
(1981).
21. Nemunaitis, R. R., Eckert, J. S., Foote,
E. H., Rollison, L. R. Chem. Eng. Progr. 67
(11) 60 (1971).
22. Todd, D. B. Chem. Eng. 69 (14) 156 (July 9,
1962).
23. Laddha, G. S., Degaleesan, T. E. Transport
Phenomena in Liquid Extraction, New Delhi,
India: Tata McGraw-Hill Publ. Co., 1978.
24. Morrelo, V. S., Poffenberger, N. Ind. Eng.
Chem. 42, 1021 (1950).
25. Lo, T. C. "Commercial Liquid-Liquid Extrac-
tion Equipment," Section 1.10 in Handbook of
Separation Techniques for Chemical Engineers,
P. A. Schweitzer, Ed., New York: McGraw-Hill,
1979.
26. Mayfield, F. D., Church, W. L. Ind. Eng.
Chem. 44, 2253 (1952).
27. Null, H. R. Chem. Eng. Progr. 76 (8) 42
(1980).
ESL-IE-83-04-130

Proceedings from the Fifth Industrial Energy Technology Conference Volume II, Houston, TX, April 17-20, 1983

Potrebbero piacerti anche